Content area
This article is the 17th in the Fungal Diversity Notes series which allows the researchers to publish fungal collections with updated reports of fungus-host and fungus-geography. Herein we report 97 taxa with four new genera distributed in three phyla (Ascomycota, Glomeromycota and Mucoromycota), 11 classes, 38 orders and 62 families collected from various regions worldwide. This collection is further classified into taxa from 69 genera with four novel genera namely Jinshana, Lithophyllospora, Parapolyplosphaeria and Stegonsporiicola. Furthermore, 71 new species, 21 new records, one new combination and four novel phylogenetic placements are provided. The new species comprise Acrocalymma estuarinum, Aggregatorygma isidiatum, Alleppeysporonites elsikii, Amphibambusa aquatica, Apiospora hongheensis, Arthrobotrys tachengensis, Calonectria potisiana, Collariella hongheensis, Colletotrichum squamosae, Corynespora chengduensis, Diaporthe beijingensis, Dicellaesporites plicatus, Dicellaesporites verrucatus, Dictyoarthrinium endophyticum, Distoseptispora chiangraiensis, Dothiora eucalypti, Epicoccum indicum, Exesisporites chandrae, Fitzroyomyces pseudopandanicola, Fomitiporia exigua, Fomitiporia rondonii, Fulvifomes subthailandicus, Gigaspora siqueirae, Gymnopus ailaoensis, Hyalorbilia yunnanensis, Hygrocybe minimiholatra, H. mitsinjoensis, H. parviholatra, H. solis, H. vintsy, Helicogermslita kunmingensis, Jinshana tangtangiae, Kirschsteiniothelia dujuanhuensis, Lamproderma subcristatum, Leucoagaricus madagascarensis, Leucocoprinus mantadiaensis, Lithophyllospora australis, Marasmius qujingensis, Melomastia aquilariae, Monoporisporites jansoniusii, M. pattersonii, Monoporisporites valdiyae, Mucispora maesotensis, Mucor soli, Muyocopron yunnanensis, Nigrospora tomentosae, Ocellularia psorirregularis, Ophiocordyceps duyunensis, Oxneriaria nigrodisca, Oxydothis aquatica, O. filiforme, Phacidiella xishuangbannaensis, Phlebiopsis subgriseofuscescens, Pleurothecium takense, Pleurotus tuber-regium, Pseudochaetosphaeronema puerensis, Pseudodactylaria guttulate, Racheliella chinensis, Rhexoacrodictys fangensis, Roussoella neoaquatica, Rubroboletus pruinosus, Sanghuangporus subzonatus, Scytalidium assmuthi, Shrungabeeja kudremukhensis, Spirographa skorinae, Stanjehughesia bambusicola, Stegonsporiicola aurantiaca, Umbelopsis hingganensis, Vararia tenuata, Verruconis pakchongensis, Wongia bandungensis, and Zygosporium cymodoceae. The new combination is Parapolyplosphaeria thailandica (≡ Polyplosphaeria thailandica). The 21 new hosts, geographical and habitat records comprise Acrocalymma fici, Apiculospora spartii, Aspergillus subramanianii, Camposporium ramosum, Clonostachys rogersoniana, Colletotrichum brevisporum, C. plurivorum, Collybiopsis gibbosa, Dictyosporium tratense, Distoseptispora adscendens, Exosporium livistonae, Ganoderma gibbosum, Graphis mikuraensis, Gymnosporangium paraphysatum, Lasiodiplodia thailandica, Moesziomyces bullatus, Penicillium cremeogriseum, P. echinulonalgiovense, P. javanicum, P. lanosocoeruleum, P. polonicum, and Pleurotus tuber-regium. Graphis chlorotica, G. panhalensis and G. parilis are given as novel phylogenetic placements. In addition, we provide the morphology of Tarzetta tibetensis which was missing in the previous Fungal Diversity Notes 1611–1716. Identification of characterization of all these taxa are supported by morphological and multigene phylogenetic analyses.
Table of contents
Ascomycota R.H. Whittaker
Subphylum Pezizomycotina O.E. Erikss. & Winka
Dothideomycetes O.E. Erikss. & Winka
Dothideomycetes orders incertae sedis
Botryosphaeriales C.L. Schoch, Crous & Shoemaker
Botryosphaeriaceae Theiss. & H. Syd. (= Endomelanconiopsidaceae Tao Yang & Crous)
Lasiodiplodia Ellis & Everh.
1818. Lasiodiplodia thailandica Trakun., L. Lombard & Crous (2014) new record (Contributed by Tennakoon DS)
Dyfrolomycetales K.L. Pang, K.D. Hyde & E.B.G. Jones (2013)
Pleurotremataceae Walt. Watson (1929)
Melomastia Nitschke ex Sacc. (1875)
1819. Melomastia aquilariae T.Y. Du & Karun. sp. nov. (Contributed by Du TY)
Kirschsteiniotheliales Hern. -Restr., R.F. Castañeda, Gené & Crous (2017)
Kirschsteiniotheliaceae Boonmee & K.D. Hyde (2012).
Kirschsteiniothelia D. Hawksw. (1985)
1820. Kirschsteiniothelia dujuanhuensis H.W. Shen & Z.L. Luo, sp. nov. (Contributed by Shen HW and Luo ZL)
Muyocopronales Mapook, Boonmee & K.D. Hyde (2016)
Muyocopronaceae K.D. Hyde (2013)
Muyocopron Speg. (1881)
1821. Muyocopron yunnanensis L. Lu, K.D. Hyde & Tibpromma, sp. nov. (Contributed by Lu L)
Venturiales Y. Zhang ter, C.L. Schoch & K.D. Hyde (2011)
Sympoventuriaceae Y. Zhang ter, C.L. Schoch & K.D. Hyde (2011)
Verruconis Samerp., H.J. Choi, van den Ende, Horré & de Hoog (2013)
1822. Verruconis pakchongensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov. (Contributed by Chuaseeharonnachai C, Nuankaew S, Somrithipol S, Boonyuen N and Sri-indrasutdhi V)
Subclass Dothideomycetidae
Dothideales Lindau (1897)
Dothideaceae Chevall. (1826).
Dothiora Fr. (1836)
1823. Dothiora eucalypti Lacerda, Gusmão, G.G. Barreto & A. Rodrigues, sp. nov. (Contributed by Lacerda LT, Rodrigues A, Kooij PW, Barreto GG, and Gusmão LFP)
Mycosphaerellales (Nannf.) P.F. Cannon (2001)
Mycosphaerellaceae Lindau (1897)
Exosporium Link (1809)
1824. Exosporium livistonae S. Anushree, M. André, D. Guillaume & F. Frédéric, (2017) (Contributed by Xia G and Manawasinghe IS)
Pleosporales Luttr. ex M.E. Barr (1987)
Acrocalymmaceae Crous & Trakun. (2014)
Acrocalymma Alcorn & J.A.G. Irwin (1987)
1825. Acrocalymma fici Crous & Trakun. (2014) new record (Contributed by Senanayake IC and Kularathnage NDI)
1826. Acrocalymma estuarinum M.S. Calabon, E.B.G. Jones & K.D. Hyde, sp. nov. (Contributed by Calabon MS)
Corynesporascaceae Sivan. (1996)
Corynespora Güssow (1905)
1827. Corynespora chengduensis Y.P. Chen & Maharachch., sp. nov. (Contributed by Chen YP and Maharachchikumbura SSN)
Didymellaceae Gruyter, Aveskamp & Verkley (2009)
Epicoccum Link (1816)
1828. Epicoccum indicum S. Rajwar & Raghv. Singh, sp. nov. (Contributed by Rajwar S and Singh R)
Didymosphaeriaceae Munk A. (1953)
Dictyoarthrinium S. Hughes (1952)
1829. Dictyoarthrinium endophyticum R.M.F. Silva, T.G.L. Oliveira & G.A. Silva, sp. nov. (Contributed by Silva RMF, Oliveira TGL and Silva GA)
Dictyosporiaceae Boonmee & K.D. Hyde (2016)
Dictyosporium Corda (1836)
1830. Dictyosporium tratense J. Yang & K.D. Hyde (2018) new record (Contributed by Shu YX)
Macrodiplodiopsidaceae Voglmayr, Jaklitsch & Crous (2015)
Pseudochaetosphaeronema Punith. (1979)
1831. Pseudochaetosphaeronema puerensis R.F. Xu, K.D. Hyde & Tibpromma, sp. nov. (Contributed Xu RF and Tibpromma S)
Melanommataceae G. Winter [as ‘Melanommeae’] (1885)
Camposporium Harkn., (1884)
1832. Camposporium ramosum Whitton, McKenzie & K.D. Hyde (2002) new record (Contributed by Xiong YR and Manawasinghe IS)
Roussoellaceae Jian K. Liu, Phook., D.Q. Dai & K.D. Hyde (2014)
Roussoella Sacc (1888)
1833. Roussoella neoaquatica W.H. Tian & Maharachch., sp. nov. (Contributed by Tian WH and Maharachchikumbura SSN)
Tetraplosphaeriaceae Kaz. Tanaka & K. Hiray. (2009)
1834. Parapolyplosphaeria Rajeshk., C.G. Lin, Dong Wei & K.D. Hyde., gen. nov. (Contributed by Rajeshkumar KC, Lin CG, Dong W and Hyde KD)
1835. Parapolyplosphaeria thailandica (C.G. Lin, Yong Wang bis & K.D. Hyde) Rajeshk., C.G. Lin, Dong Wei & K.D. Hyde. comb. nov. (Contributed by Rajeshkumar KC, Lin CG, Dong W and Hyde KD)
Shrungabeeja V.G. Rao & K.A. Reddy (1981)
1836. Shrungabeeja kudremukhensis O.P. Sruthi & Rajeshk., sp. nov. (Contributed by Sruthi OP and Rajeshkumar KC)
Eurotiomycetes O.E. Erikss. & Winka (1997)
Chaetothyriales M.E. Barr (1987)
Trichomeriaceae Chomnunti & K.D. Hyde, (= Strelitzianaceae Crous & M.J. Wingf 2013)
1837. Lithophyllospora Selbmann, Coleine gen. nov. (Contributed by Selbmann L. and Coleine C)
1838. Lithophyllospora australis Selbmann & Coleine, sp. nov. (Contributed by Selbmann L and Coleine C)
Eurotiales G.W. Martin ex Benny & Kimbr. (1980)
Aspergillaceae Link (1826)
1839. Aspergillus subramanianii Visagie, Frisvad & Samson, (2014) new record (Contributed by Rajeshkumar KC and Sruthi OP)
Penicillium Link (1809)
1840. Penicillium cremeogriseum Chalab. (1950) new record (Contributed by Rajeshkumar KC and Ashtekar N)
1841. Penicillium echinulonalgiovense S. Abe ex Houbraken & R.N. Barbosa (2018) new record (Contributed by Rajeshkumar KC and Varghese S)
1842. Penicillium javanicum J.F.H. Beyma (1929) new record (Contributed by Rajeshkumar KC and Harikrishnan K)
1843. Penicillium lanosocoeruleum Thorn (1930) new record (Contributed by Rajeshkumar KC and Ashtekar N)
1844. Penicillium polonicum K.W. Zaleski (1927) new record (Contributed by Rajeshkumar KC and Ashtekar N)
Lecanoromycetes O.E. Erikss. & Winka (1997)
Graphidales Bessey (1907)
Diploschistaceae Zahlbr. [as ‘Diplochistaceae’] (1905)
Aggregatorygma M. Cáceres, Aptroot & Lücking (2014)
1845. Aggregatorygma isidiatum Aptroot, L.A. Santos & M. Cáceres, sp. nov. (Contributed by Aptroot A, Santos LA dos and Cáceres MES)
1846. Ocellularia psorirregularis Aptroot, L.A. Santos & M. Cáceres, sp. nov. (Contributed by Aptroot A, Santos LA dos and Cáceres MES)
Graphidaceae Dumort. [as ‘Graphineae’] (1822)
Graphis Adans., (1763)
1847. Graphis chlorotica A. Massal., (1871) new phylogenetic placement (Contributed by Ansil PA and Rajeshkumar KC)
1848. Graphis mikuraensis Y. Ohmura & M. Nakan. (2016) new phylogenetic placement (Contributed by Ansil PA and Rajeshkumar KC)
1849. Graphis panhalensis (Patw. & C.R. Kulk.) Chitale, Makhija & B.O. Sharma, (2011) new phylogenetic placement (Contributed by Ansil PA and Rajeshkumar KC)
1850. Graphis parilis Kremp. (1876) new phylogenetic placement (Contributed by Ansil PA and Rajeshkumar KC)
Ostropales Nannf. (1932)
Spirographaceae Flakus, Etayo & Miądl., (2019)
Spirographa Zahlbr., (1903)
1851. Spirographa skorinae Tsurykau, Brackel, Flakus & Kukwa sp. nov. (Contributed by Tsurykau A, Brackel Wv, Flakus A and Kukwa M)
Stictidaceae Fr. [as 'Stictei'] (1849)
1852. Fitzroyomyces pseudopandanicola S.C. He & D.P. Wei, sp. nov. (Contributed by He SC and Wei DP)
1853. Phacidiella xishuangbannaensis D.P. Wei, sp. nov. (Contributed by Wei DP)
Pertusariales M. Choisy ex D. Hawksw. & O.E. Erikss (1986)
Megasporaceae Lumbsch (1994)
Oxneriaria S.Y. Kondr. & Lőkös (2017)
1854. Oxneriaria nigrodisca Usman & Khalid sp. nov. (Contributed by Usman M and Khalid AN)
Leotiomycetes O.E. Erikss. & Winka (1997)
Helotiaceae Rehm [as ‘Helotieae’], (1892)
Scytalidium Pesante (1957)
1855. Scytalidium assmuthi G Mane, R Avchar, R Morey, R Sharma, sp. nov. (Contributed by Mane G and Avchar R).
Vibrisseaceae Korf (1972)
Apiculospora Wijayaw., Camporesi, A.J.L. Phillips & K.D. Hyde (2016)
1856. Apiculospora spartii Wijayaw., W.J. Li, Camporesi, A.J.L. Phillips & K.D. Hyde (2016) new record (Contributed by Dissanayake AJ)
Orbiliomycetes O.E. Erikss. & Baral (2003)
Orbiliales Baral, O.E. Erikss., G. Marson & E. Weber (2003)
Orbiliaceae Nannf. (1932)
Arthrobotrys Corda, (1839)
1857. Arthrobotrys tachengensis F. Zhang & X.Y. Yang sp. nov. (Contributed by Zhang F & Yang YX)
Hyalorbilia Baral & G. Marson (2001)
1858. Hyalorbilia yunnanensis C.J.Y. Li & K.D. Hyde sp. Nov (Contributed by Li C)
Sordariomycetes O.E. Erikss. & Winka (1997)
Subclass Diaporthomycetidae Senan., Maharachch. & K.D. Hyde (2015)
Diaporthales Nannf. (1932)
Diaporthaceae Höhn. ex Wehm. (1926)
Diaporthe Fuckel (1867)
1859. Diaporthe beijingensis Y.Y. Zhou & W. Zhang, sp. nov. (Contributed by Zhou YY and Zhang W)
Tubakiaceae U. Braun, J.Z. Groenew. & Crous (2018)
Racheliella Crous & U. Braun (2018)
1860. Racheliella chinensis Y.X. Zhang, J.Y. Lin & Manawas., sp. nov. (Contributed by Zhang YX, Lin JY and Manawasinghe IS)
Distoseptisporales Z.L. Luo, K.D. Hyde & H.Y. Su (2019)
Distoseptisporaceae Z.L. Luo, K.D. Hyde & H.Y. Su (2019)
Distoseptispora K.D. Hyde, McKenzie & Maharachch (2016)
1861. Distoseptispora adscendens (Berk.) R. Zhu & H. Zhang (2022) new record (Contributed by Sruthi OP and Rajeshkumar KC)
1862. Distoseptispora Chiangraiensis R.J. Xu, Q. Zhao & K.D. Hyde sp. nov. (Contributed by Xu RJ and Zhao Q)
Diaporthomycetidae families incertae sedis
Papulosaceae Winka & O.E. Erikss. (2000)
Wongia Khemmuk, Geering & R.G. Shivas (2016)
1863. Wongia bandungensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov. (Contributed by Chuaseeharonnachai C, Nuankaew S, Somrithipol S and Boonyuen N)
Subclass Hypocreomycetidae O.E. Erikss. & Winka (1997)
Glomerellales Chadef. ex Réblová, W. Gams & Seifert (2011)
Glomerellaceae Locq. ex Seifert & W. Gams in (2007)
Colletotrichum Corda (1831)
1864. Colletotrichum squamosae W.A.S. Vieira, V. Doyle & M.P.S. Câmara, sp. nov. (Contributed by Vieira WAS)
1865. Colletotrichum brevisporum Noireung, Phouliv., L. Cai & K.D. Hyde (2012) new record (Contributed by Armand A and Jayawardena RS)
1866. Colletotrichum plurivorum Damm, Alizadeh & Toy. Sato (2018) new record (Contributed by Armand A and Jayawardena RS)
Hypocreales Lindau (1897)
Bionectriaceae Samuels & Rossman (1999)
Clonostachys Corda (1839)
1867. Clonostachys rogersoniana Schroers (2001) new record (Contributed by Gajanayake AJ)
Nectriaceae Tul. & C. Tul. [as ‘Nectriei’] (1865)
Calonectria De Not. (1867)
1868. Calonectria potisiana Melo & R.F. Alfenas, sp. nov. (Contributed by Melo MP and Alfena RF)
Ophiocordycipitaceae G.H. Sung, J.M. Sung, Hywel-Jones & Spatafora (2007)
Ophiocordyceps Petch (1931)
1869. Ophiocordyceps duyunensis X.C. Peng & T.C. Wen sp. nov. (Contributed by Peng XC and Wen TC)
Microascales Luttr. ex Benny & R.K. Benj. (1980)
Halosphaeriaceae E. Müll. & Arx ex Kohlm. (1972)
1870. Jinshana K.L. Pang, M.W.L. Chiang & E.B.G. Jones, gen. nov. (Contributed by Pang KL, Chiang MWL and Jones EBG)
1871.Jinshana tangtangiae K.L. Pang, M.W.L. Chiang & E.B.G. Jones, sp. nov. (Contributed by Pang KL, Chiang MWL and Jones EBG)
Subclass Savoryellomycetidae Hongsanan, K.D. Hyde & Maharachch.
Fuscosporellales Z.L. Luo, K.D. Hyde & H.Y. Su Jing Yang, Bhat & K.D. Hyde (2016)
Fuscosporellaceae Jing Yang, Bhat & K.D. Hyde (2016)
Mucispora Jing Yang, Bhat & K.D. Hyde (2016)
1872. Mucispora maesotensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov. (Contributed by Chuaseeharonnachai C, Nuankaew S, Somrithipol S and Boonyuen N)
Pleurotheciales Réblová & Seifert (2015)
Pleurotheciaceae Réblová & Seifert (2015)
Pleurothecium Höhn., (1923)
1873. Pleurothecium takense Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov. (Contributed by Chuaseeharonnachai C, Nuankaew S, Somrithipol S, Boonyuen N and Kwantong P)
Savoryellales Boonyuen, Suetrong, Sivichai, K.L. Pang & E.B.G. Jones (2011)
Savoryellaceae Jaklitsch & Réblová (2015)
Rhexoacrodictys W.A. Baker & Morgan-Jones (2002)
1874. Rhexoacrodictys fangensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov. (Contributed by Chuaseeharonnachai C, Nuankaew S, Somrithipol S, Boonyuen N and Luangsa-ard JJ)
Subclass Sordariomycetidae O.E. Erikss & Winka (= Meliolomycetidae P.M. Kirk & K.D. Hyde)
Pseudodactylariales Crous (2017)
Pseudodactylariaceae Crous (2017)
Pseudodactylaria Crous (2017)
1875. Pseudodactylaria guttulate J. Ma & Y.Z. Lu, sp. nov. (Contributed by Ma J and Lu YZ)
Sordariales Chad. ex D. Hawksw. & O.E. Erikss (1986)
Chaetomiaceae G. Winter [as ‘Chaetomieae’] (1885)
Collariella X. Wei Wang, Samson & Crous (2016)
1876. Collariella hongheensis X.F. Liu, Karun. & Tibpromma sp. nov. (Contributed by Liu XF)
Subclass Xylariomycetidae O.E. Erikss & Winka (1997)
Amphisphaeriales D. Hawksw. & O.E. Erikss (1986)
Apiosporaceae K.D. Hyde, J. Fröhl., Joanne E. Taylor & M.E. Barr (1998)
Apiospora Sacc. (1875)
1877. Apiospora hongheensis X. Zhang & Tibpromma, sp. nov. (Contributed by Zhang X)
Nigrospora Zimm. (1902)
1878. Nigrospora tomentosae M. Luo, & J.W. Liu, sp. nov. (Contributed by Liu JW and Luo M)
Oxydothidaceae Konta & K.D. Hyde (2016)
Oxydothis Penz. & Sacc. (1898)
1879. Oxydothis aquatica Y.Y. Yang & K.D. Hyde, sp. nov. (Contributed by Yang YY)
1880. Oxydothis filiforme L.W. Li & Jian K. Liu, sp. nov. (Contributed by Li WL and Liu JK)
Xylariales Nannf. (1932)
Cainiaceae J.C. Krug (1978)
Amphibambusa D.Q. Dai & K.D. Hyde (2015)
1881. Amphibambusa aquatica Doilom sp. nov. (Contributed by Doilom M and Dong W)
Xylariaceae Tul. & C. Tul. [as ‘Xylariei’] (1863)
Helicogermslita Lodha & D. Hawksw., (1983)
1882. Helicogermslita kunmingensis Y. Gao & H. Gui, sp. nov. (Contributed by Gao Y and Gui H)
Zygosporiaceae J.F. Li, Phookamsak & K.D. Hyde (2017)
Zygosporium Mont. (1842)
1883. Zygosporium cymodoceae Guerra, Baulin, Cano & Gené, sp. nov. (Contributed by Guerra-Mateo D, Baulin V, Cano-Lira JF and Gené J)
Sordariomycetes genera incertae sedis
Stanjehughesia Subram. (1992)
1884. Stanjehughesia bambusicola X.D. Yu & Jian K. Liu, sp. nov. (Contributed by Yu XD and Liu JK)
Basidiomycota R.T. Moore (1980)
Subphylum Agaricomycotina Doweld (2001)
Agaricales Underw. (1899)
Agaricaceae Chevall. (1826)
Leucoagaricus Locq. Ex Singer (1948)
1885. Leucoagaricus madagascarensis Ralaiv., Liimat., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, and Niskanen T)
Leucocoprinus Pat. (1988)
1886. Leucocoprinus mantadiaensis Ralaiv., Liimat., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, & Niskanen T)
Hygrophoraceae Lotsy (1907)
Hygrocybe P. Kumm. (1871)
1887. Hygrocybe minimiholatra Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, & Niskanen T)
1888. Hygrocybe mitsinjoensis Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, & Niskanen T)
1889. Hygrocybe parviholatra Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, & Niskanen T)
1890. Hygrocybe solis Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, & Niskanen T)
1891. Hygrocybe vintsy Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov. (Contributed by Ralaiveloarisoa A, Liimatainen K, Ainsworth AM, Niskanen T)
Marasmiaceae Roze ex Kühner (1980)
Marasmius Fr. (1836)
1892. Marasmius qujingensis W.H. Lu., Karun. & Tibpromma, sp. nov. (Contributed by Lu W)
Omphalotaceae Bresinsky (1985)
Collybiopsis (J. Schröt.) (1909)
1893. Collybiopsis gibbosa (Corner) R.H. Petersen (2021) (Contributed by Yuwei Hu)
Gymnopus (Pers. 1801: 302) Roussel (1806)
1894. Gymnopus ailaoensis Y. Hu sp. nov. (Contributed by Yuwei Hu)
Pleurotaceae Kühner (1980)
Pleurotus (Fr.) P. Kumm (1871)
1895. Pleurotus tuber-regium (Fr.) Singer (1951) (Contributed by Phonemany M and Thongklang N)
Boletales E.J. Gilbert (1931)
Boletaceae Chevall (1826)
Rubroboletus Kuan Zhao & Zhu L. Yang (2014)
1896. Rubroboletus pruinosus Salna Nanu & T.K.A. Kumar, sp. nov. (Contributed by Nanu S and Kumar TKA)
Corticiales K.H. Larss. (2007)
Corticiaceae Herter (1910)
1897. Stegonsporiicola Voglmayr, gen. nov. (Contributed by Voglmayr H)
1898. Stegonsporiicola aurantiaca Voglmayr, sp. nov. (Contributed by Voglmayr H)
Hymenochaetales Oberw. (1977)
Hymenochaetaceae Donk (1948)
Fomitiporia Murrill (1907)
1899. Fomitiporia exigua Alves-Silva, Góes-Neto & Drechsler-Santos, sp. nov. (Contributed by Alves-Silva G, Drechsler-Santos ER, da Silveira RMB & Góes-Neto A)
1900. Fomitiporia rondonii Alves-Silva & Drechsler-Santos, sp. nov. (Contributed by Alves-Silva G, Drechsler-Santos ER, da Silveira RMB & Góes-Neto A)
Fulvifomes Murrill (1914)
1901. Fulvifomes subthailandicus E. Arumugam, S. Gunaseelan, K. Kezo & M. Kaliyaperumal sp. nov. (Contributed by Contributed by Kaliyaperumal M, Arumugam E, Gunaseelan S & Kezo K)
Sanghuangporus Sheng H. Wu, L.W. Zhou & Y.C. Dai (2015)
1902. Sanghuangporus subzonatus S. Gunaseelan, M. Kaliyaperumal & K. Kezo sp. nov. (Contributed by Kaliyaperumal M, Gunaseelan S and Kezo K)
Polyporales Gäum. (1926)
Ganodermataceae (Donk) Donk (1948)
Ganoderma P. Karst. (1881)
1903. Ganoderma gibbosum (Blume and T. Nees) Pat. (1897) new record (Contributed by Luangharn T)
Phanerochaetaceae Jülich (1981)
Phlebiopsis Jülich (1978)
1904. Phlebiopsis subgriseofuscescens Kezo, K., Kaliyaperumal, M. & Gunaseelan, S., sp. nov. (Contributed by Kaliyaperumal M, Kezo K and Gunaseelan S)
Russulales Kreisel ex P.M. Kirk, P F. Cannon & J.C. David (2001)
Peniophoraceae Lotsy (1907)
Vararia P. Karst. (1898)
1905. Vararia tenuata Ghobad-Nejhad, sp. nov. (Contributed by Ghobad-Nejhad M and Langer E)
Pucciniomycetes R. Bauer, Begerow, J.P. Samp., M. Weiss & Oberw. (2006)
Pucciniales Clem. & Shear (1881)
Gymnosporangiaceae Chevall. (1826)
Gymnosporangium R. Hedw. ex DC. (1805)
1906. Gymnosporangium paraphysatum Vienn. -Bourg., Revue de Mycologie (Paris) 25 (5): 304 (1961) new record (Contributed by Shen YM)
Ustilaginomycetes Warm. (1884)
Ustilaginales Bek. (1864)
Ustilaginaceae Tul. & C. Tul. (1847)
Moesziomyces Vánky (1977)
1907. Moesziomyces bullatus (J. Schröt.) Vánky, Bot. Notiser 130(2): 133 (1977) new record (Contributed by Denchev TT, Denchev CM, Begerow D and Kemler M)
Glomeromycota C. Walker & A. Schüssler
Glomeromycetes Caval. -Sm. emend. Oehl, G.A. Silva, B.T. Goto & Sieverd.
Gigasporales S.P. Gautam & U.S. Patel (= Gigasporales Sieverd., G.A. Silva, B.T. Goto & Oehl)
Gigasporaceae J.B. Morton & Benny, Mycotaxon 37: 483 (1990)
Gigaspora Gerd. & Trappe Mycol. Mem. 5: 25 (1974)
1908. Gigaspora siqueirae F.A. de Souza, Barros-Barreto, Magurno, B.T. Goto, sp. nov. (Contributed by Goto BT)
Mucoromycota Doweld (2001)
Mucormycetes Doweld (2001)
Mucorales Dumort. [as ‘Mucorarieae’] (1829)
Mucoraceae Fr. [as ‘Mucoroidei’] (1821)
Mucor Fresen. (1850)
1909. Mucor soli C.A. de Souza, E.V. de Medeiros & R.J.V de Oliveira, sp. nov. (Contributed by de Souza CA, de Medeiros EV and de Oliveira RJV)
Umbelopsidomycetes Tedersoo (2018)
Umbelopsidales Spatafora, Stajich & Bonito (2016).
Umbelopsidaceae W. Gams & W. Mey. (2003)
Umbelopsis Amos & H.L. Barnett (1966)
1910. Umbelopsis hingganensis Tong Wu, T. Du, M. M. Ding & L.J. Xu, sp. nov. (Contributed by Wu T, Du T, Ding M, and Xu L)
Fungus-like taxa
Myxomycetes G. Winter (1880)
Physarales T. Macbr., N. Amer (1922)
Lamprodermataceae T. Macbr. [as “Lamprodermeae”] (1899)
Lamproderma Rostaf. (1873)
1911. Lamproderma subcristatum G. Moreno, López-Vill. & A. Sánchez, sp. nov. (Contributed by Moreno G, López-Villalba Á, and Sánchez A.)
Fungi Imperfecti
1912. Exesisporites chandrae R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK & Kumar A)
1913. Monoporisporites jansoniusii R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK and Kumar A)
1914. Monoporisporites pattersonii R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK and Kumar A)
1915. Monoporisporites valdiyae R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK and Kumar A)
1916. Dicellaesporites plicatus R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK and Kumar A)
1917. Dicellaesporites verrucatus R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK and Kumar A)
1918. Alleppeysporonites elsikii R.K. Saxena & A. Kumar sp. nov. (Fossil) (Contributed by Saxena RK and Kumar A)
Introduction
Fungi are one of the most diverse groups of organisms that play a vital role in ecosystem stability (Schimann et al. 2017). Despite their role in the ecosystem as decomposers, pathogens, and symbionts, fungi are also widely utilised in food, biochemical, and biotechnological fields (Hyde et al. 2020a, b, c). Fungi have amazing diversity in which their lifestyles vary as saprobes, epiphytes, endophytes, pathogens, or symbionts, and they are morphologically diverse from huge mushrooms to single-cell yeasts (Naranjo-Ortiz and Gabaldón 2019). The estimation of the number of fungi could range from 1.5 to 12 million species (Hyde et al. 2020d; Bhunjun et al. 2022; Niskanen et al. 2023), whereas only 155,000 species have been identified to date (Lücking et al. 2021). Thus, it reflects that many species remain to be discovered.
With the advancement of DNA-based techniques, the number of novel taxa described increased rapidly. In modern fungal taxonomy, species definition is based on polyphasic approaches in which authors cooperate morphology, multigene phylogeny, habitat variations, and biochemical compounds to introduce novel species (Chethana et al. 2021a, b; Jayawardena et al. 2021a; Manawasinghe et al. 2021; Maharachchikumbura et al. 2021). Furthermore, other than novel species, re-collection of previously described taxa to provide living cultures and sequence data is also necessary for present fungal taxonomy (Ariyawansa et al. 2014a, b; Chethana et al. 2021a). In addition to that new host records and geographical records also play an important role especially in disease management (Dugan et al. 2009). Even though these introductions have a significant impact on fungal classification, publishing a single new species or new host record is always challenging.
Different fungal collection series are available to facilitate mycologists publishing their valuable collections. These series include Fungal Diversity notes (Liu et al. 2015a, b; Ariyawansa et al. 2014a; Hyde et al. 2017, 2019, 2020a, b, c; Tibpromma et al. 2018; Wanasinghe et al. 2018; Phookamsak et al. 2019; Boonmee et al. 2021; Jayawardena et al. 2023; Senanayake et al. 2023), Mycosphere notes (Thambugala et al. 2017; Hyde et al. 2018, 2021; Manawasinghe et al. 2022; Hyde et al. 2023b; Dong et al. 2023; Li et al 2023a, b), Fungal Planet description sheets (Crous et al. 2015a, b, 2019b) and AJOM new records and collections of fungi (Hyde et al. 2019; Chethana et al. 2021b, 2023). These publications have resulted in numerous novel species, new genera, new combinations, new host records and geographical records. Moreover, these publications will continuously update the taxonomic status of many genera and families. As the 17th paper of the Fungal Diversity Notes series here, we provide new data including morphological, geographical and sequence data for a stable taxonomy and phylogeny.
Materials and methods
Materials and methods follow the previous fungal diversity notes (Hyde et al. 2016, 2020a, b, c; Tibpromma et al. 2018; Wanasinghe et al. 2018; Phookamsak et al. 2019; Boonmee et al. 2021) and Senanayake et al. (2020). Based on Dissanayake et al. (2020) phylogenetic analyses were performed by maximum likelihood (ML), maximum parsimony (MP) and Bayesian posterior probability (BYPP). For accurate species identification, Cao et al. (2021), Chethana et al. (2021a), Manawasinghe et al. (2021), Maharachchikumbura et al. (2021), Jayawardena et al. (2021b) and Pem et al. (2021) were followed. All taxonomic novelties are submitted to Index Fungorum, MycoBank and Faces of Fungi (Jayasiri et al. 2015). Sequence data has been deposited in the GenBank and accession numbers are provided with each species description. Descriptions and illustrations are provided for all identified genera and species.
Taxonomy
Ascomycota R.H. Whittaker
We follow the treatments and updated accounts of Ascomycota in Wijayawardene et al. (2020, 2022).
Subphylum Pezizomycotina O.E. Erikss. & Winka
Dothideomycetes O.E. Erikss. & Winka
Notes: For the taxonomic treatment of Dothideomycetes, we follow Hongsanan et al. (2020a, b) and Wijayawardene et al. (2020, 2022).
Dothideomycetes orders incertae sedis
Notes: For the taxonomic treatments, of Dothideomycetes orders incertae sedis, we follow Hongsanan et al. (2020b) and Wijayawardene et al. (2020, 2022).
Botryosphaeriales C.L. Schoch, Crous & Shoemaker
Notes: Botryosphaeriales species are ecologically diverse and widely distributed species. They are commonly associated with cankers, die-back, leaf spots and root rots of diverse woody plants (Yang et al. 2017; Wu et al. 2022). For the taxonomic treatment of Botryosphaeriales, we follow Zhang et al. (2021) and Wu et al. (2021).
Botryosphaeriaceae Theiss. & H. Syd. (= Endomelanconiopsidaceae Tao Yang & Crous).
Notes: Botryosphaeriaceae was described by Theissen and Sydow (1918). These species have worldwide distribution as endophytes, saprobes, and pathogens, especially on woody crops and forest plants as pathogens (Hilário et al. 2020; Zhou et al. 2023). Botryosphaeriaceae encompasses 22 genera (Phillips et al. 2013). For the taxonomic treatment of this family, we follow Zhang et al. (2021).
Lasiodiplodia Ellis & Everh., Bot. Gaz. 21: 92 (1896)
Notes: Ellis and Everhart (1894) introduced this genus to include L. tubericola as the type and formally described by Clendenin (1896). Lasiodiplodia species have a cosmopolitan distribution worldwide and occur on a wide range of monocotyledonous, dicotyledonous and gymnosperm hosts (Correia et al. 2016; Rosado et al. 2016; de Silva et al. 2019). Lasiodiplodia species have been recorded as endophytes, saprobes and plant pathogens that cause cankers, die-back, fruit or root rot, branch blight or discoloration on a wide range of woody hosts (Alves et al. 2008; Phillips et al. 2013; de Silva et al. 2019; Gómez et al. 2021a, b). This genus can be distinguished from other Botryosphaeriaceae genera in having pycnidial paraphyses and longitudinal striations on the mature conidia (Phillips et al. 2008, 2013; Rathnayaka et al 2023). Up to date, 65 Lasiodiplodia species are listed in Species Fungorum (2024) (Fig. 1).
[See PDF for image]
Fig. 1
The best scoring RAxML tree with a final likelihood value of − 4936.454266 for the combined dataset of ITS, tef1-α and tub2 sequence data. The topology and clade stability of the combined gene analyses were compared to the single gene analyses. The tree is rooted with Diplodia mutila (CMW7060). The matrix had 373 distinct alignment patterns with 17.69% undetermined characters and gaps. Estimated base frequencies were as follows; A = 0.208520, C = 0.302466, G = 0.256954, T = 0.232060; substitution rates AC = 1.235624, AG = 3.293289, AT = 1.444962, CG = 1.060100, CT = 5.061546, GT = 1.000000; gamma distribution shape parameter α = 0.676319. Ex-type strains are in bold and newly generated sequences are in red. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.95 are given above the nodes
Lasiodiplodia thailandica Trakun., L. Lombard & Crous, Persoonia 34: 95 (2014)
Index Fungorum number: IF810169; Facesoffungi number: FoF09333, Fig. 2
[See PDF for image]
Fig. 2
Lasiodiplodia thailandica (NCYU 19-0402, new host record) a A dead leaf of Celtis tetrandra. b, c Appearance of conidiomata on host. d Section of conidioma. e Conidioma wall. f Conidiogenous cells with developing conidia. g, h Conidia. i Germinated conidium. j Colony from above. k Colony from below. Scale bars: d = 50 µm, e = 10 µm, f–i = 12 µm
Saprobic on dead leaves of Celtis tetrandra Roxb. (Cannabaceae). Sexual morph: Not observed. Asexual morph: Coelomycetous. Conidiomata 150–250 µm high, 200–300 µm diam., pycnidial, semi-immersed, becoming erumpent, solitary or gregarious, brown to black, globose to subglobose, uniloculate, ostiolate. Conidiomatal wall 15–20 μm, consisting of 4–5 layers, thin-walled, of equal thickness, pale brown to dark brown pseudoparenchymatous cells, cells towards the inside light brown, arranged in a textura angularis, fusing and indistinguishable from the host tissues. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 7–10 × 2–4 μm (x̄ = 8 × 3 μm, n = 30), hyaline, smooth, thin-walled, discrete, cylindrical, holoblastic, proliferating percurrently from hyaline inner conidiomatal wall. Conidia 22–26 × 12–14 μm (x̄ = 24 × 13 μm), initially hyaline, turning pale brown at maturity, aseptate, with a single median septum and longitudinal striations after discharge from the pycnidia, oblong to ovoid, straight, both ends broadly rounded, thin-walled.
Culture characteristics: Colonies on PDA reaching 30 mm diam. after 4 days at 20–25 °C, colonies medium sparse, circular, raised, surface slightly rough with entire edge, cottony to fairly fluffy with sparse aspects, colony from above: greenish to grey; reverse: greenish to pale brown. Mycelium whitish grey with tufting; not producing pigments in PDA.
Material examined: China, Chiayi, Fanlu Township area, dead leaves of Celtis tetrandra (Cannabaceae), 11 June 2019, D.S. Tennakoon, GSPD061R (NCYU 19-0402, new host record), living culture, NCYUCC 19-0392.
GenBank Numbers: ITS: OR759966, tub2: OR767663, tef1-α: OR767664
Notes: The morphological characteristics of our collection (NCYU 19-0402) tally well with the type of Lasiodiplodia thailandica in having similar size range of conidiogenous cells (7–10 × 2–4 μm vs 8–9 × 2–4 µm), conidia (22–26 × 12–14 μm vs 22–25 × 13–15 μm) and other conidial characters (e.g., initially hyaline, aseptate, turning pale brown with maturity, with a single median septum and longitudinal striations after discharge from the pycnidia, oblong to ovoid) (Trakunyingcharoen et al. 2014). Multigene phylogeny (ITS, tef1-α and tub2) also directs that our collection clustered with L. thailandica species in a well-supported clade (Fig. 1, 80% ML and 1.00 BYPP). Thus, we report our collection as a new host record of L. thailandica from Celtis tetrandra (Cannabaceae).
Dyfrolomycetales K.L. Pang, K.D. Hyde & E.B.G. Jones, in Hyde et al., Fungal Diversity 63: 7 (2013)
Notes: Dyfrolomycetales consist of marine, terrestrial, and wood-inhabiting taxa. They are characterized by immersed, ostiolate, clypeate, papillate ascomata and bitunicate, cylindrical, short pedicellate asci, with a distinct ocular chamber along with a ring-like subapical ring, and overlapping uni-seriate, broadly fusiform, symmetrical, hyaline, multi-septate ascospores. Only Pleurotremataceae is accepted in this order (Hongsanan et al. 2020a; Wijewardena et al. 2021).
Pleurotremataceae Walt. Watson, New Phytol. 28: 113 (1929)
Notes: Pleurotremataceae was established by Watson (1929) to accommodate Pleurotrema Müll. Arg., with P. polysemum (Nyl.) Müll. Arg. as the type species. Wijayawardene et al. (2022) listed three genera viz. Dyfrolomyces K. D. Hyde, Melomastia Nitschke ex Sacc and Pleurotrema in Pleurotremataceae (Dyfrolomycetales, Dothideomycetes). However, Li et al. (2022) synonymized Dyfrolomyces under Melomastia based on multigene phylogeny of combined LSU, SSU and tef1-α and morphological evidence. Subsequently, Dyfrolomyces was maintained in Pleurotremataceae by Kularathnage et al. (2023) based on Dyfrolomyces and Melomastia showing differences in the morphology and septation of ascospore. Pleurotremataceae is characterized by immersed ascomata, bitunicate asci, and multi-septate ascospores (Watson 1929; Barr 1994; Li et al. 2022; Wen-Li et al. 2022). Members of this family are saprobes distributed on decaying wood in terrestrial, mangrove, and freshwater habitats (Li et al. 2022; Kularathnage et al. 2023).
[See PDF for image]
Fig. 3
Phylogram generated from maximum likelihood analysis based on combined LSU, SSU and tef1-α sequence data of 42 taxa, which comprised 2606 base pairs of LSU = 764, SSU = 968, tef1-α = 874. The best scoring RAxML tree with a final likelihood value of − 10,654.707840 is presented. The matrix had 720 distinct alignment patterns, with 25.58% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.235408, C = 0.264506, G = 0.293087, T = 0.206999; substitution rates: AC = 0.801663, AG = 1.998748, AT = 1.048958, CG = 0.946613, CT = 8.369181, GT = 1.000000; gamma distribution shape parameter α = 0.590876. Bootstrap support values for maximum likelihood (ML) equal to or greater than 50% and clade credibility values equal to or greater than 0.90 from Bayesian inference analysis are labelled at each node. The new isolates are indicated in blue bold. The tree is rooted to Anisomeridium phaeospermum (MPN539) and A. ubianum (MPN94)
Melomastia Nitschke ex Sacc., Atti Soc. Veneto-Trent. Sci. Nat., Padova, Sér. 44: 90 (1875), Fig. 3
Notes: Melomastia was introduced by Saccardo (1875) and typified by M. mastoidea (Fr.) J. Schröt. Melomastia was classified under Ascomycota genera incertae sedis due to the lack of molecular data, and it was difficult to determine the placement of Melomastia based on only morphology (Maharachchikumbura et al. 2015). Later, Norphanphoun et al. (2017) assigned Melomastia in Pleurotremataceae based on molecular data (LSU and SSU). Subsequently, Li et al. (2022) synonymized Dyfrolomyces under Melomastia based on molecular phylogeny and morphology and transferred 11 Dyfrolomyces species to Melomastia. However, Kularathnage et al. (2023) reinstated Dyfrolomyces to accommodate two species (D. tiomanensis (K.L. Pang, Alias, K.D. Hyde, Suetrong & E.B.G. Jones) W.L. Li, Maharachch. & Jian K. Liu (type) and D. chromolaenae (Mapook & K.D. Hyde) W.L. Li, Maharachch. & Jian K. Liu). Currently, 66 epithets of Melomastia are listed in Index Fungorum (http://www.indexfungorum.org/Names/Names.asp, accessed on 614 September 2021, May 2024). Melomastia is characterized by ascomata solitary, coriaceous to carbonaceous, with conical, periphysate papilla; septate pseudoparaphyses; asci bitunicate, cylindrical, short pedicel; ascospores overlapping uni-seriate, hyaline, ellipsoid to fusiform, 1–10-septate, mucilaginous sheath with or without; and asexual morph unknown (Li et al. 2022). Members of this genus are saprobic on branches, twigs, and culms of decaying wood, distributed in terrestrial, freshwater, and mangrove habitat (Li et al. 2022).
Melomastia aquilariae T.Y. Du & Karun. sp. nov.
Index Fungorum number: IF 902121; Facesoffungi number: FoF 15849; Fig. 4
[See PDF for image]
Fig. 4
Melomastia aquilariae (HKAS 126527 holotype). a, b Appearance of ascoma on the host. c Vertical sections through the ascoma. d Ostiole with periphyses. e–h Asci (h ascus stained with Melzer’s reagent). i Pseudoparaphyses. j Peridium. k–n Ascospores. o Germinated ascospore. p, q Colony on PDA surface and reverse view. Scale bars: c = 300 µm, d = 50 µm, e–h = 50 µm, i = 2 µm, j–o = 20 µm
Etymology: named after the host genus, Aquilaria.
Holotypy: HKAS 126527
Saprobic on dead stems of Aquilaria sinensis. Sexual morph: Ascomata (excluding neck) 350–700 µm high × 300–450 µm diam. (x̄ = 524 × 377 µm, n = 10), solitary, semi-immersed to immersed, visible on the host surface as dark, raised spots, dark brown to black, uniloculate, globose to subglobose (wide at the base), carbonaceous. Ostiolar canal 100–150 µm high, black, conical, carbonaceous, papillate, with periphyses. Peridium 35–150 µm wide (x̄ = 85 µm, n = 10), comprising dense, brown to dark brown cells of textura angularis to textura prismatica, fusion with host tissue. Hamathecium 1–3 µm wide, comprising numerous, hyaline, branched, septate pseudoparaphyses, longer than asci, attached at the base and between the asci. Asci 145–220 × 7.5–9 µm (x̄ = 186 × 8.5 µm, n = 30), bitunicate, 8-spored, cylindrical, cylindrical pedicellate 7–14 µm long, rounded in apex, J- apical ring. Ascospores 21.5–28 × 7–8 µm (x̄ = 25 × 7.5 µm, n = 30), 3-septate, overlapping-uniseriate, hyaline, fusiform with acute ends, slightly constricted at the septum, smooth-walled, with a large guttule in each cell when mature, not surrounded by a mucilaginous sheath. Asexual morph: Not observed.
Culture characteristics: Colonies on PDA reaching 6 cm diam., after one month at 28℃; grey, soft, irregular shape, middle protrusion, filiform margin; pale yellow to dark grey, smooth in reverse (Fig. 4).
[See PDF for image]
Fig. 5
Phylogram generated from maximum likelihood analysis based on combined LSU, SSU and ITS sequence datasets representing Kirschsteiniotheliales and other related families. The updated sequence dataset was derived from Sun et al. (2021). Forty-one strains are included in the combined analyses which comprise 2175 characters including gaps (814 characters for LSU, 874 characters for SSU, and 487 characters for ITS). Pseudorobillarda phragmitis (CBS:398.61) and P. eucalypti (MFLUCC 12–0422) were selected as the outgroup taxa. Phylogenetic trees generated from maximum likelihood and Bayesian inference analyses were similar in overall topologies. The best scoring RAxML tree with a final likelihood value of − 16,923.208917 is presented. The matrix had 1079 distinct alignment patterns, with 28.13% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.235516, C = 0.242244, G = 0.301606, T = 0.220634; substitution rates AC = 0.932755, AG = 2.403934, AT = 0.957895, CG = 1.206625, CT = 5.593191, GT = 1.000000; Tree-Length = 2.712710; gamma distribution shape parameter α = 0.375105. Bayesian posterior probabilities (BYPP) from MCMC were evaluated with a final average standard deviation of split frequencies less than 0.01. Bootstrap support values for maximum likelihood (ML) equal to or greater than 70% and Bayesian posterior probabilities (BYPP) equal to or greater than 0.97 are defined above the nodes as ML/BYPP. The type strains are indicated in bold and newly generated sequence are shown in blue
Specimen examined: China, Yunnan Province, Xishuangbanna, Jinghong City, Naban River Nature Reserve, 22° 7′ 50″ N, 100° 40′ 29″ E, on dead stems of Aquilaria sinensis (Thymelaeaceae), 14 September 2021, T.Y, YNA52, (HKAS 126527, holotype), ex-type ZHKUCC:23-0073, ex-isotype ZHKUCC:23-0088.
GenBank numbers: ZHKUCC:23-0073: LSU = OR807856; SSU = OR807854; tef1-α = OR832867; ZHKUCC:23-0088: LSU = OR807857; SSU = OR807855; tef1-α = OR832868.
Notes: Based on the results of BLAST analysis in NCBI GenBank, in LSU, Melomastia aquilariae gave 99.65% closest match to M. oleae (CGMCC 3.20619), 99.90% for SSU with the closest match M. fusispora (CGMCC 3.20618) and 98.86% for tef1-α with the closest matches M. winteri (CGMCC 3.20621). In the phylogenetic analyses of this study, M. aquilariae formed a sister branch with M. winteri (CGMCC 3.20621) with 97% ML bootstrap support (Fig. 3). However, M. aquilariae is different from M. winteri in having solitary, globose to subglobose ascomata, branched pseudoparaphyses, and colonies grey, soft, irregular shape in PDA; while M. winteri has solitary, gregarious, globose ascomata, unbranched pseudoparaphyses, and colonies white, dense, circular in PDA (Li et al. 2022). In addition, M. aquilariae can be distinguished from M. winteri by its wider ascomata (524 × 377 µm vs. 352 × 387 µm), wider range of peridium (35–150 µm vs. 55–62.5 µm), larger asci (186 × 8.5 µm vs. 177 × 7.5 µm) and longer cylindrical pedicellate (7–14 µm vs. 4.8–6.5 µm) (Li et al. 2022). Furthermore, a comparison of tef1-α nucleotides between M. aquilariae and M. winteri (CGMCC 3.20621) resulted in 1.18% differences (10/845 bp, without gaps). In this study, M. aquilariae was collected from Aquilaria sinensis in China, while M. winteri was introduced from Olea europaea in China. Therefore, we introduce our collection as a new species based on both morphological study and phylogenetic analyses.
Kirschsteiniotheliales Hern. -Restr., R.F. Castañeda, Gené & Crous, in Hernández-Restrepo et al., Studies in Mycology. 86: 72 (2017)
Notes: Kirschsteiniotheliales accommodate Kirschsteiniotheliaceae and one genus incertae sedis (Wijayawardene et al. 2022).
Kirschsteiniotheliaceae Boonmee & K.D. Hyde in Boonmee et al., Mycologia 104: 705 (2012).
Notes: Boonmee et al. (2012) introduced Kirschsteiniotheliaceae, to accommodate K. aethiops allied taxa based on morphology and phylogeny. Later Hernandez-Restrepo et al. (2017) accepted monotypic order Kirschsteiniotheliales for Kirschsteiniotheliaceae in Dothideomycetes.
Kirschsteiniothelia D. Hawksw., Bot. J. Linn. Soc. 91: 182 (1985)
Notes: Kirschsteiniothelia was introduced by Hawksworth (1985). These species have been reported from tropical or subtropical regions in terrestrial and freshwater habitats (Bao et al. 2018; Sun et al. 2021; Hyde et al. 2023b). For the taxonomic treatments of this genus (Fig. 5), we followed Sun et al. (2021).
Kirschsteiniothelia dujuanhuensis H.W. Shen & Z.L. Luo, sp. nov.
Index Fungorum number: IF557949; Facesoffungi number: FoF 09289; Fig. 6
[See PDF for image]
Fig. 6
Kirschsteiniothelia dujuanhuensis (KUN-HKAS 129177, holotype) a, b Colonies on natural substrates. c–e, i Conidiophores with conidia. f, g Conidiogenous cells with conidia and sheath. h Conidium with sheath. j Germinated conidium. k, l Culture, k from above, l from reverse. Scale bars: c–e, h, i = 30 µm, f, g = 20 µm, j = 40 µm
Etymology: Name reflects the location Dujuanhu Lake, Yunnan Province, China, from where the holotype was collected.
Holotype: KUN-HKAS 129177
Saprobic on unknown submerged wood in a freshwater lake. Sexual morph not observed. Asexual morph Hyphomycetous. Colonies scattered, effuse, brown to dark brown, glistening, hairy on natural substrate. Conidiophores 29–74(–119) × 9–11 μm (x̄ = 51 × 10 µm, n = 15), macronematous, mononematous, dark brown to dark, solitary, erect, cylindrical, straight or slightly curved, slightly narrower towards the apex, unbranched, septate. Conidiogenous cells 8–10 × 7–8 μm (x̄ = 9 × 7 µm, n = 15), monoblastic, cylindrical, integrated, terminal, brown to dark brown. Conidia (114–)122–155(–170) × 10–13(–16) μm (x̄ = 138 × 12 µm, SD = 17 × 2, n = 15), brown to dark brown, obclavate, subcylindrical, solitary, straight, acrogenous, 6–15-distoseptate, constricted at septum, truncate at base, tapering towards apex, smooth-walled, with a spherical transparent mucilaginous sheath at the apex.
Culture characteristics: Conidia germinated on PDA within 24 h from single-spore isolation, germ tubes from both ends. Colony on PDA reaching 2.6 cm after 10 days at room temperature in dark, circular, brown on the surface rough, with dense, velvety mycelium, brown to dark brown from below.
Material examined: China, Yunnan Province, Puer City, Jingdong Yi Autonomous County, Dujuanhu Lake, 24° 32.53ʹ N, 101° 1.63ʹ E, on submerged decaying wood, 25 February 2022, H.W. Shen, YJ 30-51-1(KUN-KHAS 129177, holotype), ex-type KUNCC 22 12 671.
GenBank numbers: ITS = OQ874971, LSU = OQ732682, SSU = OQ875039.
Notes: Phylogenetic analysis combining LSU, SSU and ITS sequence data shows that Kirschsteiniothelia dujuanhuensis clustered as a sister clade with K. bulbosapicalis (GZCC 23-0732, unpublished data) with 100% ML and 1.00 PP (Fig. 5). Morphologically, K. dujuanhuensis differs from K. bulbosapicalis in having longer conidia (122–155 μm vs. 58.5 –128 μm) but shorter conidiophores (29–74 μm vs. 118–236.5 μm). Comparison of the ITS, LSU and SSU sequence data between K. dujuanhuensis and K. bulbosapicalis revealed 8.9% (55/508 bp, including 10 gaps), 1.2% (13/810 bp, including 3 gaps), and 0.2% (2/1003 bp) base differences, respectively. Here in we introduce K. dujuanhuensis as a new species, based on morphological and phylogenetic evidence.
[See PDF for image]
Fig. 7
Phylogram generated from maximum likelihood analysis based on combined ITS, LSU and SSU sequence data. Related sequences were obtained from Kandawatte et al. (2021). Seventy-eight strains are included in the combined sequence analysis, which comprises 2483 characters with gaps. Lophium mytilinum (AFTOL-ID 1609) and Mytilinidion rhenanum (EB 0341) were used as the outgroup taxa. The tree topology of the ML analysis was similar to the BYPP. The best-scoring RAxML tree with a final likelihood value of − 21,247.317842 is presented. The matrix had 1250 distinct alignment patterns, with 31.45% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.243361, C = 0.241210, G = 0.293208, T = 0.222222; substitution rates AC = 1.761418, AG = 3.375009, AT = 1.973830, CG = 1.330507, CT = 7.087473, GT = 1.000000; gamma distribution shape parameter α = 0.248188. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given above the nodes. Newly generated sequences are in blue and type strains are in bold
[See PDF for image]
Fig. 8
Muyocopron yunnanensis (MHZU 22-0124, holotype). a, b Appearance of ascomata on host. c, d Section of ascomata. e Section of peridium. f Hamathecium. g–i Asci. j–m Ascospores. n Germinated ascospore. o, p Culture on PDA from above and reverse. Scale bars: c, d = 30 μm, e, f, j–n = 10 μm, g–i = 20 μm
Muyocopronales Mapook, Boonmee & K.D. Hyde in Mapook et al., Phytotaxa 265: 230 (2016)
Notes: Muyocopronales was introduced by Mapook et al. (2016) to accommodate the Muyocopronaceae as the type family. To date, Muyocopronales comprises only two families, Muyocopronaceae and Palawaniaceae (Mapook et al. 2020a, b; Tennakoon et al. 2021).
Muyocopronaceae K.D. Hyde in Hyde et al., Fungal Diversity 63: 164 (2013)
Notes: Muyocopronaceae was introduced by Luttrell (1951) with Muyocopron as the type genus and later it was found illegitimate since they introduced it without a Latin diagnosis (Spegazzini 1881). Eriksson (1981) placed this family in the order Hemisphaeriales, and later Hyde et al. (2013) formally introduced Muyocopronaceae as a distinct family incertae sedis in Dothideomycetes and with the monotypic genus Muyocopron. However, Mapook et al. (2016) provided new collections of Muyocopron with LSU and SSU sequence data to clarify its taxonomic placement. Ten genera are currently accepted in Muyocopronaceae (Wijayawardene et al. 2022).
Muyocopron Speg., Anal. Soc. cient. argent. 12(3): 113 (1881)
Notes: Muyocopron was erected by Spegazzini (1881) with M. corrientinum as the type species. There are 79 epithets (70 species) listed in Index Fungorum (2024) which are distributed worldwide and Muyocopron species mostly are saprobic and a few endophytic or pathogenic fungi on a wide variety of plant substrates (Mapook et al. 2016, Tibpromma et al. 2016; Hernández-Restrepo et al. 2019; Tennakoon et al. 2021). Since 2020, a total of six new species have been introduced in Muyocopron, and they are from China or Thailand (Hongsanan et al. 2020a, b, c; Mapook et al. 2020a, b; Chethana et al. 2021a, b; Tennakoon et al. 2021). The genus sexual morph is characterized by superficial, black spots, sub carbonaceous, central ostiolate ascomata, cylindrical to filiform, septate pseudoparaphyses, bitunicate, 4–8-spored asci, and oval to obovoid, hyaline ascospores (Spegazzini 1881; Mapook et al. 2016; Tibpromma et al. 2016; Tennakoon et al. 2021). Asexual morph is characterized by sporodochium-like and irregular conidiomata, enteroblastic, monophialidic, ellipsoidal to ampulliform conidiogenous cells, fusiform or fusoid-ellipsoid, hyaline conidia (Hernández-Restrepo et al. 2019). There are eight species of Muyocopron have been found in China viz. M. garethjonesii and M. lithocarpi (Yunnan Province) (Tibpromma et al. 2016; Phookamsak et al. 2019); M. cinnamomi, M. celtidis, M. dipterocarpi, M. fcinum and M. taiwanense (Taiwan Province) (Hernández-Restrepo et al. 2019; Chethana et al. 2021a, b; Tennakoon et al. 2021); Muyocopron laterale (Anhui Province) (Liang et al. (2020)). In this study, we found a new species, M. yunnanensis, on unidentified dead wood in Yunnan Province, China (Fig. 7).
[See PDF for image]
Fig. 9
Phylogram generated from maximum likelihood analysis (RAxML) of Verruconis and its allied genus in the Sympoventuriaceae based on the combined SSU, LSU and ITS sequence data. Fifty taxa are included in the combined analyses, which comprise 3,275 characters with gaps. Pseudosigmoidea excentrica (CBS 469.95) and Sympoventuria capensis (CBS 120136) were chosen as the outgroup. The best scoring RAxML tree with a final likelihood value of − 26,937.672831 is presented. The matrix had 1282 distinct alignment patterns, with 24.62% of undetermined characters or gaps. The proportion of invariable sites was 0.488513. Estimated base frequencies were as follows: A = 0.240109, C = 0.231110, G = 0.302179, T = 0.226602; substitution rates: AC = 0.814969, AG = 1.615578, AT = 0.983656, CG = 0.991532, CT = 3.660408, GT = 1.00000; gamma distribution shape parameter α = 0.411956. Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.95 are given near the nodes. T = ex-type strain. The newly generated sequences are indicated in blue bold
Muyocopron yunnanensis L. Lu, K.D. Hyde & Tibpromma, sp. nov.
Index Fungorum number: IF 900169; Facesoffungi number: FoF 13910; Fig. 8
Etymology: The specific epithet refers to the location, Yunnan, from where the holotype was collected.
Holotype: MHZU 22-0124
Saprobic on dead wood of unknown host. Sexual morphAscomata 60–90 µm high × 180–400 µm wide (x̄ = 75 × 255 µm, n = 10), superficial, solitary or scattered, conspicuous at the surface, appearing as irregular or circular, flattened, yellowish-green and surrounded by one black bands spots, covering the host, without a subiculum, short ostiole. Peridium 17–22 µm wide, thick-walled; outer layer comprising brown to black, pseudoparenchymatous cells, inner layer comprising light brown cells of textura angularis. Hamathecium comprises numerous, dense, 1.5–2 µm wide, filiform, filamentous, branched, septate pseudoparaphyses. Asci 45–80 × 25–35 μm (x̄ = 56 × 29 μm, n = 20), 8-spored, bitunicate, obovoid or pyriform, hyaline, with ocular chamber when young. Ascospores 18–22 × 10–12 µm (x̄ = 19 × 11 µm, n = 20), overlapping 1–2 seriate or crowded, oval to ellipsoid, or ovoid, hyaline, aseptate, smooth-walled, with oil guttules when immature. Asexual morph Not observed.
Culture characteristics: Ascospores germinated within 12 h on PDA. Colonies on PDA, 35 mm diam. after two weeks, colonies medium dense, circular, flatted, entire edge, colony from above, white to cream; from below, yellowish at margin, light brown at centre.
Material examined: China, Yunnan Province, Jinghong City, on an unidentified dead wood, 14 September 2021, Li Lu, JH6 (MHZU 22-0124, holotype), ex-type ZHKUCC 22-0216, ZHKUCC 22-0217.
GenBank numbers: ITS: OQ064500; LSU: OQ064502; SSU: OQ064504 (ZHKUCC 22-0216), ITS: OQ064501; LSU: OQ064503; SSU: OQ064505 (ZHKUCC 22-0217).
Notes: Muyocopron yunnanensis is morphologically similar to characteristics of Muyocopron species, in having superficial, circular, flattened ascomata, bitunicate, 8-spored asci, and oval to obovoid, hyaline ascospores (Spegazzini 1881; Mapook et al. 2016). In multigene phylogeny, M. yunnanensis formed a distinct clade in Muyocopron with 0.90 BYPP values (Fig. 7). Based on the blast results of sequence, ITS 91.7% was similar to M. laterale (MK487741), LSU 98.5% was similar to M. coloratum (MK487710), but M. laterale and M. coloratum were reported as asexual from leaf spot of Alloteropsis semialata and Cattleya sp. in Australia respectively, thus, it is unable to compare morphology with M. yunnanensis (Hernández-Restrepo et al. 2019), while SSU was similar to M. dipterocarpi up to 99% (KU726969). Muyocopron yunnanensis and M. dipterocarpi have similar shapes and sizes of ascospore (hyaline, oval to ellipsoid, or ovoid, 18–22 × 10–12 µm vs. hyaline, oval to obovoid, (12–)15–18 × (6–)7–9(–11), but they differ in that M. yunnanensis having yellowish-green and surrounded by one black bands spots on the substrate and obovoid or pyriform, without pedicellate asci, while M. dipterocarpi is brown to dark brown spots on the substrate and saccate or broadly obpyriform and pedicellate asci. Therefore, we introduce M. yunnanensis as a new species in Muyocopron.
Venturiales Y. Zhang., C.L. Schoch & K.D. Hyde, in Zhang et al., Fungal Diversity 51(1): 251 (2011)
Notes: Venturiales species are widely distributed as saprobes and plant and animal pathogens (Shen et al. 2020) and the families belonging to this order are Cylindrosympodiaceae, Sympoventuriaceae and Venturiaceae are accepted in this order with two incertae sedis genera (Wijayawardene et al. 2022).
Sympoventuriaceae Y. Zhang ter, C.L. Schoch & K.D. Hyde, in Zhang et al., Fungal Diversity 51(1): 251 (2011)
Notes: Sympoventuriaceae comprises saprophytes, endophytes, plant pathogens, and animal or human opportunistic pathogens. They have a wide geographical distribution and can be found in diverse ecologies. For the taxonomic treatment of this family, Wijayawardene et al. (2022) is followed Fig. 9.
Verruconis Samerp., H.J. Choi, van den Ende, Horré & de Hoog, in Samerpitak et al., Fungal Diversity 65: 117 (2013) [2014]
Notes: Verruconis species are well-known neurotropic opportunistic fungal pathogens (Seyedmousavi et al. 2014). Verruconis gallopava (= Ochroconis gallopava) infects patients with severe immunodeficiency (Murata et al. 2022).
Verruconis pakchongensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov.
Index Fungorum number: IF 900202; Facesoffungi number: FoF 15103; Fig. 10
[See PDF for image]
Fig. 10
Verruconis pakchongensis (BBH 49597, holotype). a, b Colonies with conidia on conidiophore of unidentified fungus. c–f Upper and reverse views of culture on PDA (the top) and MEA (the bottom), after 14 days at 25 °C. g–k, m Conidiophores, conidiogenous cells and conidia on PDA, arrow indicates a gelatinous brown sheath. l Chlamydospore. Scale bars: a = 20 μm, b = 5 μm, g–m = 10 μm
Etymology: The specific epithet “pakchongensis” refers to the Pak Chong District, Nakhon Ratchasima Province, Thailand where the holotype was collected.
Holotype: BBH 49597
Saprobic or opportunistic pathogens on other fungi. Sexual morph: Not observed. Asexual morph: Hyphomycetous. Colonies forming on conidiophores of unidentified fungus, caespitose, effuse, loose to dense, hairy, brown. Mycelium 1.5–2.8 μm diam, superficial, composed of branched, septate, smooth-walled, pale brown hyphae, occasionally forming anastomosis hyphae. Conidiophores 21.5–34.2 × 2.1–2.7 μm (x̄ = 28.4 × 2.5 μm, n = 20), arising laterally from hyphae, mononematous, solitary to loosely aggregated, simple or rarely branched, subcylindrical, straight or slightly flexuous, 2–4-septate, smooth-walled, pale brown. Conidiogenous cells up to 13.2 × 1.8–2.7 μm (x̄ = 10 × 2.4 μm, n = 20), holoblastic, polyblastic, integrated, terminal, with sympodial extensions, subcylindrical, denticulate, pale brown, with 1–3 cylindrical denticles at the apical region, 0.6–2.2 × 0.7–1.3 μm. Conidia 6.8–8.9 × 4–4.7 μm (x̄ = 8 × 4.3 μm, n = 10), acropleurogenous, solitary, broadly ellipsoidal with rounded ends, 1-septate, deeply constricted at the septum, rough-walled, greyish brown. Conidial secession rhezolytic.
Culture characteristics: Colonies after 14 days at 25 °C: On MEA reaching 22–26 mm diam., cottony, brown, round, margins mostly entire, soluble pigment absent, exudates absent, reverse brown. On PDA reaching 21–26 mm diam., velvety, greyish brown, round, margins mostly entire, soluble pigment yellowish brown, exudates absent, reverse brown. Vegetative hyphae on PDA medium 1.5–4 μm diam., composed of branched, septate, smooth or rough-walled, pale brown. Conidiophores 5.2–35 × 1.8–2.8 μm (x̄ = 13 × 2.3 μm, n = 20), arising laterally from vegetative hyphae, mononematous, solitary, simple or occasionally branched, subcylindrical, straight or flexuous, 0–3-septate, smooth-walled, pale brown, or reduced to a single conidiogenous cell that arise from assimilative hyphae. Conidiogenous cells up to 12.5 × 1–3.7 μm (x̄ = 8.3 × 2.3 μm, n = 15), holoblastic, mostly monoblastic, integrated, terminal, with sympodial extensions, or lateral, subcylindrical or ampulliform, denticulate, pale brown, with 1–3 cylindrical denticles at the apical region, 1–1.2 × 0.5–0.7 μm. Conidia 6.2–9 × 3.6–6.1 μm (x̄ = 7.8 × 4.6 μm, n = 30), acrogenous or pleurogenous, solitary, broadly ellipsoidal, broadly obovoid to subglobose, rounded ends, 0–1-septate, often constricted at the septum, rough-walled, greyish brown, sometimes with a gelatinous brown sheath. Conidial secession rhezolytic. Chlamydospores 10.7–14 μm, terminal or intercalary, solitary, subglobose, aseptate, smooth-walled, brown.
Material examined: Thailand, Nakhon Ratchasima Province, Pak Chong District, Khao Yai National Park, Pha Diao Dai Nature Trail, isolated from conidiophores of unidentified fungus, on decaying wood of an unidentified plant in a small pond, 27 July 2011, C. Chuaseeharonnachai, isolate FF00304 (BBH 49597, holotype), ex-type TBRC-BCC 51642.
GenBank numbers: TBRC-BCC 51642: act = OQ116769, ITS = OQ121928, LSU OQ121946, rpb2 = OQ116751, SSU: OQ121937, tef1-α = OQ116760, tub2 = OQ116768.
Notes: Our isolate (TBRC-BCC 51642) belongs to Verruconis and is phylogenetically close to the clade that includes V. cylindricalis and V. guizhouensis, with 90% ML bootstrap and 1.00 BYPP value (Fig. 9). They are morphologically similar in that they possess integrated and terminal or sympodial conidiogenous cells on solitary conidiophores which produce 1-septate and rough-walled conidia (Shao et al. 2022; Wei et al. 2022a). However, V. pakchongensis is distinguished from these two species mainly by the presence of unicellular conidia and, sometimes, brown gelatinous sheaths on its conidia. Morphologically, V. pakchongensis has more variable in conidial shape with broadly ellipsoidal and broadly obovoid to subglobose (Fig. 10), while the conidia of V. cylindricalis and V. guizhouensis are regular in broadly ellipsoidal and ellipsoidal (Shao et al. 2022; Wei et al. 2022a), respectively. Furthermore, V. pakchongensis differs from V. cylindricalis and V. guizhouensis in producing terminal or intercalary, subglobose, aseptate, and smooth-walled chlamydospores. Chlamydospores of V. guizhouensis are intercalary, ellipsoidal, 1-septate, and verrucose (Shao et al. 2022), whereas those of V. cylindricalis are intercalary, cylindrical or clavate, 2-septate, and verruculose (Wei et al. 2022a). In addition, reverse colonies of V. cylindricalis produce blood-colored diffusible pigments on both MEA and PDA media (Wei et al. 2022a), whereas V. pakchongensis produces yellowish brown on PDA and V. guizhouensis produces brown diffusible pigments on MEA (Shao et al. 2022), based on observations after 14 days of incubation at 25–26 °C. Both morphology and DNA sequence data support the distinction of V. pakchongensis as the novel taxon in this study.
Subclass Dothideomycetidae P.M. Kirk, P.F. Cannon, J.C. David & Stalpers ex C.L. Schoch, Spatafora, Crous & Shoemaker, in Schoch et al., Mycologia 98(6): 1045 (2007) [2006]
Notes: For the taxonomic treatment of Dothideomycetes, we follow Hongsanan et al. (2020a) and Wijayawardene et al. (2022).
Dothideales Lindau, in Engler & Prantl, Nat. Pflanzenfam., Teil. I (Leipzig) 1(1): 373 (1897)
Dothideaceae Chevall. [as ‘Dothideae’], Fl. gén. env. Paris (Paris) 1: 446 (1826).
Notes: Dothideaceae species are characterized by ascostromata which are immersed to erumpent or superficial, uniloculate to multiloculate and without an ostioles. They are eight or poly-spored with bitunicate asci. Ascospores and hyaline or brown, transversely septate, or muriform, and often guttulate ascospores (Thambugala et al. 2014).
Dothiora Fr., Summa veg. Scand., Sectio Post. (Stockholm): 418 (1849).
Notes: Dothiora is usually associated with dead branches of wood, although some species were found on leaf litter (Crous and Groenewald 2017). Here, we reported the occurrence on dead and living leaves (the latter as endophytic) and from a leaf-cutting ant fungus garden. In culture, the morphology is similar to several genera in Dothideaceae producing dothichiza-like asexual morphs (Crous and Groenewald 2017), and hormonema-like synasexual morphs (Thambugala et al. 2014; Crous and Groenewald 2016; Humphries et al. 2017). The asexual morph in Dothiora is represented by its typical globose, central ostiolate, unilocular pycnidia with textura angularis walls, conidiophores reduced to ampulliform to dolliform, hyaline phialidic conidiogenous cells, and cylindrical or ellipsoid, obpyriform, aseptate, smooth, hyaline conidia (Gao et al. 2021) (Fig. 11).
[See PDF for image]
Fig. 11
Phylogram generated based on combined ITS and LSU data of Dothiora eucalypti and allied taxa in Dothideales. The tree is rooted on Myriangium hispanicum CBS 247.33 and M. haraearum CBS 300.34. The newly generated sequences re indicated in blue bold. The alignment consisted of 69 sequences of Dothideomycetes, which had 1,424 characters, with 621 distinct patterns, 382 parsimony-informative, 109 singleton sites and 932 constant sites. The models for ML analysis were GTR + I + G for LSU and SYM + I + G for ITS. Bayesian inference was implemented with the same models as ML analysis. The best score tree with a final likelihood value of − 10,496.107 is presented. The tree was inferred with Bayesian Inference (BI) and Maximum Likelihood (ML) algorithms. Only posterior probabilities ≥ 0.90 and bootstrap values ≥ 70% are shown
Dothiora eucalypti Lacerda, Gusmão, G.G. Barreto & A. Rodrigues, sp. nov.
Index Fungorum number: IF 902386; Facesoffungi number: FoF 13860; Fig. 12
[See PDF for image]
Fig. 12
Dothiora eucalypti (HUEFS263169 holotype). a, b Aggregate and solitary pycnidia in culture with detail of the protuberances. c, d Longitudinal sections of pycnidia. e Detail of conidiogenous cells. f Conidia from pycnidia. g Hormonema-like hyphae with conidiogenous cells and conidia. Scale bars: a = 100 μm, b = 50 μm, c–d = 20 μm, e–g = 10 μm
Etymology: In reference to the host Eucalyptus microcorys.
Holotype: HUEFS263169
Associated with leaves of Eucalyptus microcorys. Sexual morph: Not observed. Asexual morph: Coelomycetous. Colonies on PDA grow initially yeast-like, cream-colored, hyaline hyphae, presenting conidia with yeast-like budding. After seven days hyphae become melanized, giving rise to a hormonema-like synasexual morph, thick-walled, branched, smooth, septate, slightly constricted at septa, with integrated conidiogenous cells, producing hyaline, smooth, ellipsoidal to irregular conidia 5–10 × 1.5–3 μm (x̄ = 7.9 × 2.1 μm, n = 30). Chlamydospores-like structures present. Conidiomata 100 µm diam., pycnidial, superficial to semi-immersed, solitary or aggregated, globose to irregular, unilocular, with central papillate ostiole, with or without protuberances at apex, dark brown to black, up to; wall 2–3 layers of brown textura angularis. Conidiophores 7–9 × 2–4 μm (x̄ = 8.1 × 2.9 μm, n = 30), reduced to phialidic conidiogenous cells lining the inner cavity, obpyriform to obclavate, smooth, thin-walled hyaline. Conidia 4–12 × 1–2.8 μm (x̄ = 9.8 × 2.3 μm, n = 30), aseptate, subcylindrical, oblong to irregular, smooth, sometimes with slightly truncate basis, hyaline.
Culture characteristics: Colonies on PDA at 28 °C after 7 days: initially yeast-like, cream-colored, but reaching 8.5 mm in diam., radial, olivaceous-black, black, filamentous, sometimes wrinkled with white borders; on PCA at 28 °C after 7 days: filamentous, black to brown, circular, lacking aerial mycelium, reaching 7 mm in diam.; on CMA at 28 °C after 7 days: flat, radial, olivaceous-black in the center, with sepia borders, reaching 8 mm diam.; on PDA at 15 °C: initially yeast-like, cream-colored, but after seven days filamentous, black to olivaceous-black, irregular to circular, reaching 4 mm diam.; on PCA at 15 °C after seven days: circular, filamentous, olivaceous-black, grey to black, olivaceous, with reddish borders, reddish soluble pigment release in the culture medium, reaching 4 mm diam.; on CMA at 15 °C after seven days: flat, circular to radial, black to olivaceous-black, with sepia borders, reaching 4 mm diam. No growth at 45 °C in any culture medium tested.
Material examined: Brazil. São Paulo, Rio Claro, leaf litter of E. microcorys F. Muell., 12 April 2015. L.T. Lacerda, LTL 15 (HUEFS263169, holotype as dry culture), ex-type CCMB 0728.
Additional material examined: Brazil. Botucatu, São Paulo, fungus garden of the leaf-cutting ant Atta sexdens (Linnaeus, 1758) on eucalyptus plantation, April 2012. J.S. Pereira and A. Rodrigues, (JSP06B5.2); id., Rio Claro, living leaves of E. microcorys, October 2015. L.T. Lacerda, (LTL 144; LTL 5_13); id., leaf litter of E. microcorys, April 2016. L.T. Lacerda, (LTL 6_8).
GenBank numbers: LTL 15 (ITS: MH777080; LSU MH777077); LTL 5_13 (ITS: MH777082; LSU: MH777075); LTL 6_8 (ITS: MH777081; LSU: MH777074); LTL 114 (ITS: MH777079; LSU: MH777076); JSP 06 B5.2 (ITS: KR093869; LSU: MH777078).
Notes: Based on a megablast search of NCBI GenBank nucleotide database, the closest hits using highly similar sequences for ITS are Kabatina thujae R. Schneid. & Arx (GenBank MH858857; Identities = 482/525 (92%), 18 gaps (3%)), K. juniperi (GenBank AY183367; Identities = 482/526 (92%), 19 gaps (3%)) and Dothidea eucalypti Crous (GenBank NR156390; Identities = 479/524 (91%), 17 gaps (3%)). Closest hits using LSU sequences are Dothiora mahoniae (A.W. Ramaley) Crous (GenBank MH874022; Identities = 773/795 (97%), 4 gaps (0%)) Dothiora pistaciae (Quaedvl., Verkley & Crous) Crous (GenBank NG057996; Identities = 771/793 (97%), 2 gaps (0%)) and Endoconidioma carpetanum (Bills, Peláes & Ruibal) Crous (GenBank MF611880; Identities = 770/796 (97%), 5 gaps (0%)). Dothiora eucalypti can easily be distinguished from the closely related species by pycnidia with stronger protuberances at apex and the synasexual morph with smooth hormonema-like hyphae slightly constricted at septa. The phylogenetic analyses support that the five strains are distinct species (Fig. 11). Therefore, D. eucalypti is introduced as a novel species in the family Dothioraceae.
Dothiora eucalypti was found on leaf litter (strains LTL 6_8; LTL 15) and living leaves (strains LTL 144; LTL 5_13) of E. microcorys and in a leaf-cutting ant fungus garden on eucalyptus plantation (JSP06B5.2). Regarding the isolation of D. eucalypti from ant colonies, this event sheds some lights on the ecology of the fungus. Leaf-cutting ant species in the genus Atta have the habit of foraging on fresh leaves to nourish their mutualistic fungus, Leucoagaricus gongylophorus (Möller) Singer (De Fine Licht and Boomsma 2014). Thus, we assume that strain JSP06B5.2 probably was foraged by ant workers and transported to the colony (Fig. 12).
[See PDF for image]
Fig. 13
Phylogram generated from maximum likelihood analysis based on combined large subunit (LSU) and internal transcribed spacers (ITS) for Exosporium. The ML bootstrap values equal to or greater han 75% are indicated at the respective nodes. The newly generated sequence is indicated in red
Mycosphaerellales P.F. Cannon, in Kirk, Cannon, David & Stalpers, Ainsworth & Bisby's Dictionary of the Fungi, Edn 9 (Wallingford) (2001)
Notes: Abollahzadeh et al. (2020) re-validated Mycosphaerellales based on LSU, tef1-α and rpb2 as a separate order in Dothideomycetes. This order includes eight families viz. Cystocoleaceae, Dissoconiaceae, Extremaceae, Mycosphaerellaceae, Neodevriesiaceae, Phaeothecoidiellaceae, Schizothyriaceae and Teratosphaeriaceae (Abdollahzadeh et al. 2020; Wijayawardena et al. 2022).
Mycosphaerellaceae Lindau., in Engler & Prantl, Nat. Pflanzenfam., Teil. I (Leipzig) 1(1): 421 (1897)
Notes: Mycosphaerellaceae comprised a plant pathogenic species on a wide range of hosts (Videira et al. 2017). Videira et al. (2017) accepted 120 genera in Mycosphaerellaceae, and this was re- evaluated by Hongsanan et al. (2020a, b, c) accepted 112 genera in based on molecular data. Following the treatment by Abdollahzadeh et al. (2020) the current outline of fungi (Wijayawardena et al. 2022) 122 genera are accepted. However, further studies are needed to understand the taxonomic status of remaining genera based on more collections and sequence data.
Exosporium Link, Mag. Gesell. naturf. Freunde, Berlin 3(1–2): 9 (1809)
Notes: Exosporium was proposed by Link (1809) based on E. tiliae (from Tilia in Germany) (Ellis 1961). A strain lodged in CBS as E. tiliae (CBS 484.77, CBS H-713, Québec, Canada) clustered in Pleosporales, and was shown to be a Corynespora species in the C. olivacea complex occurring on Tilia. Corynespora olivacea is commonly confused with E. tiliae, but is distinct by having short, 0–2-septate conidiophores with a single apical pore (Ellis 1960a, b). Wijewardena et al. (2022) accepted Exosporium in Mycosphaerellaceae based on molecular data (Fig. 13).
Exosporium livistonae Crous & Summerell, in Crous, Summerell, Shivas, Romberg, Mel'nik, Verkley & Groenewald, Persoonia 27: 145 (2011).
Index Fungorum number: IF489895; Facesoffungi number: FoF 16037; Fig. 14
[See PDF for image]
Fig. 14
Exosporium livistonae (ZHKUCC 24–0793, new host record) a Host photo. b, c Fruiting bodies. d Front and back view of colony at 25 °C on PDA. e, f Germinating spore (100x). h, i, k, l Single spore under 100 magnification. Scale bar: E–J = 10um
Saprophytic from the yellowed and dead leaves of Trachycarpus fortunei. Sexual morph: Not observed. Asexual morph: Colonies on PDA discrete, hairy, brown or white. Mycelium immersed. Stroma usually present, and often very well-developed. Setae and hyphopodia absent. Conidiophores 140–285 × 4–6 µm (x̄ = 180 × 4 μm, n = 50), macronematous, mononematous, often caespitose, straight, or flexuous, unbranched, or very rarely branched, mid to dark brown or olivaceous brown, smooth or verruculose. Conidia 30–70 × 4–6 µm (x̄ = 47 × 5 μm, n = 50), usually solitary, short catenate in one species, a crop leurogenous, simple, mostly obclavate, pale to dark brown or olivaceous brown, smooth, verrucose or echinulate, distoseptate, generally with a thick, dark hilum at the base.
Culture characteristics: colonies on PDA reach 7 cm diam. at 25 ℃ after 5 days. Upper view wrinkled, filamentous, entire margin, flat, cloudy, fluffy for aerial hyphae, become gray black with time, dense for aerial hyphae, reverse becomes black.
Material examined: China, Guangdong Province, Guangzhou City, South China Botanical Garden, yellow leaves of the palm (Trachycarpus fortunei), 10 June 2021, G.Y. Xia, (new host record)—living culture in ZHKUCC 24-0793.
GenBank numbers: ITS = PQ060474, LSU = PQ060475.
Note: In the present study one isolate (ZHKUCC 24-0793) obtained from Trachycarpus fortunei clustered together with Exosporium livistonae (CBS 131313 = CPC 19357) with 98% ML values. The nucleotide difference between the isolate (ZHKUCC 24-0793) in this study and Exosporium livistonae (CBS 131313 = CPC 19357) was 0.01% in LSU (9/780 bases), and 0.04% in ITS (18/430 bases) excluding gaps. Morphologically, the conidia have some differences from Exosporium livistonae (CBS 131313 = CPC 19357) and the colour distribution and morphology of the fruiting body of the Exosporium livistonae (CBS 131313 = CPC 19357). However, we do not consider these slight differences to introduce this as a new species. Therefore, herein we report Exosporium livistonae as a novel host species on Trachycarpus fortunei from China (Fig. 14).
[See PDF for image]
Fig. 15
Phylogram generated from maximum likelihood analysis based on combined ITS, LSU and SSU sequence data comprised 3,732 characters (ITS = 537, LSU = 1337, SSU = 978, tef1-α = 880). The best scoring RAxML tree with a final likelihood value of − 10,550.568 is presented. Estimated base frequencies were as follows: A = 0.250, C = 0.250, G = 0.250, T = 0.250; substitution rates: AC = 1.56440, AG = 1.60521, AT = 1.56440, CG = 1.00000, CT = 4.40412, GT = 1.00000; gamma distribution shape parameter α = 0.681. Bootstrap support values for maximum likelihood (ML) equal to or greater than 75% and clade credibility values greater than 0.95 (the rounding of values to 2 decimal proportions) from Bayesian inference analysis are labelled at each node. Ex-type strains are in bold, while the new isolate is indicated in blue bold. The tree is rooted to Boeremia exigua (CBS 431.74) and B. foveata (CBS 341.67)
Pleosporales Luttrell ex M.E. Barr, Prodr. Cl. Loculoasc. (Amherst): 67 (1987).
Notes: For taxonomic treatment of Pleosporales we follow Hongsanan et al. (2020a) and Wijewardena et al. (2022).
Acrocalymmaceae Crous & Trakun., in Trakunyingcharoen et al., IMA Fungus 5(2): 404 (2014)
Notes: The monotypic family Acrocalymmaceae was introduced to accommodate the type genus Acrocalymma Alcorn & J.A.G. Irwin (Trakunyingcharoen et al. 2014). This family is morphologically diagnosed by globose ascomata with central papilla, narrowly fusoid, pale brown, 1-septate, ascospores with a mucoid sheath, then becoming transversely 3-septate and asexual morph with papillate or rostrate, globose, conidiomata, 0–3-septate conidia with flaring mucoid apical and basal appendages. Acrocalymma species are saprobes or root pathogens and mostly occur in soil (Trakunyingcharoen et al. 2014).
Acrocalymma Alcorn & J.A.G. Irwin, Trans. Br. mycol. Soc. 88(2): 163 (1987)
Notes: Acrocalymma was introduced and typified by A. medicaginis (Alcorn and Irwin 1987). Acrocalymma includes eleven species viz. Acrocalymma ampeli, A. aquaticum, A. bipolare, A. cycadis, A. fici, A. hongheense, A. medicaginis, A. pterocarpi, A. vagum, A. walker, and A. yuxiense. Most Acrocalymma species are saprobes that occur in terrestrial habitats (Hongsanan et al. 2020a, b, c; Mortimer et al. 2021; Tennakoon et al. 2021), while A. aquaticum and A. bipolare are freshwater species (Zhang et al. 2012; Dong et al. 2020). Acrocalymma medicaginis and A. vagum are reported as root pathogens (Alcorn and Irwin 1987). This study reports a new species and new host association of Acrocalymma fici (Fig. 15).
Acrocalymma fici Crous & Trakun., in Trakunyingcharoen et al., IMA Fungus 5(2): 405 (2014)
Index Fungorum number: IF810838; Facesoffungi number: FoF09155; Fig. 16
[See PDF for image]
Fig. 16
Acrocalymma fici (ZHKU 22-0127, new host record). a Appearance of ascomata on the substrate. b Cross section of conidiomata. c Conidiomatal wall. d, e Conidiogenous cells with conidia. f–m Conidia. n Surface view of colony on PDA. o Reverse view of colony on PDA. Scale bars: b = 200 µm, c–m = 20 µm
Saprobic on leaves of grass (Poaceae) plant. Sexual morph: Not observed. Asexual morph: coelomycetous. Conidiomata up to 150–200 µm high, 100–150 µm diam. (x̄ = 180 × 120 µm, n = 10), pycnidial, globose, solitary, scattered, erumpent, dark brown, papillate, ostiolate. Conidiomatal wall 3–5 layers of brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 9–12 × 5–7 µm (x̄ = 10 × 6 µm, n = 10), ampulliform to doliiform, hyaline, smooth, inconspicuous percurrent proliferation visible at apex. Conidia 16–20 × 2–3 µm (x̄ = 19 × 2.5 µm, n = 10), cylindrical with subobtuse apex, acutely tapered at base to a small flattened central scar, hyaline, smooth, guttulate, medianly 1-septate, not constricted at septum, with flaring mucoid apical appendage 3–5 µm diam.
Culture characteristics: Colonies on MEA flat, spreading, with moderate aerial mycelium, and smooth, even margins; surface smoke-grey in centre, pale olivaceous grey in outer region, smoke-grey in reverse.
Material examined: China, Guangdong Province, Guangzhou city, Nansha district, near Bai huitian reservoir, on leaves of grass plant (Poaceae), 18 June 2021, Senanayake I.C., GZ50, (ZHKU 22-0127, new host record), living culture ZHKUCC 22-0222.
Hosts and distribution: on Ficus sp. in India (New Delhi) (Trakunyingcharoen et al. 2014), on Pterocladiella capillacea in Taiwan (China) (Cha et al. 2021), on Calamus castaneus in Malaysia (Azuddin et al. 2021), on decaying wood in Thailand (Boonmee et al. 2021), on leaf of grass in Guangdong, China (This study).
GenBank numbers: ITS: PP892073; LSU: PP892072; SSU: PP892074
Notes: The combined ITS, LSU and SSU sequence analysis (Fig. 15) showed that our isolate (ZHKUCC 22-0222) clustered with the ex-type strain of Acrocalymma fici (CBS 317.76) with 97% ML bootstrap support and 0.95 BYPP value. Our collection resembles A. fici in morphological characters and size of the conidiomata, conidiogenous cells and conidia (Fig. 16). Moreover, there are no base pair differences of the ITS and SSU locus while LSU comprises two nucleotide differences. However, Acrocalymma fici has not been collected from a grass plant (Poaceae) in China. Therefore, we report this collection as a new host record of A. fici.
Acrocalymma estuarinum M.S. Calabon, E.B.G. Jones & K.D. Hyde, sp. nov.
Index Fungorum Number: IF 902387; Facesoffungi number: FoF 15062; Fig. 17
[See PDF for image]
Fig. 17
Acrocalymma estuarinum (MFLU 24-0263, holotype). a Appearance of immersed conidiomata on wood surface. b, c Vertical section of conidioma. (c) Section through the peridium. (d)–(g) Conidiogenous cells. (h)–(l) Conidia with appendages. (m) Germinated conidium. Colony on MEA: (n) obverse. Scale bars: (a) = 200 μm, (b) = 50 μm, (c)–(g) = 10 μm, (h) = 20 μm, (i)–(l) = 5 μm. Scale bars: d, e, g = 50 μm, f = 20 μm, h–j = 30 μm, k–n = 5 μm, o = 10 μm
Etymology: In reference to the estuarine habitat where the fungus was found.
Holotype: MFLU 24-FT1207 24-0263
Saprobic on submerged decaying wood. Sexual morph: Not observed. Asexual morph: Ceolomycetous. Conidiomata 120–185 × 50–75 μm, white, solitary to gregarious, immersed, pycnidial, globose to subglobose, unilocular, glabrous and ostiolate. Ostiole 30–45 μm diam., centrally located. Peridium 8.5–17 μm thick, composed of thick-walled, dark brown cells of textura globulosa in the outer layer, become hyaline cells of textura angularis to globosa in the inner layer. Conidiophores reduced to conidiogenous cells or a supporting cell. Conidiogenous cells 12–15 × 2–3 µm, hyaline, enteroblastic, ampulliform to doliiform, smooth-walled. Conidia 17–22 × 2.5–3.5 (x̅ = 19.2 × 3.1 μm, n = 20), hyaline, cylindrical with sub-obtuse apex and base, straight, aseptate, smooth-walled, guttulate, with flaring mucoid apical appendage (2–4 μm diam.), visible in water mounts.
Culture characteristics: colonies on MEA reaching 30–35 mm diam. at 25 °C after 1 month, flat, dry, rough, margin entity to undulate and irregular; from above, light yellow at the margin, light brown in the middle; from below, light yellow at the margin, brown in the middle; not producing pigmentation in the culture.
Material examined: Thailand, Samut Songkhram, Mueng Samut Songkhram, on submerged decaying wood, 8 September 2020, M.S. Calabon, SS109 (MFLU 24-0263, holotype); ex-type MFLUCC 24-0276.
GenBank accession numbers: ITS: PP886241; LSU: PP886242; tef1-α: PP908504
Notes: The combined SSU, LSU, ITS, and tef1-α phylogenetic tree shows that Acrocalymma estuarinum clustered within Acrocalymma and a sister taxon to A. ampeli (MFLUCC 20-0159; MFLUCC 19-0288). However, Acrocalymma estuarium differs from the latter in having a longer conidiomata (120–185 × 50–75 μm vs. 80–120 × 150–180 µm) and narrower conidia (17–22 × 2.5–3.5 vs. 17–19 × 5.5–6.5 μm) with sub-obtuse apex and base. Acrocalymma estuarinum is the first species of Acrocalymma associated with estuarine environment and the fifth species to be reported in aquatic habitats (Zhang et al. 2012; Dong et al. 2020; Boonmee et al. 2021; Calabon et al. 2022, 2023a, b).
Corynesporascaceae Sivan. Mycol. Res. 100(7): 786 (1996)
Notes: Corynesporascaceae was initially referred to Melanommatales sensu and later was accepted in Pleosporales by Hyde et al. (2013). Corynesporascaceae species are available in the tropics and subtropics and can cause foliar diseases in plants (Stone and Jones 1960; Dixon et al. 2009; Hyde et al. 2013).
Corynespora Güssow, J. Royal Agric. Soc. England 65: 272 (1905) [1904]
Notes: Corynespora is an old asexual genus introduced by Güssow (1906). Currently, more than 200 epithets have been recorded (Index Fungorum 2024). However, there are only 16 species in this genus with DNA sequence data (Fig. 18), some of which are not from type materials. Corynespora species are well-known leaf spot pathogens in the tropics and subtropics (Voglmayr and Jaklitsch 2017). For the taxonomic treatment of this genus, we follow Voglmayr and Jaklitsch (2017) and Liu et al. (2023a, b2017).
[See PDF for image]
Fig. 18
The Phylogram of the best ML tree based on a combined dataset (SSU, ITS, LSU, tef1-α, and rpb2) of Corynespora. The scale bar indicates 0.04 changes. The tree is rooted with Periconia pseudodigitata (CBS 139699) (Periconiaceae, Pleosporales). The best ML tree with a final likelihood value of − 12,581.670 is presented. The alignment dataset comprises 21 taxa with 4478 sites, including 766 distinct patterns, 526 parsimony-informative sites, 429 singleton sites and 3473 constant sites. Ultrafast bootstrap values for the maximum likelihood analysis equal to or greater than 95% and posterior probability for Bayesian analysis equal to or greater than 0.95 are indicated at the nodes. Novel isolates are indicated in red. Isolates from type materials are in bold
Corynespora chengduensis Y.P. Chen & Maharachch., sp. nov.
Index Fungorum number: IF901380; Facesoffungi number: FoF 15113; Fig. 19
[See PDF for image]
Fig. 19
Corynespora chengduensis (HKAS 127168, holotype). a–c Colonies on the natural substrate. d–f Conidiophores. The arrow points to the apical pore. g, h Conidiophore bases and stroma cells. i–k Conidiophores, conidiogenous cells and apical conidia. l–y Conidia. z Germinating conidium on PDA. a1, b1 Front and back views of the colony on PDA (1 week). Scale bars: d, z = 50 μm, g, i, l, r = 20 μm. D applies to e and f. g applies to h. r applies to s–z
Etymology: The name refers to Chengdu, the city where the fungus was collected.
Holotype: HKAS 127168
Saprobic on the decaying wood of an unidentified plant. Sexual morph: not observed. Asexual morph: Hyphomycetous. Colony on natural substrate effuse, dark brown, velvety or spongy, forming widely effused patches. Mycelium partly superficial, immersed in the substrate, composed of branched, septate, subhyaline to dark brown, curved hyphae. Stromata partly superficial, immersed, brown, irregular in shape, pseudoparenchymatous, composed of brown cells. Conidiophores 100–710 μm (x̅ = 350, n = 20) long, 7–10 μm (x̅ = 8, n = 20) wide, arising singly or more often in dense tufts from superficial hyphae or from cells of the stromata, erect or ascending, simple, straight or flexuous, pale brown to dark brown, septate. Conidiogenous cells 15–31 × 8–13 μm (x̅ = 25 × 10, n = 20), monotretic, cylindrical, pale brown to brown, often with proliferation through the apical pore and formation of another conidium at the apex of the proliferation. Conidia 54–219 × 12–17 μm (x̅ = 140 × 14, n = 40), with a 4–8 μm (x̅ = 6, n = 40) wide blackish-brown scar at the base, formed singly or in a short chain through a wide pore at the apex of the conidiophore, often with proliferation through the apical pore and formation of another conidium at the apex of the proliferation, almost cylindrical but usually slightly and gradually tapering towards the rounded apex and more abruptly towards the truncate base, straight or slightly curved, smooth, subhyaline to golden brown, 0–26-distoseptate, with angular lumina; wall up to 5 μm thick (Fig. 20).
[See PDF for image]
Fig. 20
Phylogram generated from maximum likelihood analysis based on a three-locus dataset (LSU, ITS and RPB2) representing Epicoccum indicum (NFCCI 5322) and related species. The scale bar indicates 0.08 changes. The tree is rooted in Neocucurbitaria aquatica (CBS 297.74). Species sampling was based on Hou et al. (2020). One hundred-eleven sequences are included in the analysis which comprises 1850 characters after alignment. ML analysis was performed using a GTR model of site substitution, including GAMMA + P-Invar model of rate heterogeneity and a proportion of invariant sites. Bayesian inference was implemented with the GTR + I + G model. The best RAxML tree with a final likelihood value of − 9265.764120 is presented. Estimated base frequencies were as follows: A = 0.243820, C = 0.240756, G = 0.279920, T = 0.235504; substitution rates: AC = 1.874052, AG = 6.412815, AT = 2.187577, CG = 1.253908, CT = 17.584563, GT = 1.000000; gamma distribution shape parameter α = 0.138623. Bootstrap support values obtained in a complementary Maximum Likelihood analysis (MLBS, right) with RAxML using 1000 pseudoreplicates are provided after the BYPP values (left). Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.95 are given above the nodes. For each terminal, the species name and the voucher/herbarium code are indicated, and new isolate is in blue
Culture characteristics: Colony on PDA 19 mm diam after 1 week in an incubator under dark conditions at 20 °C, white, circular, surface velvety, with dense mycelium and entire margin; reverse creamy white.
Material examined: China: Sichuan Province, Chongzhou City, Baiyungou, 30°47′52″ N, 103°24′19″ E, elevation 990 m, 27 September 2021, on a decaying branch of an unidentified host, Y.P. Chen, BY13 (HKAS 127168, holotype), ex-type culture UESTCC 22.0153.
GenBank numbers: ITS = OR762653, SSU = OR762654, LSU = OR762655, tef1-α = OR771929, RPB2 = OR771930.
Notes: In this study, we isolated and identified a new Corynespora species UESTCC 22.0153, by morphological examinations and multi-locus phylogenetic analysis. The phylogenetic tree shows that the isolate UESTCC 22.0153 forms a separated branch, and C. olivacea (CBS 136917) is the only species close to it. Our species is morphologically different from the C. olivacea by having a longer conidium (54–219 μm vs. 50–105 μm) and more distosepta (0–26 vs. 5–14) (Ellis 1960a, b). In addition, BLASTn analyses of them showed 93% sequence identity (463/500, 6 gaps) for ITS and 98% sequence identity (814/828, 7 gaps) for LSU. Considering the significant differences in morphology and molecular data, we introduce the isolate UESTCC 22.0153 as a new species Corynespora chengduensis.
Didymellaceae Gruyter, Aveskamp & Verkley, Mycol. Res. 113(4): 516 (2009)
Notes: Didymellaceae species are widely distributed as endophytes, saprophytes, and human and animal pathogens (Valenzuela-Lopez et al. 2018; Wanasinghe et al. 2018; Bracale et al. 2020). Among these taxa, there are pathogens on economically important crops leading to a significant decline in yield and quality (Aveskamp et al. 2010; Chen et al. 2015). Forty-four genera and more than 5400 species are accepted in this family (Wijayawardene et al. 2022).
Epicoccum Link, Mag. Gesell. naturf. Freunde, Berlin 7: 32 (1816) [1815] (Fig. 20)
Notes: Epicoccum species are often found in the soil, and phyllosphere, as endophytes and pathogens (Manawasinghe et al. 2020). This genus is characterized by hyphomycetous and coelomycetous synanamorphs (Jayasiri et al. 2017; Thambugala et al. 2017). The hyphomycetous anamorph is characterized by having dark sporodochia with branched conidiophores and mono- to polyblastic, colourless conidiogenous cells that produce coloured, sometimes verruculose, dictyoconidia and the coelomycetous synanamorph is characterized by the formation of conidia in pycnidial conidiomata (Chen et al. 2015). This genus comprises with taxa which are known as pathogens, endophytes and saprobes (Abeywickrama et al. 2023).
Epicoccum indicum S. Rajwar & Raghv. Singh, sp. nov.
Index Fungorum number: IF900238; Facesoffungi number: FoF16038; Fig. 21
[See PDF for image]
Fig. 21
Epicoccum indicum (AMH 10466, holotype). a Host plant in its natural habitat, b Symptoms on inflorescence, c Appearance of superficial hyphae bearing conidiophores and conidia on host surface, d Germination of conidia on PDA, e, f Reverse and front view of colony on PDA after 7 days (NFCCI 5322, ex-type), g–n Superficial hyphae bearing conidiophores and conidia, o–u conidia. Scale Bars: e, f = 10 mm, g–u = 10 μm
Etymology: indicum, refers to India, the country where the fungus was discovered.
Holotype: AMH 10466
Associated with diseased inflorescence of Chrysopogon zizanioides.Sexual morph: Not observed. Asexual morph:Conidiomata or conidiomata-like structure absent. Symptoms develop on inflorescence, brown to dark blackish brown, later on covering the entire inflorescence surface. Colonies brown to dark blackish brown, velvety. Mycelium 2–3.5 µm wide (x̅ = 2.5 μm, n = 20), external, septate, hyaline to light olivaceous brown, sometimes loosely arranged to form prosenchymatous-like structures, 12–20 µm wide (x̅ = 15 μm, n = 20). Stromata absent. Conidiophores 7–42 × 1.5–4.5 µm (x̅ = 19.5 × 3 μm, n = 20), macronematous, mononematous, later on develop as a lateral branch of superficial hyphae, initially unbranched, later on highly branched, claviform, erect to procumbent, hyaline to light olivaceous brown, smooth to slightly rough, thin-walled to thick-walled, 0–5-euseptate. Conidiogenous cell indeterminate. Conidia 10–20 × 7–17 µm (x = 15.5 × 12.5 µm, n = 40), formed singly, dry, light olivaceous brown to mid brown, variable in shape, mostly oval to elliptical, rarely globose to sub globose, sometimes irregular, 1-many celled, constricted at septa, slightly roughened at maturity, thick-walled, 0.5–2 µm (x = 1.2 µm, n = 20).
Culture characteristics: Conidia germinating on PDA within 24 h. Colonies cottony, pinkish white, lower surface off pale brown to blackish brown, form reddish pigments, reaching 20–25 mm diam. in 7 days at 25 ± 5 °C. Mycelia are superficial, effuse and radially striate with regular edges. Sporulation was not observed after 30 days under 25 ± 5 °C.
Material examined: India, Uttar Pradesh, Varanasi, Botanical Garden of Banaras Hindu University, living inflorescence of Chrysopogon zizanioides (L.) Roberty (Poaceae), 10 October 2021, Raghvendra Singh, (AMH 10466, holotype; MH-BHU 66, isotype), ex-type NFCCI 5322.
GenBank numbers: ITS: ON627840, LSU: OP850270, rpb2: OP946257
Notes: Epicoccum andropogonis (Ces.) Schol-Schwarz was described (Schol-Schwarz 1959) on infected grasses with ergot. In 2019, Vu et al. (2019) assigned CBS 195.55 and CBS 193.55 as E. andropogonis which was collected from South Africa. Similarly, Hatami Rad et al. (2019) assigned IRAN 3738C as E. andropogonis collected from Iran, based on ITS-rDNA sequence data that clustered closer to CBS 195.55 and CBS 193.55. Detailed morphological description for IRAN 3738C was provided only from ex-type culture (Hatami et al. 2019). In the phylogenetic analysis of the present study, we observed that CBS 195.55 and CBS 193.55 (assigned for E. andropogonis) are clustered together with Epicoccum indicum (NFCCI 5322) and the morphological characters of NFCCI 5322 are different from E. andropogonis (Ellis 1971). In E. andropogonis well-developed sporodochia bearing densely aggregated conidiophores are found which are lacking in Epicoccum indicum. In Epicoccum indicum conidiophores are mononematous and later on develop as a lateral branch of superficial hyphae, initially unbranched, later on highly branched. Moreover, E. indicum is reported on new host. Since all these strains (CBS 195.55, CBS 193.55 and IRAN 3738C), along with our newly collected strain NFCCI 5322 clustered together with 98% ML bootstrap support and 1.00 BYPP values, we identified these strains as a sister lineage of E. dendrobii (Fig. 20). Since, the microphotographs and dimensions of conidia from the ex-type culture of IRAN 3738C fit into the frame of E. indicum, therefore, it is worthwhile to recombine E. andropogonis assigned for CBS 195.55, CBS 193.55 and IRAN 3738C under E. indicum and we choose Epicoccum indicum as none of these strains not fitting in the morphological frame of E. andropogonis. Epicoccum indicum also differs from closely related E. dendrobii as later form well-developed sporodochia bearing densely aggregated conidiophores and conidia are smaller (11–19 μm) and have basal cells (Chen et al. 2017).
Didymosphaeriaceae Munk, Dansk bot. Ark. 15(no. 2): 128 (1953)
Notes: Munk (1953) introduced Didymosphaeriaceae, and typified by Didymosphaeria, in Pleosporales. Didymosphaeriaceae is characterized by brown, thick-walled, septate ascospores and trabeculate pseudoparaphyses, which anastomose above the asci in a gelatinous matrix (Aptroot 1995; Hyde et al. 2013; Ariyawansa et al. 2014a, b). Didymosphaeriaceae is a large and diverse family of Pleosporales, that currently includes 33 genera (Wijayawardene et al. 2022). Species of the family are found as endophytes, pathogens and saprobes associated with various plant substrates worldwide (Maharachchikumbura et al. 2021).
Dictyoarthrinium S. Hughes, Mycological Papers 48: 29 (1952).
Notes: Dictyoarthrinium was introduced by Hughes (1952) with D. quadratum as the type species and was collected on dead leaves of Borassus antipoem from Ghana. The genus is characterized by mononematous or synnematous conidiophores with integrated conidiogenous cells and conidia of square-to-spherical, subspherical or oblong, pale-to-dark-brown, often 4-celled, and sometimes 16-celled (Hughes 1952). Currently, ten species (Fig. 22) are accepted in Dictyoarthrinium (Ren et al. 2022).
Dictyoarthrinium endophyticum R.M.F. Silva, T.G.L. Oliveira & G.A. Silva, sp. nov.
Index Fungorum number: IF900386; Facesoffungi number: FoF 14102 Fig. 23
[See PDF for image]
Fig. 22
Phylogenetic tree generated from Bayesian inference (BI) analysis based on combined LSU rDNA, ITS and tef1-α sequences. The matrix contained 2346 characters (LSU rDNA = 841, ITS = 599, and tef1-α = 906), inclusive of gaps. The optimal nucleotide substitution model for the BI analysis was GTR + G for LSU rDNA and tef1-α, and HKY + G for ITS. GTR + G + I was used as the ideal model for maximum likelihood analysis. Sequences from type species are indicated in bold. The new species is indicated in blue. Bayesian posterior probabilities (above 0.92) and maximum likelihood bootstrap (above 70%) values are shown near nodes. The tree is rooted with Spegazzinia radermacherae MFLUCC 17–2285 and Spegazzinia tessarthra SH 287
Etymology: In reference to the endophytic lifestyle in eggplant
Holotype: URM 95262.
Endophyte from leaves of Solanum melongena. Sexual morph: Not observed. Asexual morph: Mycelium composed of septate, anastomosing, branched, 3–4 μm in width, pale brown hyphae. Conidiophores 61–111 × 3–5 μm, (x̅ = 71 × 3.4 μm, n = 15) macronematous, basauxic, cylindrical, narrow, straight or flexuous, pale brown, rough and thin-walled terminating in conidiophore mother cells, the transverse septa dark brown with distances of 3.6–8.7 μm, rough-walled. Conidiogenous cells 5–7 × 3.5–5 μm (x̅ = 5.8 × 4.1 μm, n = 10), blastic, integrated, intercalary and terminal, cylindrical, hyaline, Conidia 5.4–12.5 × 6.4–12 μm., (x̅ = 8.3 × 8.5 μm, n = 30) arising from the lateral or apical part of conidiophores, hyaline, subhyaline to pale brown when young and brown to dark brown when mature, verrucose, spherical or subspherical, septate with 4-cells.
Culture characteristics: Colonies after 7 days on PDA growing up to 6 mm diam at 25 °C with surface yellow buff and reverse brownish yellow with white margins.
Material examined: Brazil, Pernambuco State, Chã Grande municipality, isolated as endophyte from leaves of Solanum melongena (Solanaceae), 18 July 2021, R.M.F. Silva (URM 95262, holotype); ex-type URM 8697.
GenBank numbers: ITS: OQ534286; LSU: OQ534288; tef1-α: OQ544569.
Notes: Dictyoarthrinium endophyticum is introduced as a new species based on its distinct morphology and the phylogeny of the combined LSU, ITS, and tef1-α dataset. The phylogenetic tree showed that the sequences obtained from Dictyoarthrinium endophyticum clustered, forming a single well-supported clade, in the genus Dictyoarthrinium close to D. hydei and D. sacchari (Fig. 22). However, D. hydei, described as saprobic on decaying wood submerged in freshwater, produces larger conidia (9–17 × 8–13 μm) and conidiophores (400 × 3–5.5 μm) (Maharachchikumbura et al. 2021), while D. endophyticum produces conidia (5.4–12.5 × 6.4–12 μm) and conidiophores ((30)61–111 × 3–5 μm). Dictyoarthrinium sacchari, found as saprobic on dead leaves of Musa sp., different from D. endophyticum produces acropleurogenic conidia, sometimes square with 2 or 4 cells (Samarakoon et al. 2020a, b). Many Dictyoarthrinium species are saprobes that colonize dead plants (Somrithipol 2007; Leão-Ferreira and Gusmão 2010; Tarda et al. 2019). In the present study, D. endophyticum was described from the healthy leaves of S. melongena.
Dictyosporiaceae Boonmee & K.D. Hyde, in Boonmee et al., Fungal Diversity: (2016)
Notes: Dictyosporiaceae is introduced to accommodate species with cheiroid, digitate, palmate and dictyosporous conidia. The type genus is Dictyosporium. This family was first mentioned, but not formally introduced by Liu et al. (2015a, b) as Dictyosporaceae. Boonmee et al. (2016) established this family and introduced their sexual morphs that form a monophyletic clade in Dothideomycetes.
Dictyosporium Corda, Weitenweber’s Beitr. Nat.: 87 (1836)
Notes: the type species of Dictyosporium is D. elegans. There are 88 species epithets available for Dictyosporium, but most species lack molecular data (Index Fungorum 2024). Dictyosporium species are recorded from wood and other plant matter in terrestrial and aquatic habitats. The asexual morphs of this genus are characterized by micronematous conidiophores and euseptate, complanate conidia with 4–7 rows of cells (Corda 1836). The sexual morphs are characterized by subglobose ascomata, cylindrical asci and hyaline, fusiform, uniseptate ascospores, with or without a sheath (Tibpromma et al. 2018). According to morphological characteristics and molecular evidence, the new isolate found in this study is a new record in China (Fig. 24).
[See PDF for image]
Fig. 23
Dictyoarthrinium endophyticum (URM 95262, holotype). a Colony on PDA after 7 d at 25 °C. b, c Conidiophores. d Conidia. e Beginning of conidial germination. f–g Conidia. Scale bars: = 10 µm
[See PDF for image]
Fig. 24
Maximum likelihood consensus tree inferred from the combined SSU, ITS, LSU, and tef1-α multiple sequence alignments. Bootstrap support values for maximum likelihood (ML, first value) equal to or greater than 80% and Bayesian posterior probabilities from MCMC analyses (BYPP, second value) equal to or greater than 0.95 are given above the nodes. The scale bar indicates expected changes per site. The tree is rooted to Immotthia bambusae (KUN-HKAS 112012C) and Immotthia bambusae (KUN-HKAS 112012D). Ex-type strains are bold. The newly generated sequences are indicated in red
Dictyosporium tratense J. Yang & K.D. Hyde, MycoKeys 36: 96 (2018)
Index Fungorum number: IF 554772; Facesoffungi number: FoF04678; Fig. 25
[See PDF for image]
Fig. 25
Dictyosporium tratense (MHKU 24-043, new record in China). a Colonies on the host. b, d–i Conidia. c Conidia with conidiophores. j conidia germinating on PDA. Scale bars: b–i = 20 μm
Saprobic on submerged decaying wood. Sexual morph: Not observed. Asexual morph: Hyphomycetes. Colonies punctiform, superficial, gregarious, black. Mycelium composed of immersed, pale brown, septate, smooth, thin-walled hyphae. Conidiophores cylindrical, micronematous, mononematous, septate, hyaline to brown, smooth-walled. Conidiogenous cells cylindrical, monoblastic, hyaline to pale brown. Conidia 40–55 × 20–30 μm (x̅ = 47.5 × 25 μm, n = 20), cheiroid, acrogenous, brown to dark brown, complanate, consisting of 30–52 cells arranged in (2–)4–5 (mostly 5) closely compact columns, 5–10-euseptate in each column, guttulate, usually with 2–3 central columns longest and of equal length, 2–3 peripheral columns shorter and of equal length, sometimes with hyaline globose appendages at the apical cells of outer columns.
[See PDF for image]
Fig. 26
Phylogram generated from maximum likelihood analysis based on combined LSU, SSU, ITS and tef1-α sequence data. Related sequences are taken from Xu et al. (2024) and additions according to the BLAST searches in NCBI. Sixty-one strains are included in the combined analyses which comprised 3382 characters (860 characters for LSU, 1033 characters for SSU, 573 characters for ITS, and 916 characters for tef1-α) after alignment. Leptosphaeria doliolum (CBS 505.75) was used as the outgroup taxa. The best-scoring RAxML tree with a final likelihood value of − 22,870.768793 is presented. The matrix had 1295 distinct alignment patterns, with 26.98% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.239877, C = 0.248600, G = 0.271623, T = 0.239900; substitution rates: AC = 1.424996, AG = 2.985437, AT = 1.629409, CG = 1.504009, CT = 7.211381, GT = 1.000000; gamma distribution shape parameter α = 0.189842. Bootstrap support values for ML equal to or greater than 60% are given above the nodes. Bayesian posterior probabilities (BYPP) equal to or greater than 0.90 are given above the nodes. Ex-type strains are in bold and newly generated sequence is in blue bold
Material examined: China, Guangxi Province, Qinzhou City, on submerged wood in a river, 26 February 2023, X.Y. Shu GX5.1 (MHZU 24-0435); living culture ZHKUCC 24-0792.
GenBank numbers: ITS: PP760087, LSU: PP760086, SSU: PP760088, tef1-α: PP764532.
Notes: Phylogenetic analysis showed that our new collection D. tratense (MHZU 24-0435) clustered with the type strain of Dictyosporium tratense (MFLUCC 17-2052) with 97% ML bootstrap and 0.99 BYPP support. Our collection fits well with the morphological characteristics of D. tratense in having cheiroid, acrogenous, solitary, and euseptate conidia consisting of five rows of cells and hyaline globose appendages (Yang et al. 2018). However, the holotype (MFLUCC 17-2052) has hyaline cloud-shaped mucilaginous sheath at the apical cells of outer columns, which is lacking in our collection (Yang et al. 2018). There are two nucleotide differences across ITS and four nucleotide differences in tef1-α between our new strain (MHZU 224-0435) and the type strain (MFLUCC 17-2052). Based on the morphological and phylogenetic analysis, we therefore identified our new strain as D. tratense. Dictyosporium tratense initially was reported in Thailand (Yang et al. 2018), this study expands the known geographical distribution of this species to China.
Macrodiplodiopsidaceae Voglmayr, Jaklitsch & Crous, in Crous et al., IMA Fungus 6(1): 178 (2015)
Notes: Macrodiplodiopsidaceae was introduced by Crous et al. (2015a, b) with Macrodiplodiopsis as the type genus. Two genera; Macrodiplodiopsis and Pseudochaetosphaeronema are accepted in this family (Wijayawardene et al. 2022). In this study, the phylogenetic analyses were following the latest treatment in Xu et al. (2024).
Pseudochaetosphaeronema Punith., Nova Hedwigia 31(1–3): 126 (1979)
Notes: Pseudochaetosphaeronema was introduced by Punithalingam (1979), with Pseudochaetosphaeronema larense as the type species. For Pseudochaetosphaeronema species, both asexual and sexual morphs have been reported (Zhang et al. 2016; Hyde et al. 2020a, b, c; Boonmee et al. 2021; de Silva et al. 2022; Tan et al. 2022; Li et al. 2023a, b; Xu et al. 2024). The asexual morph of Pseudochaetosphaeronema is characterized by dark brown to black, nearly globose, scattered or gregarious, surfcial conidiomata, monophialidic, cylindrical conidiogenous cells, hyaline, subglobose to oval, aseptate conidia and grey colonies on PDA (Jayasiri et al. 2019). And the asexual morph of Pseudochaetosphaeronema is characterized by solitary, scattered, immersed, uni-loculate, black ascomata; textura angularis peridium; cellular, unbranched, septate pseudoparaphyses; 8-spored, bitunicate, fssitunicate, with obovoid, short distinct pedicel with a rounded end, apex rounded with a minute ocular chamber asci and 2–3-seriate, hyaline, fusiform, with pointed ends, 1-septate at the center, constricted at the septa, guttulate, thick and smooth-walled ascospore. Eight species epithets are listed in Index Fungorum (2024). In this study, we introduce a novel species of Pseudochaetosphaeronema collected from Yunnan, China (Fig. 26). The tree topology of our multigene phylogenetic analyses is similar to the latest analysis performed by Xu et al. (2024).
Pseudochaetosphaeronema puerensis R.F. Xu, K.D. Hyde & Tibpromma, sp. nov.
Index Fungorum number: IF901412; Facesoffungi number: FoF 15836 Fig. 27
[See PDF for image]
Fig. 27
Pseudochaetosphaeronema puerensis (ZHKU 22–0153, holotype). a, b Appearance of ascomata on host substrate. c Section of ascoma. d Peridium. e Pseudoparaphyses. f–j Asci. k–q Ascospores. s Germinated ascospore. s Ascospore stained with Indian ink. t, u Colonies on PDA. Scale bars: c = 100 μm, d–f = 20 μm, g–j, s = 30 μm, q–r = 10 μm
Etymology: The name refers to the location from which species was collected, Pu’er, Yunnan, China.
Holotype: MHZU 22-0153
Saprobic on a dead branch of Hevea brasiliensis. Sexual morphAscomata 120–350 μm high, 160–340 diam., (x̅ = 233 × 300 μm, n = 10), solitary, scattered, immersed, erumpent on host, uni-loculate, black, globose to subglobose. Peridium 15–30 μm wide, thin-walled, composed of several layers of, brown to pale brown cells of textura angularis. Hamathecium comprises numerous, 1–2 μm wide, cellular, unbranched, pseudoparaphyses, septate, without constrictions at the septa. Asci 80–115 × 15–25 μm (x̅ = 98 × 21 μm, n = 20), 8-spores, cylindric-clavate obovoid, short pedicel with rounded bitunicate, end, apex rounded with a minute ocular chamber. Ascospores 20–30 × 8–10 μm (x̅ = 27 × 10 μm, n = 20), overlapping, 2–3-seriate, fusiform, with pointed ends, hyaline, 1-septate at the center (become 3-septate when germinated), constricted at the septa, containing up to four guttulate, thick-walled, surrounded by a mucilaginous sheath. Asexual morph Not observed.
Culture characteristics: culture on PDA, colonies slow growing, umbonate, smooth, edges brown, fat or efuse, from above, brown, dark brown in reverse side.
Material examined: China, Yunnan Province on dead branch of Hevea brasiliensis, 16 September 2021, Ruifang Xu, XPER-24 (MHZU 22–0153, holotype); ex-type ZHKUCC 22–0288; ibid. (ZHKUCC 22–0289, ex-isotype).
GenBank numbers: ZHKUCC 22-0288: ITS = OR807846, LSU = OR807850, SSU = OR807848, tef1-α = OR966288; ZHKUCC 22–0289: ITS = OR807847, LSU = OR807851, SSU = OR807849, tef1-α = OR966289.
Notes: In the multigene phylogeny, Pseudochaetosphaeronema puerensis (ZHKUCC 22–0288 and ZHKUCC 22-0289) clustered as a sister taxon to P. xishuangbannaensis (ZHKUCC 23-0804, ZHKUCC 23-0805) with 100% ML, 1.00 BYPP statistical support (Fig. 26). Pairwise nucleotide comparison of ITS showed that Pseudochaetosphaeronema puerensis differs from P. xishuangbannaensis in 15/559 bp (2.68%, 1 gap) (Jeewon and Hyde 2016). Moreover, Pseudochaetosphaeronema puerensis shares similar morphologies with P. xishuangbannaensis, but can be differentiated by having the peridium with the cells of textura angularis, smaller ascomata (120–354 × 166–340 μm vs. 270–410 × 370–480 μm), smaller (22–30 × 8–11 μm vs. 30–50 × 10–20 μm) and 1–3-septate ascospores with a mucilaginous sheath. Therefore, we introduce Pseudochaetosphaeronema puerensis as a new species.
Melanommataceae G. Winter [as 'Melanommeae'], Rabenh. Krypt. -Fl., Edn 2 (Leipzig) 1.2: 220 (1885)
Notes: Melanommataceae was established by Winter (1885) and this family is characterized by globose or depressed perithecial ascomata, bitunicate and fissitunicate asci, pigmented and phragmosporous ascospores (Sivanesan 1984; Zhang et al. 2012; Hyde et al. 2013). For the taxonomic treatments of this family, we follow Wijayawardene et al. (2022).
Camposporium Harkn., Bull. South. Calif. Acad. Sci. 1: 37 (1884)
Notes: Camposporium was proposed by Harkness (1884) to introduce a single species, C. antennatum. Camposporium species are characterized by dematiaceous conidiophores, terminal, integrated, denticulate conidiogenous cells, and cylindrical and elongate, multiseptate conidia with one or more cylindrical appendages at the apex (Hughes 1951a, b; Ellis 1971; Ichinoe 1971; Whitton et al. 2002a, b; Calabon et al. 2021). Based on previous research, Koukol and Delgado (2021) proposed that Fusiconidium and Camposporium are morphologically similar, and based on this viewpoint, Fusiconidium is used as a synonym for Camposporium. The phylogenetic results of this study also support this viewpoint and provide evidence that C. ramosum provides a new host of C. ramosum from decaying petiole on Livistona chinensis from China (Fig. 28).
Camposporium ramosum Whitton, McKenzie & K.D. Hyde, Fungal Divers. 11: 177–187 (2002)
Index Fungorum number: IF489895; Facesoffungi number: FoF 16039 (Fig. 29)
Saprobic on submerged decaying petiole on Livistona chinensis. Sexual morph: Not observed. Asexual morph: Hyphomycetous. Colonies on natural substrate, effuse, brown, velvety. Mycelium immersed and superficial, septate, fasciculate, unbranched. Conidiophores 55–90 × 5–8 μm (x̅ = 71.6 × 6.9 μm, n = 10) macronematous, mononematous, unbranched, erect, irregularly cylindrical, dematiaceous towards the base, fading to pale brown towards the apex, septate, thick-walled. Conidiogenous cells holoblastic, polyblastic, integrated into the apical region of the conidiophores, lighter the color towards the top, from brown to hyaline, denticulate; cylindrical denticles act as separating cells, 1–3 denticles per conidiophore. Conidia 40–90 μm long × 8–15 μm (at the widest point) (x̅ = 51.9 × 9.0 μm, n = 40), solitary, easy to dry, cylindrical, elongate, brown or pale brown, concolorous or with 2 cells at each end paler in pigmentation, slightly thickened walls and septa, 7–9-septate, apex rounded, basal cell truncate, often with a persistent portion of the denticle attached, the apical cell gives rise to a single, simple or branched appendage; appendage hyaline, aseptate, smooth, tapering from the base to the apex, appendage is located near the base and is usually divided into 2 (sometimes 1) terminal branches, individual branches 20–40 μm long.
Culture characteristics: colonies on PDA reach 7 cm diam. at 25℃ after five days. Upper view wrinkled, filamentous, entire margin, flat, cloudy, fluffy for aerial hyphae, become grey-black with time, dense for aerial hyphae, the reverse becomes black.
Material examined: China, Yunnan Province, Qujing City, QuJing Normal University, decaying petiole on Livistona chinensis, 5 August 2022, Y.R. Xiong, XG075 (MHZU 23-0132)—living culture ZHKUCC 24-0053 = ZHKUCC 24-0054.
GenBank numbers: ZHKUCC 24–0053: ITS = PP620704; LSU = PP620707; SSU = PP620709; tef1-α = PP668964. ZHKUCC 24–0054: ITS = PP620705; LSU = PP620708; SSU = PP620710; tef1-α = PP668965.
Notes: In the present study, our two isolates (ZHKUCC 24–0053 and ZHKUCC 24-0054) obtained from Livistona chinensis, clustered together with C. ramosum (CBS 132483, KUNCC 10792, KUNCC 10793) with 99% ML values and 1.00 BYPP bootstrap support (Fig. 26). The nucleotide difference between the strain of (ZHKUCC 24-0053) and C. ramosum (CBS 132483) revealed that 0.01% in LSU (1/850 bases), and 0.05% in ITS (3/522 bases) excluding gaps. The conidia size of ZHKUCC 24-0053 are not as large as C. ramosum (CBS 132483), and only 7–9-septate (40–90 × 8–15 μm vs. 80–112 × 6.4–9.6 μm; 7–9-septate vs. 8–15-septate). However, the color distribution and morphology of the apical appendages of the C. ramosum (CBS 132483) and ZHKUCC 24-0053 are similar. Based on morphology and molecular data analysis, we identified our collections are C. ramosum and this collection represents a novel host record of the fungus on L. chinensis from China (Figs. 28, 29, 30).
[See PDF for image]
Fig. 28
Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, SSU and tef1-α sequence data representing the species of Melanommataceae. Related sequences are taken from Calabon et al. (2021). Twenty-four strains are included in the combined sequence analysis, which comprised 3302 characters with gaps (ITS = 519 bp, LSU = 829 bp, SSU = 990 bp, tef1-α = 964 bp). The tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RAxML tree with a final likelihood value of − 14,767.515593 is presented. The evolutionary model GTRGAMMA is applied for all the genes. The matrix had 876 distinct alignment patterns with 28.01% of undetermined characters or gaps. Bootstrap support values for ML equal or greater than 70% and Bayesian posterior probabilities greater than 0.90 are given near nodes respectively. The tree is rooted with Cyclothyriella rubronotata (CBS 419.85; TR9) in Cyclothyriellaceae. The ex-type strains are indicated in bold. The newly generated sequence is indicated in blue
[See PDF for image]
Fig. 29
Camposporium ramosum (MHZU 23-0132 new host record). a Substrate; b–c colonies on wood; d–f conidiophores; h–j conidia; k germinating conidium. Scale bars: b, c 50 mm; d–f 10 μm; h–j 20 μm; (k) 50 μm
[See PDF for image]
Fig. 30
The phylogram of the best ML tree based on a combined dataset (LSU, ITS, rpb2 and tef1-α) of Roussoella. The scale bar indicates 0.04 changes. The tree is rooted with Torula hollandica (CBS 220.69). The best ML tree with a final likelihood value of − 17,448.644 is presented. The alignment dataset comprises 40 taxa with 3343 sites, including 1181 distinct patterns, 717 parsimony-informative, 281 singleton sites and 2345 constant sites. Ultrafast bootstrap values for the maximum likelihood analysis equal to or greater than 75% and posterior probability for Bayesian analysis equal to or greater than 0.95 are indicated at the nodes. Type isolates are in bold, and the new taxa is in blue
Roussoellaceae Jian K. Liu, Phook., D.Q. Dai & K.D. Hyde, in Liu et al., Phytotaxa 181(1): 7 (2014)
Notes: Roussoellaceae was introduced by Liu et al. (2014) to accommodate roussoella-like species in Pleosporales. For taxonomic treatment of this family, we follow Hongsanan et al. (2020a) and Wijayawardene et al. (2022).
Roussoella Sacc., in Saccardo & Paoletti, Atti Inst. Veneto Sci. lett., ed Arti, Sér. 6 6: 410 (1888).
Notes: Roussoella species are characterized by raised, ascostromata which are immersed, cylindrical to clavate asci with ellipsoidal to fusiform, brown to dark brown, 1-septate, ornamented ascospores (Dai et al. 2022). The asexual morphs of Roussoella species are cytospora-like with enteroblastic conidiogenous cells and hyaline to dark brown conidia (Aptroot 1995). They are predominant fungal species on monocotyledons, especially bamboo (Dai et al. 2022). For the taxonomic treatment of this study (Fig. 30), we follow Dai et al. (2022).
Roussoella neoaquatica W.H. Tian & Maharachch., sp. nov.
Index Fungorum number: IF9901379; Facesoffungi number: FoF 15112 Fig. 31
[See PDF for image]
Fig. 31
Roussoella neoaquatica (HKAS 126513, holotype). a–c Conidiomata on bamboo host. d Section of conidiomata. e Peridium. f Conidiophores. g, h, p–s Conidiogenous cells. i–m, t–v Conidia. n–o Culture characteristics. Scale bars: d = 50 μm, e, f, i = 10 μm, g, h, j–m = 2 μm, p–v = 5 μm
Etymology: Refers to the morphological similarity to Roussoella aquatica.
Holotype: HKAS 126513
Saprobic on dead culms of bamboo. Sexual morph Not observed. Asexual morph coelomycetous. Conidiomata 230–260 × 110–115 mm, black, scattered, forming under black, round-shape areas on the host surface, with a short neck, flattened at the base. Ostiolar neck black, short, with a large ostiole on the surface of conidiomata. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 5–10 × 1.5–3 μm (x = 7.1 × 2.2 μm, n = 20) enteroblastic, annellidic, integrated, determinate, clavate or ampulliform to lageniform, hyaline, smooth-walled. Conidia 3.1–4.2 × 2.1–3 μm (x = 3.7 × 2.6 μm, n = 20), ellipsoidal, aseptate, straight, rounded at both ends, hyaline to yellow, smooth-walled.
Culture characteristics: On PDA, colony circular, covering the entire PDA tablet in 7 days at 25 °C, initially white, becoming yellowish brown after 20 days, surface rough, felt, with dense mycelium, dry, edge entire.
Material examined: China: Sichuan Province, Chengdu, Guoxue Park, on dead culms of bamboo, 103° 55′ 23″ E, 30° 45′ 24″ N, 16 September 2021, W.H. Tian, W20-2 (HKAS 126513, holotype), ex-type UESTCC 22.0120.
GenBank numbers: ITS: OR762065; LSU: OR762066; tef1-α: OR763024.
Notes: The phylogenetic tree shows that the isolate UESTCC 22.0120 clusters with the ex-type strain of R. aquatica (MFLUCC 18–1040), which was introduced by Dong et al. (2020) from submerged wood in a stream in Yunnan Province, China. The BLASTn analysis of ITS of our isolate UESTCC 22.0120 showed 97% identity (534/548 bp, no gaps) with R. aquatica. Our collection shares similar morphological characteristics in the shape of conidiophores and conidia with R. aquatica (Dong et al. 2020). However, it differs from R. aquatica by having black short neck and large ostiole on the round-shape conidiomata, uni-loculate, bigger conidiogenous cells (5.4–9.6 × 1.4–2.9 μm vs. 3–4 × 1.5–2 μm; (Dong et al. 2020), bigger hyaline to yellow conidia (3.1–4.2 × 2.1–3 μm vs. 2.7–3.5 × 2–2.5 μm (Dong et al. 2020). Therefore, Roussoella neoaquatica is introduced as a new species.
Tetraplosphaeriaceae Kaz. Tanaka & K. Hiray., Stud. Mycol. 64: 177 (2009)
Notes: Tetraplosphaeriaceae was established by Tanaka et al. (2009) to accommodate five genera, Polyplosphaeria, Pseudotetraploa, Quadricrura, Tetraplosphaeria and Triplosphaeria. Later, Hyde et al. (2013) synonymised Tetraplosphaeria under Tetraploa as per the nomenclature priority and designated later as the type genus of the family. Hitherto, nine genera are accepted in Tetraplosphaeriaceae viz Aquatisphaeria, Byssolophis, Ernakulamia, Polyplosphaeria, Pseudotetraploa, Quadricrura, Shrungabeeja, Tetraploa and Triplosphaeria (Wijayawardene et al. 2022). Sexual morphs in Tetraplosphaeriaceae are massarina-like and characterized by hyaline, 1–3-septate ascospores surrounded by a sheath, and asexual morphs are characterized by conidia with setose appendages (Tanaka et al. 2009; Hyde et al. 2013; Tibpromma et al. 2018). Members of the family are mainly saprobes isolated from bamboo or grasses in aquatic and terrestrial habitats. In this study, we introduce a new genus Parapolyplosphaeria for a paraphyletic lineage of Polyplosphaeria thailandica C.G. Lin, Yong Wang & K.D. Hyde uses consolidated morphological and molecular data. A new species of Shrungabeeja is also proposed from the natural forests of Kudremukh, Karnataka State, India (Fig. 32).
Parapolyplosphaeria Rajeshk., C.G. Lin, Dong Wei & K.D. Hyde., gen. nov.
Index Fungorum number: IF 901424; Facesoffungi number: FoF 16040
Etymology: epithet Parapolyplosphaeria is derived from Polyplosphaeria due to the morphological similarities with the genus.
Type species: Parapolyplosphaeria thailandica (C.G. Lin, Yong Wang bis & K.D. Hyde) Rajeshk., C.G. Lin, Dong Wei & K.D. Hyde.
Saprobic on bamboo culms. Mycelium superficial. Sexual morph Not observed. Asexual morphConidiophores absent. Conidiogenous cells monoblastic. Conidia solitary, dry, acrogenous, muriform, globose, obovoid, pyriform, ellipsoidal, occasionally two conidial body fused together at the basal cell, brown, 2–5 appendages, grey to brown, straight, septate, rounded at the apex, basal cell usually cylindrical, obconical, dark brown, smooth-walled.
Note: Polyplosphaeria was introduced by Tanaka et al. (2009) with the type species Polyplosphaeria fusca Kaz. Tanaka & K. Hiray defined based on its sexual and asexual morphs. The anamorphic state of Polyplosphaeria produces globose to subglobose conidia composed of numerous internal hyphae, thin peel-like outer wall, and three to eight setose appendages (Tanaka et al. 2009). Our multi-locus sequencing data analyses (Fig. 32) of Tetraplosphaeriaceae Kaz. Tanaka & K. Hiray. delineated Polyplosphaeria thailandica C.G. Lin, Yong Wang & K.D. Hyde, as a paraphyletic lineage that form a distant lineage from Polyplosphaeria sensu stricto Clade representing its type Polyplosphaeria fusca. After Ernakulamia and Shrungabeeja sequences were introduced into the phylogenetic analysis of Tetraplosphaeriaceae, Polyplosphaeria thailandica was consistently form a paraphyletic lineage (Hyde et al. 2020a, b, c; Phookamsak et al. 2022; Liao et al. 2022). In our analysis (IQ tree, RAxML and MrBayes), P. thailandica formed a separate clade sister to the members of Shrungabeeja (Fig. 32). The morphological comparison of P. fusca with P. thailandica also supported differences in conidial characteristics such as size, shape and appendage number. P. thailandica has some unique features like grey to brown conidia and occasionally two conidia associated together at the basal cell, which are not reported in other members of Polyplosphaeria (Tanaka et al. 2009; Tibpromma et al. 2018). Likewise, only 2–5 appendages were reported in P. thailandica while more than 5 both long and short appendages were key characters of Polyplosphaeria. Considering these consolidated evidences, we propose a new genus Parapolyplosphaeria, for Polyplosphaeria thailandica as Parapolyplosphaeria thailandica. The genera Aquatisphaeria, Ernakulamia, Parapolyplosphaeria and Shrungabeeja were delineated consistently as sister clades in all the analyses (IQ-TREE (ML-BS), RAxML (R-BS) and Bayesian analysis (PP).
[See PDF for image]
Fig. 32
Phylogenetic tree obtained from an IQ-TREE analysis of species from Tetraplosphaeriaceae based on combined LSU, ITS, SSU, and tef1-α sequences. Related sequences are taken from Phookamsak et al. (2022) and additions according to the BLAST searches in NCBI. 45 sequences are included in the analysis which comprises 3385 characters (1462 characters for LSU, 588 characters for ITS, 1366 characters for SSU, 918 characters for tef1-α) after alignment. The evolutionary model SYM + I + G4 was used for IQ tree analysis. RAxML analysis and Bayesian inference were implemented with the GTR + I + G model. Muritestudina chiangrainensis (MFLUCC 17-2551) was used as the outgroup taxa. The best-scoring IQ tree with a final likelihood value of − 11,834.995 is presented. The matrix had 728 distinct alignment patterns, with 28.55% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.250, C = 0.250, G = 0.250, T = 0.250; substitution rates: AC = 3.75859, AG = 4.80170, AT = 2.27679, CG = 1.64189, CT = 10.27537, GT = 1.00000; gamma distribution shape parameter α = 0.569. Branch support values from 1000 non-parametric bootstraps for IQ-TREE (ML-BS) and RAxML (R-BS) and posterior probability values from the Bayesian analysis (BYPP) are shown at the nodes. Bootstrap support values for ML equal to or greater than 50% are given above the nodes. Bayesian posterior probabilities (BYPP) equal to or greater than 0.90 are given above the nodes. Ex-type strains are in bold and newly generated sequences as well as new combinations are in blue bold
Parapolyplosphaeria thailandica (C.G. Lin, Yong Wang bis & K.D. Hyde) Rajeshk., C.G. Lin, Dong Wei & K.D. Hyde., comb. nov.
Index Fungorum number: IF 559658; Facesoffungi number: FoF 10840 Fig. 33
[See PDF for image]
Fig. 33
Polyplosphaeria thailandica (MFLU 15–3273 holotype) a Host (decaying bamboo) b, c Conidiophores on the host surface d–g Conidiophores, conidiogenous cell and conidia h Germinating conidium i, j Colonies on PDA culture. Scale bars: b = 200 μm, c = 100 μm, d–h = 20 μm (the figures are reproduced with the permission of original authors Li et al. 2016)
Etymology: epithet thailandica after the name of the country where this fungus is native.
Holotype: MFLU 15–3273
Basionym: Polyplosphaeria thailandica C.G. Lin, Yong Wang bis & K.D. Hyde, Fungal Diversity 78: 55 (2016)
Saprobic on bamboo culms. Sexual morph Not observed. Asexual morph hyphomycetous. Conidiophores absent. Conidiogenous cells monoblastic. Conidia solitary, dry, acrogenous, muriform, globose, obovoid, pyriform, ellipsoidal, occasionally two conidial body fused together at the basal cell, brown, 20.5–43 μm long excluding the appendages, 17.5–54 μm wide at the broadest part, verrucose; with 2–5 appendages, grey to brown, straight, septate, 23–117 μm long, 2–4.5 μm thick, rounded at the apex; basal cell usually cylindrical, obconical, dark brown, smooth-walled.
Note: The morphological comparison of Polyplosphaeria fusca with P. thailandica corroborated that the differences in conidial characteristics such as size of conidia, P. thailandica; 20.5–43 × 17.5–54 μm vs. P. fusca; 43–100(–125) μm diam (av. 71.2 μm, n = 58). The fused conidial body at the basal cell is a uniqueness in P. thailandica and appendages are lesser in number (2–5 appendages) compared to three to eight setose appendages in P. fusca. A concatenated analyses of SSU, LSU, ITS and tef1-α formed a paraphyletic lineage of Polyplosphaeria thailandica distant from Polyplosphaeria sensu stricto Clade, that endorse the recombination of P. thailandica under a new genus as Parapolyplosphaeria thailandica.
Shrungabeeja V.G. Rao & K.A. Reddy, Indian J. Bot. 4(1): 109 (1981)
Notes: Shrungabeeja was established by Rao and Reddy (1981) with S. vadirajensis as the type species. The genus is characterized by macronematous, mononematous conidiophores, monoblastic conidiogenous cells, aseptate, subglobose or turbinate conidia with filiform or horn-like appendages. Sexual morphs of Shrungabeeja are unknown. Zhang et al. (2009) described S. begonia and S. melicopes from China based on morphological characteristics. Later, Kirschner et al. (2017) also introduced a new species S. piepenbringiana merely based on morphology. Ariyawansa et al. (2015a) generated sequence data for S. longiappendiculata and assigned Shrungabeeja in Tetraplosphaeriaceae. Recently, the recollection of the type species S. vadirajensis and a new species S. aquatica was reported from a freshwater habitat and placed under Tetraplosphaeriaceae allied to S. longiappendiculata (Dong et al. 2020). Recently, S fluviatilis was also introduced in this genus based on morphology and phylogeny.
Shrungabeeja kudremukhensis O.P. Sruthi & Rajeshk., sp. nov.
Index Fungorum Number: IF901423; Facesoffungi number: FoF 16041; Figs. 34, 35
[See PDF for image]
Fig. 34
Shrungabeeja kudremukhensis (AMH 10634 SEM from holotype) a Conidia attached on conidiophores. b Conidial from top view. c Conidial base shown prominent verruculose ornamentation around the base. d–e Smooth conidial body in zoom, f–l Conidial variation. Scale bars: a = 10 μm, b, c = 2 μm, d–e = 10 μm, f–l = 10 μm
[See PDF for image]
Fig. 35
Shrungabeeja kudremukhensis (AMH 10634, stereomicroscopy holotype) a Conidia on host. b Conidiophore with conidial attachment in vivo. c–d Conidia with appendages in vivo. e Conidial body showing internal mycelia. f ex-type living culture on MEA (obverse). g Conidia and conidiophores in culture (in vitro). Scale bars: a = 100 μm, b–d = 50 μm, e = 10 μm, g = 50 μm
Etymology: Name refers to the place, where the fungus was collected, Kudremukh.
Holotype: AMH 10634
Saprobic on decaying bamboo culms in terrestrial habitat. Sexual morph: Not observed. Asexual morph: hyphomycetous. Colonies effuse, brown to dark brown. Mycelium partly superficial, partly immersed in the substratum, composed of branched, septate, pale brown, smooth-walled hyphae. Conidiophores macronematous, mononematous, erect, straight or flexuous, unbranched, smooth, thick-walled, 157.5–429 7–15 μm (=289 10 μm, n = 15) 4 –5 μm wide at the apex, 3–5 septate. Conidiogenous cells monoblastic, terminal, determinate or percurrent, pale brown to brown, smooth, 10.5–20 6–8 μm. Conidia solitary, dry, acrogenous, globose to sub globose, aseptate, hollow, pedicellate, with 3–5 filiform appendages in the anterior region continuous or sometimes 3–25 septate, verruculose at basal part around the attachment and smooth over rest of the conidial body, pale brown to brown, 30.5–55 44–70 μm (=44 57 μm, n = 30) broader than long. Appendages filiform, continuous or sometimes 3–25 septate, smooth, pale brown to brown, 156–463 3–4 μm (=313 3.6 μm, n = 30).
Culture characteristics: Colonies on PDA attaining 3.5–4.8 cm diam. after 20 days at room temperature (25 °C), cottony, grey, with entire margin, reverse almost black, no pigment produced, sporulating. Conidia 30.4–54 46–70 µm (x̄ = 45 56 µm, n = 30). Appendages 4–5, 593–1125 3–4 μm (x̄ = 744.52 3.2 µm, n = 30).
Material examined: India, Karnataka, Kudremukh (13° 08′ 28″ N, 75° 15′ 75″ E), on dead culm of Bambusa sp. (Poaceae), 17 September 2022, Sruthi O. P. and Rajeshkumar K. C. (AMH 10634, holotype); ex-type NFCCI 5693.
GenBank numbers: NFCCI 5693: ITS: OR815994, LSU: OR815980, SSU: OR815981, tef1-α: OR824934.
Notes: Phylogenetic analyses of combined SSU, LSU, ITS and tef1-α gene region sequence data showed that Shrungabeeja kudremukhensis (NFCCI 5693) is related to Shrungabeeja and forms a distinct linage with 93% ML-BS, 86% R-BS and 0.94 BYPP values (Fig. 32). Shrungabeeja species have dark conidiophores with integrated, terminal, monoblastic, determinate or percurrent conidiogenous cells and lageniform, acrogenous, solitary, aseptate, globose or turbinate conidia, with filiform or horn-like appendages (Rao and Reddy 1981). There are no known sexual morphs available for this genus. Shrungabeeja kudremukhensis shares some similarities to Shrungabeeja appendiculata in its conidial size and appendage length. However, our new species (30.5–54 46–70 μm conidia) has larger conidia than S. appendiculata (30.5–54 46–70 μm conidia). Appendage length of S. appendiculata (100–360 μm, 4–21 septate) is shorter than S. kudremukhensis, which has very long filiform appendages (156–463 μm long in natural substrate and 593–1125 μm long in culture media, aseptate to 3–25 septate). Additionally, our new species has longer conidiophores (157.5–429 7–15 μm) than any other reported Shrungabeeja species. Even though, S. kudremukhensis has some overlapping morphological characters with S. appendiculata in its conidial dimension, phylogenetic analysis delineated it as a separate lineage. Based on the above morphological and molecular characteristics, Shrungabeeja kudremukhensis is introduced as a new species and placed in Tetraplosphaeriaceae.
Eurotiomycetes Tehler ex O.E. Eriksson & K. Winka., Myconet 1(1): 6 (1997)
Notes: For taxonomic treatments, we follow Wijayawardene et al. (2022).
Chaetothyriales M.E. Barr, Mycotaxon 29: 502 (1987)
Notes: Chaetothyriales species are known as black yeasts and filamentous relatives. They cause opportunistic infections in humans and has a diverse distribution including extreme environments (Quan et al. 2020). For taxonomic treatments, we follow Quan et al. (2020) and Wijayawardene et al. (2022).
Trichomeriaceae Chomnunti & K.D. Hyde, (= Strelitzianaceae Crous & M.J. Wingf.) Fungal Diversity 56: 66 (2013)
Notes: Trichomeriaceae was introduced by Chomnunti et al. (2012) based on LSU and ITS rDNA to accommodate Trichomerium species in Chaetothyriales. For the updated taxonomic treatments of this family, we follow Quan et al. (2020) and Wijayawardene et al. (2022) (Fig. 36).
[See PDF for image]
Fig. 36
Simplified phylogram showing the best maximum likelihood tree obtained from the combined multi-gene (ITS, LSU) Matrix of 58 taxa including related genera in the family Trichomeriaceae (Sun et al. 2020) and 8 strains of related families in the same order Chaeothyriales, as outgroup. The matrix comprises 2015 characters with gaps. The tree is rooted with Vonarxia vagans and Cyphellophora spp. The best-scoring ML tree with a final likelihood value of − 12,199.13 is presented. Estimated base frequencies were as follows; A = 0.2465107, C = 0.2365431, G = 0.2783904, T = 0.2385557; substitution rates AC = 0.1245685, AG = 0.1252483, AT = 0.1321545, CG = 0.1003591, CT = 0.4422091, GT = 0.0754546. ML bootstrap support (first set) equal to or greater than 88% are given near each branch. The new isolates are in blue. The type strain is in (for this study) and T
Lithophyllospora Selbmann, Coleine gen. nov.
Index Fungorum number: IF 902388; Facesoffungi number: FoF 16042
Etimology: Lithophyllospora (from Greek lithos meaning rock, phyllo meaning leaf and spora meaning seed, spore). Named after the two different substrata where it has been isolated: Protea leaves (Crous et al. 1998, 2008) and rock (this contribution).
Colonies blackish to olivaceaous, aerial mycelium absent. Sterile hyphae branched, septate, smooth to finely verruculose. frequently constricted at septa, liberating conidia by arthric secession, brownish to olivaceous, 3–6 µm wide; hyphal cells 3–6 µm wide.
Lithophyllospora australis Selbmann & Coleine, sp. nov.
Index Fungorum number: IF902389; Facesoffungi number: FoF 07391; Fig. 37
[See PDF for image]
Fig. 37
Lithophyllospora australis (MNA-CCFEE 6417, holotype). a Frontal view of the culture. b Reverse view of the culture. c, d Smooth hyphae formed by small sub-globose or cylindrical catenate cells. e Melanized septate conidium enteroblastically germinating; 2 celled conidium and hypha undergone to rhexolithic secession. f, l 1 celled arthroconidia liberated by schizolithic secession. g, i 1- and 2-celled arthroconidia liberated via rhexolitic secession. h, m endoconidia development within a conidiogenous cell. Scale bars: a, b = 2 cm, c–m = 10 µm
Etymology: with reference to the South Pole where the fungus was isolated from.
Holotype: MUT 6702; MNA-CCFEE6417
Fungus inhabiting cryptoendolithic communities in Antarctica. Sexual morph: Not observed. Asexual morph:Mycelium composed of brown melanized smooth hyphae, formed by small roundish or cylindrical catenate cells, 4.2–6 μm, diam. Apical growth by enteroblastic elongation. 1–2 celled arthroconidia liberated by rhexolithic or schyzolithic secession or produced by enteroblastic gemination (4.3 × 13 μm, 4.3 × 15 μm, diam). Rare conidiogenous cells (13 to 15 μm, diam) produced by micronematous conidiophores liberating endoconidia brown smooth broadly spherical, 6 to 7 μm diam.
Culture characteristics: Psychrophilic fungus, no growth at 25 °C. Cauliflower-like colonies black obverse and reverse, compact, slightly glittering, with irregular margin, eventually penetrating into the agar. Very slow growing, reaching 12–13 mm diam. in 12 weeks at 15 °C (optimal T).
Material examined: Antarctica, Mt. New Zealand, Northern Victoria Land, collected by Laura Selbmann, January 2016, MUT 6702, holotype = MNA-CCFEE 6417, Culture preserved at − 150 °C and in dried condition.
GenBank numbers: ITS: OQ102327, LSU: OP912972
Notes: Lithophyllospora is now an additional genus in Trichomeriaceae. Similarly, to others, it is very slow growing comparably to Bradymyces and differently from Knufia generally grows more rapidly. Similarly, to the genus Bradymyces, Lithophyllospora australis (type) produces endoconidia, despite not frequently, but differently to any other genus in the same family displays arthroconodiogenesis both through schyzolithic and rexolithic secession. Cladophialophora proteae CPC 14902 isolated from dead leaf tissue of Encephalartos altensteinii (Crous et al. 2008), here groups together with L. australis and similarly displays conidial chains with arthric secession. Based on phylogenetic placement and morphological peculiarities here we introduce a new genus in the Trichomeriaceae and the type species Lithophyllospora australis. Cladophialophora proteae CPC 14902 is here proposed as Lithophyllospora proteae, a different species in the new genus accordingly to its phylogenetic position ecology and geography.
Eurotiales G.W. Martin ex Benny & Kimbr., Mycotaxon 12(1): 23 (1980)
Aspergillaceae Link, Abh. Königl. Akad. Wiss. Berlin: 165. 1826 [1824].
Notes: Trichocomaceae was introduced by Fischer (1897) based on the morphological key characters. Traditional morphological taxonomy was further evolved until Raper, and Fenne (1965) redefined the key morphological characteristics of and conidiophore branching based delimitation of Penicillium (Pitt 1980). However, Houbraken and Samson (2011) followed modern taxonomic approaches and segregated the family Trichocomaceae into three different families: Aspergillaceae, Thermoascaceae and Trichocomaceae based on morphology, multigene sequencing, and metabolites profiling. Following one fungus = one name concept, all the teleomorph names were subsequently merged under the priority names Aspergillus and Penicillium. Lately, Houbraken et al. (2020) reevaluated the relationship between the families and genera and revised the entire classification of order Eurotiales to even series level. They accommodated 15 genera within Aspergillaceae using a nine-gene sequencing dataset. Aspergillus contains 06 subgenera, 27 sections and 96 series that accommodate 446 species. Similarly, Penicillium comprises 02 subgenera, 32 sections and 100 series that accommodate 494 species (Crous et al. 2021; Houbraken et al. 2020). A list of accepted species and associated reference data are well maintained and updated, with 494 Penicillium species in current use.
Aspergillus P. Micheli ex Haller, Hist. stirp. Helv. (Bernae) 3: 113 (1768)
Notes: Aspergillus is the largest genus in Aspergillaceae contains more than 700 accepted species and is divided into 25 sections (Houbraken et al. 2020; Species Fungorum 2022). Aspergillus section Circumdati belongs to subgenus Circumdati and is typified by A. alutaceus and divided into three series, viz. Circumdati, Sclerotiorum and Steyniorum (Houbraken et al. 2020). Series Sclerotiorum typified by A. sclerotiorum are characterized by light yellow to ochre conidia, biseriate conidiophores and white, cream, or yellow sclerotia (Visagie et al. 2014; Houbraken et al. 2020). Secondary metabolites such as cyclopenins, radarins, secalonic acid A, secopenitrem D and sulphinines are only found in species of the series Circumdati. We follow the latest treatment and update accounts of Aspergillus in Houbraken et al. (2020) (Fig. 38).
[See PDF for image]
Fig. 38
Phylogram generated from IQ-TREE maximum likelihood analysis based on combined BenA, CaM and ITS sequence data for Aspergillus subramanianii and related species within the ser. Sclerotiorum. 11 sequences are included in the analysis which comprises 1902 characters (547 characters for BenA, 800 characters for CaM, 555 characters for ITS) after alignment. The evolutionary model TNe + G4 was used for IQ tree analysis. RAxML analysis and Bayesian inference was implemented with the GTR + I + G model. A. robustus (CBS 428.77) was used as the outgroup taxa. The best-scoring IQ tree with a final likelihood value of − 5622.236 is presented. The matrix had 349 distinct alignment patterns, with 9.96% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.250, C = 0.250, G = 0.250, T = 0.250; substitution rates: AC = 1.00000, AG = 2.88720, AT = 1.00000, CG = 1.00000, CT = 5.76158, GT = 1.00000; gamma distribution shape parameter α = 0.388. Branch support values from 1000 non-parametric bootstraps for IQ-TREE (ML-BS) and RAxML (R-BS) and posterior probability values from the Bayesian analysis (BYPP) are shown at the nodes. Bootstrap support values for ML equal to or greater than 70% are given above the nodes. Bayesian posterior probabilities (BYPP) equal to or greater than 0.90 are given above the nodes. Ex-type strains are in bold and newly generated sequences are in blue bold
Aspergillus subramanianii Visagie, Frisvad & Samson, Stud. Mycol. 78: 55 (2014)
Index Fungorum number: IF809203; Facesoffungi number: FoF 16043 Fig. 39
[See PDF for image]
Fig. 39
Aspergillus subramanianii (NFCCI 10632, new record) ab Colonies on natural substrate (on insect). cd Conidiophores. e Conidia. f Rough walled conidiophores. g Colony on MEA (obverse). h Colony on MEA (reverse). Scale bars: ab = 100 μm, c = 20 μm, df = 10 μm
Associated with dead spider Myrmaplata sp. Sexual morph: undetermined. Asexual morph: Hyphomycetes, Conidiophores biseriate, Stipes hyaline to brown, rough walled, 373–1310 × 6.5–10 μm (x̄ = 840 × 8, n). Vesicles globose, 30–34 μm wide; Metulae 8–12 × 3.5–5.5 μm, covering 100% of head, minor proportion only 75%; Phialides ampulliform, 7.2–10.6 × 2.2–3.5 μm. Conidia globose, smooth, 2.4–3 × 2.4–3 μm (2.4 ± 0.1 × 2.4 ± 0.1, n = 30).
Culture characteristics: Colonies growing after 7 days at 25 °C on MEA: 45–48 mm in diameter; Colony surface floccose to somewhat velutinous; mycelial areas white; sporulation light yellow (4A5); soluble pigment absent; exudate absent; reverse light brown to brown (5D7–6D7).
Material examined: India, Kasaragod, Karadka (12°32′28″N 75°08′11″E), on dead spider Myrmaplata sp., 23 July 2022, Rajeshkumar K. C. and Sruthi O. P., AMH 10632, living culture NFCCI 5694.
GenBank numbers: NFCCI 5694: ITS: OR807407, BenA: OR824932, CaM: OR824933.
Notes: In the phylogenetic analysis of a combination of ITS-CaM-BenA shows our isolates clustered with A. subramanianii with 100% ML-BS (IQ TREE), 100% R-ML, BYPP = 1.00 support (Fig. 33). Aspergillus subramanianii was introduced by Visagie et al. (2014) from shelled Brazil nuts, in Canada. Our strain colonised on dead insects was collected from Karadka village of Kasargod, India. Macro and micromorphological characters fit well with the description of A. subramanianii in the biseriate conidiophores with yellow conidia. Hence, we determine our isolate as a new collection of A. subramanianii from a new host and a new geographical record to India.
Penicillium Link, Mag. Ges. Naturf. Freunde Berlin 3: 16. 1809.
Notes: Link (1809) initially coined the generic name Penicillium and introduced Penicillium expansum as the type species. The morphotaxonomic taxonomic concept in this genus evolved with the revisions of Thom (1930), Raper & Thom (1949), and Ramírez (1982). However, Pitt (1979) set a benchmark in the taxonomy of Penicillium. With the introduction of DNA sequencing technology in 1990 and anticipation of the changes in ICN resulting in ‘One fungus: One name’ (McNeil et al. 2012), Houbraken and Samson (2011) transferred the genus Penicillium from Trichocomaceae to Aspergillaceae. A significant change that happened while revisiting the genus Penicillium was the delimitation of Biverticillate Penicillium (Penicillium Subgenus Biverticillium) under the Talaromyces in Trichocomaceae along with the traditional aceroid phialide Talaromyces strains and their respective teleomorphs. Visagie et al. (2014) redefined the circumscription of Penicillium following gold standard polyphasic taxonomic approaches that assured the accuracy in the determination of species boundaries. Stringent protocols were set for media, macro and micromorphology and genes for identification and phylogeny. Recently, the classification of Penicillium is well-defined by Houbraken et al. (2020). They segregated the genus into 2 subgenera, 32 sections and 100 sections (Fig. 40).
[See PDF for image]
Fig. 40
The best-scoring RAxML tree for the combined dataset of ITS, BenA, CaM, and rpb2 sequence data of Penicillium section Lanata-Divaricata and the topology and clade stability of the combined gene analyses was not significantly different from the single gene analyses. The matrix had 631 distinct alignment patterns with 4.32% undetermined characters and gaps. Estimated base frequencies were as follows; A = 0.225695, C = 0.282364, G = 0.251464, T = 0.240477; substitution rates AC = 1.069213, AG = 3.823620, AT = 1.167338, CG = 0.781656, CT = 5.725837, GT = 1.000000; gamma distribution shape parameter α = 0.231382. Bootstrap support values for ML equal to or greater than 70%, Bayesian posterior probabilities (PP) equal to or greater than 0.90 are shown as ML/PP at the nodes. The tree is rooted in Penicillium citrinum (CBS 139.45). The newly generated sequences are indicated in blue bold
Penicillium cremeogriseum Chalab., Bot. Mater. Otd. Sporov. Rast. 6:168. 1950
Index Fungorum number: IF 302390; Facesoffungi number: FoF 16044 Fig. 41
[See PDF for image]
Fig. 41
Penicillium cremeogriseum (NFCCI 5293). a, b Colonies after 7d at 25 ± 2 °C on CYA and MEA obverse and reverse. c CREA obverse. d CYAS obverse. e CZA obverse. f DG18 obverse. g OA (natural) obverse. h YES obverse. i–j terverticillate conidiophore. k Conidia. Scale bar: i–k = 10 μm
Holotype: CBS 223.66
Saprobic on soil. Sexual morph: Not observed. Asexual morph: Conidiophores predominantly biverticillate, rarely with monoverticillate; Stipes long, smooth-walled, 150–400 × 2.5–3.0 µm; Metulae two, divergent, 10–25 × 2.5–3.2 μm; Phialides ampulliform, 3–7 per metulae, 6–9 × 2.5–3.0 µm; Conidia subspheroidal to subellipsoidal, smooth to finely rougned ornamentation, 2.5 − 3.0 × 2.2 − 3.0; Sclerotia not observed.
Culture characteristics: Colonies growing after 7 days at 25 ± 2 °C on following agar media: CYA colonies fast growing, slight radially sulcations, floccose, mycelia white (1A1); 34–37 mm in diam.; margins regular; no sporulation; exudate absent; soluble pigment absent; reverse filamentous, golden (4B5) at centre, light yellow (4A4) to yellowish white (4A2) towards periphery. MEA colonies fast growing, velutinous, with a concentric circle, slightly raised, mycelia white (1A1) entire, 35–46 mm in diam.; margins regular, deep; no sporulation; exudates absent; soluble pigments absent; reverse yellowish white (1A2) entire. CYAS colonies medium-growing, floccose, raise, mycelia white (1A1), 26–27 mm in diam.; margin regular, deep; no sporulation; exudates absent; soluble pigments absent; reverse light yellow (4A4) at centre, yellowish white (4A2) towards periphery. OA colonies fast-growing, velutinous, with concentric rings, sunken at centre; mycelia white (1A1), 40–45 mm in diam.; margin regular and raised; sporulation greyish white (1B1); exudates present at centre in form of colourless droplets; soluble pigments absent; reverse yellowish white (1A2) at centre, white (1A1) towards periphery. CZ colonies medium-growing, mycelial, white (1A1) entire, 21–24 mm in diam.; margin irregular; exudates absent; soluble pigments absent; reverse white (1A1) entire. DG18 colonies slow-growing, powdery, mycelium white (1A1), 9–10 mm diam.; margin irregular; exudates absent; soluble pigments absent; reverse white (1A1) entire. YES, colonies fast-growing, velutinous, radially sulcate, slightly umbonate, mycelia white (1A1); 36–40 mm diam.; margin irregular, highly raised; sporulation greyish white (1B1); exudates absent; soluble pigments absent; reverse chamcis (4C5) at centre, light yellow (4A4) towards periphery. CREA colonies slow-growing, velutinous, umbonate, mycelia white (1A1) entire, 17–21 mm in diam.; margin irregular; reverse white (1A1) entire; acid production absent.
Material examined: India, Uttarakhand, Bhowali (29° 25′ 22″ N, 74° 53′ 19″ E), from garden soil sample, 21 March 2019, Nikhil Ashtekar and Rajeshkumar K.C., living culture NFCCI 5293.
Hosts and distribution: China, Korea, Russia, South Africa, Sri Lanka and Ukraine (Visagie et al. 2015; Diao et al. 2019; Milanović et al. 2019; Choi et al. 2020; Sang et al. 2021; Silva et al. 2021).
GenBank numbers: NFCCI 5293: ITS: OK342120, BenA: OL652653, CaM: OM948800, rpb2: OL652657.
Notes: Based on a concatenated phylogenetic analysis of the ITS, BenA, CaM, and rpb2 gene regions, the isolated strain of P. cremeogriseum (NFCCI 5293) aligns with the type strain of P. cremeogriseum (CBS 223.66). Morphologically, both strains share the same characteristics such as predominantly biverticillate conidiophores with a smooth-walled stipe, ampulliform phialides, spheroidal to subspheroida conidia with smooth ornamentation. Further, both the colonies show floccose texture with slight radial sulcations on CYA media. Hence, the macromorphological and micromorphological characteristics of the isolated species match with the type species, P. cremeogriseum (CBS 223.66). However, P. echinulonalgiovense has not been reported from India and this is the first report.
Penicillium echinulonalgiovense S. Abe ex Houbraken & R.N. Barbosa Antonie van Leeuwenhoek 111: 1895. 2018.
Index Fungorum number: IF822213; Facesoffungi number: FoF 16045 Fig. 42
[See PDF for image]
Fig. 42
Penicillium echinulonalgiovense (NFCCI RKC SVWS4). a, b Colonies after 7d at 25 ± 2 ºC on CYA and MEA obverse and reverse. c CREA obverse. d CYAS obverse. e CZA obverse. f DG18 obverse. g OA (natural) obverse. h YES obverse. i–j terverticillate conidiophore. k Conidia. Scale bar: i–k = 10 μm
Holotype: CBS H-23172
Saprobic on the soil. Sexual morph: Not observed. Asexual morph: Conidiophores predominantly biverticillate, rarely with monoverticillate; Stipes long, smooth-walled, 60–200 × 2.5–3.0 µm; Metulae 2–3, divergent, 11–17 × 2.4–3.0 μm; Phialides ampulliform, 3–6 per metulae, 7–11 × 2.5–3.0 µm; Conidia spheroidal to subspheroidal, smooth ornamentation, 2.5 − 3.5 × 2.5 − 3.5.
Culture characteristics: Colonies growing after 7 days at 25 ± 2 °C on following agar media: CYA colonies medium- growing, radially sulcate, velutinous, mycelia dense white (1A1); 14–15 mm in diam.; margins regular; no sporulation; exudate absent; soluble pigment absent; reverse orange, white (5A2) entire. MEA colonies medium-growing, velutinous, floccose, mycelia white (1A1), 15–16 mm in diam.; margins regular; sporulation dense, pale yellow (4A3); exudates absent; soluble pigments absent; reverse deep yellow (4A8) at centre, maize yellow (4A6) towards periphery. CYAS colonies medium-growing, velutinous, mycelia white (1A1) entire, 20–22 mm in diam.; margin regular; no sporulation; exudates absent; soluble pigments absent; reverse light yellow (4A4) at centre, yellowish white (4A2) towards periphery. OA colonies fast-growing, velutinous, raised; mycelia white (1A1), 45–50 mm in diam.; margin irregular; sporulation dense, grey (1D1); exudates absent; soluble pigments absent; reverse white (1A1) entire. CZ colonies medium-growing, mycelial, yellowish white (1A1) entire, 24–25 mm in diam.; margin regular; exudates absent; soluble pigments absent; reverse yellowish white (1A1) entire. DG18 no growth. YES colonies fast-growing, velutinous, radially sulcate, mycelia white (1A1); 46–49 mm diam.; margin regular, raised; sporulation light yellow (4A5); exudates absent; soluble pigments absent; reverse light yellow (4A4) entire CREA colonies medium-growing, velutinous, mycelia white (1A1) entire, 23–21 mm in diam.; margin irregular; reverse white (1A1) entire; acid production absent.
Material examined: India, Kerala, Ernakulam (9° 58′ 48″ N, 76° 17′ 16″ E), an endophyte from Withania somnifera (L.) Dunal (Solanaceae), 27 July 2021, Sherin Varghese and Rajeshkumar K.C., living culture NFCCI RKC SVWS4.
Hosts and distribution: Brazil, Malaysia and Nigeria (Barbosa et al. 2018; Freire et al. 2020; Houbraken et al. 2020).
GenBank numbers: NFCCI RKC SVWS4: ITS = OP942450, BenA = OP961926, CaM = OP961927.
Notes: Based on a concatenated phylogenetic analysis of the ITS, BenA, CaM, and rpb2 gene regions, of the isolated endophytic fungal strain, P. echinulonalgiovense (NFCCI RKC SVWS4) aligns with the type stain of P. echinulonalgiovense (CBS H-23172). The isolated strain is well placed in the phylogenetic circumscription of series Simplicissima under section Lanata-Divaricata. Morphologically, P. echinulonalgiovense (NFCCI RKC SVWS4) share the same characteristics as the type strain of P. echinulonalgiovense such as biverticillate penicilli, having ampulliform phialides and conidia echinulate, globose to subglobose. Further, both the colonies show floccose texture with radial sulcations, and absence of exudates on CYA media. Hence, the macromorphological and micromorphological characteristics of the isolated species match with the type species, P. echinulonalgiovense (CBS H-23172). However, Penicillium echinulonalgiovense has not been reported from India and this is the first report.
Penicillium javanicum J.F.H. Beyma, Verh. Kon. Ned. Akad. Wetensch., Afd. Natuurk. 26: 17. 1929.
Index Fungorum number: IF268394; Facesoffungi number: FoF 16046 Fig. 43
[See PDF for image]
Fig. 43
Penicillium javanicum (NFCCI RKCNK78). a, b Colonies after 7d at 25 ± 2 ºC on CYA and MEA obverse and reverse. c CREA obverse. d CYAS obverse. e CZA obverse. f DG18 obverse. g OA (natural) obverse. h YES obverse. i–j terverticillate conidiophore. k Conidia. Scale bar: i–k = 10 μm
Holotype: CBS 341.48.
Saprobic on soil. Sexual morph: Not observed. Asexual morph: Conidiophores predominantly monoverticillate, rarely with solitary phialides; Stipes smooth-walled, short 40–90 × 1.5–2.0 µm; Phialides ampulliform, with a narrow neck, 8–11 × 2.0–2.5 µm, in verticils of 2–5; Conidia smooth, subspheroidal to ellipsoidal, 2.5–3.0 × 2.2–2.5, 5 µm, borne in irregular columns; Sclerotia not observed.
Culture characteristics: Colonies growing after 7 days at 25 ± 2 °C on following agar media: CYA colonies fast growing, radially sulcate, slightly umbonate towards centre, velutinous, mycelia champagne (4B4) to white (1A1), 35–42 mm in diam.; margins irregular, deep; sporulation dull green (27E3), star shaped at centre; exudate light orange (5A5) droplets, present irregularly; soluble pigment absent; reverse filamentous, butter yellow (4A5) at centre, cream (4A3) to yellowish white (4A2) towards periphery. MEA colonies fast-growing, velutinous, slightly umbonate, raised, mycelia light yellow (4A4) to white (1A1), 32–57 mm in diam.; margins irregular, deep; sporulation dull green (26D3); exudates absent; soluble pigments absent; reverse dull green (26D3) at centre, light yellow (4A4) to pale yellow (3A3) towards periphery. CYAS colonies medium-growing, velutinous, highly wrinkled, umbonate, mycelia white (1A1), 26–29 mm in diam.; margin irregular, deep; sporulation dull green (27D3); exudates absent; soluble pigments absent; reverse bottle green (26F4) at centre, butter yellow (4A5) towards periphery. OA colonies medium-growing, velutinous, mycelia pastel yellow (2A4) to white (1A1), 23–28 mm in diam.; margin irregular; sporulation jade green (27E5); exudates present in form of colourless droplets; soluble pigments absent; reverse dull green (26F4) at centre, pale yellow (1A3) to white (1A1) towards periphery. CZ colonies medium-growing, velutinous, raised, white (1A1) entire, 22–30 mm in diam.; margin irregular; exudates absent; soluble pigments absent; reverse white (1A1) entire. DG18 colonies slow-growing, velutinous, mycelium white (1A1), 15–17 mm diam.; margin irregular; sporulation greyish green (27D5) to pale green (27A3); exudates absent; soluble pigments absent; reverse jade green (27E5) to greyish green (27B5) at centre, greenish white (27A2) to white (1A1) towards periphery. YES colonies fast-growing, velutinous, highly wrinkled, umbonate, mycelia pastel yellow (1A4) to white (1A1), 40–58 mm diam.; margin regular; sporulation jade green (27E5); exudates absent; soluble pigments absent; reverse olive brown (4E5) at centre, chinese yellow (4B7) to butter yellow (4A5) towards periphery. CREA colonies medium-growing, velutinous, greenish white (28A2) mycelia, 24–32 mm in diam.; margin irregular; reverse white (1A1) entire; acid production present.
Material examined: India, Karnataka, Mangalore (12°53′57″N, 74°53′19″E), from soil sample, 30 March 2019, Nikhil Ashtekar and Rajeshkumar K.C., living culture NFCCI RKCNK78.
Hosts and distribution: China, Colombia, French Guiana, Indonesia, Korea, Pakistan and Brasil (Khan et al. 1992; Houbraken et al. 2011; Amaria et al. 2016; Lunardelli et al. 2016; Park et al. 2016; Ramos et al. 2018; Liang et al. 2020)
GenBank numbers: NFCCI RKCNK78: ITS = OK342120, BenA = OL652653, CaM = OM948800, rpb2 = OL652657.
Notes: Based on a concatenated phylogenetic analysis of the ITS, BenA, CaM, and rpb2 gene regions, the isolated strain of P. javanicum (NFCCI RKCNK78) aligns with the type strain of P. javanicum (CBS 341.48). Morphologically, both strains share the same characteristics such as strictly monoverticillate conidiophores with smooth-walled stipes, phialides ampulliform, ellipsoidal to pyriform conidia with smooth to finely rough ornamentation. Further, both the colonies show floccose texture with radial sulcations, and the presence of exudates on CYA media. Hence, the macromorphological and micromorphological characteristics of the isolated species match with the type species, P. javanicum (CBS 341.48). However, P. javanicum has not been reported from India and this is the first report (Fig. 44).
[See PDF for image]
Fig. 44
The best scoring RAxML tree for the combined dataset of ITS, BenA, CaM, and rpb2 sequence data of Penicillium section Chrysogena and the topology and clade stability of the combined gene analyses was not significantly different from the single gene analyses. The matrix had 146 distinct alignment patterns with 1.04% undetermined characters and gaps. Estimated base frequencies were as follows; A = 0.235786, C = 0.262842, G = 0.264881, T = 0.236491; substitution rates AC = 1.257946, AG = 4.575327, AT = 0.804181, CG = 1.096188, CT = 9.062628, GT = 1.000000; gamma distribution shape parameter α = 0.747497. Bootstrap support values for ML equal to or greater than 70%, Bayesian posterior probabilities (PP) equal to or greater than 0.90 are shown as ML/PP at the nodes. The tree is rooted to Penicillium roqueforti (CBS 221.30). The newly generated sequences are indicated in blue bold
Penicillium lanosocoeruleum Thom, Penicillia: 322. 1930.
Index Fungorum number: IF268949; Facesoffungi number: FoF 16047 Fig. 45
[See PDF for image]
Fig. 45
Penicillium lanosocoeruleum (NFCCI 5282). a, b Colonies after 7d at 25 ± 2 °C on CYA and MEA obverse and reverse. c CREA obverse. d CYAS obverse. e CZA obverse. f DG18 obverse. g OA (natural) obverse. h YES obverse. i–k terverticillate conidiophore. l Conidia. Scale bar: i–l = 10 μm
Holotype: CBS 215.30
Saprobic on soil. Sexual morph: Not observed. Asexual morph: Conidiophores predominantly terverticillate, borne from surface or aerial hyphae; Stipes rough-walled, commonly 100–550 µm × 1.5–2.0 µm; Metulae in verticils of 3–4, large, 8–12 × 2.5–3.0 µm, appressed; Phialides ampulliform, 6–9 × 2.0–2.5 µm, in verticils of 2–5; Conidia smooth, sub globose to ellipsoidal, commonly 3.0–4.0 × 2.5–3.0 µm, long, rather irregular columns.
Culture characteristics: Colonies growing after 7 days at 25 ± 2 °C on following agar media: CYA colonies fast growing, radially sulcate, highly umbonate, velutinous, mycelia white (1A1), 31–34 mm in diam.; margins regular, deep; sporulation cream (4A3); exudate colorless to pale yellow (1A2) droplets; soluble pigment absent; reverse radially sulcate, melon (5A6) to maize yellow (4A6) entire. MEA colonies fast growing, velutinous, slightly umbonate with presence of concentric rings, raised, mycelia white (1A1); 34–37 mm in diam.; margins regular, deep; sporulation deep green (28E8) at centre, greyish green (28C6–26C4); exudates colorless to pale yellow (1A2) droplets; soluble pigments absent; reverse radially sulcate, melon (5A6) to maize yellow (4A6) entire. CYAS colonies medium-growing, velutinous, radially sulcate, umbonate, mycelia white (1A1), 26–28 mm in diam.; margin regular, deep; sporulation dark green (27F6); exudates absent; soluble pigments absent; reverse radially sulcate, pale yellow (3A3) entire. OA colonies medium-growing, powdery with concentric rings, mycelia white (1A1), 24–25 mm in diam.; margin regular, deep; sporulation greyish green (27E6); exudates present in form of colourless droplets; soluble pigments absent; reverse clay (5D5) at centre, yellowish white (3A2), pale green (29A3), greyish green (27B4) towards periphery, white (1A1) margin. CZ colonies slow-growing, powdery, flat, mycelia white (1A1), 12–13 mm in diam.; margin irregular; sporulation cream(4A3); exudates absent; soluble pigments absent; reverse greyish orange (5B3) at centre, white (1A1) towards periphery. DG18 colonies slow-growing, lanose, mycelium white (1A1), 15–16 mm diam.; margin irregular; sporulation greenish white (26A2); exudates absent; soluble pigments absent; reverse white (1A1) entire. YES colonies medium-growing, velutinous, radially sulcate, highly umbonate, mycelia greenish white (26A2), 34–38 mm diam.; margin regular, thin; sporulation dark green (26F6) at centre, dull green (26E4) to pale green (26A3) towards periphery; exudates in form of small colourless droplets; soluble pigments absent; reverse radially sulcate, butter yellow (4A5) at centre, pastel yellow (3A4) to pale yellow (3A3) towards periphery. CREA colonies medium-growing, velutinous with concentric rings, white (1A2) mycelia, 24–25 mm in diam.; margin irregular; sporulation pastel green (26A4) to dull green (26D4); reverse cream (4A3) at centre, pale yellow (1A2) towards periphery; acid production present.
Material examined: India, Maharashtra, Mahableshwar, (17°92′18″N, 73°68′70″E), from soil sample, 18 July 2017, Nikhil Ashtekar and Rajeshkumar K.C., living culture NFCCI 5282.
Hosts and distribution: Ethiopia, Germany, Italy, Netherlands, and Poland, (Świzdor 2013; Świzdor et al. 2018; Dijksterhuis et al. 2019; Corrado et al. 2021; Allawi and Al-Taee 2022).
GenBank numbers: NFCCI 5282: ITS = OK342120, BenA = OL652653, CaM = OM948800, rpb2 = OL652657.
Notes: Based on a concatenated phylogenetic analysis of the ITS, BenA, CaM, and rpb2 gene regions, the isolated strain of P. lanosocoeruleum (NFCCI 5282) aligns with the type strain of P. lanosocoeruleum (CBS 215.30). Morphologically, both strains share the same characteristics such terverticillate conidiophores with rough-walled stipe, ampulliform phialides, ellipsoidal conidia with smooth ornamentation. Further, both the colonies show presence of exudates on YES media. Hence, the macromorphological and micromorphological characteristics of the isolated species match with the type species, P. lanosocoeruleum (CBS 215.30). However, P. lanosocoeruleum has not been reported from India and this is the first report (Fig. 46).
[See PDF for image]
Fig. 46
The best scoring RAxML tree for the combined dataset of ITS, BenA, CaM, and rpb2 sequence data of Penicillium section Fasciculata and the topology and clade stability of the combined gene analyses was not significantly different from the single gene analyses. The matrix had 294 distinct alignment patterns with 2.16% undetermined characters and gaps. Estimated base frequencies were as follows; A = 0.233025, C = 0.269312, G = 0.252538, T = 0.245125; substitution rates AC = 1.178250, AG = 2.934410, AT = 0.943193, CG = 0.732874, CT = 8.010509, GT = 1.000000; gamma distribution shape parameter α = 0.843671. Bootstrap support values for ML equal to or greater than 70%, Bayesian posterior probabilities (PP) equal to or greater than 0.90 are shown as ML/PP at the nodes. The tree is rooted to Penicillium expansum (ATCC 7861). The newly generated sequences are indicated in blue bold
Penicillium polonicum K.W. Zaleski, Bull. Int. Acad. Polon. Sci., Ser. B., Sci. Nat. 1927: 445. 1927
Index Fungorum number: IF274889; Facesoffungi number: FoF 16048 Fig. 47
[See PDF for image]
Fig. 47
Penicillium polonicum (NFCCI 5149). a, b Colonies after 7d at 25 ± 2 °C on CYA and MEA obverse and reverse. c CREA obverse. d CYAS obverse. e CZA obverse. f DG18 obverse. g OA (natural) obverse. h YES obverse. i–j terverticillate conidiophore. k biverticillate conidiophore. l Conidia. Scale bar: i–l = 10 μm,
Holotype: CBS 222.28
Saprobic on soil. Sexual morph: Not observed. Asexual morph: Conidiophores predominantly terverticillate, rarely biverticillate, borne singly from subsurface hyphae; Stipes rough-walled, 200–400 × 3.0–4.0 µm; rami 15–20 × 3.0–3.5 µm; Metulae in verticils of 3–4, 10–12 × 2.8–3.5 µm; Phialides ampulliform, with undistinguished collula, 9–10 × 2.5–2.8 µm, in verticils of 5–8 per metulae; Conidia smooth, subspheroidal to subellipsoidal, 3.5–4.0 × 2.5–3.2 µm, borne in well-defined columns.
Culture characteristics: Colonies growing after 7 days at 25 ± 2 °C on following agar media: CYA colonies medium growing, radially sulcate, umbonate, velutinous, mycelia white (1A1), 9–11 mm in diam.; margins irregular, wavy, deep; exudate absent; soluble pigment absent; reverse sulcate, yellowish white (4A2) at centre, white (1A1) towards periphery. MEA colonies fast growing, powdery, slightly umbonate with presence of concentric rings, mycelia white (1A1); 17–18 mm in diam.; margins regular, thin, deep; sporulation dark green (27F6) to nickel green (27F4); exudates absent; soluble pigments absent; reverse pale yellow (3A3) centre, butter yellow (4A5), pale yellow (3A3), olive brown (4E5) towards periphery, yellowish white (3A2) margin. CYAS colonies medium-growing, granular, radially sulcate, highly umbonate, mycelia white (1A1), 12–14 mm in diam.; margin irregular; sporulation dark green (27F6) to pale green (25A3); exudates absent; soluble pigments absent; reverse radially sulcate, corn yellow (4B5) at centre, pale yellow (3A3) towards periphery. OA colonies slow-growing, powdery, mycelia white (1A1), 8–9 mm in diam.; margin regular, raised; sporulation greyish green (25D4); exudates absent; soluble pigments absent; reverse viridine green (29A5) at centre, white (1A1) towards periphery. CZ colonies slow-growing, flat, mycelia white (1A1) entire, 3–4 mm in diam.; margin irregular, deep; exudates absent; soluble pigments absent; reverse white (1A1) entire. DG18 colonies medium-growing, lanose, mycelium white (1A1) entire, 7–10 mm diam.; margin irregular, deep; exudates absent; soluble pigments absent; reverse white (1A1) entire. YES colonies fast-growing, velutinous, floccose, high radial sulcations, highly umbonate, mycelia white (1A1), 22–24 mm diam.; margin irregular, deep; sporulation dull green (25D4) at centre, greyish green (25E5) to pale green(26A3) towards periphery; exudates absent; soluble pigments absent; reverse wrinkled, topaz (5C5) at centre, blonde (4A4) towards periphery. CREA colonies medium-growing, powdery, white (1A1) mycelia, 9–11 mm in diam.; margin irregular, wavy, deep; sporulation pale green (26A3) at centre, orange, white (5A2) towards periphery; reverse corn yellow (4B5) at centre, cream (4A3) to yellowish white (4A2) towards periphery; acid production present.
Material examined: India, Punjab, Atari, (39° 52′ 21″ N, 75° 04′ 59″ E), from soil sample, 19 July 2018, Nikhil Ashtekar and Rajeshkumar K.C., living culture NFCCI 5149.
Hosts and distribution: Algeria, Belgium, Brazil, Canada, Chile, China, Croatia, Egypt, Estonia, Iraq, Iran, Italy, Korea, Norway, Pakistan, Poland, Russia, Saudi Arabia, Serbia, Slovakia, South Africa, Spain, Turkey, United States of America (Núñez et al. 2000; Sonjak et al. 2006; Chatterton et al. 2012; Felšöciová et al. 2012; Lamsal et al. 2013; Niknejad et al. 2013; Duduk et al. 2014; Santini et al. 2014; Çakır and Maden 2015; Bashir et al. 2017; Tangni et al. 2017; Antipova et al. 2018; Prencipe et al. 2018; Basson et al. 2019; Khalil et al. 2019; Khokhar et al. 2019; Kalkan et al. 2020; Maamar et al. 2020; El-Dawy et al. 2021; Bradshaw et al. 2022; Costa et al. 2022; de Camargo et al. 2022; Gielen et al. 2022; Zadravec et al. 2023; Abd Oun and Abass 2023).
GenBank numbers: NFCCI 5149: ITS = OK342120, BenA = OL652653, CaM = OM948800, rpb2 = OL652657.
Notes: Based on a concatenated phylogenetic analysis of the ITS, BenA, CaM, and rpb2 gene regions, the isolated strain of P. polonicum (NFCCI 5149) aligns with the type strain of P. polonicum (CBS 222.28). Morphologically, both strains share the same characteristics such as having predominant terverticillate conidiophores with a finely rough-walled stipe, ampulliform phialides with undistinguished, sub spheroidal to ellipsoidal conidia with smooth ornamentation. Further, both the colonies show the presence of cottony colony texture with radial sulcations on CYA. Hence, the macromorphological and micromorphological characteristics of the isolated species match with the type species, P. polonicum (CBS 222.28). However, P. polonicum has not been reported from India and this is the first report (Fig. 48).
[See PDF for image]
Fig. 48
Phylogram was generated from maximum likelihood analyses based on mtSSU. Porinaceae, Gyalectaceae, and Coenogoniaceae were utilized as outgroups. The maximum likelihood (ML) analysis was conducted in RAxML. Bootstrap (BS) values ≥ 70% are annotated at the nodes. Newly identified species are highlighted in bold blue font
Lecanoromycetes O.E. Erikss. & Winka, Myconet 1(1): 7 (1997)
Notes: For taxonomic treatments, we follow Wijayawardene et al. (2022).
Graphidales Bessey, Univ. Stud. Nebraska 7: 298 (1907)
Notes: Graphidales species are an order of lichen-forming fungi in which six families are accepted (Wijayawardene et al. (2022).
Diploschistaceae Zahlbr. [as ‘Diplochistaceae’], in Engler & Prantl, Nat. Pflanzenfam., Teil. I (Leipzig) 1(1*): 121 (1905).
Notes: The family is usually included in the Graphidaceae Dumort. However was recently resurrected based on results obtained by the temporal banding approach (Aptroot á Feuerstein 2020; Aptroot et al. 2022).
Aggregatorygma M. Cáceres, Aptroot & Lücking, in Cáceres, Aptroot, Parnmen & Lücking, Phytotaxa 189(1): 89 (2014).
Notes: This is a recently described small genus with only three species (Cáceres et al. 2014; Aptroot & Feuerstein 2020; Aptroot et al. 2022), all so far only known from Brazil (Fig. 48).
Aggregatorygma isidiatum Aptroot, L.A. Santos & M. Cáceres, sp. nov.
Index Fungorum number: IF 902390; Facesoffungi number: FoF 16049; Fig. 49
[See PDF for image]
Fig. 49
Aggregatorygma isidiatum (ISE 27452, holotype). Width of picture = 2 mm
Etymology: Named after the isidia.
Corticolous Aggregatorygma differing from all known species in the genus by the presence of isidia and the absence of ascomata.
Holotype: ISE 27452
Thallus crustose, discontinuous, following the surface of the substratum, not corticate, dull, pale creamish white, up to 5 cm diam., under 0.1 mm thick, not surrounded by a prothallus. Isidia of thallus colour, not corticate, irregularly globose with constricted base, 0.1–0.2 mm wide and high. Photobiont trentepohlioid. Ascomata and pycnidia not observed.
Chemistry: Thallus UV–, K–, P–. TLC: confluentic acid.
Ecology and distribution: On tree bark in primary rain forest; only known from Brazil.
Material examined: Brazil. Amapá: Floresta Nacional do Amapá, upstream of the station, alt. 30 m, 0° 58ʹ N, 51° 36ʹ W, on tree bark in primary forest, 20 Aug. 2015, M.E.S. Cáceres & A. Aptroot (ISE 27452, holotype; isotype in ABL herbarium).
GenBank numbers: OR759506 (holotype) and OR759506 (paratype).
Note: This species is well characterized by its isidia, which could be called pseudisidia if one sticks to the strict definition of isidia as corticated structures. So far, only three species have been described in the genus (Cáceres et al. 2014; Aptroot & Feuerstein 2020; Aptroot et al. 2022), all from Brazil and all in the last decade. They are all fertile and miss the isidia.
Ocellularia psorirregularis Aptroot, L.A. Santos & M. Cáceres, sp. nov.
Index Fungorum number: IF 902391; Facesoffungi number: FoF 16050; Fig. 50
[See PDF for image]
Fig. 50
Ocellularia psorirregularis (ISE 42001, holotype). Width of picture = 6 mm
Etymology: Named after the chemistry and the irregular soredia.
Holotye: ISE 42001
Thallus crustose, discontinuous, following the surface of the substratum, not corticate, dull, pale creamish white, up to 5 cm diam., under 0.1 mm thick, not surrounded by a prothallus. Soralia of thallus colour or brighter white, irregular, 0.1–0.3 mm wide and high. Soredia granular. Photobiont trentepohlioid. Ascomata and pycnidia not observed.
Chemistry: Thallus UV–, K–, P + yellow. TLC: psoromic acid.
Ecology and distribution: On tree bark in primary rain forest; only known from Brazil.
Material examined: Brazil. Bahia: Chapada Diamantina, Lençois, Serrano along Rio de Lençois, alt. 500 m, 12° 34ʹ S, 41° 23ʹ 55ʹʹ W, on siliceous rock along river, 24 Jul. 2017, M.E.S. Cáceres & A. Aptroot (ISE 42001, holotype; isotype in ABL herbarium).
GenBank number: OR759505.
Note: This species is well characterized by its soredia and the presence of psoromic acid. So far, very few sorediate species are known in the genus (see e.g., Sipman et al. 2012), and only one of those, viz. O. microsorediata Rivas Plata & Lücking (Rivas Plata and Lücking 2012) contains psoromic acid. That species has however a corticate green thallus and more or less reticulte soredia. Its mtSSU sequence is identical to O. plicata Rivas Plata & Lücking (Rivas Plata and Lücking 2012), which is however a non-sorediate species. Similar cases where different species have identical mtSSU sequences have been reported before, even in the genus Ocellularia, by Kraichak et al. (2014).
Graphidaceae Dumort. [as ‘Graphineae’], Comment. bot. (Tournay): 69 (1822)
Notes: Graphidaceae Dumortier (1822) is the largest family of crustose lichen-forming fungi with more than 2000 species worldwide (Lucking et al. 2013; Rivas Plata et al. 2013). The family comprises predominantly lichenized fungi except for the monotypic genus Papilionovela which is saprophytic (Aptroot 1997). Two main types of photobionts are found in the family, chlorococcoid and trentepohlioid algae (Friedl and Gartner 1988; Nelsen et al. 2011). The genera within the family are circumscribed based on the combination of phenotypic characters as well as chemical characters (Staiger 2002). The recent molecular taxonomic investigations included previously separated families Asterothyriaceae, Gomphillaceae, Solorinellaceae and Thelotremataceae within Graphidaceae (Mangold et al. 2008; Baloch et al. 2010; Rivas Plata and Lumbsch 2011; Rivas Plata et al. 2012). Rivas Plata et al. (2012) proposed a revised classification for Graphidaceae based on recent phylogenetic studies which treated Asterothyriaceae, Gomphillaceae, and Thelotremataceae within Graphidaceae with three subfamilies Fissurinoideae, Gomphilloideae, and Graphidoideae. Cáceres et al. (2020) described a new subfamily, Rubikioideae, in which the genus Papilionovela forms the first non-lichenized, saprobic lineage within the core Graphidaceae (subfamily Graphidoideae). The type genus of Graphidaceae is Graphis Adans.
Graphis Adans., Fam. Pl. 2: 11 (1763)
Notes: Graphis was introduced by Adanson (1763). The type species is Graphis scripta (L.) Ach., The genus includes species having crustose, corticolous, or rarely saxicolous or folicolous thallus; lirelline, elongate, simple to irregularly branched lirellae; hyaline to carbonized, usually prosenchymatous excipulum; simple paraphyses; unitunicate asci; hyaline, amyloid to non-amyloid, transversely to muriform septate ascospores and variable thallus chemistry. The availability of representative molecular sequences compared to the number of described species shows a wide gap which warrants further studies in the genus using an integrated taxonomic approach using morphological and molecular tools. We report Graphis mikuraensis new from India along with its phylogenetic placement (Fig. 51).
[See PDF for image]
Fig. 51
Phylogram generated from maximum likelihood analysis based on mtSSU, LSU and rpb2 sequence dataset representing genera Graphis and Allographa and related families. Fifty-two sequences are included in the combined analyses which comprise 2644 characters after alignment. Phaeographis intricans (voucher Kalb 38864) is used as the outgroup taxa. Single gene analyses were also performed to compare the topology and clade stability with combined gene analyses. The tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best IQtree topology with a final likelihood value of − 14,101.317 is presented. The matrix had 1003 distinct alignment patterns, with 50% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.295, C = 0.183, G = 0.257, T = 0.264; substitution rates AC = 1.31948, AG = 6.10905, AT = 2.65601, CG = 1.08632, CT = 12.00548, GT = 1.00000; gamma distribution shape parameter α = 0.616. Bootstrap values for maximum likelihood (MLBS) equal to or greater than 50% and posterior probabilities (BYPP, right) equal to or greater than 0.97 (the rounding of values to 2 decimal places) from Bayesian inference analysis labelled on the nodes. The newly generated sequences are indicated in bold and blue
Graphis chlorotica A. Massal., in Krempelhuber, Verh. Kaiserl. -Königl. zool.-bot. Ges. Wien 21: 865 (1871)
Index Fungorum number: 385629; Facesoffungi number: FoF 16051 Fig. 52
[See PDF for image]
Fig. 52
Graphis chlorotica (AMH22.31). a, b Thallus. c Cross section of lirellae. d Hymenium. e Ascus showing ascospores. f I + ascospores. g–h Ascospore. Scale bars: a = 1 mm, b = 200 µm, c = 50 µm, d = 20 µm, e–h = 10 µm
Thallus crustose, corticolous, surface smooth, dull, continuous to sometimes cracked, greenish grey to white grey. Ascocarp lirellate, erumpent, short–sparsely branched, with a lateral to basal thick thalline margin, tenella type, 0.5–4 × 0.5–2 mm. Disc brown, non-pruinose, closed, getting narrowly exposed upon hydration; Labia smooth becoming striate, black, not pruinose. Exciple 6–8 striate, apically carbonized, brownish towards the base. Epithecium granular, brownish, 6–10 μm high. Hymenium hyaline, clear, KI–, 64–76 μm high. Subhymenium hyaline, 8–14 μm high. Ascospores 8 per ascus, hyaline, ellipsoid, transversely 7–13 septate, I + blue, 30–40 × 5–8 μm.
Secondary chemistry: No substances detected by TLC, K–
Material examined: INDIA, Maharashtra, Mahabaleshwar (17°56′ 15″ N, 73° 39′ 34″ E, + 1336 msl.),15 September 2022, P. A. Ansil and K. C. Rajeshkumar, (AMH22.31).
GenBank numbers: AMH22.31: mtSSU = OR807878, LSU = OR807872, rpb2 = OR806896.
Notes: This study presents the phylogenetic placement of Graphis chlorotica. G. chlorotica is similar to G. subtenella concerning apically carbonized striate exciple and transversely septate ascospores having no detected lichen acids. G. subtenella possess a white–grey thallus and ascocarp with lateral thalline margin and G. chlorotica differs by having a green-grey thallus and ascocarp with basal thalline margin. Considering the similarities, Lücking et al. (2009) treated G. subtenella as a synonym for G. chlorotica. Later, by revisiting the type material of G. chlorotica, PEÑA et al. (2014) removed G. subtenella from synonymy and accepted it as a separate species based on morphological differences. The sample studied here showed greenish grey thallus and ascocarps with apically thin complete to lateral thick thalline margin and hence identified as G. chlorotica. The phylogenetic analyses using combined mtSSU, LSU and rpb2 sequences placed G. chlorotica a weakly supported clade allied to G. proserpens. However, further multigene phylogenetic study is required to understand both G. chlorotica and G. subtenella.
Graphis mikuraensis Y. Ohmura & M. Nakan., in Ohmura et al., Bull. natn. Mus. Nat. Sci., Taiwan 42(1): 5 (2016)
Index Fungorum number: 815136; Facesoffungi number: FoF 16052; Fig. 53
[See PDF for image]
Fig. 53
Graphis mikuraensis (AMH22.233). a, b Thallus. c Cross section of lirellae. d Inspersed hymenium. e Paraphyses. f Ascus showing ascospores. g I + ascospore. h–j Ascospores. Scale bars: a = 1 mm, b = 200 µm, c = 50 µm, d–j = 10 µm
= Graphis srilankensis Weerakoon et al., Bryologist 115: 79 (2012)
Thallus crustose, corticate, surface smooth, continuous to sometimes cracked, greenish grey to pale yellowish buff. Ascocarp lirellae, erumpent to prominent, long, simple to sparsely branched, with a lateral thick thalline margin, edge round, 0.5–14 × 0.2–0.3 mm. Disc not visible from the surface. Labia smooth, black, not pruinose. Exciple convergent, entire, completely carbonized. Epithecium indistinct. Hymenium hyaline, inspersed with small oil-droplets persistent in KOH, KI–, 110–155 μm high. Subhymenium hyaline, 10–15 μm high. Ascospores 8 per ascus, hyaline, ellipsoid to sub-fusiform, transversely 11–14 septate, first surrounded by a halo, I + violet, 40–60 × 6–10 μm.
Secondary chemistry: Norstictic acid major and Salazinic acid minor detected by TLC; K + yellow turns red with cryastals.
Material examined: India, Kerala, Munnar (10° 05′ 24″ N, 77° 03′ 17″ E, + 1477 msl.), 18 December 2022, P. A. Ansil and K. C. Rajeshkumar, (AMH22.233).
GenBank numbers: AMH22.233: mtSSU = OR807879, LSU = OR807873, rpb2 = OR806897.
Notes: This study reports Graphis mikuraensis as a new record to India and its phylogenetic placement. The sample under the study slightly varies from G. mikuraensis holotype with respect to lirella length and degree of branching. However, we are not considering this single variation enough to propose a new species and identify the species as G. mikuraensis. This sample resembles G. sapii which has elongate and irregularly branched lirellae and by ascospore characters (40–44 × 6–7 μm in length, 12 locules in Zahlbruckner (1930: 40). However, G. sapii differs from our sample in having a scabrous ‘deserpense-morph’ (Lücking 2009; Lücking et al. 2009) lirellae with narrow acute edge, rimiform disc and sometimes sub granulose-uneven thallus. We have also noted the similarity of our sample with Graphis rajapakshana. However, on comparative analysis of G. rajapakshana with G. mikuraensis (syn.: G. srilankensis Weerakoon et al. (2019). The characters (prominent lirellae and ascospore size 35–50 × 7–8 µm) quoted to separate G. rajapakshana from G. mikuraensis were found overlapping with the G. mikuraensis holotype characters (sessile to erumpent lirellae and ascospore size (35–)43.7 ± 5.7(–54) × (6.7–)8.1 ± 0.8(–10) µm). Hence, we suggest synonymizing G. rajapakshana under Graphis mikuraensis Y. Ohmura & M. Nakan. (syn.: G. srilankensis Weerakoon et al. (2019). The phylogenetic analyses using combined mtSSU and LSU sequences placed in G. mikuraensis allied to the clade containing G. leptoclada, G. gracilescensis, G. lineola, G. librata.
Graphis panhalensis (Patw. & C.R. Kulk.) Chitale, Makhija & B.O. Sharma, Mycotaxon 115: 474 (2011)
Index Fungorum number: 519294; Facesoffungi number: FoF 16053; Fig. 54
[See PDF for image]
Fig. 54
Graphis panhalensis (AMH21.48). a, b Thallus. c, d Cross section of lirellae. e hymenium. f, g Ascus showing ascospores. h,i ascospore. i Ascospore. Scale bars: a = 1 mm, b = 500 µm, c–d = 50 µm, e = 40 µm, f–g = 20 µm, h–i = 10 µm
≡ Graphina panhalensis Patw. & C.R. Kulk., Norw. Jl Bot. 26(1): 47 (1979)
Thallus crustose, corticolous, corticate, surface smooth to indistinctly warty, pale glaucous green to greenish grey. Ascocarp lirellate black, emergent, long, flexuous or curved, simple to sparsely branched, with a lateral thick thalline margin, ends sub-acute, 0.5–6 × 0.1–0.4 mm. Disc concealed. Labia striate, black, not pruinose. Exciple convergent, 4–5 striate, pale brown, apically to sometimes laterally carbonized. Epithecium granular, brown. Hymenium hyaline, clear, KI–, 70–140 μm high. Subhymenium hyaline, 10–17 μm high. Ascospores 8 per ascus, hyaline, muriform, ellipsoid, I–, 45–65 × 12–16 μm.
Secondary chemistry: Stictic acid detected by TLC; K + yellow.
Material examined: India, Maharashtra, Satara, Thoseghar (17° 36′ 05″ N, 73° 52′ 17″ E, + 1106 msl.), 30 September 2021, P. A. Ansil and K. C. Rajeshkumar, (AMH21.48).
GenBank numbers: AMH21.48: mtSSU = OR807880, LSU = OR807874.
Notes: This study presents the phylogenetic placement of Graphis panhalensis. Graphis panhalensis was established as Graphina panhalensis by Patwardhan and Kulkarni (1979), Later renamed as Graphis panhalensis (Chitale et al. 2011) based on the convergent, well-developed exciple with distinct carbonized areas and colourless ascospores. This species was known only for its type. G. panhalensis is similar to G. parilis Kremp. in external morphology, thallus chemistry and spore size and septation. However, G. panhalensis differs by having 8 spores per ascus and I– ascospores when compared to G. parilis having 2–8 spores per ascus and I + ascospores. The phylogenetic analyses using combined mtSSU, LSU and rpb2 sequences placed in G. panhalensis sister to G. parilis.
Graphis parilis Kremp., Flora, Regensburg 59: 422 (1876)
Index Fungorum number: 386052; Facesoffungi number: FoF 16054; Fig. 55
[See PDF for image]
Fig. 55
Graphis parilis (AMH21.63). a, b Thallus. c Cross section of lirellae. d hymenium. e–h Ascus showing ascospores. i, j I + ascospore. k, l Ascospore. Scale bars: a = 1 mm, b = 200 µm, c–d = 50 µm, e–h = 20 µm, i–l = 10 µm
= Graphina parilis (Kremp.) Müll. Arg., Bull. Soc. R. Bot. Belg. 32(no. 1): 152 (1893)
Thallus crustose, corticolous, corticate, surface smooth, continuous to cracked, greenish yellow to greenish grey, surrounded by a thin black hypothallus. Ascocarp lirellate black, erumpent to prominent, long, flexuous to curved, branched, with a thick thalline margin up to the top, ends acute, 0.5–10 × 0.1–0.2 mm. Disc concealed. Labia striate, black, not pruinose. Exciple convergent, 3–8 striate, yellowish brown, present at the base, apically carbonized. Epithecium granular, brown. Hymenium hyaline, clear, I–, KI–,100–170 μm high. Subhymenium hyaline, 10–16 μm high. Ascospores 2–8 per ascus, hyaline, muriform, oblong, I + violet, (38–) 40–75 (–85) × 13–25 μm.
Secondary chemistry: Stictic acid detected by TLC; K + yellow.
Material examined: INDIA, Maharashtra, Satara, Thoseghar (17°˚ 36′ 05″ N, 73° 52′ 17″ E, + 1106 msl.), 30 September 2021, P. A. Ansil and K. C. Rajeshkumar, (AMH21.53, AMH21.63); Tamhini village (18° 27′ 14″ N, 73° 26′ 04″ E, + 628 msl.), 05 September 2021, P. A. Ansil and K. C. Rajeshkumar (AMH21.28).
GenBank numbers: AMH21.53: mtSSU = OR807881, LSU = OR807875, rpb2 = OR806898; AMH21.63: mtSSU = OR807882, LSU = OR807876; AMH21.28: mtSSU = OR807883, LSU = OR807877, rpb2 = OR806899.
Notes: This study presents the phylogenetic placement of Graphis parilis. G. parilis is similar to G. panhalensis in external morphology, thallus chemistry, spore size, and septation. However, G. parilis differs by 2–8 spores per ascus and I + ascospores when compared to G. panhalensis, having 8 spores per ascus and broader I + ascospores. The phylogenetic analyses using combined mtSSU, LSU and rpb2 sequences placed G. parilis as a weakly supported clade sister to. G. panhalensis.
Ostropales Nannf., Nova Acta R. Soc. Scient. upsal., Ser. 4 8(no. 2): 68 (1932)
Notes: The latest updated account of Ostropales includes three families and 37 genera (Wijayawardene et al. 2022).
Spirographaceae Flakus, Etayo & Miądl., in Flakus et al., Plant and Fungal Systematics 64(2): 318 (2019)
Notes: The monotypic family of lichenicolous and fungicolous fungi was recently introduced by Flakus et al. (2019). The sexual state is characterized by apothecioid or perithecioid, cleistohymenial ascomata, a I– and K/I– hymenium, and 1-septate, hyaline, narrowly ellipsoidal to fusiform, curved or sigmoid ascospores. The asexual state is characterized by immersed, pycnidial conidiomata, and hyaline, Y-shaped (with a main axis and two diverging arms), triangular, or tetra- to polyhedral conidia.
Spirographa Zahlbr., in Engler & Prantl, Nat. Pflanzenfam., Teil. I (Leipzig) 1(1*): 96 (1903)
Notes: For a long time, the genus Spirographa was believed to be monospecific and included only the morphologically variable S. fusisporella (Nyl.) Zahlbr. (e.g., Diederich 2004). However, recent molecular studies significantly changed its generic concept by stating that Asteroglobulus Brackel, Cornutispora Piroz. and Pleoscutula Vouaux are congeneric with Spirographa (Flakus et al. 2019). Thus, currently, the genus comprises 22 species with apothecioid or perithecioid, cleistohymenial ascomata and 1-septate, hyaline, narrowly ellipsoidal to fusiform, curved or sigmoid ascospores and/or pycnidia producing hyaline, Y-shaped, triangular, or tetra- to polyhedral conidia. Some species were reported to show a wide host spectrum by previous authors (e.g., Brackel 2009, 2014; Etayo 2010; Schiefelbein et al. 2014; Łubek and Kukwa 2017); however, as it is known now, members of the re-circumscribed genus Spirographa seem to be strongly host-specific (Flakus et al. 2019), and the number of taxa is expected to increase after the re-examination of species records reported from a wide range of hosts. Fieldwork in Belarus, Germany and Poland revealed an anamorphic lichenicolous fungus with Cornutispora-like, Y-shaped conidia growing on corticolous Polycauliona polycarpa. The fungus appears to represent a new species of Spirographa and is described below (Fig. 56).
[See PDF for image]
Fig. 56
Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, SSU and rpb1 dataset including Spirographa skorinae (KRAM 73133) and related species. The scale bar indicates 0.02 changes. Fissurina aggregatula and F. nigrolabiata are selected as the outgroup taxa. Related sequences were taken from Flakus et al. (2019). Eighteen sequences are included in the analysis which comprise 2997 characters after alignment. Bootstrap support values for maximum likelihood (MLBS) equal to or greater than 70% are given above the nodes. The newly generated sequences are in blue
Spirographa skorinae Tsurykau, Brackel, Flakus & Kukwa sp. nov.
Index Fungorum number: IF901387; Facesoffungi number: FoF 16055; Fig. 57
[See PDF for image]
Fig. 57
Spirographa skorinae (KRAM 73133, paratype). a Habitus, showing six pycnidia immersed in the host thallus (two overmature pycnidia at the bottom of the picture are empty). b–e Conidia, in water. Scale bars: a = 100 µm, b–e = 10 µm
Etymology: Named after Francisk Skorina, Belarusian and East Slavic first printer, scientist, first professional educator, teacher, writer and linguist due to his significant contribution to the development of science, education, Belarusian language and literature, and the whole culture of the Belarusian and East Slavic people, on the occasion of his 535th anniversary and quincentenary Belarusian book printing.
Holotype: M-0330700
Parasitic on the thallus of corticolous lichen Polycauliona polycarpa. Sexual morph Not observed. Asexual morph Coelomycetous (“Cornutispora”-type). Mycelium immersed in the host thallus, hyaline. Conidiomata pycnidial, scattered or in groups of 2–3, initially immersed, becoming partially erumpent, black when dry, pale brown to almost hyaline when wet, globose, (65–)80–135(–150) μm in diam. (n = 16), opening by the disintegration of the upper wall; ostiole absent. Pycnidial wall paraplectenchymatic, composed of several layers of cells, pale brown at the top, hyaline at the base. Conidiophores hyaline, septate, branched, thin-walled, up to 35 μm long, arising from the innermost cell layer lining the pycnidial cavity. Conidiogenous cells hyaline, thin-walled, polyblastic, sympodial or synchronous, terminal to lateral, producing few conidia from minute loci. Conidiogenesis holoblastic. Conidia hyaline, aseptate, triangular (Y-shaped), composed of a thick main axis with truncate base, (7.5–)8.4–11.7(–14.5) × (2.0–)2.1–2.7(–3.0) μm, (x̄ = 10.1, SD = 1.7, n = 37; x̄ = 2.4, SD = 0.3, n = 37), with a narrow appendage arising from the truncate base, (2.0–)2.8–4.8(–6.0) μm and two gradually tapering divergent arms, (7.0–)8.0–10.8(–13.0) × (1.5–)1.8–2.4(–3.0) μm (x̄ = 9.4, SD = 1.4, n = 37; x̄ = 2.1, SD = 0.3, n = 37).
Material examined: Spirographa skorinae: Germany, Bavaria, Mittelfranken, Kreis Erlangen-Höchstadt, northern slope of the Kalchreuther Höhe, orchard, on twigs of Prunus avium, 49° 33′ 51.8″ N, 11° 07′ 18.0″ E, 370 m, on Polycauliona polycarpa, 15 October 2007, W. v. Brackel (M-0330700, holotype; GSU 2196, hb Brackel, isotypes); Bavaria, Mittelfranken, Kreis Erlangen-Höchstadt, Kalchreuther Höhe, orchard, on twigs of Prunus avium, 49° 33′ 23.8″ N, 11° 06′ 34.4″ E, 390 m, on Polycauliona polycarpa, 22 October 2007, W. v. Brackel (hb Brackel, paratype); BELARUS, Gomel region, Gomel district, the city of Gomel, close to Zhemchuzhnaja str., 52° 21′ 30″ N, 31° 01′ 35″ E, on corticolous Polycauliona polycarpa, 12 July 2019, A. Kavalenka s.n. (KRAM 73133, paratype); POLAND, Kotlina Biebrzańska, Biebrzański National Park, c. 8 km NW of Trzcianne village, Budy village, 53° 23′ 17″ N, 22° 34′ 28″ E, on Polycauliona polycarpa on branch of Syringa vulgaris, 16 September 2005, M. Kukwa, E. Bylińska, M. Seaward (UGDA L-12918, paratype). – Spirographa cf. lichenicola: Bolivia, Dept. Tarija, Prov. Aniceto Arce, close to la Mamora between Tarija and Bermejo, 22° 09′ 51″ S, 64° 40′ 03″ W, 1320 m, disturbed Tucumano-Boliviano Forest with Tillandsia, on corticolous Ochrolechia sp., 27 July 2015, A. Flakus 27,320 (KRAM, LPB).
GenBank numbers: Spirographa skorinae KRAM-L 73133 (paratype): ITS = OQ054789, LSU = OQ048258; Spirographa cf. lichenicola A. Flakus 27320 (KRAM): ITS = OQ054793, LSU = OQ034609.
Notes. The new species resembles Spirographa intermedia in its somehow similar conidial shape and dimensions. However, the latter can be differed from S. skorinae by having larger pycnidia, (102–)150–160(–170) μm in diam., thinner main axis (2–2.5 μm wide) and different host preference (Ochrolechia spp.) (Punithalingam 2003; Flakus et al. 2019). Moreover, the shape of shorter divergent arms is different: in S. skorinae they gradually taper towards the end in the majority of conidia, whereas in S. intermedia the divergent arms have usually swollen bases (3–4 μm long) and distinct narrow appendages (3.5–4.5 μm long) (Punithalingam 2003). Measured over the whole length (including the appendages) the divergent arms are shorter in S. intermedia (up to c. 8.5 μm when measurements taken from Fig. 8 in Punithalingam 2003) than in the new species.
Spirographa intermedia was described from Ochrolechia sp. and later reported from various hosts, e.g., Lecanora spp., Phaeophyscia orbicularis, Thamnolia vermicularis and Xanthoria parietina (Brackel 2009, 2014; Etayo 2010; Schiefelbein et al. 2014; Łubek and Kukwa 2017). However, Flakus et al. (2019) suggested that the species may be confined to a group of closely related host species and probably most records of S. intermedia, except for these from Ochrolechia spp., may represent undescribed species sharing similar conidial shape and measurements.
Spirographa lichenicola is also similar to S. skorinae in conidial shape but mainly differs in a narrower main axis measuring 1.5(–2) μm in width and very short divergent arms (4.5–6 μm long) (Punithalingam 2003). Like S. intermedia, this species was reported from many unrelated hosts (e.g., Punithalingam 2003; Diederich et al. 2018; Flakus et al. 2019) and records on hosts other than Parmelia sulcata may represent an assemblage of several phylogenetically distinct species with very similar conidia.
In addition to sequences from the paratype of S. skorinae fresh material of asexual Spirographa growing on corticolous Ochrolechia sp. from Bolivia was included in our phylogenetic analysis for comparison. This specimen is characterized by the main conidial axis 8–12 μm long and up to 2 μm wide, and non-swollen arms about 4–5 μm long. Despite the host selection, this specimen cannot be attributed to S. intermedia, as the conidial measurements and shape are different from this species and closer to those of S. lichenicola. Based on its morphology, we provisionally call this specimen Spirographa cf. lichenicola but further phylogenetic analyses including sequences of both S. intermedia s. str. and S. lichenicola s. str. are needed to clarify its identity.
Based on maximum likelihood analysis of a combined ITS, LSU, SSU and rpb1 dataset, Spirographa skorinae is phylogenetically distinct from all currently sequenced Spirographa species with 99% MLBS support (Fig. 52).
Stictidaceae Fr. [as 'Stictei'], Summa veg. Scand., Sectio Post. (Stockholm): 345 (1849)
Stictidaceae was introduced based on apothecioid ascomata, cylindrical asci and filiform ascospores (Wei et al. 2021a, b). This family comprises 30 genera (Wijayawardene et al. 2022) which have diverse lifestyles various from saprophytes, pathogens, endophytes, lichens to fungicolous fungi (Wei et al. 2021a, b). Among them, saprophytic species predominate in Stictidaceae, and this lifestyle was inferred as the ancestral state of this family (Thiyagaraja et al. 2021) (Fig. 58).
[See PDF for image]
Fig. 58
Phylogram generated from maximum likelihood analysis based on combined ITS, LSU, mtSSU and rpb2 sequence data representing Stictidaceae. Ninety-five strains are included in the combined analyses which comprised 3173 characters (ITS: 583 bp, LSU: 822 bp, mtSSU 714 bp, rpb2: 1054 bp) after trimming. Cyanodermella viridula, C. asteris, C. banksiae and C. oleoligni were used as the outgroup taxa. The best scoring RAxML tree with a final likelihood value of − 37,878.808868 is presented. The matrix had 2000 distinct alignment patterns, with 46.06% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.282103, C = 0.211440, G = 0.255349, T = 0.251108; substitution rates: AC = 1.504293, AG = 3.287277, AT = 2.095138, CG = 0.854219, CT = 6.789877, GT = 1.000000; gamma distribution shape parameter α = 0.636561. Bootstrap support values equal to or greater than 60% (left side) and Bayesian posterior probabilities equal to or greater than 0.90 (right side) are given near the nodes. Ex-type strains are in bold, and the newly generated species are in blue
Fitzroyomyces pseudopandanicola S. C. He & D. P. Wei, sp. nov.
Index Fungorum number: IF902253; Facesoffungi number: FoF 16056; Fig. 59
[See PDF for image]
Fig. 59
Fitzroyomyces pseudopandanicola (HKAS 134947, holotype) a–c Apothecia on substrate. d, e Vertical section of apothecia. f, g Exciple. h Hymenium. i, j Asci and paraphyses. k, l Ascospores. m Ascus cap. n Upper and lower view of culture on PDA. Scale bars: d, e, g = 250 μm, f, h–j = 50 μm, k, l = 25 μm, m = 5 μm. (i, j, l, and m were treated with Melzer’s reagent)
Etymology: Phylogeny closely related to P. pandanicola, but morphology distinct.
Holotype: HKAS 134947
Saprobic on dead woody twig. Sexual morph: Apothecia 210–405 × 308–468 μm (x̄ = 271 × 382, n = 10), cupulate, completely immersed in the substrate when immature and opening by entire pore at maturity. Disc deeply immersed, creamy yellow, splitting away from the margin when dry. Exciple 35–65 (x̄ = 46, n = 20) μm, three-layered, with a prominent accessory thalline at the outermost layer, a wall of hyaline, thick-walled cells of textura angularis at the middle layer and a crystalliferous layer at the inner. Subhymenium 7.8–21 μm (x̄ = 15, n = 25), composed of hyaline cell of textura angularis, J−. Periphysoidal layer not observed. Paraphyses 1.8–3.7 μm (x̄ = 2.6, n = 20) in the wide, numerous, hyaline, filiform, apically branched. Asci 151–163 × 6.6–9 μm (x̄ = 157 × 7.6, n = 10), 8-spored, unitunicate, cylindrical, widest in the middle, rounded apex. Ascospores 128–147 × 2.2–2.6 μm (x̄ = 137 × 2.4, n = 10), hyaline, elongated cylindrical, slightly narrowing toward both ends, multiseptate, up to 20-septa. Asexual morph: Not observed.
Culture characteristics: Colony moderately growing on PDA media, reaching 2.1 cm after incubation for 30 days at 25 C. The colony was white to pale brown, raised, margin entire, mycelia dense, reverse brown.
Material examined: China, Yunnan Province, Xishuangbanna, on dead woody twig, 12 September 2022, De-Ping Wei, ST3 (HKAS 134947, Holotype), Ex-type, KUNCC 23–15,699.
GenBank numbers: ITS = OR769026, LSU = OR808083, mtSSU = PP832019, RPB2 = OR775514.
Notes: Fitzroyomyces pseudopandanicola phylogenetically sister to F. pandanicola with strong support (Fig. 54, 100% ML and 1.00 BYPP). Fitzroyomyces pandanicola originally was recognized as a member of Stictis by Tibpromma et al. (2018). Later this species was transferred to Fitzroyomyces by Wei et al. (2021a, b) based on multigene phylogenetic analysis. Fitzroyomyces pandanicola was found on dead leave of Pandanus sp. in Xihuangbanna District, Yunnan Province, China. Our species shares similar morphology with F. pandanicola in the cupulate apothecia with a large opening. However, F. pseudopandanicola can be distinguished from F. pandanicola in the smaller asci (151–163 × 6.6–9 μm vs. 160–240 × 7.5–23 μm) and ascospores (28–147 × 2.2–2.6 μm vs. 190–265 × 4–5 μm) as well as the association with dead twig. Additionally, the sequence comparisons show 24 bp (4.4%) differences in a 541 bp fragment of ITS between F. pseudopandanicola and F. pandanicola.
Phacidiella xishuangbannaensis D. P. Wei, sp. nov.
Index Fungorum number: IF902253; Facesoffungi number: FoF 16057 Fig. 60
[See PDF for image]
Fig. 60
Phacidiella xishuangbannaensis (HKAS 134946, holotype). a Apothecia on a dead twig. b, c Enlargement of apothecia. d, e Vertical section through apothecia. f Exceplium. g, h Asci. i Paraphyses. j–l Ascospores. m, n Apex of asci and paraphyses, respectively. o, p Upper and lower view of culture on PDA media. Scale bars: d, e = 200 μm, f–k = 50 μm, l–n = 10 μm. (j–n treated with Melzer’s reagent)
Etymology: The specific epithet is derived from Xishuangbanna District, Yunnan Province, China
Holotype: HKAS 134946
Saprobic on dead woody twig. Sexual morph: Apothecia 322–518 × 530–816 µm (x̄ = 426 × 632; n = 10), cupulate, with a lager opening, immersed, gregarious, with entire, white-pruinose margin. Disc creamy yellow, splitting away from the margin when dry. Exciple 70–160 µm (x̄ = 101; n = 15), consisted of three layers: (1) a accessory thalline margin, (2) a wall extended from subhymenium, of hyaline, thick-walled cell of textura angularis, (3) a crystalliferous layer. Periphysoidal layer not observed. Paraphyses 0.8–1.6 (x̄ = 1.1; n = 20), filiform, aseptate, apically branched, circinate, not enlarged. Asci 247–289 × 5.3–8 µm (x̄ = 260 × 6.2; n = 10), cylindrical, slightly narrow toward the apex and base, 8-spored, with a thick apex. Ascus cap 2.3–4.4 × 2.6–4.8 µm (x̄ = 3.5 × 4; n = 20), hemisphere, pierced by a pore, J−. Ascospores 138–252 × 1.2–2.4 µm (x̄ = 193 × 2; n = 30), filiform, hyaline, slightly acute at both ends, aseptate, guttulate when young, becoming multiseptate, smooth-walled with age, non-disarticulating, no contraction at septa, the interval cell 3.6–7.5 μm (x̄ = 5.4, n = 25) in length. Asexual morph: Not observed.
Culture characteristics: Culture was obtained from germinating ascospores. Colony on PDA reaching 16 mm diam. after 30 days at 25 °C, brown, slightly raised, mycelia dense, cottony, margin entire, surface dotted with brown liquid drops, reverse yellow–brown at periphery, dark brown at center.
Material examined: China, Yunnan Province, Xishuangbanna, on dead woody twig, 12 September 2022, De-Ping Wei, ST6 (HKAS 134946, holotype), ex-type, KUNCC 23-15698.
GenBank numbers: ITS = OR769025, LSU = OR808082, mtSSU = PP832018, RPB2 = OR775513.
Notes: The new species Phacidiella xishuangbannaensis clusters with P. alsophilae, P. podocarpi and P. kunmingensis, forming a monophyletic clade in Stictidaceae with 94% ML and 0.99 BYPP (Fig. 58). Phacidiella alsophilae (Crous et al. 2020a, b) and P. podocarpi (Crous et al. 2014) were known exclusively from asexual morphs, making it impossible to compare their morphologies with P. xishuangbannaensis. However, the pairwise comparison of nucleotide between P. xishuangbannaensis and P. alsophilae shows that there are 12 bp (512/527) and two bp (855/857) differences in ITS and LSU regions, respectively. The nucleotide comparison between P. xishuangbannaensis and P. podocarpi indicates that there are 54 bp (500/554, ITS) and 26 bp (770/796, LSU) differences. Phacidiella kunmingensis was introduced based on its sexual morph by Wei et al. (2022a, b). Phacidiella xishuangbannaensis has similar characteristics to P. kunmingensis in terms of the morphologies and dimensions of apothecia, exciple, paraphyses, asci and ascospores. However, the former differs from the latter species in 62 bp (503/565, ITS), 84 bp (869/953, LSU), 48 bp (642/690, mtSSU) and 155 bp (855/1010, RPB2).
Pertusariales M. Choisy ex D. Hawksw. & O.E. Erikss., Syst. Ascom. 5(1): 181 (1986)
Megasporaceae Lumbsch, in Lumbsch, Feige & Schmitz, J. Hattori bot. Lab. 75: 302 (1994)
Oxneriaria S.Y. Kondr. & Lőkös, in Haji et al., Acta bot. hung. 59(3–4): 355 (2017)
Notes: Oxneriaria comprises 12 species including Oxneriaria dendroplaca, O. haeyrenii, O. rivulicola, O. mashiginensis, O. supertegens, O. nikrapensis, O. permutata, O. verruculosa and O. virginea (Haji-Moniri et al. 2017). From Pakistan, three species of genus Oxneriaria are recently described i.e., Oxneriaria iqbalii, O. kohistaniensis and O. pakistanica (Ahmad et al. 1997; Iqbal et al. 2023; Zulfiqar et al. 2023). In this study, a new species Oxneriaria nigrodisca is introduced based on evidence of morphology, anatomy, chemical properties and phylogenetic placement (Fig. 61).
[See PDF for image]
Fig. 61
Phylogram generated from maximum likelihood analyses based on ITS sequence data representing genus Oxneriaria. Related sequences are taken from Nordin et al. (2011), Iqbal et al. (2023), Zulfiqar et al. (2023) and from the NCBI database (March 2023). Thirty sequences are included in the phylogenetic analyses which comprise 486 characters for ITS after alignment. Two sequences of Megaspora cretacea Gasparyan, Zakeri & Aptroot (KX253974, KX253975) were used as outgroup taxa. The aligned final dataset comprised 486 characters including gaps; out of these, 342 characters were conserved, 142 were variable, 138 were parsimony informative and 4 were singletons. Bootstrap values for maximum likelihood (ML) equal to or greater than 70% labelled on the nodes. The newly generated sequence is indicated in bold
Oxneriaria nigrodisca Usman & Khalid sp. nov.
Index Fungorum number: 901378; Facesoffungi number: FoF 15095; Fig. 62
[See PDF for image]
Fig. 62
Oxneriaria nigrodisca (LAH37201, holotype). a Fertile areoles with apothecia. b Sterile areoles with pycnidia. c Section of thallus. D Section of apothecium in lugol solution. e Pycnidia. f. Conidia. Scale bars: a–b = 1 mm, c = 50 µm, d = 200 µm, e = 50 µm, f = 10 µm
Etymology: The specific epithet nigrodisca refers to the black disc of apothecia.
Holotype: LAH37201
The most important features of Oxneriaria nigrodisca are epruinose thallus, indeterminate margins, bullate to angular apothecia with black disc, narrowly ellipsoid simple ascospores, bacillus conidia, hymenium initially bright orange then turns dark orange after lugol solution, atranorin and Norstictic acid detected. Thallus crustose, epilithic, areolate, epruinose, 0.3–1.5 mm in diam., indeterminate, no distinct margins, irregular 2–5 cm across, effuse, sterile areoles clay grey, fertile areoles pale grey. Prothallus white. Thallus heteromerous, upper cortex paraplectenchymatous, hyaline, 30–65 µm thick, cells 5–8 µm in diam. Algal layer discontinuous, 50–109 µm thick, photobiont Trebauxia sp, coccoid cells, globose to ellipsoid, 9–21 µm in diam. Medulla and lower cortex not differentiated and consist of paraplectenchymatous up to 150 µm thick, globose hyaline cells, 3–6 µm diam. Apothecia: scattered, without stipe, aspicilioid, epruinose, one apothecium per areole, sometime more than one apothecium fused, bullate to angular, 400–750 mm in diam, with black disc 300–550 mm in diam, dull, concave, often with depressions. Proper exciple 16–25 µm wide. Thalline exciple 109–130 µm wide. Epihymenium brown, 10–20 µm thick. Hymenium hyaline, 78–100 µm thick. Subhymenium hyaline, 30–50 µm thick. Hypothecium hyaline, 41–60 µm thick. Asci clavate, 8–spored, 52–71 × 16–23 µm (x̄ = 61.5 × 19.5 μm, n = 50). Ascospores simple, hyaline, narrowly ellipsoid, 13–20 × 7–10 µm (x̄ = 16.5 × 8.5 μm, n = 20). Paraphyses propsoplectanchymatus type elongated hyphae, septate, long and short cells, short cylindrical cells 3–7 × 1.5–2 µm (x̄ = 5 × 1.75 μm, n = 20)., with light brown terminal cell. Pycnidia lobaria type, globose to sub-globose, 70–120 µm, dark brown ostiole, bacillus conidia, 2.8–3.1 × 1 µm (x̄ = 2.95 × 1 μm, n = 30). hyaline.
Chemistry: K + ve turns orangish red, C + ve turns light green, KC + ve turns from orangish red to light orange red, UV + ve turns light green, hymenium initially bright orange then turns dark orange after lugol solution. Atranorin and Norstictic acid detected by TLC.
Ecology: Saxicolous, grows with other lichens.
Material examined: Pakistan. Gilgit Baltistan: Deosai National Park, (35° 0′ 30.21″ N, 75° 12′ 29.51″ E, 4,180 m a.s.l., on rocks, 13 May 2019, M. Usman DEO222 (LAH, holotype; LAH37201), Kharmang district, near Manthokha Waterfall, 35° 4′ 11.51″ N, 75° 59′ 47.89″ E, 2,450 m a.s.l., on rocks, 15 May 2019, M. Usman MAN21 (Paratype; LAH37416).
GenBank numbers: ITS: OR760502 (holotype), OR760503 (paratype).
Notes: Phylogenetically, the closest species to Oxneriaria nigrodisca is O. verruculosa forming a separate branch with a 100% bootstrap value. Phenotypically, it also resembles the same taxon by having areolate grey thallus, convex apothecia, 8-spored asci and ellipsoid ascospores. However, O. nigrodisca significantly differs from O. verruculosa in having epruinose thallus (vs. pruinose thallus), indeterminate margins with white prothallus (vs. subplacodioid margins with black prothallus), larger apothecia up to 0.75 mm wide. (vs. smaller apothecia up to 0.5 mm wide), epithecium brown (vs. brownish green to olive), hymenium I + orangish red (vs. I + blue), bacillus conidia 2.8–3.1 × 1 µm (vs. thread-like, straight conidia 16–20 × c. 0.5 µm) and thallus K + orangish red (vs. K + yellow turning red) (Nimis 2016). Furthermore, phylogenetic analyses indicate that O. nigrodisca forms a separate lineage in the clade with high statistical support (100ML Fig. 57). To provide the recommendation to justify our new species, we follow Haji-Moniri et al. (2017), Iqbal et al. (2023) and Zulfiqar et al. (2023). A comparison of 472 nucleotides of the ITS sequences between Oxneriaria nigrodisca (LAH37201) and O. verruculosa (EU057940) reveals 26 substitutions (5.5% nucleotide differences), which we believe should be sufficient to delineate our new species.
Leotiomycetes O.E. Erikss. & Winka, Myconet 1(1): 7 (1997)
Helotiaceae Rehm [as 'Helotieae'], Rabenh. Krypt. -Fl., Edn 2 (Leipzig) 1.3(lief. 37): 647 (1892) [1896]
Scytalidium Pesante, Annali Sper. agr., N.S. 11(2, Suppl.): cclxiv (1957)
Notes: the ascomycetous genus Scytalidium is a member of the family Chaetomiaceae within the class Leotiomycetes (Crous et al. 2023). Scytalidium species have also been reported from various habitats as soil, air, decaying plant material, and termite gut (Ogel et al. 2001). They have been known to play an important ecological role in breaking down organic matter, while some engage in parasitic or mutualistic relationships with other organisms. Scytalidium assmuthi, a newly isolated species, was isolated from the gut of an Odontotermes assmuthi termite (Fig. 63).
[See PDF for image]
Fig. 63
The placement of Scytalidium assmuthi using a maximum-likelihood (ML) analysis of the concatenated gene sequence of ITS and LSU (D1/D2 domain) rRNA region using the TNe + G4 chosen according to the Bayesian Information Criterion model in IQ-TREE v. 1.6.12 (Nguyen et al. 2015). The scale bar indicates the expected number of substitutions per site. The numbers provided on branches are frequencies with which a given branch appeared in 1000 bootstrap replications. The tree was rooted with Dermea acerina CBS 161.38 and Dermea cerasi MFLU 16-0929. The new species proposed in the present study is in bold and highlighted text
Scytalidium assmuthi G Mane, R Avchar, R Morey, R Sharma, sp. nov.
Index Fungorum number: IF901341; Facesoffungi number: 15056; Fig. 64
[See PDF for image]
Fig. 64
Scytalidium assmuthi (MCC 10102, holotype) a, b Tree bark with termite colony in Kolhapur, Maharashtra, India. c Host Odontotermes assmuthi sp.; d–h mycelium with melanized chlamydospores; dark brown hyphae with intercalary chlamydospores. i Long hyphae. j colony on PDA agar, Scale bars = 20 μm (hyphae and chlamydospores)
Etymology: refers to the host termite species Odontotermes assmuthi from which this novel species was isolated.
Holotype: MCC 10102
Scytalidium assmuthi was reported from the gut of the termite Odontotermes assmuthi feeding on wood logs from northern western ghat (India). Sexual morph: Not observed. Asexual morph: Mycelia consists of hyaline to sub-hyaline light brown, smooth-walled, branched, and septate hyphae (3.04–6.39 µm in width) without conidiophores were observed after 7 days on potato dextrose agar (PDA). Few light brown short septate intercalary chlamydospores were observed after 14 days of growth on PDA. Intercalary or terminal thick-walled smooth chlamydospores are globose to subglobose, or in chains of different sizes (7.3–13.22 µm diameter).
Culture characteristics: After 10 days colony size was 68 mm diameter at 25 °C on a PDA plate, circular, flat, creamish white, spreading with an entire margin with abundant aerial mycelium. However, after two weeks the colony color changes from creamish white to pinkish white and reverse cream white. This newly isolated fungi is positive for a few hydrolytic enzymes like amylase, cellulase, pectinase, and xylanase and negative for laccase.
Material examined:India, Maharashtra, Kolhapur district, 17.1324519° N, 73.8561159° E, from the gut of an Odontotermes assmuthi termite, 5 February 2021, collected by G Mane, (holotype MCC 10102 preserved in a metabolically inactive state at the National Centre for Microbial Resource (NCMR), formerly known as a Microbial Culture Collection (MCC), National Centre for Cell Science, Pune, India, PYCC 9837 ex-type, MycoBank MB 850092.
GenBank numbers: OR415885 (LSU); OR415883 (ITS)
Notes: Phylogenetic analysis based on the sequences of ITS rDNA and D1/D2 domain of the 28S rRNA gene supported the recognition of a new species in the genus Scytalidium and placed S. assmuthi close to S. circinatum. Additionally, newly isolated species are also distinct based on colony morphology, size, and shape of chlamydospores. The colonies of S. assmuthi on PDA are creamish white (6 D) to pinkish white (14 D), whereas S. circinatum colonies are pale to dark grey (6 D) or dark brown. New species also differentiated from closed-related species (S. circinatum) as it forms light brown globose to sub-globose intercalary or terminal chlamydospores whereas, S. circinatum produces irregularly shaped dematiaceous chlamydospores (Sigler and Wang 1990).
Vibrisseaceae Korf, Mycosystema 3: 23 (1990)
Apiculospora Wijayaw., Camporesi, A.J.L. Phillips & K.D. Hyde, Fungal Diver 77: 42 (2016)
Apiculospora was introduced by Wijayawardene et al. (2016) with A. spartii Wijayaw. et al. as the type species from dead branches of Spartium junceum L. Apiculospora belongs to Leotiomycetes, genera incertae sedis (Ekanayaka et al. 2019), while Hyde et al. (2020a, b, c) reassigned Apiculospora to Rhytismatales genera incertae sedis based on phylogenetics. Apiculospora penniseti was later introduced by Karunarathna et al. (2021) from dead leaves of Pennisetum purpureum. Apiculospora presently comprises only two species (www.indexfungorum.org, accessed in April 2024) and have been reported from only three hosts, Spartium junceum (Fabaceae; Italy), Pennisetum purpureum (Poaceae; Taiwan, China) and Yucca gigantea (Asparagaceae; China) (Fig. 65).
[See PDF for image]
Fig. 65
Maximum likelihood tree based on analysis of combined ITS, and LSU sequence data. The tree is rooted with Dactylaria dimorphospora (CBS 256.70). The newly generated sequence is in blue. Bootstrap support values of maximum likelihood and parsimony analysis equal to or greater than 80% are given near the nodes, respectively
Apiculospora spartii Wijayaw, Li WJ, Camporesi E, Phillips AJL & KD Hyde, in Wijayawardene et al., Fungal Diversity (2016)
Index Fungorum number: IF551762; Facesoffungi number: FoF 01426; Fig. 66
[See PDF for image]
Fig. 66
Apiculospora spartii (MFLU 15-0909, new host record). a, b Conidiomata on dead branches of Dorycnium hirsutum. c Longitudinal section of conidioma. d Different stages of developing conidia attach to conidiogenous cells. e–h Conidia. Scale bars: c, d = 50 μm, e–h = 20 μm
Saprobic on dead branches of Dorycnium hirsutum L. Sexual morph: Not observed. Asexual morph: Conidiomata 145–170 × 130–180 μm, (x̅ = 152 × 163 μm, n = 5), solitary, scattered, immersed, unilocular, subglobose, black. Peridium 18–26 μm, comprised of thick-walled, brown cells of textura angularis; inner cell layers thin-walled, almost reduced to a conidiogenesis region. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 4–10 × 1.5–3 μm, subcylindrical to ovoid, enteroblastic, with percurrent proliferation, hyaline, smooth walled. Conidia 16–26 × 5–11 μm (x̅ = 23 × 9 μm, n = 30), subcylindrical to ellipsoid, slightly curved, conical at apex, aseptate when immature and later become 1-septate, sometimes with a dark band at septum, occasionally constricted at septum, hyaline to pale brown when immature and later dark brown, guttulate, thick walled, non-mucilaginous.
Culture characteristics: Culture on PDA, colonies slow growing, circular, spreading, flattened, flossy, smooth with entire edge, brown, reverse brown.
Material examined: Italy, dead branches of Dorycnium hirsutum (Fabaceae), 22 November 2013, E. Camporesi, IT 1533 (MFLU 15-0909); living culture MFLUCC 13-0246.
GenBank numbers: LSU: PP762101; ITS: PP762097; SSU: PP762148.
Notes: This isolate obtained from Italy fits with the morphological characteristics of Apiculospora comprising subcylindrical to ellipsoid conidia with a dark band at septum (Wijayawardene et al. 2016, 2021; Ekanayaka et al. 2019; Karunarathna et al. 2021). In the phylogenetic tree, our isolate clusters together with A. spartii (MFLU 15-3556, MFLU 18-1812, MFLU 18-1813) with 100% ML bootstrap (Fig. 61). The isolate obtained in this study is morphologically similar to the type species of A. spartii but differs by the size of conidia (Wijayawardene et al. 2016). ITS and LSU sequences did not show any base pair differences suggesting that our isolate is identical to the type species of A. spartii. As A. spartii has been isolated from a dead branch of Spartium junceum from Italy (Wijayawardene et al. 2016; Ekanayaka et al. 2019) and Yucca gigantea from China (Wijayawardene et al. 2021), here We introduce A. spartii as a new host record from Dorycnium hirsutum from Italy.
Orbiliomycetes O.E. Erikss. & Baral, in Eriksson, Baral, Currah & Hansen, Myconet 9: 96 (2003)
Orbiliales Baral, O.E. Erikss., G. Marson & E. Weber, in Eriksson, Baral, Currah & Hansen, Myconet 9: 96 (2003)
Orbiliaceae Nannf., Nova Acta R. Soc. Scient. upsal., Ser. 4 8(no. 2): 250 (1932)
Notes: Orbiliaceae was established by Nannfeldt in 1932 and initially classified under the order Helotiales (Nannfeldt 1932). According to molecular phylogenetic analyses and the identification of specific nucleotides in the SSU rDNA sequence data, Orbiliaceae was reclassified into the order Orbiliales, this reclassification revealed it the restricted family within this order (Eriksson et al. 2003; Pfister 1997). Members of Orbiliaceae are a group of small cup fungi characterized by a high dioptric and morphologically diverse (spherical, teardrop-shaped, ellipsoid, or vermiform) ascospores, unitunicate asci, brightly coloured, small, waxy, and translucent apothecium (Nannfeldt 1932). Currently, Orbiliaceae comprises three sexual genera: Hyalorbilia Baral & G. Morson (Baral et al. 2018; Eriksson et al. 2003), Orbilia Fries (Fries 1849), and Pseudorbilia Zhang (Zhang et al. 2007), as well as at least ten asexual genera including Arthrobotrys Corda, Dactylellina Morelet, Drechslerella Subram, Dactylella Grove, and Dicranodion Harkness, etc. (Baral et al. 2018). Initially, this fungal group was considered an insignificant group due to their unclear ecological and application values. It was not until Pfister (1997) reported that the anamorph of Orbilia fimicola is Arthrobotrys species, a group of nematode-trapping fungi (NTF) with significant research, application, and ecological stability value, that this group of fungi gained attention from mycologists.
Arthrobotrys Corda, Pracht-Fl. Eur. Schimmelbild.: 43 (1839)
Notes: Arthrobotrys is the largest genus of nematode-trapping fungi in Orbiliaceae. It was established by Corda (1839), with A. superba Corda as the type species (Corda 1839). These taxa are originally characterized by regularly 1-septate conidia growing on the nodes or short denticles of conidiophores (Corda 1839). Subsequently, after systematic comparative morphological studies, scholars redefined the characteristics of the genus Arthrobotrys as follows: branched or simple conidiophores; obovoid, elliptic, pyriform, 0–3-septate conidia, growing asyn-chronously on the nodes or on short denticles of conidiophores; and including species that capture nematodes with adhesive networks, constricting rings and adhesive knobs (Cooke and Dickinson 1965). With development of the molecular biology techniques, the main characteristic of Arthrobotrys is now considered to be capturing nematodes with adhesive networks (Ahrén et al. 1998; Liou & Tzean 1997; Scholler et al. 1999; Zhang and Hyde 2014). At present, 144 records of Arthrobotrys are listed in the Species Fungorum (http://www.speciesfungorum.org; accessed on 29 December 2023), which represent 78 accepted species (Fig. 67).
Arthrobotrys tachengensis F. Zhang & X.Y. Yang sp. nov.
Index Fungorum number: IF902261; Facesoffungi number: FOF 15970; (Fig. 68)
Etymology: The species name “tachengensis” refers to the name of the sample collection site: Tacheng County, Diqing Tibetan Nationality, Yunnan Province, China.
Holotype: DLUCC179
Saprobic or capture nematodes on soil. Sexual morph Not observed. Asexual morph:Colonies on PDA white, cottony, growing rapidly, reaching 50 mm diameter after 10 days in the incubator at 26 °C. Mycelium partly superficial, partly immersed, composed of septate, branched, smooth, hyaline. Conidiophores 155–345 µm (x̅ = 226.9 µm, n = 50) long, 3–6.5 µm (x̅ = 4.8 µm, n = 50) wide at the base, gradually tapering upwards to the apex with 2–3.5 µm (x̅ = 2.7 µm, n = 50) wide, erect, septate, hyaline, unbranched or sometimes produce long branches, each branch producing 2–4 short denticles at the apex, each short denticle bearing a single conidium. Conidia 23.5–45.5 × 8.5–17.5 µm (x̅ = 34.6 × 13.4 μm, n = 50), smooth and hyaline, rounded at the apex and truncated at the base; immature conidia drop-shaped, obovate, 1–2 septate, with a larger cell at the apex; mature conidia subfusiform, 2–3-septate (mostly 3-septate, 1 septa at the apex and 1–2 septa at the base), with a larger cell at the middle. Conidia germinate from the small end cells, and the larger cell not germinates. Chlamydospores 6–21.5 × 5.5–12.5 µm (x̅ = 12.9 × 8.7 μm, n = 50), cylindrical, ellipsoidal, in chains, hyaline. Capturing nematodes with adhesive networks.
Material examined: China, Yunnan Province, Diqing Tibetan Nationality, Tacheng County, N 27° 34′ 54.02″, E 99° 31′ 18.32″, from freshwater sediment, 11 May 2014, F. Zhang, DLU 89–1 (DLUCC179, holotype).
GenBnak numbers: ITS = pp868283, tef1-α = pp869332, rpb2 = pp869333
Notes: The phylogenetic analyses revealed that Arthrobotrys tachengensis is a sister to A. indica but lacking in statistical support (Fig. 67). Although A. tachengensis and A. indica differ only 2.3% (10/426 bp) in ITS sequence, they differ significantly in morphology, with A. tachengensis producing obovate and subfusiform 1–3 septa conidia and A. indica producing obovate 0–2 septa conidia (Zhang & Hyde 2014, Chowdhry & Bahl 1982). Morphologically, A. tachengensis is more similar to A. microscaphoides and A. mangrovispora, there are 4.2% (25/595 bp) and 10.9% (65/598 bp) differences between A. tachengensis and A. microscaphoides and A. mangrovispora on ITS, respectively. In addition, the conidia of A. tachengensis are significantly larger than A. microscaphoides and smaller than A. mangrovispora [A. tachengensis, 23.5–45.5 (34.6) × 8.5–17.5 (13.4) µm versus A. microscaphoides, 22.5–45 (27.2) × 10–20 (13.9) µm versus A. mangrovispora 25–50 (38.9) × 12–24 (17.3) µm] (Zhang & Hyde 2014; Swe et al. 2008; Liu & Lu 1993).
[See PDF for image]
Fig. 67
Maximum likelihood tree based on combined sequences (ITS = 627 bp, tef1-α = 542 bp and rpb2 = 822 bp, total 2038 characters, constant sites = 900 bp, variable sites = 1087 bp, parsimony informative sites = 886) from 67 species of Orbiliaceae nematode-trapping fungi (including 59 Arthrobotrys species, 4 Dactylellina species and 4 Drechslerella species). The tree is rooted by Vermispora fusarina (YXJ13-5) and V. leguminacea (CGMCC 6.0291). The best-scoring maximum likelihood tree was performed with a final ML optimization likelihood value of − 6198.356892. Bootstrap support values for maximum likelihood (black) equal or greater than 70% and Bayesian posterior probabilities values (red) equal or greater than 0.90 are indicated above the nodes. The new isolates are in blue, type strains are in bold
Hyalorbilia Baral & G. Marson, Micologia 2000 (Trento): 44 (2001)
Hyalorbilia was introduced to accommodate five species which were removed from Orbilia based on the distinct morphological features (Baral and Marson 2000). The separated phylogenetic evidence between Orbilia and Hyalorbilia was provided by Liu et al. (Liu et al. 2006). Hyalorbilia berberidis was indicated as the type species and a total of 44 species were listed in MycoBank (2022). Hyalorbilia is characterized by pale-coloured and translucent apothecia, ectal excipulum cells horizontally oriented, hymenial surface covered with a layer of gelatinous elements and multiple warts, asci apex hemispheric with thin wall and arising from croziers and ascospore usually with a homopolar guttule. Hyalorbilia yunnanensis is described here based on the morphological illustration and phylogenetic evidence (Fig. 69).
[See PDF for image]
Fig. 68
Arthrobotrys tachengensis (DLUCC179, holotype). a Colony. b Immature conidia. d Mature conidia. e Trapping-device: adhesive networks. f Germinating conidia. g Chlamydospores. c, h Conidiophores. Scale bars: a = 1 cm, b, d, e–h = 20 µm, c = 50 µ
[See PDF for image]
Fig. 69
Phylogram generated from ML analysis based on ITS sequence data of the genus Mycoceros. 29 strains are included in the analysis which comprised 700 characters for ITS. The tree topology of the maximum likelihood analysis is similar to the Bayesian analysis. The best RaxML tree with a final likelihood value of − 18,195.465988 is presented. The matrix had 541 distinct alignment patterns, with 12.77% of undetermined characters or gaps. Evolutionary models applied for the ITS gene are GTR + I + G. Bootstrap support values for ML greater than 70% and Bayesian posterior probabilities greater than 0.90 are given near nodes, respectively. The tree was rooted with Mycoceros antennatissimus (BP 105172). Ex-type strains are in bold. The newly generated sequences are indicated in blue
Hyalorbilia yunnanensis C.J.Y. Li & K.D. Hyde sp. nov.
Index Fungorum number: IF 902392; Facesoffungi number: FoF 16058 Fig. 70
[See PDF for image]
Fig. 70
Hyalorbilia yunnanensis (HKAS 126582, holotype). a Fresh ascomata on natural habitat. b Vertical section of ascomata. c Excipulum. d Anchoring hyphae at apothecial base. e Paraphyses. f–j Asci (f–i Asci in phloxine reagent). k–p Ascospores (k–n Ascospore in phloxine reagent). Scale bar: b = 800 μm, c = 130 μm, d = 25 μm, e–j = 15 μm, k–p = 5 μm
Etymology: named according to the geographical origin, Yunnan province.
Holotype: HKAS 126582
Saprobic on decayed wood. Apothecia fresh 1.5–3 mm diam., dried 0.21–0.24 mm high (x̅ = 1.97 × 0.23 mm, n = 23), scattered or gregarious, superficial, flatted or inverted discoid, pale cream yellow when fresh, yellowish orange when dry, translucent. Disc flat and circular, smooth, margin thin; almost sessile. Excipulum 140–170 μm, hyaline, wall thin, at base comprised of textura angularis cells, 14–17.9 × 11.3–15.4 μm (x̅ = 16.3 × 13.1 mm, n = 20), at middle comprised of large, loosely disordered textura angularis-prismatica cells, 20.8–29.6 × 14.9–21.2 μm (x̅ = 24.3 × 17.5 mm, n = 20), at flank of horizontally oriented textura prismatica, 15.7–26.7 × 8–11.2 μm (x̅ = 22.2 × 9.5 mm, n = 20). Hymenium 31–35 μm, hyaline. Paraphyses 1.4–1.7 μm wide at tips, filiform, hyaline, unbranched, aseptate, scarcely extending beyond the asci. Asci 26–33 × 3.1–5.4 μm (x̅ = 29.2 × 4.4 μm, n = 30), 8–spored, short clavate, hemispherical apex, short and thick stalk arising from croziers. Ascospore (1/1/4) (7.9–)8–9.7(–10.1) × (1–)1.3–1.9(–2) μm (x̅ = 8.8 × 1.6 μm, n = 50), Q = (4.3)4.6–7(–8.9) μm, Qm = 5.67 ± 0.9 μm, overlapping uniseriate to biseriate, clavate with narrowly rounded ends, thin-walled, smooth, aseptate or 1-septate. Anchoring hyphae very abundant, 3.6–5.2 μm wide, wall thin, radiating, twining, multi-septate. Exudate 0.3–0.5 μm thick, covering asci and paraphyses, continuous, smooth, pale yellow. Asexual morph: not observed.
Material examined: China, Yunnan Province, Tengchong City, altitude 1550 m, on decayed wood, 20 August 2022, Cuijinyi Li, LCJY-1129 (HKAS 126582, holotype).
GenBank numbers: ITS: OQ158979
Notes: The phylogenetic analysis of ITS data and typical morphological features indicates that our collection belongs to Hyalorbilia. The ITS region of Hyalorbilia yunnanensis (HKAS 126582) is similar to Hyalorbilia juliae (strain 6449) (449/506 with 25 gaps) and Hyalorbilia citrina (H.B. 8012, C.L. 5247) (349/414 with 29 gaps). Hyalorbilia yunnanensis formed an individual clade between Hyalorbilia arcuate (strain H.B. 8578b) and Hyalorbilia citrina (H.B. 8012, C.L. 5247) with 84% ML bootstrap and 0.99 Bayesian probability.
Hyalorbilia yunnanensis is mainly characterised by typically pale cream yellow and translucent apothecia, horizontally oriented prismatica cells and large, loosely disordered angularis to prismatica cells, low thickness proportion of hymenium, thin paraphyses, short and broad asci with long ascospores. Morphologically, Hyalorbilia yunnanensis is similar to Hyalorbilia polypori in vertical section features (hymenium and excipulum) and short broad asci, but the latter distinct Hyalorbilia yunnanensis by having shorter ascospores (8–9.7 × 1.3–1.9 μm vs. 5.5–8.5 × 1.5–1.9 μm) and septate paraphyses (Baral et al. 2020). Hyalorbilia arcuata and Hyalorbilia citrina were sister to Hyalorbilia yunnanensis based on the phylogenetic analysis but can be distinct by curved ascospores (Baral et al. 2020).
Sordariomycetes O.E. Erikss. & Winka
Notes: Sordariomycetes are a globally dispersed, morphologically diverse class. They are important as saprobes, endophytes, plant, and human diseases, and fungicolous taxa Maharachchikumbura et al. (2015), (Chen et al. 2023). For taxonomic treatments, we follow Hyde et al. (2020a, b, c) and Wijayawardene et al. (2022).
Subclass Diaporthomycetidae Senan., Maharachch. & K.D. Hyde.
Diaporthales Nannf.
Notes: For taxonomic treatments, we follow Senanayake et al. (2017), Hyde et al. (2020a, b, c) and Wijayawardene et al. (2022).
Diaporthaceae Höhn. ex Wehm., Am. J. Bot. 13: 638 (1926)
Notes: In the recent outline of fungi by Wijayawardene et al. (2022), 15 genera are accepted in Diaporthaceae.
Diaporthe Fuckel, Fungi rhenani exsic., suppl., fasc. 5: no. 1988 (1867)
Notes: Diaporthe comprises fungi which are endophytes, pathogens and saprophytes in a wide range of hosts (Manawasinghe et al. 2019). They are pathogenic on economically important crops causing devastating damage worldwide. For the recent taxonomic treatment of this genus, we follow Norphanphoun et al. (2022), Hongsanan et al. (2023), and Pereira et al. (2023) (Fig. 71).
[See PDF for image]
Fig. 71
Phylogram generated from maximum likelihood analysis based on combined ITS, tef1-α, tub, cal and his sequence data. 142 taxa were included in the combined analyses, which comprised 2151 characters in the alignment. Bootstrap support values equal to or greater than for ML ≥ 60% and BYPP ≥ 0.95 are given above the nodes. The tree is rooted with Diaporthe sojae (FAU635) and D. actinidiae (ICMP 13683). The ex–type strains are indicated in bold. The newly generated sequence is indicated in red
Diaporthe beijingensis Y.Y. Zhou, W. Zhang & J.Y. Yan. sp. nov.
Index Fungorum number: IF902257; Facesoffungi number: FoF 15928 Fig. 72
[See PDF for image]
Fig. 72
Diaporthe beijingensis (JZB320260, holotype). a Surface, b Reverse view of 4-days old colony on PDA. c Ascomata. d Asci. e Ascospores. f Conidiomata. g Conidia. Scale bars: c = 20 μm; d,e = 10 μm; f = 100 μm; g = 10 μm
Etymology: ‘beijingensis’ refers to Beijing in China from which it was isolated.
Associated with leaf spots of Prunus avium living leaves. Sexual morph: Produced on 30 days old PDA culture. Perithecia 220–397 μm diam., subglobose, black, solitary to scattered, ostiolate. Perithecial necks 38–103 μm., black, subcylindrical, tapering towards the apex. Asci 25.6–37.2 × 5–8 μm (x̅ = 31.8 × 6.7 μm, n = 30), unitunicate, 8-spored, sessile, ellipsoid to clavate, widest at centre and rounded towards the apices. Ascospores 5.6–8.5 × 2–3 μm (x̅ = 6.5 × 2.6 μm, n = 50), hyaline, often biguttulate, ellipsoid to fusoid, Asexual morph:Conidiomata 180–350 μm diam., pycnidial, globose or irregular, scattered, orange cream conidial drops exuded from the ostioles. Alpha conidia 5–7.5 × 1.5–3 μm (x̅ = 6.9 × 2.7 μm, n = 60), hyaline, biguttulate, fusoid to ellipsoid, Beta and Gamma conidia not observed.
Culture characteristics: Colonies on PDA entirely white both on the surface and reverse. Reverse becomes amber with time. Aerial mycelium felty, colonies reaching 75 mm diam., after 4 days in room temperature.
Material examined: China, Beijing, on leaf spot of Prunus avium L. (Rosaceae), 27 September 2022, Yueyan Zhou and Wei Zhang 22-HD-2-5a (JZBH320260, holotype), ex-type JZB320260.
GenBank numbers: JZB320260: ITS: PP859267; tef1-α: PP869043; tub2: PP869034; cal: PP869037; his: PP869040. JZB320261: ITS: PP859268; tef1-α: PP869044; tub2: PP869035; cal: PP869038; his: PP869041. JZB320262: ITS: PP859269; tef1-α: PP869045; tub2: PP869036; cal: PP869039; his: PP869042.
Notes: Diaporthe beijingensis forms a distinct clade within Diaporthe and is sister to D. caulivora. It forms an independent branch with 100% ML support. The morphological characters of our new collection are similar to D. caulivora (CBS 127268, ex-neotype). However, the base pair showed 2.1% (11/536) differences in ITS, 4.4% (15/343) in tef1-α, 1.6% (7/429 bp) in act, 2.2% (8/356) in cal, and 3.2% (15/475) in his between D. beijingensis (JZB320260) and D. caulivora (CBS 127268). Therefore, we introduce Diaporthe beijingensis as a new species.
[See PDF for image]
Fig. 73
Phylogenetic tree of Racheliella species based on maximum likelihood (ML) analyses of the combined DNA dataset of ITS, tef1-α and tub gene sequences. The ML bootstrap support values ≥ 75% and BYPP higher than 0.95 are indicated above the nodes and branches. The scale bar indicates 0.05 changes per site. Ex-type strains are marked in bold. Isolates of novel taxa are in red in this study. The tree is rooted with Melanconis groenlandica (CBS 143669)
Tubakiaceae U. Braun, J.Z. Groenew. & Crous, in Braun et al., Fungal Systematics and Evolution 1: 62 (2018).
Notes: Based on morphology and phylogeny, Braun et al. (2018) proposed to accept eight genera in the Tubakiaceae family, including the type genus Tubakia. For the taxonomic treatments of this family, we follow Wijayawardene et al. (2022).
Racheliella Crous & U. Braun, in Braun et al., Fungal Systematics and Evolution 1: 69 (2018)
Notes: The genus was introduced by Broun et al. (2018) with Racheliella wingfieldiana as the type species. Currently, there are two species accepted in this genus (Fig. 73).
Racheliella chinensis Y.X. Zhang, J.Y. Lin & Manawas., sp. nov.
Index Fungorum number: IF 902393; Facesoffungi number: FoF 16059; Fig. 74
[See PDF for image]
Fig. 74
Racheliella chinensis (ZHKUCC 22-0330, holotype) a Colonies on MEA (7 d). b Colonies on MEA (14 d). c, d Conidiogenous cells. e, f Conidia. Scale bars: c–d = 10 μm
Etymology: Referring to the country name China
Holotype: MHZU 22-0172
Associated with leaf spot of Livistona chinensis. Sexual morph: not observed. Asexual morph: Produced on 14 days old PDA culture. Conidiophores often reduced to conidiogenous cells. Conidiogenous cells 5–14 × 1–3.5 (x̅ = 9.1 × 2.3, n = 50), hyaline, phialidic to doliform, thin-walled, smooth. Conidia 8–15 × 3–5(x̅ = 11.6 × 4.1, n = 50), hyaline, aseptate, mostly ovoid, occasionally fusiform, with a rounded apex and narrowly truncate to pointed base, thin-walled, smooth.
Culture characteristics: Colonies growing slowly at 25 °C on MEA, circular, fluffy, with regular edge, aerial mycelium initially white, gradually becoming grey after 14 days, reverse initially light brown, becoming dark brown after 14 days.
Material examined: China, Guangdong Province, Shenzhen, on leaves of Livistona chinensis (Palmae), 30 November 2020, C.T. Chen (MHZU 22-0172, holotype), ex-type culture ZHKUCC 22-0330.
GenBank number: ITS=PP851442, PP849415, tef1-α=PQ219689, PQ219690, tub2=PQ219691, PQ219692
Notes: In the phylogeny (Fig. 73), two isolates from Racheliella chinensis clustered into a distinct clade with 0.99 in BYPP support sister to R. wingfieldiana. Our isolate differs from R. wingfieldiana and R. saprophytica in having narrow conidia (8–15 × 3–5 vs. 12–14 × 7 vs currently 46 accepted species 9–15 × 5–6 μm). The nucleotide of our strain (ZHKUCC 22-0330) differs from R. wingfieldiana (CBS 143669) by 17.53% (102/582) variations in ITS, 34.40% (218/623) variations in tef 1-α and 24.47% (127/519) variations in tub2, differs from R. saprophytica (NTCL 052-1) by 10.09% (54/535) variations in ITS (tef 1-α and tub2 not available in R. saprophytica). Thus, we introduced Racheliella chinensis as a new species based on morphology and phylogenetic evidence.
Distoseptisporales ZL Luo, KD Hyde & HY Su, in Luo et al., Fungal Diversity [32] (2019)Notes: For taxonomic treatments, we follow, Hyde et al. (2020a, b, c) and Wijayawardene et al. (2022).
Distoseptisporaceae ZL Luo, KD Hyde & HY Su, in Luo et al., Fungal Diversity [32] (2019)
Notes: Distoseptisporaceae is a monotypic family established by Su et al. (2016) with a single genus, Distoseptispora. The family accommodates 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 relatively short conidiophores (Su et al. 2016). Yang et al. (2018) provided an emendation of Distoseptispora according to much longer, percurrently proliferating conidiophores and euseptate conidia. Yang et al. (2021) recently proposed a sexual morph of Distoseptispora for the first time in the family.
Distoseptispora K.D. Hyde, McKenzie & Maharachch., in Su et al., Fungal Diversity 80: 402 (2016)
Notes: Distoseptispora was established by Su et al. (2016) with Distoseptispora fluminicola as the type species. The asexual morph is hyphomycetes in nature with macronematous conidiophores, percurrent, elongate conidiogenous cells, olivaceous, brown, yellowish, or reddish brown, euseptate or distoseptate conidia, and rarely muriform conidia (Su et al. 2016; Yang et al. 2018, 2021). The sexual morph is characterized by being solitary or gregarious, immersed to semi-immersed, perithecial, subglobose to ellipsoidal, ostiolate, dark brown ascomata with a short neck and hyaline, 0–3-septate, ascospores with a mucilaginous sheath (Yang et al. 2021). There are currently 46 accepted species in the genus (Su et al. 2016; Dong et al. 2021; Yang et al. 2021; Konta et al. 2023; Shen et al. 2024; Karimi et al. 2024; Index Fungorum 2024) (Fig. 75).
[See PDF for image]
Fig. 75
Phylogram generated from the best scoring of the RAxML tree based on combined ITS, LSU, tef1-α and rpb2 sequence dataset to indicate the new species Distoseptispora chiangraiensis and related family in Distoseptisporaceae. Ninety-nine strains are included in the combined analyses which comprise a total of 2909 characters. Pseudistanjehughesia lignicola (MFLUCC 15-0352) and P. aquitropica (MFLUCC 16-0569) are selected as the outgroup taxa. The best RAxML tree with a final likelihood value of − 33,810.150068 is presented. RAxML analysis yielded 1410 distinct alignment patterns and 25.64% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.234011, C = 0.270143, G = 0.284578, T = 0.211268, with substitution rates AC = 1.186820, AG = 3.015581, AT = 1.143201, CG = 0.749208, CT = 6.564047, GT = 1.000000; gamma distribution shape parameter alpha = 0.206169. Bootstrap support values for maximum likelihood (MLBS, left) equal to or greater than 70% are given above the nodes. Bayesian posterior probabilities (BYPP, right) equal to or greater than 0.95 are given above the nodes. Newly generated sequences are indicated in red bold
Distoseptispora adscendens (Berk.) R. Zhu & H. Zhang, Journal of Fungi 1063: 8 (2022)
Index Fungorum number: IF559917; Facesoffungi number: FoF12574; Fig. 76
[See PDF for image]
Fig. 76
Distoseptispora adscendens (NFCCI 10633 new record) a, b Colonies on natural substrate. c conidia with conidiophores. d–g conidia. h colony on PDA (obverse). i colony on PDA (reverse). Scale bars: a–b = 100 μm, c–g = 20 μm
Saprobic on unidentified decaying stems in terrestrial habitat. Sexual morph: Not observed. Asexual morph: Colonies effuse, scattered, hairy, brown to dark brown. Mycelium mostly immersed, composed of branched, septate, brown, smooth hyphae. Conidiophores macronematous, mononematous, brown to dark brown, solitary, 1–4 septate, erect, straight or slightly flexuous, unbranched, smooth, cylindrical, 22–48 μm × 6–10 μm (=34 8 μm, n = 15). Conidiogenous cells holoblastic, monoblastic, integrated, terminal, determinate, brown, cylindrical, 5.5–10 μm × 3–5 μm. Conidia 130–360 μm × 13.8–18 μm (=142 12 μm, n = 15), acrogenous, solitary, obclavate, rostrate, slightly curved, 17– 60 distoseptate, rounded apexes 5–10 μm wide and subcylindrical to conical–truncate basal cells 5–7 μm wide at the base.
Cultural characteristics: conidia germinated on PDA within 24 h. Colonies on PDA reaching 15–20 mm diam., after 2 weeks at 25 °C, circular, with fluffy, dense, dark olivaceous brown aerial mycelium on the surface, umbonate at centre; in reverse dark brown with entire margin.
Material examined: India, Karnataka, Madikeri (12°26′11″N, 75°43′40″E), on an unidentified decaying stem, 04 March 2022, Sruthi O. P. and Rajeshkumar K. C., AMH 10633, living culture NFCCI 5692.
GenBank numbers: NFCCI 5692: ITS = OR807983, LSU = OR807984, tef1-α = OR824935.
Notes: Based on a phylogenetic study, Su et al. (2016) accepted Ellisembia adscendens, which was first introduced as Sporidesmium adscendens under Distoseptispora. Yang et al. (2021) introduced Distoseptispora adscendens as a new combination with legitimacy. Our isolate NFCCI 5692, clusters within Distoseptispora adscendens with a 72% ML and 0.99 BYPP support (Fig. 75). Furthermore, the morphology of our isolate resembles the description and illustrations for Distoseptispora adscendens provided by Wu and Zhuang (2005). Morphological characteristics and key distinguishing features of our strain are identical with the Distoseptispora adscendens. Phylogeny using ITS, LSU and tef1-α delineated our strain allied to Distoseptispora adscendens HKUCC 10820 with moderate support. Distoseptispora adscendens is a new geographical record from India (Fig. 76).
Distoseptispora chiangraiensis R.J. Xu, Q. Zhao & K.D. Hyde sp. nov.
Index Fungorum number: IF 902394; Facesoffungi number: FoF 16060 Fig. 77
[See PDF for image]
Fig. 77
Distoseptispora chiangraiensis (MFLU 21–0105, holotype) a, b Colony on natural substrate. c–e Conidiophore and conidia. f Conidiogenous cell. g Conidiophore. h–j Conidia. k Germinated conidium. l Culture on MEA medium. Scale bars: f = 50 μm, g = 20 μm, c–k = 50 μm
Etymology: Referring to the collection site from Chiang Rai Province in Thailand.
Holotype: MFLU 21-0105
Saprobic on submerged decaying wood in a freshwater stream. Sexual morph: Not observed. Asexual morph:Colonies effuse, olivaceous or dark brown, hairy or velvety, glistening. Mycelium partly superficial, partly immersed in the substrate, consisting of branched, septate, smooth, subhyaline to pale brown hyphae. Conidiophores 35–64 × 4.0–6 μm ( = 51 × 5 μm, n = 16), macronematous, mononematous, unbranched, single or in groups of 2 or 3, erect, cylindrical, straight, or slightly flexuous, brown to olivaceous, 3–4-septate, smooth, truncate at the apex. Conidiogenous cells monoblastic, integrated, terminal, olivaceous or brown, cylindrical, sometimes with percurrent proliferation. Conidia 68–172 × 6.5–11 μm ( = 114 × 9 μm, n = 25), acrogenous, solitary, elongate, tapering towards at the apex, obclavate, rostrate, truncate at base, rounded at apex, 15–25-distoseptate, slightly constricted at septa, thick-walled, olivaceous to dark brown, pale brown to subhyaline towards the apex, smooth, guttulate, sometimes with percurrent proliferation which forms another conidium from the conidial apex.
Culture characteristics: Conidia germinated on MEA within 48 h and germ tubes from both ends. Colony reached 26 mm at 28 °C for 3 weeks, on MEA media, circular, flat, surface villiform, dense, dark brown mycelium in the center, celadon to grey from above, atrovirens from below, becoming sparse and paler at the entire margin, edge entire.
Material examined: Thailand, Chiang Rai Province, Nang Lae, Mueang, (99° 52′ 52.93″ E, 20° 3′ 2.52″ N), saprobic on decaying bamboo culms, submerged in a freshwater stream, 18 July 2020, R.J Xu, MD-70C (MFLU 21-0105, holotype), ex-type living culture (T20-1126). saprobic submerged decaying wood of an unidentified plant in a freshwater stream, 18 July 2020, R.J Xu, MD-70D (HKAS 115808, isotype), living culture, MFLUCC 24-0112, KUNCC 10443.
GenBank numbers: MFLU 21-0105: ITS = MZ890145, LSU = MZ890139, SSU = MZ890181, tef1-α = MZ892970. KUNCC 10443: ITS = MZ890146, LSU = MZ890140, SSU = MZ890182, tef1-α = MZ892971
Notes: In the phylogenetic analyses, our isolates developed a sister clade with D. bangkokensis (MFLUCC 18-0262) with 98% ML bootstrap and 1.00 BYPP statistical support. Morphologically, D. chiangraiensis resembles D. bambusae and D. obclavata in having macronematous, cylindrical, brown to olivaceous, septate conidiophores and obclavate, olivaceous or brown conidia (Sun et al. 2020; Luo et al. 2019). However, D. chiangraiensis can be distinguished from D. bambusae in having elongate, longer conidia (68–172 μm vs. 45–74 μm). Distoseptispora chiangraiensis differs from D. obclavata in having shorter conidiophores (35–64 × 4.0–6 μm vs. 117.5–162.5 × 5–7 μm), and longer conidia (68–172 × 6.5–11 μm vs. 46–66 × 9–11 μm) (Luo et al. 2019; Sun et al. 2020; Sheng et al. 2021). Distoseptispora chiangraiensis resembles D. bangkokensis in having macronematous, mononematous, cylindrical conidiophores, terminal, monoblastic, cylindrical conidiogenous cells, and acrogenous, solitary, elongate, multi-distoseptate conidia (Shen et al. 2021). However, it differs in having shorter conidia (68–172 × 6.5–11 μm vs. 400–568 × 13–16 μm). Additionally, comparisons of ITS sequences demonstrate a 2.9% (15/522 bp, excluding gaps) difference between D. chiangraiensis and D. bangkokensis Jeewon and Hyde (2016). Therefore, D. chiangraiensis was identified as a new species supported by both morphological and phylogenetic evidences.
Diaporthomycetidae families incertae sedis
Papulosaceae Winka & O.E. Erikss., Mycoscience 41(2): 102 (2000)
For genera, Brunneosporella, Fluminicola, Papulosa, and Wongia are accepted in Papulosaceae (Wijayawardene et al. 2022).
Wongia Khemmuk, Geering & R.G. Shivas, IMA Fungus 7(2): 249 (2016)
Wongia was introduced by Khemmuk et al. (2016). These species differ in having non-amyloid apical rings in the asci with 3-septate ascospores that have dark brown middle cells and pale brown to subhyaline shorter distal cells (Khemmuk et al. 2016). Currently, seven species are in the genus (Fig. 78).
[See PDF for image]
Fig. 78
Phylogram generated from maximum likelihood analysis (RAxML) of the Papulosaceae based on the combined ITS, LSU and tef1-α sequence data. Nineteen taxa are included in the combined analyses, which comprise 2,256 characters with gaps. Pseudostanjehughesia aquitropica (MFLUCC 16-0569) and P. lignicola (MFLUCC 15-0352) were used as the outgroup. The best scoring RAxML tree with a final likelihood value of − 7766.776319 is presented. The matrix had 567 distinct alignment patterns, with 20.20% of undetermined characters or gaps. The proportion of invariable sites was 0.348999. Estimated base frequencies were as follows: A = 0.232003, C = 0.263011, G = 0.290595, T = 0.214390; substitution rates: AC = 1.028239, AG = 2.023933, AT = 1.464845, CG = 0.895737, CT = 6.942934, GT = 1.00000; gamma distribution shape parameter α = 0.424774. Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.95 are given near the nodes. T = ex-type strain. The newly generated sequences are indicated in blue bold
Wongia bandungensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov.
Index Fungorum number: IF900201; Facesoffungi number: FoF 15104; Fig. 79
[See PDF for image]
Fig. 79
Wongia bandungensis (BBH 49604, holotype). a Colonies on natural substrate. b–c Upper and reverse views of culture on PDA after 14 days at 25 °C. d–g Conidiophores, conidiogenous cells and conidia. h–k Conidia. Scale bars: a = 100 μm, d, e = 20 μm, f–k = 10 μm
Etymology: The specific epithet “bandungensis” refers to the Ban Dung District, Udon Thani Province, Thailand where the species originated.
Holotype: BBH 49604
Saprobic on submerged twigs in freshwater habitat. Sexual morph: Not observed. Asexual morph:Colonies on natural substratum effuse, scattered, hairy, brown. Mycelium 2–2.5 μm diam, mostly immersed, composed of branched, septate, smooth-walled, pale brown to brown hyphae. Conidiophores 77.5–153.3 × 2.7–4.7 μm (x̄ = 118.3 × 3.6 μm, n = 40), macronematous, mononematous, solitary, simple, cylindrical, erect, flexuous, 5–9-septate, smooth-walled, dark brown, paler towards the apex. Conidiogenous cells 32.2–114.4 × 2.6–3.6 μm (x̄ = 71.3 × 3.1 μm, n = 25), holoblastic, polyblastic, sympodial, integrated, terminal, denticulate, subhyaline, with pale brown scar. Conidia 11.3–14 × 4.3–5.2 μm (x̄ = 12.6 × 4.8 μm, n = 30), acropleurogenous, solitary, ellipsoidal to oblong-ellipsoidal, tapering and rounded at both ends, 1–2-septate, smooth-walled, hyaline to pale brown. Conidial secession schizolytic.
Culture characteristics: Colonies on PDA reaching 38–39 mm diam after 14 days at 25 °C, velvety, brown with greyish white margins, round, margins entire, soluble pigment absent, exudates absent, reverse dark brown with brownish orange margins.
Material examined: Thailand, Udon Thani Province, Ban Dung District, on submerged twigs of an unidentified plant in a freshwater river, 7 July 2022, S. Phookongchai, isolate FF01118 (BBH 49604, holotype), ex-type, TBRC-BCC 95171; isolate FF01118.01 (BBH 49605, isotype), living culture, TBRC-BCC 95343.
GenBank numbers: TBRC-BCC 95171: ITS = OQ121929, LSU = OQ121947, rpb2 = OQ116752, SSU: OQ121938, tef1-α = OQ116761; TBRC-BCC 95343: ITS = OQ121930, LSU = OQ121948, rpb2 = OQ116753, SSU: OQ121939, tef1-α = OQ116762.
Notes: Phylogenetic analyses based on the combined ITS, LSU and tef1-α sequence showed that our new species should be classified within Wongia (Fig. 78). Phylogenetically, W. bandungensis is sister to W. suae with strong statical support. However, they can be distinguished mainly by the conidial morphology. The conidia of W. suae are 2–3-septate, clavate to fusiform with rounded at apex and acute at base and smaller (8–11 × 3–4 μm; Zhang et al. 2023a, b, c, d), while conidia of W. bandungensis are 1–2-septate, ellipsoidal to oblong-ellipsoidal with non-acute base and larger. In culture, W. bandungensis strains exhibited in faster colony growth rate on PDA (38–39 mm diam after 14 days vs. 31 mm diam after 21 days in W. suae; Zhang et al. 2023b). Wongia bandungensis is morphologically similar to W. bambusae in its conidial pigmentation, which is hyaline or colourless. However, W. bambusae differs from W. bandungensis by having longer conidiophores (50–90 × 2.5–4.0 μm, x̄ = 75 × 3.3 μm) with hyaline, more septate conidia (0–3 septa; YU et al. 2024). Wongia bandungensis is newly introduced based on morphological data and DNA sequence analyses.
[See PDF for image]
Fig. 80
Maximum likelihood (ML) tree of the Colletotrichum gloeosporioides species complex inferred on IQ-TREE from a concatenated alignment of apn2, apn2/mat-igs, cal, gapdh, gap2-igs, gs and tub2. Significant supports for ML (SH-alrt bootstrap ≥ 80) are shown above the nodes. Thickened branches indicate significant support from Bayesian inference analysis (posterior probability ≥ 0.95). The tree was rooted at the midpoint. Ex-type isolates are highlighted in bold. New isolates are highlighted in blue bold. The scale bar indicates the average number of substitutions per site
Subclass Hypocreomycetidae O.E. Erikss. & Winka
Glomerellales Chadef. ex Réblová, W. Gams & Seifert, Stud. Mycol. 68: 170 (2011)
For taxonomic treatments, we follow Hyde et al. (2020a, b, c) and Wijayawardene et al. (2022).
Glomerellaceae Locq. ex Seifert & W. Gams in Zhang, et al., Mycologia 98(6): 1083 (2007) [2006].
This monotypic family is characterised by Colletotrichum (Maharachchikumbura et al. 2016; Jayawardena et al. 2022). They are pathogenic, endophytic and saprobic on a wide host range (Maharachchikumbura et al. 2016).
Colletotrichum Corda, in Sturm, Deutschl. Fl., 3 Abt. (Pilze Deutschl.) 3(12): 41 (1831)
Colletotrichum is recognized as a highly significant group of phytopathogenic fungi with global importance in plant diseases (Talhinhas and Baroncelli 2023). Important plant diseases include Chilli (Than et al. 2003), Citrus (Huang et al. 2013), and medicinal and ornamental plants (Zhang et al. 2023a, b). Species identification using molecular data is well-established (Bhunjun et al. 2021; Jayawardena et al. 2021a; Chen et al. 2022a, b). Colletotrichum species are diverse and classified into 13 species complexes with over 300 identified species (Armand et al. 2023; Peng et al. 2023; Zhang et al. 2023a). In the present study, C. squamosae is introduced as a novel taxon in the C. gloeosporioides complex (Fig. 80), and two new host records belonging to the C. orchidearum and C. magnum species complexes are provided (Fig. 82).
Colletotrichum squamosae W.A.S. Vieira, V. Doyle and M.P.S. Câmara, sp. nov.
Index Fungorum number: IF 902395; Facesoffungi number: FoF 1606; Fig. 81
Etymology: The specific epithet refers to the host Annona squamosa from where the ex-type was isolated.
Holotype: VIC49401
[See PDF for image]
Fig. 81
Colletotrichum squamosae (COAD3520, holotype). a–b. Cultures on PDA, 7 d growth, from above (a) and below (b). c–d. Conidiophores and conidiogenous cells from PDA (c) and CMA (d). e–f. Conida from PDA (e) and CMA (f). g–j. Appressoria. Scale bars: C–F = 10 µm; G–J = 5 µm
Associated with foliar anthracnose of Annona squamosaSexual morph: not observed. Asexual morph: Conidiogenous cells hyaline, smooth-walled, cylindrical to ampulliform, often extending to form a new conidiogenous locus. Conidiophores 7.5–20.7 μm (av. 13.2 ± 3.7 μm) length on PDA, 8.4–35.5 μm (av. 16.8 ± 6.7 μm) on CMA, hyaline, smooth-walled to verruculose, aseptate, unbranched, Conidia on PDA 9.9–11.7 × 2.9–3.9 μm (av. 10.5 ± 0.6 × 3.4 ± 0.2 μm), length/width ratio 2.6–3.7 (av. 3.1 ± 0.3), CMA 9.1–12.9 × 2.6–3.8 μm (av. 10.6 ± 0.8 × 3.2 ± 0.3 μm), length/width ratio 2.7–4.5 (av. 5.1 ± 0.1), hyaline, one-celled, smooth-walled, cylindrical with rounded ends, contents appearing granular. Appressoria in slide cultures 3.9–7.5 × 3.1–5.4 μm (av. 6.1 ± 0.9 × 4.4 ± 0.5 μm), length/width ratio 0.8–2.4 (av. 1.4 ± 0.3), single, medium to dark brown, smooth-walled, clavate, irregularly shaped or rarely ovoid, solitary or in small groups. Sclerotia, chlamydospores, and setae not observed.
[See PDF for image]
Fig. 82
Phylogenetic tree of Colletotrichum orchidearum and C. magnum species complexes generated by maximum parsimony analysis of combined ITS, act, chs-1, gapdh and tub2 sequence data. The tree was rooted with C. dracaenophilum (CBS 118199) and C. tongrenense (GMBC 0209). Maximum likelihood and Maximum parsimony bootstrap support values ≥ 50% (BT) as well as Bayesian posterior probabilities ≥ 0.90 (PP) are shown respectively near the nodes. Type strains are in bold, and the newly generated isolates are in red
[See PDF for image]
Fig. 83
Colletotrichum brevisporum (MFLU 23-0444), a, b. Host leaves with symptoms. c. Acervulus and conidial mass on PDA. d, e. Conidiogenous cells. f. Conidia. g–i. Appressoria. j. Upper and reverse view of culture on PDA. Scale bars: d,e = 10 µm, f = 20 µm, g–i = 10 µm
[See PDF for image]
Fig. 84
Colletotrichum plurivorum (MFLU 23–0443), a, b. The host leaves with symptoms. c. Acervuli on the host. d. Setae. e–g. Conidiogenous cells and conidial attachment. h–j. Conidia. k. Upper and reverse view of culture on PDA. Scale bars: d = 50 µm, e–g = 20 µm, h–j = 10 µm
Culture Characters: Colonies on PDA were saffron with chestnut centre in the obverse view and salmon with a brown vinaceous centre in the reverse, aerial mycelium sparse, growth rate at 25 °C 5–5.3 mm day−1 (average 5.1 ± 0.1 mm day−1).
Material examined: Brazil, SÃO PAULO: Itapetininga, from anthracnose lesions on Annona squamosa leaves, Jun 2017, W.A.S. Vieira. VIC49401(holotype, preserved in a metabolically inactive state). Ex-type LM04, COAD3520.
Known hosts: Annona squamosa (present study), Anacardim occidentale, Carica spp. and Licania tomentosa (Lisboa et al. 2018; Veloso et al. 2018; Vieira et al. 2022).
GenBank numbers: apn2/mat-igs: OP946694, apn2: OP946695, cal: OP946696; gap2-igs: OP946697; gapdh: OP946698; GS: OP946699; tub2: OP946700.
Notes: The multilocus analyses reveal a major clade containing the Colletotrichum isolate from Annona squamosa and isolates previously identified as C. australianum and C. queenslandicum (Fig. 80). The first clade represents the species C. australianum, which includes its ex-type (UMC002) and isolates previously identified as C. queenslandicum from various hosts in Australia and is supported by 97% ML bootstrap and 0.96 BYPP support. The second clade, with maximum support in both multilocus analyses, is comprised of the ex-type of C. queenslandicum (ICMP1778) and isolate CPC17123. Finally, the third clade contains isolate LM04 from A. squamosa and isolates associated with different hosts from Brazil which were previously identified as C. queenslandicum by Veloso et al. (2018) and Lisboa et al. (2018). This strongly supported clade (ML-BS 97%; BPP 1.0), introduced as the new species Colletotrichum squamosae, did not contain an ex-type of any previously described species. APN2/MAT-IGS is the best marker to discriminate between these three species. The cal sequence places C. australianum and C. queenslandicum in a polytomous clade but separates C. squamosae from those species. Colletotrichum squamosae isolates were split among two distantly related clades in GAPDH and GS trees (polyphyletic), but these clades did not contain any isolate of C. australianum or C. queenslandicum. This issue was previously reported in GS topology for several Colletotrichcum species by Silva et al. (2012) and is first reported here for GAPDH. Although those markers are among the most informative in identifying species within the C. gloeosporioides species complex (Liu et al. 2015a, b; Vieira et al. 2017, 2020), they negatively affect species delimitation when concordance criteria are employed. The tub2 has no ability to separate C. australianum, C. queenslandicum and C. squamosae. It was not possible to evaluate APN2 and GAP2-IGS due to the absence of sequences for C. australianum and C. queenslandicum. All isolates from Australia (Weir et al. 2012; Wang et al. 2021) were distributed among C. queenslandicum and C. australianum, while isolates from Brazil (Lisboa et al. 2018; Veloso et al. 2018; Vieira et al. 2022) were re-assigned to C. squamosae. This result indicates that these species may have a distinct geographical distribution. Morphological features show that C. squamosae conida on PDA are smaller than C. australianum (14.1–14.5 × 4.5–4.7) and C. queenslandicum (14.5–16.5 × 4.5–5). Colletotrichum squamosae appressorial shape is mostly irregular, while C. australianum and C. queenslandicum produce globose appressoria.
Colletotrichum brevisporum Noireung, Phouliv., L. Cai & K.D. Hyde, Cryptog. Mycol. 33(3): 350 (2012)
Index Fungorum number: IF 564156; Facesoffungi number: FoF 15072; Fig. 83
Associated with anthracnose leaf spot on Melothria sp. Sexual morph Not observed. Asexual morphConidiomata produced on 32 days old PDA culture, acervular, dark brown, bearing conidial mass. Setae not observed. Conidiophores not observed, Conidiogenous cells hyaline, smooth-walled, cylindrical or clavate, 10–12 × 3.5–5 μm (x̅ = 11 × 4.5 µm, n = 10). Conidia hyaline, aseptate, smooth-walled, cylindrical, straight, rounded at ends, guttulate, 15–20 × 5.5–6.5 μm (x̅ = 17.5 × 6 µm, n = 30). Appressoria in slide culture, brown to dark brown, variable in shape ovoid or slightly irregular, undulate to slightly lobate, 8–13 × 7–9 μm, formed by hyphae, becoming complex with age.
Culture characteristics: Colonies on PDA 35 mm after 7 days, cottony, circular, entire in margin; aerial hyphae medium sparse, white, reverse whitish buff. Colonies on OA 35–40 mm after 7 days, flat, entire in margin, surface white, reverse same color.
Material examined: Thailand, Chiang Rai Province, Phan District, Doi Pui Sai Khao Sub-district, on Melothria sp. leaf, 27 September 2021, Alireza Armand, T41 (MFLU 23-0444; MFLUCC 23-0285).
GenBank numbers: ITS: OR793879; act: OR804450; chs-1: OR804452; gapdh: OR804454; tub: OR804456
Notes: In the phylogenetic analysis, C. brevisporum strains from this study, clustered with the type strain of C. brevisporum (CBS129958) with 95% ML, 91% MP and 1.0 BYPP values (Fig. 82). Base pair differences between C. brevisporum (MFLUCC 23-0285) and C. brevisporum (BCC 38876) revealed no differences in ITS, chs-1 and tub2, 0.4% (1/258 bp) in act, 0.8% (2/238) differences in gapdh. Morphologically, C. brevisporum (MFLU 23-0444) is similar to C. brevisporum (BCC 38876, holotype) described by Noireung et al. (2012). However, our strain produced bigger conidia (15–20 × 5.5–6.5 μm in C. brevisporum (MFLU 23–0444) vs. 12–17 × 5–6 μm in the type strain) (Noireung et al. 2012). This is a new host record for the occurrence of C. brevisporum on Melothria.
Colletotrichum plurivorum Damm, Alizadeh & Toy. Sato, in Damm, Sato, Alizadeh, Groenewald & Crous, Stud. Mycol. 92: 31 (2018)
Index Fungorum number: IF 824228; Facesoffungi number: FoF 10691; Fig. 84
Associated with anthracnose on leaf spots of Aglonema sp. Sexual morph: Not observed. Asexual morph:Conidiomata produced on leaves, acervular, dark brown, bearing conidial mass and setae. Setae brown to dark brown, verruculose, 2–4-septate, 95–145 μm long (x̅ = 127.5 µm, n = 10), base sylindrical, slightly serrate, 4.5–6.5 μm diam (x̅ = 5 µm, n = 10), tip obtuse to acute. Conidiophores rarely observed, pale brown, septate, simple or branched, 27–42 μm diam. Conidiogenous cells hyaline, cylindrical to clacate, 15–21 × 3.5–4.5 μm (x̅ = 18 × 4 µm, n = 20). Conidia hyaline, aseptate, smooth-walled, cylindrical, straight or slightly narrow at the base, rounded at ends, guttulate, 17–23 × 5–6 μm (x̅ = 21.5 × 5.7 µm, n = 30).
Culture characteristics: Colonies on PDA 70–90 mm after 7 days, cottony, circular, entire in margin; aerial hyphae medium in dense, greenish olivaceous in center and white in margin, reverse black-green-olivaceous in margin. Colonies on OA 67–73 mm after 7 days, flat, entire in margin, surface white to pale olivaceous grey, reverse olivaceous grey.
[See PDF for image]
Fig. 85
Clonostachys rogersoniana (MFLU 23–0203) a Host. b, c14 days old colony on PDA (b. above, c. below). d Sporulation on PDA. e–g Conidiophores with phialides and conidial attachments. h Hyphal coils. Hyphal coils. i Conidia. Scale bars: a = 2.5 cm, b,c = 40 mm, d = 200 μm, e,f = 45 μm, g = 30 μm, h = 25 μm, i = 5 μm
Material examined: Thailand, Chiang Rai Province, Mueang Chiang Rai District, Tha Sut Sub-district, on leaf spots of Aglonema sp., 03 January 2022, Alireza Armand, A16 (MFLU 23-0443); (MFLUCC 23-0284).
GenBank numbers: ITS: OR793878; act: OR804449; chs-1: OR804451; gapdh: OR804453; tub: OR804455
Notes: Strains of C. plurivorum clustered together with our strain in a distinct clade with a high bootstrap value (Fig. 82). Base pair differences between C. plurivorum (MFLUCC 23-0284) and C. plurivorum (CBS 125474) revealed 0.2% (1/538 bp) in ITS, 0.7% (2/258 bp) in act, 0.3% (1/273 bp) in chs-1, 0.7% (4/533) in tub2 and no differences in gapdh. Morphologically, C. plurivorum (MFLU 23-0443) is similar to C. plurivorum, described by Damm et al. (2019), in the shape of conidia and conidiophores. However, our strain produced longer conidiophores (27–42 μm diam., in C. plurivorum (MFLU 23-0443) vs. 30 μm in the type strain), longer conidiogenous cells 15–21 × 3.5–4.5 μm in C. plurivorum (MFLU 23-0443) vs. 7–19 × 4–5.5 μm in type strain) and longer conidia (17–23 × 5–6 μm in C. plurivorum (MFLU 23–0443) vs (15–)16.5–20 (–22.5) × (5–) 5.5–6.5(–8) μm in type strain) (Damm et al. 2018). This is a new host record for the occurrence of C. plurivorum on Aglonema.
Hypocreales Lindau
For taxonomic treatments, we follow Hyde et al. (2020a, b, c) Wijayawardene et al. (2022) and Perera et al. (2023)
Bionectriaceae Samuels & Rossman, Stud. Mycol. 42: 15 (1999)
Rossman et al. (1999) established Bionectriaceae with 26 genera. In a recent update, Perera et al. (2023) accepted 41 genera.
Clonostachys Corda, Pracht-Fl. Eur. Schimmelbild.: 31 (1839)
Clonostachys was introduced by Corda (1839). According to Schroers (2001), sexual morphs with globose to subglobose perithecial ascomata and narrowly clavate asci with ellipsoidal ascospores are characteristics of their species. The conidiophores are either penicillate or verticillate, with phialides grouped in whirls that result in conidia that are usually hyaline but are occasionally greenish to olivaceous or green (Schroers 2001).
Clonostachys rogersoniana Schroers, Studies in Mycology 46: 109 (2001)
Index Fungorum number: IF485131; Facesoffungi number: FoF 07755; (Fig. 85)
Fungicolous on a bolete mushroom. Sexual morph Not observed. Asexual morph: on the host, Mycelium white cottony mass. On PDA, hyphomycetous, Hyphae 1–3.5 μm (x̅ = 2.5 μm, n = 30) wide, branched, septate, hyaline, hyphal coils observed. Conidiophores scattered on the agar media or arising from the aerial hyphae, mostly dominant towards the margin, hyaline, penicillate, septate, stipes 60–170 × 2–6 μm (x̅ = 117 × 4 μm, n = 20). Penicilli 55–78 × 50–96 μm (x̅ = 66.5 × 73 μm, n = 20), solitary to gregarious, bi- to quarter verticillate, non sporodochial; branches of the penicillus divergent, each branch terminating in metulae and adpressed phialides. Phialides 10–26 × 1.7–3.6 μm (x̅ = 16.5 × 2.6 μm, n = 30), in whorls of 2–4, narrowly flask-shaped, slightly tapering toward the apex. Intercalary phialides not observed. Conidial masses white to light yellow. Conidia 5–9 × 1.5–4 μm (x̅ = 6.5 × 2.5 μm, n = 40, broadly ellipsoidal to oval, rarely minutely curved, ends broadly rounded, hilum laterally displaced, almost median or invisible, hyaline, smooth-walled.
Culture characteristics: Colonies on PDA attaining 60–70 mm diameter after 14 days at 25 °C, velvety to lanose, dense, aerial mycelium with hyphae arranged in strands, colony front initially white and develops brown spots when getting older, colony reverse white to medium brown with irregular margins.
Material examined: Thailand, Chiang Mai province: Mae Tang, Pa Pae, Ban Pha Deng, isolated from the white mycelium growing on a bolete mushroom, 07 July 2021, AJ. Gajanayake, MR 01 (inactive dry culture, MFLU 23-0484, new host record), living culture, MFLUCC 23-0203.
GenBank numbers: ITS = OR473265, tub2 = OR804262.
Notes: Our isolate (MFLUCC 23-0203), clusters within Clonostachys rogersoniana with bootstrap support of 100% ML, 1.00 BYPP (Fig 86). Furthermore, the morphology of our isolate (MFLUCC 23-0203) resembles the original description and illustrations of C. rogersoniana by Schroers (2001). However, when our strain is compared with the strain described by Schroers (2001), there are slight dimensional differences (conidiophores: 60–170 × 2–6 μm vs 60–200 μm × 3–5 μm, height of the penicilli: 55–78 μm vs. 50–100 μm, length of the phialides: 10–26 μm vs. 11.8–26.8 μm, conidia: 5–9 × 1.5–4 μm vs. 5.8–7.2 × 3–3.8 μm) in morphological structures. The reason may be the differences in the media in which the colonies were grown.
[See PDF for image]
Fig. 86
Phylogram generated from maximum likelihood analysis based on combined ITS and tub2 sequence data representing the Clonostachys species. Related reference sequences were downloaded from GenBank based on previously published data from Perera et al. (2020) and Zeng and Zhuang (2022). Seventy-six taxa are included in the combined analyses. Fusarium acutatum (CBS 402.97) and Nectria cinnabarina (CBS 18987) are used as the outgroup taxa. The best scoring RAxML tree with a final likelihood value of − 13,172.451 is presented. The matrix had 639 distinct alignment patterns, with 19.91% of undetermined characters or gaps. The proportion of invariable sites was 0.316. Estimated base frequencies were as follows: A = 0.250, C = 0.250, G = 0.250, T = 0.250; substitution rates: AC = 1.00000, AG = 2.80366, AT = 1.00000, CG = 1.00000, CT = 4.24917, GT = 1.00000; gamma distribution shape parameter α = 0.692. The value of the average standard deviation of the split frequencies of BYPP analysis was 0.008389. Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.90 are given near the nodes. The newly generated sequence is in blue. The type strains are indicated in black bold
Clonostachys rogersoniana has been commonly isolated from soil whereas it has also been isolated from bark of dead twigs, leaf litter and decaying seed pods (Schroers 2001; Perera et al. 2020). This species has shown a cosmopolitan distribution while it has been mostly reported from warmer or tropical regions (Schroers 2001). Therefore, most probably C. rogersoniana spores from the soil on the bolete mushroom would eventually lead to growth as a fungicolous fungus. Previously, Clonostachys byssicola, C. subquaternata and C. rosea have been reported as fungicolous species (Sun et al. 2019). However, C. rogersoniana has been extensively studied for its potential to produce secondary metabolites whereby, a cordyceps-colonizing strain of C. rogersoniana was among them (Wang et al. 2017a, b; Wang et al. 2019; Han et al. 2020; Tapfuma et al. 2023). To the best of our knowledge, this is the first report of Clonostachys rogersoniana as a fungicolous fungus on a bolete mushroom, from Thailand.
Nectriaceae Tul. & C. Tul. [as ‘Nectriei’], Select. fung. carpol. (Paris) 3: 3 (1865)
Nectriaceae species are characterised by uniloculate ascomata that are white, yellow, orange-red or purple. Ascomata of Nectriaceae species change colour in KOH and are not immersed in a well-developed stroma (Rossman et al. 1999; Rossman 2000). Nectriaceae includes 70 genera (Wijayawardene et al. 2022).
Calonectria De Not., Comm. Soc. crittog. Ital. 2(fasc. 3): 477 (1867)
Calonectria was introduced by De Not (1867) and typified by C. daldiniana which was later changed to C. pyrochroa (Rossman 1979). Calonectria species are distinguished by the production of multi-septate, cylindrical conidia, a vesicle with a long stipe derived from conidiophores, and a crimson perithecium (Crous et al. 1994). Species of Calonectria are important plant pathogens on a wide range of hosts (Zhang et al. 2022) (Fig. 87).
[See PDF for image]
Fig. 87
Bayesian phylogenetic tree of Calonectria. The tree was built using concatenated sequences of the genes cal, his, tef-1α and tub2. Bayesian posterior probability values ≥ 0.90 are indicated above the nodes. The sequences generated in this study are highlighted in bold. This tree is rooted with Calonectria pteridis
Calonectria potisiana Melo & R.F. Alfenas, sp. nov.
Index Fungorum number: 901065; Facesoffungi number: 14769; Fig. 88
[See PDF for image]
Fig. 88
Calonectria potisiana (COUFPI 186, ex-type). a Leaf spots on Strunthus sp. b Frontal view of the culture. c reverse view of the culture. d Protoperithecia. e cirri (red arrow). f Perithecia. g Asci with ascospores. h Ascospores. i Macroconidia. j Branched phialides. h Stipe with the presence of vesicula. Scale Bars: d, e,f: 2 µm; g, h, i: 10 µm. j, k: 20 µm
Etymology: “Potis” reference to the name of an extinct indigenous tribe that inhabited the banks of the Poti River, Teresina, Brazil, the geographic region where the fungus was collected.
Holotype: VIC 4741
Associated with Strutanthus sp., necrotic leaf spots. Sexual morphProtoperithecia 48–45 µm, globose to oval, superficial distribution in the culture medium, orange. Fertile perithecia 160–420 µm, pear-shaped, orange-red, turning red-brown with time. Asci 45–98 × 8–22 µm (x̅ = 38–88 × 8–18 μm, n = 20), 8 ascospores, clavate. Ascospores 39–42 × 5–7 µm (x̅ = 38–41 × 4–7 μm: n = 20), hyaline, guttulate, fusoides, with 1–3-septate, with constriction in the septum, Asexual morph: Produced on 18 days old MEA culture, Conidiophores 65–112 × 3–4 µm, consist of a stipe with a penicillate arrangement, 1-septate, hyaline. Phialides 22–43 × 2–4 µm (x̅ = 20–38 × 1,8–3,8 μm n = 10), branched hyaline, 3–4 per branch. Macroconidia 45–52 × 3–5 µm (x̅ = 42–48 × 2,6–4,8 µm. n = 20), cylindrical, hyaline, rounded at both ends, 3- septate. Stipe 85–132 × 2–4 µm (x̅: 82–128 × 1,6–3,5. n:10), hyaline, septate, terminal naviculata vesicles, Chlamydospores 25–38 µm globose, solitary or in short chains with thick walls,
Culture characteristics: Fast growing on MEA medium, growth rate 32 mm/day, mycelium white, brown at the centre, aerial, cottony to flaky, reverse umber. Abundant production of spores on aerial mycelium, especially in the centre of the culture.
Material examined: Brazil, State of Piauí, Teresina, Campus of the Federal University of Piauí, Center for Agricultural Sciences, Agricultural Sciences, on necrotic leaf spots of Struanthus sp., (Loranthaceae), April 20, 2018, M.P. Melo, original code (VIC 4741 holotype), ex-type, COUFPI 186.
GenBank numbers: COUFPI 186: tef1-α = OR500380, cal = OR500371, his = OR500374, tub2 = OR500377, COUFPI 196: tef1-α = OR500381, cal = OR500372, his = OR500375, tub2 = OR500378, COUFPI 299: tef1-α = OR500379, cal = OR500370, his = OR500373, tub2 = OR500376.
Notes Calonectria potisiana, is a fungus associated with necrotic leaf spots in Struanthus sp., with spores in the center of the lesions. In the phylogenetic analysis, using the tef1-α, tub2, his, and cmdA sequences, C. potisiana is placed in the C. naviculata complex, with 100% ML bootstrap and 1.00 BYPP supports, having C. multiphialidica as a sister clade, (Fig. 87). When cultivated in carrot agar culture medium, C. potisiana produces a high number of perithecia, resulting in fertile ascospores. Calonectria potisiana can be differentiated from the species of the C. naviculata complex by its 3-septate macroconidia and producing perithecia and fertile ascospores. Due to the number of morphological markers for accurate differentiation of the species in the C. naviculata complex, multigene phylogenetic analysis of the tef1-α, tub2, his, and cmdA genomic regions is necessary for correct species delimitation (Liu et al. 2020). Most species of the C. naviculata complex are saprophytic, living on soil and litter (Alfenas et al. 2015). Recently, new species of Calonectria have been described in Brazil (Sanchez-Gonzalez et al. 2022). Here we introduce C. potisiana as a new species associated with leaf spots on Strunthus sp., showing that tropical regions have a high diversity of Calonectria species.
Ophiocordycipitaceae G.H. Sung, J.M. Sung, Hywel-Jones & Spatafora, in Sung et al., Stud. Mycol. 57: 35 (2007)
Notes: Four families, Clavicipitaceae, Cordycipitaceae, Ophiocordycipitaceae, and Polycephalomycetaceae, comprise the majority of the species in Cordyceps sensu lato (Wei et al. 2021a, b). Many Cordyceps sensu lato species, including Beauveria bassiana, Cordyceps militaris, C. cicadae, Ophiocordyceps sinensis, Papiliomyces shibinensis, and Keithomyces neogunnii, are used in traditional Chinese medicine (Wen et al. 2016, 2017; Zha et al. 2018, 2019). Ophiocordycipitaceae contains eight genera with dark stromate, superficial to immersed perithecia, and cylindrical asci (Sung et al. 2007; Ban et al. 2015; Xiao et al. 2019, 2023; Wijayawardene et al. 2022).
Ophiocordyceps Petch, Trans. Br. Mycol. Soc. 16(1): 73 (1931)
Notes: Petch (1931) established Ophiocordyceps to accommodate species with distinct asci and ascospores from Cordyceps and was historically classified in the Clavicipitaceae, with Ophiocordyceps blattae as the type species. Sung et al. (2007) divided Clavicipitaceae into Clavicipitaceae, Cordycipitaceae, and Ophiocordycipitaceae in the Hypocreales based on molecular data and revised Ophiocordyceps as the type genus in Ophiocordycipitaceae. The hosts of species in Ophiocordyceps are in 15 orders of insects, which are Araneida, Blattaria, Coleoptera, Dermaptera, Diptera, Hemiptera, Homoptera, Hymenoptera, Isoptera, Lepidoptera, Mantodea, Odonata, Orthoptera, Phasmatodea, and Tylenchida (Araújo & Hughes 2016; Kepler et al. 2014; Boucias et al. 2007; Simmons et al. 2015). Hitherto, Ophiocordyceps has included 335 species epithets in Index Fungorum (June 2024) with worldwide distribution (Sung et al. 2007a; Simmons et al. 2015; Spatafora et al. 2015; Xiao et al. 2019) (Fig. 89).
[See PDF for image]
Fig. 89
Phylogram generated from maximum likelihood analysis based on combined LSU, tef1-α, and rpb1 sequence data. The maximum likelihood (ML) analysis was performed in RAxML-HPC BlackBox (8.2.12), and Bayesian inference (BI) analysis was performed the MrBayes on XSEDE (3.2.7a) (Miller et al. 2010). Bootstrap support values for maximum likelihood, greater than 75% and Posterior Probabilities from Bayesian Inference ≥ 0.80 are given above branches. Sixty-nine strains are included in the combined gene analyses, which comprise 2450 characters after alignment (836 characters for LSU, 918 characters for tef1-α, and 696 characters for rpb1). Two specimens of Cordyceps militaris (OSC 93623 and YFCC 6587) are used as the outgroup taxa. The tree topology derived from the Bayesian analysis was similar to the maximum likelihood analysis. The best RAxML tree with a final likelihood value of − 25,546.230423 is presented. The matrix had 1262 distinct alignment patterns, with 16.37% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.226310, C = 0.299137, G = 0.286722, T = 0.187831; substitution rates AC = 1.097283, AG = 3.441517, AT = 1.035933, CG = 1.353883, CT = 7.600215, GT = 1.000000; gamma distribution shape parameter α = 0.588153. The newly generated sequences are indicated in blue bold
Ophiocordyceps duyunensis X. C. Peng & T. C. Wen sp. nov.
Index Fungorum number: IF900184; Facesoffungi number: FoF 13974; Fig. 90
[See PDF for image]
Fig. 90
Ophiocordyceps duyunensis (a–b HKAS 125850, holotype. c HKAS 125850, paratype. d HKAS 125843, paratype) a Habitat.b–d Synnemata emerging from the infected insect. e–g Overview of the host. h–i Philidies. j–k Philidie with conidia. l conidia. m–n Lower side and upper side of the culture, o Synnemata growing on PDA medium, p–q Phialides on culture, r Philidies with conidia. s Conidia on culture. Scale Bars: b, m, n = 10 mm, h, j, l, p–r = 10 µm, i, k, s = 5 µm
Etymology: Referring to the collecting site “Duyun District”.
Holotype: HKAS 125850.
Parasitic on the cocoon or larva. Sexual morph: Not observed. Asexual morph:Primary synnemata 5–7.3 cm long, 0.7–1 mm wide, arising from the insect body, solitary, or several, stipitate, multiply branched, grey to brown, not smooth, cylindrical, or rhizoid, tapering gradually toward the apex. Secondary synnemata observed, without an enlarged globose fertile head on the top. Phialides hirsutella-like, scattered, smooth or rough, tapering to a neck, lageniform to conical, lower portions amygdaliform to ovoid, tapering abruptly, forming a long neck, 9–44 × 2.8–7.7 µm (x̅ = 19.6 × 4.7 µm, n = 60). Conidia fusiform to limoniform, rough, 8.7–15.5 × 2.7–6.5 μm (x̅ = 11.2 × 5 µm, n = 20). Mycelium rough or smooth, branched, hyaline, septate, 1.4–3.7 μm (x̅ = 2.5 μm, n = 20) diam.
Culture Characteristics: colonies on PDA, attaining a diameter of 27–34 mm within 22 d at 25 ℃, dense, leathery or villous, the center is brown, the edges are white, flat, reverse off-white to dark grey. Synnemate was produced in the later stage. Phialides scattered, hirsutella-like, hyaline, solitary, branched or unbranched, narrow slender, smooth, 11.5–41.5 × 1–6 µm (x̅ = 26 × 3 µm, n = 50). Conidia fusiform to limoniform, rough, 6–15 × 4–8.5 μm (x̅ = 11 × 5.5 µm, n = 80). Mycelium rough or smooth, branched, hyaline, hyaline, septate, 2.5–5.5 μm (x̅ = 3.8 μm, n = 30) diam.
Habitat and distribution: The new species grow on the ground of coniferous and broad leaf forests and is currently only found in China.
Material examined: China, Guizhou, Duyun City, Lvyinhu street, 26° 23′ 47′′ N, 107° 30′ 42′′ E, 946 m alt., parasitise on cocoon of Lepidoptera, 20 May 2021, X. C. Peng, (HKAS 125850, holotype, ex-type DY21052060, HKAS 125849, paratype). China, Guizhou, Duyun City, Yangliu street, 26° 23′ 46′′ N, 107° 30′ 43′′ E, 929 m alt., parasitise on pupa of Lepidoptera, 20 May 2021, X. C. Peng, (HKAS 125843, paratype, HKAS 125844, paratype).
GenBank numbers: HKAS 125850: ITS = OQ110574; SSU = OQ110579, LSU = OQ110572; tef1-α = OQ116922, rpb1 = OQ116925. HKAS 125843: LSU = OQ110570, tef1-α = OQ116920, rpb1 = OQ116923. HKAS 125849: LSU = OQ110571, tef1-α = OQ116921, rpb1 = OQ116924.
Notes: Three specimens were obtained from the same region in Guizhou, China. Ophiocordyceps duyunensis is phylogenetically distinct and is sister to a clade containing O. elongate (its anamorphic stage is Hirsutella gigantea; Simmons et al. 2015), H. kuankuoshuiensis, and O. alboperitheciata with strong support value (100% ML/1.00 PP) (Fig. 89). They are all parasitic on Lepidoptera, but at different metamorphosis stages. Ophiocordyceps elongate (≡ Cordyceps elongata) is on larvae and pupa (Petch 1937), H. kuankuoshuiensis and O. alboperitheciata are on larvae (Qu et al. 2021; Fan et al. 2021), and O. duyunensis is on cocoon and pupa. The morphology of O. duyunensis is different from O. elongate and H. kuankuoshuiensis since the former have secondary synnemata and bigger conidia (8.7–15.5 × 2.7–6.5 vs. 9–10 × 3–4 vs. 9.9–12.6 × 2.7–4.5) (Qu et al. 2021; Petch 1937), O. alboperitheciata is only known from its sexual morph (Qu et al. 2021) (Fig. 91).
[See PDF for image]
Fig. 91
A maximum likelihood tree based on a combined analysis of the 18S and 28S rDNA. The numbers at the nodes represent maximum likelihood bootstrap (ML), maximum parsimony (MP) bootstrap and posterior probability (BYPP). The ML analysis was performed in IQ-TREE web server (Trifinopoulos et al. 2016) with the following settings: Tamura 3-parameter model (Tamura 1992), gamma distributed with invariant sites (G + I), #rate categories set at 4. The ML branch support analysis was performed with ultrafast bootstrap (UFBoot) (Hoang et al. 2018) with the following settings: number of bootstrap alignments set at 1000, maximum iteration set at 1000, minimum correlation coefficient set at 0.99, 1000 replicates of SH-aLRT branch test. The MP bootstrap analysis was performed in MEGA11 (Tamura et al. 2021) with the following settings: Tree-Bisection-Regrafting algorithm with search level 1, initial trees through random addition of sequences (5 replicates), 1000 replicates (Felsenstein 1985). The PP was obtained through Bayesian analysis in BEAST v1.10.4 (Suchard et al. 2018) with the following settings: TN93, estimated base frequency, gamma + invariant sites, number of gamma categories set at 5, a strict clock, Coalescent: Constant Size as the speciation model, running 20 million generations with parameters and trees sampled every 1000 generations
Microascales Luttr. ex Benny & R.K. Benj., Mycotaxon 12(1): 40 (1980)
Microascales enclosed seven families; Ceratocystidaceae, Chadefaudiellaceae, Gondwanamycetaceae, Graphiaceae, Halosphaeriaceae, Microascaceae and Triadelphiaceae (Wijayawardene et al. 2022).
Halosphaeriaceae E. Müll. & Arx ex Kohlm., Can. J. Bot. 50: 1951 (1972)
Halosphaeriaceae comprises species that live in aquatic and semiaquatic marine environments (Jones et al. 2017). According to Jones et al. (2017), they comprise the majority of the marine Ascomycota group, with a small number of species also existing in freshwater and terrestrial habitats. Currently, 68 genera are accepted (Wijayawardene et al. 2022). Protein coding genes are not available for the majority of the species in the Halosphaeriaceae. Therefore, in the present study, phylogenetic analysis is based on 18S and 28S rDNA sequence data (Fig. 91).
Jinshana K.L. Pang, M.W.L. Chiang & E.B.G. Jones, gen. nov.
Index Fungorum Number: IF 902396; Facesoffungi number: FoF 14035
Etymology: In reference to the type collection location ‘Jin-Shan’.
Saprobic on trapped wood on a rocky shore. Sexual morph:Ascomata superficial, globose to subglobose, solitary to gregarious, coriaceous, yellow to brown, ostiolate, papillate. Necks short, cylindrical, white to pale yellow. Periphyses not observed. Peridium comprising an outer stratum (5–8 layers) of hyaline cells of textura angularis with large lumina and an inner stratum (3–4 layers) of hyaline and elongated cells. Catenophyses present. Asci 8-spored, unitunicate, thin-walled, clavate with a flattened apex, pedunculate. Ascospores ellipsoidal, 1-septate, with one large oil globule in each cell, not constricted at the septum, thick-walled, hyaline, appendaged. Appendages bipolar, initially adpressed to the ascospore wall, unravelling in sea water to form a long thin filament. Asexual morph: Not observed.
Type species: Jinshana tangtangiae K.L. Pang, M.W.L. Chiang & E.B.G. Jones, sp. nov.
Jinshana tangtangiae K.L. Pang, M.W.L. Chiang & E.B.G. Jones, sp. nov.
Index Fungorum number: IF 902397; Facesoffungi number: FoF 14036; Fig. 92
Holotype: F0036020
Etymology: To honour my Thai boxing coach Ya-Wen Tang (nick name ‘Tang Tang’), who has been helping me to get back to a healthy lifestyle.
[See PDF for image]
Fig. 92
Jinshana tangtangiae (holotype). a Parafilm section of ascoma showing asci developing at the base of venter. b Aperiphysate neck and peridium made of two strata of cells. c Light-coloured ascomata on wood. d Eight-spored asci with a flat apex. e Catenophyses made up of irregularly shaped cells. f–h Hyaline ascospores with bipolar, pad-like appendages which unfurl into a thread in seawater. i Long, thread-like appendage
Saprobic on trapped wood on a rocky shore. Sexual morph:Ascomata 181–245 μm high, 181–266 μm diam. (x̅ = 213 × 226 μm, n = 5), superficial, globose to subglobose, solitary to gregarious, coriaceous, yellow to brown, ostiolate, papillate (Fig. 92a, c). Necks short, 53–117 μm long, 32–64 μm diam. (x̅ = 77 × 48 μm, n = 4), cylindrical, yellow to brown (Fig. 92b). Periphyses not observed. Peridium 21–53 μm (x̅ = 34 μm, n = 12), comprising an outer stratum (5–8 layers) of hyaline cells of textura angularis with large lumina and an inner stratum (3–4 layers) of hyaline and elongated cells (Fig. 92a, b). Catenophyses present (Fig. 92e). Asci 105–142 × 19–28 μm (x̅ = 121 × 24 μm, n = 12), 8-spored, unitunicate, thin-walled, clavate with a flattened apex, pedunculate (Fig. 92d). Ascospores 21–28 × 10–11 μm (x̅ = 25 × μm, n = 10), ellipsoidal, 1-septate, with one large oil globule in each cell, not constricted at the septum, thick-walled, hyaline, appendaged (Fig. 92f–h). Appendages bipolar, initially adpressed to the ascospore wall, unravelling in sea water to form a long thin filament (Fig. 92i). Asexual morph: Not observed.
Material examined: Jin-Shan (New Taipei City), on a piece of unidentified trapped wood, 1 June 2022, S.Y. Guo and K.L. Pang, (F0036020, holotype), deposited at National Museum of Natural Science (Taipei) as dried wood.
GenBank accession no.: OQ418018 (18S rDNA), OQ418019 (28S rDNA), OQ469752 (ITS rDNA).
Notes: Jinshana tangtangiae did not grow in culture after multiple isolation attempts, and the partial 18S and 28S rDNA sequences were obtained from PCR amplifications of genomic DNA extracted from spores by boiling for 10 min. Jinshana tangtangiae did not group with the genera with unfurling ascospore appendages and grouped with Qarounispora grandiappendiculata with 100% ML, 92% MP bootstrap supports and 1.00 BYPP values (Fig. 91). Morphologically, J. tangtangiae is similar to Q. grandiappendiculata in the colour and shape of the ascomata but differs from it in having a clavate asci with a flattened apex, and the presence of bipolar ascospore appendages which unravel in seawater. Qarounispora grandiappendiculata has broad clavate asci without an apical apparatus and broadly ellipsoidal ascospores with a unipolar appendage which swells in water to form an irregular amorphous structure (Nourel-Din et al. 2022). Jinshana tangtangiae is most similar morphologically to Aniptodera aquadulcis, which has immersed/superficial ascomata, asci with an apical pore, plasmalemma retracted below the apex and bigger ascospores (Hsieh et al. 1995). Therefore, a new genus and a new species are described for this marine fungus.
New sequences of Tunicatispora australiensis grouped within Halosarpheia (H. fibrosa, H. japonica, H. trullifera, H. unicellularis) with a strong support (Fig. 91). Tunicatispora australiensis and Halosarpheia spp. have many morphological similarities, such as immersed/partly immersed, soft-textured, globose/subglobose ascomata, catenophysate, broadly ellipsoidal ascospores and the presence of bipolar ascospore appendages which unravel in seawater (Kohlmeyer and Kohlmeyer 1977; Hyde 1990). However, T. australiensis has two types of appendages, i.e., bipolar unfurling ascospore appendages and a skin-like sheath surrounding the ascospores (Hyde 1990; McKeown et al. 1996). The phylogenetic placement of T. australiensis in maximum parsimony was different from that in maximum likelihood, and other genes, such as protein genes, should be sequenced to resolve its relationship with Halosarpheia.
As shown in Fig. 91, the taxa with unfurling ascospore appendages included in the phylogenetic analysis did not form a monophyletic group, an observation also made in other studies (Campbell et al. 2003; Pang et al. 2003a, b). The unfurling type of ascospore appendages may be acquired through convergent evolution due to the evolutionary pressure in attaching to the scarce organic substrates in the marine environment for germination and subsequent growth (Jones 1994; Pang et al. 2003a). Therefore, this type of ascospore appendages is not a reliable taxonomic character for delineation of taxa in the Halosphaeriaceae. Ascospore shape, on the other hand, may be useful in separating different genera. For example, Oceanitis is distinct in having falcate ascospores (Kohlmeyer 1977; Dupont et al. 2009) while Halosarpheia (sensu stricto) is characterized by having broadly ellipsoidal ascospores (Kohlmeyer and Kohlmeyer 1977). Saagaromyces species are known to have large ascospores (Pang et al. 2003b). Pang et al. (2012) analysed shape parameters of ascospores of selected taxa of the Halosphaeriaceae with unfurling ascospore appendages (and related taxa), and the cluster analysis of the shape parameter grouped the three Saagaromyces spp. together (S. abonnis, S. ratnagiriensis and S. glitra), as well as Cucullosporella mangrovei with Aniptosporopsis lignatilis and Paraaniptodera longispora. Saagaromyces glitra is an unappendaged species and clustering of the three Saagaromyces species may suggest the importance of ascospore shape in the delineation of different species.
Subclass Savoryellomycetidae Hongsanan, K.D. Hyde & Maharachch.
Fuscosporellales Z.L. Luo, K.D. Hyde & H.Y. Su Jing Yang, Bhat & K.D. Hyde, in Yang, Maharachchikumbura, Bhat, Hyde, et al., Cryptog. Mycol. 37(4): 457 (2016)
Notes: Yang et al. (2016) introduced Fuscosporellales with four monotypic genera, viz. Fuscosporella, Mucispora, Parafuscosporella and Pseudoascotaiwania. For taxonomic treatment of this order, we follow Hyde et al. (2020a, b, c) and Wijayawardene et al. (2022).
Fuscosporellaceae Jing Yang, Bhat & K.D. Hyde, in Yang, et al., Cryptog. Mycol. 37(4): 457 (2016)
Notes: Fuscosporellaceae comprises six genera viz; Bactrodesmiastrum, Fuscosporella, Mucispora, Parafuscosporella, Plagiascoma and Pseudoascotaiwania (Wijayawardene et al. 2022).
Mucispora Jing Yang, Bhat & K.D. Hyde, in Yang, Maharachchikumbura, Bhat, Hyde, et al., Cryptog. Mycol. 37(4): 466 (2016)
Notes: Yang et al. (2016) introduced Mucispora as a monotypic genus for M. obscuriseptata. Mucispora species are macronematous, mononematous conidiophores, monoblastic, terminal and cylindrical conidiogenous cells and ellipsoidal or obovoid dark brown conidia with three septa (Fig. 93).
[See PDF for image]
Fig. 93
Phylogram generated from maximum likelihood analysis (RAxML) of the Fuscosporellaceae based on the combined ITS, LSU and rpb2 sequence data. Twenty-eight taxa are included in the combined analyses, which comprise 2515 characters with gaps. Canalisporium caribense (SS 03839) and Pleurothecium semifecundum (CBS 131271) were used as the outgroup. The best scoring RAxML tree with a final likelihood value of − 14,326.553200 is presented. The matrix had 1,199 distinct alignment patterns, with 34.92% of undetermined characters or gaps. The proportion of invariable sites was 0.224117. Estimated base frequencies were as follows: A = 0.225509, C = 0.298033, G = 0.295650, T = 0.180808; substitution rates: AC = 1.094752, AG = 2.429464, AT = 1.688220, CG = 0.843807, CT = 6.850439, GT = 1.00000; gamma distribution shape parameter α = 0.516382. Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.95 are given near the nodes. T = ex-type strain. The newly generated sequences are indicated in blue bold
Mucispora maesotensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov.
Index Fungorum number: IF 900199; Facesoffungi number: FoF 15105; Fig. 94
[See PDF for image]
Fig. 94
Mucispora maesotensis (BBH 49600, holotype). a, b Colonies on natural substrate. c–e Conidiophores with conidia. f Conidiophore. g–i Conidia. j–n Hyphae and conidiophores with conidia on PDA. o–r Conidia. Upper and reverse views of culture on PDA after 27 days at 25 °C. Scale bars: a, b = 100 μm, c–r = 20 μm
Etymology: The specific epithet “maesotensis” refers to the Mae Sot District, Tak Province, Thailand where is the type locality of this fungus.
Holotype: BBH 49600
Saprobic on submerged wood in freshwater habitat. Sexual morph: not observed. Asexual morph:Colonies on natural substratum sparse, scattered, hairy, black. Mycelium 2–2.5 μm diam, mostly immersed, composed of branched, septate, smooth-walled, pale brown to brown hyphae. Conidiophores 46–107.7 × 5–6.4 μm (x̄ = 73.3 × 5.6 μm, n = 15), macronematous, mononematous, solitary or in small groups on compactly aggregated cells, simple, cylindrical, erect, straight, 2–6-septate, smooth-walled, brown below, paler towards the apex, with up to 3 or more percurrent extensions. Conidiogenous cells holoblastic, monoblastic, integrated, terminal, cylindrical, pale brown. Conidia 24.5–37 × 14.2–22.5 μm (x̄ = 30.8 × 17.7 μm, n = 25), acrogenous, solitary, ellipsoidal to obovoid, apex rounded and base truncate, 3-septate, smooth-walled, dark brown to black, often paler at the basal cell, with unobservable septa at maturity. Conidial secession schizolytic.
Culture characteristics: Colonies on PDA reaching 12–15 mm diam., after 27 days at 25 °C, glabrous, white, round, margins mostly entire, soluble pigment absent, exudates absent, reverse white with dark grey at center. Vegetative hyphae on PDA medium 1.5–4 μm diam., composed of branched, septate, smooth-walled, subhyaline to pale brown. Conidiophores 16.1–111.5 × 4.6–6.2 μm (x̅ = 54.2 × 5.4 μm, n = 15), macronematous, mononematous, solitary, simple, cylindrical, erect, straight, 1–5-septate, smooth-walled, brown, paler towards the apex, with up to 5 or more percurrent extensions. Conidiogenous cells holoblastic, monoblastic, integrated, terminal, cylindrical, pale brown. Conidia 27.3–33.5 × 14.5–19 μm (x̅ = 29.6 × 17 μm, n = 30), acrogenous, solitary, ellipsoidal to obovoid, apex rounded and base truncate, 3-septate, smooth-walled, dark brown to black, basal cell paler. Conidial secession schizolytic.
Material examined: Thailand, Tak Province, Mae Sot District, Pha Daeng Waterfall Nature Trail, on submerged wood of an unidentified plant in a small freshwater stream, 24 May 2022, N. Boonyuen, isolate FF01097 (BBH 49600, holotype), ex-type, TBRC-BCC 95165; isolate FF01097.01 (BBH 49601, isotype), living culture, TBRC-BCC 95166.
GenBank numbers: TBRC-BCC 95165: ITS = OQ121935, LSU = OQ121953, rpb2 = OQ116758, SSU: OQ121944, tef1-α = OQ116766; TBRC-BCC 95166: ITS = OQ121936, LSU = OQ121954, rpb2 = OQ116759, SSU: OQ121945, tef1-α = OQ116767.
Notes: In the phylogenetic analysis of Mucispora, isolates from the present study, developed a particular phylogenetic relationship to M. obscuriseptata with 100% ML bootstrap and 1.00 BYPP values (Fig. 93). Our isolates share some morphological characteristics with M. obscuriseptata including mononematous, simple conidiophores, cylindrical conidiogenous cells with ellipsoidal to obovoid, dark brown, 3-septate conidia (Fig. 94). However, M. obscuriseptata differs from M. maesotensis because its conidia are predominantly surrounded by a hyaline mucilaginous sheath and have longer conidiophores (80–170 × 5–7.5 μm; Yang et al. 2016). In culture, these two species mainly differ in the size of conidiophore and conidial septation. Mucispora maesotensis has unbranched, smaller conidiophores and 3-septate conidia, while M. obscuriseptata has branched, longer (up to 367 × 4.5–9 μm) conidiophores and 2 − 3-septate with rarely uniseptate conidia (Yang et al. 2016). Thus, we propose Mucispora maesotensis as new to science.
Pleurotheciales Réblová & Seifert, in Réblová, Seifert, Fournier & Štěpánek, Persoonia 37: 63 (2015) [2016]
Pleurotheciaceae Réblová & Seifert, in Réblová, Seifert, Fournier & Štěpánek, Persoonia 37: 63 (2015) [2016]
Notes: Pleurotheciaceae is accepted in Pleurotheciales and currently, 14 genera are in the genus (Wijayawardene et al. 2022, Bao et al. (2022)).
Pleurothecium Höhn., Centbl. Bakt. ParasitKde, Abt. II 60: 26 (1923) [1924]
Notes: Pleurothecium species live in both terrestrial and freshwater environments. Currently, 15 species in the genus (Wijayawardene et al. 2022) (Fig. 95).
[See PDF for image]
Fig. 95
Phylogram generated from maximum likelihood analysis (RAxML) of the Pleurotheciaceae based on the combined ITS, LSU and rpb2 sequence data. Seventy-four taxa are included in the combined analyses, which comprise 2,700 characters with gaps. Canalisporium caribense (SS 03839) and Savoryella lignicola (NTOU791) were chosen as the outgroup. The best scoring RAxML tree with a final likelihood value of − 27,078.871595 is presented. The matrix had 1,468 distinct alignment patterns, with 37.01% of undetermined characters or gaps. The proportion of invariable sites was 0.396394. Estimated base frequencies were as follows: A = 0.215041, C = 0.319662, G = 0.285771, T = 0.179525; substitution rates: AC = 1.174315, AG = 2.403151, AT = 1.719399, CG = 0.666160, CT = 6.665075, GT = 1.00000; gamma distribution shape parameter α = 0.638780. Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.95 are given near the nodes. T = ex-type strain. The newly generated sequences are indicated in blue bold
Pleurothecium takense Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov.
Index Fungorum number: IF 900198; Facesoffungi number: FoF 15106; Fig. 96
[See PDF for image]
Fig. 96
Pleurothecium takense (BBH 49602, holotype). a Colonies on natural substrate. b–e Conidiophores, conidiogenous cells and conidia. f Conidiogenous cells. g–j Conidia. k–l Upper and reverse views of culture on PDA after 14 days at 25 °C. Scale bars: a = 200 μm, b–d, f = 20 μm, e = 50 μm, g–j = 10 μm
Etymology: The specific epithet “takense” refers to the Tak Province, Thailand where the taxon was collected.
Holotype: BBH 49602
Saprobic on submerged twigs in freshwater habitat. Sexual morph: not observed. Asexual morph:Colonies on natural substratum effuse, hairy, brown to black with visible whitish to yellowish spore masses. Mycelium 2.5–5 μm diam, partly immersed, partly superficial, composed of branched, septate, smooth-walled, pale brown to brown hyphae. Conidiophores 213–550 × 3.3–4.8 μm (x̄ = 370 × 4.1 μm, n = 20), macronematous, mononematous, solitary, simple, cylindrical, erect, straight or slightly flexuous, 6–11-septate, smooth-walled, brown to black below, paler towards the apex. Conidiogenous cells up to 35 × 3.7–7.5 μm, holoblastic, polyblastic, integrated, terminal, sometimes becoming intercalary, cylindrical, denticulate, sympodially elongating and recurved, pale brown to brown, with 3–13 or more cylindrical denticles, 1.3–2.8 × 1.2–2.3 μm. Conidiogenous locus occasionally formed an apical secondary conidiophore. Conidia 25–31 × 6.5–9 μm (x̄ = 27.7 × 7.7 μm, n = 20), acropleurogenous, solitary, ellipsoidal, slightly curved, 3-septate, smooth-walled, central cells light brown to becoming brown at maturity, end cells paler. Conidial secession schizolytic.
Culture characteristics: Colonies on PDA reaching 18–21 mm diam., after 14 days at 25 °C, cottony with sulcate, grey, round, margins mostly entire, soluble pigment absent, exudates absent, reverse dark grey.
Material examined: Thailand, Tak Province, Mae Sot District, Pha Daeng Waterfall Nature Trail, on submerged twigs of an unidentified plant in a small freshwater stream, 24 May 2022, N. Boonyuen, isolate FF01100 (BBH 49602, holotype), ex-type, TBRC-BCC 95074; isolate FF01100.01 (BBH 49603, isotype), living culture, TBRC-BCC 95075.
GenBank numbers: TBRC-BCC 95074: ITS = OQ121931, LSU = OQ121949, rpb2 = OQ116754, SSU: OQ121940; TBRC-BCC 95075: ITS = OQ121932, LSU = OQ121950, rpb2 = OQ116755, SSU: OQ121941, tef1-α = OQ116763.
Pleurothecium takense is phylogenetically related to P. floriforme, with 100% ML bootstrap and 1.00 BYPP supports (Fig. 95). Morphologically, the new taxon differs from P. floriforme in that it has pigmented conidia in maturity with light brown to brown central cells and paler end cells, whereas the latter has hyaline conidia (Hyde et al. 2017). Pleurothecium bicoloratum and P. recurvatum are most morphologically similar to P. takense, which has 3-septate, smooth-walled, pigmented conidia (Monteiro et al. 2016). Nevertheless, P. takense is different from P. recurvatum based on phylogenetic analysis (Fig. 95). Due to a lack of sequence data for P. bicoloratum, the phylogenetic position of P. bicoloratum could not be compared with our new species. Morphologically, P. bicoloratum differs in possessing strong pigment with dark olivaceous-brown to black at central cells and hyaline end cells and broadly allantoid conidia (Monteiro et al. 2016), whereas conidia in those two taxa (P. takense and P. recurvatum) are initially hyaline, later becoming pale brown to brown with central cells darker than end cells as they mature, with conidial shape varying from oblong to oblanceolate or somewhat falcate in P. recurvatum (Morgan 1895; Goos 1969), and ellipsoidal in P. takense (Fig. 96). Thus, one new species is described in this study as Pleurothecium takense.
Savoryellales Boonyuen, Suetrong, Sivichai, K.L. Pang & E.B.G. Jones
Savoryellaceae Jaklitsch & Réblová, in Jaklitsch, Index Fungorum 209: 1 (2015)
Notes: Savoryellaceae is a monotypic family in Savoryellales with six genera viz; Ascotaiwania, Canalisporium, Kaseifertia, Obliquifusoideum, Rhexoacrodictys, and Savoryella (Goh and Kuo (2021), Wijayawardene et al. 2022, Du et al. (2022)).
Rhexoacrodictys W.A. Baker & Morgan-Jones, Mycotaxon 82: 98 (2002)
Notes: Rhexoacrodictys was established by Baker & Morgan-Jones (2002) to accommodate four Acrodictys. These species are characterized by indeterminate and regeneratively percurrently extending conidiogenous cells that often acquire a narrow circumscissile dehiscence zone at the extreme apex. The conidia are ellipsoid, ovoid, subspherical to spherical, transversely, longitudinally, and obliquely septate, which undergo rhexolytic conidial secession. Currently, five species are accepted in Index Fungorum (Index Fungorum 2024) (Fig. 97).
[See PDF for image]
Fig. 97
Phylogram generated from maximum likelihood analysis (RAxML) of the Pleurotheciaceae based on the combined ITS, LSU and rpb2 sequence data. Seventy-four taxa are included in the combined analyses, which comprise 2700 characters with gaps. Canalisporium caribense (SS 03839) and Savoryella lignicola (NTOU791) were chosen as the outgroup. The best scoring RAxML tree with a final likelihood value of − 27,078.871595 is presented. The matrix had 1468 distinct alignment patterns, with 37.01% of undetermined characters or gaps. The proportion of invariable sites was 0.396394. Estimated base frequencies were as follows: A = 0.215041, C = 0.319662, G = 0.285771, T = 0.179525; substitution rates: AC = 1.174315, AG = 2.403151, AT = 1.719399, CG = 0.666160, CT = 6.665075, GT = 1.00000; gamma distribution shape parameter α = 0.638780. Bootstrap support values for ML equal to or greater than 70% and BYPP equal to or greater than 0.95 are given near the nodes. T = ex-type strain. The newly generated sequences are indicated in blue bold
Rhexoacrodictys fangensis Chuaseehar., Nuankaew, Somrith. & Boonyuen, sp. nov.
Index Fungorum number: IF 900200; Facesoffungi number: FoF 15107; Fig. 98
[See PDF for image]
Fig. 98
Rhexoacrodictys fangensis (BBH 49598, holotype). a, b Colonies on natural substrate. c–f Conidiophores and conidia. g conidiophore. h–k Conidia. l–o Upper and reverse views of culture on PDA (the top) and MEA (the bottom), after 14 days at 25 °C. Scale bars: a = 100 μm, b = 25 μm, c = 20 μm, d–k = 10 μm
Etymology: The specific epithet “fangensis” refers to the Fang District, Chiang Mai Province, Thailand where is the type was collected.
Holotype: BBH 49598
Saprobic on submerged wood in freshwater habitat. Sexual morph: Not observed. Asexual morph:Colonies on natural substratum effuse, scattered, granulate, black. Mycelia 2–2.5 μm diam, mostly immersed, composed of branched, septate, smooth-walled, pale brown to brown hyphae. Conidiophores 12.5–37.5 × 3.7–5 μm (x̄ = 20 × 4.3 μm, n = 20), macronematous, mononematous, solitary, simple, cylindrical, erect, straight, 1–5-septate, smooth-walled, brown, with 1 or more percurrent extensions at the upper part. Conidiogenous cells holoblastic, monoblastic, integrated, terminal, cylindrical, brown. Conidia 20–30 × 12.7–22.5 μm (x̄ = 26.2 × 15.4 μm, n = 20), acrogenous, solitary, obovoid to obpyriform, oblong-ellipsoidal, muriform, with multiple longitudinal or oblique and transverse septa, smooth-walled, brown to dark brown and basal cell paler, apex rounded, basal cell protruding, cylindrical, with a distinct, 1–7 × 3.5–4 μm basal frill. Conidial secession rhezolytic.
Culture characteristics: Colonies after 14 days at 25 °C: On MEA attaining 13–14 mm diam, glabrous, white, round, margins entire, soluble pigment absent, exudates absent, reverse yellowish white with dark grey at center. On PDA reaching 16–25 mm diam, velvety with sulcate, olive grey with white margins, irregular, margins undulate, soluble pigment absent, exudates clear droplets, reverse medium grey with white margins.
Material examined: Thailand, Chiang Mai Province, Fang District, Khlong Pong Nam Dang, on submerged wood of an unidentified plant in a freshwater stream, 24 November 2021, N. Boonyuen and S. Nuankaew, isolate FF01084 (BBH 49598, holotype), ex-type, TBRC-BCC 94389; isolate FF01084.01 (BBH 49599, isotype), living culture, TBRC-BCC 94390.
GenBank numbers: TBRC-BCC 94389: ITS = OQ121933, LSU = OQ121951, rpb2 = OQ116756, SSU: OQ121942, tef1-α = OQ116764; TBRC-BCC 94390: ITS = OQ121934, LSU = OQ121952, rpb2 = OQ116757, SSU: OQ121943, tef1-α = OQ116765.
Notes: Rhexoacrodictys fangensis differs from the other accepted Rhexoacrodictys species (Species Fungorum; accessed 23 May 2024) by having the shortest conidiophores with one or more percurrent extensions at the apical part. In multi-gene phylogenetic analysis, based on the combined ITS, LSU, and rpb2 dataset (Fig. 97), the new species forms a sister relationship with R. fimicola. However, the support values for separation are low. Rhexoacrodictys fimicola is distinguished from R. fangensis by producing conidia which is evenly pigmented or without a dark brown basal cell and having longer conidiophores (up to 75 × 5–7 μm; Baker et al. 2002), while R. fangensis has unevenly pigmented conidia in brown to dark with basal cell paler and shorter conidiophores. Thus, our collection is different from known Rhexoacrodictys species, and it formed a distinct subclade in phylogenetic analyses. Based on the morphology and molecular data, we introduce our collection here as a new species.
[See PDF for image]
Fig. 99
Phylogram generated from maximum likelihood analysis based on combined ITS and LSU sequence data. Twenty-nine taxa were included in the combined analyses, which comprised 1479 characters (ITS: 605, LSU: 874) after alignment. Bootstrap support values for ML and MP equal to or greater than 75% and BYPP equal to or greater than 0.95 are given above the nodes. Arthrinium arundinis AFTOL-ID 951 and A. arundinis CBS 133509 were used as the outgroup taxa. The newly generated strain is shown in blue and bold. Ex-type strains are indicated by black and bold
Subclass Sordariomycetidae O.E. Erikss & Winka (= Meliolomycetidae P.M. Kirk & K.D. Hyde)
Pseudodactylariales Crous in Crous et al., Persoonia 39: 421 (2017)
Notes: Crous et al. (2017a) introduced Pseudodactylariales to accommodate Pseudodactylariaceae, and Pseudodactylaria (Fig. 99).
Pseudodactylariaceae Crous, in Crous et al., Persoonia 39: 421 (2017)
Notes: Pseudodactylariaceae is a monotypic family in Pseudodactylariales (Wijayawardene et al. 2022).
Pseudodactylaria Crous, Persoonia 39: 421 (2017)
Notes: Pseudodactylaria was introduced by Crous et al. (2017a, b, c) with P. xanthorrhoeae as the type species. Currently, 10 species have been accepted in Pseudodactylaria (Bao et al. 2021; Crous et al. 2017a, b, c; Hyde et al. 2020a, b, c; Lin et al. 2018; Lu et al. 2020; Boonmee et al. 2021). Most of the Pseudodactylaria species have been reported from Asia, primarily in Thailand and China, while one species viz P. xanthorrhoeae is from Australia (Bao et al. 2021; Crous et al. 2017a, b, c; Hyde et al. 2020a, b, c; Lin et al. 2018; Lu et al. 2020; Boonmee et al. 2021). It should be noted that only Pseudodactylaria xanthorrhoeae has been reported as an endophytic species (Crous et al. 2017a, b, c).
Pseudodactylaria guttulate J. Ma & Y.Z. Lu, sp. nov.
Index Fungorum number: IF 901521; Facesoffungi number: FoF 14265 Fig. 100
[See PDF for image]
Fig. 100
Pseudodactylaria guttulate (GZAAS 22-2037, holotype). a, b Colonies on decaying wood; c, d Conidiophores and conidiogenous cells; e–f Conidiogenous cells with attached conidia; g–i Conidiogenous cells; j Germinated conidia; k–n Conidia; o, p Colony on PDA from above and below. Scale bars: c–d = 20 μm; e–j = 10 μm; k–n = 5 μm
Etymology: “guttulate” referring to the guttulate conidia.
Holotype: GZAAS 22-2037.
Saprobic on decaying wood. Sexual morph: Not observed. Asexual morph: hyphomycetous. Colonies on the substratum superficial, effuse, gregarious, white. Mycelium composed of partly immersed, hyaline, septate, branched hyphae. Conidiophores 45–110 × 3–4 μm (x ̅ = 82 × 3.5 μm, n = 20), macronematous, mononematous, solitary, erect, occasionally branched, subcylindrical, straight to slightly flexuous, hyaline, 1–3-septate. Conidiogenous cells 20–30 × 3–3.5 μm (x ̅ = 24 × 3.2 μm, n = 20), holoblastic, polyblastic, integrated, terminal, sympodial, subcylindrical, straight or flexuous, hyaline, apical part forming a rachis with numerous, aggregated, cylindrical denticles, 1.0–1.2 × 0.6–1.0 μm. Conidia 14–20 × 3–3.5 μm (x ̅ = 18 × 3 μm, n = 30).solitary, acropleurogenous, smooth, prominently guttulate, 0–1-septate, fusiform, hyaline.
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 23 mm in 40 days at 25 °C, greyish white.
Material examined: China, Guizhou Province, Guiyang City, Changpoling National Forest Park, on decaying wood in a terrestrial habitat, 20 August 2022, Yong-Zhong Lu, JW3.1 (GZAAS 22–2037, holotype), ex-type, GZCC 22–2037.
GenBank accession numbers: LSU: OR807743, ITS: OR807742
Notes: Pseudodactylaria guttulate shares a sister relationship to P. camporesiana with strong support (98% ML/ 98% MP/ 1.00 PP). However, P. guttulate differs from P. camporesiana in having longer conidiophores (45–110 × 3–4 μm vs. 35–45 × 3.5–5 μm) (Hyde et al. 2020a, b, c). Particularly, P. guttulate is distinguished from P. camporesiana by distinctly branched conidiophores (Hyde et al. 2020a, b, c). Based on a pairwise comparison of ITS nucleotides, P. guttulate differs from P. camporesiana by 26/693 bp (3.8%). Following the guidelines for defining species boundaries of Chethana et al. (2021a, b), we therefore introduce GZCC 22–2037 as a new species.
Sordariales Chad. ex D. Hawksw. & O.E. Erikss.
Chaetomiaceae G. Winter [as ‘Chaetomieae’], Rabenh. Krypt.-Fl., Edn 2 (Leipzig) 1.2: 153 (1885)Chaetomiaceae species are ecological diversity and well-known indoor contaminants such as Chaetomium globosum (Ahmed et al. 2002; Samson et al. 2019). These species have worldwide distribution as saprobes and occur in soil, air, decaying plant materials, and indoor environments. Currently, 50 genera are accepted in this family (Wang et al. 2022a, b, Yang et al. 2024).
Collariella X. Wei Wang, Samson & Crous, Stud. Mycol. 84: 177. 2016.
Collariella was introduced by Wang et al. (2016), and typified by C. bostrychodes. This is a cosmopolitan genus consisting of 13 legitimate species that were accepted in Index Fungorum (2023), many of which are saprobic, and are widely distributed in air, dung, dust, insects, soil and plant (Wang et al. 2016. 2022; Crous et al. 2017a, b, c; Zhang et al. 2017; Aghyl et al. 2020). In this study, we introduced C. hongheensis as a new species that was isolated from living leaves of Mangifera indica (Figs. 101, 102).
[See PDF for image]
Fig. 101
Phylogram generated from maximum likelihood analysis based on combined LSU, ITS, rpb2 and tub sequences data. Related sequences were obtained from Wang et al. (2016, 2022a, b). Twenty-seven strains are included in the combined sequence analysis, which comprises 2626 characters with gaps. Chaetomium globosum (CBS 148.51) and C. globosum (CBS 160.62) were used as the outgroup taxa. The tree topology of the ML analysis was similar to the BYPP. The best-scoring RAxML tree with a final likelihood value of − 9562.850095 is presented. The matrix had 647 distinct alignment patterns, with 17.07% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.226172, C = 0.289966, G = 0.278318, T = 0.205544; substitution rates AC = 0.911184, AG = 2.964801, AT = 1.311409, CG = 1.266535, CT = 6.151905, GT = 1.000000; proportion of invariable sites I = 0.558684; gamma distribution shape parameter α = 1.168086. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given above the nodes. Newly generated sequences are in red. Type strains denoted by bold
[See PDF for image]
Fig. 102
Collariella hongheensis (KUNCC 22-10749 ex-type). a, b Colony characteristics on PDA (60 days old culture). c, d, f Ascomata. e Ascomatal hairs. g Masses of ascospores. h, j–m Asci. i Hamathecium. n, o Ascospores. Scale bars: d–g = 200 μm; h = 50 μm; i, m = 30 μm; j–l, n, o = 20 μm (h–m, n treated with Melzer’s reagent)
Collariella hongheensis X.F. Liu, Karun. & Tibpromma sp. nov.
Index Fungorum number: IF 902097; Facesoffungi number: FoF 15838, Fig. 102
Etymology: Named after the location Honghe, where the fungus was first discovered.
Endophytic on Mangifera indica leaves. Sexual morph: Produced on 60 days old PDA culture, Vegetative hyphae hyaline, septate, branched, smooth-walled, 1–5 µm diam wide. Ascomata 250–500 µm high, 200–300 µm diam., superficial, subglobose, oval to fusiform, pale mouse grey to black, with rounded base, ostiolate, neck inconspicuous. Ascomatal wall brown, textura globulosa to angularis in surface view. Terminal hairs arising from the apical collar, conspicuously rough, dark brown, septate, erect in the lower part, 2.5–6 μm near the base, spirally coiled in the upper part. Lateral hairs seta-like, tapering and fading towards the tips. Hamathecium 2–4 μm wide (x̄ = 3.19 μm, n = 20), hyaline, septate paraphyses. Asci 30–63.5 × 5–13 μm (x̄ = 42.30 × 8.65 μm, n = 30), fasciculate, clavate or fusiform, eight-spored, evanescent, hyaline, long-stalked. Ascospores 5–7.5 × 4–7 μm (x̄ = 6.47 × 5.57 μm, n = 90), limoniform to broadly limoniform, olivaceous to brown, usually biapiculate at both ends, bilaterally flattened, with an apical germ pore.
Asexual morph: Not observed.
Culture characteristics: Colonies on PDA attaining 30–40 mm diam after 15 d, felty, margin entire, white to black, forming dense and pale mouse grey to mouse grey ascomata. Reverse brown to black.
Material examined: China, Yunnan Province, Honghe Menglong Village, on healthy leaves of Mangifera indica, (102° 50′ 11″ E, 23° 41′ 01″ N, 500 m), 24 July 2019, E.F. Yang, EFA (MHZU 23-0263, holotype), ex- type KUNCC 22-10749; ibid. EFB, living culture, KUNCC 22-10757.
GenBank numbers: KUNCC 22-10749 = LSU: PQ014642, ITS: PQ014644, rpb2: PP766708, tub2: PP766710; KUNCC 22-10757 = LSU: PQ014643, ITS: PQ014645, rpb2: PP766709, tub2: PP766711.
Notes: Collariella hongheensis is phylogenetically closely related to C. pachypodioides (CBS 164.52, type), C. carteri (CBS 128.85, type) and C. capillicompacta (IRAN 3496C, type) (Fig. 101). Based on the sequence comparisons, Collariella hongheensis (KUNCC 22-10749 holotype) is different from C. pachypodioides (CBS 164.52 Type) in 0.53% (3/563 bp) of ITS, 0.22% (2/892 bp) of LSU, 2.23% (19/852 bp) of rpb2 and 3.29% (22/669 bp) of tub2; different from C. carteri (CBS 128.85 Type) in 0.36% (2/562 bp) of ITS, 0% (0/561 bp) of LSU, 1.92% (10/522 bp) of rpb2 and 4.20% (28/667 bp) of tub2; different from C. capillicompacta (IRAN 3496C Type) in 2.29% (11/481 bp) of RPB2 and 3.94% (21/533 bp) of tub2. Collariella hongheensis differs from C. pachypodioides by the latter having elongated ovate ascomata and only produces spirally coiled terminal hairs (Greathouse and Ames 1945; Wang et al. 2022a, b). Collariella hongheensis differs from C. carteri by the latter having short and seta-like terminal ascomatal hairs, smaller hyaline asci (30–63.5 × 5–13 μm vs. 17–26 × 8.5–11.5 μm), and ascospores limoniform, olivaceous when mature (Wang et al. 2016). Collariella hongheensis differs from C. capillicompacta by the latter having high-density and very compacted terminal hairs, significantly smaller hyaline asci (30–63.5 × 5–13 μm vs 21–28 × 10–11 μm), and smaller ascospores (5–7.5 × 4–7 μm vs. 5.2–6.2 × 4.2–5.2 μm) (Aghyl et al. 2020). Based on the phylogenetic analyses and morphological characteristics, our species is a distinct new species. In addition, in the previous study, the hamathecium of Collariella species was not observed and described. Therefore, this is the first observation of hamathecium in Collariella. Meanwhile, Chaetomiaceae species are a dominant group of endophytic fungi in mango leaves in Yunnan, China; they can also fight against plant fungal or bacterial pathogens (Yang et al. 2023a, b, 2024). Our new species was also isolated from living leaves of mangoes in Yunnan, and it might have potential antagonistic effects. Pathogen antagonistic testing needs to be conducted in the future to confirm its ability to control pathogens.
SubclassXylariomycetidae O.E. Erikss & Winka
Amphisphaeriales D. Hawksw. & O.E. Erikss.
Notes: For taxonomic treatments, we follow Hyde et al. (2020a, b, c) and Wijayawardene et al. (2022).
Apiosporaceae K.D. Hyde, J. Fröhl., Joanne E. Taylor & M.E. Barr, in Hyde, et al., Sydowia 50(1): 23 (1998).
Notes: Apiosporaceae was introduced by Hyde et al. (1998) and typified by Apiospora (Ellis 1971; Seifert et al. 2011; Hyde et al. 2020a, b, c), in the order Amphisphaeriales (Wijayawardene et al. 2020). Apiosporaceae species have a worldwide distribution and contain important pathogens and saprophytes in a wide range of hosts (Kwon et al. 2021; Pintos and Alvarado 2022). The divergence time of Apiosporaceae has been estimated at 69 Mya (50–130) in the late Cretaceous (Hyde et al. 2017).
Apiospora Sacc., Atti Soc. Veneto-Trent. Sci. Nat., Padova, Sér. 4 4: 85 (1875)
Notes: Apiospora was introduced by Saccardo (1875), with A. montagnei as the type species. This genus is reported as having sexual and asexual morphs. The sexual morph of Apiospora is characterized by hyaline ascospores surrounded by a gelatinous sheath. The asexual morph of Apiospora is characterized by conidia lenticular with a paler equatorial slit in the side view, globose to ellipsoid conidiogenous cells hyaline to pale brown to brown (Dai et al. 2017; Feng et al. 2021; Kwon et al. 2021; Tian et al. 2021). Apiospora species are distributed worldwide and can be found in different hosts (Dai et al. 2017; Wang et al., 2018; Hyde et al. 2020a, b, c; Tian et al. 2021; Liu et al. 2022, 2023a, b). This study introduces a new Apiospora species based on morphology and molecular phylogenetic analyses (Fig. 103).
[See PDF for image]
Fig. 103
The RAxML tree is based on a combined dataset of ITS, LSU, tef1-α, and tub2 gene sequences. Bootstrap support values for maximum likelihood (ML) equal to or higher than 60% and Bayesian posterior probability (BYPP) equal to or higher than 0.90 are indicated above the branches. Newly generated sequences are shown in red, while the ex-type species are shown in bold black. The RAxML analyses of the combined dataset yielded the best scoring tree, with a final ML optimization likelihood value of − 47,864.244581. The matrix had 1797 distinct alignment patterns, with 44.78% of undetermined characters or gaps. Parameters for the GTR + I + G model of the combined ITS, LSU, tef1-α, and tub2 were as follows: estimated base frequencies A = 0.235844, C = 0.256476, G = 0.253489, T = 0.254191; substitution rates AC = 1.148171, AG = 2.788122, AT = 1.092424, CG = 1.194513, CT = 4.146085, GT = 1.0; proportion of invariable sites I = 0.452913; and gamma distribution shape parameter α = 0.728911
Apiospora hongheensis X. Zhang & Tibpromma, sp. nov.
Index Fungorum number: IF 901411; Facesoffungi number: FoF 15837; Fig. 104
[See PDF for image]
Fig. 104
Apiospora hongheensis (ZHKU 23–0100, holotype). a–c Appearance of the fungus on dead culms of an unidentified plant. d–e Conidia with conidiophores. f–h Conidiogenous cells bearing conidia. i–k Conidia. l Germinated conidium. m, q Colony on PDA media (m reversed, q forward). n Sporulation on PDA. o Conidia. p Conidiogenous cells and conidia. Scale bars: d, e = 40 μm; f–k, p = 20 μm; l, o = 50 μm
Etymology: Named after Honghe City in China, where the fungus was first discovered.
Holotype: MHZU 23-0100
Saprobic on dead culms of unidentified branches. Sexual morph: Not observed. Asexual morph:Colonies on natural substrate surface, powdery, dark brown to black, gregarious, dull with conidia readily liberated when disturbed. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 10–30 × 4–7 µm (x̄ = 21 × 5 µm, n = 30), cylindrical, hyaline, smooth-walled, unbranched, erect or slightly curved. Conidia 15–25 × 10–15 µm (x̄ = 19 × 12.5 µm, n = 40), subglobose to ellipsoidal, hyaline when immature, brown to dark brown when mature, smooth-walled to finely roughened, guttulate, lacking germ-slit.
Culture characteristics: Conidia germinated on PDA within 24 h. Colonies on PDA at 25–28 ℃ after one month, irregular, flat, undulate edge with white to gray at the margin, the center is black, wrinkled in above, the reverse is white in the middle and gray the margin. Sporulation after two months.
Material examined: China, Yunnan Province, Honghe City, on dead culms of an unidentified plant,17 July 2021, HH03 (holotype, MHZU 23-0100); ex-type, ZHKUCC 23-0792 = ZHKUCC 23-0793.
GenBank numbers: ZHKUCC 23-0792 = ITS: OR936320, LSU: OR936322, tef1-α: PP778354, tub2: PP778365; ZHKUCC 23-0793 = ITS: OR936321, LSU: OR936323, tef1-α: PP778355, tub2: PP778366.
Notes: Notes: In the present phylogenetic analyses, Apiospora hongheensis (ZHKUCC 23-0792, 23-0793) forms a separate lineage within Apiospora with 86% ML statistical support and is sister to A. malaysiana (CBS 102053) and A. euphorbiae (IMI 285638b). In morphology, A. hongheensis is distinct from A. malaysiana (CBS 102053) and A. euphorbiae (IMI 285638b) by the morphology of colonies, conidiogenous cells, and conidia. Conidiogenous cells of A. malaysiana aggregate in clusters on hyphae, hyaline to pale brown, doliiform to clavate to ampulliform, conidia globose in surface view, lenticular inside view, with pale equatorial slit (Ellis, 1963, 1965; Crous et al. 2013), while conidiogenous cells of A. hongheensis are cylindrical, hyaline, erect, or slightly curved; conidia are subglobose to ellipsoidal and thin walled. Based on nucleotide comparisons, A. malaysiana is slightly different from A. hongheensis in 70/605 bp (11.67%) of the ITS, 19/831 (2.29%) of the LSU, 12/549 bp (2.19%) of tub2 and 10/393 bp (2.54%) of tef1-α. Based on nucleotide comparisons, A. euphorbiae is slightly different from A. hongheensis in 10/605 bp (1.65%) of the ITS, 12/831 (1.44%) of the LSU, and 13/549 bp (2.37%) of tub2. Additionally, colonies of A. euphorbiae are effused, and conidiophores are erect or ascending, simple, flexuous, cylindrical, colourless, except for the numerous brown transverse septa, smooth-walled, thick lenticular conidia, while colonies of A. hongheensis are powdery, dark brown to black, gregarious, dull with conidia readily liberated when disturbed, conidiophores reduced to conidiogenous cells, and conidia are subglobose to ellipsoidal, thin-walled.
Nigrospora Zimm., Centbl. Bakt. ParasitKde, Abt. I 8: 220 (1902)
Nigrospora was introduced by Zimmerman in 1902 and typified as Nigrospora panici. Nigrospora species have a cosmopolitan distribution with a wide range of hosts and are capable of existing as saprobes, endophytes, or plant and human pathogens (Chen et al. 2022a, b). These species are species are characterized by branched, smooth, micronematous or semi-macronematous conidiophores, monoblastic conidiophores, globose, or sub-globose, pale to black and aseptate conidia (Zhang et al. 2020). Currently, 45 species are accepted under this genus in Species Fungorum (2023). In this study, one new species N. tomentosae is described based on ITS, tef1-α and tub2 phylogeny (Fig. 105).
[See PDF for image]
Fig. 105
The best-scoring RAxML tree for the combined dataset of ITS, tub2 and tef1-α sequence data of Nigrospora following Chen et al. (2022a, b). The tree is rooted in Arthrinium malaysianum (CBS 102053). The matrix had 501 distinct alignment patterns with 13.35% undetermined characters and gaps. Estimated base frequencies were as follows; A = 0.215381, C = 0.290983, G = 0.253160, T = 0.240476; substitution rates AC = 0.865842, AG = 2.403689, AT = 0.630710, CG = 0.956682, CT = 4.524366, GT = 1.00000; gamma distribution shape parameter α = 0.527542 with a final likelihood value of − 6763.967328. For the Bayesian inference, SYM + I + G model was selected for ITS, HKY + G for tub2, TPM2uf + I + G for tef1-α. Ex-type strains with T and newly generated sequences are in blue. Bootstrap support values for ML equal to or greater than 50% and BYPP equal to or greater than 0.90 are given at the nodes
[See PDF for image]
Fig. 106
Nigrospora tomentosae (ZHKUCC22-0339, ex-type). a, b. Upper surface and reverse overview of culture 7d after inoculation on PDA medium; c. colony on SNA; d–g. conidiogenous cells giving rise to conidia; h, i. conidia. Scale bars: d–i = 10 μm
Nigrospora tomentosae M. Luo, & J.W. Liu, sp. nov.
Index Fungorum number: IF902096, Facesoffungi number: FoF15835 Fig. 106
Etymology: In reference to its host Citrus grandis cv. ‘Tomentosa’.
Endophytic on Citrus grandis cv. ‘Tomentosa’. Sexual morph: Not observed. Asexual morph: Conidiophores reduced to conidiogenous cells and aggregated in clusters on hyphae. Conidiogenous cells hyaline to pale brown, globose to pot-shaped or clavate, 9–12 × 5–7 μm, aseptate. Conidia 12–17 × 12–17 μm (x̄ = 14.4 ± 1.5 × 14.1 ± 1.3, n = 50), solitary, globose or subglobose, smooth, aseptate, dark brown to black. Chlamydospore thick-walled, rough surface, pale brown, subglobose to globose, 9–11 × 6–10 μm (x̄ = 9.8 ± 0.7 × 8.1 ± 1.2, n = 50).
Culture characteristics: On PDA, colonies blanket, edge entire, off-white, reaching 9 cm diam after 7 d at 25 ℃, reverse pale yellow. On SNA, colonies flat, aerial mycelia dense with patches of sporulation, surface olive and reverse grey.
Material examined: China, Guangdong Province, Huazhou city, healthy fruit of Citrus grandis cv. ‘Tomentosa’, May 2021, B. Y. Cen, (MHKU 22-0173; holotype as dry culture), ex type ZHKUCC 22-0339 and living culture ZHKUCC 22-0340.
GenBank numbers: ITS = PP759659, tef1-α = PP763294, tub2 = PP763296
Notes: Two strains representing N. tomentosae clustered in a well-supported clade with 76% ML bootstrap and 1.0 BYPP which is sister to N. camelliae-sinensis (Fig. 105). Morphologically, N. tomentosae differs from N. camelliae-sinensis (Wang et al. 2017a, b) in its conidiogenous cells (9–12 × 5–7 μm vs. 6–11 × 4.5–8.5 μm). Also, N. tomentosae differs from N. pyriformis (Wang et al. 2017a, b) in conidial size (12–17 × 12–17 μm vs. 17.5–27.5 × 10–18.5 μm) and shape (globose or subglobose vs. water prop shapes). In pairwise nucleotide comparison, N. tomentosae is different from N. camelliae-sinensis (CGMCC 3.18125 Type) in 0.4% (2/513 bp) of tub2, and 1.7% (7/404 bp) of tef1-α.
Oxydothidaceae Konta & K.D. Hyde, Mycosphere 7 (9): 1425–1446 (2016)
Oxydothidaceae was introduced by Konta et al. (2016) to accommodate the monotypic genus Oxydothis. Following the updated outline of fungi, Oxydothidaceae is placed on Amphisphaeriales (Wijayawardena et al. 2022). Oxydothidaceae species are found as saprobes or endophytes and occur mainly on the leaves, trunks and branches of palms. However, they are occasionally also found on other monocotyledons (Fröhlich & Hyde 1994; Hyde 1994; Pinnoi et al. 2006).
Oxydothis Penz. & Sacc., Malpighia 11: 505 (1898)
Oxydothis was described by Penzig & Saccardo (1897) to accommodate O. grisea (Penz et al. 1898). The taxonomic status of Oxydothis has been variously classified in the Amphisphaeriaceae, Clypeosphaeriaceae, and Hyponectriaceae (Eriksson and Hawksworth 1991; Hawksworth et al. 1995; Kang et al. 1998, 1999a, b). It is presently placed in Oxydothidaceae within Xylariales (Konta et al. 2016). Oxydothis species have darkened blister-like ascomata, cylindrical asci with J+, subapical rings and fusiform, hyaline ascospores. Eighty-seven species of Oxydothis are listed in Index Fungorum (2024), but most Oxydothis species lack sequence data (Hyde et al. 2023a) (Fig. 107).
[See PDF for image]
Fig. 107
Phylogram generated from maximum likelihood analysis (RAxML) of Oxydothidaceae, and its related family based on ITS, LSU and SSU sequence data. The tree is rooted in Arthrinium hydei (CBS 114990) and A. phaeospermum (HKUCC 3395). Maximum likelihood bootstrap values equal to or above 75%, and Bayesian posterior probabilities equal to or above 0.99 (MLBS/PP) are given at the nodes. Hyphen (–) represents support values below 75% MLBS and 0.95 PP. The original isolate number is noted after the species name. The ex-type strains are indicated in bold. Isolates from the present study are in given in blue
Oxydothis aquatica Y. Y. Yang & K.D. Hyde, sp. nov.
Index Fungorum number: IF 902262; Facesoffungi number: FoF 15184; Fig. 108
[See PDF for image]
Fig. 108
Oxydothis aquatica (HKAS 131230, holotype) a. Appearance of fruiting bodies on host substrate. b–c. Close-up of ascomata. d–e. Section of ascoma. f–g. Asci. h–m. Ascospores. Scale bars: d–e = 50 μm, f = 100 μm, g = 50 μm, h–m = 20 μm
Etymology: “aquatica” refers to the collection of from aquarium habitats of this fungus.
Holotype: HKAS 131230.
Saprobic on the dead stem of Trachycarpus fortunei. Asexual morph: Undetermined. Sexual morph: Ascomata 220–260 × 38–48 µm (x¯ = 243 × 45, n = 5), solitary or clustered in groups, immersed, black, ellipsoidal, forming under slightly raised, split at maturity and contents overflow. Peridium 2–5 µm wide, outer cells merging with the host epidermal cells, comprising dark brown to hyaline, walled cells of textura angularis. Asci 102–187 × 7–9 μm (x̅ = 148 × 8 μm, n = 20), unitunicate, pedicellate, fasciculate, cylindrical, smooth-walled, subapical ring, J+, with a wedg-eshaped. Ascospores 54–69 × 3–4 µm (x̅ = 62 × 3.8, n = 30), hyaline, fusiform, smooth-walled, guttulate, tapering gradually, 1–2-seriate (delete), tapering gradually from the center to the ends, centrally 1-septate, constricted at the septum, with pointed ends.
Material examined: China, Yunnan Province, Kunming City, Panlong District, on a dead stem of Trachycarpus fortunei in aquarium. 9 December 2021, Rong-Ju Xu, YYY183 (holotype: HKAS 131230).
GenBank numbers: ITS: OR809208, LSU: OR809209, SSU: OR809210
Notes: The phylogeny based on multi-gene analysis indicates that Oxydothis aquatica is a distinct clade from O. inaequalis with 100% ML,1.00 PP statistical support. Morphologically, O. aquatica differs from O. inaequalis by having short and narrow asci (102–187 × 7–9 μm vs. 200–285 × 11.25–12.5 µm), and small ascospores (54–69 × 3–4 µm vs. 78–87.5 × 5–7.5 µm). Additionally, O. aquatica having bundle arrangement ascospores, but O. inaequalis having 1–2 seriate ascospores. Oxydothis aquatica is similsr to O. calamicola and O. rhapidicola in its in asci and ascospore shape. However, Oxydothis aquatica differs from both species in forming subcircular blackened regions on the host surface, and ascospore more than slender. Oxydothis calamicola forms a black hollow oval on the host surface, O. rhapidicola forming globose to subglobose rims on the host surface, with dark ostiolar dots at the center.
Oxydothis filiforme L.W. Li & Jian K. Liu, sp. nov.
Index Fungorum number: IF 902399; Facesoffungi number: FoF 16062; Fig. 109
[See PDF for image]
Fig. 109
Oxydothis filiforme (HKAS 125889, holotype): a–c. Ascomata on host substrate. d, e Transverse section of ascomata. f Peridium. g, h, j Asci. i Asci in Congo red. k J + reaction of apical ring in Melzer’s reagent. l–n Ascospores. o Ascospores in Congo red. p Germinated ascospore. q Colonies from above (on PDA). r Colonies from below (on PDA). Scale bars: d, e, g = 100 μm, f, h–p = 10 μm
Etymology: The name reflects the shape of Asci
Holotype: HKAS 125889
Saprobic on decaying branches of Trachycarpus fortunei. Sexual morph: Ascomata 283–292 × 305–323 μm (x̅ = 287.5 × 314 μm, n = 10), solitary or aggregated, immersed, under slightly raised areas, visible as blister-like regions on the host surface, long axis horizontal to that of the host, ostiolate. Ostioles 102–158 μm wide, eccentric, conspicuous, curving upwards and piercing the host cuticle. Peridium 31–38.5 (x̅ = 35 μm, n = 10), comprising 8–10 layers of brown-walled cells of textura angularis. Hamathecium few, filamentous, septate, tapering distally. Asci 102–128 × 11–13 μm (x̅ = 115 × 12 μm, n = 20), 8-spored, fasciculate, unitunicate, cylindrical-clavate, pedicellate, with a J + , wedge-shaped, subapical ring, 2.6–3.6 μm high, 3.2–3.5 diam, and faint canal leading to the tip. Ascospores 67.5–77 × 4.5–5.5 μm (x̅ = 72 × 5 μm, n = 20), filiform, tapering towards both ends, centrally 1-septate, not constricted at the septum, hyaline, with multi-guttule in each cell, smooth-walled. Asexual morph: Not observed.
[See PDF for image]
Fig. 110
Phylogram generated from maximum likelihood analysis based on combined LSU and ITS sequence data. The tree is rooted with Barrmaelia macrospora (CBS 142768) and B. rhamnicola (CBS 142772). ML bootstrap values equal to or greater than 75% and posterior probabilities equal to or greater than 95% are given above the nodes (ML/BYPP). Ex-type strains are in bold, and the new isolate is indicated in blue
Culture characteristics: Ascospores germinating on PDA within three days. Colonies on PDA, white to smoky-grey, rough on the surface, margins smooth, dense at the center, with fairly fluffy.
Material examined: China, Sichuan Province, Chengdu, University of Electronic Science and Technology of China (UESTC) campus (Qingshuihe), on the dead branch of Trachycarpus fortunei (Hook.) H. Wendl. (Arecaceae), 25 Nov. 2020, W.L. Li, WD25 (HKAS 125889, holotype; HUEST 22.0062, isotype), ex-type living cultures CGMCC 3.24160 = UESTCC 22.0061. WD25b (HUEST 22.0063, paratype), living cultures UESTCC 22.0062.
GenBank numbers: UESTCC 22.0061: ITS: OQ171260; LSU: OQ171229; SSU: OQ171231. UESTCC 22.0062: ITS: OQ171261; LSU: OQ171230; SSU: OQ171231.
Notes: In the phylogenetic analysis, our stain of Oxydothis filiforme (UESTCC 22.0061, UESTCC 22.0062) clustered together with the ex-type strain of O. inaequalis with absolute bootstrap support (ML 100%). Oxydothis filiforme is morphologically similar to O. inaequalis as of the report by Hidayat et al. (2006) with immersed, ellipsoid ascomata and fusiform, 1-septate, hyaline ascospores. However, O. inaequalis differs from O. filiforme in having longer asci (200–285 vs. 102–128) and larger ascospores (78–87.5 × 5–7.5 μm vs. 67.5–77 × 4.5–5.5 μm). Additionally, ascospores of O. inaequalis are arranged 1–2 seriate or partially overlapping within asci but arranged in bundles in those of O. filiforme. Oxydothis filiforme resembles previously known but without molecular data species O. maculosa and O. luteaspora in having filiform, straight, 1-septate ascospore, J + asci and their ascomata parallel to the host surface. However, Oxydothis filiforme has shorter and thinner ascospores than O. luteaspora (long: 67.5–77 μm vs. 84–100 μm; wide: 4.5–5.5 μm vs. 6.5–8.5 μm). Oxydothis maculosa can be distinguished from O. filiforme in its ascal ring size (2–2.5 × 2–3 μm vs. 2.6–3.6 × 3.2–3.5 μm) and morphology (discoid vs. rectangular).
Xylariales Nannf., Nova Acta R. Soc. Scient. upsal., Ser. 4 8(no. 2): 66 (1932)
For taxonomic treatments, we follow Wijayawardene et al. (2022).
Cainiaceae J.C. Krug, Sydowia 30(1–6): 123 (1978) [1977]
Notes: Cainiaceae was introduced by Krug (1978) with Cainia as the type genus. Cainiaceae was described based on the unique structure of the apical apparatus and the unique ascospore characteristics. The family was later revived by Kang (1999) to accommodate Atrotorquata, Arecophila, Cainia, Ceriophora and Reticulosphaeria. The important features of the family are ascomata which are generally immersed below a clypeus, unitunicate asci with a complex J+, apical apparatus comprising sets of rings, and brown ascospores which may have longitudinal germ slits, striations, reticulate ornamentations and/or germ pores. Currently, ten genera are included in Cainiaceae (Hyde et al. 2020a, b, c).
Amphibambusa D.Q. Dai & K.D. Hyde, in Liu et al., Fungal Diversity: (2015)
Notes: Amphibambusa was introduced by Liu et al. (2015a, b) with A. bambusicola as the type species. Amphibambusa is characterized by immersed ascomata surrounded by a small, blackened clypeus and ostiolar opening surrounded by a white margin, and cylindrical asci with fusiform ascospores surrounded by a wide gelatinous sheath. In this study, we introduce a new species, Amphibambusa aquatica, based on morphological and phylogenetic evidence (Fig. 110).
Amphibambusa aquatica Doilom sp. nov.
Index Fungorum number: IF901394; Facesoffungi number: FoF 16063; Fig. 111
[See PDF for image]
Fig. 111
Amphibambusa aquatica (MFLU 18-1172, holotype). a, b Appearance of subglobose group of spores on host surface. c Immersed ascomata with opening ostiolar on host surface. d Cross section of ascoma showing perithecium. e Structure of peridium. f–h Unitunicate asci. i Paraphyses. j–n Ascospores. o Germinated ascospore. p Colony on PDA (from front). q Colony on PDA (from below). Scale bars: e–i = 20 μm, j–n = 10 μm, o = 30 μm
Etymology: In reference to the freshwater habitat
Holotype: MFLU 18-1172
Saprobic on decaying wood submerged in freshwater. Sexual morph: Ascomata 400–900 μm high, 500–1000 μm diam., black, scattered, immersed, subglobose, unilocular, thin-walled, ostiolate, ostiolar opening with a mass of spores accumulated on the surface of substratum, forming a black, glistening, subglobose group of spores. Peridium 15–45 μm thick, comprising several layers of dark brown to black, thin-walled cells of textura angularis. Paraphyses even in width, 2–4 μm diam., numerous, cylindrical, branched, hyaline, aseptate. Asci 190–240 × 12–16 μm (x̅ = 215 × 15 μm, n = 10), 8-spored, unitunicate, cylindrical, thin-walled, pedicellate, apically rounded with a cylindrical, refractive, subapical ring, 1–2 μm high, 2–3 μm diam. Ascospores 32–40 × 6–7.5 μm (x̅ = 35 × 6.5 μm, n = 20), mostly uni-seriate or overlapping uni-seriate, occasionally biseriate in the middle part, straight or slightly curved, hyaline, 1–2(-3)-septate, fusiform, pointed at both ends, thin-walled, smooth, guttulate, with a longitudinally striated wall and surrounded by 2–4(-7) μm thick, gelatinous sheath. Asexual morph: Not observed.
Culture characteristics: On PDA, colony circular, reaching 50 mm in 25 days at 25 °C, white from above, pale brown from below, surface rough, dry, raised, edge entire.
Material examined: China, Dehong, Yunnan Province, on submerged wood in a small river, 25 November 2017, G.N. Wang, H12A-1 (MFLU 18-1172, holotype); ex-type MFLUCC 18-1046; ibid., H12A-2 (HKAS 101747, isotype); ex-isotype KUMCC 18-0099.
GenBank numbers: ITS: PP584660; LSU: PP584659.
Notes: Amphibambusa aquatica shares morphology with the type species of Amphibambusa (A. bambusicola) in having immersed ascomata with an opening ostiolar, cylindrical asci, subapical ring, fusiform ascospores with a longitudinally striated wall, and being surrounded by a wide gelatinous sheath. However, A. aquatica has immersed ascomata with the opening ostiolar covered by a black, glistening, subglobose group of spores, while A. bambusicola has immersed ascomata under a small blackened clypeus and is surrounded by a white margin. Paraphyses of A. aquatica are even in width with aseptate, while they are uneven in width with septate of A. bambusicola. Amphibambusa aquatica has longer and narrower asci than those of A. bambusicola (190–240 × 12–16 μm vs. 150–200 × 17.5–20 μm). The ascospores of A. aquatica are mostly uni-seriate, 1–2(-3)-septate being surrounded by 2–4(-7) μm thick sheath, while they are biseriate, 1-septate asco surrounded by 10 μm thick sheath A. bambusicola (Liu et al. 2015a, b). In addition, A. aquatica has different culture characteristics on PDA from A. bambusicola (entire at the margin vs. irregular at the margin). The phylogenetic analysis of combined LSU and ITS showed that A. aquatica clustered in a sister branch with A. bambusicola resided in the family Cainiaceae (Fig. 110). Based on the differences in morphological characteristics and phylogenetic analysis, we introduce A. aquatica, a freshwater taxon, as a new species in Amphibambusa.
[See PDF for image]
Fig. 112
The phylogenetic tree obtained inferred from a combined LSU, ITS, tub, and rpb2 sequence dataset representing the species of Xylariaceae strains were included in the combined analyses. Bootstrap support values for ML equal to or greater than 70%. The best RAxML tree with a final likelihood value of − 140614.819024 is presented. The matrix had 2627 distinct alignment patterns, with 40.86% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.237758, C = 0.272136, G = 0.249814, T = 0.240293; substitution rates AC = 1.293484, AG = 4.807374, AT = 1.149005, CG = 1.072288, CT = 7.777795, GT = 1.000000; gamma distribution shape parameter α = 0.776984. The tree is rooted to Hypoxylon fragiforme (HAST 383). The ex-type strains are in bold, and the new isolate of this study is in blue. Bar = 0.06 estimated number of nucleotide substitutions per site per branch
[See PDF for image]
Fig. 113
Helicogermslita kunmingensis (HKAS 125556, holotype). a, b. Ascomata on dead woody twigs. c, d. Vertical section of the ascoma. e. Ostiole. f. Peridium. g. Hermathecium. h–k. Asci. l, m. Tubular ascal apical apparati in Melzer’s reagent. n–q. Ascospores. r. Germinating ascospore. s. Front colony on PDA. t. Reverse colony on PDA. Scale bars: d = 200 μm. e = 100 μm. f = 30 μm. g = 40 μm. h–k = 30 μm. l–q = 10 μm. r = 30 μm
Xylariaceae Tul. & C. Tul. [as ‘Xylariei’], Select. fung. carpol. (Paris) 2: 3 (1863)
Xylariaceae is a well-known family of Ascomycota which comprises wood inhabitants (Samarakoon et al. 2023; Li et al. 2024). They were traditionally classify based on stromal form, stromal colour, and ascospore shape and dimensions (Samarakoon et al. 2023). Thirty-six genera are accepted (Wijayawardene et al. 2022).
Helicogermslita Lodha & D. Hawksw., Trans. Br. mycol. Soc. 81(1): 91 (1983)
Helicogermslita was introduced to accommodate a xylarialean species by Hawksworth and Lodha (1983) with ascospores having spiral germ slits and typified by H. celastri. Apart from this, Helicogermslita species share stromatic ascomata with white ectostroma and asci with J + , apical rings (Petrini 2003; Daranagama et al. 2018). Nine species have been described based only on morphology (Fig. 112).
Helicogermslita kunmingensis Y. Gao & H. Gui, sp. nov.
Index Fungorum number: IF 902400; Facesoffungi number: FoF 15063; Fig. 113
[See PDF for image]
Fig. 114
Phylogram generated from a RaxML analysis based on LSU sequence data of Zygosporium isolates. Related sequences were obtained from NCBI´s GenBank. Fifteen taxa are included in the analysis, which represents 822 characters including gaps. The tree is rooted to Xylaria arbuscula CBS 126416 and Iodosphaeria tongrenensis MFLU 15-0393. Here is represented the best RaxML tree using the GTR model with a final likelihood of − 2334.079572. The matrix had 126 alignment patterns, with 1.82% of undetermined characters or gaps. Estimate base frequencies were as follows: A = 0.250237, C = 0.213354, G = 0.301698, T = 0.234711; substitution rates: AC = 1.189322, AG = 2.931397, AT = 0.638000, CG = 0.896786, CT = 10.246864, GT = 1.000000; gamma distribution shape parameter α = 0.107590. RaxML bootstrap support values ≥ 70% (MLBS) and Bayesian posterior probabilities ≥ 0.95 (BYPP) are given at the nodes. The scale bar indicates 0.02 changes. The novel species proposed in this study is in blue and ex-types strains of different species are in bold
Etymology: The specific epithet “kunmingensis” refers to Kunming City where the holotype was collected.
Holotype: HKAS 125556.
Saprobic on dead woody twigs. Sexual morph: Ascomata 570–950 μm diam × 430–830 μm high (x̅ = 822 × 629 μm, n = 15), scattered, immersed, globose to subglobose or irregular, ostiolate. Peridium 72–100 μm wide, (x̅ = 86 μm, n = 25), comprising an outer, brown, thick-walled, stratum of cells of textura angularis and inner, hyaline, thick-walled, stratum of textura irregularis cells. Hamathecium 3.5 –6 μm wide, (x̅ = 4.6 μm, n = 20), septate, hyaline, unbranched, broad at base, tapering upwards. Asci 146–185 × 17–23 μm (x̅ = 160 × 19 μm, n = 30), 8-spored, cylindrical, cylindric-clavate, unitunicate, short pedicellate with foot-like pedicel, thick-walled at the apex, with apical apparatus tubular, bluing in Melzer’s reagent. Ascospores 20–29 × 11–13 μm (x̅ = 24 × 12 μm, n = 30), initially hyaline, becoming dark brown at maturity, guttulate, fusiform or ellipsoid-inequilateral, with narrowly rounded ends, aseptate, spiral germ slit extending over the full length. Asexual morph: Not observed.
Culture characteristics: Ascospores germinated on PDA within 20 h and germ tube initially produced from both ends of the ascospores. Colonies on PDA reaching 20 mm in 4 weeks at room temperature (25–27 °C), failamentous, mycelium grows on the surface of PDA, white, Asexual spores and sexual spores were not formed on PDA within 60 days.
Material examined: China, Yunnan Province, Kunming City (25° 8′ 19″ N, 102° 44′ 25″ E), On dead woody twigs, 20 June 2021, Ying Gao, CCSG11 (HKAS 125556, holotype), ex-type, KUNCC 22-12535.
GenBank numbers: ITS: PQ036918, ITS: OP912879, rpb2: OP918020
Note: Based on our phylogenetic analysis of the combined LSU, ITS, rpb2 and tub sequence data, our novel species, Helicogermslita kunmingensis, is closely related to the strain of Helicogermslita clypeata with BYPP bootstrap supports (Fig. 112). Helicogermslita clypeata was introduced by Samarakoon et al. (2022), which have detailed morphological analysis and can be compared with our species Helicogermslita kunmingensis in peridium (29–45 μm vs. 72–100 μm), asci (115–145 × 9–11.5 μm vs. 146–185 × 17–23 μm), ascospores (13–18.5 × 5.7–7.5 μm, broadly ellipsoidal vs. 20–29 × 11–13 μm, fusiform or ellipsoid-inequilateral). and, a nucleotide pairwise comparison showed that Helicogermslita kunmingensis differs from Helicogermslita clypeata (MFLU 18–0852) in 75/545 bp of ITS (13.76%, with 19 gaps), 179/1042 bp of rpb2 (17.18%, with 6 gaps), 24/858 bp of LSU (2.8%, without gaps), Therefore, Helicogermslita kunmingensis is introduced here as a novel taxon based on the phylogeny with the guidelines for species delineation described in Chethana et al. (2021a, b) and Maharachchikumbura et al. (2021).
Zygosporiaceae J.F. Li, Phookamsak & K.D. Hyde, in Li, Phookamsak, Jeewon, Tibpromma, et al., Mycosphere 8(10): 1860 (2017)
Zygosporiaceae is a monotypic family in Xylariales (Wijayawardene et al. 2022). This family was introduced to accommodate the genus Zygosporium based on morphology and phylogeny (Li et al. 2017). Zygosporium species are cosmopolitan, with a distribution that ranges from temperate to tropical areas. They are frequently recovered as saprobes occurring on plant litter from gymnosperms and angiosperms, but especially on monocotyledon plants from the families Arecaceae and Pandanaceae (Whitton et al. 2002a, b; Li et al. 2017). In fact, this association with plant material has been detected in fossil records (Bera et al. 2023).
Zygosporium Mont., Annls Sci. Nat., Bot., sér. 2 17: 120 (1842)
Zygosporium is a diverse genus that comprises anamorphic fungi characterized by dark pigmented, curved (hooked) vesicular cells. Two major groups can be differentiated: i) species with a setiform conidiophore, where the vesicular cells arise laterally from a seta-like structure, and ii) species with a vesicular conidiophore, where vesicular cells are arranged in chains or usually single, arising directly from the mycelium. The vesicular cells produce between 2–4 ampulliform conidiogenous cells. The conidia range from ellipsoid to globose, smooth, or ornamented with different patterns. The type species is Zygosporium oscheoides, which produces setiform conidiophores (Montagne 1842; Hughes 1951a, b; Whitton et al. 2002a, b). Herein we introduce a new Zygosporium based on morphology and phylogeny (Fig. 114)
Zygosporium cymodoceae Guerra, Baulin, Cano & Gené, sp. nov.
Index Fungorum number: IF 902401; Facesoffungi number: FoF 14004; Fig. 115
[See PDF for image]
Fig. 115
Zygosporium cymodoceae (FMR 19851). a, b Conidiophores. c, d Vesicular cells from superior view. e, f Vesicular cells from lateral view, bearing verruculose conidiogenous cells. g Conidia. h Colony on OA after 14 days at 25ºC. Scale bars: 10 µm
Etymology. The name refers to the substrate where this fungus was isolated, Cymodocea nodosa.
Holotype: CBS H-25261
Saprobic in Cymodocea nodosa.Sexual morph not observed. Asexual morph on OA: Mycelium 1.5–2 µm wide, hyaline, smooth-walled to verrucose, branched, septate. Conidiophores (35–)40–50(–55) × 1.5–2.5 µm (x̅ = 46.9 × 1.9 µm, n = 30), arising directly from the superficial mycelium, setiform, erect, 3(–4) septate, unilaterally branched near the base, brown and smooth-walled at the base, subhyaline to hyaline and often roughened towards the apex, lateral branch consisting in a short stalk, 0(–1)-septate, brown to pale brown, 3–5 × 2–3 µm (x̅ = 4 × 2.5 µm, n = 30), with a terminal vesicular cell bearing two conidiogenous cells at the apex. Vesicular cells 6–7.5 × 4–5.5 µm (x̅ = 6.9 × 5 µm, n = 30), in the broadest part, solitary, dark brown, smooth-walled, clavate, curved (hooked). Conidiogenous cells 4–4.5(–6) × 2.5–3 µm (x̅ = 4.5 × 2.6 µm, n = 30), holoblastic, hyaline, verruculose, ampulliform, curved. Conidia 4–6 µm diam (x̅ = 4.6 µm, n = 30), solitary, pale brown, verrucose, globose.
Culture characteristics: after 2 weeks at 25 °C, colonies on PDA attaining 8 mm diam., raised, umbonate, grey (4F1, Kornerup & Wanscher 1978) at the centre, white (4A1) towards the periphery, margin entire; reverse yellowish white (4A2). On MEA, colonies reaching 8 mm diam., slightly raised, crateriform, radially sulcate, velvety, grey (4F1) at the centre, light orange (5A4) towards the periphery, margin entire and lobated; reverse grey (4F1) to light orange (5A4). On OA, colonies reaching 16–18 mm diam., flat, velvety, hyaline with black spots of sporulation at the centre; reverse colourless. Cardinal temperatures for growth: minimum at 10 °C, optimum at 25 °C, maximum at 30 °C.
Material examined: Spain, Catalonia, Tarragona, Platja del Miracle, N 41° 6′ 37″, E 1° 15′ 37″, on Cymodocea nodosa Asch. (Cymodoceae), at 6 m of depth, 14 October 2021, G. Quiroga-Jofre and D. Guerra-Mateo (CBS H-25261, holotype), ex-type CBS 149942, FMR 19851, MycoBank MB847540.
GenBank numbers: ITS = OQ589930, LSU = OQ589929.
Notes: Zygosporium cymodoceae is a novel species phylogenetically related to Z. oscheoides, the type species of the genus. Since DNA data of the species accepted in Zygosporium is very limited, morphological comparison is essential to distinguish species. Although the conidiogenous apparatus of Z. cymodoceae is similar to that of its closely related species, Z. minus and Z. oscheoides, they can be distinguished mainly by their conidial morphology. The conidia in the new species are globose, conspicuously verrucose and measure 4–6 µm diam. (Fig. 115g), while in Z. minus, although globose, they are larger (5.5–10 µm diam.) and verruculose, and in Z. oscheoides are ellipsoid, smooth to verruculose and measure 7–12.5 × 4–9 µm diam. (Hughes 1951a; Whitton et al. 2002a). Other Zygosporium species that can be easily distinguished from Z. cymodoceae by the conidial features are Z. bioblitzi, Z. deightonii, Z. geminatum, Z. pacificum and Z. verruciferum, all with ellipsoid conidia (Whitton et al. 2002a; Mckenzie et al. 2007; Lucena and Fernández-Valencia 2017); Z. anupamvarmae shows falcate conidia with ridges (Monoharachary et al. 2006); the conidia of Z. majus, Z. tuberculatum and Z. pandanicola are globose but considerably larger (13–18, 12.5–16.5 and 11–14.5 µm diam., respectively); and Z. chartarum and Z. verticilatum show globose and ovoid conidia, respectively, but hyaline and smooth (Whitton et al. 2002a). Other features that distinguish Z. cymodoceae from the other species in the genus is the presence of a setiform conidiophore (Fig. 115a–f), which is absent in Z. cocos, Z. dilleniae, Z. gibbum, Z. masonii, Z. mycophilum, Z. tonellianum, Z. pseudogibbum and Z. pseudomasonii (Whitton et al. 2002a; Dubey 2014; Crous et al. 2018, 2019a), and the production of conidiophores unilaterally branched at the base bearing a single vesicular cell, in contrast to Z. verticillatum which produces 2–3 vesicular cells per conidiophore (Whitton et al. 2002a).
[See PDF for image]
Fig. 116
RAxML tree generated from combined SSU, ITS, LSU, rpb2 and tef1-α sequence data of Xylariales. Bootstrap values for ML equal to or greater than 75% are placed above the branches. Branches with Bayesian posterior probabilities (PP) from MCMC analysis equal to or greater than 0.95 were in bold. The tree was rooted with Amphisphaeria sorbi (MFLUCC 13-0721) and A. thailandica (MFLU 18-0794). The ex-type strains were indicated in bold and newly generated sequences were indicated in red
Phylogenetically, Z. cymodoceae represents an independent lineage with strong statistical support using the LSU region (99% MLBS, 0.99 BYPP) (Fig. 114). The ITS region was excluded from the phylogenetic tree due to the great phylogenetic distance between FMR 19851 and related taxa. The alignment of the ITS sequences revealed a 94% of similarity with Z. oscheoides (CBS 195.79, GenBank MH861194), and a 93% with Z. minus (HKAS99625, MF621586). BLAST searches on the NCBIs GenBank nucleotide database using the ITS region did not provide significant results. According to our results and previous works (Crous et al. 2020a), some Zygosporium species should be accommodated in other genera. Since Z. cymodoceae clusters with Z. oscheoides, we consider it a species of Zygosporium s. str.
Sordariomycetes genera incertae sedis
Stanjehughesia Subram., Proc. Indian natn Sci. Acad., Part B. Biol. Sci. 58(4): 184 (1992)
Stanjehughesia is polyphyletic, and most of the species are distributed in Chaetosphaeriales, and a few species are distributed in Xylariales (Hsieh et al. 2021; Wu and Diao 2022). Type species, S. hormiscioides lacks molecular data and the phylogeny of the genus is uncertain (Fig. 116).
Stanjehughesia bambusicola X.D. Yu & Jian K. Liu, sp. nov.
Index Fungorum number: IF 902402; Fig. 117
[See PDF for image]
Fig. 117
Stanjehughesia bambusicola (HKAS 127156, holotype) a–b Colonies on the natural substratum. d–e Conidiophores and conidia. f–j Conidia. k Germinating conidium. l, m Colony on PDA from above and below. Scale bars: d–k = 10 μm
Etymology: The epithet “bambusicola” refers to the bamboo host, on which the fungus was collected.
Holotype: HKAS 127156
Saprobic on dead branches of bamboo. Sexual morph: Not observed. Asexual morph: Hyphomycetous. Colonies on wood effuse, hairy, black, scattered or in small groups, glistening. Mycelium mostly immersed, composed of septate, smooth, brown to hyaline hyphae. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 7.5–10.0 × 3.0–3.5 µm (x̅ = 8.2 × 3.2 µm, n = 10), monoblastic, holoblastic, integrated, terminal, determinate, erect, solitary or caespitose, straight, cylindrical or lageniform, aseptate, smooth-walled, unbranched, dark brown, truncated at the apex. Conidia 35–50 × 8.5–10.0 µm (x̅ = 42.0 × 9.5 µm, n = 30), acrogenous, solitary, 6–8-euseptate, verrucose, fusiform or ellipsoidal, brown to dark brown, pale brown to subhyaline at the apex of apical cells, guttulate, rounded at the apex and truncated at the base.
Culture characteristics: Conidium germinated on PDA within 24 h. Colonies growing well on PDA and attaining a diameter about 15 mm after 3 months at 25 °C, circular, medium dense, flat to raised, entire margins, greyish-white, brown to dark brown from the reverse.
Material examined: China, Sichuan Province, Chengdu City, Qionglai County, Lugou Bamboo Sea Area, 30° 22′ 37″ N, 103° 16′ 45″ E, 730 m Elevation, on dead branches of bamboo in terrestrial habitat, 12 October 2021, X.D. Yu, A33 (HKAS 127156, holotype); ex-type CGMCC 3.24359.
GenBank numbers: ITS = OR822007, LSU = OR822023, tef1-α = OR855673, rpb2 = OR855672.
Notes: Morphologically, S. bambusicola shares similar characteristics with S. minima in having fusiform or ellipsoidal, rounded at the apex, 6–8-euseptate conidia, but it differs in conidiogenous cell size (7.5–10.0 × 3.0–3.5 µm vs. 8–11 × 6–8) and conidiogenous cells habit (solitary or grouped vs. solitary) (Wu 2005). Phylogenetically, our new collection forms a separate lineage basal to S. aquatica and S. polypora with 87% ML/1.00 BYPP support within Xylariales (Fig. 117). Stanjehughesia bambusicola differs from S. aquatica and S. polypora in having no rostrate (Wu 2005; Yang et al. 2023). Therefore, Stanjehughesia bambusicola is introduced as a new species based on morphological characters and phylogenetic evidence.
[See PDF for image]
Fig. 118
The best scoring RAxML tree with a final likelihood value of − 9504.567668 of the genus Leucoagaricus and Leucocoprinus based on the ITS region. Related sequences were retrieved from GenBank. Forty-nine specimens were included in the analysis of the ITS region which comprises 798 characters after alignment. The tree is rooted with Cystolepiota seminuda (GenBank OM212892 and AY176359). The matrix had 534 distinct alignment patterns, with 19.55% of undetermined characters or gaps. Estimated base frequencies were A = 0.232236, C = 0.219433, G = 0.233142, T = 1.000000; substitution rates AC = 1.252315, AG = 3.790459, AT = 1.412007, CG = 0.564967, CT = 3.937227, GT = 1.000000. Maximum likelihood bootstrap values equal to or greater than 70% are near each branch. The holotype specimens retrieved from the GenBank are in bold and black. The new species are in bold and blue. The subgenus of the species is indicated after the collection/GenBank no. of each specimen
[See PDF for image]
Fig. 119
Leucoagaricus madagascarensis (K-M000251887, holotype). a Basidioma. b Basidiospores in Melzer’s. c Basidiospores in Congo red. d Basidiospores in KOH. e Basidia in Congo red. f Cystida in Melzer’s. g Pileipellis in Melzer’s. Photographs P.A. Hagl, Anna Ralaiveloarisoa. Scale bars: a = 1 cm; b–g = 10 µm
Basidiomycota R.T. Moore, Bot. Mar. 23(6): 371 (1980)
Subphylum Agaricomycotina Doweld, Prosyllabus Tracheophytorum, Tentamen Systematis Plantarum Vascularium (Tracheophyta) (Moscow): LXXVIII (2001)
Agaricales Underw., Moulds, mildews and mushrooms. A guide to the systematic study of the Fungi and Mycetozoa and their literature (New York): 97 (1899)
Agaricaceae Chevall., Fl. gén. env. Paris (Paris) 1: 121 (1826)
Notes: Agaricaceae is a well-known family of Agaricales with global distribution. Agaricaceae s.l. includes five families – Agaricaceae s.str., Coprinaceae, Lepiotaceae (nom. inval.), Lycoperdaceae and Tulostomataceae – as well as five genera with unknown phylogenetic position and no existing sequence data (Endolepiotula, Hiatulopsis, Janauaria, Phyllogaster, and Rugosospora) (Kalichman et al. 2020). Genera in this group are agaricoid, gasteroid or secotioid, and the agaricoid species generally have pluteoid, lepiotoid or tricholomatoid, rarely collybioid or mycenoid basidiomata. Typically, they also consist of free lamellae, a veil commonly forming an annulus and/or characteristic squamules on the pileal and stipe surfaces, hyaline to pigmented basidiospores seen as white, very pale pink or dark/black spore prints, and a regular to subregular and non-gelatinized trama (Pegler 1986; Vellinga 2004). The family is of great scientific interest comprising important decomposers but also symbionts as well as edible, poisonous and medicinal species (Singer 1986; Alexopoulos et al. 1996; Didukh et al. 2003)(Fig. 118).
Leucoagaricus Locq. ex Singer, Sydowia 2(1–6): 35 (1948)
Notes: The traditional circumscription of the genus includes small- to medium-sized species with thin to fleshy basidiomata, a fibrillose or ± squamulose to scaly pileus, free lamellae, simple annulus, dextrinoid and metachromatic basidiospores, and cheilocystidia and hyphae without clamp connections (Singer 1986; Vellinga 2001). The species with plicate pilei, spores with or without apical germ pore, and a hymenium with pseudoparaphyses have been placed in the genus Leucocoprinus (Singer 1986). The phylogenetic studies have, however, shown that Leucoagaricus (type species L. rubrotinctus (Peck) Singer) is polyphyletic and intermixes with Leucocoprinus Pat. (type species L. cepistipes (Sowerby) Pat.) and the relationships of both genera are still unresolved (Vellinga 2003, 2004; Vellinga et al. 2011; Justo et al. 2021). The morphologically delimited Leucoagaricus is widely distributed and contains approximately 150 species (Hussain et al. 2018). The species are saprotrophic occurring on soil, wood chips, and sawdust in a wide range of habitats.
Leucoagaricus madagascarensis Ralaiv., Liimat., Niskanen sp. nov.
Index Fungorum number: IF 902403; Facesoffungi number: FoF 16064; Fig. 119a–g
[See PDF for image]
Fig. 120
Leucocoprinus mantadiaensis (K-M000251856, holotype). a Basidioma. b Basidiospores in Melzer’s. c Basidiospores in Congo red. d Basidiospores in KOH. e Basidia in Congo red. f Cystidia in Congo red. g Pileipellis in Melzer’s. Photographs Robert Patrick Brown, Anna Ralaiveloarisoa. Scale bar: a = 1 cm; b–g = 10 µm
Etymology: The species epithet is derived from the name of the country (Madagascar) from which the type specimen was collected.
Holotype: K-M 000251887
Pileus 5 cm in diam., at first low convex, then almost plane with a low and broad umbo; brownish pinkish orange, fibrillose-striate, black at the centre, background cream white. Lamellae free, crowded, white. Stipe 6.5 cm, white, clavate, annulus membraneous. Odour strong mushroom smell. Basidiospores (6–)6.5–9.5(–10) × 3.5–5(–6.5) µm, amygdaloid to ellipsoid, without germ pore, smooth, hyaline in KOH, moderately dextrinoid in Melzer’s. Basidia (in Congo red) 12–15 × 6.5–9 µm, sterigmata 1.5–2.5 µm long, broadly clavate, usually 4-spored, rarely 2-spored, not clamped. Cheilocystidia (11.5–)22–50.5 × (7.5–)10–20 µm, cylindrical to clavate. Pileipellis cutis, hyphae up to 3.5–6 µm wide, not clamped.
Habitat: Humid, tropical, evergreen primary forest with no disturbance, on organic soil.
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Toamasina, Andasibe, Mantadia Nature Reserve, tropical humid forest, 25 Feb. 2018, P.A. Hagl 000251887 (K-M, holotype), MAF18-249 (TAN, isotype).
GenBank numbers: OR785982
Notes: Leucoagaricus madagascarensis groups with 100% bootstrap support with the type species of Leucoagaricus. The species is characterised by brownish pinkish orange, at the centre blackish pileus, membraneous annulus and amygdaloid to ellipsoid 6.5–9.5 × 3.5–5 µm basidiospores. ITS sequence is distinct from other species of Leucoagaricus s. str. and deviating from them by > 3% (> 20 substitutions and indels). It is the first Leucoagaricus species described from Madagascar and is currently not known to occur elsewhere.
Leucocoprinus Pat., J. Bot., Paris 2: 16 (1888)
Notes: Leucocoprinus belongs to Agaricaceae s.str. (Kalichman et al. 2020). The morphologically delimited genus does not form a monophyletic group and intermixes with Leucoagaricus in the phylogenetic analysis (Vellinga et al. 2011; Justo et al. 2021; see also above the notes of the genus Leucoagaricus) and a consensus has yet to be reached regarding the taxonomy of these genera. Of the generic names, Leucocoprinus is older and has priority in case the two genera are merged.
Leucocoprinus mantadiaensis Ralaiv., Liimat., Niskanen sp. nov.
Index Fungorum number: IF 902404; Facesoffungi number: FoF 16065; Fig. 120a–g
Etymology: The species epithet is derived from the name of the protected area “Mantadia” where the type specimen was collected.
Holotype: K-M 000251856
Pileus 1.5 cm in diam., convex to plano-convex, with an umbo in the centre, white, brown at the centre and very finely brown squamulose on the other parts. Lamellae free, crowded, white. Stipe 4 cm, white, cylindrical, with an annulus. Odour indistinct. Basidiospores 7–9(–10) × (3.5–)4–5 µm, amygdaloid to ellipsoid, without a germ pore, smooth, hyaline in KOH, strongly dextrinoid in Melzer’s. Basidia 17.5–22.5 × 8–9 µm, sterigmata 2.5–4 µm long, clavate, 4-spored, rarely 2-spored, not clamped. Cheilocystidia 14–42 × 9–15 µm, clavate. Pileipellis cutis, hyphae 3–6 µm wide, not clamped.
Habitat: Humid, tropical, evergreen primary forest with no disturbance, on organic soil. Dominated by Cunnoniaceae, Eugenia, Euphorbiaceae, Rubiaceae, Clusiaceae, Phyllanthaceae.
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Toamasina, Andasibe, Mantadia Nature Reserve, tropical humid forest, 24 Feb. 2018, Robert Patrick Brown 000251856 (K-M, holotype), MAF18-231.
GenBank numbers: OR785983
Notes: Leucocoprinus mantadiaensis is placed, but without support, in a big group containing the type species of Leucocoprinus in our phylogenetic analysis. In addition, it does not belong to the well-supported Leucoagaricus s. stricto group with the type species of the genus Leucoagaricus. Therefore, we currently place the species in Leucocoprinus. Leucocoprinus mantadiaensis is a small, white species with a non-plicate pileus and a brown pileus centre. The spores are amygdaloid to ellipsoid 7–9(–10) × (3.5–)4–5 µm and the cheilocystidia are clavate 14–42 × 9–15 µm. ITS sequence (GenBank OR785983, holotype) is distinct from other members of Leucoagaricus/Leucocoprinus group and deviating from the other species (Leucoagaricus infuscatus GenBank no. EU141943, Lepiota flammeotincta GenBank no. OM474006, Leucocoprinus sp GenBank no. OL653095, and Leucoagaricus viriditinctus GenBank EU419375, AF482873) placed in the same well-supported (BS 99%) clade by more than 5%. The species is currently only known from Madagascar.
Hygrophoraceae Lotsy, Vortr. bot. Stammesgesch. 1: 705 (1907)
Notes: Hygrophoraceae is one of the largest families in Agaricales. The family comprises 29 accepted genera and over 900 species (Kalichman et al. 2020). They are mostly biotrophic and represent a wide ecological spectrum including ectomycorrhizal, bryophilous, pteridicolous and lichen-forming species. However, the details of the nutritional strategies of many genera remain unresolved (Lodge et al. 2014). Historically, the family was morphologically characterised by basidiomata with thick, widely-spaced, waxy lamellae; basidiospores that are usually smooth, hyaline and inamyloid; and basidial lengths exceeding those of their spores by at least a factor of five. However, the current view is that such characteristics are losing some of their former taxonomic value (Lodge et al. 2013). Family members show a worldwide distribution, from the tropics to the subpolar and alpine regions. Although regarded as woodland fungi in several countries, some are associated with natural or semi-natural grasslands, particularly in Europe. These show extreme sensitivity to elevated environmental nitrogen levels, e.g., from the application of agricultural fertilizers and air pollution. This makes them good indicator species for grasslands of high conservation value (Boertmann 1995; Lodge et al. 2013).
Hygrocybe (Fr.) P. Kumm., Führ. Pilzk. (Zerbst): 26 (1871)
Notes: Hygrocybe s.s. is a strongly supported monophyletic genus based on a four-gene analysis by Lodge et al. (2013). They upheld the recognition of two subgenera, Hygrocybe and Pseudohygrocybe, but these had differing and weaker levels of molecular support. If H. helobia was excluded from their four-gene and supermatrix analyses, subgenus Hygrocybe was resolved as a well-supported monophyletic clade, however, subgenus Pseudohygrocybe could only be resolved as a paraphyletic grade (Lodge et al. 2013). The species of Hygrocybe are characterized by bright pigments in their basidiomata, thick lamellae and smooth basidiospores (Lodge et al. 2013). The genus contains over 300 species (Kalichman et al. 2020) and has a worldwide distribution. The species are mostly terrestrial (Lodge et al. 2013) and can be found in forests, grasslands, marshes, sand dunes, fens and bogs (Boertmann 2010). They are considered to be endophytic and may potentially form other biotrophic associations with host plants (Halbwachs et al. 2018) (Fig. 121).
[See PDF for image]
Fig. 121
The best-scoring RAxML tree with a final likelihood value of − 12,831.577315 of Hygrocybe is based on the ITS region. Related sequences were retrieved from GenBank. Forty-nine specimens were included in the analysis of the ITS region which comprises 919 characters after alignment. The tree is rooted with Hygroaster nodulisporus (KM181781) and Hygroaster madagascarensis (KM255431). The matrix had 693 distinct alignment patterns, with 36.20% of undetermined characters or gaps. Estimated base frequencies were A = 0.236482, C = 0.229418, G = 0.227898, T = 0.306202; substitution rates AC = 1.008294, AG = 2.079633, AT = 0.845056, CG = 0.700696, CT = 3.205370, GT = 1.000000. Maximum likelihood bootstrap values equal to or greater than 70% are given near each branch. The type specimens are in bold. The new species are in blue. Names, other than those originating from the type specimens, are only indicative and not confirmed via comparison with sequences from type specimens. The subgenus of the species is indicated after the collection/GenBank no. of each specimen
Hygrocybe minimiholatra Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov.
Index Fungorum number: IF 902405; Facesoffungi number: FoF 16066; Figs. 122a, 123a, 124a, 125a
[See PDF for image]
Fig. 122
Basidiomata of the Hygrocybe species: aHygrocybe minimiholatra 000251949 (holotype, K-M), bHygrocybe parviholatra 000251897 (holotype, K-M), cHygrocybe mitsinjoensis 000251832 (holotype, K-M), dHygrocybe solis 000251894 (holotype, K-M), eHygrocybe solis 000252075 (K-M), f, gHygrocybe vintsy 000252026 (K-M). Photographs a Javier Galan, b Paul Eguia, c Tuula Niskanen, d Paul Eguia, e Tuula Niskanen, f, g Meilinda Sulastri. Scale bar 1 cm
[See PDF for image]
Fig. 123
Basidiospores Hygrocybe species in Melzer’s. a Hygrocybe minimiholatra (K-M000251949, holotype). b Hygrocybe mitsinjoensis (K-M000251832, holotype). c Hygrocybe parviholatra (K-M000251897, holotype). d Hygrocybe solis (K-M000251894, holotype). e Hygrocybe vintsy (K-M000252026, holotype). Photographs Anna Ralaiveloarisoa. Scale bar = 10 µm
[See PDF for image]
Fig. 124
Basidia of the species of the genus Hygrocybe in Melzer’s. a Hygrocybe minimiholatra (K-M000251949, holotype). b Hygrocybe mitsinjoensis (K-M000251832, holotype). c Hygrocybe parviholatra (K-M000251897, holotype). d Hygrocybe solis (K-M000251894, holotype). e Hygrocybe vintsy (K-M000252026, holotype). Photographs Anna Ralaiveloarisoa. Scale bar = 10 µm
[See PDF for image]
Fig. 125
Pileipellis of the species of the genus Hygrocybe in Melzer’s. aHygrocybe minimiholatra (K-M000251949, holotype). bHygrocybe mitsinjoensis (K-M000251832, holotype). cHygrocybe parviholatra (K-M000251897, holotype). dHygrocybe solis (K-M000252026, holotype). eHygrocybe vintsy (K-M000251894, holotype). Photographs Anna Ralaiveloarisoa. Scale bar = 10 µm
Etymology: The name is derived from the Malagasy word holatra meaning a fungus and the Latin word minimus denoting the smallest size and refers to the very small size of the basidiomata.
Holotype: K-M000251949
Pileus 0.3–0.6 cm in diam., convex, with a small depression in the centre, red. Lamellae subdecurrent, subdistant, thick, pale red. Stipe 1–2 cm long, 0.5–1.5 mm thick at the apex, mostly red and then pale red at the base, cylindrical. Odour indistinct. Basidiospores 7–8 × 4–5.5 µm, ± ellipsoid, smooth, hyaline in Melzer’s. Basidia 17–38 × 5–8.5(–10) µm, sterigmata 2–5 µm long, clavate, (2-)4-spored, clamped. Ratio of basidial to basidiospore length < 5. Lamellar trama regular. Pileipellis cutis, clamped, hyphae 2–4 µm wide. ITS sequence (GenBank OR785988, holotype) distinct from other sequenced members of Hygrocybe and close to a species from Zimbabwe (UDB025759) with interspecific difference > 4% (20 indels and substitutions).
Habitat: Dry, deciduous, tropical forest, on sandy soil.
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Mahajanga, Ankarafantsika National Park, tropical dry forest, 12 Feb 2017, J. Galan 000251949 (K-M, holotype), MAF17-033.
GenBank numbers: OR785988
Notes: A minute red species with ellipsoid spores (7–8 × 4–5.5 µm). It belongs to Hygrocybe subgenus Hygrocybe and has the typical characteristics of that subgenus: the mean ratio of basidial to basidiospore length is < 5. In addition, the lamellar trama is regular and the basidia and spores are monomorphic. It differs from its closest species Hygrocybe from Zimbabwe with 20 indels and substitutions (> 4% interspecific difference).
Hygrocybe mitsinjoensis Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov.
Index Fungorum number: IF 902406; Facesoffungi number: FoF 16067; Figs. 122c, 123b, 124b, 125b
Etymology: The name is derived from the type locality “Mitsinjo Forest” of the species.
Holotype: K-M000251832
Pileus 0.5–1 cm in diam., convex, bright red, smooth. Lamellae adnate to adnexed or subdecurrent, subdistant, thick, white, becoming dark pink or red near the pileal attachment. Stipe 2.5–4.5 cm long, 1–2 mm thick at the apex, light red, becoming yellow towards the base, cylindrical. Odour not recorded. Basidiospores 7–8 × 4–5(–6) µm, somewhat ovoid to ellipsoid, smooth, hyaline in Melzer’s, monomorphic. Basidia (27.5)37–53 × 6–7(8.5) µm, sterigmata 3.5–7 µm long, clavate, (1-)4-spored, some clamped, monomorphic. Ratio of basidial to basidiospore length > 5. Lamellar trama subregular. Pileipellis a cutis, some clamps seen, hyphae 3–5 µm wide. ITS sequences (GenBank OR785984, holotype): The three specimens studied here as well as a published sequence from UNITE (UDB015243/SH1243293.09FU, Taolagnaro, Madagascar) form one clade with an intraspecific variation < 0.5% and are distinct from other species of Hygrocybe. The sister species from Zambia and Gabon (UDB013832, GenBank no JQ657787) differs by 2.3% (15 indels and substitutions).
Habitat: Humid, tropical, Humid, tropical, evergreen primary forest with no disturbance, dominated by Eugenia, Sarcolaena, Rubiaceae, Fabaceae, Phyllantaceae, on organic soil.
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Toamasina, Andasibe, Mitsinjo Forest, tropical humid forest, 23 Feb 2018, T. Niskanen 000251832 (K-M, holotype), MAF18-193. Madagascar. Toamasina. Andasibe, Mitsinjo Forest, tropical humid forest, 23 Feb 2018, T. Niskanen 000251833 (K-M), MAF18-180. Madagascar. Toamasina. Foulpointe, Analalava Forest Reserve, tropical humid forest, 19 Feb 2017, L. Chen 000252098 (K-M), MAF17-242.
GenBank numbers: OR785984
Notes: The species can be identified by the combination of small basidiomata with the red pileus, two-coloured stipe that is pale red at the top and yellow towards the base, and ovoid to ellipsoid basidiospores (7–8 × 4–5(–6) µm). It belongs to Hygrocybe subgenus Pseudohygrocybe and has the typical characteristics of that subgenus: the mean ratio of basidial to basidiospore length is > 5. In addition, the lamellar trama is subregular and the basidia and spores are monomorphic.
Hygrocybe parviholatra Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov.
Index Fungorum number: IF 902407; Facesoffungi number: FoF 16068; Figs. 122b, 123c, 124c, 125c
Etymology: The name is derived from the Malagasy word holatra meaning a fungus and the Latin word parvus means small and refers to the small size of the basidiomata.
Holotype: K-M000251897.
Pileus 0.4–0.8 cm in diam, convex, red. Lamellae subdecurrent to decurrent, subdistant, thick, white. Stipe 1.3–2.3 cm long, 0.5–2 mm thick at the apex, mostly red and then white to brownish white at the base, cylindrical to somewhat clavate. Odour not recorded. Basidiospores 6–8 × 3.5–4.5 µm, ± ellipsoid, smooth, hyaline in Melzer’s, monomorphic. Basidia 29–38.5 × 5.5–7 µm, sterigmata 3–4.5 µm long, subcylindrical to clavate, (1-)4-spored, clamped, monomorphic. Ratio of basidial to basidiospore length < 5. Lamellar trama subregular. Pileipellis cutis, clamped, hyphae 3.5–5 µm wide. ITS sequences (GenBank OR785989, holotype): distinct from other species of Hygrocybe. Forming a well-supported clade (100% BS) with species from USA and Gabon and deviating from them by > 4% (> 25 substitutions and indels). They are placed outside as the basal lineage of the genus Hygrocybe with (100% BS).
Habitat: Humid, tropical, evergreen primary forest with no disturbance, on organic soil and with Lauraceae, Monimiliaceae, Cunnoniaceae, Rubiaceae and Euphorbiaceae
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Toamasina, Andasibe, Mantadia Nature Reserve, tropical humid forest, 25 Feb 2018, P. Eguia 000251897 (K-M, holotype), MAF18-257. Madagascar. Toamasina, Andasibe, Mantadia Nature Reserve, tropical humid forest, 23 Feb 2018, P. Eguia 000251837 (K-M), MAF18-199.
GenBank numbers: OR785989
Notes: The species can be identified by its small and fleshy basidiomata, red and convex pileus, two coloured-stipe that are mostly red and then white to brownish white at the base, and ellipsoid to ovoid basidiospores (6–8 × 3.5–4.5 µm). The mean ratio of basidial to basidiospore length < 5. In addition, the lamellar trama is subregular and the basidia and spores are monomorphic.
Hygrocybe solis Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov.
Index Fungorum number: IF 902408; Facesoffungi number: FoF 16069; Figs. 122d–e, 123d, 124d, 125d
Etymology: Solis (Latin) means sun and refers to the colour of the basidiomata.
Holotype: K-M000251894
Pileus 1.5–2.5 cm in diam., convex and depressed in the centre, yellow to orange yellow. Lamellae adnexed to emarginate with a decurrent tooth, subdistant, thick, yellow. Stipe 4.5–8.5 cm, 4–5 mm thick at the apex, yellow to orange yellow, cylindrical. Odour not recorded. Basidiospores 9.5–12 × 6–8 µm, ellipsoid to ovoid, smooth, hyaline in Melzer’s, monomorphic. Basidia 34–61 × 6–15 µm, sterigmata 5–9.5 µm long, subclavate to clavate, (2-)4-spored, clamped, monomorphic. Ratio of basidial to basidiospore length ≤ 5. Lamellar trama subregular. Pileipellis cutis, some clamps seen, hyphae 3.5–6.5 µm wide. ITS sequences (GenBank OR785991, holotype): the two sequences are identical and form a well-supported clade (99% BS) with Hygrocybe macambrarensis from São Tomé and Príncipe and differing from it by 6.1% (39 indels and substitutions).
Habitat: Humid, tropical, evergreen primary forest with no disturbance, on organic soil, dominated by Lauraceae, Monimiliaceae, Cunnoniaceae, Rubiaceae and Euphorbiaceae.
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Toamasina, Andasibe, Mantadia Nature Reserve, tropical humid forest, 25 Feb 2018, P. Eguia 000251894 (K-M, holotype), MAF18-254. Madagascar. Toamasina, Andasibe, Mantadia Nature Reserve, tropical humid forest, 23 Feb 2018, T. Niskanen 000252075 (K-M), MAF18-179.
GeneBank numbers: OR785991
Notes: The yellow to orange-yellow basidiomata and large spores (9.5–12 × 6–8 µm) are typical for the species. Based on our phylogenetic analysis the species clusters together with species from Hygrocybe subgenus Pseudohygrocybe, but the mean ratio of basidial to basidiospore length is < 5 like in the species of H. subgenus Hygrocybe. The species differs from its closest Hygrocybe macambrarensis by 39 indels and substitutions (6.1% interspecific differences).
Hygrocybe vintsy Ralaiv., Liimat., A.M. Ainsw., Niskanen sp. nov.
Index Fungorum number: IF 902409; Facesoffungi number: FoF 16070; Figs. 122f–g, 123e, 124e, 125e
Etymology: Named after the Malagasy magazine and club Vintsy. They specialize in raising awareness of the protection and conservation of the environment and are designed especially for young people. Vintsy is the local name of the Madagascar kingfisher (Corythornis vintsioides).
Holotype: K-M000252026
Pileus 0.7–1.4 cm in diam., plano-convex and depressed in the centre, orange to reddish-orange, yellow towards the margin. Lamellae subdecurrent to decurrent, subdistant, thick, white, becoming orange and whitish with age. Stipe 2.7–3.5 cm long, 2–2.5 mm thick at the apex, yellow-orange at the top, paler and more yellow towards the base, more or less clavate. Odour indistinct. Basidiospores 6.5–7.5(8) × (4–)4.5–5.5 µm, ellipsoid to ovoid, smooth, hyaline in Melzer’s, monomorphic. Basidia 36–55 × 6–8 µm, sterigmata 3–6 µm long, clavate, (1-)4-spored, clamped or not, monomorphic. The ratio of basidial to basidiospore length > 5. Lamellartrama subregular. Pileipellis cutis, clamped irregularly, hyphae 2–4 µm wide. ITS sequence (GenBank OR785987): Distinct from all other species of Hygrocybe and differing from e.g., H. sp. (GenBank no HM020688) and H. reidii (GenBank no. EU784347), placed in the same clade, by more than 10% (> 70 substitutions and indels).
Habitat: Dry, deciduous, tropical forest, on sandy soil with Euphorbiaceae-Cephalocroton, Rubiaceae, Grevia, Salacia madagascariensis and Diospyros.
Distribution: So far only known from Madagascar.
Material examined: Madagascar. Mahajanga, Boeny, Ankarafantsika National Park, tropical dry forest, 13 Feb 2017, Meilinda Sulastri 000252026 (K-M, holotype), MAF17- 091.
GeneBank numbers: OR785987
Notes: A plano-convex pileus with a depression in the centre that is orange to reddish-orange and yellow towards the margin as well as two coloured stipe that is paler towards the base and ellipsoid to ovoid spores 6.5–7.5(–8) × (4–)4.5–5.5 µm are typical for this species. It belongs to Hygrocybe subgenus Pseudohygrocybe and has the typical characteristics of that subgenus: the mean ratio of basidial to basidiospore length is > 5. In addition, the lamellar trama is subregular and the basidia and spores are monomorphic.
Marasmiaceae Roze ex Kühner, Bull. mens. Soc. linn. Lyon 49: 76 (1980)
Marasmiaceae is a family of basidiomycete mushrooms, and the members of the family have characteristics dry cap with white, yellow, grey, brown or reddish-brown. Stipe fibrous, lacks a universal veil, spore elliptical to elongated, with thin, smooth, non-starchy, and clamp connection. These species are saprobic and are typically found in forest, and shrubbery, on fallen branches, leaf litter, tree trunks, stumps, twigs, and dead grass (Kuhner, 1980).
Marasmius Fr., Fl. Scan.: 339 (1836) [1835]
Marasmius is a mushroom genus that plays an important ecological role as a decomposer of leaf and wood litter and is reported worldwide, mostly from the tropics (Singer 1976, 1986; Desjardin 1989; Lodge et al. 1995), with most species being saprotrophic. Members of this genus are usually found on decaying leaves, trunks, branches, twigs, and wood debris (Tamur et al. 2019). Amoako-Attah et al. (2020) introduced a few parasitic species that cause diseases like horsehair blight in economically important plants like cocoa. Fries (1835) introduced the genus, but a complete diagnosis was later published without specifying the type species (Fries 1838). Subsequently, Marasmius rotula (Sco) was proposed as the lectotype by Singer and Smith (1946). The genus is usually characterized by thin and small or sometimes large and robust fruiting bodies, convex or campanulate, striated to the furrowed, white or intensely pigmented cap, pale and narrow gills, stipe thin and typically tough, horse-hair-like, often brown or darkly fuscous, pale cream to white spore-print, thin-walled, hyaline, smooth and non-dextrinoid basidiospores (Singer 1976, 1986; Antonín 2007; Oliveira et al. 2020, Antonίn and Noordeloos 2010) (Fig. 126).
[See PDF for image]
Fig. 126
Phylogram generated from maximum likelihood analysis based on ITS sequences data. Related sequences were obtained from Bhunjun et al. (2022), Oliveira (2020), and Zhang et al. (2023d). Eighty-seven strains are included in the sequence analysis, which comprises 627 characters with gaps. Marasmius yunnanensis (Dai 19782) and M. yunnanensis (Dai 19857) were used as the outgroup taxa. The tree topology of the ML analysis was similar to the BYPP. The best-scoring RAxML tree with a final likelihood value of − 6270.324271 is presented. The matrix had 364 distinct alignment patterns, with 3.67% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.247688, C = 0.198222, G = 0.217049, T = 0.337041; substitution rates AC = 0.884748, AG = 3.076732, AT = 0.975543, CG = 0.390992, CT = 3.644064, GT = 1.000000; gamma distribution shape parameter α = 0.727557. Bootstrap support values for ML equal to or greater than 65% and BYPP equal to or greater than 0.90 are given above the nodes. Newly generated sequences are in bold and blue
Marasmius qujingensis W.H. Lu., Karun. & Tibpromma, sp. nov.
Index Fungorum number: IF 900394; Facesoffungi number: FOF 14111 Fig. 127
[See PDF for image]
Fig. 127
Marasmius qujingensis (HKAS 127151, holotype). a Basidiomata. b Basidia. c Basidia and Basidioles. d Basidioles e Cheilocystidia. f Clamp connection. g Caulocystidia. h, i Basidiospores. Bars: a = 1 cm. b, c, d, g = 10 μm. e, f = 20 μm. h, i = 5 μm
Etymology: “qujingensis” refers to Qujing City where the species was first collected.
Holotype: HKAS 127151.
Pileus 9 to 30 mm in diam., hemispherical, convex with dark brown conical papilla at the center when young, expanding to broadly convex with age; smooth and entire, moisture-laden, glabrous; lead grey (2B2) at dark disc spot (2F2) when young, dull yellow and yellowish grey (3B2–3) at the disc with age. Lamellae adnate to shallowly adnate, medium spaced to rather distant, yellowish white (1A2) with a white edge. Stipe 25–70 × 5 mm, cylindrical, central, glabrous, waterlogged, moist, white, light orange (5A4) with age, hollow, some stipes arising directly from pale orange rhizomorphs. Basidiospores 5.5–10 × 3–3.5 μm (n = 25), thin-walled, smooth, hyaline, pyriform, ellipsoid. Basidia (16)21–35 × 4–6 μm, clavate, inamyloid, thin-walled, with 4 sterigmata. Cheilocystidia of Siccus-type broom cells; main body 8–12 × 6–8 μm, clavate to pyriform, hyaline, inamyloid, thin-walled. Pileipellis hymeniform, mottled, with Siccus-type broom cells, clavate to pyriform or turbinate, hyaline, inamyloid, and thin-walled. Pileus trama inamyloid, interwoven. Lamellar trama hyphae interwoven, cylindrical, smooth, hyaline, inamyloid, thin-walled, non-gelatinous. Clamp connections present.
Material examined: China, Yunnan Province, Qujing Normal University, 25° 31′ 16′′ N 103° 45′ 12′′ E, elev. 1628 m, 14 June 2022, QJ130, W. Lu & S. Karunarathna.
Additional specimens examined: China, Yunnan Province, Qujing Normal University, 25° 31′ 16′′ N 103° 45′ 12′′ E, elev. 1628 m, 14 June 2022, HKAS 128153(QJ309), W. Lu & S. Karunarathna.
GenBank numbers: ITS: OQ755409 (HKAS 127151), OQ755410 (HKAS 128153(QJ309).
Notes: Marasmius qujingensis is morphologically similar to M. torquescens and M. wynneae. However, M. qujingensis differs from M. wynneae by having beige gray, dry, wavy, and depressed at the center of the cap, and stipe, black at the base (Berkeley & Broome 1859). Marasmius qujingensis differs from M. torquescens by having light brown, yellowish-brown, reddish-brown, and dark brown in between, ridged to wrinkled edge pileus; pileipellis composed of mixed siccus-type broom cells and non-setulose cells (Quél 1872). In our phylogenetic study (Fig. 126), M. qujingensis formed a separate branch from all other known Marasmius species with relatively good statistical support (100% ML/1.00 PP). In addition, M. qujingensis differs from M. coarctatus and M. trinitatis in the comparison of ITS gene fragments (M. coarctatus: 40/627, 6.37%; M. trinitatis: 74/627, 11.8%, without gaps). Therefore, M. qujingensis is introduced as a distinct new species based on morphological and molecular evidence.
[See PDF for image]
Fig. 128
Phylogenetic tree generated by ML analysis based on combined ITS and LSU sequence data of Gymnopus, Collybiopsis and its closely related taxa. The analyses included 121 strains and the 599 tree is rooted with Omphalotus olearious (CBS102282), Omphalotus japonicus (CBS374.51), Omphalotus ovilascens (TENN56257), and Omphalotus flagelliformis (HKAS 76645). The tree topology of the ML analysis was similar to the BYPP. The best-scoring RAxML tree with a final likelihood value of − 20,386.607306 is presented. The matrix had 787 distinct alignment patterns, with 21.27% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.246718, C = 0.181088, G = 0.257374, T = 0.314820; substitution rates AC = 0.796329, AG = 5.315752, AT = 1.623571, CG = 0.450011, CT = 5.454699, GT = 1.000000; gamma distribution shape parameter α = 0.256828. RAxML bootstrap support ≥ 50% and Bayesian posterior probabilities ≥ 0.50 (BYPP) are indicated above the nodes. The scale bar indicates 0.04 changes per site. New sequences recovered in this study are in bold black
[See PDF for image]
Fig. 129
Collybiopsis gibbosa (HYW117 new geographical record). a, basidiocarps in natural environment, b, basidiocarps in laboratory, c–e, basidiospores, f, basidia, g, basidioles, h cheilocystidia, i, caulocystidia. Scale bar: b = 5 cm, c–e = 3 μm, f–g = 10 μm, h–i = 20 μm
Omphalotaceae Bresinsky, in Kämmerer, Besl & Bresinsky, Pl. Syst. Evol. 150(1–2): 113 (1985)
For taxonomic treatment of this family, we followed Matheny et al. (2006) and Oliveira et al. (2019).
Collybiopsis (J. Schröt.) Earle, Bull. New York Bot. Gard. 5: 415 (1909)
Notes: For the taxonomic treatment of Collybiopsis, we follow Petersen & Hughes (2021) in Figure 128.
Collybiopsis gibbosa (Corner) R.H. Petersen, in Petersen & Hughes, Mycotaxon 136(2): 342 (2021)
Index Fungorum number: IF 556187; Facesoffungi number: FoF 16480 Fig. 129,
[See PDF for image]
Fig. 130
Gymnopus ailaoensis (HKAS 131326, holotype). a, basidiocarps on dead wood, b, basidiocarps in laboratory, c–e, basidiospores, f, basidioles, g, Pileocystidia, h cheilocystidia, i, caulocystidia. Scale bar: b = 5 cm, c–e = 5 μm, f–i = 20 μm
Basionym: Marasmius gibbosus Corner. 1996. Nova Hedwigia, Beih. 111: 53.
= Marasmiellus gibbosus (Corner) J.S. Oliveira. 2019. Mycol. Prog. 18: 734.
= Gymnopus gibbosus (Corner) A.W. Wilson, Desjardin & E. Horak, Sydowia 56: 175. 2004.
Basidiocarps medium-sized, Pileus 7.0‒33.4 mm in diam, convex when young, becoming undulating with age; margin inflexed or slightly reflexed, striated; surface smooth, hygrophanous, center light brown (6D6) to gray, orange (6B5), orange, white (6A2) to whitish (6A1) at margin. Lamellae subdistant, adnexed, white (1A1), margin even. Stipe 26.4‒55.6 × 2.8‒5.8 mm, central, cylindrical, glabrous, tapered toward apex, swollen to the side at base, whitish (1A1) to greyish yellow (4B4). Basidiospores 4.5‒8.3 × 2.8‒6.0 μm, average 6.00 × 4.06 μm, Q = 1.0‒2.1 (mean = 1.49), subglobose to obovoid to lacrymoid, smooth, thin-walled, with drops. Basidia 14.7‒24.8 × 3.7‒6.8 μm, 4-spored, clavate to narrowly clavate. Pleurocystidia absent. Cheilocystidia 30.5‒128.1 × 5.7‒13.3 μm, narrowly cylindrical, subinflated, thin-walled, smooth, branched. Caulocystidia 16.1‒50.3 × 3.2‒6.0 μm, narrowly cylindrical to flexuose, irregular. Pileipellis hyphae 3.6‒8.3 μm wide, thick-walled, intricate trichoderm. Stipitipellis hyphae 3.1‒7.2 μm, curved trichoderm. Clamp connection present in all tissues. Odour and taste not recorded.
Habitat: Gregarious in soil covered with leaf litter under Samanea saman.
Materials examined: Thailand, Chiang Rai, Muang District, Mae Fah Luang University, 20° 03′ 14″ N, 99° 53′ 37″ E, alt. 413 m, 11 August 2020, Yuwei Hu, HYW117(MFLU 24-0038). Chiang Rai, Muang District, Mae Fah Luang University, 20° 03′ 17″ N, 99° 53′ 43″ E, alt. 411 m, 13 August 2020, Yuwei Hu, HYW127(MFLU 24-0039).
GenBank numbers: ITS: OR815366(MFLU24-0038),OR815367 (MFLU 24-0039), LSU: OR815384(MFLU24-0038),OR815385 (MFLU 24-0039)
Notes: This species was first described from Singapore, Java, and Republic of São Tomé and Príncipe as Gymnopus gibbosus (Corner 1996; Wilson et al. 2004; Desjardin & Perry 2017). The collections from this study are morphologically similar to Gymnopus gibbosus. Thai collections of Collybiopsis gibbosa has moderately-sized basidiocarps; light brown disc and whitish margin; white lamellae; grayish yellow stipe (brown stipe for Gymnopus gibbosus); basidiospore mean 6.00 × 4.06 μm (8.5 × 4.7 for Gymnopus gibbosus); cylindrical cystidia; pileipellis and stipitipellis hyphae intricate trichoderm (cutis-type for Gymnopus gibosus). Based on Petersen & Hughes (2021), Gymnopus gibbosus is synonymous with Collybiopsis gibbosa. In ITS sequence phylogenetic analysis, HYW117 and HYW127 showed 100% similarity to Brazil collections of URM90006 and URM90012, and 99.72% to Australian collection MEL: 2382838. In combined ITS-nrLSU phylogenetic analysis, Collybiopsis gibbosa is sister to Collybiopsis luxurians and Collybiopsis pseudoluxurians clade with 80% BS and 1.0 PP support (Fig. 128). The collections of this study represent a new geographical record of Thailand.
Gymnopus (Pers. 1801: 302) Roussel (1806: 62)
For the taxonomic treatment of Gymnopus, we follow Hu et al. (2022).
Gymnopus ailaoensis Y. Hu, sp. nov.
Index Fungorum number: IF 902410; Facesoffungi number: FOF 16092 Fig. 130
[See PDF for image]
Fig. 131
Maximum likelihood phylogenetic tree based on ITS sequences of Pleurotus species. Bootstrap support values (BS ≥ 70%) and Bootstrap values of BI ≥ 0.90 are shown on the branches. The newly generated sequences are indicated in bold. Hohenbuehelia atrocoerulea (KU355304) and H. Petaloides (KU355346) are outgroup. Sixty-three strains are included in the analysis, which comprises 580 characters after trimming with the Trimal program in CIPRES. Bayesian analysis, the best-fit substitution models were selected from jModelTest2 version 2.1.6 (Darriba et al. 2012) on XSEDE. The best-fit model for ITS was HKY + I + G. Bayesian analysis was performed in MrBayes version 3.2.7a (Ronquist et al. 2012). Two runs of five chains were run for 500,000 generations and sampled every 200 generations. The average standard deviation of the split frequencies was 0.081120 at the end of the runs. The burn-in phase (25%) was estimated by checking the stationarity in the plot generated by the sump command
Etymology: ‘‘ailaoensis’ refers to the locality “Ailao Mountains” which the species was recorded.
Holotype: HKAS 131326
Basidiocarps medium-sized, Pileus 5.3‒30.6 mm in diam, 2.2‒10.3 in height, hemispherical to slightly campanulate when young, becoming convex to plano-convex with age; margin straight to slightly reflexed, translucent striate, surface smooth, hygrophanous, light orange (6A5) to yellowish gray (6C6), margin white (1A1). Lamellae crowded, adnexed, white (1A1), margin even. Stipe 11.6‒38.6 × 1.9‒3.9 mm, central, cylindrical to compressed, hygrophanous, pliant, tapered toward apex, swollen to the side at base, with white basal mycelium, yellowish white (2A2) when young, light orange (6A5) to yellowish gray (6C6) with age. Basidiospores 3.5‒7.8 × 2.4‒3.9 μm, average 5.27 × 3.03 μm, Q = 1.4‒2.6 (mean = 1.75), ellipsoid to oblong to lacrymoid, smooth, thin-walled, with drops. Basidia not seen, Basidioles narrowly clavate, curved. Pileocystidia 34.2‒103.5 × 3.6‒8.8 μm, cylindrical, curved, slightly swollen toward apex. Cheilocystidia 17.7‒27.8 × 3.1‒5.6 μm, subcylindrical to narrowly utriform, tapered toward apex. Caulocystidia 27.0‒57.9 × 3.7‒8.2 μm, narrowly cylindrical, irregular. Pileipellis hyphae 3.3‒7.2 μm wide, thin-walled, intricate trichoderm. Stipitipellis hyphae 4.9‒8.1 μm, cutis, thick-walled. Clamp connection present in all tissues. Odour and taste not recorded.
Materials examined: CHINA, Puer, Jingdong, Ailao Mountains, 24° 32′ 45″ N, 101° 01′ 48″ E, alt. 2,503 m, 8 June 2021, Yuwei Hu, ALS33 (HKAS 131326, holotype).
Other materials examined: CHINA, Puer, Jingdong, Ailao Mountains, 24° 32′ 35″ N, 101° 01′ 39″ E, alt. 2,491 m, 8 June 2021, Peter E. Mortimer, ALS70 (HKAS 131325, paratype); Puer, Jingdong, Ailao Mountains, 24° 32′ 19″ N, 101° 01′ 07″ E, alt. 2,483 m, 8 June 2021, Peter E. Mortimer, ALS71 (HKAS 131324, paratype).
Habitat: Gregarious on dead wood.
GenBank numbers: ITS: OR815363 (HKAS 131326), OR815364 (HKAS 131325), OR815365 (HKAS 131324); nrLSU: OR815381 (HKAS 131326), OR815382 (HKAS 131325), OR815383 (HKAS 131325).
Notes: Phylogenetically, Gymnopus ailaoensis belongs to Gymnopus sect. Levipedes and forms a sister lineage with G. earleae and G. bicolor. Gymnopus ailaoensis is similar to Gymnopus earleae by basidiocarp size (cap 0.5‒2 cm vs. 1‒3.5 cm; stipe 11.6‒38.6 × 1.9‒3.9 vs. 20–50 × 5 mm) and basidiospore size (3.5‒7.8 × 2.4‒3.9 vs. 4.5‒7 × 2‒3.5). However, Gymnopus ailaoensis is distinct by orange to yellowish gray pileus and stipe while Gymnopus earleae presents dark brown to dark reddish-brown cap (Murrill 1916). Gymnopus ailaoensis is morphological similar to G. bicolor, while G. ailaoensis has an entire subcylindrical to narrowly utriform cheilocystidia (17.7‒27.8 × 3.1‒5.6 μm,) and G. bicolor has a broadly clavate and lobulate cheilocystidia (17.5‒25.5 × 6.5‒9.5 μm) (Wilson et al. 2004).
Pleurotaceae Kühner, Bull. mens. Soc. linn. Lyon 49: 184 (1980)
Pleurotus (Fr.) P. Kumm., Führ. Pilzk. (Zerbst): 24 (1871)
Pleurotus belongs to Pleurotaceae Kühner in Agaricales Underw. He et al. (2019) accpeted 25 species in this genus. The typical characteristics of this genus are pleurotoid shape, white basidiospores, a dimitic hyphal system, and the presence of clamp connections (Largent 1986), and P. ostreatus (Jacq.) P. Kumm. is the type species (He et al. 2019). Pleurotus species are mostly white-rot fungi that grow on logs, hardwood trees, and decaying wood; and this genus has been found in both tropical and temperate regions (Chang and Miles 2004). All members of Pleurotus areconsidered edible, and they are well known as various oyster mushrooms. They have been cultivated on an economic scale throughout the world (Sánchez 2010). Pleurotus species are not only edible; but they also provide good bioactivities from various extracts such as anti-cancer, anti-diabetic, antimicrobial, antiviral, antioxidant, anti-inflammatory, anti-lipidemic, hyperglycemic, and hypotensive (Cohen et al. 2002; De Silva et al. 2012). In addition, the mycelium of some Pleurotus species can to trap nematodes by paralyzing them with a toxin such as P. cornucopiae (Paulet) Quél., P. ostreatus, and P. subareolatus Peck, (Barron and Thorn 1987; Thorn et al. 2000).
In previous records of Pleurotus in Laos, only five species of Pleurotus have been reported, namely, P. djamor (Rumph. ex Fr.) Boedijn, P. eryngii (DC.) Quél., P. giganteus (Berk.) Karun. & K.D. Hyde, P. ostreatus (Jacq.) P. Kumm., and P. pulmonarius (Fr.) Quél. (Læssøe et al. 2019, Łuczaj et al. 2021, Phonemany et a. 2021). Here we report a new distribution record of P. tuber-regium (Fr.) Singer from Laos (Fig. 131).
Pleurotus tuber-regium (Fr.) Singer, Lilloa 22: 271 (1951) [1949]
Index Fungorum number:198804, Facesoffungi number: 13861 Fig. 132
[See PDF for image]
Fig. 132
Basidiomata of Pleurotus tuber-reguim in the field (a HNL501189; b, c HNL503400). d Basidia, e Basidiospores, f Irregular cutis, g Cheilocystidia, h Skeletal hyphae, i Generative hyphae. Scale bars a–c = 2 cm; d = 10 µm; e = 5 µm; f = 50 µm; g = 20; h, i = 10 µm
Basidiomata medium to large size, tuber-liked. Pileus 120–150 mm diam., deeply infundibuliform, brown to light brown (6C5) to greyish yellow (4B4), fuliginous and minutely scurfy squamulose, with white in the center, then fading with age to amber yellow (4B6), remaining dark and dry; margin incurved or involute when mature, appendiculate and wavy. Lamellae decurrent, very crowded, more than 5 lamellulae of more than five, narrow, brownish orange (5C4), pale-yellow to cream-yellow (4A3, 4B3), with the entire fuscous-grey edge, very sinuous on drying. Stipe 60–150 × 6–10 mm, wider at base zone (15–20 mm), central, cylindrical to subcylindrical, solid, surface minutely scurfy, covered with concolorous squamulose as pileus. Annulus not distinct, with greyish white to greyish (1A1) subfloccose at apex, then becoming ruptured and disappearing. Context thick in the center of the pileus, white, when young soft and fleshy-spongy, then becoming hard coriaceous with age; white and solid in stipe. Odor not severed. Taste not severed. Spore print white.
Basidiospores [150,3,3] (5.1–)5–6–7(–7.3) × (2.5–)2.6–3–4(–5) µm, Q = (1.41–)1.45–1.9–2.49(–2.67), cylindrical to subcylindrical, inamyloid, smooth, white, thin-walled. Basidia (22.2–)22–26–27.5(–27.8) × (5–)5–6–7.5(–7.9) µm, clavate, 4-spored, hyaline, thin-walled. Cheilocystidia (23.1–)23–34–49(–36.7) × (5.6–)6–7–10.5(–10.5) µm, submoniliform, subcylindrical, flexuose, subventricose, protruding, smooth, hyaline, thin-walled. Pleurocystidia absent. The hyphal system consists of a dimitic hyphal system with skeletal hyphae and generative hyphae, similar in the stipe context and the pileus context; skeletal hyphae 4–8 µm thick-walled, rarely branching and narrow, with clamp connections; generative hyphae 4–6 µm, not bloat, slightly thick-walled, branched, with prominent clamp connections. Pileus covering a cutis composed of hyaline hyphae and cylindrical terminal elements, 3–5 × 90–100 µm, thin-walled. Stipe covering same as on pileus. Clamp connections are present in all tissues.
Habitat and distribution: grow in rotten buried wood in a rain forest, solitary, distributed in many countries such as China (Karunarathna et al. 2016), Laos (this study), Madagascar (Vizzini et al. 2019), Malaysia as Panus tuber-regium (Fr.) (Corner 1981), and Nigeria (Oso 1977).
Materials examined: Laos, Xieng Khouang Province, Phoukout District, Gnotphe, Na Phouang hamlet, 23 May 2016, Phongeun Sysouphanthong (HNL501189); ibidem collected Vientiane Capital, Xaythany District, Houay Yang Preserve Forest, date 15 September 2015, Phongeun Sysouphanthong (HNL503400).
GenBank numbers: ITS: MK894133, MK894134.
Notes: Pleurotus tuber-regium is edible and has been cultivated in many countries (Okhuoya and Okogbo 1990; Okjuoya 1991; Isikhuemhen and Okhyoya 1996). This species is not only edible, but also good in bioactivities such as antioxidant, antitumor, antihyperglycemic, and antihypertensive agents (Lin et al. 2020). Pleurotus tuber-regium is found in both tropical and subtrop. Earlier, P. tuber-regium was originally described as Agaricus tuber-regium Fr., and later, many synonyms were reported in the Species Fungorum (http://www.speciesfungorum.org) e.g., Lentinus tuber-regium (Fr.) Fr., Pachyma tuber-regium Fr., Panus tuber-regium (Fr.), and Pocillaria tuber-regium (Fr.) Kuntze. The morphology of specimens of P. tuber-regium from Laos is very similar to the descriptions of the epitype described from China by Karunarathna et al. (2016), and it is consistent with those described from Malaysia as Panus tuber-regium by Corner (1981). Lao specimens are smaller than specimens from other areas, and the size of the species might depend on the condition of habitats (Hewitt et al. 2016). However, in the phylogenetic analysis (Fig. 132), P. tuber-regium (MK894133, MK894134) from Laos grouped with P. tuber-regium from other regions with high bootstrap support (ML 100%, BI 0.96). In this work, we report a new country record of P. tuber-regium from Laos (Fig. 133).
[See PDF for image]
Fig. 133
Phylogram generated using maximum likelihood analysis method based on ITS, 28S and RPB2 sequences. The aligned dataset consisted of 152 sequences from 88 representative taxa with 2043 characters including gaps (ITS = 45, LSU = 66, RPB2 = 41). Caloboletus rubripes (Thiers) Vizzini, Caloboletus xiangtoushanensis Ming Zhang, T.H. Li & X.J. Zhong and Caloboletus inedulis (Murrill) Vizzini were used as outgroup taxa following Zhao et al. (2014), Zhao and Shao (2017) and Zhang et al. (2022). Type strains are in bold, and the newly generated sequence is in blue
Boletales E.-J. Gilbert, Les Livres du Mycologue Tome I-IV, Tom. III: Les Bolets: 83 (1931)
Notes: Boletales is a monophyletic group of typically fleshy, pored mushrooms that also contains gilled, gasteromycete, and resupinate or crust-like forms (Binder & Hibbett 2006). The majority of the species in this order have a stipitate-pileate habit with tubular hymenophore (Cao et al. 2021). Six major lineages are recognized (Boletineae, Paxillineae, Sclerodermatineae, Suillineae, Tapinellineae and Coniophorineae) within the order Boletales (Binder & Hibbett 2006). He et al. (2019) listed 141 genera in 16 families of Boletales based on morphology and phylogeny. According to Kirk (2019) and Cao et al. (2021), 2173 described species are placed in this order.
Boletaceae Chevall., Fl. gén. env. Paris (Paris) 1: 248 (1826)
Notes: Boletaceae is one of the most studied families of Boletales (Yang 2011; Wu et al. 2014). The species of Boletaceae are mainly characterized by a fleshy context and tubulose or lamellate or loculate hymenophore (Wu et al. 2014). Most of the species have obligate ectomycorrhizal association with different vascular plants. The family includes approximately 1000 species in 89 genera, with worldwide distribution (He et al. 2019). For the updated treatments we follow Wu et al. (2023) and Mao et al (2023).
Rubroboletus Kuan Zhao & Zhu L. Yang, Phytotaxa 188(2): 67 (2014)
Notes: Rubroboletus is a recently erected taxa typified by R. sinicus (W.F. Chiu) Kuan Zhao & Zhu L. Yang (Zhavo et al. 2014). The genus is characterized by a brown, pinkish to reddish pileal surface, an orange-red to the blood-red surface of the hymenophore, yellow tubes, pink to red reticula or spots on the stipe, a blue colour change of tissues when injured, and smooth basidiospores (Zhang et al. 2022). There are edible, medicinal, and poisonous species in the genus Rubroboletus. For the latest updated account of Rubroboletus, see Zhao et al. (2015), and Zhang et al. (2022).
Rubroboletus pruinosus Salna Nanu & T. K. A. Kumar, sp. nov.
Index Fungorum number: IF 902411; Facesoffungi number: FoF 14043; Fig. 134
[See PDF for image]
Fig. 134
Rubroboletus pruinosus (ZGCSN100, holotype). A, B, C, D, E Basidiomata. F Basidiospores. G Basidia. H Pleurocystidia. I Cheilocystidia. J Pileipellis. K Stipitipellis Scale bars: A, B, C, D, E = 10 mm, F, G, H, I = 10 µm, J, K = 20 µm
Etymology: The epithet “pruinosus” refers to the pruinose nature of pileus and stipe surface.
Holotype: ZGCSN100
Basidiomata small- to medium-sized. Pileus 15–25 mm in diam., hemispherical when young becoming applanate at maturation; surface grayish red to dark red, dry, pruinose, dark immediately turning blue to dark blue on bruising. Pileal context 5–15 mm thick, cream to white, turning blue to dark blue on bruising. Hymenophore somewhat depressed around stipe apex; surface pinkish to reddish, fading with aging, immediately changing to blue with bruising; pores round to angular, 1–2 per mm; tubes up to 3.5 mm long, yellowish, rapidly changing to blue on bruising. Stipe 20–40 × 5–15 mm, reddish, paler towards the middle, with scattered darker pruinose patches, changing to dark blue on bruising, subcylindrical, broadening towards the base. Stipe context pale yellow, turning blue quickly when cut. Basal mycelium white. Odour not distinct. Spore print not obtained.
Basidiospores 8–10 × 4.5–5.5 µm (Q = 1.6–2 µm, Qm = 1.74 µm), broadly ellipsoid, smooth, with a large guttule, thin- to thick-walled, hyaline to pale yellow, inamyloid. Basidia 24–31 × 9–13 µm, clavate, thin-walled, 4-spored, hyaline, inamyloid, sterigmata up to 4 µm long. Pleurocystidia 28–60 × 7–9 µm, lageniform, mostly with apical protrusions up to 11 µm long, thin-walled, yellow or yellowish brown in water, inamyloid. Cheilocystidia 26–40 × 6–10.5 µm, lageniform, utriform, abundant, thin-walled, hyaline to pale yellow in water, inamyloid. Pileipellis a cutis disrupted by trichodermial patches of ascending to erect hyphae with cylindrical end cells; end cells 2–6 µm wide, thick-walled (wall thickness up to 1.5 µm), yellowish to bright yellow in water; hyphae 3–6 µm wide, thick-walled (wall thickness up to 1.5 µm), with reddish brown plasmatic pigment. Pileal trama interwoven; hyphae 4–10 µm wide, thin-walled, hyaline to pale yellow, inamyloid. Stipitipellis a cutis disrupted by trichodermial patches; terminal elements 2–4 µm wide, slightly thick-walled (wall thickness up to 0.5 µm), cylindrical, with brownish yellow plasmatic pigments. Stipe trama interwoven, hyphae 3–10 µm, hyaline, thin-walled.
Ecology and distribution: on soil, scattered under Xanthophyllum arnottianum Wight.
Material examined: India, Kerala State, Kollam District, near Ammayambalam temple, 10 August 2021, Salna Nanu (ZGCSN100, holotype).
GenBank numbers: ZGCSN100: ITS = OQ504748: LSU = OQ472490: RPB2 = OQ689073
Notes: Rubroboletus pruinosus is characterized by basidiomata with grayish red to dark red, pruinose pileus, pinkish to reddish pore surface, yellowish tubes, colour change of the pileal and stipe context to blue immediately on bruising, pruinose stipe surface, medium-sized basidiospores, and a pileipellis that is basically a cutis disrupted by trichodermial patches. Rubroboletus pruinosus differs from the closely related R. legaliae in having pinkish to reddish pore surface, pruinose stipe surface, smaller basidiospores, and the absence of caulocystidia. A comparative study of the ITS, 28S and rpb2 gene sequences reveals its genetic distinction. Rubroboletus legaliae shows resemblances with R. pruinosus in having depressed hymenial region around the stipe apex, yellowish tubes, colour change of the pileal context to blue when bruised, and pleurocystidia and cheilocystidia with identical shapes and dimensions. However, R. legaliae differs in having larger basidiomata, reticulated stipe surface, larger basidiospores (10–15 × 4.5–5 µm), pileipellis that is an intricate trichodermium and by the presence of caulocystidia. Rubroboletus rhodosanguineus shows similar characters with R. pruinosus like pileus colour, context of the pileus and stipe changing to blue when hurt and subcylindrical stipe. However, R. rhodosanguineus has a smell of overripe fruit, yellowish pileal context, glabrous pileal surface and is mostly distributed in North and Central America and Europe.
Phylogenetically, R. pruinosus is clustering as a separate taxon within the Rubroboletus clade of the Pulveroboletus group in the Boletaceae family and is distinct from all other Rubroboletus species. The species is described here as novel based on the molecular and morphological evidence.
Corticiales K.H. Larss., in Hibbett et al., Mycol. Res. 111(5): 540 (2007)
Corticiales K.H. Larss., in Hibbett et al., Mycol. Res. 111(5): 540 (2007)
Notes: Based on molecular phylogenetic analyses, the order was established by Karl-Henrik Larsson (in Hibbett et al. 2007) and, apart from the type family Corticiaceae, includes the families Dendrominiaceae, Punctulariaceae, and Vuilleminiaceae (Ghobad-Nejhad et al. 2010, 2021; He et al. 2019; Wijayawardene et al. 2022). Corticiales comprises genera with corticioid effused to discoid basidiomata with a smooth hymenophore, monomitic hyphal system with usually clamped hyphae and smooth basidiospores which are white to pink in mass; ecologically they are saprobionts, parasites or lichenicolous (Hibbett et al. 2007).
Corticiaceae Herter, Krypt. -Fl. Brandenburg (Leipzig) 6(1): 70 (1910)
Notes: Corticiaceae is a well-known family of the Agaricomycetes which, in a traditional, morphological concept, contained the majority of basidiomycetes producing resupinate, corticioid basidiomata. However, molecular phylogenetic analyses revealed that most of the genera traditionally classified within Corticiaceae belong to other families, and the family Corticiaceae was restricted to few genera, i.e. Basidiodesertica, Corticium, Disporotrichum, Erythricium, Giulia, Laetisaria, Marchandiomyces, Mycobernardia, Tretopileus, and Waitea (Ghobad-Nejhad et al. 2021). Except for Mycobernardia, all genera of Corticiaceae are known to produce sclerotia, bulbils, pycnidial or hyphomycetous anamorphs (Ghobad-Nejhad et al. 2021). Ecologically, the Corticiaceae contain saprobionts and parasites of plants, fungi and lichens (Ghobad-Nejhad et al. 2021).
Stegonsporiicola Voglmayr, gen. nov.
Index Fungorum number: IF 902412; Facesoffungi number: FoF 16071
Etymology: Referring to its host genus Stegonsporium (Stilbosporaceae, Diaporthales).
Fungicolous. Sexual morph: Not observed. Asexual morph: Hyphomycetous. Conidiomata sporodochial. Mycelium immersed in the substrate, composed of septate, branched, smooth, hyaline hyphae. Conidiophores branched, hyaline, with clamps at septa. Conidiogenous cells holoblastic, monoblastic, terminal, hyaline. Conidia acrogenous, solitary, subglobose to globose, multicellular, dictyosporous, bulbil-like, consisting of several swollen globose conidial cells, hyaline when immature, becoming orange when mature.
Type species: Stegonsporiicola aurantiaca Voglmayr
Notes: The phylogenetic analyses based on an SSU-ITS-LSU matrix place Stegonsporiicola as a distinct clade within Corticiaceae (Fig. 135), in an unsupported sister-group relationship to the plant-parasitic genus Waitea. Within Corticiaceae, Stegonsporiicola is morphologically well characterised by sporodochial conidiomata producing comparatively small bulbil-like dictyoconidia consisting of few swollen conidial cells, while the bulbils (sclerotia) of other genera of Corticiaceae are much larger and not produced in sporodochia (DePriest et a. 2005; Diederich and Lawrey 2007; Lawrey et al. 2008; Maharachchikumbura et al. 2021). Based on its phylogenetic position and its distinct morphology, the new genus Stegonsporiicola is therefore established. The orange colour of the conidia of Stegonsporiicola relates well to the known asexual morphs of other Corticiaceae which are commonly bright orange or pink (DePriest et al. 2005). With Stegonsporiicola, a fungicolous genus is added to Corticiaceae, which so far contains saprotrophs, plant parasites, lichenicolous and endolichenic members (Ghobad-Nejhad et al. 2021) (Fig. 135).
[See PDF for image]
Fig. 135
Phylogram showing the best RAxML maximum likelihood tree obtained from the SSU-ITS-LSU matrix of Corticiales, showing the phylogenetic position of Stegonsporiicola aurantiaca within Corticiaceae. Representatives of Corticiales were selected from Ghobad-Nejhad et al. (2021). The matrix comprises 3844 characters with gaps. The tree is rooted with 5 species of Gloeophyllales. The best scoring RAxML tree with a final likelihood value of − 22,818.1338 is presented. The matrix had 1279 distinct alignment patterns, with 46.16% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.253797, C = 0.210673, G = 0.270671, T = 0.264859; substitution rates AC = 1.106791, AG = 3.174807, AT = 1.815956, CG = 0.494322, CT = 5.447541, GT = 1.000000. ML bootstrap support equal or higher than 50% is given near to each branch. The new isolates are in blue. The ex-type strains are in bold
Stegonsporiicola aurantiaca Voglmayr, sp. nov.
Index Fungorum number: IF 902413; Facesoffungi number: FoF 16072; Fig. 136
[See PDF for image]
Fig. 136
Stegonsporiicola aurantiaca (a–c, g, n, o WU-MYC 0049454, holotype; d WU-MYC 0049460; e, f CBS 149659, ex holotype; h WU-MYC 0049457; i–m WU-MYC 0049461). a, b Frontal view of the sporodochial conidiomata. c Conidioma in section, growing on an effete conidioma of Stegonsporium (arrows). d Conidioma in section, growing on effete stroma of Stegonsporium; arrows denoting perithecia of the host. e Frontal view of the culture. f Reverse view of the culture. g Conidioma in section with conidia on top and dead dark brown conidia of Stegonsporium mostly at the base. h, o Conidia. i, j Clamps on hyphae. k–n Developing conidia on conidiophores with clamps (arrows). All microscope mounts in water. Scale bars: a = 1 cm, b–d = 200 µm, g = 50 µm, h = 20 µm, i, j = 5 µm, k–o = 10 µm
Etymology: With reference to the orange conidiomata and conidia.
Holotype: WU-MYC 0049454
Fungicolous on effete conidiomata and stromata of Stegonsporium. Sexual morph: Not observed. Asexual morph:Conidiomata 0.5–1.5 mm diam, 200–350 µm high, sporodochial, immersed to erumpent, oval, polyangular to circular, surrounded by bark flaps. Conidiophores hyaline, branched, with clamps at septa, 2.5–6.5 µm wide, wall distinctly swelling to 2.5 µm with age. Conidiogenous cells holoblastic, monoblastic, terminal, hyaline, without scars or denticles. Conidia (13–)16–21(–27) × (11–)13.8–17.6(–21.6) μm (x̅ = 18.6 × 15.7 μm, n = 162), subglobose to globose, young subhyaline, pale yellow to orange with age, dictyosporous, bulbil-like, composed of 5–10 swollen cells of 5–12 µm diam., with a hyaline, smooth wall and fine pale orange granular contents, which sometimes fuse to an orange guttule of 1.3–4 μm diam.
Culture characteristics: Colonies on CMD cottony, first cream, soon becoming bright orange, reverse bright orange, becoming orange brown in the centre, reaching 55–60 mm diam. in 28 days at 22 °C. Mycelia are immersed at margins, superficial in the centre, effuse with diffuse regular edge. No sporulation observed in pure culture.
Material examined: (all on effete conidiomata or stromata of Stegonsporium sp., Stilbosporaceae, on Acer pseudoplatanus, Aceraceae)—AUSTRIA, Oberösterreich, Raab, Wetzlbach, 24 Dec. 2017, H. Voglmayr, BOS (WU-MYC 0049454, holotype), ex-type, CBS 149659; Niederösterreich, Gießhübl, Wassergspreng, 11 February 2009, H. Voglmayr (WU-MYC 0049455, paratype); Niederösterreich, Hardegg, Hammerschmiede, 8 May 2008, H. Voglmayr (WU-MYC 0049456, paratype); Niederösterreich, Lahnsattel, Donaudörfl, 23 April 2008, H. Voglmayr, LO (WU-MYC 0049457, paratype); Niederösterreich, Mariensee, 24 May 2010, H. Voglmayr (WU-MYC 0049458, paratype); Oberösterreich, Raab, Rotes Kreuz, Rothmayrberg, 15 May 2021, H. Voglmayr (WU-MYC 0049459, paratype); Oberösterreich, St. Willibald, Große Sallet, 30 April 2022, H. Voglmayr (WU-MYC 0049460, paratype); Steiermark, Gröbming, Öfen, 31 May 2009, H. Voglmayr (WU-MYC 0049461, paratype); Steiermark, Kleines Sölktal, Schwarzensee, 1 June 2009, H. Voglmayr (WU-MYC 0049462, paratype).
GenBank numbers: SSU-ITS-LSU: OQ102368 (WU-MYC 0049454, ex holotype), OQ102369 (WU-MYC 0049457, ex paratype).
Notes: The bright orange sporodochial conidiomata producing bulbil-like dictyoconidia with hyaline walls and orange granular content are characteristic for the species. Upon closer examination, S. aurantiaca was constantly found growing on effete conidiomata and stromata of Stegonsporium sp. (Stilbosporaceae, Diaporthales), indicating a mycoparasitic habit. In suitable habitats, it seems to be rather common but easily overlooked on dead corticated branches of Acer pseudoplatanus in spring, following the development of its Stegonsporium host which mostly occurs during the winter months, and it is apparently host-specific on Stegonsporium.
Hymenochaetales Oberw., in Frey, Hurka & Oberwinkler, Beitr. Biol. Pfl.: 89 (1977)
Notes: Oberwinkler (1977) raised the order Hymenochaetales by including the characteristics of xanthochoroid polypores in Série des Igniaires (Patouillard 1900) and the concept of subfamily Hymenochaetoidae (Donk 1948) of Hymenochaetaceae. Hymenochaetales is one of the largest orders of the class Agaricomycotina. Order Hymenochaetales includes mushrooms with various basidiomatal characters, from resupinate, effused-reflexed, imbricate, pileate, stipitate, coral-like to spathulate pilei with smooth, poroid, hydnoid or lamellate hymenophore. Microscopically, mono, mono-di, di or trimitic hyphal system, absence of clamp connections, and globose to cylindrical, smooth to finely ornamented basidiospores, with or without sterile elements such as cystidioles and setae are the features (Wang et al. 2023a, b). MycoBank recorded 19 associated families (http://www.mycobank.org) whereas 1653 taxa related to Hymenochaetales, submitted in GenBank (https://www.ncbi.nlm.nih.gov) (January 2024).
Type species: Hymenochaetaceae Imazeki and Toki, Bulletin of the Government Forest Experimental Station Meguro 67: 24 (1954) [MB#80883].
Hymenochaetaceae Donk, Bull. bot. Gdns Buitenz. 17(4): 474 (1948)
Notes: Donk (1948) proposed Hymenochaetaceae to include wood decaying xanthochoric causing white rot Angiosperms and Gymnosperms. Hymenochaetaceae is characterized by annual to perennial, yellow to brownish xanthochroic polypores with smooth, poroid or corticioid hymenophore and microscopic features such as monomitic, mono-dimitic or dimitic hyphal system, clampless generative hyphae, presence or absence of cystidioles and setae and thin to thick walled, hyaline to brown, globose to cylindrical basidiospores (Ryvarden 1991; Dai 2010).
Hymenochaetaceae sensu Donk is a species-rich family with more than 900 species worldwide (Index Fungorum 2024; https://www.indexfungorum.org/) with three major genera Phellinus, Hymenochaete and Inonotus. Recent investigations and phylogenetic analyses have discovered many new taxa in the family (Rajchenberg et al. 2015; Zhou et al. 2016a, b; Ryvarden 2018; Wu et al. 2020). Zhou et al. (2016a, b) introduced two new genera, Sanghuangporus and Tropicoporus from Inonotus linteus complex. Recently, Wu et al. (2022) revised the systematics of 672 poroid Hymenochaetoid fungi, and classified them under 34 genera, including seven new genera, 37 new species and 108 new combinations. In total, 71 associated genera were documented in MycoBank (http://www.mycobank.org) and 1,194 taxa were submitted in GenBank (https://www.ncbi.nlm.nih.gov) under the family Hymenochaetaceae (January 2024).
Fomitiporia Murrill, N. Amer. Fl. (New York) 9(1): 7 (1907)
Notes: Fomitiporia Murrill species are wood rotting characterized by their resupinate to pileate basidiomata, the presence of hymenial setae in some species, dextrinoid basidiospores, and a dimitic hyphal system throughout the basidioma (Decock et al. 2007; Dai 2010). The genus comprises approximately 70 described species (Index Fungorum 2023), with many of them being collected on living hosts, indicating some level of host-exclusivity or host-recurrence (Alves-Silva et al. 2020a, b; Amalfi et al. 2012, 2014; Dai et al. 2008). Due to limited morphological variation, several taxa represent morphological complexes of cryptic species. Therefore, phylogenetic reconstructions based on molecular data and data on host relationships and geographic distribution have played a crucial role in elucidating the diversity of the genus (Alves-Silva et al. 2020b; Decock et al. 2007; Morera et al. 2017; Vlasák and Kout 2011). Studies conducted in the Neotropical region have unveiled a high species richness (Alves-Silva 2020a, b; Amalfi and Decock 2013; Li et al. 2016; Morera et al. 2017). In this study, we describe two new pileate species of Fomitiporia from midwestern Brazil, and the phylogenetic tree for Fomitiporia is presented in Fig. 137.
[See PDF for image]
Fig. 137
Phylogenetic tree of Fomitiporia based on ML analysis of combined ITS, 28S, tef1-α, and rpb2 sequences. The matrix comprises 3816 characters with gaps. The tree is rooted with Phellinus uncisetus (MUCL47061 and MUCL46231). The best-scoring RAxML tree with a final likelihood value of − lnL 28,484.23 is presented. The matrix had 1948 distinct alignment patterns, with 9.55% of undetermined characters or gaps. The best partition scheme and evolution models were estimated in PartitionFinder 2 (Guindon et al. 2010; Lanfear et al. 2017) under the linked model of branch lengths, greedy search algorithm (Lanfear et al. 2012), and Akaike information criterion for model selection and. The best-fit models, matrices and trees are available at Alves-Silva et al. (2023). In the BI analysis, runs converged to stable likelihood values (− lnL = 28,067.41 and 28,075.73), and after the 50% burn-in, 7502 trees were used to compute a 50% majority-rule consensus tree and to estimate BPPs. ML bootstrap support (first set) ≥ 70% and Bayesian posterior probabilities ≥ 0.95 are given near each branch. The new species proposed are in blue
Fomitiporia exigua Alves-Silva, Góes-Neto & Drechsler-Santos, sp. nov.
Index Fungorum number: 900086; Facesoffungi number: FoF 13980; Fig. 138
[See PDF for image]
Fig. 138
Fomitiporia exigua (FLOR58558, holotype) a–e Basidiomata ex-situ. f–g Slightly dextrinoid basidiospores. Scale bars: a–c, e = 5 mm, d = 1 mm, f–g = 5 µm
Etymology: exigua, from exiguus (Latin) = little, small in all parts; it refers to the small basidiomata.
Holotype: FLOR 58558
White-rot fungus, found growing on living Protium heptaphyllum (Aubl.) Marchand, (Burseraceae). It is currently known in Brazil, in the Imerí Province, Amazonas state. Basidiomata perennial, pileate, sessile, and mostly broadly attached, subdimidiate, semicircular symmetric to asymmetric, spatulate, applanate to convex (in older basidiomata), obtriquetrous to rarely triquetrous, projecting 2–18.5 mm, 4–13 mm wide and 2–8 mm thick, with woody consistency when dried. Pileus glabrous, concentrically zonate with multiple narrow bands, moderately sulcate, brown, dark brown [6EF (6–8)] to black; margin acute, folded, well delimited around pore surface, sterile, light brown, yellowish brown to brown [5DE (5–8)]. Pore surface light greyish brown (5D8) to greyish brown (6F3), (8F3); pores round to angular, 8–12/mm, (60–)70–120 µm diam. (x̅ = 84 µm); dissepiments entire, 20–90 µm (x̅ = 51.4 µm) thick. Tubes distinct stratified, up to 6 layers interleaved with context, individual tube layers thin, up to 0.5 mm thick, brown [5EF (3–4)] to grayish brown (5F3), the older layers filled with whitish mycelium. Context simple, thin, concentrically zonate, with dense texture and woody consistency, brown to dark brown [6EF (6–8)], with a distinct dark line near to surface. Hyphal system dimitic in all parts; generative hyphae simple septate, hyaline to pale yellow, sparingly branched, 2–2.5 μm diam.; skeletal hyphae golden brown to reddish brown, unbranched, thick-walled, in the context and hymenophoral trama 3–4 μm diam., the lumen 1–2 μm wide. Hymenial setae absent. Cystidioles rare, fusoid, lanceolate, hyaline, and thin walled. Basidia subglobose to globose, hyaline, tetrasporic, 7–9 × 6–8 µm (ave = 7.7 × 6.9 µm), Q = 1–1.3 µm (x̅ = 1.12 µm); basidioles identical in shape but slightly smaller than basidia. Basidiospores subglobose to globose, 4–4.5 × 4–4.5 µm (x̅ = 4.2 × 4.1 µm), Q = 1.0–1.3 µm (aveQ = 1.02 µm), hyaline, undextrinoid to slightly dextrinoid, cyanophilous, thick-walled, smooth. Crystals rhomboid, of variable size.
Material examined: Brazil, Amazonas: Novo Airão, Parque Nacional Anavilhanas, Igarapé Santo Antônio, 2° 24′ 42.5″ S 60° 58′ 08.9″ W, on a living Protium heptaphyllum (Aubl.) Marchand (Burseraceae), 6 Dec 2013, E.R. Drechsler-Santos 1256 (FLOR 58558 holotype).
GenBank numbers: LSU: KU663296, ITS: KU663323.
Notes: Fomitiporia exigua is mainly characterized by having perennial, small (less than 20 mm long and wide) and thin (less than 10 mm thick) basidiomata; convex pileus in the older basidiomata; and small pores (8–12/mm). Microscopically, it has small basidiospores, 4.2 × 4.1 µm on average and variable reactions in Melzer, from undextrinoid to slightly dextrinoid (Fig. 137f–g). Phylogenetically, it was recovered in F. langloisii–F. castilloi clade, sister to F. rondonii. As F. rondonii, it has pileate basidiomata, which is different from most of the species of this clade. Fomitiporia rondonii differs by having bigger basidiomata and basidiospores (5 × 4.8). In addition, F. exigua differs from all pileate species by having the smallest basidiomata in the genus and undextrinoid to slightly dextrinoid basidiospores.
Fomitiporia rondonii Alves-Silva & Drechsler-Santos, sp. nov.
Index Fungorum number: 900085; Facesoffungi number: FoF 13979; Fig. 139
[See PDF for image]
Fig. 139
Fomitiporia rondonii (FLOR58557, holotype) a Basidiomata in-situ. b Black line near to the surface. c Obtriquetrous basidioma, context, and indistinct tube layers. (ICN202298) d Whitish abhymenial surface. e Triquetrous basidioma. (FLOR58557 holotype) f Strongly dextrinoid basidiospore. (ICN202298) g–h Moderately to strongly dextrinoid basidiospores. Scale bars: a–e = 30 mm, f–h = 5 µm
Etymology: rondonii, named in honor of Marechal Cândido Mariano da Silva Rondon (1865–1958), a nominated for the Nobel Peace Prize in 1957, by his scientific and cultural contribution to the region in which this species occurs, through expeditions in 1900–1930.
Holotype: FLOR 58557
White-rot fungus found growing on dead unidentified angiosperm. It is currently known in Brazil, in the Cerrado and Xingu-Tapajós Provinces, Mato Grosso state. Basidiomata perennial, pileate, sessile, semicircular, obtriquetrous to triquetrous, occasionally concave, projecting 42–115 mm, 41–160 mm wide, and 25–85 mm thick at the base, with a woody consistency. Pileus glabrous, concentrically zonate with multiple narrow bands and up to 6 broad bands, moderately sulcate, faintly to strongly cracked, dull, dark brown [7F (4–8)] to black, greenish by algae and greyish to whitish in older basidiomata. Margin round, folded, thick, sterile, light orange [5A(5–6)] to golden brown, light brown, yellowish brown [5D(6–8)]. Pore surface greyish brown, brownish grey [7F (2–3)], dark brown [7F(4–5)] to cinnamon; pores round to angular, (5–)6–9(–10)/mm, (122–)127–155(–168) µm diam. (ave = 141.46 µm); dissepiments entire, (59–)67–223(–238) µm (ave = 102 µm) thick. Tubes distinct to mostly indistinctly stratified, with several layers, interleaved with context in older tube layers, brown [5EF (4–5)] to grayish brown (5E3), and the older layers filled with whitish mycelium. Context simple, up to 10 mm thick, concentrically zonate, with dense texture and with a woody consistency, brownish orange [6C (7–8)], light brown to brown [6DE (7–8)], with a distinct thicker dark line near to surface. Hyphal system dimitic in all parts; generative hyphae simple septate, hyaline to pale yellow, sparingly branched, 1.5–2.5 μm diam.; skeletal hyphae golden brown to reddish brown, unbranched, thick-walled, in the context and hymenophoral trama 3–4.5 μm diam., the lumen 0.8–2.5(–2.8) μm wide. Hymenial setae absent. Cystidioles rare, fusoid, lanceolate, hyaline, and thin walled. Basidia subglobose to globose, hyaline, tetrasporic, 8–9(–10) × 7–8 µm (ave = 8.5 × 7.3 µm), Q = 1–1.4 µm (aveQ = 1.18 µm); basidioles identical in shape but slightly smaller than basidia. Basidiospores subglobose to globose, (4–)5–6 × (4–)4.5–5(–6) µm (ave = 5 × 4.8 µm), Q = 1–1.1 µm (aveQ = 1.2 µm), hyaline, moderately to strongly dextrinoid, cyanophilous, thick-walled, smooth. Crystals rhomboid, of variable size.
Material examined: BRAZIL, Mato Grosso: Cuiabá, Parque Nacional de Chapada dos Guimarães, Sítio Véu de Noiva, 15° 24′ 23.19″ S 55° 50′ 12.14″ W, on dead trunk, unidentified angiosperm, 8 Feb 2015, G. Alves-Silva 726 (FLOR 58557 holotype); ibid., Itaúba, Rio Teles Pires, 11° 03′ 54.2″ S 55° 19′ 25.2″ W, on dead standing unidentified angiosperm, 01 Apr 2017, M. E. Engels (ICN202298 paratype).
GenBank numbers: FLOR 58557 holotype (LSU: KU663295, ITS: KU663322, tef1-α: KU663346, RPB2: KU663371); ICN202298 paratype (LSU: OQ148660, ITS: OQ148661, tef1-α: OQ148662, RPB2: OQ148663).
Notes: Fomitiporia rondonii is characterized mainly by having obtriquetrous basidiomata, mostly indistinctly stratified tube layers, and greyish to whitish pileus in older basidiomata. Microscopically, basidiospores are moderately to strongly dextrinoid, 5 × 4.8 µm on average. Phylogenetically, this species was recovered in F. langloisii–F. castilloi clade, in which only F. castilloi, F. bambusipileata, and F. exigua (sp. nov. also proposed in this paper) are pileate species retrieved among all neotropical resupinate species. Fomitiporia castilloi differs by having hymenial setae and F. bambusipileata that only occurs on bamboo culm (Amalfi and Decock 2013; Alves-Silva et al. 2020a). Concerning the remaining neotropical pileate species, F. rondonii is morphologically closely related to the species of F. apiahyna s.lat. clade and F. subtilissima clade. Fomitiporia subtilissima has thinner and distinctly reddish brown pileate basidiomata. Fomitiporia nubicola and F. apiahyna differ by having particular ecological requirements, the former occurs exclusively on Drimys spp. in Cloud forests and the latter in Araucaria Forest province, at 800–1000 m.a.s.l. (Alves-Silva et al. 2020b), whereas F. rondonii appears to be restricted to Cerrado–Amazon distribution. Fomitiporia conyana and F. murrillii differ by having slightly bigger pores [F. conyana: 6–8(–9)/mm and F. murrillii: (4–)5–7(–8)/mm], moreover, F. murrillii also has slightly bigger basidiospores [5–6(–7) × 5–6(–7) µm] (Alves-Silva et al. 2020b).
Fulvifomes Murrill, North. Polyp. (New York): 49 (1914)
Notes: Fulvifomes comprise species with annual to perennial, resupinate to pileate, applanate to ungulate basidiomata with homogenous to duplex context and poroid hymenial layer, bearing round to angular pores. Microscopically mono-dimitic to dimitic hyphal system, smooth, thick walled, colored, globose to ellipsoidal inamyloid basidiospores, presence or absence of cystidioles and setal elements are the distinguishing characteristics (Murrill 1914; Wagner and Fischer 2002; Dai 2010; Wu et al. 2022) (Fig. 140).
[See PDF for image]
Fig. 140
Phylogram generated from combined LSU and ITS sequence dataset of Fulvifomes spp. Based on the MegaBLAST search of NCBIs GenBank nucleotide database. Forty-one sequences were included in the combined analyses which comprises 1977 characters (932 characters for LSU, 1045 characters for ITS) alignment. Fomitiporella caryophylli (CBS44) and Phellinotus neoaridus (URM80362) were used as outgroup. Maximum likelihood tree was performed using MEGA X (Kumar et al. 2018) and the same data were used for Bayesian analysis using MrBayes v. 3.2.7 (Ronquist et al. 2012). Branches are labeled with Bayesian posterior probabilities more than 0.5 and maximum likelihood bootstrap higher than 50%. The novel species are in bold
Fulvifomes Murrill, was treated as a synonym of Phellinus Quél. (Ryvarden and Johansen 1980; Ryvarden 1991; Núñez and Ryvarden 2000). Later, Fiasson and Niemelä (1984) studied the nuclear behaviour and molecular data of Phellinus s.l. complex and supported the generic rank of Fulvifomes. Further, nuLSU rDNA sequence data supported the generic status of Fulvifomes, to include taxa with pileate basidiomata, dimitic hyphal system and coloured spores without setal elements (Wagner and Fischer 2002). Dai (2010) included a few resupinate basidiomes and species with hymenial setae in Fulvifomes. Nucleotide sequences for 32 different species are available in NCBI (https://www.ncbi.nlm.nih.gov). In total, 62 associated taxa are recorded in MycoBank (http://www.mycobank.org), and 90 species are documented under Fulvifomes in Index Fungorum. Type species: Fulvifomes robiniae Murrill, Bulletin of the Torrey Botanical Club 30 (2): 114 (1903) [MB#152311] (Fig. 140)
Fulvifomes subthailandicus E. Arumugam, S. Gunaseelan, K. Kezo & M. Kaliyaperumal sp. nov.
Index Fungorum number: IF900218; Facesoffungi number: FOF13988; Fig. 141
[See PDF for image]
Fig. 141
Fulvifomes subthailandicus (MUBL4025, holotype) Microscopic structures of Fulvifomes subthailandicus (drawn from the holotype). a Basidiomata. b Pilear surface. c Pore surface. d Context. e Stratified tube layer. f Tramal hyphae. g Context hyphae. h Basidioles. i Basidia. j Basidiospores k. phloxine. l Water. m Cotton blue. n Melzer. Scale bars: c = 1 mm, d–e = 10 mm, f–n = 10 µm
Etymology: The species epithet “subthailandicus”, signifies the close relation with the species Fulvifomes thailandicus.
Holotype: MUBL4025
Basidiocarp perennial, solitary, pileate, sessile, woody hard when fresh, woody hard and light weight when dry. Pileus dimidiate, meagrely ungulate, projecting up to 14 cm wide and 12.5 cm in length. Pilear surface radially cracked concentrically zonate, greyish brown (7D2) near the attachment to brownish grey (7F5) towards the margin. Margin entire, acute, velutinate, brown (7E7). Pore surface brown (7E5). Pores regular, round, 4–5 per mm. Context less than 1 mm, homogenous, brown (7F7). Tubes brown (6E7), 3.3 cm in length, tube layer stratified with intermittent context, each stratum up to 2 mm.
Hyphal system dimitic, tissue darkening with KOH without swelling. Context generative hyphae thin to thick walled, hyaline to dark brown, simple septate, rarely branched, 2.5—5.2 µm dia., skeletal hyphae, thick walled with narrow to wide lumen, brown, aseptate, unbranched, 3.9—4.5 µm dia. Tramal generative hyphae thin to thick walled, hyaline to dark brown, simple septate, rarely branched, 2.5–5 µm dia., skeletal hyphae, thick walled with narrow to wide lumen, brown, aseptate, unbranched, 3.9–4.4 µm dia. Setae and cystidioles absent. Basidia clavate with four sterigmata, 9–10.8 × 3–5.8 µm. Basidioles clavate, 9–10.2 × 3.3–5.8 µm. Basidiospores thick walled, subglobose to broadly ellipsoid, pale yellow to golden yellow in water, turning brown in KOH, smooth, (5.2–) 5.4–5.9 (–6.2) × (4.9–) 5.2–5.4 µm (n = 30/2), IKI−, CB−, Q = 1.15 (Q range 1.1–1.15).
Material examined: India, Tamil Nadu, Villupuram District, Nedimozhiyanur, 1208′ 24.45″ N 7833′ 38.05″ E on a dead wood of Prosopis juliflora, 14 September 2021, Elangovan Arumugam (MUBL4025, holotype).
Additional Material examined: India, Tamil Nadu, Villupuram District, Nedimozhiyanur, 1208′ 24.38″ N 7833′ 38.26″ E on a dead wood of Prosopis juliflora, 24 December 2023, Elangovan Arumugam (MEII2_2, Paratype).
GenBank numbers: ITS: OQ062657; LSU: OQ064102 (MUBL4025, holotype); ITS: PP827158; LSU: PP816284 (MEII2_2, Paratype)
Notes: The ML and Bayesian analyses revealed that F. subthailandicus formed a new linage that shares a sister clade with F. grenadensis (726% ML, 0.9 BYPP), F. thailandicus (56% ML, 0.54 BYPP), F. hainanensis (97% ML, 1.0 BYPP) (Fig. 139). Fulvifomes subthailandicus is similar to F. thailandicus by having a cracked and concentrically zonate pilear surface, dimitic hyphal system but the later differs with the presence of cystidioles, smaller pores (6–7/mm) and basidiospores ((5–) 5.2–6.2 (–6.4) × (4–) 4.8–5.1 µm) (Zhou 2015). Fulvifomes subthailandicus shares similar characteristics with F. grenadensis by having concentrically zonate pilear surface, homogeneous context, lacks cystidioles, and stratified tubes but the later differs by having smaller pores (4–7/mm) and basidiospores (5–6.4 × 4–5.1 µm) (Ryvarden 2004). Our Indian species shares similar morphological and microscopic characteristics with F. hainanensis by having velutinate pileus, dimitic hyphal system and absence of cystidioles and setae but the later differs by the presence of uncracked pilear surface, obtuse margin, duplex context, larger pores (3–4/mm) and slightly smaller spores (5–6.4 × 4–5.1 µm) (Zhou 2014). Our macro-microscopic illustrations were consistent with the phylogenetic analyses inferred from the combined dataset of ITS and nLSU of Fulvifomes spp.
Fulvifomes subthailandicus, F. azonatus, F. boninensis, F. costaricense, F. halophilus, F. imbricatus, F. jouzaii, F. imazekii, F. merrillii, F. submerrilli F. minutiporus, F. nakasoneae, F. nonggangensis, F. tubogeneratus, and F. xylocarpicola are strictly dimitic in nature, however, the former shows significant variations in phenetic and microscopic features with others (Dai 2010; Hattori et al. 2014; Liu et al. 2020; Wu et al. 2022; Zheng et al. 2021). Fulvifomes subthailandicus differs with F. caligoporus, F. cedrelae, F. coffeatoporus, F. dracaenicola, F. elaeodendri, F. floridanus, F. popoffii, F. robinae, F. siamensis, F. squamosus and F. subindicus in having dimitic hyphal system whereas the others are having mono-dimitic hyphal system (Du et al. 2021; Hattori et al. 2014; Salvador-Montoya et al. 2018, 2022; Tchoumi et al. 2020; Wu et al. 2022).
Sanghuangporus Sheng H. Wu, L.W. Zhou & Y.C. Dai, in Zhou, Vlasák, Decock, Assefa, et al., Fungal Diversity 77: 340 (2015)
Notes: Zhou et al. (2015) studied the morphological and phylogenetic data of the Inonotus linteus complex and introduced two new genera, Sanghuangporus and Tropicoporus. Sanghuangporus refers to the popular Chinese traditional medicinal fungus “Sanghuang” (Zhu et al. 2019). Sanghuangporus includes species with perennial, resupinate, effused-reflexed to pileate basidiomata, homogeneous to duplex context, poroid hymenial layer, with round to angular pores. Microscopically monomitic to mono-dimitic hyphal system, presence or absence of cystidioles, presence of hymenial setae, absence of hyphoid setae and ellipsoid, broadly ellipsoid to subglobose, yellowish, slightly thick- to thick-walled, smooth basidiospores (Zhou et al. 2015). Mycobank reported 20 related species in the genus Sanghuangporus, 19 species were documented in Index Fungorum and Nucleotides from 16 different species were recorded in GenBank (as on 9 May 2024) (Fig. 142).
[See PDF for image]
Fig. 142
phylogram generated from Bayesian analysis based on combined ITS and LSU rDNA sequence data of Sanghuangporus spp. The combined analyses include 29 taxa, which comprised 1908 characters (LSU: 933, ITS: 966) after alignment. The new species sequences used in the study are in bold. The tree is rooted with Inocutis tamaricis (CBS 384.72) and Inonotus compositus (Wang 552). Bootstrap support values for Maximum Likelihood (ML) ≥ 65% and Bayesian posterior probability (BYPP) ≥ 0.95 are given above the nodes. Type species are bold and newly generated taxa in green
Sanghuangporus subzonatus S. Gunaseelan, M. Kaliyaperumal & K. Kezo sp. nov.
Index Fungorum number; IF 902414; Facesoffungi number: FOF13990. Figure 143
[See PDF for image]
Fig. 143
Sanghuangporus subzonatus (MUBL4023, holotype) Morphological characteristics of Sanghuangporus subzonatus a Basidiomata (holotype). b Pore surface. c Cross-section of basidiome showing obtuse margin. d Stratified tube layer. e hyphae from context. f Tramal hyphae. g Cystidioles. h Hymenia setae. i Basidia. j Basidioles. k–o Basidiospores: k Basidiospore. l Basidiopore in water. m Basidiopore in KOH. n Basidiopore in cotton blue. o Basidiopore in Melzer’s reagent. Scale bar: e–o = 5 μm
Etymology: The species epithet “subzonatus”, signifies the close phylogenetic relationship with the species Sanghuangporus zonatus
Holotype: MUBL4023
Basidiomata perennial, solitary, pileate, sessile, without odor or taste, hard corky to woody hard when fresh, become woody hard when dry. Pileus dimidiate, applanate, with no distinct crust, projecting up to 3.5 cm to 6.3 cm, 6.7 cm wide and 4.7 cm thick near attachment. Pileal surface brown (6F4), yellowish brown (5E7, 5E8) to dark brown (6F8) when young, infrequently tuberculae, concentrically zonate, on maturation pilei becoming brownish grey (5F2), narrowly sulcate with few cracks near attachment. Margin distinct, sterile, obtuse, yellowish brown (5D6) to golden brown (5D7). Pore surface yellowish brown (5D6) to golden brown (5D7); Pores round to angular, 6–9/mm; Dissepiments entire, thin. Context homogenous, up to 2.7 cm thick, yellowish brown (5D8) to brown (6D7). Tube layer golden brown (5D7), stratified, tubes upto 0.2 cm long.
Hyphal system monomitic in the context and dimitic in trama; generative hyphae simple septate; tissue darkening in KOH without hyphal swelling; Context Generative hyphae, thin to thick walled, hyaline to golden brown, simple septate, occasionally branched, 2–5.2 μm diam., Trama Generative hyphae, dominant, thin to thick-walled, hyaline to yellowish, septate, occasionally branched, 2–5 μm dia.; Skeletal hyphae, thick walled with a distinct narrow to wide lumen, yellow to yellowish brown, aseptate, unbranched, 2–5 μm dia. Hymenial setae thick-walled, dark brown, ventricose to subulate, with a sharp or obtuse tip, occasionally with lateral appendages, 11.2–39 × 3–10 μm. Cystidioles hyaline, thin walled, ventricose to fusoid with elongated tapering apical portion, or abruptly narrow apices, 5–27.5 × 3–6 μm. Basidia clavate to sub clavate, 5.5–11 × 4–5.5 μm, with four sterigmata and a simple septum at the base. Basidioles clavate, 4–11.5 × 3.5–5.2 μm. Basidiospores smooth, broadly ellipsoid to subglobose, juvenile spores thin to fairly thick-walled, hyaline to yellowish, CB+, IKI ̄.; matured spores slightly thick walled to thick walled, pale yellow to golden yellow in water, turning golden yellow to brown in KOH, CB ̄, IKI ̄, (3.9–) 4.5–4.8 (–4.9) × (3–) 3.2–3.6 (–3.8) μm (n = 50/2), Q = 1.35 (Q range 1.02–1.5). Chlamydospores, smooth, irregular, globose to subglobose, golden brown in water, turning to rust brown in KOH, 5.5–9 × 5–7.5 μm, CB(+) to CB ̄, IKI ̄.
Material examined: India, Tamil Nadu, Salem District, Kolli Hills, 11° 17′ 42.60″ N 78° 23′ 07.45″ E, on dead wood, 03 February 2018, Sugantha Gunaseelan (MUBL4023, holotype).
Additional Material examined: India, Tamil Nadu, Salem District, Kolli Hills, 11° 17′ 43.61″ N 78° 25′ 08.52″ E, on dead wood, 03 February 2018, Malarvizhi Kaliyaperumal (MSK-II-AS04, Paratype).
GenBank numbers: ITS: OP363954; LSU: OP379563 (MUBL4023, holotype); ITS: OQ225689; LSU: OQ225691 (MSK-II-AS04, Paratype).
Notes: Morpho-microscopical analyses and phylogenetic studies of the combined ITS and LSU dataset show that S. subzonatus fits well within the Sanghuangporus genus concept (Fig. 142) (Zhou et al. 2015). Sanghuangporus zonatus and S. subzonatus share similar morpho-microscopical characteristic features such as concentrically zonate pileal surface, basidiomes more or less cracked with age, mono-dimitic hyphal system. However, S. subzonatus significantly varies from S. zonatus, by having obtuse margin, presence of cystidioles and larger basidiospores (3.9–4.9 × 3–3.8 μm) (Tian et al. 2013). The present phylogenetic study indicated that S. subzonatus is closely related to S. zonatus with significant support (98% ML, 1.00 Bayesian posterior probability (BYPP)). Sanghuangporus vitexicola formed a well-resolved sister clade to S. subzonatus (98% ML, 1.00 BYPP), the former has dimitic hyphal system with radially rimose and cracked basidiome, acute margin whereas S. subzonatus is having mono-dimitic hyphal system with few cracks in near attachment (Wu et al. 2020) (Fig. 144).
[See PDF for image]
Fig. 144
Maximum likelihood phylogenetic tree ML obtained from the DNA sequence data of the ITS and LSU data sets. Bootstrap values of maximum likelihood ML in the left, MP in the middle, equal to or greater than 70%, and Bayesian posterior probabilities PP in the right, equal to or greater than 0.95, are indicated above or below the nodes as MLBS/MPBS/PP. The tree is rooted with Tomophagus colossus TC‐02. A known species obtained in this study is indicated in bold
Morpho-microscopical characteristic of S. subzonatus, varies from S. australianus, and S. lagerstroemiae (Wu et al. 2022) in resupinate basidiomata and dimitic hyphal system. Sanghuangporus lonicerinus, S. pilatii, S. sanghuang, S. toxicodendri, S. weigelae, and S. weirianus differs from S. subzonatus in having dimitic hyphal system (Gilbertson 1979; Tomšovsky 2015; Wu et al. 2012).
Sanghuangporus subbaumii and S. subzonatus shows similar morphological traits such as concentrically zonate, narrowly sulcate basidiomata with few cracks near attachment on maturity, mono-dimitic hyphal system and presence of cystidioles, however the former differs by having acute margin, not stratified tube layer and larger pores (5–7/mm) (Shen et al. 2021). Though S. vaninii and S. subzonatus have similar features such as the presence of zonate context, smaller pores (6–8/mm) and mono-dimitic hyphal system; matted to glabrous basidiomes, absence of cystidioles and ovoid to broadly ellipsoid basidiospores (3.5–4.6 × 2.5–4 μm) of S. vaninii differs from S. subzonatus (Dai 2010).
Sanghuangporus ligneus and S. subzonatus are comparable in mono-dimitic hyphal system, presences of cytidioles and concentrically zonate pileus but former varies in having ungulate to triquetrous and radially cracked basidiomata with larger pores (4–5/mm) (Ghobad-Nejhad 2015). Sanghuangporus alpinus, S. microcystideus and S. subzonatus shares concentrically zonate pileus, obtuse margin, homogenous context and mono-dimitic hyphal system and presence of cystidioles, however in comparison with S. subzonatus, S. alpinus is having larger pores (5–7/mm) and broadly ellipsoid to ellipsoid smaller basidiospores (3–4 × 2.5–3.3 μm) whereas S. microcystideus differs in distinctly cracked pilear and larger basidiospores (5.1–7 × 4.1–5.1 μm) (Tian et al. 2013; Zhou et al. 2016a, b). Sanghuangporus baumii, S. quercicola and S. lonicericola are consistent with S. subzonatus in concentrically zonate, sulcate basidiomes with slightly rimose with age, obtuse margin, smaller pores, presence of cystidioles and mono-dimitic hyphal system, but our Indian species significantly differs in shape and size of basidiospores (Dai 2010; Zhu et al. 2017; Wu et al. 2020).
Polyporales Gäum., Vergl. Morph. Pilze (Jena): 503 (1926)
Notes: Polyporales was proposed by Gäumann (1926) with ten families based on their morphological characteristics (Liu et al. 2023a, b; Wang et al. 2023a, b). It is one of the most intensively studied clades of fungi that includes a variety of basidiocarp types ranging from bracket-shaped (e.g., Ganoderma, Trametes), effused to resupinate (Wolfiporia, Phlebia) or stipitate (Microporus, Polyporus), and flabellate clusters of overlapping basidiocarps (Laetiporus, Sparassis). Hymenophore differs from daedaloid to hydnoid (Daedaliopsis, Flavodon), and poroid to lamellate (Coriolopsis, Lentinus) or smooth (Podoscypha) (Binder et al. 2005). Hyphal system includes monomitic, dimitic, and trimitic (Gilbertson and Ryvarden 1986) and basidiospores shape ranging from globose, sub-globose, ellipsoid, and oblong to cylindrical. In recent years, phylogenetic studies from multigene data sets have resolved 40 valid legitimate families of Polyporales (Binder et al. 2013; Justo et al. 2017). MycoBank has 81 family records as of 12th May 2024.
Ganodermataceae (Donk) Donk, Bull. bot. Gdns Buitenz. 17(4): 474 (1948)
Notes: Ganodermataceae was first introduced by Donk (1984), and is one of the main families of Polyporales, which was typified by Ganoderma P. Karst. (1881). Ganodermataceae is a white rot fungus; it is one of a large and complex family, and its family systematics have been reported over the past 100 years. Even though the proposed genus Ganodermataceae was established in 1948, this family was accepted in the Dictionary of Fungi, published in 1971 (Ainsworth and Bisty 1987). Index Fungoum reported that the Ganodermataceae were verified to be Polyporaceae. Based on morphological and molecular data, 14 genera, including Amauroderma, Amaurodermellus, Cristataspora, Foraminispora, Furtadoella, Ganoderma, Haddowia, Humphreya, Magoderna, Neoganoderma, Sanguinoderma, Sinoganoderma, Tomophagus, and Trachydermella, were included in the Ganodermataceae. This family is distributed in tropical and temperate regions. A member of this family has distinctive double-walled basidiospores with ornamented endospore walls, which are the characteristic features of Ganodermataceae. In addition, some species in the family are proposed as medicinal and economically important (Moncalvo and Ryvarden 1997; Ryvarden 2004; Hapuarachchi et al. 2019).
Ganoderma P. Karst., Revue mycol., Toulouse 3(9): 17 (1881)
Notes: Ganoderma is saprobic fungi, which are distinctive laccate and non-laccate basidiocarps (Cao et al. 2012; Wang et al. 2012). Ganoderma was established by Karsten (Karsten 1881), with Polyporus lucidus (Curtis) Fr. (= Ganoderma lucidum (Curtis) P. Karst.) as the type species (Richter et al. 2015). Some Ganoderma mushrooms have been used as traditional medicine for pharmaceutical purposes (Richter et al. 2015). Species of Ganoderma are distinctive by their sessile or stipitate, annual, dimidiate, flabelliform shape, laccate or non-laccate (dull), basidiomes, and double wall basidiospore (Cao et al. 2012; Luangharn et al. 2021). Ganoderma is widely distributed in temperate and tropical regions (Ryvarden 2000), and most Ganoderma is pathogenic and causes white rot fungi in various plant diseases (Dai et al. 2007) (Fig. 144).
Ganoderma gibbosum (Blume and T. Nees) Pat. (1897)
Index Fungorum number: 250058; Facesoffungi number: FoF 05652; Fig. 145.
[See PDF for image]
Fig. 145
Ganoderma gibbosum (MFLU23-0017, new host record). a, b Mature basidiocarps on host. c Underbasidiocarp. d Pore characteristics. e Tube layer hyphae in Melzer’s reagent. f–h Contaxt hyphae. i, j Basidiospore. Scale bars: a–c = 5 μm, d = 500 μm, f = 10 μm, g–h = 20 μm, i–j = 5 μm
≡ Polyporus gibbosus (Blume and T. Nees)., Nov. Act. Academiae Caesareae Leo‐poldino Carolinae Germanicae Naturae Curiosorum. 13: 19, 4(1–4) (1826)
≡ Fomes amboinensis var. gibbosus (Blume and T. Nees) Cooke, Grevillea. 13(68): 118 (1885)
≡ Fomes gibbosus (Blume and T. Nees) Sacc. Syll. Fung. 6: 156 (1888)
≡ Scindalma gibbosum (Blume and T. Nees) Kuntze., Revisio generum plantarum 3(2): 518 (1898)
Basidiocarps annual or perennial, sessile. Pileus 7–13 cm in length, 4–7 cm in width, and 0.8–1.0 cm thick, usually thick at the base. Pileus shape convex, imbricate, umbonate, uneven, ungulate, somewhat round, and plump when young, broadly attached to its host. Pileus surface non‐laccate (dull), furrowed, sulcate, undulating on the upper surface, somewhat spathulate to uneven, an irregularly ruptured crust overlying the surface, slightly soft at the margin, incised, woody, and cracked when old to older. Pileus colour is usually pale orange (5A3), light orange (5A4), and orange (5A5-7) when young to mature stage, homogenous with pale yellow (1A3), grayish-yellow (1B4–1B6), greyish-green (1C5-8) at the base when developing to mature, extending to greyish-orange (5B3–5B4), yellowish-brown (5D8), brownish-orange (6C4), reddish closed to margin (7A2), and white at the margin of mature fruiting bodies. Context up to 0.4–1.8 cm thick, often reddish brown (8E6-7) to dark brown (8F5-8), compact and hard, tough when dried. Hymenophore reddish-brown (8E7) with shading dark brown (7F6-7). Tube layers 0.3–1.1 cm in length with brown (7D8) to dark brown (7F7). Margin often white (8A1), wavy, and slippery when fresh close to the underside of basidiomes. Pore 4–7 in number per mm, with subcircular to circular. Pore surface usually white (6A1) and turn reddish-brown (8E4-6) when scratched or touched. Hyphal system trimitic hyphal, hyaline, with abundant thick-walled, with clamp connections, hyaline, with walls varying in thickness with simple septa, composed of narrow and sparingly branched; generative hyphae 1.2–3.8 µm broad (n = 50), often brownish-orange (6C7) in Melzer’s reagent, thin-walled, hyaline; skeletal hyphae 3.5–4.9 µm broad (n = 50), brown (7D8), abundant thick-walled, dextrinoid, mostly hyaline; binding hyphal 2.3–3.9 µm width (n = 50), reddish-brown (8E7) in Melzer’s reagent, usually thick-walled, hymenial, branched, mostly dark brown (7F7) near the layer tubes.
Basidiospores ellipsoid to broadly ellipsoid, with some oblong with double walls (4.6-)6.8–8.1–9.1(-10.6) × (5.3-)5.9–6.2–7.4(-8.1) µm (x̅ = 8.2 × 6.3 µm, n = 50) µm, with Q = 1.49–1.57, L = 8.24 µm, W = 6.27 µm (including myxosporium), (4.8-)5.2–6.1–6.9(-8.2) × (4.2-)4.9–5.6–5.9(-6.7) µm (x̅ = 6.4 × 5.7 µm, n = 50) µm, with Q = 1.08–1.14, L = 6.24 µm, W = 5.67 µm (including myxosporium), overlaid by hyaline, dextrinoid, echinulae, echinulate brown inner wall, light yellow (4A4–4A5) to grayish-yellow (4B5–4B6) in 5% KOH. Basidia not seen.
Material examined: Thailand, Chiang Rai Province, Muang District, Thasud, 20°2′50′′N, 99°53′44′′E, 20 m, 25 May 2022, T. Luangharn, LT2022-096, MFLU23-0017.
Host and habitat: solitary on the dead stump of a Bauhinia variegate tree.
GenBank numbers: ITS: OQ383906; LSU: OQ383907.
Notes: Ganoderma gibbosum was initially documented in Java; however, the original specimen has been regrettably misplaced (Moncalvo and Ryvarden 1997). This species has been synonymous with non-laccate G. applanatum, which is classified in the G. applanatum–australe complex (Moncalvo and Ryvarden 1997). Ganoderma gibbosum is widely distributed in both tropical and temperate regions. Species of this genus can be pathogens or saprobes on a wide range of hosts (Luangharn et al. 2020, 2021). Ganoderma gibbosum is characterized by its non-laccate basidiomes, annual or perennial, sessile, furrowed and sulcated to undulating, and ellipsoids with double-walled basidiospores. Luangharn et al. (2020) reported the Chinese collections of G. gibbosum and host preferences from Kunming, China. Phylogenetic analysis (Fig. 143) and macro–micro morphological characteristics examined largely overlap with those of G. gibbosum reported by Luangharn et al. (2020). In this study, we illustrate the macro- and micro-characteristics and molecular evidence of G. gibbosum, through its new host record of Bauhinia variegate from Thailand.
Phanerochaetaceae Jülich, Biblthca Mycol. 85: 384 (1982) (1981)
Note: Phanerochaetaceae was raised by Jülich (1982). Earlier the term ‘phanerochaetoid fungi’ was broadly used in two ways; ‘corticiocentric’ where the term was strictly associated with members possessing clampless hyphae (Parmasto 1986; Wu 1990) and ‘phanerochaetoid’ where the members were broad classification of aphyllophoroid fungi (Hansen and Knudsen 1997). The family has gained its prominence only in the recent years with members such as; Bjerkandera, Phaeophlebiopsis, Phanerochaete, Phanerodontia, Phanerina, Phlebiopsis, Rhizochaete, Oxychaete (Larsson 2007; Binder et al. 2013; Justo et al. 2017; Zhao et al. 2020a, b). Members of this family are characterized by their resupinate to effused-reflexed basidiomata, monomitic to rarely dimitic hyphal system, hyaline cystidia and encrusted, usually narrowly clavate basidia, and hyaline, thin-walled, smooth, cylindrical to ellipsoid basidiospores. At present Mycobank has 44 associated records as of 10th Dec. 2022 (Fig. 146).
[See PDF for image]
Fig. 146
Phylogram generated from maximum likelihood analysis based on combined LSU and ITS sequence dataset of Phlebiopsis spp. Based on MegaBLAST search of NCBIs GenBank nucleotide database, the closest hits using the ITS sequence had highest similarity to P. griseofucescens (strain Cui 12629, GenBank MT561718; Identities = 567/582 (97%), no gaps) and P. lacerate (strain SWFC00003747, GenBank MT180949; Identities = 567/582 (97%), no gaps). The closest hit using the LSU sequence had the highest similarity to P. lacerate (synonym to P. griseofucescens) (strain SWFC00003705 GenBank MT180951; Identities = 1233/1272 (97%); Gaps = 21/1272(1%)). Thirty-three sequences were included in the combined analyses which comprised 2071 characters (1381 characters for LSU, 689 characters for ITS) alignment. Bjerkandera adusta (HHB-12826-Sp) was used as the outgroup. The maximum likelihood tree was performed using MEGA X (Kumar et al 2018) and the same data were used for Bayesian analysis using MrBayes v. 3.2.7 (Ronquist et al 2012). Branches are labelled with Bayesian posterior probabilities more than 0.7 and maximum likelihood bootstrap higher than 60%. The type species are in bold and new taxa in blue
Phlebiopsis Jülich, Persoonia 10 (1): 137 (1978)
Note: Phlebiopsis was raised by Jülich (1978). Member of this genera are characterized by their resupinate to effused-reflexed basidiomata which are membranaceous to subceraceous when fresh that crack upon drying. The hymenophore may vary from smooth to poroid to odontoid pores. Hyphal system mostly monomitic and rarely dimitic with simple-septate generative hyphae. Presence of encrusted cystidia, clavate basidia and basidioles. Hyaline thin-walled, smooth, cylindrical to ellipsoid, acyanophilous and negative in Melzer’s reagent (Jülich 1978; Bernicchia and Gorjón 2010). An in-depth study of phylogeny by Zhao et al. (2021) revealed 24 lineages resolved in the ITS phylogenetic tree. At present MycoBank has recorded 36 species as of 10 May 2024.
Phlebiopsis subgriseofuscescens K. Kezo, M. Kaliyaperumal & S. Gunaseelan sp. nov.
Index Fungorum number: IF 902415; Facesoffungi number: FOF13989; Fig. 147
[See PDF for image]
Fig. 147
Phlebiopsis subgriseofuscescens (MUBL4024, holotype) Microscopic structures of Phlebiopsis subgriseofuscescens (drawn from the holotype). a Basidiomata. b1 Duplex context. b2 Tube layer. c Encrusted cystidia. d Encrusted cystidia and basidia. e Thin and thick-walled generative hyphae. f Encrusted cystidia. g Basidioles. h Basidiospores. i Basidiospores phloxine. j Water. k Cotton blue. l Melzer. Scale bars: b = 1 mm, c–l = 10 μm
Etymology: The species epithet “subgriseofuscescens”, signifies the close relation with the species Phlebiopsis griseofuscescens.
Holotype: MUBL4024
Description: Basidiomata annual, resupinate to effuse reflexed, subceraceous when fresh, turning leathery on drying, up to 20 cm long, 7 cm wide, 3 mm thick. Pilei laterally fused, pileal surface tomentose, white (5A1) when fresh turning orange white (5A2) on drying. Context duplex, greyish orange (5B3) towards tomentum and greyish drown (7D3) towards tube, up to 1 mm in thickness. Hymenial surface brownish violet (11D7) when fresh turning greyish brown (7D3) on drying. Pores hydnoid, projecting up to 1 mm, and angular pores towards margin, 1–2 per mm. Margin white (5A1) when fresh becoming creamy to orange white (5A2) on drying.
Hyphal system monomitic; context generative hyphae, thin to thick walled, hyaline, simple-septate, branched, 2.5–5 μm dia., CB ̄, IKI ̄. Trama generative hyphae, thin to thick walled, hyaline, simple-septate, branched, 2.5–6 μm dia., CB ̄, IKI ̄. Cystidia clavate, thick-walled, coarsely encrusted, projecting above the hymenium, 48–83 × 6–7.2 μm. Cystidioles absent. Basidioles narrowly clavate, 14–20 × 3.5–5.5 μm. Basidia narrowly clavate, with a simple septum at the base and four sterigmata, 16–24.2 × 3.5–6.2 μm. Basidiospores smooth, hyaline, thin-walled, allantoid to cylindrical, (6–)6.5–7.1(–7.5) × (2.2–)2.4–2.6(–3.1) μm (n = 30/2), Q = 2.7; CB ̄, IKI ̄.
Materials examined: India, Tamil Nadu, Thiruvannamalai District, 12° 29′ 14.4″ N 78° 55′ 06.0″ E on dead angiosperm wood, 14th Nov. 2019, Kezhocuyi Kezo (MUBL4024, holotype).
Additional materials examined. India, Tamil Nadu, Salem District, 11° 47′ 40.1″ N 78° 12′ 44.9″ E, on dead angiosperm wood, 23rd Jan. 2018, Kezhocuyi Kezo (Isotype KSM-HRS31), same collection site (Isotype KSM-HRS30, KSM-HRS9). India, Tamil Nadu, Thiruvannamalai district, 12° 29′ 13.5″ N 78° 55′ 07.0″ E on dead angiosperm wood, 27th Mar. 2018, Kezhocuyi Kezo (Isotype KSM-PK6). INDIA, Tamil Nadu, Trichy District, 11° 17′ 21.4″ N 78° 36′ 12.3″ E on dead angiosperm wood, 14th Feb. 2018, Kezhocuyi Kezo (Isotype KSM-KK15).
GenBank numbers: MUBL4024; ITS: OP101632; LSU: OP379562, KSM-HRS31; ITS: OP101633, KSM-KK15; ITS: OP379520
Notes: Phlebiopsis subgriseofuscescens is characterised by their resupinate to effuse reflexed basidiocarp, brownish violet, hydnoid hymenophore and cylindrical spore. Our Indian species shares similar characters with P. griseofuscescens except for the spore size, as the latter has smaller ellipsoid spores ((4–5.5)–6 × 2.5– (3–3.5) μm) (Xu et al. 2020; Chen et al. 2021; Zhao et al. 2021). Phlebiopsis dregeana differs from our Indian species having smooth to glabrous hymenophore and larger ellipsoid spores (6.5–8 × 4–5 μm) (Hjortstam and Ryvarden 1990; Zhao et al. 2021). Our observations on morpho-microscopic illustrations are consistent with phylogenetic analyses. ML and Bayesian analyses inferred from combined dataset of ITS and nLSU of Phlebiopsis spp. revealed that our Indian species formed a new lineage to P. griseofuscescens (0.99 BYPP /100% ML) and P. dregeana (1 BYPP/100% ML) clade (Fig. 2). Our strain shows variation in hymenophore and basidiospore size and shape with P. cylindrospora (smooth hymenophore and cylindrical spore ((5–)5.5–7.5(–8) × 1.8–2.8 (–3)).
[See PDF for image]
Fig. 148
Phylogenetic tree of selected Vararia species and related taxa sampled from Peniophoraceae (green box), based on the Bayesian analyses of the ITS and LSU gene regions. Taxon sampling was optimized by selecting the type isolates (marked with an asterisk) as well as sampling the generic types. Sequences were analysed using MrBayes v. 3.2.7a (Ronquist et al. 2012). Bayesian analyses were run at the CIPRES Science Gateway (Miller et al. 2010), applying the GTR substitution model with invariable sites and gamma-distribution rate variation. Peniophora rufa was used as an outgroup. Bayesian posterior probabilities (BYPP) are given close to the nodes. GenBank accession numbers are indicated (ITS/LSU)
Russulales Kreisel ex P. M. Kirk, P. F. Cannon & J. C. David, in Kirk et al., Ainsworth & Bisby's Dictionary of the Fungi, Edn 9 (Wallingford): xi (2001)
Peniophoraceae Lotsy, Vortr. bot. Stammesgesch. 1: 687 (1907)
Vararia P. Karst., Kritisk Öfversigt af Finlands Basidsvampar, Tillägg 3: 32 (1898)(Fig. 148)
Vararia tenuata Ghobad-Nejhad, sp. nov.
Index Fungorum number: IF901609; Facesoffungi number: FoF 15298 Fig. 149
[See PDF for image]
Fig. 149
Vararia tenuata (RAMK, holotype). a Habitat. b Basidiomata. c Squashed hyphae in Melzer’s reagent. d, e Basidia. f Gloeocystidium. g Dichophyses. h, i Basidiospore in KOH (h) and cotton blue (i). Scale bars: b = 2 cm, c = 50 µm, d–g = 10 µm, h, i = 5 µm
Etymology: Referring to the thin basidiomata of the new species.
Holotype: RAMK. Isotype: KAS, ICH.
Basidiomata annual, resupinate, thin, ceraceous, tightly adnate, confluent, hymenial surface cream, uniform, smooth, margin determinate to slightly thinning out. Hyphal system dimitic with only few binding hyphae, tissues generally strongly colored in CB, generative hyphae with simple septae, colorless, thin-walled, tightly interwoven, 2–3 μm diam., binding hyphae (dichohyphae) infrequent, IKI− to moderately dextrinoid, 2–4 μm diam. Dendrophyses abundant, bush-like at apex, strongly dextrinoid, upper branches narrow and thin-walled, 0.5–1 μm wide, stem 2–3 μm wide. Dendrohyphidia numerous, little to moderately branched, width uneven, in general 2–2.5 μm wide at upper branches and 3–6 μm wide in lower parts. Gloeocystidia numerous in subhymenium, irregular in size and shape, with refractive contents, subcylindrical to clavate with pleural base, 28–40(–60) × 7–10 μm, walls thickened up to 2–2.5 μm. Basidia subcylindrical to utriform, with four stout sterigmata, without clamp connection, 37–55 × 7.5–10 μm (mature basidia, excluding sterigmata), walls thickened. Basidiospores fusiform, more or less bent near apiculus, with a small, distinct apiculus, colorless, thin-walled, smooth, with irregular guttules, IKI−, CB−, 13.5–15 × 5–6 μm.
Material examined: Thailand, Chiang Mai, Mae Rim District, Mae Raem, Queen Sirikit Botanical Garden, in dry deciduous forest with Dipterocarpus obtusifolius, D. tuberculatus, Shorea obtusa, and Hopea odorata, on fallen branches of Bambusa Schreb. (Poaceae), 15 January 2012, Ghobad-Nejhad 3018 (RAMK holotype, isotypes KAS, ICH).
GenBank numbers: ITS: OQ612724, LSU: OQ607809
Notes: Vararia tenuata is characterized by thin, light cream, tightly adnate basidiomata, dextrinoid, bush-like dendrophyses, and fusiform basidiospores. It resembles V. breviphysa with regard to its basidiomata texture and shape of dendrophyses and basidiospores. The two species also cluster closely in the phylogenetic tree. However, V. tenuata differs from V. breviphysa by the color of basidiomata, being isabelle grey to beige-cinnamon in the latter, and smaller basidiospores [(14–)16–20 × 4.8–6 µm in V. breviphysa; Boidin & Lanquetin 1975)]. The two species show only 93% similarity in Blast searches.
Based on a megablast search of NCBI’s GenBank nucleotide database (as of March 2023), the closest hits using the ITS sequence had the highest similarity to Vararia breviphysa (strain CBS 644.81 from Type material, GenBank NR_175095.1; identities = 650/701 (93%), gaps 27/701 (3%)), V. breviphysa (strain CBS 643.81, GenBank MH861395.1; identities = 641/697 (92%), gaps 21/697 (3%)), and V. pirispora (strain CBS 721.86, GenBank MH862017.1; identities = 603/659 (92%), gaps 27/659 (4%)). The closest hits using the LSU sequence were V. pirispora (strain CBS 721.86, GenBank MH873707.1; identities = 891/906 (98%), gaps 4/906 (0%)), V. breviphysa (strain CBS 644.81 from Type material, GenBank NG_078651.1; identities = 867/876 (99%), gaps 0/876 (0%)), and V. breviphysa (strain CBS 643.81, GenBank MH873144.1; identities = 867/877(99%), gaps 1/877(0%)).
Pucciniomycetes R. Bauer, Begerow, J.P. Samp., M. Weiss & Oberw., Mycol. Progr. 5(1): 48 (2006)
Pucciniales Caruel, Atti R. Accad. Naz. Lincei, Mem. Cl. Sci. Fis. Matem. Nat., sér. 5: 246 (1881)
Gymnosporangiaceae Chevall., Fl. gén. env. Paris (Paris) 1: 422 (1826)
Notes: This family was first introduced by Chevall (1826), with the type genus Gymnosporangium R. Hedw. ex DC. The recent treatment is by Zhao et al. (2020a, b) and further amended by Aime and McTaggart (2021). Currently, the genera Gymnosporangium and Gymnotelium are accommodated in the family.
Gymnosporangium R. Hedw. ex DC., in Lamarck & de Candolle, Fl. franç., Edn 3 (Paris) 2: 216 (1805)
Notes: Gymnosporangium was introduced by Lamarck and de Candolle (1805) with Gymnosporangium fuscum on Juniperus sabina as the type species (Zhao et al. 2020a, b). Species of Gymnosporangium are generally heteroecious and demicyclic, producing the telial stage on plants in Cupressaceae and the spermogonial and aecial stages on plants in Rosaceae (Yun et al. 2009; Shen et al. 2018; Aime and McTaggart 2021). Currently, 153 epithets are listed for Gymnosporangium in Index Fungorum (2024). Recent studies suggest there are around 70 known Gymnosporangium species worldwide (Shen et al. 2018; Zhao et al. 2020a, b; Tao et al. 2020; Wang et al. 2022a).
Gymnosporangium paraphysatum Vienn. -Bourg., Revue de Mycologie (Paris) 25 (5): 304 (1961)
Index Fungorum number: IF331602; Facesoffungi number: FOF Fig. 150
[See PDF for image]
Fig. 150
Gymnosporangium paraphysatum (TNM F0030451). a Collection of the specimen on Calocedrus formosana in the field. b, c Uredinia and urediniospores on the leaves. d Urediniospores. e Paraphyses. Scale bars: b = 2 mm, c = 0.5 mm, d, e = 20 µm
Holotype: BPI 844997.
Parasitic on leaves of Calocedrus formosana. Sexual morph: Not observed. Asexual morph:Uredinia foliicolous, amphigenous, solitary and scattered, 0.2–0.6 mm across, subepidermal, becoming erumpent, surrounded by discolored margins. Paraphyses numerous, intermixed, stalks hyaline, filiform, 117–172 μm long, 2.5–5 μm wide, apex brown, clavate, 38–58 × 16–25 μm. Urediniospores orange to yellowish, broadly ellipsoidal to oval, 28–43 × 16–22 μm, tapered towards the apex, wall hyaline, 1.2–2.5 μm, thickened at the apex, 2–8 μm.
Material examined: Taipei, Taichung City, Dongshi District, Tungshih Forestry Station, on leaves of Calocedrus formosana, 21 Apr 2016, Yuan-Min Shen (TNM F0030451)
GenBank numbers: SSU-ITS-LSU: PP923925.
Notes: The rust produces urediniospores and teliospores on leaves of Calocedrus but has no known spermogonial and aecial stages. It is characterized by the presence of numerous conspicuous paraphyses (Kern 1973). The type specimen of this species was collected from the host plant Calocedrus macrolepis (Libocedrus macrolepis) in Vietnam. Although Hiratsuka & Chen (1991) listed the fungus on C. formosana, no characterization of the rust on this host had been made. This note provides the first description of G. paraphysatum on C. formosana and represents the first confirmed report of G. paraphysatum. New sequence data have been generated based on our collection. The species may be closely related to G. libocedri due to the phylogenetic relationship and their ability to parasitize the plant genus Calocedrus (Kern 1973; Kasuya et al. 2024).
Ustilaginomycotina Doweld
Notes: We follow the latest treatment and updated account of Ustilaginomycotina in Begerow and McTaggart (2018).
Ustilaginomycetes Warm.
Notes: The classification of the orders in Ustilaginomycetes follows He et al. (2024).
Ustilaginales Bek.
Notes: There are seven families in this order: Anthracoideaceae, Clintamraceae, Ustilaginaceae, Websdaneaceae, Geminaginaceae, Pericladiaceae, and Melanotaeniaceae (He et al. 2024).
Ustilaginaceae Tul. & C. Tul.
Notes: The family Ustilaginaceae was introduced by Tulasne and Tulasne (1847). Twenty-six genera are currently recognized in this family (He et al. 2024).
Moesziomyces Vánky
Notes: The species of Moesziomyces are parasites on grasses but anamorphic stages were isolated as yeasts from plant surfaces. The sori of the parasitic species are in ovaries, covered by green later brown peridium of host tissue. The spores are arranged in many-spored balls, firmly agglutinated and mixed with sterile cells (Vánky 2011; He et al. 2024). During an examination of grasses in the herbarium of the Botanic Garden and Botanical Museum Berlin (B), a specimen of Echinochloa oryzoides from Greece, infected by Moesziomyces bullatus, was found. It is reported here as a new record for Greece (Fig. 151).
[See PDF for image]
Fig. 151
Most likely tree generated using maximum likelihood analysis (RAxML-NG, Kozlov et al. 2019) implemented in RAxMLGUI (Edler et al. 2021) based on a MAFFT v7.450 (Katoh and Standley 2013) alignment of ITS. The tree is rooted with Triodiomyces altilis (Syd.) McTaggart & R.G. Shivas and Ustilago echinata J. Schröt. Values at nodes indicate bootstrap values inferred by 1000 replicates; only values ≥ 60% are shown
Moesziomyces bullatus (J. Schröt.) Vánky, Bot. Notiser 130(2): 133 (1977)
Index Fungorum number: IF 317784; Facesoffungi number: FoF 15929, Fig. 152
[See PDF for image]
Fig. 152
Moesziomyces bullatus on Echinochloa oryzoides (B s.n.). a Habit (sori in single ovaries). b Spore ball in LM. c Spores attached together by remnants of broken sterile cells. Scale bars: a = 1 cm, b = 20 μm, c = 10 μm
Parasitic on Echinochloa oryzoides. Infection local. Sori in single ovaries of infected plant, visible between spreading glumes, subglobose, broadly ellipsoidal or ovoid, 3–5 × 2–4 mm, covered by a thick greenish brown to yellowish brown peridium that later ruptures exposing a powdery, very dark brown mass of spore balls. Spore balls variable in shape and size, broadly ellipsoidal, subglobose, ovoid or globose, sometimes irregular, (55–)80–240(–290) × (40–)60–160(–200) μm, very dark reddish brown, opaque, composed of hundreds, rather firmly united spores and sterile cells. Spores irregularly polygonal, subglobose, ovoid or broadly ellipsoidal, (6.5–)7–10(–11) × 6.5–8(–9) (8.7 ± 0.8 × 7.4 ± 0.6) μm (n = 100), light yellow–brown, attached together by cylindrical protuberances, 1.5–2.5 μm wide, 1–2 μm high; spore wall 0.5–0.8 μm thick, with irregular meshes and wings, representing remnants of broken sterile cells.
Material examined: GREECE, Central Macedonia Region, Thessaloniki, in a rice field near Axion, 40°30′N 22°34′E, 10 msl, on Echinochloa oryzoides (Ard.) Fritsch (Poaceae), 8 October 1992, Th. Raus & Ch. Schiers, s.n. (B s.n.).
GenBank numbers: ITS: PP860524.
Notes: Moesziomyces bullatus is a cosmopolitan species known on Echinochloa spp. Echinochloa oryzoides is reported here as a new host for M. bullatus.
The phylogenetic analysis of Moesziomyces, based on ITS sequence data (Fig. 151), confirmed previous phylogenetic relationships within the genus. The newly studied specimen of M. bullatus groups together with other M. bullatus specimens from other hosts. This species was inferred as sister taxon to a group containing M. antarcticus and M. kimberleyensis.
Glomeromycota C. Walker & A. Schüßler, in Schüßler, Schwarzott & Walker, Mycol. Res. 105(12): 1416 (2001)
Glomeromycetes Caval. -Sm., Biol. Rev. 73: 246 (1998)
Gigasporales S.P. Gautam & U.S. Patel, The Mycorrhizae, Diversity, Ecology and Applications (Delhi): 7 (2007)
Gigasporaceae J.B. Morton & Benny, Mycotaxon 37: 483 (1990)
Notes: Gigasporaceae was introduced by Morton and Benny (1990) with two genera, Gigaspora, the type of the family, and Scutellospora. Currently, the family present only one genus (Gigaspora) whose spores possess a unique germination pattern where the germination tubes originate from warts on a germinal wall, and do not present inner walls and germination shield similarly to other genera in the Gigasporales (Oehl et al. 2011a,b; Wijayawardene et al. 2022b).
Gigaspora Gerd. & Trappe Mycol. Mem. 5: 25 (1974)
Notes: Originally, Gigaspora was erected with five species (G. calospora (T.H. Nicolson & Gerd.) Gerd. & Trappe, G. coralloidea Trappe, Gerd. & I. Ho, G. gigantea (T.H. Nicolson & Gerd.) Gerd. & Trappe, G. gilmorei Trappe & Gerd., G. heterogama (T.H. Nicolson & Gerd.) Gerd. & Trappe) in Endogonaceae, which also comprised saprotrophic and ectomycorrhizal fungi, e.g., Endogone species (Gerdemann and Trappe 1974). The type species of Gigaspora is G. gigantea, which was originally described as Endogone gigantea T.H. Nicolson & Gerd. (Nicolson and Gerdemann 1968). The unique morphological characteristic linking the five species was the formation of spores at the top of a subtending hypha called sporogenous cell or bulbous suspensor-like cell. However, spores of G. gigantea had only one three–layered spore wall, while spores of the other four species had two or three spore walls. Moreover, G. gigantea spores germinated from germ warts, present on the inner surface of the third spore wall layer, and produced auxiliary cells ornamented with spines in the extraradical mycelium. Instead, the other species produced germination hyphae from a plate-like structure, called germination shield, formed on the upper surface of the innermost spore wall, and knobby auxiliary cells in the extraradical mycelium. Therefore, Walker and Sanders (1986) transferred the latter species to a new genus, Scutellospora, in Endogonaceae. Morton and Benny (1990) accommodated Gigaspora and Scutellospora in a new family, Gigasporaceae, established in a new order, Glomales, orthographically corrected into Glomerales (Schüßler et al. 2001), which included only fungi producing (or hypothetically producing) arbuscular mycorrhiza. Schüßler et al. (2001) transferred Gigasporaceae to a new order, Diversisporales, in a newly established phylum, Glomeromycota. Oehl et al. (2008) retained only Gigaspora in Gigasporaceae and distributed the known Scutellospora species in three new families with five new genera, which later were placed in Gigasporales, a new order introduced into Glomeromycota (Oehl et al. 2008, 2011a, 2011b; Goto et al. 2012). Differently from most Glomeromycota species, members of Gigasporales do not produce intraradical vesicles.
Currently, Gigaspora comprises eight species: G. albida N.C. Schenck & G.S. Sm., G. candida Bhattacharjee, Mukerji, J.P. Tewari & Skoropad, G. decipiens I.R. Hall & L.K. Abbott, G. gigantea, G. margarita W.N. Becker & I.R. Hall, G. polymorphira L.M. Yao, G.Y. Tao & L. Jiang, G. ramisporophora Spain, Sieverd. & N.C. Schenck, and G. rosea T.H. Nicolson & N.C. Schenck. Three other species have been described, but two do not possess the characteristics of Gigaspora (G. lazzarii Montecchi, Ruini & G. Gross—Montecchi et al. 1996 and G. tuberculata Neeraj, Mukerji, B.C. Sharma & A.K. Varma—Neeraj et al. 1993), and one, G. alboaurantiaca W.N. Chou, was synonymized with G. candida (Schüßler and Walker 2010). However, the latter synonymization is incongruent as G. alboaurantiaca description clearly states that the sporogenous cell is pale brown (Chou et al. 1991), while it is white in G. candida (Bhattacharjee et al. 1982) (Fig. 153).
[See PDF for image]
Fig. 153
Phylogram generated from Maximum Likelihood (ML) and Bayesian Inference (BI) analyses displaying the phylogenetic relationships between the Gigaspora siqueirae (in bold, blue) and members of the Gigasporaceae and Intraornatosporaceae. The family Dentiscutataceae, as sister to the clade Gigasporaceae–Intraornatosporaceae, was used to root the consensus tree. The trees were inferred using a dataset that includes concatenated sequences divided into five partitions (18S: 1–1786, ITS1: 1787–1883, 5.8S: 1884-2042, ITS2: 2043–2251, 28S: 2252–3047). For species in the Intraornatosporaceae only partial 28S sequences were available. Overall, the dataset included 15 species in six genera represented by 55 sequences. Sequence alignment was performed using MAFFT 7.243 (Katoh et al. 2019), strategy E-INS-i. In both ML and BI analyses, GTR + I + G was chosen as a nucleotide substitution model for each nucleotide partition (Abadi et al. 2019). The ML tree was estimated using RAxML-NG 1.0.1 (Kozlov et al. 2019), with a maximum likelihood/1000 bootstrapping run, and ML estimated proportion of invariable sites and base frequencies. In the BI analysis, four Markov chains were run over ten million generations in MrBayes 3.2 (Ronquist et al. 2012), sampling every 500 generations, with a burn-in at 30% of sampled trees. All parameters of the convergence diagnostic (Potential Scale Reduction Factor) indicated that convergence was obtained (Ronquist et al. 2012; Miller et al. 2015). The tree topology obtained from the ML analysis was identical to that generated in the BI analysis. Support values and posterior probabilities greater than 60% and 0.95, respectively, are indicated above or below the nodes. The bar indicates 0.005 expected change per site per branch
General data and phylogeny: Following the analysis of DGGE rDNA profiles of 51 strains of seven Gigaspora species, de Souza et al. (2004) recognized a strain (named UFLA872), initially identified from morphology as G. gigantea, that diverged from the other species, including seven G. gigantea strains. This strain, described here as G. siqueirae, also had pigmented spores’ cytoplasm, as in G. gigantea Séjalon-Delmas et al. (1998). The cytoplasm colour can be observed in a supernatant obtained from 5–20 spores crushed in 20 µL of water. However, the spore colour changes during aging, from bright yellow at maturity, then brownish yellow to dark yellowish brown at old age (Fig. 154), but the cytoplasm colour persists in aged healthy spores.
[See PDF for image]
Fig. 154
Gigaspora siqueirae (UFRN-Fungos 3674, holotype).a Glomerospores with bright yellow colour typical of the new species obtained on a plant host culture. b Mature dark yellowish brown glomerospores. c Glomerospores with bright yellow colour produced root-organ-culture. d Bulbous sporogenous cell with two wall layers attached to spore. e–f Spore wall with three layers (swl1–3). f Germ hypha (gh) developed from a wart formed on the lower surface of swl 3 and emerged from swl 2 and 1. g–i Glomerospore, intraradical hypha and auxiliary cells with spiny ornamentation produced by pigmented extraradical hyphae colonizing carrot root-organ-culture. a‒b. Spores in water. c‒f. Spores in PVLG. g‒i. Mycorrhizal structures obtained from in vitro cultures. Scale bars: a, c, g = 200 μm, b,h = 100 μm, d = 20 μm, e–f = 5 μm, i = 50 μm
Our phylogenetic analyses placed G. siqueirae in an autonomous clade in relation to all other sequenced Gigaspora species (Fig. 153). Gigaspora siqueirae can be distinguished from the other Gigaspora spp. by the DNA signature CGCGTG in the V9 region of the SSU nrDNA gene (de Souza et al. 2004; Bago et al. 1998). BLASTn searches, using the ITS (ITS1–5.8S–ITS2) region of the two sequences of G. siqueirae, showed that the identity of the most closely related species, G. margarita, was only 92.41–94.68% (accessions KP756509, KP756511; search on 31.01.2023). This also suggests G. siqueirae as a rare species.
Gigaspora siqueirae F.A. de Souza, Barros-Barreto, Magurno, B.T. Goto, sp. nov.
Index Fungorum number: IF 902416; Facesoffungi number: FoF 15874 Fig. 154
Etymology: in honor of Prof. José Oswaldo Siqueira, collector and keeper of the species, member of the Brazilian Academy of Science, soil microbiologist, for his enormous contribution to mycorrhizal research, especially in Brazil.
Holotype: UFRN Herbarium Fungos 3674; isotype: UFRN Herbarium Fungos 3675, 3676.
Glomerospores (= spores) formed singly in the soil, terminally on a sporogenous cell. Spores globose, (240–)367(–411) µm diam (n = 100), subglobose to oblong, 220–420 × 300–460 µm; yellow cream when young, bright yellow to dark yellow at maturity (3B4–2C8, most 2A8–2C8, Fig. 154A–C), dark yellowish brown (5D8–6F8) when older (Fig. 154B); with yellow (1A7–2A8) cytoplasm. Spore wall structure composed of three layers, spore wall layers (swl) 1–3 (Fig. 154E–F): swl1, forming the spore surface, smooth, rigid, hyaline, < 1.0–1.5 µm thick; swl2 yellow to dark yellow, laminate (each sublayer 1.0–1.2 µm thick), (7–)13(–28) µm thick (n = 54); swl3, a germinal layer, concolorous with and tightly adherent to swl2, < 1.0–1.5 µm thick, with many round papillae/warts on the inner surface formed prior to germination (Fig. 154F). Spore wall layers 1 and 2 turn brownish (6D8–6F7) in Melzer’s reagent. Sporogenous cell (Fig. 154D) formed terminally on a coenocytic subtending hypha, becoming sparsely septate at attaining full spore development, globose to subglobose, (30–)49(–68) µm wide. Wall of sporogenous cell (9.0–)12.5(–14.0) µm thick near the spore base, 2.5–7.0 µm thick 30 µm below the spore base, composed of two layers continuous with swl1 and swl2; rarely with a peg-like projection, up to 50 µm long and 2.5–12.5 µm wide. Sporogenous cells loosely associated with spores, found only in ca. 27% of mature spores. Germination by hyphae growing from the papillae of swl3 and penetrating through swl2 and swl1, usually near the sporogenous cell. Auxiliary cells borne in clusters of ca. 10, with spiny projections, sometimes attached to the host plant root surface (Fig. 154I). Mycorrhiza with arbuscules, straight, and coiled hyphae. Intraradical mycelium without vesicles. Arbuscules with a trunk, 5–7 µm wide, and numerous branches narrowing abruptly towards their tips. In carrot root-organ-culture, straight hyphae 5–11 µm wide, coiled hyphae 5–7 µm wide (Fig. 154H–I).
Material examined: Brazil. Boa Esperança municipality, in cultures (multi-spore) on Urochloa decumbens (Stapf) R.D. Webster (J.O. Siqueira, 27 May 2003); deposited in DCS-UFLA (holotype). ISOTYPE deposited at the International Culture Collection of Glomeromycota (CICG, Blumenau, Santa Catarina, Brazil, http://www.furb.br/cicg/index.php?lang=EN) under the accession MGR253. The fungus was originally cultured from a soil sample collected in March 1986 from a coffee (Coffea arabica L.) plantation in Cerrado Bioma, the Brazilian Savanna, in the south of Minas Gerais State in SE Brazil. The locality is around 900 m above sea level and consists of cerrado fragments and coffee crops. The soils are highly weathered, clayey, acidic, low-fertility Dusky Red Latossol (Oxisol), with a pH (in water) of around 5.7, available P (Mehlich) of 2 mg‧dm−3, and Ca + Mg of 431.0 mg‧dm−3. Mean precipitation and temperature in this area is 1,400 mm and approx. 19 °C, respectively. The trap culture in which the species was first established was set up on 27 Aug. 1987 by sowing seeds of U. decubens grass in a substrate composed of fumigated soil and sand (1:1 v/v) that received spore suspension. The trap cultures were fertilized with nutrient solution lacking P. Subsequently, multi-spore, single species cultures were established using purified spore suspensions poured onto the roots of mycorrhiza-free U. decumbens seedlings at transplanting. The species has not been established in single-spore cultures, but it has been repeatedly subcultured since its original culture in 1987. Four different multi-spore lines have been maintained, with the reference numbers UFLA 175, 178, 179, and 872.
GenBank accession numbers: OQ302571, OQ302575-OQ302576, OQ302579-OQ302583, OQ302587, OQ302590-OQ302591 and OQ680681-OQ680684. The primer combination used for amplification and cloning of the near clomplete SSU and complete ITS region were NS1 5′-GTAGTCATATGCTTGTCTC-3′ and ITS4 5′-TCCTCCGCTTATTGATATGC-3′ (de Souza et al. 2004). For the LSU we used the primer combination LR1 5′-GCATATCAATAAGCGGAGGA-3′- van Tuinen et al, (1998) and FLR4 (5′-TACGTCAACATCCTTAACGAA-3′- Gollotte et al., (2004). The DNA extraction, amplification, cloning and sequencing was carried out by de Souza et al. (2004) for amplicons covering the SSU and ITS regions. Similarly, the LSU amplicons were obtained by the authors.
Notes: Gigaspora siqueirae is mainly distinguished by its unique phylogeny, which accommodated its sequences in a fully supported clade, inside a wider polyphyletic clade hosting all sequenced Gigaspora, Intraornatospora and Paradentiscutata species (Fig. 153). Despite its phylogenetic placement, we decide to retain the new species as Gigaspora, based on the morphological features of spores, convincingly resembling those of the genus. Potential emendations inside of the polyphyletic clade will be discussed once additional molecular data will be available for the Intraornatospora and Paradenticutata genera. Morphologically, G. siqueirae is indistinguishable from G. gigantea. Both species produce spores of similar size, the thickness of their spore wall overlaps, and, most importantly, the spore color—“bright yellow”—of these two species come from the pigmentation of the cytoplasm rather than the spore wall. That characteristic was diagnostic to differentiate G. gigantea from all previously described Gigaspora species. Interestingly, the origin of distinctly coloured spores with a colourless or at most lightly coloured spore wall from the pigmentation of the spore cytoplasm has so far been identified only in one other species of the Glomeromycota, i.e., Desertispora omaniana (Symanczik, Błaszk. & Al-Yahya'ei) Symanczik, Błaszk., Kozłowska & Al-Yahya’ei (Symanczik et al. 2018), originally described as Diversispora omaniana Symanczik, Błaszk. & Al-Yahya’ei (Symanczik et al. 2014). Phylogenetically G. gigantea behaves differently, forming a sister relationship with G. albida and G. candida as in the tree depicted in Fig. 153. The relationship of G. gigantea with G. albida and G. candida and the distinctiveness of G. siqueirae was also evident from a PCR-Denaturing Gradient Gel Electrophoresis (DGGE) profiling of inter- and intraspecies sequence heterogeneity of the V9 region of the 18S rRNA gene of 51 of Gigaspora strains, which included G. albida (6), G. candida (1), G. decipiens (2), G. gigantea (9), G. margarita (11), G. ramisporophora (1), G. rosea (19) and G. siqueirae (1), represented by the strain UFLA872 and one Gigaspora sp. (TW1-1) (de Souza et al. 2004).
[See PDF for image]
Fig. 155
The phylogenetic tree of Mucoralean species (Mucor flavus group), constructed using LSU rDNA sequences, includes Pilaira anomala as an outgroup. Sequences are labelled with their database accession numbers. Bootstrap support values from maximum likelihood (ML-BS) greater than or equal to 60% and Bayesian posterior probabilities (BPP) greater than or equal to 0.60 are indicated at the nodes. The scale bar indicates expected changes per site. Type and ex-neotype strains are marked with T and NT, respectively. The newly generated sequences are indicated in bold and blue (Mucor soli)
In summary, recognition of Gigaspora species is difficult because their spores have few diagnostic morphological characteristics, of which, e.g., spore size and colour, especially in field-collected specimens, often overlap (Bentivenga and Morton 1995; Bago et al. 1998; de Souza et al. 2004). Therefore, the only reliable way to distinguish G. siqueirae from G. gigantea and other Gigaspora species is to reveal their differences in molecular phylogenies (Fig. 153).
Mucoromycota Doweld Prosyllabus Tracheophytorum, Tentamen Systematis Plantarum Vascularium (Tracheophyta) (Moscou): LXXVII (2001)
Notes: Species of Mucoromycota are distributed across the classes Endogonomycetes Doweld, Mucormycetes Doweld, and Umbelopsidomycetes Tedersoo et al. (Spatafora et al. 2016). They form a diverse and cosmopolitan group of saprobic fungi and some of them can be associated with aquatic hosts (Voigt 2016). Species in this phylum are also reported as saprophytes, endocommensals, symbionts (endomycorrhizal), or even parasites (Benny et al. 2016). See Wijayawardene et al. (2022) for the most up-to-date information on Mucoromycota.
Mucoromycotina Benny, in Hibbett et al., Mycol. Res. 111(5): 517 (2007)
Notes: The subphylum Mucoromycotina is characterized by well-developed, coenocytic mycelium with septa present only at the base of reproductive structures or old mycelium. Asexual reproduction occurs through the production of spores formed in sporangia, sporangioles and merosporangia, or via the formation of chlamydospores. Less commonly, it occurs through the fragmentation of the mycelium and the formation of budding cells. In cases where sexual reproduction is documented, it is recognized by the production of zygospores, which can originate from opposite or apposed gametangia (Benny et al. 2016). See Tedersoo et al. (2018) and Wijayawardene et al. (2022) for the most recent treatments and updated accounts of Mucoromycotina.
Mucormycetes Doweld Prosyllabus Tracheophytorum, Tentamen Systematis Plantarum Vascularium (Tracheophyta) (Moscou): LXXVII (2001)
Notes: The class Mucoromycetes (Doweld 2001) has fungi commonly found in various environmental conditions and can colonize various organic substrates. Several taxa within this class are very important in biotechnological processes owing to their capacity to produce enzymes and organic acids (Ekanayaka et al. 2022). See Wijayawardene et al. (2022) for the most recent information and updated accounts of Mucoromycetes.
Mucorales Dumort. [as ‘Mucorarieae’], Analyse des familles des plantes: avec l'indication des principaux genres qui s’y rattachent: 73 (1829)
Notes: Mucorales encompass the largest number of genera (55) and species (over 300) distributed across 14 families of Mucoromycota (Wijayawardene et al. 2022). Members of Mucorales mainly exhibit rapid mycelial growth and generate numerous of asexual structures, resulting in colonies with a cottony appearance and having different colors (Alvarez 2013). Specialized structures for nutrient fixation and absorption, such as rhizoids, and the presence of stolons and chlamydospores are commonly observed (Kirk et al. 2008). See Tedersoo et al. (2018) and Wijayawardene et al. (2022) to the most recent treatments and updated accounts of Mucorales.
Mucoraceae Fr. [as ‘Mucoroidei’], Syst. mycol. (Lundae) 1: xlix (1821).
Notes: Mucoraceae encompasses the majority of Mucorales taxa. Studies based on DNA have changed the classification of this family (e.g., Hoffmann et al. 2013). See Wijayawardene et al. (2022) for the most recent information and updated accounts of the family.
Mucor Fresen., Beitr. Mykol. 1: 7 (1850)
Notes: Mucor is the largest genus in number of species (98 valid names) in Mucoromycota (de Souza et al. 2022; Wijayawardene et al. 2022). Traditionally, species in this genus were delineated based on morpho-physiological characteristics, and recently the combination of traditional methods and molecular approaches has facilitated the resolution of species in this genus (Lima et al. 2020; Zhao et al. 2023; Gajanayake et al. 2023) (Figs. 155 and 156).
[See PDF for image]
Fig. 156
The phylogenetic tree of Mucor-related species, constructed using ITS rDNA sequences, includes Pilaira anomala as an outgroup. Sequences are labelled with their database accession numbers. Bootstrap support values from maximum likelihood (ML-BS) greater than 60% and Bayesian posterior probabilities (BPP) greater than 0.60 are indicated at the nodes. The scale bar indicates expected changes per site. Type and ex-neotype strains are marked with T and NT, respectively. The newly generated sequences are indicated in bold and blue (Mucor soli)
Mucor soli C.A. de Souza, E.V. de Medeiros & R.J.V de Oliveira, sp. nov.
Index Fungorum number: IF 902417; Facesoffungi number: FoF 09923; Fig. 157
[See PDF for image]
Fig. 157
Mucor soli (URM 95261, holotype). a Colonies (verse and reverse) on MEA and PDA at 25 °C after 1 week. b Observed columellae types. c‒d Simple sporangiophore with sporangia. e Simple branched sporangiophore with sporangia. f Sympodially branched sporangiophore. g Sporangiospores. Scale bars: b and g = 20 µm, c‒f = 50 µm
Etymology: Referring to the substrate, soil, from which the fungus was first isolated.
Holotype: URM 95261
Saprobic on submerged decaying wood. Sexual morph: Not observed. Asexual morph: Colonies fast growing (9 cm diam. × 0.5 cm in height after four days at 25 ℃ on PDA), initially white then becoming yellowish to cream (MP 19D1), reverse brownish (MP 12K13) with irregular margins. Odour acid. Sporangiophores growing directly from the substrate, cylindrical, up to 22.5 μm diam., simple and sympodially branched (up to four times), with short or long branches, hyaline, presenting a texture that varies between smooth-walled and slightly rough (striated). Occasionally exhibiting one to three septate below sporangia (either in sporangiophores or in sporangiophore branches), some showing a swelling close to the columellae, especially on short branches sporangiophores. Sporangia globose (15.5‒) 20–80 (–80.5) µm diam. very rarely reach 95 μm diam., and subglobose to slightly dorsiventrally flattened 12.5–55 × 18.5–60 µm, wall evanescent, light brown, slightly echinulate. Columellae variable in shape, smooth-walled, hyaline, frequently globose to subglobose, (13.5‒) 15.5–72.5 (‒76.5) µm diam., cylindrical, 50–75 × 40–65 µm diam., obovoid to slightly ellipsoidal, 12.5–40 × 10–50 µm, ovoid 20–65 × 15.5–50 µm, conical or oblong, 15–30 × 10.5–25 µm, pyriform 20.5–50 × 10.5–40 and rarely applanate, 12–25 × 15–25 µm. Collar evident. Sporangiospores globose and subglobose, (2.5–) 3–6 (–7.5) µm diam. Chlamydospores infrequent, barrel shaped.
Culture characteristics: colonies PDA. At 10 °C—very limited growth (3.2 cm in diam. in 168 h); total lack of reproductive structures; At 15 °C—low colonies (< 1 mm in height) with slow growth (6 cm in diam. in 168 h); good sporulation. At 20 °C—colonies up to 3 mm diam. with good growth (9 cm in 120 h) and good sporulation. At 25 °C—better growth (9 cm in 96 h) and excellent sporulation. At 30 °C—good growth (8 cm in 96 h) and excellent sporulation. At 35ºC—lack of growth and sporulation. Mucor soli exhibited a better growth rate and sporulation on PDA than on MEA at all tested temperatures. At 15 °C, on PDA, the production of reproductive structures was good, and the sporangiophores were simple or slightly sympodially branched, whereas on MEA the production of reproductive structures was very poor and the sporangiophores were simply branched. The columellae were mostly globose (up to 24 μm diam.) at 15 °C on PDA and MEA. At 36 °C after 1 week in the dark, colonies on PDA and MEA growing fast and attaining a diameter of 90 mm and morphologically similar as described at 25 °C.
Material examined: BRAZIL, Pernambuco State, Quipapá Municipally, Serra do Quipapá, 8° 48′ 21.15″ S, 36° 20′ 80″ W, from soils of a semi-arid region, 15 April 2022, C.A. de Souza, LEMA 77962 (URM 95261, holotype); ex-type, URM 8698; living culture ibid. URM 8699.
GenBank accession numbers: URM 8698, ex-type: ITS = OQ939972 and LSU: OQ939973; URM 8699: ITS = OQ944938 and LSU: OQ944939.
Notes: Phylogenetic analysis of two sequence datasets (ITS and LSU rDNA) revealed that Mucor soli sp. nov. formed a distinct lineage among other species of Mucor. The new species is phylogenetically positioned in a well-supported clade (ML = 100% and BY = 1) among species of the Mucor flavus group (Figs. 2, 3). Morphologically, M. soli sp. nov. exhibits a strong resemblance to most species of the M. flavus group (Schipper 1975). Mucor soli sp. nov. predominantly produces globose sporangia, with diameters ranging from 20 to 80 µm, occasionally reaching up to 95 µm diam. In contrast, M. merdophylus exhibits globose to slightly dorsally flattened sporangia, with 35.5 to 95 µm diam., and occasionally reaching up to (–110) μm diam. (Lima et al. 2020). Mucor minutus, on the other hand, is characterized by globose sporangia reaching up to 160 µm in length (Alves et al. 2021). Mucor flavus has larger sporangia than those observed in the present study, with a maximum of 175 µm diam., whereas M. aligarensis and M. saturninus have sporangia with a maximum diameter of 100 µm. These differences in sporangia are distinctive morphological characteristics that differentiate M. soli sp. nov. from the other species in the group. Mucor soli also exhibit variable columella shapes and sizes, measuring 15.5 to 72.5 µm diam. Despite the close morphological and phylogenetic affinity of M. soli with species of the M. flavus group, differences in the formation of columellae and their sizes are distinctive morphological characteristics. Mucor aligarensis, M. flavus, M. minutus, and M. saturninus have been described as producers of columellae up to 130 μm diam. (Schipper 1975), larger than those observed for M. soli sp. nov. Mucor soli sp. nov. also differ from other related species by the production of globose and subglobose sporangiospores ranging from 3 to 6 µm diam. having thin and smooth-walled. Mucor merdophylus has larger sporangiospores (5 to 26 µm diam.) with granular content and smooth-walled (Lima et al. 2020), while M. saturninus has sporangiospores with up to 8 µm diam. and thick-walled and predominantly ellipsoidal (Schipper 1975). The sporangiospores of M. aligarensis and M. flavus are also larger (up to 12 × 6.5 μm) and M. minutus has subglobose to ellipsoidal sporangiospores of up to 7 μm diam. (Schipper 1975; Alves et al. 2021). These differences in sporangiospore sizes and shapes provide additional criteria for distinguishing Mucor soli sp. nov. from the other species in the group.
Umbelopsidomycetes Tedersoo, Sánchez-Ramírez, Kõljalg, Bahram, Döring, Schigel, T. May, M. Ryberg & Abarenkov, Fungal Diversity 90: 152 (2018)
Umbelopsidales Spatafora, Stajich & Bonito, in Spatafora, Chang, Benny, Lazarus, et al., Mycologia 108(5): 1035 (2016).
Umbelopsidaceae W. Gams & W. Mey., in Meyer & Gams, Mycol. Res. 107(3): 348 (2003)
Notes: Umbelopsidaceae was introduced by Meyer and Gams (2003). For the taxonomic treatment of this genus, we follow Wijayawardena et al. (2022).
Umbelopsis Amos & H.L. Barnett, Mycologia 58(5): 807 (1966)
Notes: Umbelopsis was introduced Amos and Barnett (1966) and later Meyer and Gams (2003) included Micromucor species based on molecular evidence. Morphologically Umbelopsis species are characterised by velvety and often coloured colonies, aerial hyphae, small or absent columellae, sporangiophores which often arise erectly from a vesicle; and sporangiospores which vary in shape and are similar in colour to the sporangia (Linnemann 1941; Linnemann 1969; Meyer and Gams 2003). For the taxonomic treatment of this genus, we follow Wang et al. (2022a, b) (Fig. 158).
[See PDF for image]
Fig. 158
Phylogram generated from RAxML analysis based on combined ITS, SSU, and LSU sequence data of Umbelopsis isolates. Related sequences were obtained from GenBank. Twenty-eight taxa are included in the analyses, which comprise 2264 characters including gaps. Tree is rooted to Mortierella antarctica CBS 609.70. Tree topology of the ML analysis was similar to the BI. The best scoring RAxML tree with a final likelihood value of − 7485.315022 is presented. The matrix had 320 distinct alignment patterns, with 0.36% of undetermined characters or gaps. Estimated base frequencies were as follows; A = 0.282345, C = 0.186030, G = 0.243519, T = 0.288107; substitution rates AC = 0.885256, AG = 3.840307, AT = 1.754600, CG = 0.329086, CT = 6.911236, GT = 1.000000; gamma distribution shape parameter α = 0.179154 Bootstrap support values for ML equal to or greater than 70% and Bayesian posterior probabilities equal to or greater than 0.7 BYPP are given above the nodes. The scale bar indicates 0.02 changes. The isolates obtained in this study are in blue and ex-types are in black bold
Umbelopsis hingganensis Tong Wu, T. Du, M. M. Ding & L.J. Xu, sp. nov.
Index Fungorum number: IF 902418; Facesoffungi number: FoF 16073; Fig. 159
[See PDF for image]
Fig. 159
Umbelopsis hingganensis (CCTCC AF 2022084, holotype). a–c Upper view and reverse view of colony on PDA (a), OA (b), and CMA (c). d–h Sporangia. i Sporangiospores. Scale bar: d, e = 100 μm, f–i = 10 μm
Etymology: Referring to the location where the type material was collected.
Holotype: HMPHU 1269A.
Saprobic on fallen dead leaves. Sexual morph Not observed. Asexual morph Sporangiophores arising from a vesicle, verticillately, simply branched, hyaline, 26.8–76.6 μm long, 2.1–4.2 μm wide. Sporangia globose to subglobose, 7.1–11.8 μm in diameter, reddish brown to reddish purple when mature, multi-spored. Sporangiospores irregularly angular, (2.4–) 3.5–3.7 (− 6.1) × (1.9–) 2.7–2.8 (− 3.6) μm, reddish brown to reddish purple, without oil droplet. Columellae absent or slightly convex. Collars absent. Chlamydospores globose to subglobose, single or bunching, 5.3–14.8 μm in diameter, containing abundant oil droplets (Fig. 160).
[See PDF for image]
Fig. 160
Phylogram generated from maximum likelihood (ML) analysis based on a combined tef1-α and SSU sequence data for species in Lamproderma. Two taxa of the genus Meriderma were selected as the outgroup. Related sequences were obtained from Genbank. In total, fifty-two sequences were included in the analyses, which comprise 1390 characters including gaps (745 for tef1-α, 645 for SSU). Single gene analyses were also performed to compare the topology and clade stability with combined gene analyses. The ML analysis with 1000 bootstrap replicates gave a best tree with a final likelihood value of − 10,624.2876 is presented. The matrix had 625 distinct alignment patterns, with 37.34% undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.2389, C = 0.2852, G = 0.272, T = 0.2039; substitution rates AC = 0.9272, AG = 2.3616, AT = 1.7787, CG = 0.8366, CT = 5.6937, GT = 1.0000; gamma distribution shape parameter α = 1.054. Bootstrap values for ML equal to or greater than 75% and clade credibility values greater than 0.90 (the rounding of values to 2 decimal proportions) from Bayesian-inference analysis labeled on the nodes. The new species is indicated in bold and blue
Culture characteristics: Colonies growing on PDA, cottony, at first white for the hyaline hyphae and when old pinkish due to the occurrence of pigmented sporangia, reaching 30–32 mm in 14 days at 25 °C. Colonies growing on OA, velutinous, yellowish white, reaching 9–11 mm in 14 days at 25 °C. Colonies growing on CMA, flat, hairy or cottony, at first white for the hyaline hyphae and when old reddish purple due to the occurrence of pigmented sporangia, reaching 36–38 mm in 14 days at 25 °C.
Material examined: China, Heilongjiang Province, Greater Hinggan Mountains, on fallen dead leaves of Larix gmelinii, September 2018, Zeyu Li and Zhedong Zhang (HMPHU 1269A, holotype), ex-type CCTCC AF 2022084.
GenBank numbers: ITS = OM838288, LSU = OM838391, SSU = OQ147016.
Notes: Umbelopsis hingganensis differs from its closely related species in that it has reddish brown to reddish purple, globose, 7.1–11.8 μm diameter sporangia and angular sporangiospores. Umbelopsis hingganensis and the other three closely related species U. isabelline, U. ovata, and U. autotrophica formed a monophyletic group with 90% ML and 1.00 BYPP support based on the phylogenetic analysis on their ITS, LSU and SSU sequences. Umbelopsis isabelline and U. autotrophica differ by having globose sporangiospores from U. hingganensis (Bezerra et al. 2018; Meyer and Gams 2003). Umbelopsis ovata differ differs from U. hingganensis fusiform sporangia (Yip 1986).
Fungus-like taxa
Myxomycetes G. Winter, Rabenh. Krypt. -Fl., Edn 2 (Leipzig) 1.1: 32 (1880) [1884]
Physarales T. Macbr., N. Amer. Slime-Moulds, Edn 2 (New York): 22 (1922)
Lamprodermataceae T. Macbr. [as “Lamprodermeae”], N. Amer. Slime-Moulds: 136 (1899)
Notes: This family was introduced by Macbride (1922) with Lamproderma Rostaf. as the type genus. Recent treatment follows Fiore-Donno et al. (2012), later amended by Leontyev et al. (2019). Currently the genera Diacheopsis Meyl., Colloderma G. Lister, and Elaeomyxa Hagelst. are accommodated in the family, but it is possible that these three genera would be treated as synonyms of Lamproderma when further molecular data are obtained, or even that several new genera will be created (Leontyev et al. 2019).
Lamproderma Rostaf., Vers. Syst. Mycetozoen (Strassburg): 7 (1873)
Notes: Lamproderma was introduced by Rostafiński (1873) with Lamproderma columbinum as the type species. Lamproderma species are generally nivicolous, so they are mainly found at high, snowy elevations. This is possibly due to their adaptation to radiation thanks to the dark, melanized spores. Currently, 57 epithets are listed and considered valid for Lamproderma according to Lado (2005–2023) (Fig. 160).
Lamproderma subcristatum G. Moreno, López-Vill. & A. Sánchez, sp. nov.
Index Fungorum Number: IF 902419; Facesoffungi number: FoF 16074; Figs. 161, 162, 163
[See PDF for image]
Fig. 161
Lamproderma subcristatum (AH 55703, holotype): a, b Sporocarps, c Sporocarp under transmitted light, d Detail of the apex of the columella and capillitium, e Spores under LM, f–g Spores under SEM, h Detail of the spore ornamentation under SEM. Bars: a–b = 1 mm, c–d = 100 µm, e = 10 µm, f–g = 2 µm, h = 1 µm
[See PDF for image]
Fig. 162
Lamproderma subcristatum (AH 55704, paratypus) a, b Sporocarps, c Sporocarp under transmitted light, d Detail of the apex of the columella and capillitium, e Spores under LM, f–g Spores under SEM, h Detail of the spore ornamentation under SEM. Bars: a–b = 1 mm, c–d = 100 µm, e = 10 µm, f–g = 2 µm, h = 1 µm
[See PDF for image]
Fig. 163
Lamproderma spinulosporum (M. Meyer 3265, holotype): a Sporocarps, b Sporocarps under transmitted light, c Detail of the apex of the columella and capillitium, d Capillitium, e Spores under LM, f–g Spores under SEM, h. Detail of the spore ornamentation under SEM. Bars: a = 1 mm, b = 0.5 mm, c–d = 100 µm, e = 10 µm, f–g = 2 µm, h = 1 µm
Etymology: the epithet subcristatum refers to the ornamentation of the spores which is formed by short crests.
Holotype: AH 55703.
= Lamproderma spinulosporum Mar. Mey., Nowotny & Poulain, sensu Sánchez & Moreno, Bol. Soc. Micol. Madrid 40: 53–55 (2016).
Fructifications isolated or in groups, usually piled up, sessile to subsessile. Sporocarps 0.8–1.5 mm diam., globose to subglobose, even ovoid, flattened laterally, rarely elongated. Peridium simple, membranous, persistent at the base of the sporotheca, iridescent, blue to violet- or grey-blue, dehiscence irregular. Hypothallus membranous, discoid, orange brown to reddish brown, translucent. Stalk usually absent, very short when present, 0.04–0.1 mm long, cylindrical erect, dark reddish brown. Columella reaching 1/3–1/2 the height of the sporotheca, reddish brown, with a blunt apex from which the threads of the capillitium radiate. Capillitium dense, branched, anastomosed, creating a three-dimensional net, pale under the stereomicroscope, reddish brown on transmitted light, paler towards the tips, threads with polygonal, membranous expansions, concolorous or paler than the main branches. Spores globose to subglobose, blackish in mass, brown under transmitted light, 10–12 µm, av. 11 µm, Qav = 1 (n = 20), with short, low crests. Under SEM the crests are irregular and sinuous, variable in length. Plasmodium unknown.
Material examined: Spain, Segovia, Sierra de Guadarrama National Park, road from Cotos to Navacerrada, on twigs of Juniperus communis ssp. nana (Gaudin) Endl., 30 March 2005, 1825, m, AH 33624, AH 50788. 5 April 2005, AH 45876. 16 April 2005, 1850, m, AH 33627. 20 April 2005, AH 33621, AH 33622, AH 33629, AH 33631. 22 April 2005, AH 33626, AH 45878, AH 45879. 30 April 2005, 1825, m, AH 33625. 2 May 2005, 1850 m, AH 45881, AH 55622. 15 April 2006, 1875, m, AH 45877. On bark of Pinus sylvestris L., 6 May 2007, 1850, m, AH 45882. Navacerrada mountain pass, on needles of Pinus sylvestris, 27 April 2007, 1950, m, AH 45880. Plant debris of Festuca sp., 15 June 2005, 2075, m, AH 33628. Sierra de Guadarrama National Park, road from Cotos to Valdesquí, on needles of Juniperus communis ssp. nana (Gaudin) Endl., 09 May 2009, AH 55543, AH 55551. Sierra de Guadarrama National Park, Cabeza de Hierro, among bryophytes on woody remnants of Cytisus oromediterraneus Rivas Mart., T.E. Díaz, Fern. Prieto, Loidi & Penas, 2100 m, leg. A. Sánchez, 13 May 2013, AH 55703 (holotype), AH 55704. On woody remnants of Cytisus oromediterraneus, AH 49113. 2100 m, AH 55702, AH 55697.
Other material examined: France, Lamproderma spinulosporum St. Paul/Isère-73. Savoie, 1300 m, à proximité de la neige fondante, sur végétaux vivants, 8 May 1988, M. Meyer 3265 Holotype dans l’ herbier du Jardin Botanique National de Meise, Belgique (BR). Isotype in AH.
Genbank numbers: tef1-α = OQ222156-OQ222168, SSU = OQ241222-OQ241234.
Notes: Lamproderma subcristatum sp. nov. did not appear in previous works or monographs written by Kowalski (1968, 1970), Lister (1894, 1911, 1925), Macbride (1899, 1922), Martin and Alexopoulos (1969), Nannenga-Bremekamp (1991), Neubert et al. (2000), and Poulain et al. (2011). This new species is characterised by the sessile or shortly stalked sporocarps, the eye-catching blue, iridescent colour, the pale capillitium under the stereomicroscope that turns out to be reddish brown under transmitted light, and by the 10–12 µm diam. spores with spinules joined into short, irregular crests.
Macroscopically similar species are Lamproderma cacographicum Bozonnet, Mar. Mey. & Poulain, Lamproderma argenteobrunneum A. Ronikier, Lado & Mar. Mey., and Lamproderma spinulosporum Mar. Mey., Nowotny & Poulain. The spores of L. cacographicum are larger [(10.5–)12–15 µm diam] and, although they also have crests, the crests are broader, longer, and more prominent (Bozonnet et al. 1997; Moreno et al. 2002).
Lamproderma argenteobrunneum has spores similar in size [(8–)9–11(–12) µm diam], but the spores are ornamented with loosely arranged, short, curved ridges (Ronikier et al. 2010). The peridium also differs, since the colour is silvery brown with golden reflections at the base and with an areolate pattern visible only under transmitted light. In contrast, the peridium of Lamproderma subcristatum sp. nov. is devoid of any type of ornamentation.
Lamproderma spinulosporum is the most morphologically similar species to the new species described herein. However, the spore ornamentation is distinctly spinulose, with more evident spinules that are rarely joined into crests (Fig. 163). The new records of this species for Spain given by Sánchez & Moreno (2016) must be corrected to the new species proposed herein. Phylogenetic analyses based on tef1-α, SSU, and both genes combined support the separation of Lamproderma subcristatum from any other already described species within the genus. The separation from the closest taxon, L. cristatum, is significantly supported (99% ML, 1.00 BYPP, Fig. 160).
Fungi Imperfecti
Amerosporae
Exesisporites Elsik, Trans. Gulf Coast Assoc. Geol. Societies 19: 516 (1969)
Notes: The monotypic genus Exesisporites (Type: E. neogenicus Elsik) was proposed by Elsik (1969) with the following diagnosis: “Unicellular, aseptate, psilate, monoporate fungal spores of circular outline with lenticular to spherical shape. The centrally located pore in most specimens is surrounded by a dark circular patch which is interpreted as a thickened wall. This polar area is occasionally found free of the spore”. The genus was differentiated from Monoporisporites Hammen 1954a, b, Lacrimasporonites Clarke 1965 and Basidiosporites Elsik 1968 on the character of the pore or hilum and the orientation of the spores in mounted residues. It is differentiated from Reticulatisporonites Elsik 1968 by the lack of reticulate ornamentation. Glass et al. (1986) cited possible affinity of Exesisporites to the extant imperfect fungus Nigrospora Zimm. However, this affinity is uncertain.
[See PDF for image]
Fig. 164
Map of the Arabian Peninsula showing various countries and Red Sea coast showing the location of area of study. (Modified after https://commons.wikimedia.org/wiki/File:Saudi_Arabia_Topography.png)
[See PDF for image]
Fig. 165
A‒D Exesisporites chandrae R.K. Saxena & A. Kumar sp. nov. A. Slide M7a, 163 × 9.5; size: 43 µm in diameter. B Slide M7a, 140.5 × 10 (holotype); size: 43 × 36.4 µm. C Slide M7b, 140.5 × 5; size: 43.5 × 32.7 µm; D Slide M2b, 133.5 × 5.3; size: 57 × 37 µm. EMonoporisporites jansoniusii R.K. Saxena & A. Kumar sp. nov. Slide AM1b,134.8 × 12.8 (holotype); size: 88 × 86 µm. F‒HMonoporisporites pattersonii R.K. Saxena & A. Kumar sp. nov. F Slide L2c, 164.5 × 10; size: 42 × 38 µm; G Slide L2c, 167 × 7 (holotype); size: 39 × 44 µm. H Slide L1c, 149.5 × 2; size: 46 × 41 µm. I‒JMonoporisporites valdiyae R.K. Saxena & A. Kumar sp. nov. I Slide AM2c,158.5 × 18 (holotype); size: 39.6 × 28.2 µm; J Slide AM1b,154 × 16.5; size: 40.4 × 30.3 µm. KDicellaesporites plicatus R.K. Saxena & A. Kumar sp. nov. Slide WH1d, 158.5 × 8 (holotype); size: 27.8 × 13.5 µm. LDicellaesporites verrucatus R.K. Saxena & A. Kumar sp. nov. Slide L3c, 135.3 × 18 (holotype); size: 65 × 38 µm. MAlleppeysporonites elsikii R.K. Saxena & A. Kumar sp. nov. Slide M1c, 160.8 × 16.5 (holotype); size: 95 µm, hyphae 81 × 10 µm
Exesisporites chandrae R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901367; Facesoffungi number: FoF 16075; Fig. 165A‒D.
[See PDF for image]
Fig. 166
Tarzetta tibetensis (HKAS 127118, holotype). a–c Typical mature specimens. d Anatomic structure of a portion of an apothecium. e, f Asci and paraphyses. g–j Asci. k Paraphyses. l Apex of ascus. m Ascospores. Scale bars: a–c = 0.5 cm, d–f = 200 μm, g, h = 100 μm, i = 50 μm, j = 20 μm
Etymology: The specific epithet honours Dr. Anil Chandra, Birbal Sahni Institute of Palaeosciences, Lucknow, India.
Holotype: Fig. 165B, slide M7a; 140.5 × 10; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Fungal spores spherical to oval in shape. Size 43–57 × 32–43 µm. Monoporate, pore small, surrounded by 2–4 µm thick pore margin; pore associated with a prominent fold. Spore wall 1–2.5 µm thick, psilate.
Notes: Exesisporites neogenicus Elsik 1969 can be distinguished from the present species by its smaller size (15–20 μm in diameter) and thinner spore wall (0.5 μm). Exesisporites verrucatus Kumar 1990 is conspicuously distinct in being smaller in size (25–27 × 28–30 μm). It also has two layered, verrucate spore wall imparting angularity to the ambitus. Exesisporites psilatus Saxena 2000 is larger in size (50–82 × 50–71 μm). This species can be distinguished from all the other species of Exesisporites by presence of a prominent fold associated with the pore.
Monoporisporites Hammen 1954
Monoporisporites (Type: M. minutus Hammen) was proposed by van der Hammen (1954a, b). Subsequently, Elsik (1968), Sheffy & Dilcher (1971) and Kalgutkar and Jansonius (2000) emended the diagnosis of this genus, Kalgutkar and Jansonius (2000) considered Ornatisporites M.G. Parsons & G. Norris (Parsons and Norris 1999), Polyporisporites Hammen 1954a, b, Psiammopomopiospora Sal.-Cheb. & Locq. 1980, Psiamspora Sal. -Cheb. & Locq. 1980 and Reticulatisporonites Elsik 1968 as later synonyms of Monoporisporites.
Monoporisporites jansoniusii R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901368; Facesoffungi number: FoF 16076; Fig. 165 E.
Etymology: This specific epithet honours Dr. Jan Jansonius, Geological Survey of Canada, Calgary, Alberta, Canada.
Holotype: Fig. 165 E, slide AM1b;134.8 × 12.8; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Fungal spore triangular with broadly rounded angles. Size 88 × 86 µm. Monoporate, pore angular, clearly discernible, not surrounded by pore margin; Spore wall 1–2.5 µm thick, psilate, having a few folds.
Notes: This species can be distinguished from the other species of Monoporisporites by its triangular shape and large size.
Monoporisporites pattersonii R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901369; Facesoffungi number: FoF 16077; Fig. 165 F–H
Etymology: The specific epithet honours Professor R.T. Patterson, Department of Earth Sciences, Carleton University, Ottawa, Canada.
Holotype: Fig. 165 G, slide L2c; 167 × 7; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Fungal spores spherical-subspherical in shape, Size 39‒46 × 38‒44 µm. Monoporate, pore circular, 6‒9 µm in diameter, not surrounded by wall thickening. Spore wall 1‒2.5 µm thick, reticulate.
Notes: The present species can be distinguished from the other species of Monoporisporites by its large pore and reticulate spore wall.
Monoporisporites valdiyae R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901370; Facesoffungi number: FoF 16078; Fig. 165 I–J
Etymology: The specific epithet honours Professor K.S. Valdiya, Jawaharlal Nehru Centre for Advanced Scientific Research, Bangalore, India.
Holotype: Fig. 165 I, slide AM2c;158.5 × 18; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Fungal spores oval‒elliptical in shape. Size 39‒41 × 28‒31 µm. Monoporate, pore terminal, 4‒7 µm in diameter, surrounded by 1.5‒2.5 µm thick wall thickening. Spore wall 2‒2.5 µm thick, coarsely but faintly reticulate.
Notes: The present species can be distinguished from the other species of Monoporisporites by its large pore surrounded by prominent wall thickening and coarsely reticulate spore wall.
Didymosporae
Dicellaesporites Elsik, Pollen Spores 10(2): 269 (1968)
Notes: The monotypic genus Dicellaesporites (Type: D. popovii Elsik) was proposed by Elsik (1968) for inaperturate, dicellate, uniseptate psilate fungal spores having variable shapes. Subsequently, Sheffy & Dilcher (1971) and Norris (1986) emended the diagnosis of this genus, Norris (1986) restricted this genus to dicellate, aporate spores with isopolar, equilateral cells. Kalgutkar and Jansonius (2000), however, do not agree with Norris (1986) because in many species, viz. Dicellaesporites paradoxus Ke & Shi 1978, D. inaequabilis Mart.-Hern. & Tom. -Ort. 1989, D. keralensis Kumar 1990, the spores are aporate, dicellate with unequal cells.
Dicellaesporites plicatus R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901365; Facesoffungi number: FoF 16079; Fig. 165 K
Etymology: The species is named after its folded spore wall.
Holotype: Fig. 165 K, slide WH1d; 158.5 × 8; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Fungal spore oval-elongated with one end broadly rounded and the other end pointed. Size 26‒29 × 13‒15 µm; dicellate, cells slightly unequal; single septum, 1‒1.5 μm thick; spore wall ca. 1‒2 μm thick, irregularly folded, folds thin and delicate.
Notes: This species can be distinguished from the other species of the genus by its spore wall having thin, delicate irregular folds.
Dicellaesporites verrucatus R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901366; Facesoffungi number: FoF 16080; Fig. 165L
Etymology: The species is named after its verrucate spore wall.
Holotype: Fig. 165L, slide L3c; 135.3 × 18; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Fungal spore elliptical with both ends narrowly rounded. Size 65 × 38 µm. Dicellate, cells equal in size. Single septum, 3 μm thick; spore wall ca. 1‒2 μm thick, verrucate, verrucae low and flat, irregular in shape but evenly distributed.
Notes: This species can be distinguished from the other species of the genus by its verrucate spore wall.
Phragmosporae
Alleppeysporonites Ramanujam & K.P. Rao 1978
Notes: Ramanujam and Rao (1978) commented that the branched nature and the presence of appendages are the important features of this spore type and that the fossil taxon shows a striking similarity to the dematiaceous fungus Grallomyces F. Stevens (Barnett 1956; Ellis 1971; Subramanian 1971). The appendages of the fossil are similar to the stalked attachment organs of Grallomyces conidia. The spore wall in Grallomyces is minutely verrucate whereas in the fossil spores, it is psilate to scabrate. Grallomyces is common in moist tropics. Ramanujam and Rao (1978) illustrated two specimens of the type species (A. scabratus Ramanujam & Rao 1978). Of these, the holotype (pl. 3, Fig. 40) complies with the specific diagnosis whereas the other specimen (pl. 3, Fig. 41) appears to be different.
Alleppeysporonites elsikii R.K. Saxena & A. Kumar sp. nov. (Fossil)
Index Fungorum number: IF 901364; Facesoffungi number: FoF 16081; Fig. 165M
Etymology: This specific epithet honours Dr. William C. Elsik, Exxon Oil Company, Houston, Texas, U.S.A.
Holotype: Fig. 165M, slide M1c; 160.8 × 16.5; Repository: Carleton Climate and Environment Research Group, Department of Earth Sciences, Carleton University, Ottawa, Canada. Location: West of Abha town on the southern Red Sea coast of Saudi Arabia between Jeddah in the north and Jizan in the south (Fig. 164).
Diagnosis: Spores light brown. Size 95 µm. Branched, branches straight, uniformly wide throughout their length and of varying lengths, three per spore, multicellate, transverse septa 2‒6 per branch, cells rectangular, 6 × 2 μm, nonaperturate. Spore wall thin, psilate. Appendages are simple, 12‒23 μm long.
Notes: The present species differs from Alleppeysporonites scabratus Ramanujam & K.P. Rao 1978 in having psilate spore wall and fewer cells per branch.
Correction to: Fungal diversity notes 1611–1716: taxonomic and phylogenetic contributions on fungal genera and species emphasis in south China.
Figure 90 in “Fungal diversity notes 1611–1716” should be illustrate the morphology of Tarzetta tibetensis. However, it was mistakenly replaced by the phylogenetic analyses of Tarzetta. Therefore, here we provide the morphology of Tarzetta tibetensis (Fig. 166).
Table 1. List of higher taxa with notes in Fungal Diversity Notes from 1817 to 1
Name | Paper | Page number |
|---|---|---|
Dothideomycetes | Manawasinghe et al. (2024) (this study) | 9 |
Botryosphaeriales | Manawasinghe et al. (2024) (this study) | 9 |
Botryosphaeriaceae | Manawasinghe et al. (2024) (this study) | 9 |
Dyfrolomycetales | Manawasinghe et al. (2024) (this study) | 10 |
Pleurotremataceae | Manawasinghe et al. (2024) (this study) | 10 |
Kirschsteiniotheliales | Manawasinghe et al. (2024) (this study) | 13 |
Kirschsteiniotheliaceae | Manawasinghe et al. (2024) (this study) | 13 |
Muyocopronales | Manawasinghe et al. (2024) (this study) | 15 |
Muyocopronaceae | Manawasinghe et al. (2024) (this study) | 15 |
Venturiales | Manawasinghe et al. (2024) (this study) | 19 |
Sympoventuriaceae | Manawasinghe et al. (2024) (this study) | 20 |
Dothideales | Manawasinghe et al. (2024) (this study) | 21 |
Dothideaceae | Manawasinghe et al. (2024) (this study) | 21 |
Mycosphaerellales | Manawasinghe et al. (2024) (this study) | 24 |
Mycosphaerellaceae | Manawasinghe et al. (2024) (this study) | 25 |
Pleosporales | Manawasinghe et al. (2024) (this study) | 27 |
Acrocalymmaceae | Manawasinghe et al. (2024) (this study) | 27 |
Corynesporascaceae | Manawasinghe et al. (2024) (this study) | 30 |
Didymellaceae | Manawasinghe et al. (2024) (this study) | 33 |
Didymosphaeriaceae | Manawasinghe et al. (2024) (this study) | 36 |
Dictyosporiaceae | Manawasinghe et al. (2024) (this study) | 38 |
Macrodiplodiopsidaceae | Manawasinghe et al. (2024) (this study) | 40 |
Pseudochaetosphaeronema | Manawasinghe et al. (2024) (this study) | 40 |
Melanommataceae | Manawasinghe et al. (2024) (this study) | 42 |
Roussoellaceae | Manawasinghe et al. (2024) (this study) | 45 |
Tetraplosphaeriaceae | Manawasinghe et al. (2024) (this study) | 47 |
Eurotiomycetes | Manawasinghe et al. (2024) (this study) | 51 |
Chaetothyriales | Manawasinghe et al. (2024) (this study) | 51 |
Trichomeriaceae | Manawasinghe et al. (2024) (this study) | 51 |
Eurotiales | Manawasinghe et al. (2024) (this study) | 52 |
Aspergillaceae | Manawasinghe et al. (2024) (this study) | 52 |
Lecanoromycetes | Manawasinghe et al. (2024) (this study) | 68 |
Graphidales | Manawasinghe et al. (2024) (this study) | 68 |
Diploschistaceae | Manawasinghe et al. (2024) (this study) | 68 |
Graphidaceae | Manawasinghe et al. (2024) (this study) | 69 |
Ostropales | Manawasinghe et al. (2024) (this study) | 75 |
Spirographaceae | Manawasinghe et al. (2024) (this study) | 75 |
Stictidaceae | Manawasinghe et al. (2024) (this study) | 79 |
Pertusariales | Manawasinghe et al. (2024) (this study) | 83 |
Megasporaceae | Manawasinghe et al. (2024) (this study) | 83 |
Leotiomycetes | Manawasinghe et al. (2024) (this study) | 86 |
Helotiaceae | Manawasinghe et al. (2024) (this study) | 86 |
Vibrisseaceae | Manawasinghe et al. (2024) (this study) | 86 |
Orbiliomycetes | Manawasinghe et al. (2024) (this study) | 88 |
Orbiliales | Manawasinghe et al. (2024) (this study) | 88 |
Orbiliaceae | Manawasinghe et al. (2024) (this study) | 89 |
Sordariomycetes | Manawasinghe et al. (2024) (this study) | 91 |
Diaporthales | Manawasinghe et al. (2024) (this study) | 91 |
Diaporthaceae | Manawasinghe et al. (2024) (this study) | 91 |
Tubakiaceae | Manawasinghe et al. (2024) (this study) | 93 |
Distoseptisporales | Manawasinghe et al. (2024) (this study) | 98 |
Distoseptisporaceae | Manawasinghe et al. (2024) (this study) | 98 |
Papulosaceae | Manawasinghe et al. (2024) (this study) | |
Hypocreomycetidae | Manawasinghe et al. (2024) (this study) | 105 |
Glomerellales | Manawasinghe et al. (2024) (this study) | |
Glomerellaceae | Manawasinghe et al. (2024) (this study) | 105 |
Hypocreales | Manawasinghe et al. (2024) (this study) | 101 |
Bionectriaceae | Manawasinghe et al. (2024) (this study) | 101 |
Nectriaceae | Manawasinghe et al. (2024) (this study) | 114 |
Ophiocordycipitaceae | Manawasinghe et al. (2024) (this study) | 115 |
Microascales | Manawasinghe et al. (2024) (this study) | 118 |
Halosphaeriaceae | Manawasinghe et al. (2024) (this study) | 118 |
Savoryellomycetidae | Manawasinghe et al. (2024) (this study) | 121 |
Fuscosporellales | Manawasinghe et al. (2024) (this study) | 121 |
Fuscosporellaceae | Manawasinghe et al. (2024) (this study) | 121 |
Pleurotheciales | Manawasinghe et al. (2024) (this study) | 123 |
Pleurotheciaceae | Manawasinghe et al. (2024) (this study) | 123 |
Savoryellales | Manawasinghe et al. (2024) (this study) | 125 |
Savoryellaceae | Manawasinghe et al. (2024) (this study) | 125 |
Sordariomycetidae | Manawasinghe et al. (2024) (this study) | 128 |
Pseudodactylariales | Manawasinghe et al. (2024) (this study) | 128 |
Pseudodactylariaceae | Manawasinghe et al. (2024) (this study) | 128 |
Sordariales | Manawasinghe et al. (2024) (this study) | 131 |
Chaetomiaceae | Manawasinghe et al. (2024) (this study) | 131 |
Xylariomycetidae | Manawasinghe et al. (2024) (this study) | 133 |
Amphisphaeriales | Manawasinghe et al. (2024) (this study) | 133 |
Apiosporaceae | Manawasinghe et al. (2024) (this study) | 133 |
Oxydothidaceae | Manawasinghe et al. (2024) (this study) | 139 |
Xylariales | Manawasinghe et al. (2024) (this study) | 141 |
Cainiaceae | Manawasinghe et al. (2024) (this study) | 141 |
Xylariaceae | Manawasinghe et al. (2024) (this study) | 145 |
Zygosporiaceae | Manawasinghe et al. (2024) (this study) | 145 |
Basidiomycota | Manawasinghe et al. (2024) (this study) | 151 |
Agaricomycotina | Manawasinghe et al. (2024) (this study) | 151 |
Agaricales | Manawasinghe et al. (2024) (this study) | 151 |
Agaricaceae | Manawasinghe et al. (2024) (this study) | 151 |
Hygrophoraceae | Manawasinghe et al. (2024) (this study) | 155 |
Marasmiaceae | Manawasinghe et al. (2024) (this study) | 163 |
Omphalotaceae | Manawasinghe et al. (2024) (this study) | 165 |
Pleurotaceae | Manawasinghe et al. (2024) (this study) | 169 |
Boletales | Manawasinghe et al. (2024) (this study) | 173 |
Boletaceae | Manawasinghe et al. (2024) (this study) | 173 |
Corticiales | Manawasinghe et al. (2024) (this study) | 177 |
Corticiaceae | Manawasinghe et al. (2024) (this study) | 177 |
Hymenochaetales | Manawasinghe et al. (2024) (this study) | 179 |
Hymenochaetaceae | Manawasinghe et al. (2024) (this study) | 181 |
Polyporales | Manawasinghe et al. (2024) (this study) | 190 |
Ganodermataceae | Manawasinghe et al. (2024) (this study) | 190 |
Phanerochaetaceae | Manawasinghe et al. (2024) (this study) | 192 |
Russulales | Manawasinghe et al. (2024) (this study) | 195 |
Peniophoraceae | Manawasinghe et al. (2024) (this study) | 195 |
Pucciniomycetes | Manawasinghe et al. (2024) (this study) | 196 |
Pucciniales | Manawasinghe et al. (2024) (this study) | 196 |
Gymnosporangiaceae | Manawasinghe et al. (2024) (this study) | 196 |
Ustilaginomycetes | Manawasinghe et al. (2024) (this study) | 197 |
Ustilaginales | Manawasinghe et al. (2024) (this study) | 197 |
Ustilaginaceae | Manawasinghe et al. (2024) (this study) | 197 |
Glomeromycota | Manawasinghe et al. (2024) (this study) | 201 |
Glomeromycetes | Manawasinghe et al. (2024) (this study) | 201 |
Gigasporales | Manawasinghe et al. (2024) (this study) | 201 |
Gigasporaceae | Manawasinghe et al. (2024) (this study) | 201 |
Mucoromycota | Manawasinghe et al. (2024) (this study) | 204 |
Mucormycetes | Manawasinghe et al. (2024) (this study) | 204 |
Mucorales | Manawasinghe et al. (2024) (this study) | 204 |
Mucoraceae | Manawasinghe et al. (2024) (this study) | 204 |
Umbelopsidomycetes | Manawasinghe et al. (2024) (this study) | 209 |
Umbelopsidales | Manawasinghe et al. (2024) (this study) | 209 |
Umbelopsidaceae | Manawasinghe et al. (2024) (this study) | 209 |
Fungus-like taxa | Manawasinghe et al. (2024) (this study) | 214 |
Myxomycetes | Manawasinghe et al. (2024) (this study) | 211 |
Physarales | Manawasinghe et al. (2024) (this study) | 211 |
Lamprodermataceae | Manawasinghe et al. (2024) (this study) | 211 |
Lamproderma | Manawasinghe et al. (2024) (this study) | 211 |
Dothideomycetes | Liu et al. (2024) | 6 |
Muyocopronales | Liu et al. (2024) | 6 |
Muyocopronaceae | Liu et al. (2024) | 6 |
Pleosporales | Liu et al. (2024) | 8 |
Lentitheciaceae | Liu et al. (2024) | 8 |
Nigrogranaceae | Liu et al. (2024) | 12 |
Pleosporaceae | Liu et al. (2024) | 17 |
Trypetheliales | Liu et al. (2024) | 20 |
Trypetheliaceae | Liu et al. (2024) | 20 |
Tubeufiales | Liu et al. (2024) | 22 |
Tubeufiaceae | Liu et al. (2024) | 22 |
Eurotiales | Liu et al. (2024) | 25 |
Thermoascaceae | Liu et al. (2024) | 25 |
Laboulbeniales | Liu et al. (2024) | 29 |
Laboulbeniaceae | Liu et al. (2024) | 29 |
Lecanoromycetes | Liu et al. (2024) | 35 |
Lecanorales | Liu et al. (2024) | 35 |
Parmeliaceae | Liu et al. (2024) | 35 |
Leotiomycetes | Liu et al. (2024) | 37 |
Helotiales | Liu et al. (2024) | 38 |
Rutstroemiaceae | Liu et al. (2024) | 38 |
Sordariomycetes | Liu et al. (2024) | 42 |
Amphisphaeriales | Liu et al. (2024) | 42 |
Amphisphaeriaceae | Liu et al. (2024) | 42 |
Beltraniaceae | Liu et al. (2024) | 44 |
Hypocreales | Liu et al. (2024) | 49 |
Bionectriaceae | Liu et al. (2024) | 49 |
Cordycipitaceae | Liu et al. (2024) | 51 |
Nectriaceae | Liu et al. (2024) | 52 |
Microascales | Liu et al. (2024) | 58 |
Halosphaeriaceae | Liu et al. (2024) | 58 |
Pleurotheciales | Liu et al. (2024) | 59 |
Pleurotheciaceae | Liu et al. (2024) | 59 |
Agaricales | Liu et al. (2024) | 64 |
Agaricaceae | Liu et al. (2024) | 64 |
Cortinariaceae | Liu et al. (2024) | 74 |
Cystostereaceae | Liu et al. (2024) | 87 |
Lyophyllaceae | Liu et al. (2024) | 92 |
Marasmiaceae | Liu et al. (2024) | 94 |
Pleurotaceae | Liu et al. (2024) | 107 |
Psathyrellaceae | Liu et al. (2024) | 110 |
Atheliales | Liu et al. (2024) | 115 |
Atheliaceae | Liu et al. (2024) | 115 |
Byssocorticiaceae | Liu et al. (2024) | 117 |
Boletales | Liu et al. (2024) | 120 |
Boletaceae | Liu et al. (2024) | 121 |
Cantharellales | Liu et al. (2024) | 130 |
Botryobasidiaceae | Liu et al. (2024) | 130 |
Gomphales | Liu et al. (2024) | 132 |
Gomphaceae | Liu et al. (2024) | 133 |
Hymenochaetales | Liu et al. (2024) | 136 |
Chaetoporellaceae | Liu et al. (2024) | 136 |
Hymenochaetaceae | Liu et al. (2024) | 140 |
Schizoporaceae | Liu et al. (2024) | 147 |
Polyporales | Liu et al. (2024) | 160 |
Irpicaceae | Liu et al. (2024) | 160 |
Laetiporaceae | Liu et al. (2024) | 162 |
Meruliaceae | Liu et al. (2024) | 163 |
Polyporaceae | Liu et al. (2024) | 166 |
Russulales | Liu et al. (2024) | 171 |
Hericiaceae | Liu et al. (2024) | 172 |
Russulaceae | Liu et al. (2024) | 173 |
Trechisporales | Liu et al. (2024) | 189 |
Hydnodontaceae | Liu et al. (2024) | 189 |
Dyfrolomycetales | Senanayake et al. (2023) | 167 |
Pleurotremataceae | Senanayake et al. (2023) | 167 |
Hysteriales | Senanayake et al. (2023) | 169 |
Hysteriaceae | Senanayake et al. (2023) | 169 |
Jahnulales | Senanayake et al. (2023) | 173 |
Aliquandostipitaceae | Senanayake et al. (2023) | 173 |
Kirschsteiniotheliaceae | Senanayake et al. (2023) | 174 |
Patellariales | Senanayake et al. (2023) | 175 |
Patellariaceae | Senanayake et al. (2023) | 175 |
Pleosporales | Senanayake et al. (2023) | 177 |
Didymellaceae | Senanayake et al. (2023) | 177 |
Didymosphaeriaceae | Senanayake et al. (2023) | 188 |
Ageratinicolaceae | Senanayake et al. (2023) | 193 |
Lentitheciaceae | Senanayake et al. (2023) | 196 |
Nigrogranaceae | Senanayake et al. (2023) | 199 |
Occultibambusaceae | Senanayake et al. (2023) | 199 |
Parabambusicolaceae | Senanayake et al. (2023) | 203 |
Pleosporaceae | Senanayake et al. (2023) | 211 |
Tetraplosphaeriaceae | Senanayake et al. (2023) | 213 |
Thyridariaceae | Senanayake et al. (2023) | 221 |
Abrothallales | Senanayake et al. (2023) | 223 |
Lichenoconiaceae | Senanayake et al. (2023) | 223 |
Botryosphaeriaceae | Senanayake et al. (2023) | 229 |
Botryosphaeriales | Senanayake et al. (2023) | 229 |
Trypetheliales | Senanayake et al. (2023) | 240 |
Polycoccaceae | Senanayake et al. (2023) | 240 |
Geoglossales | Senanayake et al. (2023) | 252 |
Geoglossaceae | Senanayake et al. (2023) | 253 |
Laboulbeniales | Senanayake et al. (2023) | 255 |
Laboulbeniaceae | Senanayake et al. (2023) | 255 |
Ostropales | Senanayake et al. (2023) | 264 |
Stictidaceae | Senanayake et al. (2023) | 264 |
Pezizellaceae | Senanayake et al. (2023) | 268 |
Pezizales | Senanayake et al. (2023) | 272 |
Sarcosomataceae | Senanayake et al. (2023) | 273 |
Tarzettaceae | Senanayake et al. (2023) | 273 |
Coniochaetales | Senanayake et al. (2023) | 277 |
Coniochaetaceae | Senanayake et al. (2023) | 278 |
Diaporthales | Senanayake et al. (2023) | 282 |
Cryphonectriaceae | Senanayake et al. (2023) | 282 |
Cytosporaceae | Senanayake et al. (2023) | 283 |
Diaporthaceae | Senanayake et al. (2023) | 285 |
Togniniales | Senanayake et al. (2023) | 297 |
Togniniaceae | Senanayake et al. (2023) | 297 |
Glomerellales | Senanayake et al. (2023) | 299 |
Plectosphaerellaceae | Senanayake et al. (2023) | 299 |
Bionectriaceae | Senanayake et al. (2023) | 301 |
Hypocreaceae | Senanayake et al. (2023) | 306 |
Microascaceae | Senanayake et al. (2023) | 309 |
Chaetosphaeriales | Senanayake et al. (2023) | 309 |
Chaetosphaeriaceae | Senanayake et al. (2023) | 310 |
Amphisphaeriales | Senanayake et al. (2023) | 311 |
Apiosporaceae | Senanayake et al. (2023) | 311 |
Xylariales | Senanayake et al. (2023) | 323 |
Cainiaceae | Senanayake et al. (2023) | 323 |
Diatrypaceae | Senanayake et al. (2023) | 325 |
Oxydothidaceae | Senanayake et al. (2023) | 330 |
Xylariaceae | Senanayake et al. (2023) | 336 |
Agaricales | Senanayake et al. (2023) | 338 |
Agaricaceae | Senanayake et al. (2023) | 338 |
Hydnangiaceae | Senanayake et al. (2023) | 344 |
Inocybaceae | Senanayake et al. (2023) | 349 |
Lyophyllaceae | Senanayake et al. (2023) | 355 |
Porotheleaceae | Senanayake et al. (2023) | 358 |
Hymenochaetales | Senanayake et al. (2023) | 369 |
Hymenochaetaceae | Senanayake et al. (2023) | 369 |
Polyporales | Senanayake et al. (2023) | 373 |
Ganodermataceae | Senanayake et al. (2023) | 373 |
Peronosporales | Senanayake et al. (2023) | 376 |
Peronosporaceae | Senanayake et al. (2023) | 376 |
Mycosphaerellales | Jayawardena et al. (2022) | 7 |
Mycosphaerellaceae | Jayawardena et al. (2022) | 7 |
Kirschsteiniotheliales | Jayawardena et al. (2022) | 9 |
Kirschsteiniotheliaceae | Jayawardena et al. (2022) | 9 |
Amorosiaceae | Jayawardena et al. (2022) | 18 |
Bambusicolaceae | Jayawardena et al. (2022) | 21 |
Coniothyriaceae | Jayawardena et al. (2022) | 26 |
Didymellaceae | Jayawardena et al. (2022) | 28 |
Didymosphaeriaceae | Jayawardena et al. (2022) | 32 |
Lindgomycetaceae | Jayawardena et al. (2022) | 33 |
Lophiostomataceae | Jayawardena et al. (2022) | 39 |
Phaeoseptaceae | Jayawardena et al. (2022) | 45 |
Phaeosphaeriaceae | Jayawardena et al. (2022) | 48 |
Pleosporaceae | Jayawardena et al. (2022) | 53 |
Tetraplosphaeriaceae | Jayawardena et al. (2022) | 58 |
Torulaceae | Jayawardena et al. (2022) | 62 |
Periconiaceae | Jayawardena et al. (2022) | 65 |
Tubeufiaceae | Jayawardena et al. (2022) | 71 |
Wiesneriomycetaceae | Jayawardena et al. (2022) | 71 |
Asterinales | Jayawardena et al. (2022) | 77 |
Asterinaceae | Jayawardena et al. (2022) | 78 |
Botryosphaeriales | Jayawardena et al. (2022) | 82 |
Aplosporellaceae | Jayawardena et al. (2022) | 82 |
Laboulbeniales | Jayawardena et al. (2022) | 86 |
Laboulbeniaceae | Jayawardena et al. (2022) | 86 |
Caliciaceae | Jayawardena et al. (2022) | 90 |
Phacidiales | Jayawardena et al. (2022) | 94 |
Phacidiaceae | Jayawardena et al. (2022) | 95 |
Diaporthales | Jayawardena et al. (2022) | 95 |
Diaporthaceae | Jayawardena et al. (2022) | 97 |
Melanconiellaceae | Jayawardena et al. (2022) | 103 |
Pararamichloridiales | Jayawardena et al. (2022) | 105 |
Pararamichloridiaceae | Jayawardena et al. (2022) | 105 |
Distoseptisporales | Jayawardena et al. (2022) | 108 |
Distoseptisporaceae | Jayawardena et al. (2022) | 108 |
Glomerellales | Jayawardena et al. (2022) | 112 |
Hypocreales | Jayawardena et al. (2022) | 118 |
Nectriaceae | Jayawardena et al. (2022) | 118 |
Microascaceae | Jayawardena et al. (2022) | 123 |
Coniochaetales | Jayawardena et al. (2022) | 137 |
Coniochaetaceae | Jayawardena et al. (2022) | 137 |
Pleurotheciales | Jayawardena et al. (2022) | 139 |
Pleurotheciaceae | Jayawardena et al. (2022) | 139 |
Amphisphaeriales | Jayawardena et al. (2022) | 143 |
Sporocadaceae | Jayawardena et al. (2022) | 146 |
Apiosporaceae | Jayawardena et al. (2022) | 151 |
Xylariales | Jayawardena et al. (2022) | 154 |
Diatrypaceae | Jayawardena et al. (2022) | 154 |
Hypoxylaceae | Jayawardena et al. (2022) | 156 |
Xylariaceae | Jayawardena et al. (2022) | 162 |
Agaricaceae | Jayawardena et al. (2022) | 166 |
Atheliales | Jayawardena et al. (2022) | 175 |
Atheliaceae | Jayawardena et al. (2022) | 175 |
Hymenochaetales | Jayawardena et al. (2022) | 176 |
Hymenochaetaceae | Jayawardena et al. (2022) | 177 |
Hymenogastraceae | Jayawardena et al. (2022) | 186 |
Marasmiaceae | Jayawardena et al. (2022) | 188 |
Physalacriaceae | Jayawardena et al. (2022) | 192 |
Polyporales | Jayawardena et al. (2022) | 194 |
Polyporaceae | Jayawardena et al. (2022) | 195 |
Thelephorales | Jayawardena et al. (2022) | 201 |
Thelephoraceae | Jayawardena et al. (2022) | 202 |
Ustilaginales | Jayawardena et al. (2022) | 216 |
Ustilaginaceae | Jayawardena et al. (2022) | 216 |
Botryosphaeriales | Boonmee et al. (2021) | 8 |
Botryosphaeriaceae | Boonmee et al. (2021) | 8 |
Dissoconiaceae | Boonmee et al. (2021) | 9 |
Dothideaceae | Boonmee et al. (2021) | 14 |
Pleurotremataceae | Boonmee et al. (2021) | 17 |
Jahnulales | Boonmee et al. (2021) | 17 |
Aliquandostipitaceae | Boonmee et al. (2021) | 20 |
Minutisphaerales | Boonmee et al. (2021) | 22 |
Minutisphaeraceae | Boonmee et al. (2021) | 22 |
Acrocalymmaceae | Boonmee et al. (2021) | 22 |
Dictyosporiaceae | Boonmee et al. (2021) | 26 |
Camarosporidiellaceae | Boonmee et al. (2021) | 28 |
Cryptocoryneaceae | Boonmee et al. (2021) | 30 |
Didymellaceae | Boonmee et al. (2021) | 34 |
Didymosphaeriaceae | Boonmee et al. (2021) | 37 |
Hermatomycetaceae | Boonmee et al. (2021) | 45 |
Lentitheciaceae | Boonmee et al. (2021) | 49 |
Lindgomycetaceae | Boonmee et al. (2021) | 53 |
Longipedicellataceae | Boonmee et al. (2021) | 58 |
Lophiostomataceae | Boonmee et al. (2021) | 62 |
Lophiotremataceae | Boonmee et al. (2021) | 69 |
Macrodiplodiopsidaceae | Boonmee et al. (2021) | 76 |
Nigrogranaceae | Boonmee et al. (2021) | 83 |
Occultibambusaceae | Boonmee et al. (2021) | 83 |
Paradictyoarthriniaceae | Boonmee et al. (2021) | 86 |
Phaeoseptaceae | Boonmee et al. (2021) | 90 |
Phaeosphaeriaceae | Boonmee et al. (2021) | 90 |
Pleosporaceae | Boonmee et al. (2021) | 95 |
Testudinaceae | Boonmee et al. (2021) | 98 |
Tetraplosphaeriaceae | Boonmee et al. (2021) | 102 |
Thyridariaceae | Boonmee et al. (2021) | 102 |
Torulaceae | Boonmee et al. (2021) | 106 |
Wicklowiaceae | Boonmee et al. (2021) | 117 |
Tubeufiaceae | Boonmee et al. (2021) | 121 |
Herpotrichiellaceae | Boonmee et al. (2021) | 140 |
Dactylosporaceae | Boonmee et al. (2021) | 147 |
Erysiphaceae | Boonmee et al. (2021) | 156 |
Diaporthales | Boonmee et al. (2021) | 162 |
Coryneaceae | Boonmee et al. (2021) | 162 |
Fuscosporellaceae | Boonmee et al. (2021) | 175 |
Nectriaceae | Boonmee et al. (2021) | 178 |
Pleurotheciaceae | Boonmee et al. (2021) | 188 |
Pseudodactylariaceae | Boonmee et al. (2021) | 196 |
Savoryellaceae | Boonmee et al. (2021) | 199 |
Lasiosphaeriaceae | Boonmee et al. (2021) | 202 |
Juncigenaceae | Boonmee et al. (2021) | 204 |
Xylariales | Boonmee et al. (2021) | 207 |
Diatrypaceae | Boonmee et al. (2021) | 210 |
Xylariaceae | Boonmee et al. (2021) | 213 |
Agaricales | Boonmee et al. (2021) | 222 |
Agaricaceae | Boonmee et al. (2021) | 222 |
Cortinariaceae | Boonmee et al. (2021) | 234 |
Hydnangiaceae | Boonmee et al. (2021) | 242 |
Hygrophoraceae | Boonmee et al. (2021) | 247 |
Marasmiaceae | Boonmee et al. (2021) | 249 |
Mycenaceae | Boonmee et al. (2021) | 257 |
Nidulariaceae | Boonmee et al. (2021) | 261 |
Psathyrellaceae | Boonmee et al. (2021) | 266 |
Geastraceae | Boonmee et al. (2021) | 269 |
Schizoporaceae | Boonmee et al. (2021) | 274 |
Thelephoraceae | Boonmee et al. (2021) | 276 |
Hyphodermataceae | Boonmee et al. (2021) | 280 |
Irpicaceae | Boonmee et al. (2021) | 284 |
Phanerochaetaceae | Boonmee et al. (2021) | 288 |
Polyporaceae | Boonmee et al. (2021) | 289 |
Russulaceae | Boonmee et al. (2021) | 291 |
Capnodiales | Hyde et al. (2020a) | 13 |
Dissoconiaceae | Hyde et al. (2020a) | 13 |
Mycosphaerellaceae | Hyde et al. (2020a) | 15 |
Hysteriales | Hyde et al. (2020a) | 18 |
Hysteriaceae | Hyde et al. (2020a) | 18 |
Pleosporales | Hyde et al. (2020a) | 24 |
Amorosiaceae | Hyde et al. (2020a) | 24 |
Camarosporidiellaceae | Hyde et al. (2020a) | 25 |
Coniothyriaceae | Hyde et al. (2020a) | 28 |
Dictyosporiaceae | Hyde et al. (2020a) | 30 |
Didymellaceae | Hyde et al. (2020a) | 35 |
Didymosphaeriaceae | Hyde et al. (2020a) | 45 |
Fuscostagonosporaceae | Hyde et al. (2020a) | 57 |
Halotthiaceae | Hyde et al. (2020a) | 61 |
Lentitheciaceae | Hyde et al. (2020a) | 63 |
Leptosphaeriaceae | Hyde et al. (2020a) | 71 |
Lophiostomataceae | Hyde et al. (2020a) | 73 |
Macrodiplodiopsidaceae | Hyde et al. (2020a) | 75 |
Melanommataceae | Hyde et al. (2020a) | 78 |
Parabambusicolaceae | Hyde et al. (2020a) | 90 |
Periconiaceae | Hyde et al. (2020a) | 91 |
Phaeosphaeriaceae | Hyde et al. (2020a) | 93 |
Tetraplosphaeriaceae | Hyde et al. (2020a) | 112 |
Torulaceae | Hyde et al. (2020a) | 123 |
Trematosphaeriaceae | Hyde et al. (2020a) | 126 |
Botryosphaeriaceae | Hyde et al. (2020a) | 129 |
Muyocopronales | Hyde et al. (2020a) | 134 |
Muyocopronaceae | Hyde et al. (2020a) | 134 |
Tubeufiales | Hyde et al. (2020a) | 137 |
Tubeufiaceae | Hyde et al. (2020a) | 139 |
Venturiales | Hyde et al. (2020a) | 142 |
Sympoventuriaceae | Hyde et al. (2020a) | 142 |
Eriomycetaceae | Hyde et al. (2020a) | 146 |
Lecanoromycetes | Hyde et al. (2020a) | 148 |
Caliciales | Hyde et al. (2020a) | 150 |
Caliciaceae | Hyde et al. (2020a) | 150 |
Leotiomycetes | Hyde et al. (2020a) | 150 |
Helotiales | Hyde et al. (2020a) | 152 |
Heterosphaeriaceae | Hyde et al. (2020a) | 152 |
Ploettnerulaceae | Hyde et al. (2020a) | 153 |
Vibrisseaceae | Hyde et al. (2020a) | 156 |
Pezizomycetes | Hyde et al. (2020a) | 159 |
Pezizales | Hyde et al. (2020a) | 159 |
Helvellaceae | Hyde et al. (2020a) | 159 |
Pyronemataceae | Hyde et al. (2020a) | 160 |
Sordariomycetes | Hyde et al. (2020a) | 163 |
Diaporthales | Hyde et al. (2020a) | 163 |
Cryphonectriaceae | Hyde et al. (2020a) | 165 |
Myrmecridiaceae | Hyde et al. (2020a) | 187 |
Phomatosporales | Hyde et al. (2020a) | 187 |
Phomatosporaceae | Hyde et al. (2020a) | 187 |
Glomerellales | Hyde et al. (2020a) | 191 |
Glomerellaceae | Hyde et al. (2020a) | 191 |
Hypocreales | Hyde et al. (2020a) | 195 |
Bionectriaceae | Hyde et al. (2020a) | 195 |
Nectriaceae | Hyde et al. (2020a) | 203 |
Conioscyphales | Hyde et al. (2020a) | 204 |
Conioscyphaceae | Hyde et al. (2020a) | 204 |
Pleurotheciales | Hyde et al. (2020a) | 207 |
Pleurotheciaceae | Hyde et al. (2020a) | 207 |
Savoryellales | Hyde et al. (2020a) | 207 |
Savoryellaceae | Hyde et al. (2020a) | 209 |
Coniochaetales | Hyde et al. (2020a) | 211 |
Coniochaetaceae | Hyde et al. (2020a) | 211 |
Pseudodactylariales | Hyde et al. (2020a) | 215 |
Pseudodactylariaceae | Hyde et al. (2020a) | 215 |
Amphisphaeriales | Hyde et al. (2020a) | 219 |
Apiosporaceae | Hyde et al. (2020a) | 219 |
Pseudotruncatellaceae | Hyde et al. (2020a) | 221 |
Sporocadaceae | Hyde et al. (2020a) | 224 |
Xylariales | Hyde et al. (2020a) | 226 |
Diatrypaceae | Hyde et al. (2020a) | 226 |
Fasciatisporaceae | Hyde et al. (2020a) | 227 |
Xylariaceae | Hyde et al. (2020a) | 232 |
Saccharomycetes | Hyde et al. (2020a) | 237 |
Saccharomycetales | Hyde et al. (2020a) | 237 |
Agaricales | Hyde et al. (2020a) | 244 |
Cortinariaceae | Hyde et al. (2020a) | 244 |
Auriculariales | Hyde et al. (2020a) | 253 |
Auriculariaceae | Hyde et al. (2020a) | 254 |
Umbelopsidales | Yuan et al. (2020) | 7 |
Umbelopsidaceae | Yuan et al. (2020) | 7 |
Botryosphaeriales | Yuan et al. (2020) | 10 |
Aplosporellaceae | Yuan et al. (2020) | 11 |
Botryosphaeriaceae | Yuan et al. (2020) | 11 |
Neodevriesiaceae | Yuan et al. (2020) | 15 |
Coniothyriaceae | Yuan et al. (2020) | 16 |
Didymellaceae | Yuan et al. (2020) | 20 |
Lophiostomataceae | Yuan et al. (2020) | 23 |
Paradictyoarthriniaceae | Yuan et al. (2020) | 24 |
Phaeosphaeriaceae | Yuan et al. (2020) | 29 |
Pleomassariaceae | Yuan et al. (2020) | 29 |
Roussoellaceae | Yuan et al. (2020) | 32 |
Sporormiaceae | Yuan et al. (2020) | 34 |
Tubeufiales | Yuan et al. (2020) | 40 |
Tubeufiaceae | Yuan et al. (2020) | 40 |
Graphidales | Yuan et al. (2020) | 47 |
Graphidaceae | Yuan et al. (2020) | 47 |
Erysiphales | Yuan et al. (2020) | 49 |
Erysiphaceae | Yuan et al. (2020) | 49 |
Pezizaceae | Yuan et al. (2020) | 56 |
Pyronemataceae | Yuan et al. (2020) | 70 |
Apiosporaceae | Yuan et al. (2020) | 70 |
Chaetosphaeriales | Yuan et al. (2020) | 74 |
Chaetosphaeriaceae | Yuan et al. (2020) | 74 |
Diaporthales | Yuan et al. (2020) | 75 |
Diaporthaceae | Yuan et al. (2020) | 75 |
Hypocreales | Yuan et al. (2020) | 77 |
Clavicipitaceae | Yuan et al. (2020) | 77 |
Hypocreaceae | Yuan et al. (2020) | 85 |
Ophiocordycipitaceae | Yuan et al. (2020) | 85 |
Glomerellales | Yuan et al. (2020) | 88 |
Glomerellaceae | Yuan et al. (2020) | 88 |
Junewangiaceae | Yuan et al. (2020) | 92 |
Rhamphoriaceae | Yuan et al. (2020) | 93 |
Xylariales | Yuan et al. (2020) | 97 |
Agaricales | Yuan et al. (2020) | 99 |
Agaricaceae | Yuan et al. (2020) | 101 |
Hygrophoraceae | Yuan et al. (2020) | 115 |
Amylocorticiales | Yuan et al. (2020) | 116 |
Amylocorticiaceae | Yuan et al. (2020) | 118 |
Cantharellales | Yuan et al. (2020) | 120 |
Clavulinaceae | Yuan et al. (2020) | 121 |
Gomphales | Yuan et al. (2020) | 124 |
Lentariaceae | Yuan et al. (2020) | 124 |
Hymenochaetales | Yuan et al. (2020) | 128 |
Hymenochaetaceae | Yuan et al. (2020) | 129 |
Polyporales | Yuan et al. (2020) | 136 |
Fomitopsidaceae | Yuan et al. (2020) | 136 |
Polyporaceae | Yuan et al. (2020) | 141 |
Russulales | Yuan et al. (2020) | 145 |
Russulaceae | Yuan et al. (2020) | 146 |
Thelephorales | Yuan et al. (2020) | 148 |
Thelephoraceae | Yuan et al. (2020) | 148 |
Trechisporales | Yuan et al. (2020) | 239 |
Hydnodontaceae | Yuan et al. (2020) | 240 |
Tremellales | Yuan et al. (2020) | 245 |
Phaeotremellaceae | Yuan et al. (2020) | 245 |
Pseudoberkleasmiaceae | Hyde et al. (2019) | 6 |
Arthoniales | Hyde et al. (2019) | 7 |
Lecanographaceae | Hyde et al. (2019) | 8 |
Dothideomycetes | Hyde et al. (2019) | 10 |
Teratosphaeriaceae | Hyde et al. (2019) | 10 |
Pleosporales | Hyde et al. (2019) | 11 |
Amniculicolaceae | Hyde et al. (2019) | 12 |
Amorosiaceae | Hyde et al. (2019) | 17 |
Camarosporidiellaceae | Hyde et al. (2019) | 17 |
Cucurbitariaceae | Hyde et al. (2019) | 21 |
Dictyosporiaceae | Hyde et al. (2019) | 23 |
Didymellaceae | Hyde et al. (2019) | 29 |
Hermatomycetaceae | Hyde et al. (2019) | 40 |
Lophiostomataceae | Hyde et al. (2019) | 41 |
Massariaceae | Hyde et al. (2019) | 45 |
Phaeosphaeriaceae | Hyde et al. (2019) | 49 |
Pseudoberkleasmiaceae | Hyde et al. (2019) | 59 |
Pyrenochaetopsidaceae | Hyde et al. (2019) | 63 |
Tetraplosphaeriaceae | Hyde et al. (2019) | 65 |
Torulaceae | Hyde et al. (2019) | 70 |
Trematosphaeriaceae | Hyde et al. (2019) | 72 |
Minutisphaerales | Hyde et al. (2019) | 75 |
Acrogenosporaceae | Hyde et al. (2019) | 75 |
Asterinales | Hyde et al. (2019) | 78 |
Asterinaceae | Hyde et al. (2019) | 78 |
Botryosphaeriales | Hyde et al. (2019) | 83 |
Botryosphaeriaceae | Hyde et al. (2019) | 83 |
Jahnulales | Hyde et al. (2019) | 91 |
Aliquandostipitaceae | Hyde et al. (2019) | 91 |
Tubeufiales | Hyde et al. (2019) | 93 |
Tubeufiaceae | Hyde et al. (2019) | 93 |
Chaetothyriales | Hyde et al. (2019) | 95 |
Herpotrichiellaceae | Hyde et al. (2019) | 95 |
Eurotiales | Hyde et al. (2019) | 97 |
Aspergillaceae | Hyde et al. (2019) | 97 |
Laboulbeniales | Hyde et al. (2019) | 108 |
Laboulbeniaceae | Hyde et al. (2019) | 108 |
Caliciales | Hyde et al. (2019) | 120 |
Caliciaceae | Hyde et al. (2019) | 120 |
Lecanorales | Hyde et al. (2019) | 123 |
Lecanoraceae | Hyde et al. (2019) | 123 |
Pilocarpaceae | Hyde et al. (2019) | 123 |
Ostropomycetidae | Hyde et al. (2019) | 127 |
Ostropales | Hyde et al. (2019) | 127 |
Porinaceae | Hyde et al. (2019) | 127 |
Graphidaceae | Hyde et al. (2019) | 129 |
Leotiomycetes | Hyde et al. (2019) | 130 |
Helotiales | Hyde et al. (2019) | 130 |
Chaetomellaceae | Hyde et al. (2019) | 131 |
Pezizomycetes | Hyde et al. (2019) | 132 |
Pezizales | Hyde et al. (2019) | 133 |
Ascodesmidaceae | Hyde et al. (2019) | 133 |
Diaporthomycetidae | Hyde et al. (2019) | 136 |
Diaporthales | Hyde et al. (2019) | 136 |
Gnomoniaceae | Hyde et al. (2019) | 139 |
Distoseptisporaceae | Hyde et al. (2019) | 145 |
Hypocreomycetidae | Hyde et al. (2019) | 148 |
Plectosphaerellaceae | Hyde et al. (2019) | 149 |
Hypocreales | Hyde et al. (2019) | 150 |
Cordycipitaceae | Hyde et al. (2019) | 151 |
Pleurotheciaceae | Hyde et al. (2019) | 155 |
Savoryellaceae | Hyde et al. (2019) | 159 |
Chaetosphaeriales | Hyde et al. (2019) | 159 |
Chaetosphaeriaceae | Hyde et al. (2019) | 160 |
Linocarpaceae | Hyde et al. (2019) | 169 |
Amphisphaeriales | Hyde et al. (2019) | 169 |
Amphisphaeriaceae | Hyde et al. (2019) | 169 |
Sporocadaceae | Hyde et al. (2019) | 173 |
Castanediellaceae | Hyde et al. (2019) | 175 |
Diatrypaceae | Hyde et al. (2019) | 177 |
Xylariales | Hyde et al. (2019) | 181 |
Cortinariaceae | Hyde et al. (2019) | 181 |
Psathyrellaceae | Hyde et al. (2019) | 183 |
Cantharellales | Hyde et al. (2019) | 189 |
Botryobasidiaceae | Hyde et al. (2019) | 189 |
Hymenochaetaceae | Hyde et al. (2019) | 205 |
Polyporales | Hyde et al. (2019) | 212 |
Polyporaceae | Hyde et al. (2019) | 212 |
Russulales | Hyde et al. (2019) | 215 |
Russulaceae | Hyde et al. (2019) | 215 |
Rhizophydiales | Hyde et al. (2019) | 220 |
Rhizophydium | Hyde et al. (2019) | 220 |
Teratosphaeriaceae | Phookamsak et al. (2019) | 6 |
Lentitheciaceae | Phookamsak et al. (2019) | 14 |
Leptosphaeriaceae | Phookamsak et al. (2019) | 21 |
Lophiotremataceae | Phookamsak et al. (2019) | 31 |
Occultibambusaceae | Phookamsak et al. (2019) | 36 |
Parabambusicolaceae | Phookamsak et al. (2019) | 39 |
Periconiaceae | Phookamsak et al. (2019) | 45 |
Phaeosphaeriaceae | Phookamsak et al. (2019) | 47 |
Roussoellaceae | Phookamsak et al. (2019) | 65 |
Thyridariaceae | Phookamsak et al. (2019) | 71 |
Asterinaceae | Phookamsak et al. (2019) | 75 |
Botryosphaeriaceae | Phookamsak et al. (2019) | 78 |
Muyocopronaceae | Phookamsak et al. (2019) | 84 |
Tubeufiaceae | Phookamsak et al. (2019) | 87 |
Cyphellophoraceae | Phookamsak et al. (2019) | 87 |
Herpotrichiellaceae | Phookamsak et al. (2019) | 93 |
Trichocomaceae | Phookamsak et al. (2019) | 95 |
Micropeltidaceae | Phookamsak et al. (2019) | 98 |
Lachnaceae | Phookamsak et al. (2019) | 101 |
Pezizaceae | Phookamsak et al. (2019) | 109 |
Conlariaceae | Phookamsak et al. (2019) | 113 |
Cytosporaceae | Phookamsak et al. (2019) | 115 |
Melanconiellaceae | Phookamsak et al. (2019) | 118 |
Pseudoplagiostomataceae | Phookamsak et al. (2019) | 121 |
Schizoparmaceae | Phookamsak et al. (2019) | 122 |
Distoseptisporaceae | Phookamsak et al. (2019) | 126 |
Plectosphaerellaceae | Phookamsak et al. (2019) | 131 |
Cordycipitaceae | Phookamsak et al. (2019) | 133 |
Hypocreaceae | Phookamsak et al. (2019) | 135 |
Savoryellaceae | Phookamsak et al. (2019) | 143 |
Chaetosphaeriaceae | Phookamsak et al. (2019) | 144 |
Coniochaetaceae | Phookamsak et al. (2019) | 150 |
Phyllachoraceae | Phookamsak et al. (2019) | 152 |
Lasiosphaeriaceae | Phookamsak et al. (2019) | 155 |
Amphisphaeriaceae | Phookamsak et al. (2019) | 158 |
Sporocadaceae | Phookamsak et al. (2019) | 161 |
Diatrypaceae | Phookamsak et al. (2019) | 167 |
Agaricaceae | Phookamsak et al. (2019) | 172 |
Amanitaceae | Phookamsak et al. (2019) | 180 |
Hygrophoraceae | Phookamsak et al. (2019) | 190 |
Marasmiaceae | Phookamsak et al. (2019) | 193 |
Psathyrellaceae | Phookamsak et al. (2019) | 199 |
Boletaceae | Phookamsak et al. (2019) | 200 |
Clavulinaceae | Phookamsak et al. (2019) | 207 |
Phanerochaetaceae | Phookamsak et al. (2019) | 212 |
Russulaceae | Phookamsak et al. (2019) | 216 |
Stereaceae | Phookamsak et al. (2019) | 233 |
Hydnodontaceae | Phookamsak et al. (2019) | 239 |
Auriculariaceae | Phookamsak et al. (2019) | 242 |
Dacrymycetaceae | Phookamsak et al. (2019) | 244 |
Mycosphaerellaceae | Tibpromma et al. (2018) | 7 |
Mycosphaerellaceae | Tibpromma et al. (2018) | 7 |
Dictyosporiaceae | Tibpromma et al. (2018) | 8 |
Didymosphaeriaceae | Tibpromma et al. (2018) | 22 |
Hermatomycetaceae | Tibpromma et al. (2018) | 26 |
Melanommataceae | Tibpromma et al. (2018) | 29 |
Occultibambusaceae | Tibpromma et al. (2018) | 32 |
Pleosporaceae | Tibpromma et al. (2018) | 33 |
Roussoellaceae | Tibpromma et al. (2018) | 39 |
Tetraplosphaeriaceae | Tibpromma et al. (2018) | 42 |
Torulaceae | Tibpromma et al. (2018) | 45 |
Botryosphaeriaceae | Tibpromma et al. (2018) | 53 |
Pseudofusicoccumaceae | Tibpromma et al. (2018) | 64 |
Tubeufiaceae | Tibpromma et al. (2018) | 66 |
Sympoventuriaceae | Tibpromma et al. (2018) | 75 |
Stictidaceae | Tibpromma et al. (2018) | 77 |
Rhytismataceae | Tibpromma et al. (2018) | 79 |
Distoseptisporaceae | Tibpromma et al. (2018) | 79 |
Glomerellaceae | Tibpromma et al. (2018) | 85 |
Malaysiascaceae | Tibpromma et al. (2018) | 88 |
Plectosphaerellaceae | Tibpromma et al. (2018) | 88 |
Bionectriaceae | Tibpromma et al. (2018) | 91 |
Nectriaceae | Tibpromma et al. (2018) | 104 |
Stachybotryaceae | Tibpromma et al. (2018) | 114 |
Microascaceae | Tibpromma et al. (2018) | 118 |
Savoryellaceae | Tibpromma et al. (2018) | 122 |
Chaetosphaeriaceae | Tibpromma et al. (2018) | 125 |
Chaetomiaceae | Tibpromma et al. (2018) | 131 |
Beltraniaceae | Tibpromma et al. (2018) | 134 |
Sporocadaceae | Tibpromma et al. (2018) | 138 |
Vermiculariopsiellaceae | Tibpromma et al. (2018) | 146 |
Mycosphaerellaceae | Wanasinghe et al. (2018) | 13 |
Dothideaceae | Wanasinghe et al. (2018) | 15 |
Pleosporomycetidae | Wanasinghe et al. (2018) | 17 |
Pleosporales | Wanasinghe et al. (2018) | 17 |
Cucurbitariaceae | Wanasinghe et al. (2018) | 21 |
Didymellaceae | Wanasinghe et al. (2018) | 27 |
Lentitheciaceae | Wanasinghe et al. (2018) | 43 |
Lophiostomataceae | Wanasinghe et al. (2018) | 59 |
Melanommataceae | Wanasinghe et al. (2018) | 82 |
Paradictyoarthriniaceae | Wanasinghe et al. (2018) | 103 |
Phaeosphaeriaceae | Wanasinghe et al. (2018) | 104 |
Sporormiaceae | Wanasinghe et al. (2018) | 157 |
Teichosporaceae | Wanasinghe et al. (2018) | 159 |
Thyridariaceae | Wanasinghe et al. (2018) | 161 |
Eurotiales | Wanasinghe et al. (2018) | 171 |
Trichocomaceae | Wanasinghe et al. (2018) | 171 |
Lecanoromycetidae | Wanasinghe et al. (2018) | 175 |
Caliciaceae | Wanasinghe et al. (2018) | 177 |
Lecanoraceae | Wanasinghe et al. (2018) | 180 |
Diaporthales | Wanasinghe et al. (2018) | 180 |
Diaporthaceae | Wanasinghe et al. (2018) | 181 |
Coniocessiaceae | Wanasinghe et al. (2018) | 199 |
Coniochaetales | Wanasinghe et al. (2018) | 209 |
Backusellaceae | Wanasinghe et al. (2018) | 214 |
Mucoraceae | Wanasinghe et al. (2018) | 217 |
Aliquandostipitaceae | Hyde et al. (2017) | 6 |
Amniculicolaceae | Hyde et al. (2017) | 9 |
Amorosiaceae | Hyde et al. (2017) | 9 |
Botryosphaeriaceae | Hyde et al. (2017) | 12 |
Capnodiaceae | Hyde et al. (2017) | 14 |
Dictyosporiaceae | Hyde et al. (2017) | 18 |
Didymellaceae | Hyde et al. (2017) | 20 |
Dothideaceae | Hyde et al. (2017) | 28 |
Dyfrolomycetaceae | Hyde et al. (2017) | 32 |
Fuscostagonosporaceae | Hyde et al. (2017) | 34 |
Hermatomycetaceae | Hyde et al. (2017) | 39 |
Hysteriaceae | Hyde et al. (2017) | 40 |
Kirschsteiniotheliaceae | Hyde et al. (2017) | 45 |
Lentitheciaceae | Hyde et al. (2017) | 45 |
Lindgomycetaceae | Hyde et al. (2017) | 54 |
Micropeltidaceae | Hyde et al. (2017) | 59 |
Microthyriaceae | Hyde et al. (2017) | 59 |
Mytilinidiaceae | Hyde et al. (2017) | 64 |
Nigrogranaceae | Hyde et al. (2017) | 66 |
Periconiaceae | Hyde et al. (2017) | 70 |
Phaeothecoidiellaceae | Hyde et al. (2017) | 86 |
Pleosporaceae | Hyde et al. (2017) | 89 |
Pseudoastrosphaeriellaceae | Hyde et al. (2017) | 93 |
Pseudoperisporiaceae | Hyde et al. (2017) | 100 |
Teichosporaceae | Hyde et al. (2017) | 105 |
Torulaceae | Hyde et al. (2017) | 112 |
Vizellaceae | Hyde et al. (2017) | 119 |
Verrucariaceae | Hyde et al. (2017) | 123 |
Pyronemataceae | Hyde et al. (2017) | 130 |
Cytosporaceae | Hyde et al. (2017) | 133 |
Diatrypaceae | Hyde et al. (2017) | 139 |
Glomerellaceae | Hyde et al. (2017) | 144 |
Meliolaceae | Hyde et al. (2017) | 148 |
Ophiocordycipitaceae | Hyde et al. (2017) | 151 |
Plectosphaerellaceae | Hyde et al. (2017) | 155 |
Pleurotheciaceae | Hyde et al. (2017) | 160 |
Stachybotryaceae | Hyde et al. (2017) | 166 |
Trichosphaeriaceae | Hyde et al. (2017) | 171 |
Xylariaceae | Hyde et al. (2017) | 173 |
Agaricaceae | Hyde et al. (2017) | 185 |
Lyophyllaceae | Hyde et al. (2017) | 189 |
Meruliaceae | Hyde et al. (2017) | 191 |
Polyporaceae | Hyde et al. (2017) | 194 |
Asterinales | Tibpromma et al. (2017) | 10 |
Botryosphaeriales | Tibpromma et al. (2017) | 11 |
Botryosphaeriaceae | Tibpromma et al. (2017) | 11 |
Mycosphaerellaceae | Tibpromma et al. (2017) | 14 |
Dothioraceae | Tibpromma et al. (2017) | 18 |
Amorosiaceae | Tibpromma et al. (2017) | 26 |
Biatrosporaceae | Tibpromma et al. (2017) | 33 |
Lophiostomataceae | Tibpromma et al. (2017) | 33 |
Lophiotremataceae | Tibpromma et al. (2017) | 40 |
Melanommataceae | Tibpromma et al. (2017) | 49 |
Roussoellaceae | Tibpromma et al. (2017) | 52 |
Latoruaceae | Tibpromma et al. (2017) | 57 |
Didymellaceae | Tibpromma et al. (2017) | 68 |
Leptosphaeriaceae | Tibpromma et al. (2017) | 82 |
Phaeosphaeriaceae | Tibpromma et al. (2017) | 87 |
Mucorales | Tibpromma et al. (2017) | 109 |
Gnomoniaceae | Tibpromma et al. (2017) | 113 |
Hypocreales | Tibpromma et al. (2017) | 124 |
Clavicipitaceae | Tibpromma et al. (2017) | 124 |
Clavulinaceae | Tibpromma et al. (2017) | 204 |
Fomitopsidaceae | Tibpromma et al. (2017) | 208 |
Asterinales | Hyde et al. (2017) | 13 |
Asterinaceae | Hyde et al. (2017) | 13 |
Botryosphaeriales | Hyde et al. (2017) | 14 |
Mycosphaerellaceae | Hyde et al. (2017) | 19 |
Dothideaceae | Hyde et al. (2017) | 30 |
Hysteriaceae | Hyde et al. (2017) | 32 |
Dictyosporiaceae | Hyde et al. (2017) | 34 |
Didymellaceae | Hyde et al. (2017) | 37 |
Lentitheciaceae | Hyde et al. (2017) | 50 |
Leptosphaeriaceae | Hyde et al. (2017) | 55 |
Lindgomycetaceae | Hyde et al. (2017) | 59 |
Lophiostomataceae | Hyde et al. (2017) | 62 |
Lophiotremataceae | Hyde et al. (2017) | 65 |
Massariaceae | Hyde et al. (2017) | 74 |
Massarinaceae | Hyde et al. (2017) | 77 |
Occultabambusaceae | Hyde et al. (2017) | 81 |
Phaeosphaeriaceae | Hyde et al. (2017) | 83 |
Pleosporaceae | Hyde et al. (2017) | 117 |
Roussoellaceae | Hyde et al. (2017) | 119 |
Torulaceae | Hyde et al. (2017) | 121 |
Trypetheliaceae | Hyde et al. (2017) | 131 |
Trichomeriaceae | Hyde et al. (2017) | 145 |
Pezizales | Hyde et al. (2017) | 149 |
Helvellaceae | Hyde et al. (2017) | 149 |
Coronophorales | Hyde et al. (2017) | 165 |
Glomerellaceae | Hyde et al. (2017) | 176 |
Reticulascaceae | Hyde et al. (2017) | 177 |
Ophiocordycipitaceae | Hyde et al. (2017) | 179 |
Meliolaceae | Hyde et al. (2017) | 191 |
Cainiaceae | Hyde et al. (2017) | 201 |
Sporidesmiaceae | Hyde et al. (2017) | 214 |
Clavariaceae | Hyde et al. (2017) | 225 |
Dothideales | Li et al. (2016) | 19 |
Dothideaceae | Li et al. (2016) | 19 |
Saccotheciaceae | Li et al. (2016) | 22 |
Hysteriaceae | Li et al. (2016) | 24 |
Didymosphaeriaceae | Li et al. (2016) | 28 |
Lentitheciaceae | Li et al. (2016) | 31 |
Melanommataceae | Li et al. (2016) | 39 |
Parabambusicolaceae | Li et al. (2016) | 44 |
Phaeosphaeriaceae | Li et al. (2016) | 46 |
Testudinaceae | Li et al. (2016) | 52 |
Tetraplosphaeriaceae | Li et al. (2016) | 55 |
Kirschsteiniotheliaceae | Li et al. (2016) | 68 |
Graphidaceae | Li et al. (2016) | 71 |
Chaetosphaeriaceae | Li et al. (2016) | 74 |
Gnomoniaceae | Li et al. (2016) | 75 |
Valsaceae | Li et al. (2016) | 77 |
Glomerellales | Li et al. (2016) | 78 |
Clavicipitaceae | Li et al. (2016) | 87 |
Ophiocordycipitaceae | Li et al. (2016) | 90 |
Halosphaeriaceae | Li et al. (2016) | 94 |
Xylariaceae | Li et al. (2016) | 105 |
Cortinariaceae | Li et al. (2016) | 141 |
Tricholomataceae | Li et al. (2016) | 151 |
Russulaceae | Li et al. (2016) | 177 |
Neocallimastigaceae | Li et al. (2016) | 187 |
Amniculicolaceae | Ariyawansa et al. (2015) | 35 |
Anteagloniaceae | Ariyawansa et al. (2015) | 45 |
Ascocylindricaceae | Ariyawansa et al. (2015) | 45 |
Caryosporaceae | Ariyawansa et al. (2015) | 54 |
Cucurbitariaceae | Ariyawansa et al. (2015) | 57 |
Didymosphaeriaceae | Ariyawansa et al. (2015) | 62 |
Floricolaceae | Ariyawansa et al. (2015) | 69 |
Halotthiaceae | Ariyawansa et al. (2015) | 71 |
Latoruaceae | Ariyawansa et al. (2015) | 77 |
Lentitheciaceae | Ariyawansa et al. (2015) | 81 |
Melanommataceae | Ariyawansa et al. (2015) | 91 |
Microthyriaceae | Ariyawansa et al. (2015) | 93 |
Phaeosphaeriaceae | Ariyawansa et al. (2015) | 97 |
Pleosporaceae | Ariyawansa et al. (2015) | 115 |
Roussoellaceae | Ariyawansa et al. (2015) | 118 |
Tetraplosphaeriaceae | Ariyawansa et al. (2015) | 122 |
Wicklowiaceae | Ariyawansa et al. (2015) | 126 |
Mycocaliciaceae | Ariyawansa et al. (2015) | 126 |
Acarosporaceae | Ariyawansa et al. (2015) | 129 |
Graphidaceae | Ariyawansa et al. (2015) | 134 |
Lobariaceae | Ariyawansa et al. (2015) | 137 |
Annulatascaceae | Ariyawansa et al. (2015) | 142 |
Cordycipitaceae | Ariyawansa et al. (2015) | 152 |
Xylariaceae | Ariyawansa et al. (2015) | 174 |
Entolomataceae | Ariyawansa et al. (2015) | 196 |
Hygrophoraceae | Ariyawansa et al. (2015) | 198 |
Inocybaceae | Ariyawansa et al. (2015) | 201 |
Boletaceae | Ariyawansa et al. (2015) | 204 |
Corticiales | Ariyawansa et al. (2015) | 217 |
Cortinariaceae | Ariyawansa et al. (2015) | 218 |
Dendrominiaceae | Ariyawansa et al. (2015) | 224 |
Neoantrodiellaceae | Ariyawansa et al. (2015) | 228 |
Mucorales | Ariyawansa et al. (2015) | 245 |
Neocallimastigales | Ariyawansa et al. (2015) | 255 |
Amphisphaeriaceae | Liu et al. (2015, b) | 7 |
Diaporthaceae | Liu et al. (2015, b) | 22 |
Diatrypaceae | Liu et al. (2015, b) | 22 |
Meliolaceae | Liu et al. (2015, b) | 39 |
Sydowiellaceae | Liu et al. (2015, b) | 39 |
Xylariaceae | Liu et al. (2015, b) | 47 |
Bambusicolaceae | Liu et al. (2015, b) | 65 |
Didymellaceae | Liu et al. (2015, b) | 79 |
Didymosphaeriaceae | Liu et al. (2015, b) | 85 |
Hysteriaceae | Liu et al. (2015, b) | 98 |
Pleomassariaceae | Liu et al. (2015, b) | 147 |
Saccotheciaceae | Liu et al. (2015, b) | 159 |
Venturiaceae | Liu et al. (2015, b) | 159 |
Chaetothyriales | Liu et al. (2015, b) | 160 |
Chaetothyriaceae | Liu et al. (2015, b) | 160 |
Trichomeriaceae | Liu et al. (2015, b) | 162 |
Squamarinaceae | Liu et al. (2015, b) | 164 |
In addition to the novel taxa and records, from the current Fungal Diversity Notes, we will provide the list of genera and higher taxa with notes in Fungal Diversity in Table 1.
Acknowledgements
Biao Xu would like to thank to National Natural Science Foundation of China (Nos. 32370021 and 31860008), the Innovative team program of the Department of Education of Guangdong Province (Nos. 2022KCXTD015 and 2022ZDJS020) and the Project of Fungi Investigation in Tomur Mountains National Nature Reserve (2021-01-139-2). Jiage Song would like to thank the National Natural Science Foundation of China (No. 32100012), and the Science and Technology Bureau of Guangzhou City (202201011618). Ishara S. Manawasinghe would like to acknowledge Zhongkai University of Agriculture and Engineering, talent funding (Grant number KA210319288) and the Guangzhou Science and Technology Plan Project (2023A04J1427). Mingkwan Doilom thanks the Key Laboratory of Green Prevention and Control on Fruits and Vegetables in South China, Guangdong (KA21031C502), and Zhongkai University of Agriculture and Engineering, Guangzhou, Guangdong, China (KA22016B746) for financial research support. MS Calabon is grateful to the UP System Balik PhD Program (OVPAA-BPhD-2022-02). Dhanushka Wanasinghe gratefully acknowledges Yunnan Department of Sciences and Technology of China (Grant No: 202101AS070045, 202205AM070007, 202302AE090023, 202303AP140001) and the financial support provided by the Distinguished Scientist Fellowship Program (DSFP) at King Saud University in Riyadh, Saudi Arabia. Raghvendra Singh thanks Science & Engineering Research Board (SERB), Department of Science & Technology (DST), Govt. of India (Scheme No. CRG/2020/006053) and Institution of Eminence (IoE) Scheme, Ministry of Human Resource and Development (MHRD), Govt. of India (No. R/Dev/D/IoE/Incentive/2021-22/32387) for providing financial support. Genivaldo Alves-Silva, Elisandro R. Drechsler-Santos, Rosa M. B. da Silveira, and Aristóteles Góes-Neto are supported by Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) (Grant No. 153025/2022-0; 310150/2022-1; 308122/2019-4; 308880/2022-6; respectively). Genivaldo Alves-Silva and Elisandro R. Drechsler-Santos would like to thank the CNPq and FAPESC under the PROTAX program (Grant No. FAPESC 2021TR390, Grant No. CNPq 441821/2020-0) and M. E. Engels for collections. The authors thank Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), Brazil, that provided research grants to B.T. Goto (proc. 306632/2022-5). The financial support from the National Science and Technology Council is acknowledged (101-2621-B-019-001-MY3). G. Moreno, Á. López-Villalba & A. Sánchez express our gratitude to S.L. Stephenson for the revision of the manuscript. We also thank L. Monje and A. Pueblas (Gabinete de Fotografía, Univ. Alcalá) for their help with the digital treatment of the photographs, J.A. Pérez and A. Priego (Servicio de Microscopía Electrónica, Univ. Alcalá) for their invaluable help with SEM, and Dr. J. Rejos, curator of the AH herbarium, for his assistance with the specimens examined in the present study. Acknowledgements are extendable to A. Klahr and O. Shchepin (University of Greifswald, Germany) since they helped get some of the sequences for the phylogenetic study and with technical, and computer problems, respectively. The molecular study was supported by Prof. Dr. M. Schnittler (University of Greifswald, Germany), through the DFG project RESPONSE (RTG2010). This study by Ralaiveloarisoa A. et al. was supported by the Today’s Flora for Tomorrow project funded by a generous donor through the Kew Foundation, and by a grant from the Bentham-Moxon Trust. The study of Teodor T. Denchev and Cvetomir M. Denchev was supported by the Bulgarian National Science Fund (Grant no. KP-06-N51/10/16.11.2021). Their visit to the herbarium at the Botanic Garden and Botanical Museum Berlin received support from the SYNTHESYS Plus Project http://www.synthesys.info, which is financed by the H2020 Research Infrastructures Programme (Grant no. DE-TAF-8193). Li Lu, Xiang-Fu Liu, Tian-Ye Du, and Rui-Fang Xu thank Mae Fah Luang University for providing tuition fee scholarship. The Center for Yunnan Plateau Biological Resources Protection and Utilization, College of Biological Resource and Food Engineering, Qujing Normal University is thanked for the facilities provided for the research work. This work was supported by the National Natural Science Foundation of China (No. 32060012) Muhammad Usman and Abdul Nasir Khalid would like to thank Dr. Kamran Habib, Dr. Muhammad Ali, Mr. Mohammad Aijaz Ahmad and Mr. Muhammad Shafiq for accompanying during the collection survey. Kunhiraman C. Rajeshkumar would like to thank the Science and Engineering Research Board (SERB), the Department of Science and Technology, Government of India, for their financial support through CRG/2020/000668 project. RKC also thank Dr. P. K. Dhakephalkar, Director of the MACS Agharkar Research Institute in Pune, for providing the lab resources and motivating us in our research work. Nattawut Boonyuen and Jennifer Luangsa-ard would like to thank “Southeast-Asia Europe Joint Funding Scheme for Research and Innovation grant for the project “Discovery of new antivirals using cultures of filamentous fungi collected in Europe and Thailand as compound sources (JFS20ST-127 Antiviralfun; P2150844)” and Veera Sri-Indrasutdhi would like to thank BIOTEC-Novartis collaboration for microbial bioprospecting project (P20-52031). Sruthi O. P. would like to express her gratitude to CSIR-HRDG, India, for providing her with financial assistance as part of the JRF fellowship (09/0670(13602)/2022-EMR-I). Andrei Tsurykau is cordially thankful to Javier Etayo (Pamplona) for his valuable suggestions. Adam Flakus is greatly indebted to all Herbario Nacional de Bolivia staff, Instituto de Ecología, Universidad Mayor de San Andrés, La Paz, for their generous long-term cooperation. Lijian Xu thanks National Science Foundation of China (No. 31870528). M. Ghobad-Nejhad acknowledges the support from Iran National Science Foundation (INSF, no. 4000655). This study was financed in part by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior—Brasil (CAPES)—Finance Code 001 who provided a visiting professorship to the first author. The Stichting Hugo de Vries Fonds kindly gave a grant for his fieldwork. Felix Schumm is warmly thanked for making the pictures of the new lichen species. Robert Lücking is warmly thanked for discussions about the new lichen species. The authors thank Conselho Nacional de Pesquisa e Desenvolvimento (CNPQ) for financially supporting this work with a PhD’s scholarship to LAS (140847/2019-7), a research grant to MESC (307569/2019-5) and for financial support in the Universal project (Process: 437097/2018-8). Daniel Guerra-Mateo and co-authors are partly financed by the Spanish Ministerio de Ciencia, Innovación y Universidades, AEI, https://doi.org/10.13039/501100011033, and ERDF-A way of making Europe (Grant PID2021-128068NB-100); and the authors thank to Gabriel Quiroga Jofre (Diving Center Tarraco) for his support in the marine sampling. Salna Nanu acknowledges the support from Kerala State Council for Science Technology and Environment in the form of a research fellowship. Salna Nanu and T.K. Arun Kumar are thankful to the Chief Conservator of Forests & Chief Wildlife Warden, Kerala, for permission for field work in the forest areas of Kerala. Rameshwar Avchar would like to thank the Department of Biotechnology (DBT), Government of India (Grant no. BT/PR/0054/NDB/52/94/2007), for financial support under the project ‘Establishment of Microbial Culture Collection (NCMR-NCCS).’ Gajanan Mane is thankful to the University Grants Commission, Delhi (India) for the senior research fellowship (File No. 16-6(Dec. 2017)/2018(NET/CSIR). Rohit Sharma thanks the Department of Biotechnology (DBT), Government of India (Grant no. BT/PR25490/NER/95/1220/2017 dated 28.06.2018), for financial support. Mei Luo would like to thank the grant from the Guangdong Rural Science and Technology Commissioner project (KTP20210313), the Key Laboratory of Green Prevention and Control on Fruits and Vegetables in South China, Guangdong (KA21031C501), and the Innovative Team program of the Department of Education of Guangdong Province (2023KCXTD018/2022KCXTD015). Malarvizhi Kaliyaperumal and Sugantha Gunaseelan thank Extramural Research-SERB, DST (EMR/2016/003078), Government of India for the financial assistance. They are also grateful to ‘The PCCF’ of the Tamil Nadu Forest Department for providing permission (E2/20458/2017), assistance and support during field visit in the Eastern Ghats. Malarvizhi Kaliyaperumal and Kezhocuyi Kezo thank RUSA 2.0 (Theme-1, Group-1/2021/49) for providing grant. Malarvizhi Kaliyaperumal and Elangovan Arumugam thank the Tamil Nadu State Council for Higher Education, Chennai (RGP/2019-20/MU/HECP-0040) for financial assistance. Lijian Xu thanks the National Science Foundation of China (No. 31870528). Adam Flakus received additional support under statutory funds from the W. Szafer Institute of Botany, Polish Academy of Sciences. The authors Luís F. P. Gusmão, Gabriel G. Barreto, Andre Rodrigues, Lorena T. Lacerda and Pepijn W. Kooij are grateful to ICMBio (Instituto Chico Mendes de Conservação da Biodiversidade) and IF (Instituto Florestal) for the collecting permits #38466-2 and #260108-001.102/2015, respectively. This study was financed in part by Coordination of Improvement of Higher Education Personnel—Brazil (CAPES)—Finance Code 001 and by the National Council for Scientific and Technological Development (to LFPG and Proc. 305269/2018-6 to AR) to LFPG and Proc. 305269/2018-6 to AR). The contribution to this paper by PK was possible thanks to the scholarship granted from CAPES, in the scope of the Program CAPES-PrInt, process number 88887.310463/2018–00, Mobility numbers #88887.468939/2019-00 and #88887.571230/2020–00. Carlos A. F. de Souza and colleagues are grateful to the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior CAPES (processes numbers CAPES 88887.360774/2019-00) and Conselho Nacional Desenvolvimento Científico e Tecnológico CNPq (ONDACBC: 465764/2014-2 and NEXUS: 441305/2017-2). Rejane M. F. Silva; Thays G. L. Oliveira and Gladstone A. da Silva are grateful to the Fundação de Amparo à Ciência e Tecnologia de Pernambuco—FACEPE (BFP-0046-5.01/20, APQ-0350-2.12/19 and APQ 1527-5.01/22), and the Conselho Nacional de Desenvolvimento Científico e Tecnológico—CNPq (Proc. 312606/2022-2). Hong Wei Shen and Zong Long Luo are grateful to the National Natural Science Foundation of China (Project ID: 32060005) and the Yunnan Fundamental Research Project (202201AW070001). Samantha C. Karunarathna and Saowaluck Tibpromma thank the National Natural Science Foundation of China (No. 32260004), Yunnan Revitalization Talents Support Plan (High-End Foreign Experts Program) and the Key Laboratory of Yunnan Provincial Department of Education of the Deep-Time Evolution on Biodiversity from the Origin of the Pearl River for their support. Xing-Can Peng and Ting-Chi Wen acknowledge the support by the National Natural Science Foundation of China (No. 32060012). Yuwei Hu thanks the Yunnan Department of Sciences and Technology of China (No. 202202AE090091). Wei Dong thanks the National Natural Science Foundation of China (Grant No. 32200015) and the foundation of the Guangzhou bureau of science and technology (Grant No. 2023A04J1425). Ruvishika S. Jayawardena and Alireza Armand are grateful to Thailand Science Research and Innovation (TSRI) for the grant “Biodiversity, taxonomy, phylogeny and evolution of Colletotrichum on Avocado, Citrus, Durian and Mango in northern Thailand” (Grant no. 652A01003). Xian Zhang and co-authors are financed by the National Natural Science Foundation of China (No. NSFC 32260004) and the Yunnan Revitalization Talents Support Plan (Young Talents Program and High-End Foreign Experts Program), and thanks The Center for Yunnan Plateau Biological Resources Protection and Utilization, College of Biological Resource and Food Engineering, Qujing Normal University for the facilities provided for the research work. Yun Xia Zhang would like to thank the National Natural Science Foundation of China (Grant no. 31600019), the Modern Agricultural Industry Technology System Flower Innovation Team of Guangdong Province (Grant no. 2023KJ121) and the Project of Educational Commission of Guangdong Province of China (Grant no. 2021KTSCX045). Marcos Paz Saraiva Câmara acknowledges CNPq for the research productivity fellowship (Grant No. 303834/2020-0). RS Jayawardena would like to thank the Eminent scholar offered by Kyun Hee University. Yanyan Yang thanks the Chinese Research Fund, Grant number E1644111K1, titled “Flexible introduction of the high-level expert program, Kunming Institute of Botany, Chinese Academy of Sciences” for financial support. Laura Selbmann and Claudia Coleine wish to thank the Italian National Antarctic Research Program for funding Antarctic campaings. The Italian Antarctic National Museum (MNA) is kindly acknowledged for financial support to the Mycological Section of the MNA and the Culture Collection of Antarctic fungi (MNA-CCFEE), University of Tuscia, Italy.
Funding
The authors have not disclosed any funding.
Declarations
Conflict of interest
The authors declare that they have no conflict of interest. The author list includes members of the Editorial Board of Fungal Diversity. They were not involved in the journal’s review of, or decisions related to, this manuscript.
Sinang Hongsanan
References
Abadi, S; Azouri, D; Pupko, T; Mayrose, I. Model selection may not be a mandatory step for phylogeny reconstruction. Nat Commun; 2019; 10, 934.1:CAS:528:DC%2BC1MXns1GksrY%3D [DOI: https://dx.doi.org/10.1038/s41467-019-08822-w] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30804347][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6389923]
Abeywickrama, PD; Q ian, N; Jayawardena, RS; Li, Y; Zhang, W; Guo, K; Zhang, L; Zhang, G; Yan, J; Li, X; Guo, Z; Hyde, KD; Peng, Y; Zhao, W. Endophytic fungi in green manure crops; friends or foe?. Mycosphere; 2023; 14, pp. 1-106. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/1]
Adanson M (1763) Familles des plantes, vol II. Paris http://www.biodiversitylibrary.org/bibliography/271#/summary. Accessed 1 Aug 2023
Abd Oun AG, Abass MH (2023) Disease Note First report of Didymella Glomerata (Corda) and Penicillium Polonicum Zaleski on Wheat (Triticum aestivum L.) Seeds in Iraq. Bas J Sci 41(1)
Abdollahzadeh, J; Groenewald, JZ; Coetzee, MP; Wingfield, MJ; Crous, PW. Evolution of lifestyles in Capnodiales. Stud Mycol; 2020; 95, pp. 381-414.1:STN:280:DC%2BB38bhsVKktg%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2020.02.004] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32855743][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7426231]
Aghyl, H; Mehrabi-Koushki, M; Esfandiari, M. Chaetomium iranicum and Collariella capillicompacta spp. nov. and notes to new hosts of Amesia species in Iran. Sydowia; 2020; 73, pp. 21-30.
Ahrén, D; Ursing, BM; Tunlid, A. Phylogeny of nematode-trapping fungi based on 18S rDNA sequences. FEMS Microbiol Lett; 1998; 158,
Ahmad S, Iqbal SH, Khalid AN (1997) Fungi of Pakistan. Mycol Soc Pak Uni. Punjab. 1–248
Aime, MC; McTaggart, AR. A higher-rank classification for rust fungi, with notes on genera. Fungal Syst Evol; 2021; 7, pp. 21-47.1:STN:280:DC%2BB2c7nvFWmtw%3D%3D [DOI: https://dx.doi.org/10.3114/fuse.2021.07.02] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34124616]
Alcorn, JL; Irwin, JAG. Acrocalymma medicaginis gen. et sp. nov. causing root and crown rot of Medicago sativa in Australia. Trans Br Mycol Soc; 1987; 88, pp. 163-167. [DOI: https://dx.doi.org/10.1016/S0007-1536(87)80211-5]
Alexopoulos, CJ; Mims, CW; Blackwell, M. Introductory mycology; 1996; 4 Hoboken, Wiley:
Alfenas, RF; Lombard, L; Pereira, OL; Alfenas, AC; Crous, PW. Diversity and potencial impact of Calonectria in Eucalyptus plantations in Brazil. Stud Mycol; 2015; 80, pp. 89-130.1:STN:280:DC%2BC28josVGgtw%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2014.11.002] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26955192][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4779794]
Allawi, MY; Al-Taee, WS. Investigation of a new local isolate of Penicillium lanosocoeruleum that produces the antifungal griseofulvin. Pak J Med Health Sci; 2022; 16,
Alves, A; Crous, PW; Correia, A; Phillips, AJ. Morphological and molecular data reveal cryptic speciation in Lasiodiplodia theobromae. Fungal Divers; 2008; 28, pp. 1-3.
Alves-Silva, G; Drechsler-Santos, ER; da Silveira, RM. Bambusicolous Fomitiporia revisited: multilocus phylogeny reveals a clade of host-exclusive species. Mycologia; 2020; 112, pp. 633-648. [DOI: https://dx.doi.org/10.1080/00275514.2020.1741316] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32412354]
Alves-Silva, G; Reck, MA; da Silveira, RM; Bittencourt, F; Robledo, GL; Góes-Neto, A; Drechsler-Santos, ER. The Neotropical Fomitiporia (Hymenochaetales, Basidiomycota): the redefinition of F. apiahyna ss allows revealing a high hidden species diversity. Mycol Prog; 2020; 19, pp. 769-790. [DOI: https://dx.doi.org/10.1007/s11557-020-01593-5]
Alves-Silva G, Drechsler-Santos ER, Góes-Neto A (2023) Data for two new species of Fomitiporia from Brazil, Harvard Dataverse, V1. https://doi.org/10.7910/DVN/VC82MZ
Amalfi, M; Decock, C. Fomitiporia castilloi sp. nov. and multiple clades around F. apiahyna and F. texana in Meso- and South America evidenced by multiloci phylogenetic inferences. Mycologia; 2013; 105, pp. 873-887. [DOI: https://dx.doi.org/10.3852/11-423] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/23709522]
Amalfi M, Decock C (2014) Fomitiporia expansa, an undescribed species from French Guiana. Cryptogamie, Mycologie 35:73–85. https://doi.org/10.7872/crym.v35.iss1.2014.73
Amaria, W; Harni, R; Samsudin, S. Evaluation of antagonistic fungi in inhibiting the growth of Rigidoporus microporus causing white root disease in rubber plants. J Ind Beverage Crops; 2016; 2,
Amoako-Attah, I; Shahin, AS; Aime, MC; Odamtten, GT; Cornelius, E; Nyaku, ST; Kumi-Asare, E; Yahaya, HB; Bailey, BA. Identification and characterization of fungi causing thread blight diseases on cacao in Ghana. Plant Dis; 2020; 104, pp. 3033-3042. [DOI: https://dx.doi.org/10.1094/PDIS-03-20-0565-RE] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32822261]
Antipova, TV; Zhelifonova, VP; Baskunov, BP; Kochkina, GA; Ozerskaya, SM; Kozlovskii, AG. Exometabolites the Penicillium fungi isolated from various high-latitude ecosystems. Microbiology; 2018; 87, pp. 642-651.1:CAS:528:DC%2BC1cXhvVGisLjF [DOI: https://dx.doi.org/10.1134/S002626171805003X]
Antonín V (2007) Monograph of Marasmius, Gloiocephala, Palaeocephala and Setulipes in tropical Africa. In: Fungus Flora of Tropical Africa, vol 1. National Tropical Botanic Garden, Belgium, 177
Antonín, V; Noordeloos, ME. A monograph of marasmioid and collybioid fungi in Europe: with 131 The collections of this studyures and 130 coloured plates; 2010; Eching, IHW-Verlag: 480.
Aptroot, A. Redisposition of some species excluded from Didymosphaeria (Ascomycotina). Nova Hedwig; 1995; 60, pp. 325-379.
Aptroot, A. Notes on the saprobic ascomycetes from Papua New Guinea, with the new genus Papilionovela. Mycol Res; 1997; 101, pp. 266-268. [DOI: https://dx.doi.org/10.1017/S0953756296002511]
Aptroot, A; Feuerstein, S. New Graphidaceae from South and Central Brazil. Arch Lichenol; 2020; 16, pp. 1-10.
Aptroot, A; de Souza, MF; dos Santos, LA; Junior, IO; Barbosa, BM; da Silva, ME. New species of lichenized fungi from Brazil, with a record report of 492 species in a small area of the Amazon Forest. The Bryologist; 2022; 125, pp. 433-465. [DOI: https://dx.doi.org/10.1639/0007-2745-125.3.433]
Araújo, JPM; Hughes, DP. Diversity of entomopathogenic fungi: which groups conquered the insect body?. Adv Genet; 2016; 94, pp. 1-39. [DOI: https://dx.doi.org/10.1016/bs.adgen.2016.01.001] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27131321]
Armand, A; Hyde, KD; Huanraluek, N; Wang, Y; Jayawardena, RS. Identification and characterization of Colletotrichum species associated with durian fruit in northern Thailand. Mycosphere; 2023; 14, pp. 107-129. [DOI: https://dx.doi.org/10.5943/mycosphere/14/si2/2]
Ariyawansa, HA; Camporesi, E; Thambugala, KM; Mapook, A; Kang, JC; Alias, SA; Chukeatirote, E; Thines, M; Mckenzie, EHC; Hyde, KD. Confusion surrounding Didymosphaeria—phylogenetic and morphological evidence suggest Didymosphaeriaceae is not a distinct family. Phytotaxa; 2014; 176, pp. 102-119. [DOI: https://dx.doi.org/10.11646/phytotaxa.176.1.12]
Ariyawansa, HA; Hawksworth, DL; Hyde, KD; Jones, EG; Maharachchikumbura, SS; Manamgoda, DS; Thambugala, KM; Udayanga, D; Camporesi, E; Daranagama, A; Jayawardena, R; Liu, JK; McKenzie, EHC; Phookamsak, R; Senanayake, IC; Shivas, RG; Tian, Q; Xu, JC. Epitypification and neotypification: guidelines with appropriate and inappropriate examples. Fungal Divers; 2014; 69, pp. 57-91. [DOI: https://dx.doi.org/10.1007/s13225-014-0315-4]
Ariyawansa HA, Hyde KD, Jayasiri SC, Buyck B, Chethana KWT,Dai DQ, Dai YC, Daranagama DA, Jayawardena RS, Lu¨cking 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- AzizFA, Wedin M, Westberg M, Ammirati JF, Bulgakov TS, Lima DX, Callaghan TM, Callac P, Chang CH, Coca LF, Dal-FornoM, Dollhofer V, Fliegerova´ K, Greiner K, Griffith GW, Ho HM, Hofstetter V, Jeewon R, Kang JC, Wen TC, Kirk PM, Kyto¨vuori I, Lawrey JD, Xing J, Li H, Liu ZY, Liu XZ, Liimatainen K, Thorsten Lumbsch H, Matsumura M, Moncada B, Nuankaew S,Parnmen S, Santiago ALCMA, Sommai S, Song Y, de Souza CAF, de Souza- Motta CM, Su HY, Suetrong S, Wang Y, FongWS YH, Zhou LW, Re´blova´ 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, Lee HB, Battistin E, Marsico O, Vizzini A, Vila J, Ercole E, Eberhardt U, Simonini G, Wen HA, Chen XH, Miettinen O, Spirin V, Hernawati (2015c) Taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 75:27–274
Aveskamp, MM; de Gruyter, J; Woudenberg, JHC; Verkley, GJM; Crous, PW. Highlights of the Didymellaceae: a polyphasic approach to characterise Phoma and related pleosporalean genera. Stud Mycol; 2010; 65, pp. 1-60.1:STN:280:DC%2BC3czmsVenuw%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/20502538][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2836210][DOI: https://dx.doi.org/10.3114/sim.2010.65.01]
Azuddin, NF; Mohd, MH; Rosely, NFN; Mansor, A; Zakaria, L. Molecular phylogeny of endophytic fungi from rattan (Calamus Castaneus Griff) spines and their antagonistic activities against plant pathogenic fungi. J Fungi; 2021; 7, 301.1:CAS:528:DC%2BB3MXhvFOgsb7E [DOI: https://dx.doi.org/10.3390/jof7040301]
Bago, B; Bentivenga, SP; Brenac, V; Dodd, JC; Piche, Y; Simon, L. Molecular analysis of Gigaspora (Glomales, Gigasporaceae). New Phytol; 1998; 139, pp. 581-588.1:CAS:528:DyaK1cXlsleqtr0%3D [DOI: https://dx.doi.org/10.1046/j.1469-8137.1998.00212.x]
Baker, WA; Partridge, EC; Morgan-Jones, G. Notes on hyphomycetes LXXXVII. Rhexoacrodictys, a new segregate genus to accommodate four species previously classified in Acrodictys. Mycotaxon; 2002; 82, pp. 95-113.
Baloch, E; Lücking, R; Lumbsch, HT; Wedin, M. Major clades and phylogenetic relationships between lichenized and non-lichenized lineages in Ostropales (Ascomycota: Lecanoromycetes). Taxon; 2010; 59, pp. 1483-1494. [DOI: https://dx.doi.org/10.1002/tax.595013]
Ban, S; Sakane, T; Nakagiri, A. Three new species of Ophiocordyceps and overview of anamorph types in the genus and the family Ophiocordycipitaceae. Mycol Prog; 2015; 14, pp. 1-12. [DOI: https://dx.doi.org/10.1007/s11557-014-1017-8]
Bao, DF; Hyde, KD; McKenzie, EHC; Jeewon, R; Su, HY; Nalumpang, S; Luo, ZL. Biodiversity of lignicolous freshwater hyphomycetes from China and Thailand and description of sixteen species. J Fungi; 2021; 7, pp. 669-715. [DOI: https://dx.doi.org/10.3390/jof7080669]
Bao, DF; Bhat, DJ; Boonmee, S; Hyde, KD; Luo, ZL; Nalumpang, S. Lignicolous freshwater ascomycetes from Thailand: introducing Dematipyriforma muriformis sp. nov., one new combination and two new records in Pleurotheciaceae. MycoKeys; 2022; 93, pp. 57-79. [DOI: https://dx.doi.org/10.3897/mycokeys.93.87797] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36761914][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9836478]
Baral HO, Marson G (2000) Monographic revision of Gelatinopsis and Calloriopsis Calloriopsideae, Leotiales. Monographic revision of Gelatinopsis and CalloriopsisCalloriopsideae, Leotiales 23–46
Baral, HO; Weber, E; Gams, W; Hagedorn, G; Liu, B; Liu, X; Marson, G; Marvanová, L; Stadler, M; Weiß, M. Generic names in the Orbiliaceae (Orbiliomycetes) and recommendations on which names should be protected or suppressed. Mycol Prog; 2018; 17, pp. 5-31. [DOI: https://dx.doi.org/10.1007/s11557-017-1300-6]
Baral HO, Weber E, Marson G (2020) Monograph of Orbiliomycetes (Ascomycota) based on vital taxonomy. National Museum of Natural History Luxembourg
Barbosa, RN; Bezerra, JDP; Souza-Motta, CM et al. (2018) New Penicillium and Talaromyces species from honey, pollen and nests of stingless bees. AntonieVan Leeuwenhoek; 2018; 111,
Barnett, HL. Illustrated genera of imperfect fungi; 1956; Minneapolis, Burgess Publishing Company: 218.
Barr, ME. Notes on the Amphisphaeriaceae and related families. Mycotaxon; 1994; 51, pp. 191-224.
Basson, E; Meitz-Hopkins, JC; Lennox, CL. Morphological and molecular identification of fungi associated with South African apple core rot. Eur J Plant Pathol; 2019; 153, pp. 849-868.1:CAS:528:DC%2BC1cXhslSrsLvO [DOI: https://dx.doi.org/10.1007/s10658-018-1601-x]
Bentivenga, SP; Morton, JB. A monograph of the genus Gigaspora, incorporating developmental patterns of morphological characters. Mycologia; 1995; 87, pp. 719-731. [DOI: https://dx.doi.org/10.1080/00275514.1995.12026590]
Begerow D, McTaggart A (2018) Ustilaginomycotina. In: Begerow D et al (eds) Syllabus of plant families—A. Engler’s Syllabus der Pflanzenfamilien, Part 1/3: Basidiomycota and Entorrhizomycota. Schweizerbart Science Publishers, Stuttgart, pp 92–129
Berkeley, MJ; Broome, CE. Notices of British Fungi. Ann Mag Nat Hist Zool Bot Geol; 1859; 3, 358.
Bernicchia A, Gorjón SP (2010) Fungi Europaei 12: Corticiaceae s.l. Edizioni Candusso. Lomazzo 1–1007
Bezerra, JD; Machado, AR; Firmino, AL; Rosado, AW; Souza, CD; Souza-Motta, CM; Freire, KT; Paiva, LM; Magalhães, OM; Pereira, OL; Crous, PW; Oliveira, TG; Abreu, V; Fan, X. Mycological diversity description I. Acta Bot Bras; 2018; 32, pp. 656-666. [DOI: https://dx.doi.org/10.1590/0102-33062018abb0154]
Bhattacharjee, M; Mukerji, KG; Tewari, JP; Skoropad, WP. Structure and Hyperparasitism of a new species of Gigaspora. Trans Br Mycol Soc; 1982; 78,
Bhunjun, CS; Phukhamsakda, C; Jayawardena, RS; Jeewon, R et al. Investigating species boundaries in Colletotrichum. Fungal Divers; 2021; 107, pp. 107-127.1:CAS:528:DC%2BB38XitlCitr7N [DOI: https://dx.doi.org/10.1007/s13225-021-00471-z]
Bhunjun, CS; Niskanen, T; Suwannarach, N; Wannathes, N; Chen, YJ; McKenzie, EH; Maharachchikumbura, SS; Buyck, B; Zhao, CL; Fan, YG; Zhang, JY; Dissanayake, AJ; Marasinghe, DS; Jayawardena, RS; Kumla, J; Padamsee, M; Chen, YY; Liimatainen, K; Ammirati, JF; Phukhamsakda, C; Liu, JK; Phonrob, W; Randrianjohany, E; Hongsanan, S; Cheewangkoon, R; Bundhun, D; Khuna, S; Yu, WJ; Deng, LS; Lu, YZ; Hyde, KD; Lumyong, S. The numbers of fungi: are the most speciose genera truly diverse?. Fungal Divers; 2022; 114, pp. 387-462.1:CAS:528:DC%2BB38XhslWqtL7P [DOI: https://dx.doi.org/10.1007/s13225-022-00501-4]
Binder, M; Hibbett, DS. Molecular systematics and biological diversification of Boletales. Mycologia; 2006; 98, pp. 971-981. [DOI: https://dx.doi.org/10.1080/15572536.2006.11832626] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/17486973]
Binder, M; Hibbett, DS; Larsson, K-H; Larsson, E; Langer, E. The phylogenetic distribution of resupinate forms in the homobasidiomycetes. Syst Biodivers; 2005; 3, pp. 113-157. [DOI: https://dx.doi.org/10.1017/S1477200005001623]
Binder, M; Justo, A; Riley, R; Salamov, A; Lopez-Giraldez, F; Sjökvist, E; Copeland, A; Foster, B; Sun, H; Larsson, E; Larsson, KH; Townsend, J; Grigoriev, IV; Hibbett, DS. Phylogenetic and phylogenomic overview of the Polyporales. Mycologia; 2013; 105, pp. 1350-1373.1:CAS:528:DC%2BC3sXhvF2ku7zO [DOI: https://dx.doi.org/10.3852/13-003] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/23935031]
Boertmann D (1995) The genus Hygrocybe. Fungi of Northern Europe vol. 1, 1st edn. Danish Mycological Society, Greve
Boertmann D (2010) The genus Hygrocybe. Fungi of Northern Europe vol. 1, 2nd edn. Danish Mycological Society, Denmark
Boidin, J; Lanquetin, P. Vararia subgenus Vararia (Basidiomycètes Lachnocladiaceae): étude spéciale des espèces d'Afrique intertropicale. Bull Soc Mycol France; 1975; 91, pp. 457-513.
Boonmee, S; D’souza, MJ; Luo, Z; Pinruan, U; Tanaka, K; Su, H; Bhat, DJ; McKenzie, EH; Jones, EG; Taylor, JE; Phillips, AJ. Dictyosporiaceae fam. nov. Fungal Divers; 2016; 80, pp. 457-482. [DOI: https://dx.doi.org/10.1007/s13225-016-0363-z]
Boonmee, S; Wanasinghe, DN; Calabon, MS; Huanraluek, N; Chandrasiri, SKU; Jones, GEB; Rossi, W; Leonardi, M; Singh, SK; Rana, S; Singh, PN; Maurya, DK; Lagashetti, AC; Choudhary, D; Dai, YC; Zhao, CL; Mu, YH; Yuan, HS; He, SH; Phookamsak, R; Jiang, HB; Martín, MP; Dueñas, M; Telleria, MT; Kałucka, IL; Jagodziński, AM; Liimatainen, K; Pereira, DS; Phillips, AJL; Suwannarach, N; Kumla, J; Khuna, S; Lumyong, S; Potter, TB; Shivas, RG; Sparks, AH; Vaghefi, N; Abdel-Wahab, MA; Abdel-Aziz, FA; Li, GJ; Lin, WF; Singh, U; Bhatt, RP; Lee, HB; Nguyen, TTT; Kirk, PM; Dutta, AK; Acharya, K; Sarma, VV; Niranjan, M; Rajeshkumar, KC; Ashtekar, N; Lad, S; Wijayawardene, NN; Bhat, DJ; Xu, RJ; Wijesinghe, SN; Shen, HW; Luo, ZL; Zhang, JY; Sysouphanthong, P; Thongklang, N; Bao, DF; Aluthmuhandiram, JVS; Abdollahzadeh, J; Javadi, A; Dovana, F; Usman, M; Khalid, AN; Dissanayake, AJ; Telagathoti, A; Probst, M; Peintner, U; Garrido-Benavent, I; Bóna, L; Merényi, Z; Boros, L; Zoltán, B; Stielow, JB; Jiang, N; Tian, CM; Shams, E; Dehghanizadeh, F; Pordel, A; Javan-Nikkhah, M; Denchev, TT; Denchev, CM; Kemler, M; Begerow, D; Deng, CY; Harrower, E; Bozorov, T; Kholmuradova, T; Gafforov, Y; Abdurazakov, A; Xu, JC; Mortimer, PE; Ren, GC; Jeewon, R; Maharachchikumbura, SSN; Phukhamsakda, C; Mapook, A; Hyde, KD. Fungal diversity notes 1387–1511: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers; 2021; 111, pp. 1-335. [DOI: https://dx.doi.org/10.1007/s13225-021-00489-3] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34899100][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8648402]
Boucias, DG; Scharf, DW; Breaux, SE; Purcell, DH; Mizell, RE. Studies on the fungi associated with the glassy-winged sharpshooter Homalodisca coagulata with emphasis on a new species Hirsutella homalodiscae nom. prov. Biocontrol; 2007; 52, pp. 231-258. [DOI: https://dx.doi.org/10.1007/s10526-006-9019-3]
Bozonnet, J; Meyer, M; Poulain, M. Lamproderma cacographicum une nouvelle espèce nivale de Myxomycetes. Bull Féd Dauphiné-Savoie; 1997; 144, pp. 117-121.
Bracale, MF; Nóbrega, TF; Barreto, RW. Fungal diseases of non-conventional food plants: first report of Stagonosporopsis caricae causing leaf spots on Vasconcellea monoica. Aust Plant Dis Notes; 2020; 15, pp. 1-4. [DOI: https://dx.doi.org/10.1007/s13314-020-00391-y]
Bradshaw, MJ; Bartholomew, HP; Lichtner, F; Gaskins, VL; Jurick, WM. First report of blue mold caused by Penicillium polonicum on apple in the United States. Plant Dis; 2022; 106,
Brackel, WV. Weitere Funde von flechtenbewohnenden Pilzen in Bayern—Beitrag zu einer Checkliste IV. Ber Bayer Bot Ges; 2009; 79, pp. 5-55.
Brackel, WV. Kommentierter Katalog der flechtenbewohnenden Pilze Bayerns. Bibl Lichenol; 2014; 109, pp. 1-476.
Caceres, MES; Aptroot, A; Parnmen, S; Lücking, R. Remarkable diversity of the lichen family Graphidaceae in the Amazon rain forest of Rondônia, Brazil. Phytotaxa; 2014; 189, pp. 87-136. [DOI: https://dx.doi.org/10.11646/phytotaxa.189.1.8]
Cáceres, MES; Lücking, R; SchummAptroot, FA. A lichenized family yields another renegade lineage: Papilionovela albothallina is the first non-lichenized, saprobic member of Graphidaceae subfam. Graphidoideae. Bryologist; 2020; 123, pp. 144-154. [DOI: https://dx.doi.org/10.1639/0007-2745-123.2.144]
Çakır, E; Maden, S. First report of Penicillium polonicum causing storage rots of onion bulbs in Ankara province, Turkey. New Dis Rep; 2015; 32,
Calabon, MS; Jones, EBG; Boonmee, S; Doilom, M; Lumyong, S; Hyde, KD. Five novel freshwater Ascomycetes indicate high undiscovered diversity in lotic habitats in Thailand. J Fungi; 2021; 7, 117. [DOI: https://dx.doi.org/10.3390/jof7020117]
Calabon MS, Jones EG, Boonmee S, Li WJ, Xiao YP, Hyde KD (2023a) Acrocalymmaceae (Pleosporales) from freshwater habitats in Thailand with the introduction of Acrocalymma bilobatum sp. nov. Stud Fungi 8(1)
Calabon, MS; Hyde, KD; Jones, EBG; Bao, DF; Bhunjun, CS; Phukhamsakda, C; Shen, HW; Gentekaki, E; Al Sharie, AH; Barros, JL; Chandrasiri, KSU. Freshwater fungal biology. Mycosphere; 2023; 14,
Campos-Santana M, Amalfi M, Robledo G, da Silveira RMB, Decock C (2014) Fomitiporia neotropica, a new species from South America evidenced by multilocus phylogenetic analyses. Mycological progress 13:601–615. https://doi.org/10.1007/s11557-013-0943-1
Campbell, J; Anderson, JL; Shearer, CA. Systematics of Halosarpheia based on morphological and molecular data. Mycologia; 2003; 95, pp. 530-552.1:CAS:528:DC%2BD3sXmslantr4%3D [DOI: https://dx.doi.org/10.1080/15572536.2004.11833100] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/21156644]
Cao, Y; Wu, SH; Dai, YC. Species clarification of the prize medicinal Ganoderma mushroom “Lingzhi”. Fungal Divers; 2012; 56, pp. 49-62. [DOI: https://dx.doi.org/10.1007/s13225-012-0178-5]
Cao, B; Haelewaters, D; Schoutteten, N; Begerow, D; Boekhout, T; Giachini, AJ; Gorjón, SP; Gunde-Cimerman, N; Hyde, KD; Kemler, M; Li, GJ. Delimiting species in Basidiomycota: a review. Fungal Divers; 2021; 1, pp. 1-57.1:CAS:528:DC%2BB3sXhvVOrt77I [DOI: https://dx.doi.org/10.1007/s13225-021-00479-5]
Cha, HJ; Chiang, MW; Guo, SY; Lin, SM; Pang, KL. Culturable fungal community of Pterocladiella capillacea in Keelung, Taiwan: effects of surface sterilization method and isolation medium. J Fungi; 2021; 7, 651.1:CAS:528:DC%2BB3MXit1Kkt7%2FE [DOI: https://dx.doi.org/10.3390/jof7080651]
Chatterton, S; Wylie, AC; Punja, ZK. Fruit infection and postharvest decay of greenhouse tomatoes caused by Penicillium species in British Columbia. Can J Plant Path; 2012; 34,
Chen, Q; Jiang, JR; Zhang, GZ; Cai, L; Crous, PW. Resolving the Phoma enigma. Stud Mycol; 2015; 82, pp. 137-217.1:STN:280:DC%2BC28josVGntw%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26955202][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4774273][DOI: https://dx.doi.org/10.1016/j.simyco.2015.10.003]
Chen, Q; Hou, LW; Duan, WJ; Crous, PW; Cai, L. Didymellaceae revisited. Stud Mycol; 2017; 87, pp. 105-159.1:STN:280:DC%2BC1cjnsVeisw%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2017.06.002] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28706324][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5498420]
Chen, CC; Chen, CY; Wu, SH. Species diversity, taxonomy and multi-gene phylogeny of phlebioid clade (Phanerochaetaceae, Irpicaceae, Meruliaceae) of Polyporales. Fungal Divers; 2021; 111, pp. 337-442.1:CAS:528:DC%2BB38XhslWqt7rO [DOI: https://dx.doi.org/10.1007/s13225-021-00490-w]
Chen, YP; Wu, T; Tian, WH; Ilyukhin, E; Hyde, KD; Maharachchikumbura, SSN. Comparative genomics provides new insights into the evolution of Colletotrichum. Mycosphere; 2022; 13,
Chen, Q; Bakhshi, M; Balci, Y; Broders, KD; Cheewangkoon, R; Chen, SF; Fan, XL; Gramaje, D; Halleen, F; Horta JungMmJiang, N; Jung, T; Májek, T; Marincowitz, S; Milenković, I; Mostert, L; Nakashima, C; Nurul Faziha, I; Pan, M; Raza, M; Scanu, B; Spies, CFJ; Suhaizan, L; Suzuki, H; Tian, CM; Tomšovský, M; Úrbez-Torres, JR; Wang, W; Wingfield, BD; Wingfield, MJ; Yang, Q; Yang, X; Zare, R; Zhao, P; Groenewald, JZ; Cai, L; Crous, PW. Genera of phytopathogenic fungi: GOPHY 4. Stud Mycol; 2022; 101, pp. 417-564.1:STN:280:DC%2BB287htFKjtA%3D%3D [DOI: https://dx.doi.org/10.3114/sim.2022.101.06] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36059898][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9365048]
Chen, YP; Su, PW; Hyde, KD; Maharachchikumbura, SSN. Phylogenomics and diversification of sordariomycetes. Mycosphere; 2023; 14, pp. 414-451. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/5]
Chethana, KWT; Niranjan, M; Dong, W; Samarakoon, MC; Bao, DF; Calabon, MS; Chaiwan, N; Chuankid, B; Dayarathne, MC; de Silva, NI; Devadatha, B; Dissanayake, AJ; Goonasekara, ID; Huanraluek, N; Jayawardena, RS; Karunarathna, A; Luo, ZL; Marasinghe, DS; Ma, XY; Norphanphoun, C; Pem, D; Perera, RH; Rathnayaka, AR; Raspé, O; Samarakoon, BC; Senwanna, C; Sun, YR; Tang, X; Thiyagaraja, V; Tennakoon, DS; Zeng, M; Zeng, XY; Zhang, JY; Zhang, SN; Bulgakov, TS; Camporesi, E; Sarma, VV; Wang, Y; Bhat, DJ; Hyde, KD. AJOM new records and collections of fungi: 101–150. Asian J Mycol; 2021; 4, pp. 113-260. [DOI: https://dx.doi.org/10.5943/ajom/4/1/8]
Chethana, KWT; Manawasinghe, IS; Hurdeal, VG; Bhunjun, CS; Appadoo, MA; Gentekaki, E; Raspé, O; Promputtha, I; Hyde, KD. What are fungal species and how to delineate them?. Fungal Divers; 2021; 109, pp. 1-25. [DOI: https://dx.doi.org/10.1007/s13225-021-00483-9]
Chitale, G; Makhija, U; Sharma, B. Additional species of Graphis from Maharashtra, India. Mycotaxon; 2011; 115, pp. 469-480. [DOI: https://dx.doi.org/10.5248/115.469]
Choi, DH; Kim, YG; Lee, IS; Hong, SB; Kim, JG. Identification and characterization of unreported Penicillium species in Korea. Korean J Mycol; 2020; 48,
Chomnunti, P; Bhat, DJ; Jones, EG; Chukeatirote, E; Bahkali, AH; Hyde, KD. Trichomeriaceae, a new sooty mould family of Chaetothyriales. Fungal Divers; 2012; 56, pp. 63-76. [DOI: https://dx.doi.org/10.1007/s13225-012-0197-2]
Chou, WN; Yen, CH; Chung, HH. Species of Gigaspora and Scutellospora (Endogonaceae) in Taiwan. Trans Mycol Soc Rep China; 1991; 6, pp. 1-17. [DOI: https://dx.doi.org/10.7099/TMSRC.199112.0001]
Clarke, RT. Fungal spores from Vermejo Formation coal beds (Upper Cretaceous) of Central Colorado. Mt Geol; 1965; 2, pp. 85-93.
Clendenin, I. Lasiodiplodia E. and E., n. gen. Bot Gazette; 1896; 21, pp. 92-93. [DOI: https://dx.doi.org/10.1086/327307]
Cohen, R; Persky, L; Hadar, Y. Biotechnological applications and potential of wood-degrading mushrooms of the genus Pleurotus. Appl Microbiol Biotechnol; 2002; 58, pp. 582-594.1:CAS:528:DC%2BD38Xjs12ktLs%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/11956739][DOI: https://dx.doi.org/10.1007/s00253-002-0930-y]
Cooke, RC; Dickinson, CH. Nematode-trapping species of Dactylella and Monacrosporium. Trans Br Mycol Soc; 1965; 48, pp. 621-629. [DOI: https://dx.doi.org/10.1016/S0007-1536(65)80040-7]
Corda AC (1836) Mykologische Beobachtungen. Weitenweber’s Beitr Naturges Heilwiss, Prague
Corner, EJH. The agaric genera Marasmius, Chaetocalathus, Crinipellis, Heimiomyces, Resupinatus, Xerula and Xerulina in Malesia. Beiheft Nova Hedwigia; 1996; 111, pp. 1-175.
Corrado, I; Cascelli, N; Ntasi, G; Birolo, L; Sannia, G; Pezzella, C. Optimization of inulin hydrolysis by Penicillium lanosocoeruleum inulinases and efficient conversion into polyhydroxyalkanoates. Front Bioeng Biotechnol; 2021; 9, [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33732688][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7959777][DOI: https://dx.doi.org/10.3389/fbioe.2021.616908] 616908.
Correia, KC; Silva, MA; de Morais Jr, MA; Armengol, J; Phillips, AJ; Cômara, MP; Michereff, SJ. Phylogeny, distribution and pathogenicity of Lasiodiplodia species associated with dieback of table grape in the main Brazilian exporting region. Plant Pathol; 2016; 65, pp. 92-103.1:CAS:528:DC%2BC2MXitVCrsbbF [DOI: https://dx.doi.org/10.1111/ppa.12388]
Costa, J; Santos, C; Soares, C; Rodríguez, R; Lima, N; Santos, C. Occurrence of Aflatoxins and Ochratoxin A during Merkén pepper powder production in Chile. Foods; 2022; 11,
Crous, PW; Groenewald, JZ. The genera of fungi—G 4: Camarosporium and Dothiora. IMA Fungus; 2017; 8, pp. 131-152. [DOI: https://dx.doi.org/10.5598/imafungus.2017.08.01.10] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28824845][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5493531]
Crous, PW; Wingfield, MJ; Schumacher, RK; Summerell, BA; Giraldo, A; Gené, J; Guarro, J; Wanasinghe, DN; Hyde, KD; Camporesi, E; Gareth Jones, EB; Thambugala, KM; Malysheva, EF; Malysheva, VF; Acharya, K; Álvarez, J; Alvarado, P; Assefa, A; Barnes, CW; Bartlett, JS; Blanchette, RA; Burgess, TI; Carlavilla, JR; Coetzee, MP; Damm, U; Decock, CA; den Breeÿen, A; de Vries, B; Dutta, AK; Holdom, DG; Rooney-Latham, S; Manjón, JL; Marincowitz, S; Mirabolfathy, M; Moreno, G; Nakashima, C; Papizadeh, M; Shahzadeh Fazeli, SA; Amoozegar, MA; Romberg, MK; Shivas, RG; Stalpers, JA; Stielow, B; Stukely, MJ; Swart, WJ; Tan, YP; van der BankWood, MAR; ZhangGroenewald, YJZ. Fungal Planet description sheets: 281–319. Persoonia; 2014; 33, pp. 212-289.1:STN:280:DC%2BC2MngvFyjuw%3D%3D [DOI: https://dx.doi.org/10.3767/003158514X685680] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25737601][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4312934]
Crous, PW; Carris, LM; Giraldo, A; Groenewald, JZ; Hawksworth, DL; Hernández-Restrepo, M; Wood, AR. The genera of fungi fixing the application of the type species of generic names—G2: Allantophomopsis, Latorua, Macrodiplodiopsis, Macrohilum, Milospium, Protostegia, Pyricularia, Robillarda, Rotula, Septoriella, Torula and Wojnowicia. IMA Fungus; 2015; 6, pp. 163-198. [DOI: https://dx.doi.org/10.5598/imafungus.2015.06.01.11] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26203422][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4500082]
Crous, PW; Wingfield, MJ; Guarro, J; Hernández-Restrepo, M; Sutton, DA; Acharya, K; Barber, PA; Boekhout, T; Dimitrov, RA; Dueñas, M; Dutta, AK; Gené, J; Gouliamova, DE; Groenewald, M; Lombard, L; Morozova, OV; Sarkar, J; Smith, MTH; Stchigel, AM; Wiederhold, NP; Alexandrova, AV; Antelmi, I; Armengol, J; Barnes, I; Cano-Lira, JF; Ruiz, RFC; Contu, M; Courtecuisse, PR; da Silveira, AL; Decock, CA; de Goes, A; Edathodu, J; Ercole, E; Firmino, AC; Fourie, A; Fournier, J; Furtado, EL; Geering, ADW; Gershenzon, J; Giraldo, A; Gramaje, D; Hammerbacher, A; He, XL; Haryadi, D; Khemmuk, W; Kovalenko, AE; Krawczynski, R; Laich, F; Lechat, C; Lopes, UP; Madrid, H; Malysheva, EF; Marín-Felix, Y; Martín, MP; Mostert, L; Nigro, F; Pereira, OL; Picillo, B; Pinho, DB; Popov, ES; Peláez, CA; Rodas Rooney-Latham, S; Sandoval-Denis, M; Shivas, RG; Silva, V; Stoilova-Disheva, MM; Telleria, MT; Ullah, C; Unsicker, SB; van der Merwe, NA; Vizzini, A; Wagner, HG; Wong, PTW; Wood, AR; Groenewald, JZ. Fungal Planet description sheets: 320–370. Persoonia-Mol Phylogeny Evol Fungi; 2015; 34, pp. 167-266.1:STN:280:DC%2BC28%2FpvFSquw%3D%3D [DOI: https://dx.doi.org/10.3767/003158515X688433]
Crous, PW; Wingfield, MJ; Burgess, TI; Carnegie, AJ; Hardy, GS; Smith, D; Summerell, BA; Cano-Lira, JF; Guarro, J; Houbraken, J; Lombard, L. Fungal Planet description sheets: 625–715. Persoonia; 2017; 39, pp. 270-467.1:STN:280:DC%2BC1MrnsFOhsQ%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2017.39.11] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29503478][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5832955]
Crous, PW; Wingfield, MJ; Burgess, TI; Hardy, GESJ et al. Fungal Planet description sheets: 558–624. Persoonia; 2017; 38, pp. 240-384.1:STN:280:DC%2BC1M3isF2mtg%3D%3D [DOI: https://dx.doi.org/10.3767/003158517X698941] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29151634][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5645186]
Crous PW, Wingfield MJ, Burgess TI, Carnegie AJ, Hardy GS (2017c) Fungal Planet description sheets: 625–715. Persoonia 39:270. https://doi.org/10.3767/persoonia.2017.39.11
Crous, PW; Schumacher, RK; Wingfield, MJ; Akulov, A; Denman, S; Roux, J; Braun, U; Burgess, TI; Carnegie, AJ; Váczy, KZ. New and interesting fungi 1. FUSE-Fungal Syst Evol; 2018; 1, pp. 169-215.1:STN:280:DC%2BB38rltFGhuw%3D%3D [DOI: https://dx.doi.org/10.3114/fuse.2018.01.08]
Crous PW, Wingfield MJ, Burgess TI, Hardy GS, Barber PA, Alvarado P, Barnes CW, Buchanan PK, Heykoop M, Moreno G, Thangavel R (2019a) Foliar pathogens of eucalypts. Stud Mycol 94:125–298. https://doi.org/10.3767/003158517X698941
Crous, PW; Wingfield, MJ; Lombard, L; Roets, F; Swart, WJ; Alvarado, P; Carnegie, AJ; Moreno, G; Luangsaard, J; Thangavel, R; Alexandrova, AV; Baseia, IG; Bellanger, JM; Bessette, AE; Bessette, AR; De la Peña-Lastra, S; García, D; Gené, J; Pham, THG; Heykoop, M; Malysheva, E; Malysheva, V; Martín, MP; Morozova, OV; Noisripoom, W; Overton, BE; Rea, AE; Sewall, BJ; Smith, ME; Smyth, CW; Tasanathai, K; Visagie, CM; Adamčík, S; Alves, A; Andrade, JP; Aninat, MJ; Araújo, RVB; Bordallo, JJ; Boufleur, T; Baroncelli, R; Barreto, RW; Bolin, J; Cabero, J; Caboň, M; Cafà, G; Caffot, MLH; Cai, L; Carlavilla, JR; Chávez, R; de Castro, RRL; Delgat, L; Deschuyteneer, D; Dios, MM; Domínguez, LS; Evans, HC; Eyssartier, G; Ferreira, BW; Figueiredo, CN; Liu, F; Fournier, J; Galli-Terasawa, LV; Gil-Durán, C; Glienke, C; Gonçalves, MFM; Gryta, H; Guarro, J; Himaman, W; Hywel-Jones, N; Iturrieta-González, I; Ivanushkina, NE; Jargeat, P; Khalid, AN; Khan, J; Kiran, M; Kiss, L; Kochkina, GA; Kolařík, M; Kubátová, A; Lodge, DJ; Loizides, M; Luque, D; Manjón, JL; Marbach, PAS; Massola, NS, Jr; Mata, M; Miller, AN; Mongkolsamrit, S; Moreau, PA; Morte, A; Mujic, A; Navarro-Ródenas, A; Németh, MZ; Nóbrega, TF; Nováková, A; Olariaga, I; Ozerskaya, SM; Palma, MA; Petters-Vandresen, DAL; Piontelli, E; Popov, ES; Rodríguez, A; Requejo, Ó; Rodrigues, ACM; Rong, IH; Roux, J; Seifert, KA; Silva, BDB; Sklenář, F; Smith, JA; Sousa, JO; Souza, HG; De Souza, JT; Švec, K; Tanchaud, P; Tanney, JB; Terasawa, F; Thanakitpipattana, D; Torres-Garcia, D; Vaca, I; Vaghefi, N; van Iperen, AL; Vasilenko, OV; Verbeken, A; Yilmaz, N; Zamora, JC; Zapata, M; Jurjević, Ž; Groenewald, JZ. Fungal Planet description sheets: 951–1041. Persoonia; 2019; 43, pp. 223-425.1:STN:280:DC%2BB383ns1KhsQ%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2019.43.06] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32214501][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7085856]
Crous, PW; Wingfield, MJ; Schumacher, RK; Akulov, A; Bulgakov, TS; Carnegie, AJ; Jurjević, Ž; Decock, C; Denman, S; Lombard, L; Lawrence, DP; Stack, AJ; Gordon, TR; Bostock, RM; Burgess, T; Summerell, BA; Taylor, PWJ; Edwards, J; Hou, LW; Cai, L; Rossman, AY; Wöhner, T; Allen, WC; Castlebury, LA; Visagie, CM; Groenewald, JZ. New and Interesting Fungi 3. FUSE-Fungal Syst Evol; 2020; 6, pp. 157-231.1:STN:280:DC%2BB38bmtlSitw%3D%3D [DOI: https://dx.doi.org/10.3114/fuse.2020.06.09]
Crous, PW; Wingfield, MJ; Chooi, YH; Gilchrist, CLM; Lacey, E; Pitt, JI; Roets, F; Swart, WJ; Cano-Lira, JF; ValenzuelaLopez, N; Hubka, V; Shivas, RG; Stchigel, AM; Holdom, DG; Jurjević, Ž; Kachalkin, AV; Lebel, T; Lock, C; Martín, MP; Tan, YP; Tomashevskaya, MA; Vitelli, JS; Baseia, IG; Bhatt, VK; Brandrud, TE; De Souza, JT; Dima, B; Lacey, HJ; Lombard, L; Johnston, PR; Morte, A; Papp, V; Rodríguez, A; Rodríguez-Andrade, E; Semwal, KC; Tegart, L; Abad, ZG; Akulov, A; Alvarado, P; Alves, A; Andrade, JP; Arenas, F; Asenjo, C; Ballarà, J; Barrett, MD; Berná, LM; Berraf-Tebbal, A; Bianchinotti, MV; Bransgrove, K; Burgess, TI; Carmo, FS; Chávez, R; Čmoková, A; Dearnaley, JDW; de Santiago, A; FreitasNeto, JF; Denman, S; Douglas, B; Dovana, F; Eichmeier, A; Esteve-Raventós, F; Farid, A; Fedosova, AG; Ferisin, G; Ferreira, RJ; Ferrer, A; Figueiredo, CN; Figueiredo, YF; Reinoso-Fuentealba, CG; Garrido-Benavent, I; Cañete-Gibas, CF; GilDurán, C; Glushakova, AM; Gonçalves, MFM; González, M; Gorczak, M; Gorton, C; Guard, FE; Guarnizo, AL; Guarro, J; Gutiérrez, M; Hamal, P; Hien, LT; Hocking, AD; Houbraken, J; Hunter, GC; Inácio, CA; Jourdan, M; Kapitonov, VI; Kelly, L; Khanh, TN; Kisło, K; Kiss, L; Kiyashko, A; Kolařík, M; Kruse, J; Kubátová, A; Kučera, V; Kučerová, I; Kušan, I; Lee, HB; Levicán, G; Lewis, A; Liem, NV; Liimatainen, K; Lim, HJ; Lyons, MN; Maciá-Vicente, JG; Magaña-Dueñas, V; Mahiques, R; Malysheva, EF; Marbach, PAS; Marinho, P; Matočec, N; McTaggart, AR; Mešić, A; Morin, L; MuñozMohedano, JM; Navarro-Ródenas, A; Nicolli, CP; Oliveira, RL; Otsing, E; Ovrebo, CL; Pankratov, TA; Paños, A; Paz-Conde, A; Pérez-Sierra, A; Phosri, C; Pintos, Á; Pošta, A; Prencipe, S; Rubio, E; Saitta, A; Sales, LS; Sanhueza, L; Shuttleworth, LA; Smith, J; Smith, ME; Spadaro, D; Spetik, M; Sochor, M; Sochorová, Z; Sousa, JO; Suwannasai, N; Tedersoo, L; Thanh, HM; Thao, LD; Tkalčec, Z; Vaghefi, N; Venzhik, AS; Verbeken, A; Vizzini, A; Voyron, S; Wainhouse, M; Whalley, AJS; Wrzosek, M; Zapata, M; Zeil-Rolfe, I; Groenewald, JZ. Fungal Planet description sheets: 1042–1111. Persoonia; 2020; 44, pp. 301-459.1:STN:280:DC%2BB3s7nsVWquw%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2020.44.11] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33116344][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7567971]
Crous PW, Cowan DA, Maggs‐Kölling, G, Yilmaz N, Thangavel R, Wingfield MJ, Noordeloos ME, Dima B, Brandrud TE, Jansen GM, Morozova OV, Vila J, Shivas RG, Tan YP, Bishop‐Hurley S, Lacey E, Marney TS, Larsson E, Le Floch G, Lombard L, Nodet P, Hubka V, Alvarado P, Berraf‐Tebbal A, Reyes JD, Delgado G, Eichmeier A, Jordal, JB, Kachalkin AV, Kubátová A, Maciá‐Vicente JG, Malysheva, EF, Papp V, Rajeshkumar KC, Sharma A, Spetik M, Szabóová D, Tomashevskaya MA, Abad JA, Abad ZG, Alexandrova AV, Anand G, Arenas F, Ashtekar N, Balashov S, Bañares Á, Baroncelli R, Bera I, Biketova A Yu, Blomquist CL, Boekhout T, Boertmann D, Bulyonkova TM, Burgess TI, Carnegie AJ, Cobo‐Diaz JF, Corriol G, Cunnington JH, da Cruz MO, Damm U, Davoodian N, de A Santiago ALCM, Dearnaley J, de Freitas LWS, Dhileepan K, Dimitrov R, Di Piazza S, Fatima S, Fuljer F, Galera H, Ghosh A, Giraldo A, Glushakova AM, Gorczak M, Gouliamova DE, Gramaje D, Groenewald M, Gunsch CK, Gutiérrez A, Holdom D, Houbraken J, Ismailov AB, Istel Ł, Iturriaga T, Jeppson M, Jurjević Ž, Kalinina LB, Kapitonov VI, Kautmanova I, Khalid AN, Kiran M, Kiss L, Kovács Á, Kurose D, Kusan I, Lad S, Læssøe T, Lee HB, Luangsa‐ard JJ, Lynch M, Mahamedi AE, Malysheva VF, Mateos A, Matočec N, Mešić A, Miller AN, Mongkolsamrit S, Moreno G, Morte A, Mostowfizadeh‐Ghalamfarsa R, Naseer A, Navarro‐Ródenas A, Nguyen TTT, Noisripoom W, Ntandu JE, Nuytinck J, Ostrý V, Pankratov TA, Pawłowska J, Pecenka J, Pham THG, Polhorský A, Posta A, Raudabaugh DB, Reschke K, Rodríguez A, Romero M, Rooney‐Latham S, Roux J, Sandoval‐Denis M, Smith M Th, Steinrucken TV, Svetasheva TY, Tkalčec Z, van der Linde EJ, v d Vegte M, Vauras J, Verbeken A, Visagie CM, Vitelli JS, Volobuev SV, Weill A, Wrzosek M, Zmitrovich IV, Zvyagina EA, Groenewald JZ (2021) Fungal Planet description sheets: 11821283 Persoonia: Molecular Phylogeny and Evolution of Fungi.46:313. https://doi.org/10.3767/persoonia.2021.46.11
Da Silva, MC; Naozuka, J; da Luz, JMR; de Assunção, LS; Oliveira, PV; Vanetti, MC; Bazzolli, DM; Kasuya, MC. Enrichment of Pleurotus ostreatus mushrooms with selenium in coffee husks. Food Chem; 2012; 131, pp. 558-563. [DOI: https://dx.doi.org/10.1016/j.foodchem.2011.09.023]
Dai, DQ; Wijayawardene, NN; Dayarathne, MC; Kumla, J; Han, LS; Zhang, GQ; Zhang, X; Zhang, TT; Chen, HH. Taxonomic and phylogenetic characterizations reveal four new species, two new asexual morph reports, and six new country records of bambusicolous Roussoella from China. J Fungi; 2022; 8, 532.1:CAS:528:DC%2BB38XhsFSjsbvO [DOI: https://dx.doi.org/10.3390/jof8050532]
Da Silva MDC, Nunes MD, da Luz JMR, Kasuya MCM (2013) Mycelial growth of Pleurotus spp. in Se-enriched culture media. Adv Microbiol
Dai, YC. Hymenochaetaceae (Basidiomycota) in China. Fungal Divers; 2010; 45, pp. 131-343. [DOI: https://dx.doi.org/10.1007/s13225-010-0066-9]
Dai, YC; Cui, BK; Yuan, HS; Li, BD. Pathogenic wood-decaying fungi in China. For Pathol; 2007; 37, pp. 105-120. [DOI: https://dx.doi.org/10.1111/j.1439-0329.2007.00485.x]
Damm, U; Sato, T; Alizadeh, A; Groenewald, JZ; Crous, PW. The Colletotrichum dracaenophilum, C. magnum and C. orchidearum species complexes. Stud Mycol; 2018; 90, pp. 71-118. [DOI: https://dx.doi.org/10.1016/j.simyco.2018.04.001]
Damm, U; Sato, T; Alizadeh, A; Groenewald, JZ; Crous, PW. The Colletotrichum dracaenophilum, C. magnum and C. orchidearum species complexes. Stud Mycol; 2019; 92, pp. 1-46.1:STN:280:DC%2BB3c%2Fnt1Kmsw%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2018.04.001] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29997400]
de Camargo BR, Takematsu HM, Ticona AR, da Silva LA, Silva FL, Quirino BF, Hamann PR, Noronha EF (2022) Penicillium polonicum a new isolate obtained from Cerrado soil as a source of carbohydrate-active enzymes produced in response to sugarcane bagasse. 3 Biotech 12(12):348
Decock, C; Figueroa, SH; Robledo, G; Castillo, G. Fomitiporia punctata (Basidiomycota, Hymenochaetales) and its presumed taxonomic synonyms in America: taxonomy and phylogeny of some species from tropical/subtropical areas. Mycologia; 2007; 99, pp. 733-752.1:CAS:528:DC%2BD1cXhslOls7c%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/18268907][DOI: https://dx.doi.org/10.1080/15572536.2007.11832537]
De Fine Licht HH, Boomsma JJ (2010) Forage collection, substrate preparation, and diet composition in fungus-growing ants. Ecol Entomol 35:259–269
De Fine Licht, HH; Boomsma, JJ. Variable interaction specificity and symbiont performance in Panamanian Trachymyrmex and Sericomyrmex fungus-growing ants. BMC Evol Biol; 2014; 14, pp. 1-10. [DOI: https://dx.doi.org/10.1186/s12862-014-0244-6]
De Silva, NI; Phillips, AJ; Liu, JK; Lumyong, S; Hyde, KD. Phylogeny and morphology of Lasiodiplodia species associated with Magnolia forest plants. Sci Rep; 2019; 9, pp. 1-11.1:CAS:528:DC%2BC1MXhvFersL3O [DOI: https://dx.doi.org/10.1038/s41598-019-50804-x]
de Silva, NI; Hyde, KD; Lumyong, S; Phillips, AJL; Bhat, DJ; Maharachchikumbura, SSN; Thambugala, KM; Tennakoon, DS; Suwannarach, N; Karunarathna, SC. Morphology, phylogeny, host association and geography of fungi associated with plants of Annonaceae, Apocynaceae and Magnoliaceae. Mycosphere; 2022; 13, pp. 955-1076. [DOI: https://dx.doi.org/10.5943/mycosphere/13/1/12]
de Souza, FA; Kowalchuk, GA; Leeflang, P; van Veen, JA; Smit, E. PCR-denaturing gradient gel electrophoresis profiling of inter- and intraspecies 18S rRNA gene sequence heterogeneity is an accurate and sensitive method to assess species diversity of arbuscular mycorrhizal fungi of the genus Gigaspora. Appl Environ Microbiol; 2004; 70, pp. 1413-1424.1:CAS:528:DC%2BD2cXisVKjurg%3D [DOI: https://dx.doi.org/10.1128/AEM.70.3.1413-1424.2004] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/15006761][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC368397]
DePriest, PT; Sikaroodi, M; Lawrey, JD; Diederich, P. Marchandiomyces lignicola sp. nov. shows recent and repeated transition between a lignicolous and a lichenicolous habit. Mycol Res; 2005; 109, pp. 57-70.1:CAS:528:DC%2BD2MXhtlers7w%3D [DOI: https://dx.doi.org/10.1017/S0953756204001601] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/15736863]
Desjardin DE (1989) The genus Marasmius from the southern Appalachian Mountains. PhD dissertation, University of Tennessee, Knoxville
Desjardin, DE; Perry, BA. The gymnopoid fungi (Basidiomycota, Agaricales) from the Republic of São Tomé and Príncipe, West Africa. Mycosphere; 2017; 8, pp. 1317-1391. [DOI: https://dx.doi.org/10.5943/mycosphere/8/9/5]
Diao, YZ; Chen, Q; Jiang, XZ; Houbraken, J; Barbosa, RN; Cai, L; Wu, WP. Penicillium section Lanata-divaricata from acidic soil. Cladistics; 2019; 35,
Didukh, M; Wasser, SP; Nevo, E; Ur, Y. New records of Leucocoprineae and Lepioteae in Israel. Documents Mycologiques; 2003; 23, pp. 39-58.
Diederich P (2004) Spirographa. In: Nash TH, Ryan BD, Diederich P, Gries C, Bungartz F (eds) Lichen Flora of the Greater Sonoran Desert Region, vol 2. Tempe, Arizona State University: Lichens Unlimited, Arizona State University, pp 702–703
Diederich, P; Lawrey, JD. New lichenicolous, muscicolous, corticolous and lignicolous taxa of Burgoa s. l. and Marchandiomyces s. l. (anamorphic Basidiomycota), a new genus for Omphalina foliacea, and a catalogue and a key to the non-lichenized, bulbilliferous basidiomycetes. Mycol Prog; 2007; 6, pp. 61-80. [DOI: https://dx.doi.org/10.1007/s11557-007-0523-3]
Diederich, P; Lawrey, JD; Ertz, D. The 2018 classification and checklist of lichenicolous fungi, with 2000 non-lichenized, obligately lichenicolous taxa. Bryologist; 2018; 121, pp. 340-425. [DOI: https://dx.doi.org/10.1639/0007-2745-121.3.340]
Dijksterhuis, J; Meijer, M; van Doorn, T; Houbraken, J; Bruinenberg, P. The preservative propionic acid differentially affects survival of conidia and germ tubes of feed spoilage fungi. Int J Food Microbiol; 2019; 306, 1:CAS:528:DC%2BC1MXhsVGgsrnM [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31362161][DOI: https://dx.doi.org/10.1016/j.ijfoodmicro.2019.108258] 108258.
Dissanayake, AJ; Bhunjun, CS; Maharachchikumbura, SS; Liu, JK. Applied aspects of methods to infer phylogenetic relationships amongst fungi. Mycosphere; 2020; 11, pp. 2652-2676. [DOI: https://dx.doi.org/10.5943/mycosphere/11/1/18]
Dixon, LJ; Schlub, RL; Pernezny, K; Datnoff, LE. Host specialization and phylogenetic diversity of Corynespora cassiicola. Phytopathology; 2009; 99, pp. 1015-1027.1:STN:280:DC%2BD1MrkvFWmuw%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/19671003][DOI: https://dx.doi.org/10.1094/PHYTO-99-9-1015]
Dong, W; Wang, B; Hyde, KD; McKenzie, EHC; Raja, HA; Tanaka, K; Abdel-Wahab, MA; Abdel-Aziz, FA; Doilom, M; Phookamsak, R; Hongsanan, S; Wanasinghe, DN; Yu, XD; Wang, GN; Yang, H; Yang, J; Thambugala, KM; Tian, Q; Luo, ZL; Yang, JB; Miller, AN; Fournier, J; Boonmee, S; Hu, DM; Nalumpang, S; Zhang, H. Freshwater Dothideomycetes. Fungal Divers; 2020; 105, pp. 319-575. [DOI: https://dx.doi.org/10.1007/s13225-020-00463-5]
Dong, W; Hyde, KD; Jeewon, R; Doilom, M; Yu, XD; Wang, GN; Liu, NG; Hu, DM; Nalumpang, S; Nalumpang, S. Towards a natural classification of annulatascaceae-like taxa II: introducing five new genera and eighteen new species from freshwater. Mycosphere; 2021; 12,
Donk, M. Notes on Malesian fungi I. Bull Bot Gard Buitenzorg III; 1948; 17, pp. 473-482.
Du, P; Cao, TX; Wu, YD; Zhou, M; Liu, ZB. Two new species of Hymenochaetaceae on Dracaena cambodiana from tropical China. MycoKeys; 2021; 80, pp. 1-17. [DOI: https://dx.doi.org/10.3897/mycokeys.80.63997] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34007241][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8116325]
Du, HZ; Yang, J; Liu, NG; Cheewangkoon, R; Liu, JK. Morpho-phylogenetic evidence reveals new species of Fuscosporellaceae and Savoryellaceae from freshwater habitats in Guizhou Province, China. J Fungi; 2022; 8, 1138. [DOI: https://dx.doi.org/10.3390/jof8111138]
Duduk, N; Vasić, M; Vico, I. First report of Penicillium polonicum causing blue mold on stored onion (Allium cepa) in Serbia. Plant Dis; 2014; 98,
Dubey, R. Two new species of Zygosporium Mont. from Indian subcontinent. Indian J for; 2014; 37, pp. 165-168.
Dugan, FM; Glawe, DA; Attanayake, RN; Chen, W. The importance of reporting new host-fungus records for ornamental and regional crops. Plant Health Prog; 2009; 10, 34. [DOI: https://dx.doi.org/10.1094/PHP-2009-0512-01-RV]
Dumortier BC (1822) Commentationes Botanicae. Observationes Botanique, dédiées a la Société d'horticulture de Tournay. Casterman-Dieu, Tournay, p 116. https://doi.org/10.5962/bhl.title.10534
Dupont, J; Magnin, S; Rousseau, F; Zbinden, M; Frebourg, G; Samani, S; De Forges, BR; Jones, EBG. Molecular and ultrastructural characterization of two ascomycetes found on sunken wood off Vanuatu Islands in the deep Pacific Ocean. Mycol Res; 2009; 113, pp. 1351-1364.1:CAS:528:DC%2BC3cXhtFejuro%3D [DOI: https://dx.doi.org/10.1016/j.mycres.2009.08.015] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/19737615]
Edler, D; Klein, J; Antonelli, A; Silvestro, D. raxmlGUI 2.0: A graphical interface and toolkit for phylogenetic analyses using RAxML. Methods Ecol Evol; 2021; 12, pp. 373-377. [DOI: https://dx.doi.org/10.1111/2041-210X.13512]
Ekanayaka, AH; Hyde, KD; Gentekaki, E; McKenzie, EHC; Zhao, Q; Bulgakov, TS; Camporesi, E. Preliminary classification of Leotiomycetes. Mycosphere; 2019; 10, pp. 310-489. [DOI: https://dx.doi.org/10.5943/mycosphere/10/1/7]
El-Dawy EG, Gherbawy YA, Hassan S, Hussein MA (2021) Molecular identification of Penicillium sp. isolated from citrus fruits. Curr Microbiol 78:1981–1990
Ellis, M. Dematiaceous hyphomycetes. Mycol Pap; 1960; 76, 32.
Ellis, MB. Dematiaceous Hyphomycetes. I. Mycol Pap; 1960; 76, pp. 1-36.
Ellis, MB. Dematiaceous Hyphomycetes. III. Mycol Pap; 1961; 82, pp. 1-55.
Ellis MB (1971) Dematiaceous Hyphomycetes. Commonwealth Mycological Institute, Kew, England, p 608
Ellis, JB; Everhart, BM. New species of fungi from various localities. Proc Acad Nat Sci Philadelphia; 1894; 46, pp. 322-386.
Elsik, WC. Palynology of a Palaeocene Rockdale lignite, Milam County, Texas. 1 Morphology and Taxonomy. Pollen Spores; 1968; 10, pp. 263-314.
Elsik, WC. Late Neogene palynomorph diagram, northern Gulf of Mexico. Trans Gulf Coast Assoc Geol Soc; 1969; 19, pp. 509-528.
Eriksson O (1981) The families of bitunicate Ascomycetes. Nordic J Bot 60: 800–800. https://doi.org/10.1111/j.1756-1051.1981.tb01167.x
Eriksson, E; Baral, HO; Currah, RS; Hansen, K; Kurtzman, CP; Rambold, G; Lassøe, T. Outline of Ascomycota 2003. Myconet; 2003; 7, pp. 1-89.
Etayo, J. Hongos liquenícolas de Perú: Homenaje a Rolf Santesson. Bull Soc Linn Provence; 2010; 61, pp. 83-128.
Fan, Q; Wang, Y; Zhang, G; Tang, D; Yu, H. Multigene Phylogeny and Morphology of Ophiocordyceps alboperitheciata sp. nov., A New Entomopathogenic Fungus Attacking Lepidopteran Larva from Yunnan China. Mycobiology; 2021; 49, pp. 133-141. [DOI: https://dx.doi.org/10.1080/12298093.2021.1903130] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/37970184][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10635235]
Felsenstein, J. Confidence limits on phylogenies: an approach using the bootstrap. Evolution; 1985; 39, pp. 783-791. [DOI: https://dx.doi.org/10.1111/j.1558-5646.1985.tb00420.x] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28561359]
Felšöciová S, Kačániová M, Hleba L, Petrová J, Pavelková A, Džugan M, Grabek-Lejko D (2012) Microscopic fungi isolated from Polish honey: 1040–1049
Feng, Y; Liu, JK; Lin, CG; Chen, YY; Xiang, MM; Liu, ZY. Additions to the genus Arthrinium (Apiosporaceae) from bamboos in China. Front Microbiol; 2021; 12, [DOI: https://dx.doi.org/10.3389/fmicb.2021.661281] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33936017][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8086194]661281.
Fiasson, JL; Niemelä, T. The Hymenochaetales: a revision of the European poroid taxa. Karstenia; 1984; 24, pp. 14-28. [DOI: https://dx.doi.org/10.29203/ka.1984.224]
Flakus, A; Etayo, J; Miadlikowska, J; Lutzoni, F; Kukwa, M; Matura, N; Rodriguez-Flakus, P. Biodiversity assessment of ascomycetes inhabiting Lobariella lichens in Andean cloud forests led to one new family, three new genera and 13 new species of lichenicolous fungi. Plant Fungal Syst; 2019; 64, pp. 283-344. [DOI: https://dx.doi.org/10.2478/pfs-2019-0022]
Freire KT, Araújo-Magalhães GR, Nascimento SS, Paiva LM, Barbosa RN, Bezerra JD, Souza-Motta CM (2020) First report of. Penicillium brasilianum Bat., P. cluniae Quintan., and P. echinulonalgiovense S. Abe ex Houbraken & R.N. Barbosa (Eurotiales, Aspergillaceae) as endophytes from a bromeliad in the Caatinga dry forest in Brazil. Check List 16 (4): 1055–1061
Friedl, T; Gartner, G. Trebouxia (Pleurastrales, Chlorophyta) as a phycobiont in the lichen genus Diploschistes. Archiv Far Protistenkunde; 1988; 135, pp. 147-158. [DOI: https://dx.doi.org/10.1016/S0003-9365(88)80061-7]
Fries EM (1835) Corpus florarum provincialium Sueciae I. Floram scanicam. Typis Palmblad, Sebell & Company, Upsaliae
Fries EM (1838) Epicrisis systematis mycologici. seu synopsis hymenomycetum. Typographia Academica, Upsaliæ
Fröhlich, J; Hyde, KD. New Oxydothis species associated with palm leaf spots in north Queensland, Australia. Mycol Res; 1994; 98, pp. 213-218. [DOI: https://dx.doi.org/10.1016/S0953-7562(09)80188-3]
Gao Y, Monkai J, Gentekaki E, Ren GC, Wanasinghe DN, Xu JC, Gui H (2021) Dothidea kunmingensis, a novel asexual species of Dothideaceae on Jasminum nudiflorum (winter jasmine) from Southwestern China. Phytotaxa 529:43–56. https://doi.org/10.11646/phytotaxa.529.1.3
Gäumann E (1926) Vergleichende morphologie der pilze. 1–626
Gerdemann JW, Trappe JM (1974) The Endogonaceae in the Pacific Northwest
Ghobad-Nejhad, M. Collections on Lonicera in northwest Iran represent an undescribed species in the Inonotus linteus complex (Hymenochaetales). Mycol Prog; 2015; 14, pp. 1-5. [DOI: https://dx.doi.org/10.1007/s11557-015-1100-9]
Ghobad-Nejhad, M; Nilsson, RH; Hallenberg, N. Phylogeny and taxonomy of the genus Vuilleminia (Basidiomycota) based on molecular and morphological evidence, with new insights into Corticiales. Taxon; 2010; 59, pp. 1519-1534. [DOI: https://dx.doi.org/10.1002/tax.595016]
Ghobad-Nejhad, M; Langer, E; Nakasone, K; Diederich, P; Nilsson, RH; Rajchenberg, M; Ginns, J. Digging up the roots: taxonomic and phylogenetic disentanglements in Corticiaceae s.s. (Corticiales, Basidiomycota) and evolution of nutritional modes. Front Microbiol; 2021; 12, [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34512580][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8425454][DOI: https://dx.doi.org/10.3389/fmicb.2021.704802] 704802.
Gielen, R; Põldmaa, K; Tammaru, T. In search of ecological determinants of fungal infections: a semi-field experiment with folivorous moths. Ecol Evol; 2022; 12,
Gilbertson, RL. The genus Phellinus (Aphyllophorales: Hymenochaetaceae) in western North America. Mycotaxon; 1979; 9, pp. 51-89.
Gilbertson RL, Ryvarden L (1986) North American Polypores, vol 1, Fungiflora
Glass, DLE; Brown, DD; Elsik, WC. Fungal spores from the Upper Eocene Manning Formation, Jackson Group, east and south-central Texas, U.S.A. Pollen Spores; 1986; 28, pp. 403-420.
Goh, TK; Kuo, CH. Reflections on Canalisporium, with descriptions of new species and records from Taiwan. Mycol Progress; 2021; 20, pp. 647-680. [DOI: https://dx.doi.org/10.1007/s11557-021-01689-6]
Gollotte, A; van Tuinen, D; Atkinson, D. Diversity of arbuscular mycorrhizal fungi colonizing roots of the grass species Agrostis capillaries and Lolium perenne in a field experiment. Mycorrhiza; 2004; 14, pp. 111-117. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/12768382][DOI: https://dx.doi.org/10.1007/s00572-003-0244-7]
Gómez, OC; Moreira, DMB; Luiz, JHH. Medicinal potentialities and pathogenic profile of Lasiodiplodia genus. World J Microbiol Biotechnol; 2021; 37, pp. 1-12. [DOI: https://dx.doi.org/10.1007/s11274-021-03137-9]
Gómez OC, Moreira DM, Hernández IL, Lemes RM, Luiz JH (2021) Antimicrobial activity improvement after fractionating organic extracts from Lasiodiplodia sp. fermentation. Braz J Dev 7:3795–816. https://doi.org/10.34117/bjdv7n1-257
Goos, RD. The Genus Pleurothecium. Mycologia; 1969; 61, pp. 1048-1053. [DOI: https://dx.doi.org/10.1080/00275514.1969.12018832]
Goto, BT; Silva, GA; Assis, D; Silva, DK; Souza, RG; Ferreira, AC; Jobim, K; Mello, C; Vieira, HE; Maia, LC; Oehl, F. Intraornatosporaceae (Gigasporales), a new family with two new genera and two new species. Mycotaxon; 2012; 119, pp. 117-132. [DOI: https://dx.doi.org/10.5248/119.117]
Greathouse, GA; Ames, LM. Fabric deterioration by thirteen described and three new species of Chaetomium. Mycologia; 1945; 37, pp. 138-155.1:CAS:528:DyaH28Xhtlaj [DOI: https://dx.doi.org/10.1080/00275514.1945.12023976]
Guindon, S; Dufayard, JF; Lefort, V; Anisimova, M; Hordijk, W; Gascuel, O. New algorithms and methods to estimate maximum-likelihood phylogenies: assessing the performance of PhyML 3.0. Syst Biol; 2010; 59, pp. 307-321.1:CAS:528:DC%2BC3cXks1Kms7s%3D [DOI: https://dx.doi.org/10.1093/sysbio/syq010] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/20525638]
Güssow H (1906) Uber eine neue Krankeit aliquot Gurken in England (Corynespora mazei, Güssow gen. et sp. nov.). eitschrift für Pflanzenkran-kheiten 16:10–13
Haji-Moniri, M; Gromakova, AB; Lőkös, L; Kondratyuk, SY. New members of the Megasporaceae (Pertusariales, lichen—forming Ascomycota): Megaspora iranica spec. nova and Oxneriaria gen. nova. Act Bot Hung; 2017; 59, pp. 343-370. [DOI: https://dx.doi.org/10.1556/034.59.2017.3-4.5]
Halbwachs, H; Easton, GL; Bol, R; Hobbie, EA; Garnett, MH; Persoh, D; Dixon, L; Ostle, N; Karasch, P; Griffith, GW. Isotopic evidence of biotrophy and unusual nitrogen nutrition in soil-dwelling Hygrophoraceae. Environ Microbiol; 2018; 20, pp. 3573-3588.1:CAS:528:DC%2BC1cXhvFeqsrrP [DOI: https://dx.doi.org/10.1111/1462-2920.14327] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30105856][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6849620]
Han, P; Zhang, X; Xu, D; Zhang, B; Lai, D; Zhou, L. Metabolites from Clonostachys fungi and their biological activities. J Fungi; 2020; 6, 229.1:CAS:528:DC%2BB3MXos12isLs%3D [DOI: https://dx.doi.org/10.3390/jof6040229]
Hansen L, Knudsen H (1997) Nordic Macromycetes. Heterobasidioid, aphyllophoroid and gastromycetoid Basidiomycetes. Nordsvamp, Copenhagen, Denmark. Illus 3:444
Harkness, HW. New species of California fungi. Bull South Calif Acad Sci; 1884; 1, pp. 29-47.
Hatami Rad S, Ebrahimi L, Shahbazi H (2019) First record of Epicoccum andropogonis growing on Paspalum dilatatum ergot in Iran. Mycol Iran 6:49–54. https://doi.org/10.22043/mi.2020.121093
Hattori, T; Sakayaroj, J; Jones, EBG; Suetrong, S; Preedanon, S; Klaysuban, A. Three species of Fulvifomes (Basidiomycota, Hymenochaetales) associated with rots on mangrove tree Xylocarpus granatum in Thailand. Mycoscience; 2014; 55, pp. 344-354. [DOI: https://dx.doi.org/10.1016/j.myc.2014.01.001]
Hawksworth, DL. Some new combinations in Xylaria Hill ex Grev. Trans Br Mycol Soc; 1973; 61, pp. 199-200. [DOI: https://dx.doi.org/10.1016/S0007-1536(73)80107-X]
Hawksworth, DL; Kirk, PM; Sutton, BC; Pegler, DN. Ainsworth and Bisby’s Dictionary of the Fungi; 1995; London, CAB International:
He, MQ; Zhao, RL; Hyde, KD; Begerow, D; Kemler, M; Yurkov, A; McKenzie, EHC; Raspé, O; Kakishima, M; Sánchez-Ramírez, S; Vellinga, EC; Halling, R; Papp, V; Zmitrovich, IV; Buyck, B; Ertz, D; Wijayawardene, NN; Cui, BK; Schoutteten, N; Liu, XZ; Li, TH; Yao, YJ; Zhu, XY; Liu, AQ; Li, GJ; Zhang, MZ; Ling, ZL; Cao, B; Antonín, V; Boekhout, T; da Silva, BDB; De Crop, E; Decock, C; Dima, B; Dutta, AK; Fell, JW; Geml, J; Ghobad-Nejhad, M; Giachini, AJ; Gibertoni, TB; Gorjón, SP; Haelewaters, D; He, SH; Hodkinson, BP; Horak, E; Hoshino, T; Justo, A; Lim, YW; Menolli, N, Jr; Mešić, A; Moncalvo, JM; Mueller, GM; Nagy, LG; Nilsson, RH; Noordeloos, M; Nuytinck, J; Orihara, T; Ratchadawan, C; Rajchenberg, M; Silva-Filho, AGS; Sulzbacher, MA; Tkalčec, Z; Valenzuela, R; Verbeken, A; Vizzini, A; Wartchow, F; Wei, TZ; Weiß, M; Zhao, CL; Kirk, PM. Notes, outline and divergence times of Basidiomycota. Fungal Divers; 2019; 99, pp. 105-367. [DOI: https://dx.doi.org/10.1007/s13225-019-00435-4]
He MQ, Cao B, Liu F, Boekhout T, Denchev TT, Schoutteten N, Denchev CM, Kemler M, Gorjón SP, Begerow D, Valenzuela R, Davoodian N, Niskanen T, Vizzini A, Redhead SA, Ramírez-Cruz V, Papp V, Dudka VA, Dutta AK, García-Sandoval R, Liu X-Z, Kijpornyongpan T, Savchenko A, Tedersoo L, Theelen B, Trierveiler-Pereira L, Wu F, Carlos Zamora J, Zeng X-Y, Zhou L-W, Liu S-L, Ghobad-Nejhad M, Giachini AJ, Li G-J, Kakishima M, Olariaga I, Haelewaters D, Sulistyo B, Sugiyama J, Svantesson S, Yurkov A, Alvarado P, Antonín V, Felipe da Silva A, Druzhinina I, Gibertoni TB, Guzmán-Dávalos L, Justo A, Karunarathna SC, Galappaththi MCA, Toome-Heller M, Hosoya T, Liimatainen K, Márquez R, Mešić A, Moncalvo J-M, Nagy LG, Varga T, Orihara T, Raymundo T, Salcedo I, Silva-Filho AGS, Tkalčec Z, Wartchow F, Zhao C-L, Bau T, Cabarroi Hernández M, Cortés-Pérez A, Decock C, De Lange R, Weiß M, Menolli Jr N, Nilsson RH, Fan Y-G, Verbeken A, Gafforov Y, Meiras-Ottoni A, Mendes-Alvarenga RL, Zeng N-K, Wu Q, Hyde KD, Kirk PM, Zhao RL (2024) Phylogenomics, divergence times and notes of orders in Basidiomycota. Fungal Divers 126:127–406. https://doi.org/10.1007/s13225-024-00535-w
Hernãndez-Restrepo, M; Bezerra, JD; Tan, YP; Wiederhold, N; Crous, PW; Guarro, J; Gené, J. Re-evaluation of Mycoleptodiscus species and morphologically similar fungi. Persoonia-Mol Phylogeny Evol Fungi; 2019; 42, pp. 205-227. [DOI: https://dx.doi.org/10.3767/persoonia.2019.42.08]
Hibbett, DS; Binder, M; Bischoff, JF; Blackwell, M; Cannon, PF; Eriksson, OE; Huhndorf, SM; James, T; Kirk, PM; Lücking, R; Lumbsch, HT. A higher-level phylogenetic classification of the Fungi. Mycol Res; 2007; 111, pp. 509-547. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/17572334][DOI: https://dx.doi.org/10.1016/j.mycres.2007.03.004]
Hidayat, I; Jeewon, R; To-anun, C; Hyde, KD. The genus Oxydothis: New palmicolous taxa and phylogenetic relationships within Xylariales. Fungal Divers; 2006; 23, 159179.
Hilário, S; Lopes, A; Santos, L; Alves, A. Botryosphaeriaceae species associated with blueberry stem blight and dieback in the Centre Region of Portugal. Eur J Plant Pathol; 2020; 156, pp. 31-44. [DOI: https://dx.doi.org/10.1007/s10658-019-01860-6]
Hiratsuka, N; Chen, ZC. A list of Uredinales collected from Taiwan. Trans Mycol Soc Jpn; 1991; 32, pp. 3-22.
Hjortstam, K; Ryvarden, L. Lopharia and Porostereum (Corticiacae). Synop Fungorum; 1990; 4, pp. 1-68.
Hoang, DT; Chernomor, O; von Haeseler, A; Minh, BQ; Vinh, LS. UFBoot2: Improving the ultrafast bootstrap approximation. Mol Biol Evol; 2018; 35, pp. 518-522.1:CAS:528:DC%2BC1cXitlyjs7rK [DOI: https://dx.doi.org/10.1093/molbev/msx281] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29077904]
Hongsanan, S; Hyde, KD; Phookamsak, R; Wanasinghe, DN et al. Refined families of Dothideomycetes: Dothideomycetidae and Pleosporomycetidae. Mycosphere; 2020; 11, pp. 1553-2107. [DOI: https://dx.doi.org/10.5943/mycosphere/11/1/13]
Hongsanan, S; Hyde, KD; Phookamsak, R et al. Refined families of Dothideomycetes: orders and families incertae sedis in Dothideomycetes. Fungal Divers; 2020; 105, pp. 17-318. [DOI: https://dx.doi.org/10.1007/s13225-020-00462-6]
Hongsanan S, Norphanphoun C, Thambugala KM, Boonyuen N, Huanraluek N, Cheewangkoon R (2020c) Taxonomy and phylogeny of Muyocopron thailandica sp. nov. Phytotaxa. 456:195–202. https://doi.org/10.11646/PHYTOTAXA.456.2.7
Hongsanan, S; Norphanphoun, C; Senanayake, IC; Jayawardena, RS; Manawasinghe, IS; Abeywickrama, PD; Khuna, S; Suwannarach, N; Senwanna, C; Monkai, J; Hyde, KD; Gentekaki, E; Bhunjun, CS. Annotated notes on Diaporthe species. Mycosphere; 2023; 14,
Hou, L; Hernández-Restrepo, M; Groenewald, JZ; Cai, L; Crous, PW. Citizen science project reveals high diversity in Didymellaceae (Pleosporales, Dothideomycetes). MycoKeys; 2020; 65, 49. [DOI: https://dx.doi.org/10.3897/mycokeys.65.47704] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32206025][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7078340]
Houbraken, JA; Samson, R. Phylogeny of Penicillium and the segregation of Trichocomaceae into three families. Stud Mycol; 2011; 70, pp. 1-51.1:STN:280:DC%2BC383gsFKnsQ%3D%3D [DOI: https://dx.doi.org/10.3114/sim.2011.70.01] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22308045][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3233907]
Houbraken J, Lopez-Quintero CA, Frisvad JC, Boekhout T, Theelen B, Franco-Molano AE, Samson RA (2011) Penicillium araracuarense sp. nov., Penicillium elleniae sp. nov., Penicillium penarojense sp. nov., Penicillium vanderhammenii sp. nov. and Penicillium wotroi sp. nov., isolated from leaf litter. Int J Syst Evolut Microbiol 61(6):1462–1475
Houbraken, J; Kocsubé, S; Visagie, CM; Yilmaz, N; Wang, XC; Meijer, M; Kraak, B; Hubka, V; Bensch, K; Samson, RA; Frisvad, JC. Classification of Aspergillus, Penicillium, Talaromyces and related genera (Eurotiales): An overview of families, genera, subgenera, sections, series and species. Stud Mycol; 2020; 95, pp. 51-69. [DOI: https://dx.doi.org/10.1016/j.simyco.2020.05.002]
Hsieh, SY; Chang, HS; Jones, EBG; Read, SJ; Moss, ST. Halosarpheia aquadulcis sp. nov., a new lignicolous, freshwater ascomycete from Taiwan. Mycol Res; 1995; 99, pp. 49-53. [DOI: https://dx.doi.org/10.1016/S0953-7562(09)80315-8]
Hsieh SY, Goh TK, Kuo CH (2021) A taxonomic revision of Stanjehughesia (Chaetosphaeriaceae, Sordariomycetes), with a novel species S. kaohsiungensis from Taiwan. Phytotaxa 484:261–280. https://doi.org/10.11646/PHYTOTAXA.484.3.2
Hu, J; Zhao, G; Tuo, Y; Rao, G; Zhang, Z; Qi, Z; Yue, L; Liu, Y; Zhang, T; Li, Y; Zhang, B. Morphological and molecular evidence reveal eight new species of Gymnopus from Northeast China. J Fungi; 2022; 8,
Huang, F; Chen, GQ; Hou, X; Fu, YS; Cai, L; Hyde, KD; Li, HY. Colletotrichum species associated with cultivated citrus in China. Fungal Divers; 2013; 61, pp. 61-74. [DOI: https://dx.doi.org/10.1007/s13225-013-0232-y]
Hughes, SJ. Studies on micro-fungi X. Zygosporium. Mycol Pap; 1951; 44, pp. 1-18.
Hughes, SJ. Studies on micro-fungi. III. Mastigosporium, Camposporium, and Ceratophoram. Mycol Pap; 1951; 36, pp. 1-43.
Hughes, SJ. Fungi from the Gold Coast. I Mycol Pap; 1952; 48, pp. 1-91.
Humphries, Z; Seifert, KA; Hirooka, Y; Visagie, CM. A new family and genus in Dothideales for Aureobasidium-like species isolated from house dust. Ima Fungus; 2017; 8, pp. 299-315. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29242777][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5729714][DOI: https://dx.doi.org/10.5598/imafungus.2017.08.02.05]
Hussain, S; Jabeen, S; Khalid, AN; Ahmad, H; Afshan, NuS; Sher, H; Pfister, DH. Underexplored regions of Pakistan yield five new species of Leucoagaricus. Mycologia; 2018; 110, pp. 387-400. [DOI: https://dx.doi.org/10.1080/00275514.2018.1439651] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29737930]
Hyde, KD. Intertidal fungi from warm temperate mangroves of Australia, including Tunicatispora australiensis, gen. et. sp. nov. Austr Syst Bot; 1990; 3, pp. 711-718. [DOI: https://dx.doi.org/10.1071/SB9900711]
Hyde, KD. Fungi from palms. XIII. The genus Oxydothis, a revision. Sydowia; 1994; 46, pp. 265-314.
Hyde, KD; Jones, EBG; Liu, JK; Ariyawansa, HA; Boehm, E; Boonmee, S; Braun, U; Chomnunti, P; Crous, PW; Dai, DQ; Diederich, P; Dissanayake, AJ; Doilom, M; Doveri, F; Hongsanan, S; Jayawardena, R; Lawrey, JD; Li, YM; Liu, YX; Lücking, R; Monkai, J; Muggia, L; Nelsen, MP; Pang, KL; Phookamsak, R; Senanayake, I; Shearer, CA; Seutrong, S; Tanaka, K; Thambugala, KM; Wijayawardene, NN; Wikee, S; Wu, HX; Zhang, Y; Aguirre-Hudson, B; Alias, SA; Aptroot, A; Bahkali, AH; Bezerra, JL; Bhat, JD; Camporesi, E; Chukeatirote, E; Gueidan, C; Hawksworth, DL; Hirayama, K; De Hoog, S; Kang, J-C; Knudsen, K; Li, WJ; Li, X; 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; Yan, JY; Yacharoen, S; Zhang, M. Families of Dothideomycetes. Fungal Divers; 2013; 63, pp. 1-313. [DOI: https://dx.doi.org/10.1007/s13225-013-0263-4]
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-Ruíz, RF; Chomnunti, P; Doilom, M; Dovana, F; Gibertoni, TB; Jadan, M; Jeewon, R; Jones, EBG; Kang, JC; Karunarathna, SC; Lim, YW; Liu, JK; Liu, ZY; Plautz, HL, Jr; Lumyong, S; Maharachchikumbura, SSN; Matočec, N; McKenzie, EHC; Mešić, 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. Fungal diversity notes 603–708: taxonomic and phylogenetic notes on genera and species. Fungal Divers; 2017; 87, pp. 1-235. [DOI: https://dx.doi.org/10.1007/s13225-017-0391-3]
Hyde, KD; Chaiwan, N; Norphanphoun, C; Boonmee, S; Camporesi, E; Chethana, KW; Dayarathne, MC; De Silva, NI; Dissanayake, AJ; Ekanayaka, AH; Hongsanan, S. Mycosphere notes 169–224. Mycosphere; 2018; 9, pp. 271-430. [DOI: https://dx.doi.org/10.5943/mycosphere/9/2/8]
Hyde, KD; Tennakoon, DS; Jeewon, R; Bhat, DJ et al. Fungal diversity notes 1036–1150: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers; 2019; 96, pp. 1-242. [DOI: https://dx.doi.org/10.1007/s13225-019-00429-2]
Hyde, KD; Dong, Y; Phookamsak, R; Jeewon, R; Bhat, DJ; Jones, EBG; Liu, NG; Abeywickrama, PD; Mapook, A; Wei, DP; Perera, RH; Manawasinghe, IS; Pem, D; Bundhun, D; Karunarathna, A; Ekanayaka, AH; Bao, DF; Li, JF; Samarakoon, MC; Chaiwan, N; Lin, CG; Phutthacharoen, K; Zhang, SN; Senanayake, IC; Goonasekara, ID; Thambugala, KM; Phukhamsakda, C; Tennakoon, DS; Jiang, HB; Yang, J; Zeng, M; Huanraluek, N; Liu, JK; Wijesinghe, SN; Tian, Q; Tibpromma, S; Brahmanage, RS; Boonmee, S; Huang, SK; Thiyagaraja, V; Lu, YZ; Jayawardena, RS; Dong, W; Yang, EF; Singh, SK; Singh, SM; Rana, S; Lad, SS; Anand, G; Devadatha, B; Niranjan, M; Sarma, VV; Liimatainen, K; Aguirre- Hudson, B; Niskanen, T; Overall, A; Alvarenga, RLM; Gibertoni, TB; Pfiegler, WP; Horváth, E; Imre, A; Alves, AL; da Silva Santos, AC; Tiago, PV; Bulgakov, TS; Wanasinghe, DN; Bahkali, AH; Doilom, M; Elgorban, AM; Maharachchikumbura, SSN; Rajeshkumar, KC; Haelewaters, D; Mortimer, PE; Zhao, Q; Lumyong, S; Xu, JC; Sheng, J. Fungal diversity notes 1151–1276: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers; 2020; 100, pp. 5-277. [DOI: https://dx.doi.org/10.1007/s13225-020-00439-5]
Hyde, KD; de Silva, NI; Jeewon, R; Bhat, DJ; Phookamsak, R; Doilom, M; Boonmee, S; Jayawardena, RS; Maharachchikumbura, SSN; Senanayake, IC; Manawasinghe, IS; Liu, NG; Abeywickrama, PD; Chaiwan, N; Karunarathna, A; Pem, D; Lin, CG; Sysouphanthong, P; Luo, ZL; Wei, DP; Wanasinghe, DN; Norphanphoun, C; Tennakoon, DS; Samarakoon, MC; Jayasiri, SC; Jiang, HB; Zeng, XY; Li, JF; Wijesinghe, SN; Devadatha, B; Goonasekara, ID; Brahmanage, RS; Yang, EF; Aluthmuhandiram, JVS; Dayarathne, MC; Marasinghe, DS; Li, WJ; Dissanayake, LS; Dong, W; Huanraluek, N; Lumyong, S; Liu, JK; Karunarathna, SC; Jones, EBG; Al-Sadi, AM; Xu, JC; Harishchandra, D; Sarma, VV. AJOM new records and collections of fungi: 1–100. Asian J Mycol; 2020; 3, pp. 22-294. [DOI: https://dx.doi.org/10.5943/ajom/3/1/3]
Hyde, KD; Norphanphoun, C; Maharachchikumbura, SSN; Bhat, DJ; Jones, EBG; Bundhun, D; Chen, YJ; Bao, DF; Boonmee, S; Calabon, MS et al. Refined families of Sordariomycetes. Mycosphere; 2020; 11, pp. 305-1059. [DOI: https://dx.doi.org/10.5943/mycosphere/11/1/7]
Hyde KD, Jeewon R, Chen YJ, Bhunjun CS, Calabon MS, Jiang HB, Lin CG, Norphanphoun C, Sysouphanthong P, Pem D, Tibpromma S (2020d) The numbers of fungi: is the descriptive curve flattening?. Fungal Divers 219–271
Hyde, KD; Suwannarach, N; Jayawardena, RS; Manawasinghe, IS; Liao, CF; Doilom, M; Cai, L; Zhao, P; Buyck, B; Phukhamsakda, C; Su, WX. Mycosphere notes 325–344: novel species and records of fungal taxa from around the world. Mycosphere; 2021; 13, pp. 1101-1156. [DOI: https://dx.doi.org/10.5943/mycosphere/12/1/14]
Hyde KD, Bao DF, Hongsanan S, Chethana KWT, Yang J, Suwannarach N (2021b) Evolution of freshwater Diaporthomycetidae (Sordariomycetes) provides evidence for five new orders and six new families. Fungal Diversity 107:71105. https://doi.org/10.1007/s13225-021-00469-7
Hyde, KD; Amuhenage, TB; Apurillo, CCS; Asghari, R; Aumentado, HD; Bahkali, AH; Bera, I; Bhunjun, CS; Calabon, MS; Chandrasiri, S; Chethana, KWT; Doilom, M; Dong, W; Fallahi, M; Gajanayake, AJ; Gomdola, D; Gunarathne, A; Hongsanan, S; Huanraluek, N; Jayawardena, RS; Kapov, SA; Khyaju, S; Le, L; Li, CJY; Li, QR; Li, YX; Lin, CG; Linn, MM; Liu, JK; Liu, NG; Luangharn, T; Madagammana, AD; Manawasinghe, IS; Marasinghe, DS; McKenzie, EHC; Meakala, N; Meng, Q F; Mukhopadhyay, S; Norphanphoun, C; Pem, D; Phukhamsakda, C; Sarma, VV; Selcuk, F; Senanayake, IC; Shah, S; Shu, YX; Silva, HVS; Su, HL; Tavakol, M; Thakshila, SAD; Thiyagaraja, V; Thongklang, N; Tian, Q; Tibpromma, S; Tun, ZL; Ulukapi, M; Wang, Y; Wannasawang, N; Wijayawardene, NN; Wimalasena, SDMK; Xiao, Y; Xiong, YR; Yasanthika, WAE; Li, Q; Dai, DQ. Fungalpedia, an illustrated compendium for the fungi and fungus-like taxa. Mycosphere; 2023; 14, pp. 1835-1959. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/22]
Hyde, KD; Norphanphoun, C; Ma, J; Yang, HD; Zhang, JY; Du, TY; Gao, Y; Gomes de Farias, AR; He, SC; He, YK; Li, CJY; Li, JY; Liu, XF; Lu, L; Su, HL; Tang, X; Tian, XG; Wang, SY; Wei, DP; Xu, RF; Xu, RJ; Yang, YY; Zhang, F; Zhang, Q; Bahkali, AH; Boonmee, S; Chethana, KWT; Jayawardena, RS; Lu, YZ; Karunarathna, SC; Tibpromma, S; Wang, Y; Zhao, Q. Mycosphere notes 387–412 – novel species of fungal taxa from around the world. Mycosphere; 2023; 14,
Ichinoe, M. Camposporium species from Japan. Trans Mycol Soc Jpn; 1971; 12, pp. 79-88.
Index Fungorum (2024) Index Fungorum. https://www.indexfungorum.org [May 10, 2024]
Index Fungorum (2024) https://www.indexfungorum.org/names/Names.asp
Iqbal MS, Usman M, Habib K, Khalid AN (2023) Oxneriaria pakistanica sp. nov. (Megasporaceae, Pertusariales, Ascomycota) from Darel Valley, Gilgit Baltistan, Pakistan. Phytotaxa 579:125–131. https://doi.org/10.11646/phytotaxa.579.2.6
Jayasiri, SC; Hyde, KD; Ariyawansa, HA et al. The Faces of Fungi database: fungal names linked with morphology, phylogeny and human impacts. Fungal Divers; 2015; 74, pp. 3-18. [DOI: https://dx.doi.org/10.1007/s13225-015-0351-8]
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. Diversity, morphology and molecular phylogeny of Dothideomycetes on decaying wild seed pods and fruits. Mycosphere; 2019; 10, pp. 1-186. [DOI: https://dx.doi.org/10.5943/mycosphere/10/1/1]
Jayawardena, RS; Bhunjun, CS; Hyde, KD; Gentekaki, E; Itthayakorn, P. Colletotrichum: lifestyles, biology, morpho-species, species complexes and accepted species. Mycosphere; 2021; 12, pp. 519-669. [DOI: https://dx.doi.org/10.5943/mycosphere/12/1/7]
Jayawardena, RS; Hyde, KD; de Farias, AR; Bhunjun, CS; Fernandez, HS; Manamgoda, DS; Udayanga, D; Herath, IS; Thambugala, KM; Manawasinghe, IS; Gajanayake, AJ. What is a species in fungal plant pathogens?. Fungal Divers; 2021; 109, pp. 239-266. [DOI: https://dx.doi.org/10.1007/s13225-021-00484-8]
Jayawardena, RS; Hyde, KD; Wang, S; Sun, YR; Suwannarach, N; Sysouphanthong, P; Abdel-Wahab, MA; Abdel-Aziz, FA; Abeywickrama, PD; Abreu, VP; Armand, A. Fungal diversity notes 1512–1610: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers; 2022; 117, pp. 1-272. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36852303][DOI: https://dx.doi.org/10.1007/s13225-022-00513-0]
Jeewon, R; Hyde, KD. Establishing species boundaries and new taxa among fungi: recommendations to resolve taxonomic ambiguities. Mycosphere; 2016; 7, pp. 1669-1677. [DOI: https://dx.doi.org/10.5943/mycosphere/7/11/4]
Jiang H, Zhang M, Zhou Y, Li X, Li J, Song C (2022) First report of Muyocopron laterale causing a new leaf disease of Camellia sinensis in China. Beverage Plant Res 2:1–4. https://doi.org/10.48130/BPR-2022-0010
Jones, EBG. Fungal adhesion. Mycol Res; 1994; 98, pp. 961-981. [DOI: https://dx.doi.org/10.1007/978-3-319-46082-6_2]
Jülich, W. Studies in resupinate Basidiomycetes V. Some new genera and species. Persoonia; 1978; 10, pp. 137-140.
Jülich, W. Higher taxa of basidiomycetes. Bibl Mycol; 1982; 85, pp. 1-485.
Justo, A; Miettinen, O; Floudas, D; Ortiz-Santana, B; Sjökvist, E; Lindner, D; Nakasone, K; Nieneläb, T; Larsson, KH; Ryvarden, L; Hibbett, DS. A revised family-level classification of the Polyporales (Basidiomycota). Fungal Biol; 2017; 121, pp. 798-824. [DOI: https://dx.doi.org/10.1016/j.funbio.2017.05.010] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28800851]
Justo, A; Ci, CA; Bizzi, A. The genera Leucoagaricus and Leucocoprinus in the Dominican Republic. Mycologia; 2021; 113, pp. 348-389. [DOI: https://dx.doi.org/10.1080/00275514.2020.1819142] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33481687]
Kalgutkar, RM; Jansonius, J. Synopsis of fungal spores, mycelia and fructifications. AASP Contrib Ser; 2000; 39, pp. 1-423.
Kalichman, J; Kirk, PM; Matheny, PB. A compendium of generic names of agarics and Agaricales. Taxon; 2020; 69, pp. 425-447. [DOI: https://dx.doi.org/10.1002/tax.12240]
Kalkan, SO; Bozcal, E; Hames Tuna, EE; Uzel, A. Characterisation of a thermostable and proteolysis resistant phytase from Penicillium polonicum MF82 associated with the marine sponge Phorbas sp. Biocatal Biotransform; 2020; 38,
Kang, JC; Kong, RYC; Hyde, KD. Studies on the Amphisphaeriales 1. Amphisphaeriaceae (sensu stricto) and its phylogenetic relationships inferred from 5.8S rDNA and ITS2 sequences. Fungal Divers; 1998; 1, pp. 147-157.
Kang, JC; Hyde, KD; Kong, RYC. Studies on Amphisphaeriales: the Amphisphaeriaceae (sensu stricto). Mycol Res; 1999; 103, pp. 53-64. [DOI: https://dx.doi.org/10.1017/S0953756298006650]
Kang, JC; Hyde, KD; Kong, RYC. Studies on Amphisphaeriales: the Cainiaceae. Mycol Res; 1999; 103, pp. 1621-1627. [DOI: https://dx.doi.org/10.1017/S0953756299001264]
Karimi, O; Chethana, KT; de Farias, AR; Asghari, R; Kaewchai, S; Hyde, KD; Li, Q. Morphology and multigene phylogeny reveal three new species of Distoseptispora (Distoseptisporales, Distoseptisporaceae) on palms (Arecaceae) from peatswamp areas in southern Thailand. MycoKeys; 2024; 102, 55. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38370856][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10873808][DOI: https://dx.doi.org/10.3897/mycokeys.102.112815]
Karsten, P. Enumeratio Boletinearum et Polyporearum Fennicarum, systemate novo dispositarum. Rev Mycol Toulouse; 1881; 3, pp. 16-19.
Karunarathna A, Działak P, Jayawardena RS, Karunarathna SC, Kuo CH, Suwannarach N, Tibpromma S, Lumyong S (2021) A novel addition to the Pezizellaceae (Rhytismatales, Ascomycota). Phytotaxa 480:251–261. https://doi.org/10.11646/phytotaxa.480.3.4
Kasuya T, Hosaka K, Ji JX, Kakishima M (2024) Gymnosporangium mori comb. nov. (Pucciniales) for Caeoma mori (≡ Aecidium mori) inferred from phylogenetic evidence. Mycoscience 65:79–85. https://doi.org/10.47371/mycosci.2023.1.001
Katoh, K; Rozewicki, J; Yamada, KD. MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Brief Bioinform; 2019; 20, pp. 1160-1166.1:CAS:528:DC%2BB3cXht1Kqtb3K [DOI: https://dx.doi.org/10.1093/bib/bbx108] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28968734]
Khalil, AM; Hashem, AH; Abdelaziz, AM. Occurrence of toxigenic Penicillium polonicum in retail green table olives from the Saudi Arabia market. Biocatal Agric Biotechnol; 2019; 21, [DOI: https://dx.doi.org/10.1016/j.bcab.2019.101314] 101314.
Khan, MY; Dahot, MU; Khan, MY. Single cell protein production by Penicillium javanicum from pretreated rice husk. Med J Islamic World Acad Sci; 1992; 5,
Khokhar, I; Jia, Y; Mukhtar, I; Wang, J; Yan, Y. First report of Penicillium polonicum causing blue mold on stored pear (Pyrus bretschneideri) fruits in China. Plant Dis; 2019; 103,
Kularathnage, ND; Tennakoon, DS; Zhu, X; Zhou, J; Su, B; Xie, Y; Chen, Q; Calabon, MS; Kirk, PM; Senanayake, IC; Doilom, M; Xu, B; Dong, W; Song, J. Reinstating Dyfrolomyces and introducing Melomastia pyriformis sp. nov. (Pleurotremataceae, Dyfrolomycetales) from Guangdong Province, China. Curr Res Environ Appl Mycol; 2023; 13, pp. 426-438. [DOI: https://dx.doi.org/10.5943/cream13/1/16]
Ke Shi (1978) (pseudonym of Sung ZC, Tsao L, Chou HI, Kwang HL, Wang KT) Early Tertiary spores and pollen grains from the coastal region of the Bohai (in Chinese). Academy of Petroleum Exploration, Development and Planning Research of the Ministry of Petroleum and Chemical Industries and the Nanjing Institute of Geology, and Paleontology, Chinese Academy of Sciences, Kexue Chubanshe, Peking, p 177
Kepler, RM; Humber, RA; Bischoff, JF; Rehner, SA. Clarification of generic and species boundaries for Metarhizium and related fungi through multigene phylogenetics. Mycologia; 2014; 106, pp. 811-829. [DOI: https://dx.doi.org/10.3852/13-319] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/24891418]
Kern, FD. A revised taxonomic account of Gymnosporangium; 1973; Pennsylvania, Pennsylvania State University Press:
Kirk PM (2019) Catalogue of Life. http://www.catalogueoflife.org
Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Ainsworth & Bisby’s dictionary of the fungi 10th. CAB International 1–771
Kirschner, R; Villarreal, RV; Hofmann, TA. A new species of Shrungabeeja (Tetraplosphaeriaceae) and new records of saprobic dematiaceous hyphomycetes on Poaceae from the high mountains of Panama. Sydowia; 2017; 69, pp. 153-160.
Kohlmeyer, J. New genera and species of higher fungi from the deep sea (1615–5315 m). Revue De Mycologie; 1977; 41, pp. 189-206.
Kohlmeyer, J; Kohlmeyer, E. Bermuda marine fungi. Trans Br Mycol Soc; 1977; 68, pp. 207-219. [DOI: https://dx.doi.org/10.1016/S0007-1536(77)80010-7]
Konta, SHS; Tibpromma, S; Thongbai, B; Maharachchikumbura, SSN; Bahkali, AH; Boonmee, HKD; S, ,. An advance in the endophyte story: Oxydothidaceae fam. nov. with six new species of Oxydothis. Mycosphere; 2016; 7, pp. 1425-1446. [DOI: https://dx.doi.org/10.5943/mycosphere/7/9/15]
Konta, S; Tibpromma, S; Karunarathna, SC; Samarakoon, MC; Steven, LS; Mapook, A; Boonmee, S; Senwanna, C; Balasuriya, A; Eungwanichayapant, PD; Hyde, KD. Morphology and multigene phylogeny reveal ten novel taxa in Ascomycota from terrestrial palm substrates (Arecaceae) in Thailand. Mycosphere; 2023; 14, pp. 107-152. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/2]
Kornerup, A; Wanscher, JH. Methuen handbook of colour; 1978; 3 London, Eyre Methuen Ltd.:
Koukol, O; Delgado, G. Why morphology matters: the negative consequences of hasty descriptions of putative novelties in asexual ascomycetes. IMA Fungus; 2021; 12, 26. [DOI: https://dx.doi.org/10.1186/s43008-021-00073-z] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34551825][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8459516]
Kowalski, DT. Observations on the genus Lamproderma. Mycologia; 1968; 60, pp. 756-768. [DOI: https://dx.doi.org/10.1080/00275514.1968.12018636]
Kowalski, DT. The species of Lamproderma. Mycologia; 1970; 62, pp. 621-672.1:STN:280:DyaE3M%2FltlWgtQ%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/5530316][DOI: https://dx.doi.org/10.1080/00275514.1970.12019010]
Kozlov, AM; Darriba, D; Flouri, T; Morel, B; Stamatakis, A. RAxML-NG: A fast, scalable, and user-friendly tool for maximum likelihood phylogenetic inference. Bioinformatics; 2019; 35, pp. 4453-4455.1:CAS:528:DC%2BB3cXhtFalsL3F [DOI: https://dx.doi.org/10.1093/bioinformatics/btz305] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31070718][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6821337]
Krug, JC. The genus Cainia and a new family, Cainiaceae. Sydowia; 1978; 30, pp. 122-133.
Kuhner R (1980) Les Hyménomycètes agaricoides. Bulletin Mensuel de la Société Linnéenne de Lyon. 49:1–1027
Kumar, P. Fungal remains from the Miocene Quilon Bed of Kerala state, South India. Rev Palaeobot Palynol; 1990; 63, pp. 13-28. [DOI: https://dx.doi.org/10.1016/0034-6667(90)90003-2]
Kumar, S; Stecher, G; Li, M; Knyaz, C; Tamura, K. MEGA X: molecular evolutionary genetics analysis across computing platforms. Mol Biol Evol; 2018; 35, pp. 1547-1549.1:CAS:528:DC%2BC1MXis1Ontrc%3D [DOI: https://dx.doi.org/10.1093/molbev/msy096] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29722887][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5967553]
Kwon, H; Nguyen, QN; Na, MW; Kim, KH; Guo, Y; Yim, JH; Shim, SH; Kim, JJ; Kang, KS; Lee, D. A new α-pyrone from Arthrinium pseudosinense culture medium and its estrogenic activity in MCF-7 cells. J Antibiot; 2021; 74, pp. 893-897.1:CAS:528:DC%2BB3MXhvFyjtL%2FK [DOI: https://dx.doi.org/10.1038/s41429-021-00473-8]
Lamarck, JB; De Candolle, AP. Flore française, tome second; 1805; Paris, France:
Lamsal, K; Kim, SW; Naeimi, S; Adhikari, M; Yadav, DR; Kim, C; Lee, HB; Lee, YS. Three new records of Penicillium species isolated from insect specimens in Korea. Mycobiology; 2013; 41,
Lanfear, R; Calcott, B; Ho, SY; Guindon, S. PartitionFinder: combined selection of partitioning schemes and substitution models for phylogenetic analyses. Mol Biol Evol; 2012; 29, pp. 1695-1701.1:CAS:528:DC%2BC38Xnt1ehsbg%3D [DOI: https://dx.doi.org/10.1093/molbev/mss020] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22319168]
Lanfear, R; Frandsen, PB; Wright, AM; Senfeld, T; Calcott, B. PartitionFinder 2: New methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Mol Biol Evol; 2017; 34, pp. 772-773.1:CAS:528:DC%2BC1cXitFOmtrvM [DOI: https://dx.doi.org/10.1093/molbev/msw260] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28013191]
Larsson, KH. Re-thinking the classification of corticioid fungi. Mycol Res; 2007; 111, pp. 1040-1063. [DOI: https://dx.doi.org/10.1016/j.mycres.2007.08.001] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/17981020]
Lawrey, JD; Diederich, P; Sikaroodi, M; Gillevet, P. Remarkable nutritional diversity in the Corticiales, including a new foliicolous species of Marchandiomyces (anamorphic Basidiomycota, Corticiaceae) from Australia. Am J Bot; 2008; 95, pp. 816-823. [DOI: https://dx.doi.org/10.3732/ajb.0800078] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/21632407]
Leão-Ferreira, SM; Gusmão, LFP. Conidial fungi from the semi-arid Caatinga biome of Brazil. New species of Endophragmiella and Spegazzinia with new records for Brazil, South America, and Neotropica. Mycotaxon; 2010; 111, pp. 1-10. [DOI: https://dx.doi.org/10.5248/111.1]
Li, GJ; Hyde, KD; Zhao, RL; Hongsanan, S; Abdel-Aziz, FA; Abdel-Wahab, MA; Alvarado, P; Alves-Silva, G; Ammirati, JF; Ariyawansa, HA; Baghela, A. Fungal diversity notes 253–366: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers; 2016; 78, pp. 1-237.1:CAS:528:DC%2BC1cXis1equrk%3D [DOI: https://dx.doi.org/10.1007/s13225-016-0366-9]
Li, WL; Maharachchikumbura, SSN; Cheewangkoon, R; Liu, JK. Reassessment of Dyfrolomyces and Four New Species of Melomastia from Olive (Olea europaea) in Sichuan Province, China. J Fungi; 2022; 8, 76.1:CAS:528:DC%2BB38Xnt1Wksr8%3D [DOI: https://dx.doi.org/10.3390/jof8010076]
Li, M; Raza, M; Song, S; Hou, LW; Zhang, ZF; Gao, M; Huang, JE; Liu, F; Cai, L. Application of culturomics in fungal isolation from mangrove sediments. Microbiome; 2023; 11, pp. 1-87.1:CAS:528:DC%2BB3sXisF2isrjP [DOI: https://dx.doi.org/10.1186/s40168-023-01708-6]
Li, WL; Liang, RR; Dissanayake, AJ; Liu, JK. Mycosphere Notes 413–448: Dothideomycetes associated with woody oil plants in China. Mycosphere; 2023; 14, pp. 1436-1529. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/16]
Li, QR; Habib, K; Long, SH; Wu, YP; Zhang, X; Hu, HM; Wu, QZ; Liu, LL; Zhou, SX; Shen, XC; Kang, JC. Unveiling fungal diversity in China: New species and records within the Xylariaceae Family. Mycosphere; 2024; 15, pp. 275-364.1:CAS:528:DC%2BB2cXpslyrtLo%3D [DOI: https://dx.doi.org/10.5943/mycosphere/15/1/2]
Liang, ZY; Shen, NX; Zhou, XJ; Zheng, YY; Chen, M; Wang, CY. Bioactive indole diterpenoids and polyketides from the marine-derived fungus Penicillium javanicum. Chem Nat Compd; 2020; 56, pp. 379-382.1:CAS:528:DC%2BB3cXmtFymtL4%3D [DOI: https://dx.doi.org/10.1007/s10600-020-03039-6]
Liao CF, Dong W, Chethana KWT, Pem D, Phookamsak R, Suwannarach N, lom M (2022) Morphological and phylogenetic evidence reveal Tetraploa cylindrica sp. nov. (Tetraplosphaeriaceae, Pleosporales) from Saccharum arundinaceum (Poaceae) in Yunnan Province, China. Phytotaxa 554:189–200. https://doi.org/10.11646/PHYTOTAXA.554.2.7
Lin CG, Mckenzie EHC, Bhat DJ, Liu JK, Hyde KD, Lumyong S (2018). Pseudodactylaria brevis sp. nov. from Thailand confirms the status of Pseudodactylariaceae. Phytotaxa 369:241–250. https://doi.org/10.11646/phytotaxa.369.4.1
Link, JHF. Observationes in ordines plantarum naturales. Dissertatio Ima. Gesellschaft Naturforschender Freunde Zu Berlin Magazin; 1809; 3, pp. 3-42.
Liou, GY; Tzean, SS. Phylogeny of the genus Arthrobotrys and allied nematode-trapping fungi based on rDNA sequences. Mycologia; 1997; 89,
Lisboa, DO; Silva, MA; Pinho, DB; Pereira, OL; Furtado, GQ. Diversity of pathogenic and endophytic Colletotrichum isolates from Licania tomentosa in Brazil. Forest Pathol; 2018; 48, [DOI: https://dx.doi.org/10.1111/efp.12448] e12448.
Lister, A. A Monograph of the Mycetozoa; 1894; London, British Museum Nat. Hist: 224.
Lister A (1911) A Monograph of the Mycetozoa (ed. 2. rev. by G. Lister). British Museum Nat. Hist. London, p 302
Lister A (1925) A Monograph of the Mycetozoa (ed. 3. rev. by G. Lister). British Museum Nat. Hist. London, p 296
Liu B, Liu XZ, Zhuang WY (2006) Taxonomy and molecular phylogeny of Orbiliaceae from China. Poster presented at the 8th International Mycological Congress in August 2006 in Cairns, Australia
Liu, JK; Phookamsak, R; Dai, DQ; Tanaka, K; Jones, EG; Xu, JC; Chukeatirote, E; Hyde, KD. Roussoellaceae, a new pleosporalean family to accommodate the genera Neoroussoella gen. nov., Roussoella and Roussoellopsis. Phytotaxa; 2014; 181, pp. 1-33. [DOI: https://dx.doi.org/10.11646/phytotaxa.181.1.1]
Liu, F; Weir, BS; Damm, U; Crous, PW; Wang, Y; Liu, B; Wang, M; Zhang, M; Cai, L. Unravelling Colletotrichum species associated with Camellia: employing ApMat and GS loci to resolve species in the C. gloeosporioides complex. Persoonia; 2015; 35, pp. 63-86. [DOI: https://dx.doi.org/10.3767/003158515X687597] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26823629][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4713112]
Liu, JK; Hyde, KD; Jones, EBG; Ariyawansa, HA; Bhat, DJ; Boonmee, S; Maharachchikumbura, SS; McKenzie, EH; Phookamsak, R; Phukhamsakda, C; Shenoy, BD; Abdel-Wahab, MA; Buyck, B; Chen, J; Chethana, KWT; Singtripop, C; Dai, DQ; Dai, YC; Daranagama, DA; Dissanayake, AJ; Doilom, M; D’souza, MJ; Fan, XL; Goonasekara, ID; Hirayama, K; Hongsanan, S; Jayasiri, SV; Jayawardena, RS; Karunarathna, SC; Li, WJ; Mapook, A; Norphanphoun, C; PangKL, PRH; Peršoh, D; Pinruan, U; Senanayake, IC; Somrithipol, S; Suetrong, S; Tanaka, K; Thambugala, KM; Tian, Q; Tibpromma, S; Udayanga, D; Wijayawardene, NN; Wanasinghe, D; Wisitrassameewong, K; Zeng, XY; Abdel-Aziz, FA; Adamčík, S; Bahkali, AH; Boonyuen, N; Bulgakov, T; Callac, P; Chomnunti, P; Greiner, K; Hashimoto, A; Hofstetter, V; Kang, JC; Lewis, D; Li, XH; Liu, ZH; Liu, ZY; Matsumura, M; Mortimer, PE; Rambold, G; Randrianjohany, E; Sato, G; Sri-Indrasutdhi, V; Tian, CM; Verbeken, A; Brackel, W; Wang, Y; Wen, TC; Xu, JC; Yan, JY; Zhao, RL; Camporesi, E. Fungal Diversity Notes 1–110: taxonomic and phylogenetic contributions to fungal species. Fungal Divers; 2015; 72, pp. 1-197.1:CAS:528:DC%2BC2cXitFOrtb%2FO [DOI: https://dx.doi.org/10.1007/s13225-015-0324-y]
Liu, QL; Li, JQ; Wingfield, MJ; Duong, BD; Wingfield, BD; Crous, PW; Chen, SF. Reconsideration of species boundaries and proposed DNA barcodes for Calonectria. Stud Mycol; 2020; 97, pp. 112-156. [DOI: https://dx.doi.org/10.1016/j.simyco.2020.08.001]
Liu, S; Shen, LL; Xu, TM; Song, CG; Gao, N; Wu, DM; Sun, YF; Cui, BK. Global diversity, molecular phylogeny and divergence times of the brown-rot fungi within the Polyporales. Mycosphere; 2023; 14, pp. 1564-1664. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/18]
Liu, XF; Tibpromma, S; Hughes, AC; Chethana, KWT; Wijayawardene, NN; Dai, DQ; Du, TY; Elgorban, AM; Stephenson, SL; Suwannarach, N; Xu, JC; Lu, L; Xu, RF; Maharachchikumbura, SSN; Zhao, CL; Bhat, DJ; Sun, YM; Karunarathna, SC; Mortimer, PE. Culturable mycota on bats in central and southern Yunnan Province, China. Mycosphere; 2023; 14, pp. 497-662. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/7]
Liu SL, Wang XW, Li GJ, Deng CY, Rossi W, Leonardi M, Liimatainen K, Kekki T, Niskanen T, Smith ME, Ammirati J (2024) Taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Diversity. 124(1):1–216
Lodge, DJ; Chapela, I; Samuels, G; Uecker, FA; Desjardin, D; Horak, E; Miller, OK, Jr; Hennebert, GL; Decock, CA; Ammirati, J; Burdsall, HH, Jr; Kirk, PM; Minter, DW; Hailing, R; Laessøe, T; Mueller, G; HuhndorfBerwinkler, SOF; Pegler, DN; Spooner, B; Petersen, RH; Rogers, JD; Ryvarden, L; Watling, R; Turnbull, E; Whalley, AJS. A survey of patterns of diversity in non-lichenized fungi. Mitt Eidgenöss Forsch Anst Wald Schnee Landsch; 1995; 70, pp. 157-173.
Lodge, DJ; Padamsee, M; Matheny, PB; Aime, CM; Cantrell, SA; Boertmann, D; Kovalenko, A; Vizzini, A; Dentinger, BTM; Kirk, PM; Ainsworth, AM; Moncalvo, JM; Vilgalys, R; Larsson, E; Lücking, R; Griffith, WG; Smith, ME; Norvell, LL; Desjardin, DE; Redhead, SA; Ovrebo, CL; Lickey, EB; Ercole, E; Hughes, KW; Courtecuisse, R; Young, A; Binder, M; Minnis, AM; Lindner, DL; Ortiz-Santana, B; Haight, J; Læssøe, T; Baroni, TJ; Geml, J; Hattori, T. Molecular phylogeny, morphology, pigment chemistry and ecology in Hygrophoraceae. Fungal Divers; 2014; 64, pp. 1-99. [DOI: https://dx.doi.org/10.1007/s13225-013-0259-0]
Lu YZ, Zhang JY, Lin CG, Luo ZL (2020) Pseudodactylaria fusiformis sp. nov. from freshwater habitat in China. Phytotaxa. 446:95–102. https://doi.org/10.11646/PHYTOTAXA.446.2.2
Luangharn T, Karunarathna SC, Mortimer PE, Hyde KD, Xu JC (2020) Morphology, phylogeny and culture characteristics of Ganoderma gibbosum collected from Kunming, Yunnan Province, China. Phyton: 89:1–22. https://doi.org/10.32604/phyton.2020.09690
Luangharn, T; Karunarathna, SC; Dutta, AK; Paloi, S; Promputtha, I; Hyde, KD; Xu, JC; Mortimer, PE. Ganoderma (Ganodermataceae, Basidiomycota) species from the Greater Mekong Subregion. J Fungi; 2021; 7, 819.1:CAS:528:DC%2BB3MXivVWltr%2FN [DOI: https://dx.doi.org/10.3390/jof7100819]
Łubek, A; Kukwa, M. Additions to the mycobiota of Poland. Mycotaxon; 2017; 132, pp. 183-195. [DOI: https://dx.doi.org/10.5248/132.183]
Lucena, L; Fernández-Valencia, RC. Zygosporium verruciferum sp. nov., on Solanum lycopersicum from Venezuela. Mycotaxon; 2017; 132, pp. 585-589. [DOI: https://dx.doi.org/10.5248/132.585]
Lücking, R. The taxonomy of the genus Graphis sensu Staiger (Ascomycota: Ostropales: Graphidaceae). Lichenologist; 2009; 41, pp. 319-362. [DOI: https://dx.doi.org/10.1017/S0024282909008524]
Lücking, R; Archer, AW; Aptroot, A. A worldwide key to the genus Graphis (Ostropales: Graphidaceae). Lichenologist; 2009; 41, pp. 363-452. [DOI: https://dx.doi.org/10.1017/S0024282909008305]
Lücking, R; Tehler, A; Bungartz, F; Rivas Plata, E; Lumbsch, HT. Journey from the West: did tropical Graphidaceae (lichenized Ascomycota: Ostropales) evolve from a saxicolous ancestor along the American Pacific coast?. Am J Bot; 2013; 100, pp. 844-856. [DOI: https://dx.doi.org/10.3732/ajb.1200548] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/23594913]
Lücking, R; Aime, MC; Robbertse, B; Miller, AN; Aoki, T; Ariyawansa, HA; Cardinali, G; Crous, PW; Druzhinina, IS; Geiser, DM; Hawksworth, DL; Hyde, KD; Laszlo Irinyi, L; Jeewon, R; Johnston, PR; Kirk, PM; Elaine Malosso, E; May, TW; Meyer, W; Nilsson, HR; Öpik, M; Robert, V; Stadler, M; Thines, M; Vu, D; Yurkov, AM; Zhang, N; Schoch, CL. Fungal taxonomy and sequence-based nomenclature. Nat Microbiol; 2021; 6, pp. 540-548.1:CAS:528:DC%2BB3MXps12nsrg%3D [DOI: https://dx.doi.org/10.1038/s41564-021-00888-x] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33903746][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10116568]
Lunardelli Negreiros de Carvalho P, de Oliveira Silva E, Aparecida Chagas-Paula D, Honorata Hortolan Luiz J, Ikegaki M (2016) Importance and implications of the production of phenolic secondary metabolites by endophytic fungi: a mini-review. Mini Rev Med Chem 1;16(4):259–271
Luo, ZL; Hyde, KD; Liu, LK; Maharachchikumbura, SSN; Jeewon, R; Bao, FD; Bhat, DJ; Lin, GC; Li, WL; Yang, J; Liu, NG; Lu, YZ; Jayawardena, RS; Li, JF; Su, YH. Freshwater Sordariomycetes. Fungal Divers; 2019; 99, pp. 451-660. [DOI: https://dx.doi.org/10.1007/s13225-019-00438-1]
Luttrell, ES. Taxonomy of Pyrenomycetes. Univ Missouri Stud; 1951; 24, pp. 1-120.
Maamar, A; Lucchesi, ME; Debaets, S; Nguyen van Long, N; Quemener, M; Coton, E; Bouderbala, M; Burgaud, G; Matallah-Boutiba, A. Highlighting the crude oil bioremediation potential of marine fungi isolated from the Port of Oran (Algeria). Diversity; 2020; 12,
Macbride TH (1899) The North American Slime-Moulds. New York The Macmillan Company, p 269
Macbride TH (1922) The North American Slime-Moulds (ed. 2) New York The Macmillan Company, p 347
Maharachchikumbura, SSN; Hyde, KD; Jones, EBG; McKenzie, EHC; Huang, SK; 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. Towards a natural classification and backbone tree for Sordariomycetes. Fungal Divers; 2015; 72, pp. 199-230. [DOI: https://dx.doi.org/10.1007/s13225-015-0331-z]
Maharachchikumbura, SS; Wanasinghe, DN; Cheewangkoon, R; Al-Sadi, AM. Uncovering the hidden taxonomic diversity of fungi in Oman. Fungal Divers; 2021; 106, pp. 229-268.1:CAS:528:DC%2BB38XitlCisbfF [DOI: https://dx.doi.org/10.1007/s13225-020-00467-1]
Manawasinghe, IS; Dissanayake, AJ; Li, X; Liu, M; Wanasinghe, DN; Xu, J; Zhao, W; Zhang, W; Zhou, Y; Hyde, KD; Brooks, S. High genetic diversity and species complexity of Diaporthe associated with grapevine dieback in China. Front Microbiol; 2019; 10, 1936. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31543868][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6732904][DOI: https://dx.doi.org/10.3389/fmicb.2019.01936]
Manawasinghe, IS; Phillips, AJ; Xu, J; Balasuriya, A; Hyde, KD; Stępień, Ł; Harischandra, DL; Karunarathna, A; Yan, J; Weerasinghe, J; Luo, M. Defining a species in fungal plant pathology: beyond the species level. Fungal Divers; 2021; 109, pp. 267-282. [DOI: https://dx.doi.org/10.1007/s13225-021-00481-x]
Manawasinghe, IS; Calabon, MS; Jones, EBG; Zhang, YX; Liao, CF; Xiong, Y; Chaiwan, N; Kularathnage, ND; Liu, NG; Tang, SM; Sysouphanthong, P; Du, TY; Luo, M; Pasouvang, P; Pem, D; Phonemany, M; Ishaq, M; Chen, JW; Karunarathna, SC; Mai, ZL; Rathnayaka, AR; Samarakoon, MC; Tennakoon, DS; Wijesinghe, SN; Yang, YH; Zhao, HJ; Fiaz, M; Doilom, M; Dutta, AK; Khalid, AN; Liu, JW; Thongklang, N; Senanayake, IC; Tibpromma, S; You, LQ; Camporesi, E; Gafforov, YS; Hyde, KD. Mycosphere notes 345–386. Mycosphere; 2022; 13, pp. 454-557. [DOI: https://dx.doi.org/10.5943/mycosphere/13/1/3]
Mangold, A; Martin, MP; Lucking, R; Lumbsch, HT. Molecular phylogeny suggests synonymy of Thelotremataceae within Graphidaceae (Ascomycota: Ostropales). Taxon; 2008; 57, pp. 476-486. [DOI: https://dx.doi.org/10.2307/25066016]
Manoharachary, C; Agarwal, DK; Sureshkumar, G; Kunwar, IK; Babu, KS. Memnoniella mohanramii sp. nov. and Zygosporium anupamvarmae sp. nov. from India. Indian Phytopathol; 2006; 59, pp. 489-491.
Mao, N; Zhao, TY; Zhang, YX; Li, T; Lv, JC; Fan, L. Boletaceae from Shanxi Province of northern China with descriptions of ten new species. Mycosphere; 2023; 14, pp. 2013-2091. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/24]
Mapook A, Hyde KD, Dai DQ, Li J, Jones EG, Bahkali AH, Boonmee S (2016) Muyocopronales, ord. nov., (Dothideomycetes, Ascomycota) and a reappraisal of Muyocopron species from northern Thailand. Phytotaxa 265:225–237. https://doi.org/10.11646/PHYTOTAXA.265.3.3
Mapook, A; Hyde, KD; McKenzie, EH; Jones, EG; Bhat, DJ; Jeewon, R; Stadler, M; Samarakoon, MC; Malaithong, M; Tanunchai, B; Buscot, F. Taxonomic and phylogenetic contributions to fungi associated with the invasive weed Chromolaena odorata (Siam weed). Fungal Divers; 2020; 101, pp. 1-175. [DOI: https://dx.doi.org/10.1007/s13225-020-00444-8]
Mapook A, Macabeo AP, Thongbai B, Hyde KD, Stadler M (2020b) Polyketide-derived secondary metabolites from a Dothideomycetes fungus, Pseudopalawania siamensis gen. et sp. nov., (Muyocopronales) with antimicrobial and cytotoxic activities. Biomolecules 10:569. https://doi.org/10.3390/biom10040569
Martin, GW; Alexopoulos, CJ. The Myxomycetes; 1969; Iowa, University of Iowa Press: 560.
Matheny, PB; Curtis, JM; Hofstetter, V; Aime, MC; Moncalvo, JM; Ge, ZW; Slot, JC; Ammirati, JF; Baroni, TJ; Bougher, NL; Hughes, KW; Lodge, DJ; Kerrigan, RW; Seidl, MT; Aanen, DK; DeNitis, M; Daniele, GM; Desjardin, DE; Kropp, BR; Norvell, LL; Parker, A; Vellinga, EC; Vilgalys, R; Hibbett, DS. Major clades of Agaricales: a multilocus phylogenetic overview. Mycologia; 2006; 98, pp. 982-995. [DOI: https://dx.doi.org/10.1080/15572536.2006.11832627] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/17486974]
McKenzie, EHC; Thongkantha, S; Lumyong, S. Zygosporium bioblitzi sp. nov. on dead leaves of Cortaderia and Dracaena. N Z J Bot; 2007; 45, pp. 433-435. [DOI: https://dx.doi.org/10.1080/00288250709509724]
McKeown, TA; Moss, ST; Jones, EBG. Ultrastructure of ascospores of Tunicatispora australiensis. Mycol Res; 1996; 100, pp. 1247-1255. [DOI: https://dx.doi.org/10.1016/S0953-7562(96)80188-2]
McNeill J, Barrie FF, Buck WR et al (eds) (2012) International Code of Nomenclature for algae, fungi, and plants (Melbourne Code). Koeltz Scientific Books, Königstein [Regnum vegetabile no. 154
Meyer, W; Gams, W. Delimitation of Umbelopsis (Mucorales, Umbelopsidaceae fam. nov.) based on ITS sequence and RFLP data. Mycol Res; 2003; 107, pp. 339-350.1:CAS:528:DC%2BD3sXjs1Kmsr4%3D [DOI: https://dx.doi.org/10.1017/S0953756203007226] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/12825503]
Milanović, V; Osimani, A; Cardinali, F; Taccari, M; Garofalo, C; Clementi, F; Ashoor, S; Mozzon, M; Foligni, R; Canonico, L; Ciani, M. Effect of inoculated azotobacteria and Phanerochaete chrysosporium on the composting of olive pomace: Microbial community dynamics and phenols evolution. Sci Rep; 2019; 9,
Miller MA, Pfeiffer W, Schwartz T (2010) Creating the CIPRES Science Gateway for inference of large phylogenetic trees. Proceedings of the Gateway Computing Environments Workshop (GCE), 14 Nov. 2010, New Orleans, LA, USA, 1–8. https://doi.org/10.1109/GCE.2010.5676129
Miller, MA; Schwartz, T; Pickett, BE; He, S; Klem, EB; Scheuermann, RH; Passarotti, M; Kaufman, S; O'Leary, MA. A RESTful API for Access to Phylogenetic Tools via the CIPRES Science Gateway. Evol Bioinforma; 2015; 11, pp. 43-48.1:CAS:528:DC%2BC1cXlsVKmtrg%3D [DOI: https://dx.doi.org/10.4137/EBO.S21501]
Montagne, JPFC. Troisième Centurie de plantes cellulaires exotiques nouvelles, Décades I, II, III et IV. Fungi cubenses. Annales Des Sciences Naturelles Botanique; 1842; 17, pp. 119-128.
Montecchi, A; Ruini, S; Gross, G. Gigaspora Lazzarii nov. sp. Riv Micol; 1996; 39,
Monteiro, JS; Gusmão, LFP; Castañeda-Ruiz, RF. Pleurothecium bicoloratum & Sporidesmiopsis pluriseptata spp. nov. from Brazil. Mycotaxon; 2016; 131, pp. 145-152. [DOI: https://dx.doi.org/10.5248/131.145]
Morera G, Robledo GL, Ferreira-Lopes V, Urcelay RC (2017) South American Fomitiporia (Hymenochaetaceae, Basidiomycota) 'jump on' exotic living trees revealed by multi-gene phylogenetic analysis. Phytotaxa 321:277–286. https://doi.org/10.11646/phytotaxa.321.3.5
Moreno G, Sánchez A, Singer H, Illana C, Castillo A (2002) A study on nivicolous Myxomycetes. The genus Lamproderma I. Fungi non Deliniati 19. 65 pp. Edizioni Candusso. I- Alassio (SV)
Mortimer, PE; Jeewon, R; Xu, JC; Lumyong, S; Wanasinghe, DN. Morpho-phylo taxonomy of novel dothideomycetous fungi associated with dead woody twigs in Yunnan Province, China. Front Microbiol; 2021; 12, [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33833748][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8021917][DOI: https://dx.doi.org/10.3389/fmicb.2021.654683] 654683.
Morgan, AP. New North American fungi. J Cincinnati Soc Nat Hist; 1895; 18, pp. 36-45.
Morton, JB; Benny, GL. Revised classification of arbuscular mycorrhizal fungi (Zygomycetes): a new order, Glomales, two new suborders, Glomineae and Gigasporineae, and two new families, Acaulosporaceae and Gigasporaceae, with an emendation of Glomaceae. Mycotaxon; 1990; 37, pp. 471-491.
Munk, A. The system of the pyrenomycetes. A contribution to a natural classification of the group Sphaeriales sensu Lindau. Dan Bot Ark; 1953; 15, pp. 1-163.
Murata, K; Ogawa, Y; Kusama, K; Yasuhara, Y. Disseminated Verruconis gallopava infection in a patient with systemic lupus erythematosus in Japan: a case report, literature review, and autopsy case. Med Mycol Case Rep; 2022; 35, pp. 35-38. [DOI: https://dx.doi.org/10.1016/j.mmcr.2022.01.004] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/35096522][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8783064]
Murrill, WA. Pleurotus, Omphalia,Mycena and. Mycologia; 1916; 8, 218. [DOI: https://dx.doi.org/10.1080/00275514.1916.12018883]
MycoBank (2022) https://www.mycobank.org/. Accessed 1 Jan 2023
Nannenga-Bremekamp, NE. A guide to temperate myxomycetes; 1991; Bristol, Biopress Limited: 409.
Nannfeldt, JZ. Studien über die Morphologie und Systematik der nicht-lichenisierten inoperculaten Discomyceten. Nova Acta Regiae Soc Sci Ups; 1932; 48, 368.
Naranjo-Ortiz, MA; Gabaldón, T. Fungal evolution: major ecological adaptations and evolutionary transitions. Biol Rev Camb Philos Soc; 2019; 94, pp. 1443-1476. [DOI: https://dx.doi.org/10.1111/brv.12510] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31021528][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6850671]
Neeraj, MKG; Sharma, BC; Varma, AK. A new species of Gigaspora from desert soils of Rajasthan, India. World J Microbiol Biotechnol; 1993; 9,
Neubert, H; Nowotny, W; Baumann, K; Marx, H. Die Myxomyceten; 2000; Stemonitales, Karlheinz Baumann Verlag Gomaringen: 391.
Nicolson, TH; Gerdemann, JW. Mycorrhizal Endogone species. Mycologia; 1968; 60, pp. 313-325. [DOI: https://dx.doi.org/10.1080/00275514.1968.12018572]
Niknejad, F; Moshfegh, M; Najafzadeh, MJ; Houbraken, J; Rezaei, S; Zarrini, G; Faramarzi, MA; Nafissi-Varcheh, N. Halotolerant ability and α-amylase activity of some saltwater fungal isolates. Iranian J Pharm Res; 2013; 12,
Nimis PL (2016) ITALIC—The information system on Italian lichens. Version 5.0. University of Trieste, Department of Biology, Trieste. https://dryades.units.it/italic. Accessed 01 April 2023
Niskanen, T; Lücking, R; Dahlberg, A; Gaya, E; Suz, LM; Mikryukov, V; Liimatainen, K; Druzhinina, I; Westrip, JR; Mueller, GM; Martins-Cunha, K. Pushing the frontiers of biodiversity research: unveiling the global diversity, distribution, and conservation of fungi. Annu Rev Environ Resour; 2023; 48, pp. 149-176. [DOI: https://dx.doi.org/10.1146/annurev-environ-112621-090937]
Noireung, P; Phoulivong, S; Liu, F; Cai, L; Mckenzie, EH; Chukeatirote, E; Jones, EB; Bahkali, AH; Hyde, KD. Novel species of Colletotrichum revealed by morphology and molecular analysis. Cryptogam Mycol; 2012; 33, pp. 347-362. [DOI: https://dx.doi.org/10.7872/crym.v33.iss3.2012.347]
Nordin, A; Owe-Larsson, B; Tibell, L. Two new Aspicilia species from Fennoscandia and Russia. Lichenologist; 2011; 43, pp. 27-37. [DOI: https://dx.doi.org/10.1017/S0024282910000629]
Norphanphoun, C; Jeewon, R; Mckenzie, EHC; Wen, TC; Camporesi, E; Hyde, KD. Taxonomic Position of Melomastia italica sp. nov. and Phylogenetic Reappraisal of Dyfrolomycetales. Cryptogam Mycol; 2017; 38, pp. 507-525. [DOI: https://dx.doi.org/10.7872/crym/v38.iss4.2017.507]
Norphanphoun, C; Gentekaki, E; Hongsanan, S; Jayawardena, R; Senanayake, IC; Manawasinghe, IS; Abeywickrama, PD; Bhunjun, CS; Hyde, KD. Diaporthe: formalizing the species-group concept. Mycosphere; 2022; 13,
Norris, G. Systematic and stratigraphic palynology of Eocene to Pliocene strata in the Imperial Nuktak C–22 well, Mackenzie Delta Region, District of Mackenzie, N.W.T. Geol Surv Can Bull; 1986; 340, pp. 1-89.
Nourel-Din AAH, Abdel-Aziz FA, Abdel-Wahab MA (2022) Qarounispora grandiappendiculata gen. et sp. nov. (Halosphaeriaceae, Microascales) from Qaroun Lake, Egypt. Phytotaxa 530:086–094. https://doi.org/10.11646/phytotaxa.530.1.7
Núñez, M; Ryvarden, L. East Asian polypores 1 Ganodermataceae and Hymenochaetaceae. Synop Fungorum; 2000; 13, pp. 1-168.
Núñez F, Díaz MC, Rodríguez M, Aranda E, Martín A, Asensio MA (2000) Effects of substrate, water activity, and temperature on growth and verrucosidin production by Penicillium polonicum isolated from dry-cured ham. Journal of food protection; 63(2):231-236
Oehl, F; de Souza, FA; Sieverding, E. Revision of Scutellospora and description on five new genera and three new families in the arbuscular mycorrhizal-forming Glomeromycetes. Mycotaxon; 2008; 106, pp. 311-360.
Oehl, F; Sieverding, E; Palenzuela, J; Ineichen, K; da Silva, GA. Advances in Glomeromycota taxonomy and classification. IMA Fungus; 2011; 2, pp. 191-199. [DOI: https://dx.doi.org/10.5598/imafungus.2011.02.02.10] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22679604][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3359817]
Oehl, F; Silva, GA; Goto, BT; Maia, LC; Sieverding, E. Glomeromycota: two new classes and a new order. Mycotaxon; 2011; 116, pp. 365-379. [DOI: https://dx.doi.org/10.5248/116.365]
Oliveira, JJ; Vargas-Isla, R; Cabral, TS; Rodrigues, DP; Ishikawa, NK. Progress on the phylogeny of the Omphalotaceae: Gymnopus s. str., Marasmiellus s. str., Paragymnopus gen. nov. and Pusillomyces gen. nov. Mycol Prog; 2019; 18, pp. 713-739. [DOI: https://dx.doi.org/10.1007/s11557-019-01483-5]
Oliveira, JJS; Moncalvo, JM; Margaritescu, S; Capelari, M. A morphological and phylogenetic evaluation of Marasmius sect. Globulares (Globulares-Sicci complex) with nine new taxa from the Neotropical Atlantic Forest. Persoonia-Mol Phylogeny Evol Fungi; 2020; 44, pp. 240-277.1:STN:280:DC%2BB3s7nsVWqtQ%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2020.44.09]
Pang, KL; Vrijmoed, LLP; Kong, RYC; Jones, EBG. Polyphyly of Halosarpheia (Halosphaeriales, Ascomycota): implications on the use of unfurling ascospore appendage as a systematic character. Nova Hedw; 2003; 77, pp. 1-18. [DOI: https://dx.doi.org/10.1127/0029-5035/2003/0077-0001]
Pang, KL; Vrijmoed, LLP; Kong, RYC; Jones, EBG. Lignincola and Nais, polyphyletic genera of the Halosphaeriales (Ascomycota). Mycol Prog; 2003; 2, pp. 29-36. [DOI: https://dx.doi.org/10.1007/s11557-006-0041-8]
Pang KL, Chiang WL, Jheng JS (2012) Morphological evaluation of peridial wall, ascus and ascospore characteristics in the delineation of genera with unfurling ascospore appendages (Halosphaeriaceae). In: Raghukumar C (ed) Biology of marine fungi. Progress in molecular and subcellular biology. Springer, Berlin, pp 159–171. https://doi.org/10.1007/978-3-642-23342-5_8
Park MS, Lee S, Oh SY, Lim YW (2016) Penicillium Diversity from Intertidal Zone in Korea. 한국균학회 학술대회;11:11
Parmasto, E. On the origin of the Hymenomycetes (What are corticioid fungi?). Windahlia; 1986; 16, pp. 3-19.
Parsons, MG; Norris, G. Paleogene fungi from the Caribou Hills, Mackenzie Delta, northern Canada. Palaeontographica Abt B; 1999; 250, pp. 77-167. [DOI: https://dx.doi.org/10.1127/palb/250/1999/77]
Pegler DN (1986) Agaric flora of Sri Lanka. Kew Bull Adit Ser 12. HMSO, London
Pem, D; Jeewon, R; Chethana, KW; Hongsanan, S; Doilom, M; Suwannarach, N; Hyde, KD. Species concepts of Dothideomycetes: classification, phylogenetic inconsistencies and taxonomic standardization. Fungal Divers; 2021; 109, pp. 283-319. [DOI: https://dx.doi.org/10.1007/s13225-021-00485-7]
Peña AB, Luecking R, Miranda-Gonzalez R, de los Angeles Herrera-Campos M (2014) Three new species of Graphis (Ascomycota: Ostropales: Graphidaceae) from Mexico, with updates to taxonomic key entries for 41 species described between 2009 and 2013. The Lichenologist 46(1):69-82. https://doi.org/10.1017/S0024282913000637
Peng, XJ; Wang, QC; Zhang, SK; Guo, K; Zhou, XD. Colletotrichum species associated with Camellia anthracnose in China. Mycosphere; 2023; 14, pp. 130-157. [DOI: https://dx.doi.org/10.5943/mycosphere/14/si2/3]
Penzig, O; Saccardo, PA. Diagnoses fungorum novorum in Insula Java collectorum. Series Secunda. Malpighia; 1897; 11, pp. 491-530.
Pereira DS, Hilário S, Gonçalves MFM, Phillips AJL (2023) Diaporthe species on palms: molecular re-assessment and species boundaries delimitation in the D. arecae species complex. Microorganisms 11:2717. https://doi.org/10.3390/microorganisms11112717
Perera, RH; Hyde, KD; Maharachchikumbura, SSN; Jones, EBG; McKenzie, EHC; Stadler, M; Lee, HB; Samarakoon, MC; Ekanayaka, AH; Camporesi, E; Liu, JK. Fungi on wild seeds and fruits. Mycosphere; 2020; 11, pp. 2108-2480. [DOI: https://dx.doi.org/10.5943/MYCOSPHERE/11/1/14]
Petch, T. Notes on entomogenous fungi. Trans Br Mycol Soc; 1931; 16, pp. 55-75. [DOI: https://dx.doi.org/10.1016/S0007-1536(31)80006-3]
Petch, T. Notes on entomogenous fungi. Trans Br Mycol Soc; 1937; 21, pp. 34-67. [DOI: https://dx.doi.org/10.1016/S0007-1536(37)80005-4]
Petersen, RH; Hughes, KW. Collybiopsis and its type species, Co. ramealis. Mycotaxon; 2021; 136, pp. 263-349. [DOI: https://dx.doi.org/10.5248/136.263]
Pfister, DH. Castor, Pollux and life histories of fungi. Mycologia; 1997; 89,
Phillips, AJL; Alves, A; Pennycook, SR; Johnston, PR; Ramaley, A; Akulov, A; Crous, PW. Resolving the phylogenetic and taxonomic status of dark-spored teleomorph genera in the Botryosphaeriaceae. Persoonia; 2008; 21, pp. 29-55.1:STN:280:DC%2BC3c3ms12ktw%3D%3D [DOI: https://dx.doi.org/10.3767/003158508X340742] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/20396576][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2846129]
Phillips, AJL; Alves, A; Abdollahzadeh, J; Slippers, B; Wingfield, MJ; Groenewald, JZ; Crous, PW. The Botryosphaeriaceae: genera and species known from culture. Stud Mycol; 2013; 76, pp. 51-167.1:STN:280:DC%2BC2c3ntlegtg%3D%3D [DOI: https://dx.doi.org/10.3114/sim0021] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/24302790][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3825232]
Phookamsak R, Hyde KD, Jeewon R, Bhat DJ, Jones EBG, Maharachchikumbura SSN, Raspe O, Karunarathna SC, Wanasinghe DN, Hongsanan S, lom M, Tennakoon DS, Machado AR, Firmino AL,Ghosh A, Karunarathna A, Mesic 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, Kusˇan Cano J, Gene J, Li J, Das K, Acharya K, Raj KNA, Latha KPD, Chethana KWT, He MQ, Duenas M, Jadan M, Martı´n MP, Samarakoon MC, Dayarathne MC, Raza M, Park MS, Telleria MT, Chaiwan N, Matocec 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, Tkalcˇec 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 929–1035: taxonomic and phylogenetic contributions on genera and species of fungi. Fungal Divers 95: 1–273.https://doi.org/10.1007/s13225-019-00421-w
Phookamsak, R; Jiang, H; Suwannarach, N; Lumyong, S; Xu, J; Xu, S; Liao, CF; Chomnunti, P. Bambusicolous fungi in Pleosporales: Introducing four novel taxa and a new habitat record for Anastomitrabeculia didymospora. J Fungi; 2022; 8,
Pinnoi, A; Lumyong, S; Hyde, KD; Jones, EG. Biodiversity of fungi on the palm Eleiodoxa conferta in Sirindhorn peat swamp forest, Narathiwat, Thailand. Fungal Divers; 2006; 22, pp. 205-218.
Pitt, JI. The genus Penicillium and its teleomorphic states Eupenicillium and Talaromyces; 1979; New York, Academic Press Inc.:
Pitt JI (1980) The genus Penicillium and its teleomorphic states Eupenicillium and Talaromyces. Academic, London.:1634
Pitt JI, Samson RA (1993) Names in current use in the family Trichocomaceae. In: Names in current use in the family Trichocomaceae, Cladoniaceae, Pinaceae, and Lemnaceae (Greuter W, ed). Koeltz Scientific Books, Konigstein, Germany, pp 13–57
Plata, ER; Lücking, R. High diversity of Graphidaceae (lichenized Ascomycota: Ostropales) in Amazonian Perú. Fungal Divers; 2012; 58, pp. 13-32. [DOI: https://dx.doi.org/10.1007/s13225-012-0172-y]
Plata, ER; Lumbsch, HT. Parallel evolution and phenotypic divergence in lichenized fungi: a case study in the lichen-forming fungal family Graphidaceae (Ascomycota: Lecanoromycetes: Ostropales). Mol Phylogenet Evol; 2011; 61,
Poulain M, Meyer M, Bozonnet J (2011) Les Myxomycètes tomes 1–2. Fédération mycologique et botanique Dauphiné-Savoie, Sevrier, p 568, 544 plates
Prencipe, S; Siciliano, I; Gatti, C; Garibaldi, A; Gullino, ML; Botta, R; Spadaro, D. Several species of Penicillium isolated from chestnut flour processing are pathogenic on fresh chestnuts and produce mycotoxins. Food Microbiol; 2018; 76, pp. 396-404.1:CAS:528:DC%2BC1cXhtlSjt77J [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30166166][DOI: https://dx.doi.org/10.1016/j.fm.2018.07.003]
Punithalingam, E. Sphaeropsidales in culture from humans. Nova Hedwigia; 1979; 31, pp. 119-158.
Punithalingam, E. Nuclei, micronuclei and appendages in tri- and tetraradiate conidia of Cornutispora and four other coelomycete genera. Mycol Res; 2003; 107, pp. 917-948. [DOI: https://dx.doi.org/10.1017/S0953756203008037] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/14531616]
Qu, J; Zou, X; Cao, W; Xu, Z; Liang, Z. Two new species of Hirsutella (Ophiocordycipitaceae, Sordariomycetes) that are parasitic on lepidopteran insects from China. MycoKeys; 2021; 82, pp. 81-96. [DOI: https://dx.doi.org/10.3897/mycokeys.82.66927] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34408539][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8367965]
Quan, Y; Muggia, L; Moreno, LF; Wang, M; Al-Hatmi, AM; da Silva, MN; Shi, D; Deng, S; Ahmed, S; Hyde, KD; Vicente, VA. A re-evaluation of the Chaetothyriales using criteria of comparative biology. Fungal Divers; 2020; 103, pp. 47-85. [DOI: https://dx.doi.org/10.1007/s13225-020-00452-8]
Quél (1872) In: Mém. Soc. Émul. Montbéliard, Sér. 2 5:221
Rajchenberg, M; Pildain, MB; Bianchinotti, MV; Barroetaveña, C. The phylogenetic position of poroid Hymenochaetaceae (Hymenochaetales, Basidiomycota) from Patagonia, Argentina. Mycologia; 2015; 107, pp. 754-767.1:CAS:528:DC%2BC28XnsFeksbk%3D [DOI: https://dx.doi.org/10.3852/14-170] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25911700]
Ramanujam CGK, Rao KP (1978) Fungal spores from the Neogene strata of Kerala in South India; 291–304 in Bharadwaj DC et al (eds) Proceedings of the 4th International Palynological Conference, Lucknow. 1976–1977
Ramos SM, Cruz R, Barbosa RD, Machado AR, Costa AD, Motta CD, Oliveira ND (2018) Penicillium and Talaromyces communities of sugarcane soils (Saccharum officinarum L.): ecological and phylogenetic aspects. J Agric Sci 9;10(4)
Rao, VG; Reddy, KA. Two new hyphomycetes. Indian J Bot; 1981; 4, pp. 108-114.
Rapper KB, Thom C (1949) A manual of the Penicillia. Williams & Wilkins co., Baltimore
Rathnayaka, AR; Chethana, KWT; Manawasinghe, IS; Wijesinghe, SN; de Silva, NI; Tennakoon, DS; Phillips, AJL; Liu, JK; Jones, EBG; Wang, Y; Hyde, KD. Lasiodiplodia: Generic revision by providing molecular markers, geographical distribution and haplotype diversity. Mycosphere; 2023; 14, pp. 1254-1339. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/14]
Ren, G; Wanasinghe, DN; de Farias, ARG; Hyde, KD; Yasanthika, E; Xu, J; Balasuriya, A; Chethana, KWT; Gui, H. Taxonomic Novelties of Woody Litter Fungi (Didymosphaeriaceae, Pleosporales) from the Greater Mekong Subregion. Biology; 2022; 11, 1660.1:CAS:528:DC%2BB38XjtFWitrvL [DOI: https://dx.doi.org/10.3390/biology11111660] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36421373][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9687740]
Richter, C; Wittstein, K; Kirk, PM; Stadler, M. An assessment of the taxonomy and chemotaxonomy of Ganoderma. Fungal Divers; 2015; 71, pp. 1-15. [DOI: https://dx.doi.org/10.1007/s13225-014-0313-6]
Rivas Plata, E; Lücking, R; Lumbsch, HT. A new classification for the family Graphidaceae (Ascomycota: Lecanoromycetes: Ostropales). Fungal Divers; 2012; 52, pp. 107-121. [DOI: https://dx.doi.org/10.1007/s13225-011-0135-8]
Ronikier, A; Lado, C; Meyer, M; De Basanta, DW. Two new species of nivicolous Lamproderma (Myxomycetes) from the mountains of Europe and America. Mycologia; 2010; 102, pp. 718-728. [DOI: https://dx.doi.org/10.3852/09-223] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/20524603]
Ronquist, F; Teslenko, M; van der Mark, P; Ayres, DL; Darling, A; Höhna, S; Larget, B; Liu, L; Suchard, MA; Huelsenbeck, JP. MrBayes 3.2: efficient Bayesian phylogenetic inference and model choice across a large model space. Syst Biol; 2012; 61, pp. 539-542. [DOI: https://dx.doi.org/10.1093/sysbio/sys029] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22357727][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3329765]
Rosado, AWC; Machado, AR; Freire, FDCO; Pereira, OL. Phylogeny, identification, and pathogenicity of Lasiodiplodia associated with postharvest stem-end rot of coconut in Brazil. Plant Dis; 2016; 100, pp. 561-568.1:CAS:528:DC%2BC28XitFeqtLzK [DOI: https://dx.doi.org/10.1094/PDIS-03-15-0242-RE] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30688600]
Ryvarden, L. Genera of polypores. Nomencult Taxonomy Synop Fungorum; 1991; 5, pp. 1-363.
Ryvarden, L. Studies in Neotropical Polypores 2: a preliminary key to neotropical species of Ganoderma with a laccate pileus. Mycologia; 2000; 92, pp. 180-191. [DOI: https://dx.doi.org/10.1080/00275514.2000.12061142]
Ryvarden, L. Neotropical polypores 1. Introduction: Hymenochaetaceae and Ganodermataceae. Synop Fungorum; 2004; 19, pp. 1-227.
Ryvarden, L. Studies in African aphyllophorales 24. A first checklist of polypores from Mozambique. Synop Fungorum; 2018; 38, pp. 20-24.
Ryvarden L, Johansen I (1980) A preliminary polypore flora of East Africa. Fungiflora 1–636
Saccardo, PA. Conspectus generum pyrenomycetum italicorum additis speciebus fungorum Venetorum novis vel criticis, systemate carpologico dispositorum. Atti Soc Veneziana-Trent-Istriana Sci Nat; 1875; 4, pp. 77-100.
Salvador-Montoya, CA; Popoff, OF; Reck, M; Drechsler-Santos, ER. Taxonomic delimitation of Fulvifomes robiniae (Hymenochaetales, Basidiomycota) and related species in America: F. squamosus sp. nov. Plant Syst Evol; 2018; 304, pp. 445-459. [DOI: https://dx.doi.org/10.1007/s00606-017-1487-7]
Salvador-Montoya, CA; Elias, SG; Popoff, OF; Robledo, GL; Urcelay, C; Góes-Neto, A; Martínez, S; Drechsler-Santos, ER. Neotropical Studies on Hymenochaetaceae: unveiling the Diversity and Endemicity of Phellinotus. J Fungi; 2022; 8, pp. 216-250.1:CAS:528:DC%2BB38XhtVCmurjP [DOI: https://dx.doi.org/10.3390/jof8030216]
Samarakoon, BC; Phookamsak, R; Wanasinghe, DN; Chomnunti, P; Hyde, KD; McKenzie, EH; Promputtha, I; Xu, JC; Li, YJ. Taxonomy and phylogenetic appraisal of Spegazzinia musae sp. nov. and S. deightonii (Didymosphaeriaceae, Pleosporales) on Musaceae from Thailand. MycoKeys; 2020; 70, pp. 19-37. [DOI: https://dx.doi.org/10.3897/mycokeys.70.52043] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32742179][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7387365]
Samarakoon BC, Wanasinghe DN, Samarakoon MC, Phookamsak R, McKenzie EH, Chomnunti P, Hyde KD, Lumyong S, Karunarathna SC (2020b) Multi-gene phylogenetic evidence suggests Dictyoarthrinium belongs in Didymosphaeriaceae (Pleosporales, Dothideomycetes) and Dictyoarthrinium musae sp. nov. on Musa from Thailand. MycoKeys 71:101
Samarakoon MC, Lumyong S, Manawasinghe IS, Suwannarach N, Cheewangkoon R (2023) Addition of Five Novel Fungal Flora to the Xylariomycetidae (Sordariomycetes, Ascomycota) in Northern Thailand. Journal of Fungi, 9(11), p.1065
Sánchez, A; Moreno, G. Catálogo de Myxomycetes de Segovia. Bol Soc Micol Madrid; 2016; 40, pp. 37-68.
Sanchez-Gonzalez, EI; Soares, TD; Zarpelon, TG; Zauza, EA; Mafia, RG; Ferreira, MA. Two new species of Calonectria (Hypocreales, Nectriaceae) causing Eucalyptus leaf blight in Brazil. Mycokeys; 2022; 91, pp. 169-197. [DOI: https://dx.doi.org/10.3897/mycokeys.91.84896] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36760892][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9849068]
Sang, XY; Wang, ZJ; Yang, YB; Wang, YL; Xiang, ML; Zhu, DY; Zhou, Y; Luo, XD; Zhao, LX. Antimicrobial Natural Products Produced by Soil-Derived Fungus Penicillium cremeogriseum W1–1. Indian J Microbiol; 2021; 61, pp. 519-523.1:CAS:528:DC%2BB3MXhs1Cmt7nL [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34744207][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8542016][DOI: https://dx.doi.org/10.1007/s12088-021-00957-z]
Santini, A; Mikušová, P; Sulyok, M; Krska, R; Labuda, R; Šrobárová, A. Penicillium strains isolated from Slovak grape berries taxonomy assessment by secondary metabolite profile. Mycotoxin Res; 2014; 30, pp. 213-220.1:CAS:528:DC%2BC2cXhtlart7bI [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25109845][DOI: https://dx.doi.org/10.1007/s12550-014-0205-3]
Saxena, RK. Palynological investigation of the Sindhudurg Formation in the type area, Sindhudurg District, Maharashtra, India. ONGC Bulletin; 2000; 37, pp. 157-166.
Schenck, S; Kendrick, WB; Pramer, D. A new nematode-trapping hyphomycete and a reevaluation of Dactylaria and Arthrobotrys. Can J Bot; 1977; 55, pp. 977-985. [DOI: https://dx.doi.org/10.1139/b77-115]
Schiefelbein U, Dolnik C, Bruyn U de, Schultz M, Thiemann R, Stordeur R, van den Boom PPG, Litterski B, Sipman HJM (2014) Interesting records of lichenized, lichenicolous and saprophytic fungi from northern Germany. Herzogia 27:237–256. https://doi.org/10.13158/heia.27.2.2014.237
Schimann, H; Bach, C; Lengelle, J; Louisanna, E; Barantal, S; Murat, C; Buée, M. Diversity and structure of fungal communities in neotropical rainforest soils: the effect of host recurrence. Microbial Ecol; 2017; 73, pp. 310-320. [DOI: https://dx.doi.org/10.1007/s00248-016-0839-0]
Schol-Schwarz, MB. The genus Epicoccum. Trans Br Mycol; 1959; 42, pp. 149-173. [DOI: https://dx.doi.org/10.1016/S0007-1536(59)80024-3]
Schroers, HJ. A monograph of Bionectria (Ascomycota, Hypocreales, Bionectriaceae) and its Clonostachys anamorphs; 2001; Utrecht, Centraalbureau voor Schimmelcultures:
Schüßler A, Walker C (2010) The Glomeromycota. A species list with new families and new genera. Schüßler A, Walker C, Gloucester. Published in December 2010 in libraries at The Royal Botanic Garden Edinburgh, The Royal Botanic Garden Kew, Botanische Staatssammlung Munich, and Oregon State University
Schüßler, A; Schwarzott, D; Walker, C. A new fungal phylum, the Glomeromycota: phylogeny and evolution. Mycol Res; 2001; 105, pp. 1413-1421. [DOI: https://dx.doi.org/10.1017/S0953756201005196]
Sejalon-Delmas, N; Magnier, A; Douds, DD. Bécard G (1998) Cytoplasmic autofluorescence of an arbuscular mycorrhizal fungus Gigaspora gigantea and nondestructive fungal observations in planta. Mycologia; 1998; 90, pp. 921-926. [DOI: https://dx.doi.org/10.1080/00275514.1998.12026986]
Senanayake, IC; Rathnayaka, AR; Marasinghe, DS; Calabon, MS; Gentekaki, E; Lee, HB; Hurdeal, VG; Pem, D; Dissanayake, LS; Wijesinghe, SN; Bundhun, D. Morphological approaches in studying fungi: collection, examination, isolation, sporulation and preservation. Mycosphere; 2020; 11, pp. 2678-2754. [DOI: https://dx.doi.org/10.5943/mycosphere/11/1/20]
Senanayake, IC; Rossi, W; Leonardi, M; Weir, A; McHugh, M; Rajeshkumar, KC; Verma, RK; Karunarathna, SC; Tibpromma, S; Ashtekar, N; Ashtamoorthy, SK. Fungal diversity notes 1611–1716: taxonomic and phylogenetic contributions on fungal genera and species emphasis in South China. Fungal Divers; 2023; 122, pp. 161-403. [DOI: https://dx.doi.org/10.1007/s13225-023-00523-6]
Seyedmousavi, S; Samerpitak, K; Rijs, AJ; Melchers, WJ; Mouton, JW; Verweij, PE; de Hoog, GS. Antifungal susceptibility patterns of opportunistic fungi in the genera Verruconis and Ochroconis. Antimicrob Agents Chemother; 2014; 58, pp. 3285-3292.1:CAS:528:DC%2BC2cXhtlCntLnO [DOI: https://dx.doi.org/10.1128/aac.00002-14] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/24687495][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4068420]
Shao QY, Ren YL, Cheng WH, Liang JD, Zhang YW, Liang ZQ, Han YF (2022) Verruconis guizhouensis sp. nov. (Sympoventuriaceae, Venturiales) from soil in China. Phytotaxa 572:259–267. https://doi.org/10.11646/PHYTOTAXA.572.3.4
Sheffy, MV; Dilcher, DL. Morphology and taxonomy of fungal spores. Palaeontographica Abt B; 1971; 133, pp. 34-51.
Shen, YM; Chung, WH; Huang, TC; Rodeva, R; Hung, TH. Unveiling Gymnosporangium corniforme, G. unicorne, and G. niitakayamense sp. nov. in Taiwan. Mycoscience; 2018; 59, pp. 218-228. [DOI: https://dx.doi.org/10.1016/j.myc.2017.11.003]
Shen, M; Zhang, JQ; Zhao, LL; Groenewald, JZ; Crous, PW; Zhang, Y. Venturiales. Stud Mycol; 2020; 96, pp. 185-308.1:STN:280:DC%2BB38bmtlSisQ%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32904190][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7452091][DOI: https://dx.doi.org/10.1016/j.simyco.2020.03.001]
Shen, S; Liu, SL; Jiang, JH; Zhou, LW. Addressing widespread misidentifications of traditional medicinal mushrooms in Sanghuangporus (Basidiomycota) through ITS barcoding and designation of reference sequences. IMA Fungus; 2021; 12, 10.1:CAS:528:DC%2BB38XhvF2nu7rO [DOI: https://dx.doi.org/10.1186/s43008-021-00059-x] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33853671][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8048060]
Shen, HW; Bao, DF; Boonmee, S; Lu, YZ; Su, XJ; Li, YX; Luo, ZL. Diversity of Distoseptispora (Distoseptisporaceae) taxa on submerged decaying wood from the Red River in Yunnan, China. Mycokeys; 2024; 102, 1. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38356851][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10862349][DOI: https://dx.doi.org/10.3897/mycokeys.102.116096]
Silva, DN; Talhinas, P; Várzea, V; Cai, L; Paulo, OS; Batista, D. Application of the Apn2/MAT locus to improve the systematics of the Colletotrichum gloeosporioides complex: an example from coffee (Coffea spp.) hosts. Mycologia; 2012; 104, pp. 396-409.1:CAS:528:DC%2BC38Xmt1CisbY%3D [DOI: https://dx.doi.org/10.3852/11-145] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/22086913]
Silva, EI; Jayasingha, P; Senanayake, S; Dandeniya, A; Munasinghe, DH. Microbiological study in a gneissic cave from Sri Lanka, with special focus on potential antimicrobial activities. Int J Speleol; 2021; 50,
Simmons, DR; Lund, J; Levitsky, T; Groden, E. Ophiocordyceps myrmicarum, a new species infecting invasive Myrmica rubra in Maine. J Invertebr Pathol; 2015; 125, pp. 23-30.1:CAS:528:DC%2BC2MXhtVClu7s%3D [DOI: https://dx.doi.org/10.1016/j.jip.2014.12.010] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25555674]
Singer, R. Marasmieae (Basidiomycetes-Tricholomataceae). Flora Neotrop; 1976; 17, pp. 1-347.
Singer, R. The Agaricales in modern taxonomy; 1986; 4 Koenigstein, Koeltz Scientific Books: 981.
Singer, R; Smith, AH. Proposals concerning the nomenclature of the gill fungi including a list of proposed lectotypes and genera conservanda. Mycologia; 1946; 38, pp. 240-299.1:STN:280:DyaH28%2FgtFaruw%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/20983184][DOI: https://dx.doi.org/10.1080/00275514.1946.12024058]
Sipman, HJ; Luecking, R; Aptroot, A; Chaves, JL; Kalb, K; Tenorio, LU. A first assessment of the Ticolichen biodiversity inventory in Costa Rica and adjacent areas: the thelotremoid Graphidaceae (Ascomycota: Ostropales). Phytotaxa; 2012; 55, pp. 1-214. [DOI: https://dx.doi.org/10.11646/phytotaxa.55.1.1]
Sivanesan A (1984) The bitunicate Ascomycetes and their anamorphs, J Cramer, Vaduz, Liechtenstein. 701
Somrithipol, S. A synnematous species of Dictyoarthrinium from Thailand. Mycologia; 2007; 99, pp. 792-796. [DOI: https://dx.doi.org/10.1080/15572536.2007.11832542] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/18268912]
Sonjak, S; Frisvad, JC; Gunde-Cimerman, N. Penicillium mycobiota in Arctic subglacial ice. Microb Ecol; 2006; 52, pp. 207-216. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/16897300][DOI: https://dx.doi.org/10.1007/s00248-006-9086-0]
Spatafora, JW; Quandt, CA; Kepler, RM; Sung, GH; Shrestha, B; Hywel-Jones, NL; Luangsa-ard, JJ. New 1F1N species combinations in Ophiocordycipitaceae (Hypocreales). IMA Fungus; 2015; 6, pp. 357-362. [DOI: https://dx.doi.org/10.5598/imafungus.2015.06.02.07] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26734546][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4681259]
Species Fungorum (2024) http://www.speciesfungorum.org/names/Names.asp. Accessed 17 Jan 2023
Spegazzini, CL. Fungi argentini additis nonnullis brasiliensibus montevideensibusque. Pugillus quartus (Continuacion). An Soc Cient Arg; 1881; 12, pp. 97-117.
Stone, WJ; Jones, JP. Corynespora blight of Sesame. Phytopathology; 1960; 50, pp. 263-266.
Su, HY; Hyde, KD; Maharachchikumbura, SSN; Ariyawansa, HA; Luo, ZL; Promputtha, I; Tian, Q; Lin, CG; Shang, QJ; Zhao, YC; Chai, HM; Liu, XY; Bahkali, AH; Bhat, DJ; McKenzie, EHC; Zhou, DQ. The families Distoseptisporaceae fam. nov., Kirschsteiniotheliaceae, Sporidesmiaceae and Torulaceae, with new species from freshwater in Yunnan Province. China Fungal Divers; 2016; 80, pp. 375-440. [DOI: https://dx.doi.org/10.1007/s13225-016-0362-0]
Subramanian CV (1971) Hyphomycetes, an account of Indian species, except Cercosporae. Indian Council of Agricultural Research, New Delhi, p 930
Suchard MA, Lemey P, Baele G, Ayres DL, Drummond AJ, Rambaut A (2018) Bayesian phylogenetic and phylodynamic data integration using BEAST 1.10. Virus E 4: vey016. https://doi.org/10.1093/ve/vey016
Sun, JZ; Liu, XZ; McKenzie, EH; Jeewon, R; Liu, JK; Zhang, XL; Zhao, Q; Hyde, KD. Fungicolous fungi: terminology, diversity, distribution, evolution, and species checklist. Fungal Divers; 2019; 95, pp. 337-430. [DOI: https://dx.doi.org/10.1007/s13225-019-00422-9]
Sun YR, Goonasekara ID, Thambugala KM, Jayawardena RS, Wang Y, Hyde KD. (2020) Distoseptispora bambusae sp. nov. (Distoseptisporaceae) on bamboo from China and Thailand. Biodiversity Data Journal 8: e53678. https://doi.org/10.3897/BDJ.8.e53678
Sun YR, Jayawardena RS, Hyde KD, Wang Y (2021) Kirschsteiniothelia thailandica sp. nov. (Kirschsteiniotheliaceae) from Thailand. Phytotaxa 490:172–182. https://doi.org/10.11646/phytotaxa.490.2.3
Sung, GH; Hywel-Jones, NL; Sung, JM; Luangsa-Ard, JJ; Shrestha, B; Spatafora, JW. Phylogenetic classification of Cordyceps and the clavicipitaceous fungi. Stud Mycol; 2007; 57, pp. 5-59. [DOI: https://dx.doi.org/10.3114/sim.2007.57.01] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/18490993][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2104736]
Świzdor A. Baeyer-Villiger (2013) Oxidation of some C19 steroids by Penicillium lanosocoeruleum. Molecule 18(11):13812-13822
Symanczik, S; Błaszkowski, J; Chwat, G; Boller, T; Wiemken, A; Al-Yahya’ei, MN. Three new species of arbuscular mycorrhizal fungi discovered at one location in a desert of Oman: Diversispora omaniana, Septoglomus nakheelum and Rhizophagus arabicus. Mycologia; 2014; 106, pp. 243-259.1:CAS:528:DC%2BC2cXpvFShsrc%3D [DOI: https://dx.doi.org/10.3852/106.2.243] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/24782493]
Symanczik, S; Al-Yahya’ei, MN; Kozłowska, A; Ryszka, P; Błaszkowski, J. A new genus, Desertispora, and a new species, Diversispora sabulosa, in the family Diversisporaceae (order Diversisporales, subphylum Glomeromycotina). Mycol Prog; 2018; 17, pp. 437-449. [DOI: https://dx.doi.org/10.1007/s11557-017-1369-y]
Talhinhas, P; Baroncelli, R. Hosts of Colletotrichum. Mycosphere; 2023; 14, pp. 158-261. [DOI: https://dx.doi.org/10.5943/mycosphere/14/si2/4]
Tamur HA, Al-Janabi HJ, Abood Al-Janabi JK, Mohsin LY, Al-Yassiry ZA (2019) Characterization and antagonistic activity of new causal agent of wilt disease in Imperata cylindrica (Marasmius palmivorus). J Pure Appl Microbiol 13:1525–1536. https://doi.org/10.22207/JPAM.13.3.24
Tamura, K. Estimation of the number of nucleotide substitutions when there are strong transition-transversion and G + C-content biases. Mol Biol Evol; 1992; 9, pp. 678-687.1:CAS:528:DyaK38Xks1eqsbY%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/1630306]
Tamura, K; Stecher, G; Kumar, S. MEGA11: molecular evolutionary genetics analysis version 11. Mol Biol Evol; 2021; 38, pp. 3022-3027.1:CAS:528:DC%2BB3MXitlCktrfN [DOI: https://dx.doi.org/10.1093/molbev/msab120] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33892491][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8233496]
Tanaka, K; Hirayama, K; Yonezawa, H; Hatakeyama, S; Harada, Y; Sano, T; Shirouzu, T; Hosoya, T. Molecular taxonomy of bambusicolous fungi: Tetraplosphaeriaceae, a new pleosporalean family with Tetraploa-like anamorphs. Stud Mycol; 2009; 64,
Tangni, EK; Wambacq, E; Bastiaanse, H; Haesaert, G; Pussemier, L; De Poorter, J; Foucart, G; Van Hove, F. Survey of fungal diversity in silages supplied to dairy cattle in Belgium over a two-year period. J Anim Sci Adv; 2017; 7,
Tao, SQ; Cao, B; Kakishima, M; Liang, YM. Species diversity, taxonomy, and phylogeny of Gymnosporangium in China. Mycologia; 2020; 112, pp. 941-973. [DOI: https://dx.doi.org/10.1080/00275514.2020.1790272] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33006908]
Tapfuma, KI; Lourens, A; Baatjies, L; Allie, N; Loxton, AG; Mavumengwana, V; Malgas-Enus, R. Antimycobacterial activity of Clonostachys rogersoniana MGK33 functionalized metal-coated magnetic nanoparticles. Mater Today Commun; 2023; 34, 1:CAS:528:DC%2BB38XjtFWgtrvP [DOI: https://dx.doi.org/10.1016/j.mtcomm.2022.105098] 105098.
Tarda, AS; Saparrat, MCN; Gómez, N. Assemblage of dematiaceous and Ingoldian fungi associated with leaf litter of decomposing Typha latifolia L. (Typhaceae) in riverine wetlands of the Pampean plain (Argentina) exposed to different water quality. J Environ Manag; 2019; 250, pp. 109-409.1:CAS:528:DC%2BC1MXhslKnu7fN [DOI: https://dx.doi.org/10.1016/j.jenvman.2019.109409]
Tchoumi, JMT; Coetzee, MPA; Rajchenberg, M; Roux, J. Poroid Hymenochaetaceae associated with trees showing wood-rot symptoms in the Garden Route National Park of South Africa. Mycologia; 2020; 112, pp. 722-741.1:CAS:528:DC%2BB3cXhsFSis7rK [DOI: https://dx.doi.org/10.1080/00275514.2020.1753160] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/32574523]
Tennakoon, DS; Kuo, CH; Maharachchikumbura, SS; Thambugala, KM; Gentekaki, E; Phillips, AJ; Bhat, DJ; Wanasinghe, DN; de Silva, NI; Promputtha, I; Hyde, KD. Taxonomic and phylogenetic contributions to Celtis formosana, Ficus ampelas, F. septica, Macaranga tanarius and Morus australis leaf litter inhabiting microfungi. Fungal Divers; 2021; 108, pp. 1-215. [DOI: https://dx.doi.org/10.1007/s13225-021-00474-w]
Thambugala, KM; Ariyawansa, HA; Li, YM; Boonmee, S; Hongsanan, S; Tian, Q; Singtripop, C; Bhat, DJ; Camporesi, E; Jayawardena, R; Liu, ZY; Xu, JC; Chukeatirote, E; Hyde, KD. Dothideales. Fungal Divers; 2014; 68, pp. 105-158. [DOI: https://dx.doi.org/10.1007/s13225-014-0303-8]
Thambugala, KM; Wanasinghe, DN; Phillips, AJ; Camporesi, E; Bulgakov, TS; Phukhamsakda, C; Ariyawansa, HA; Goonasekara, ID; Phookamsak, R; Dissanayake, A; Tennakoon, DS; Tibpromma, S; Chen, YY; Liu, ZY; Hyde, KD. Mycosphere notes 1–50: grass (Poaceae) inhabiting Dothideomycetes. Mycosphere; 2017; 8, pp. 697-796. [DOI: https://dx.doi.org/10.5943/mycosphere/8/4/13]
Than PP, Prihastuti H, Phoulivong, S, Taylor PWJ, Hyde KD (2003) Chilli anthracnose disease caused by Colletotrichum species. J Zhejiang Univ-Sci B 9:764–788
Thiyagaraja, V; Lücking, R; Ertz, D; Karunarathna, SC; Wanasinghe, DN; Lumyong, S; Hyde, KD. The evolution of life modes in Stictidaceae, with three novel taxa. J Fungi; 2021; 2, 105. [DOI: https://dx.doi.org/10.3390/jof7020105]
Thom, C. The Penicillia; 1930; Baltimore, Williams & Wilkins: pp. 1-644.
Tian, XM; Yu, HY; Zhou, LW; Decock, C; Vlasák, J; Dai, YC. Phylogeny and taxonomy of the Inonotus linteus complex. Fungal Divers; 2013; 58, pp. 159-169. [DOI: https://dx.doi.org/10.1007/s13225-012-0202-9]
Tibpromma, S; McKenzie, EHC; Karunarathna, SC; Mortimer, PE; Xu, J; Hyde, KD; Hu, DM. Muyocopron garethjonesii sp. nov. (Muyocopronales, Dothideomycetes) on Pandanus sp. Mycosphere; 2016; 7, pp. 1480-1489. [DOI: https://dx.doi.org/10.5943/mycosphere/7/9/19]
Tibpromma S, Hyde KD, Jeewon R et al (2017) Taxonomic and phylogenetic contributions to fungal taxa. Fungal Diversity 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. Fungal diversity notes 840–928: Micro-fungi associated with Pandanaceae. Fungal Divers; 2018; 93, pp. 1-160. [DOI: https://dx.doi.org/10.1007/s13225-018-0408-6]
Trakunyingcharoen, T; Lombard, L; Groenewald, JZ; Cheewangkoon, R; Toanun, C; Alfenas, AC; Crous, PW. Mycoparasitic species of Sphaerellopsis, and allied lichenicolous and other genera. IMA Fungus; 2014; 5, pp. 391-414. [DOI: https://dx.doi.org/10.5598/imafungus.2014.05.02.05] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25734030][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4329322]
Trifinopoulos, J; Nguyen, LT; von Haeseler, A; Minh, BQ. W-IQ-TREE: a fast online phylogenetic tool for maximum likelihood analysis. Nucleic Acids Res; 2016; 44, pp. W232-W235.1:CAS:528:DC%2BC2sXhtV2isbfN [DOI: https://dx.doi.org/10.1093/nar/gkw256] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27084950][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4987875]
Tulasne, LR; Tulasne, C. Mémoire sur les Ustilaginées comparées aux Urédinées. Ann Sci Nat Botanique Et Biologie Végétale; 1847; 3, pp. 12-126.
Valenzuela-Lopez, N; Cano-Lira, JF; Guarro, J; Sutton, DA; Wiederhold, N; Crous, PW; Stchigel, AM. Coelomycetous Dothideomycetes with emphasis on the families Cucurbitariaceae and Didymellaceae. Stud Mycol; 2018; 90, pp. 1-69.1:STN:280:DC%2BC1MzivFClsA%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29255336][DOI: https://dx.doi.org/10.1016/j.simyco.2017.11.003]
van der Hammen, T. El desarrollo de la flora Colombiana en los periodos geológicos. I. Maestrichtiano hasta Terciario más inferior. Boletín Geológico; 1954; 2, pp. 49-106. [DOI: https://dx.doi.org/10.32685/0120-1425/bolgeol2.1.1954.387]
Van der Hammen, T. El desarrollo de la flora Colombiana en los periodos geologicos–1. Maestrichtiano hasta Terciaro mas inferior (Una investigacion Palinologica de la formacion de Guaduas y equivalentes). Boletin Geologico; 1954; 2, pp. 49-106.
van Tuinen, D; Jacquot, E; Zhao, B; Gollotte, A; Gianinazzi-Pearson, V. Characterization of root colonization profiles by a microcosm community of arbuscular mycorrhizal fungi using 25S rDNA-targeted nested PCR. Mol Ecol; 1998; 7, pp. 879-887. [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/9691489][DOI: https://dx.doi.org/10.1046/j.1365-294x.1998.00410.x]
Vánky, K. Smut fungi of the world; 2011; St. Paul, APS Press:
Vellinga, EC. Notulae ad floram agaricinam neerlandicam—XXXVIII. Leucoagaricus subgenus Sericeomyces. Persoonia-Mol Phylogeny Evol Fungi; 2001; 17, pp. 473-480.
Vellinga EC (2003) Phylogeny and taxonomy of lepiotaceous fungi [PhD dissertation]. Leiden, The Netherlands: University of Leiden, p 259
Vellinga, EC. Genera in the family Agaricaceae: evidence from nrITS and nrLSU sequences. Mycol Res; 2004; 108, pp. 354-357.1:CAS:528:DC%2BD2cXjvFartLk%3D [DOI: https://dx.doi.org/10.1017/S0953756204009700] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/15209277]
Vellinga, EC; Sysouphanthong, P; Hyde, KD. The family Agaricaceae: phylogenies and two new white-spored genera. Mycologia; 2011; 103, pp. 494-509. [DOI: https://dx.doi.org/10.3852/10-204] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/21193599]
Veloso, JS; Câmara, MPS; Lima, WG; Michereff, SJ; Doyle, VP. Why species delimitation matters for fungal ecology: Colletotrichum diversity on wild and cultivated cashew in Brazil. Fungal Biol; 2018; 122, pp. 677-691. [DOI: https://dx.doi.org/10.1016/j.funbio.2018.03.005] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29880203]
Vieira, WAS; Lima, WG; Nascimento, ES; Michereff, SJ; Câmara, MPS; Doyle, VP. The impact of phenotypic and molecular data on the inference of Colletotrichum diversity associated with Musa. Mycologia; 2017; 109, pp. 912-934.1:CAS:528:DC%2BC1cXktlOqtrk%3D [DOI: https://dx.doi.org/10.1080/00275514.2017.1418577] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29494311]
Vieira, WAS; Bezerra, PA; da Silva, AC; Veloso, JS; Câmara, MPS; Doyle, VP. Optimal markers for the identification of Colletotrichum species. Mol Phylogenet Evol; 2020; 143, 1:CAS:528:DC%2BC1MXitlehurrP [DOI: https://dx.doi.org/10.1016/j.ympev.2019.106694] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31786239]106694.
Vieira, WAS; Veloso, JS; da Silva, AC; Nunes, AS; Doyle, VP; Castlebury, LA; Câmara, MPS. Elucidating the Colletotrichum spp. diversity responsible for papaya anthracnose in Brazil. Fungal Biol; 2022; 126, pp. 623-630. [DOI: https://dx.doi.org/10.1016/j.funbio.2022.08.001] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36116894]
Videira, SIR; Groenewald, JZ; Nakashima, C; Braun, U; Barreto, RW; de Wit, PJ; Crous, PW. Mycosphaerellaceae–chaos or clarity?. Stud Mycol; 2017; 87, pp. 257-421.1:STN:280:DC%2BC1M3lsFWhtA%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29180830][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5693839][DOI: https://dx.doi.org/10.1016/j.simyco.2017.09.003]
Visagie, CM; Houbraken, J; Frisvad, JC; Hong, SB; Klaassen, CH; Perrone, G; Seifert, KA; Varga, J; Yaguchi, T; Samson, RA. Identification and nomenclature of the genus Penicillium. Stud Mycol; 2014; 78, pp. 343-371.1:STN:280:DC%2BC2Mvgt1Sqsw%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2014.09.001] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/25505353][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC4261876]
Visagie, CM; Houbraken, J; Seifert, KA; Samson, RA; Jacobs, K. Four new Penicillium species isolated from the fynbos biome in South Africa, including a multigene phylogeny of section Lanata-Divaricata. Mycol Prog; 2015; 14, pp. 1-23.
Voglmayr, H; Jaklitsch, WM. Corynespora, Exosporium and Helminthosporium revisited: new species and generic reclassification. Stud Mycol; 2017; 87, pp. 43-76.1:STN:280:DC%2BC1cjhslarsw%3D%3D [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28649153][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5473648][DOI: https://dx.doi.org/10.1016/j.simyco.2017.05.001]
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. Large-scale generation and analysis of filamentous fungal DNA barcodes boosts coverage for kingdom fungi and reveals thresholds for fungal species and higher taxon delimitation. Stud Mycol; 2019; 92, pp. 135-154.1:STN:280:DC%2BB3c%2FjtVGrsQ%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2018.05.001] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29955203]
Wagner, T; Fischer, M. Proceedings towards a natural classification of the worldwide taxa Phellinus s.l. and Inonotus s.l., and phylogenetic relationships of allied genera. Mycologia; 2002; 94, pp. 998-1016. [DOI: https://dx.doi.org/10.1080/15572536.2003.11833156] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/21156572]
Walker, C; Sanders, FE. Taxonomic concepts in the Endogonaceae: III. The separation of Scutellospora gen. nov. from Gigaspora Gerd. & Trappe. Mycotaxon; 1986; 27, pp. 169-182.
Wanasinghe, DN; Phukhamsakda, C; Hyde, KD; Jeewon, R; Lee, HB; Gareth Jones, EB; Tibpromma, S; Tennakoon, DS; Dissanayake, AJ; Jayasiri, SC; Gafforov, Y. Fungal diversity notes 709–839: taxonomic and phylogenetic contributions to fungal taxa with an emphasis on fungi on Rosaceae. Fungal Divers; 2018; 89, pp. 1-236. [DOI: https://dx.doi.org/10.1007/s13225-018-0395-7]
Wang XC, Xi RJ, Li Y, Wang DM, Yao YJ (2012) The species identity of the widely cultivated Ganoderma, “G. lucidum” (Lingzhi), in China. PLoS ONE 7:e40857. https://doi.org/10.1371/journal.pone.0040857
Wang, XW; Houbraken, J; Groenewald, JZ; Meijer, M; Andersen, B; Nielsen, KF; Crous, PW; Samson, RA. Diversity and taxonomy of Chaetomium and chaetomium-like fungi from indoor environments. Stud Mycol; 2016; 84, pp. 145-224.1:STN:280:DC%2BC1c7lt1OntA%3D%3D [DOI: https://dx.doi.org/10.1016/j.simyco.2016.11.005] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28082757][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5226397]
Wang, M; Liu, F; Crous, PW; Cai, L. Phylogenetic reassessment of Nigrospora: Ubiquitous endophytes, plant and human pathogens. Persoonia; 2017; 39, pp. 118-142.1:STN:280:DC%2BC1MrnsFOiug%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2017.39.06] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29503473][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5832950]
Wang, Y; Hu, P; Pan, Y; Zhu, Y; Liu, X; Che, Y; Liu, G. Identification and characterization of the verticillin biosynthetic gene cluster in Clonostachys rogersoniana. Fungal Genet Biol; 2017; 103, pp. 25-33.1:CAS:528:DC%2BC2sXlsV2ktLs%3D [DOI: https://dx.doi.org/10.1016/j.fgb.2017.03.007] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/28376389]
Wang, Y; Ren, J; Li, H; Pan, Y; Liu, X; Che, Y; Liu, G. The disruption of verM activates the production of gliocladiosin A and B in Clonostachys rogersoniana. Org Biomol Chem; 2019; 17, pp. 6782-6785.1:CAS:528:DC%2BC1MXht1yitL7O [DOI: https://dx.doi.org/10.1039/C9OB01102A] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31276151]
Wang, W; de Silva, DD; Moslemi, A; Edwards, J; Ades, PK; Crous, PW; Taylor, PWJ. Colletotrichum species causing anthracnose of citrus in Australia. J Fungi; 2021; 7, 47.1:CAS:528:DC%2BB3MXhs1GhtrnO [DOI: https://dx.doi.org/10.3390/jof7010047]
Wang, L; Sun, C; Jia, S; Liang, YM. Identification and characterization of two new Gymnosporangium species causing rust on Juniperus rigida in China. Mycologia; 2022; 114, pp. 857-867. [DOI: https://dx.doi.org/10.1080/00275514.2022.2094116] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/35895294]
Wang, XW; Han, PJ; Bai, FY; Luo, A; Bensch, K; Meijer, M; Kraak, B; Han, DY; Sun, BD; Crous, PW; Houbraken, J. Taxonomy, phylogeny and identification of Chaetomiaceae with emphasis on thermophilic species. Stud Mycol; 2022; 101, pp. 121-243.1:STN:280:DC%2BB287htFKjsQ%3D%3D [DOI: https://dx.doi.org/10.3114/sim.2022.101.03] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/36059895][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC9365047]
Wang, CG; Zhao, H; Liu, HG; Zeng, GY; Yuan, Y; Dai, YC. A multi-gene phylogeny clarifies species diversity, taxonomy, and divergence times of Ceriporia and other related genera in Irpicaceae (Polyporales, Basidiomycota). Mycosphere; 2023; 14, pp. 1665-1729. [DOI: https://dx.doi.org/10.5943/mycosphere/17/1/9]
Wang, XW; Liu, SL; Zhou, LW. An updated taxonomic framework of Hymenochaetales (Agaricomycetes, Basidiomycota). Mycosphere; 2023; 14, pp. 452-496. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/6]
Watson, W. The classification of lichens. New Phytol; 1929; 28, pp. 1-36. [DOI: https://dx.doi.org/10.1111/j.1469-8137.1929.tb06745.x]
Wei, DP; Wanasinghe, DN; Gentekaki, E; Thiyagaraja, V; Lumyong, S; Hyde, KD. Morphological and phylogenetic appraisal of novel and extant taxa of Stictidaceae from northern Thailand. J Fungi; 2021; 7, 880. [DOI: https://dx.doi.org/10.3390/jof7100880]
Wei, DP; Wanasinghe, DN; Xu, JC; To-Anun, C; Mortimer, PE; Hyde, KD; Elgorban, AM; Madawala, S; Suwannarach, N; Karunarathna, SC; Tibpromma, S; Lumyong, S. Three Novel Entomopathogenic Fungi from China and Thailand. Front Microbiol; 2021; 11, [DOI: https://dx.doi.org/10.3389/fmicb.2020.608991] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/33584571][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC7873960]608991.
Wei, TP; Zhang, H; Zeng, XY; Crous, PW; Jiang, YL. Re-evaluation of Sympoventuriaceae. Persoonia; 2022; 48, pp. 219-260.1:STN:280:DC%2BB1c3otVeksg%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2022.48.07] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38234692][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10792283]
Wei DP, Gentekaki E, Wanasinghe DN, Hyde KD, To-Anun C, Cheewangkoon R (2022b) Taxonomy and morphology of Phacidiella kunmingensis sp. nov. (Stictidaceae) from southwest China. Phytotaxa, 573:70–84. https://doi.org/10.11646/PHYTOTAXA.573.1.4
Weir, BS; Johnston, PR; Damm, U. The Colletotrichum gloeosporioides species complex. Stud Mycol; 2012; 73, pp. 115-180.1:STN:280:DC%2BC3s7hvVOltw%3D%3D [DOI: https://dx.doi.org/10.3114/sim0011] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/23136459][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC3458417]
Wen, TC; Wei, DP; Long, FY; Zeng, XY; Kang, JC. Multigene phylogeny and HPLC analysis reveal fake Ophiocordyceps sinensis in markets. Mycosphere; 2016; 7, pp. 853-867. [DOI: https://dx.doi.org/10.5943/mycosphere/7/6/16]
Wen, TC; Long, FY; Kang, C; Wang, F; Zeng, W. Effects of additives and bioreactors on cordycepin production from Cordyceps militaris in liquid static culture. Mycosphere; 2017; 8, pp. 886-898. [DOI: https://dx.doi.org/10.5943/mycosphere/8/7/5]
Wen-Li, L; Maharachchikumbura, SS; Cheewangkoon, R; Liu, JK. Reassessment of Dyfrolomyces and Four New Species of Melomastia from Olive (Olea europaea) in Sichuan Province, China. J Fungi; 2022; 8, 76.1:CAS:528:DC%2BB38Xnt1Wksr8%3D [DOI: https://dx.doi.org/10.3390/jof8010076]
Whitton, SR; McKenzie, EH; Hyde, KD. Microfungi on the Pandanaceae: Zygosporium, a review of the genus and two new species. Fungal Divers; 2002; 12, pp. 207-222.
Whitton, SR; McKenzie, EHC; Hyde, KD. Microfungi on the Pandanaceae: Two new species of Camposporium and key to the genus. Fungal Divers; 2002; 11, pp. 177-187.
Wijayawardene, NN; Hyde, KD; Dai, DQ; Sánchez-García, M; Goto, BT; Saxena, RK; Erdoğdu, M; Selçuk, F; Rajeshkumar, KC; Aptroot, A; Błaszkowski, J; Boonyuen, N; da Silva, GA; de Souza, FA; Dong, W; Ertz, D; Haelewaters, D; Jones, EBG; Karunarathna, SC; Kirk, PM; Kukwa, M; Kumla, J; Leontyev, DV; Lumbsch, HT; Maharachchikumbura, SSN; Marguno, F; Martínez-Rodríguez, P; Mešić, A; Monteiro, JS; Oehl, F; Pawłowska, J; Pem, D; Pfliegler, WP; Phillips, AJL; Pošta, A; He, MQ; Li, JX; Raza, M; Sruthi, OP; Suetrong, S; Suwannarach, N; Tedersoo, L; Thiyagaraja, V; Tibpromma, S; Tkalčec, Z; Tokarev, YS; Wanasinghe, DN; Wijesundara, DSA; Wimalaseana, SDMK; Madrid, H; Zhang, GQ; Gao, Y; Sánchez-Castro, I; Tang, LZ; Stadler, M; Yurkov, A; Thines, M. Outline of Fungi and fungus-like taxa—2021. Mycosphere; 2022; 13, pp. 53-453. [DOI: https://dx.doi.org/10.5943/mycosphere/13/1/2]
Wilson, A; Desjardin, D; Horak, E. Agaricales of Indonesia: 5. The genus Gymnopus from Java and Bali. Sydowia; 2004; 56, pp. 137-210.
Winter G (1885) Pilze—Ascomyceten. In: Rabenhorst GL (ed) Kryptogamen-Flora von Deutschland, Oesterreich und der Schweiz 1:65–528.
Wu, SH. The Corticiaceae (Basidiomycetes) subfamilies Phlebioideae, Phanerochaetoideae and Hyphodermoideae in Taiwan. Acta Bot Fennica; 1990; 142, pp. 11-23.
Wu, WP. Sporidesmium, Endophragmiella and related genera from China. Fungal Divers Res Ser; 2005; 15, pp. 1-351.
Wu, W; Diao, Y. Anamorphic chaetosphaeriaceous fungi from China. Fungal Divers; 2022; 116, pp. 1-546.1:CAS:528:DC%2BB38XisF2iu7nE [DOI: https://dx.doi.org/10.1007/s13225-022-00509-w]
Wu, W; Zhuang, W. Sporidesmium, Endophragmiella and related genera from China; 2005; Hong Kong, Fungal Divers Press:
Wu, SH; Dai, YC; Hattori, T; Yu, TW; Wang, DM; Parmasto, E; Chang, HY; Shih, SY. Species clarification for the medicinally valuable ‘sanghuang’ mushroom. Bot Stud; 2012; 53, pp. 135-149.
Wu, G; Feng, B; Xu, JP; Zhu, XT; Li, YC; Zeng, NK; Hosen, MI; Yang, ZL. Molecular phylogenetic analyses redefine seven major clades and reveal 22 new generic clades in the fungal family Boletaceae. Fungal Divers; 2014; 69, pp. 93-115. [DOI: https://dx.doi.org/10.1007/s13225-014-0283-8]
Wu SH, Wei CL, Chang CC (2020) Sanghuangporus vitexicola sp. nov. (Hymenochaetales, Basidiomycota) from tropical Taiwan. Phytotaxa 475:43–51. https://doi.org/10.11646/phytotaxa.475.1.4
Wu NA, Dissanayake AJ, Manawasinghe IS, Rathnayaka AR, Liu JK, Phillips AJ, Promputtha I, Hyde KD (2021) https://botryosphaeriales.org/, an online platform for up-to-date classification and account of taxa of Botryosphaeriales. Database. 2021: baab061. https://doi.org/10.1093/database/baab061
Wu, G; Li, HJ; Horak, E; Wu, K; Li, GM; Yang, ZL. New taxa of Boletaceae from China. Mycosphere; 2023; 14, pp. 745-776. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/9]
Xiao, YP; Hongsanan, S; Hyde, KD; Brooks, S; Xie, N; Long, FY; Wen, TC. Two new entomopathogenic species of Ophiocordyceps in Thailand. MycoKeys; 2019; 47, pp. 53-74.1:CAS:528:DC%2BB38XhvFCgtr7O [DOI: https://dx.doi.org/10.3897/mycokeys.47.29898]
Xiao, YP; Wang, YB; Hyde, KD; Eleni, G; Sun, JZ; Yang, Y; Meng, J; Yu, H; Wen, TC. Polycephalomycetaceae, a new family of clavicipitoid fungi segregates from Ophiocordycipitaceae. Fungal Divers; 2023; 120,
Xu TM, Zeng YF, Cheng YH, Zhao CL (2020) Phlebiopsis lacerata sp. nov. (Polyporales, Basidiomycota) from southern China. Phytotaxa 440:268–280. https://doi.org/10.11646/phytotaxa.440.4.2
Xu, RF; Karunarathna, SC; Phukhamsakda, C; Dai, DQ; Elgorban, AM; Suwannarach, N; Kumla, J; Wang, XY; Tibpromma, S. Four new species of Dothideomycetes (Ascomycota) from Pará Rubber (Hevea brasiliensis) in Yunnan Province, China. MycoKeys; 2024; 103, pp. 71-95. [DOI: https://dx.doi.org/10.3897/mycokeys.103.117580] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/38560534][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC10980880]
Yang, ZL. Molecular techniques revolutionize knowledge of basidiomycete evolution. Fungal Divers; 2011; 50, pp. 47-58. [DOI: https://dx.doi.org/10.1007/s13225-011-0121-1]
Yang, J; Maharachchikumbura, SSN; Bhat, DJ; Hyde, KD; McKenzie, EHC; Jones, EBG; Al-Sadi, AM; Lumyong, S. Fuscosporellales, a new order of aquatic and terrestrial Hypocreomycetidae (Sordariomycetes). Cryptogam Mycol; 2016; 37, pp. 449-475. [DOI: https://dx.doi.org/10.7872/crym/v37.iss4.2016.449]
Yang, J; Liu, JK; Hyde, KD; Jones, EBG; Liu, ZY. Two new species in Fuscosporellaceae from freshwater habitats in Thailand. Mycosphere; 2017; 8, pp. 1893-1903. [DOI: https://dx.doi.org/10.5943/mycosphere/8/10/12]
Yang T, Groenewald JZ, Cheewangkoon R, Jami F, Abdollahzadeh J, Lombard L, Crous PW (2017b) Families, genera, and species of Botryosphaeriales. Fungal Biol 121:322–346. https://doi.org/10.1016/j.funbio.2016.11.001
Yang, J; Maharachchikumbura, SSN; Liu, JK; Hyde, KD; Jones, EBG; Al-Sadi, AM; Liu, ZY. Pseudostanjehughesia aquitropica gen. et sp. nov. and Sporidesmium sensu lato species from freshwater habitats. Mycol Prog; 2018; 17, pp. 591-616. [DOI: https://dx.doi.org/10.1007/s11557-017-1339-4]
Yang, J; Liu, LL; Jones, EBG; Hyde, KD; Liu, ZY; Bao, DF; Liu, NG; Li, WL; Shen, HW; Yu, XD; Liu, JK. Freshwater fungi from karst landscapes in China and Thailand. Fungal Divers; 2023; 119, pp. 1-212. [DOI: https://dx.doi.org/10.1007/s13225-023-00514-7]
Yip, HY. Umbelopsis fusiformis sp. nov. from a wet sclerophyll forest and a new combination for Mortierella ovata. Trans Br Mycol Soc; 1986; 86, pp. 334-337. [DOI: https://dx.doi.org/10.1016/S0007-1536(86)80167-X]
Yu XD, Zhang SN, Liu JK (2024) Novel hyphomycetous fungi associated with bamboo from Sichuan, China. Phytotaxa 634:235–254. https://doi.org/10.11646/PHYTOTAXA.634.3.4
Yuan, HS; Lu, X; Dai, YC; Hyde, KD; Kan, YH; Kušan, I; He, SH; Liu, NG; Sarma, VV; Zhao, CL; Cui, BK. Fungal diversity notes 1277–1386: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers; 2020; 104, pp. 1-266. [DOI: https://dx.doi.org/10.1007/s13225-020-00461-7]
Yun, HY; Hong, SG; Rossman, AY; Lee, SK; Lee, KJ; Bae, KS. The rust fungus Gymnosporangium in Korea including two new species. G Monticola and g Unicorne Mycologia; 2009; 101, pp. 790-809. [DOI: https://dx.doi.org/10.3852/08-221]
Zadravec, M; Lešić, T; Brnić, D; Pleadin, J; Kraak, B; Jakopović, Ž; Perković, I; Vahčić, N; Tkalec, VJ; Houbraken, J. Regional distribution and diversity of Aspergillus and Penicillium species on Croatian traditional meat products. Int J Food Microbiol; 2023; 2,
Zeng, ZQ; Zhuang, WY. Three New Species of Clonostachys (Hypocreales, Ascomycota) from China. J Fungi; 2022; 8, 1027.1:CAS:528:DC%2BB38XivVemtLbJ [DOI: https://dx.doi.org/10.3390/jof8101027]
Zha, LS; Huang, SK; Xiao, YP; Boonmee, S; Eungwanichayapant, PD; McKenzie, EHC; Kryukov, V; Wu, XL; Hyde, KD; Wen, TC. An Evaluation of Common Cordyceps (Ascomycetes) Species Found in Chinese Markets. Int J Med Mushrooms; 2018; 20, pp. 1149-1162. [DOI: https://dx.doi.org/10.1615/IntJMedMushrooms.2018027330] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/30806296]
Zha, LS; Xiao, YP; Jeewon, R; Zou, X; Wang, X; Boonmee, S; Eungwanichayapant, PD; McKenzie, EHC; Hyde, KD; Wen, TC. Notes on the medicinal mushroom Chanhua (Cordyceps cicadae (Miq.) Massee). Chiang Mai J Sci; 2019; 46, pp. 1023-1035.
Zhang, KQ; Hyde, KD. Nematode-trapping fungi; 2014; New York, Springer:
Zhang, K; Ma, L; Zhang, X. New species and records of Shrungabeeja from southern China. Mycologia; 2009; 101,
Zhang, Y; Yu, ZF; Baral, HO; Qiao, M; Zhang, KQ. Pseudorbilia gen. nov. (Orbiliaceae) from Yunnan, China. Fungal Divers; 2007; 26, pp. 305-312.
Zhang, H; Hyde, KD; Mckenzie, EH; Bahkali, AH; Zhou, D. Sequence data reveals phylogenetic affinities of Acrocalymma aquatica sp. nov., Aquasubmersa mircensis gen. et sp. nov. and Clohesyomyces aquaticus (freshwater coelomycetes). Cryptogam Mycol; 2012; 33, pp. 333-346. [DOI: https://dx.doi.org/10.7872/crym.v33.iss3.2012.333]
Zhang, T; Deng, X; Yu, Y; Zhang, M; Zhang, Y. Pseudochaetosphaeronema ginkgonis sp. nov., an endophyte isolated from Ginkgo biloba. Int J Syst Evol Microbiol; 2016; 66, pp. 4377-4381.1:CAS:528:DC%2BC1cXitlSntr%2FM [DOI: https://dx.doi.org/10.1099/ijsem.0.00135] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/27488177]
Zhang, ZF; Liu, F; Zhou, X; Liu, XZ; Liu, SJ; Cai, L. Culturable mycobiota from Karst caves in China, with descriptions of 20 new species. Persoonia; 2017; 39, pp. 1-31. [DOI: https://dx.doi.org/10.3767/persoonia.2017.39.01] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/29503468][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC5832949]
Zhang, Y. Lasiodiplodia chinensis, a new holomorphic species from China. Mycosphere; 2017; 8, pp. 521-530. [DOI: https://dx.doi.org/10.5943/mycosphere/8/2/3]
Zhang, W; Groenewald, JZ; Lombard, L; Schumacher, RK; Phillips, AJ; Crous, PW. Evaluating species in Botryosphaeriales. Persoonia -Mol Phylogeny Evol Fungi; 2021; 46, pp. 63-115.1:STN:280:DC%2BB2MbosF2mtQ%3D%3D
Zhang, X; Qin, WQ; Liang, ZQ; Tian, R; Zeng, NK. Rubroboletus flammeus (Boletaceae, Boletales), a novel species unveiled from subtropical China based on morphological and phylogenetic evidence. Arch Microbiol; 2022; 204, 379.1:CAS:528:DC%2BB38XhsFCksb%2FM [DOI: https://dx.doi.org/10.1007/s00203-022-02998-4] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/35678898]
Zhang Q, Nizamani MM, Feng Y, Yang YQ, Jayawardena RS, Hyde KD, Wang Y, Li C (2023a) Genome-scale and multi-gene phylogenetic analyses of Colletotrichum spp. host preference and associated with medicinal plants. Mycosphere 14(2):1–106
Zhang L, Shen HW, Bao DF, Li LL, Su XJ, Luo ZL (2023b) Wongia suae sp. nov., a lignicolous freshwater fungus from Yuanjiang (Red River) Basin, China. Phytotaxa 616:258–268. https://doi.org/10.11646/PHYTOTAXA.616.3.5
Zhang, YX; Chen, JW; Manawasinghe, IS; Lin, YH; Jayawardena, RS; McKenzie, EHC; Hyde, KD; Xiang, MM. Identification and characterization of Colletotrichum species associated with ornamental plants in Southern China. Mycosphere; 2023; 14,
Zhang Q, Si J, Li HJ (2023d) A new Marasmius species (Agaricales, Marasmiaceae) with sessile basidiomata growing on wood, from Yunnan, China. Phytotaxa 578:169–179. https://doi.org/10.11646/PHYTOTAXA.578.2.3
Zhao K, Wu G, Yang ZL (2014) A new genus, Rubroboletus, to accommodate Boletus sinicus and its allies. Phytotaxa 188:61–77. https://doi.org/10.11646/phytotaxa.188.2.1
Zhao, P; Qi, XH; Crous, PW; Duan, WJ; Cai, L. Gymnosporangium species on Malus: species delineation, diversity and host alternation. Persoonia; 2020; 45, pp. 68-100.1:STN:280:DC%2BB2crgvFyksg%3D%3D [DOI: https://dx.doi.org/10.3767/persoonia.2020.45.03] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34456372][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8375348]
Zhao YN, Nakasone KK, Chen CC, Liu SL, Cao YF, He SH (2020b) A Contribution to the Phylogeny and Taxonomy of Phlebiopsis (Polyporales, Basidiomycota). Res Square. 1–59. https://doi.org/10.21203/rs.3.rs-86969/v1
Zhao, YN; He, SH; Nakasone, KK; Kumara, KLW; Chen, CC; Liu, SL; Ma, HX; Huang, MR. Global Phylogeny and Taxonomy of the Wood Decaying Fungal Genus Phlebiopsis (Polyporales, Basidiomycota). Front Microbiol; 2021; 12, pp. 1-20. [DOI: https://dx.doi.org/10.3389/fmicb.2021.622460]
Zheng, HF; Huang, FC; Liu, B; Shao, YY; Qin, PS. Fulvifomes nonggangensis and F. tubogeneratus (Hymenochaetales, Basidiomycota): two new species from Southern China based on morphological and molecular evidence. Mycobiology; 2021; 49, pp. 213-222. [DOI: https://dx.doi.org/10.1080/12298093.2021.1932162] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/34290546][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC8259859]
Zhou, LW. Fulvifomes hainanensis sp. nov. and F. indicus comb. nov. (Hymenochaetales, Basidiomycota) evidenced by a combination of morphology and phylogeny. Mycoscience; 2014; 55, pp. 70-77. [DOI: https://dx.doi.org/10.1016/j.myc.2013.05.006]
Zhou, LW. Fulvifomes imbricatus and F. thailandicus (Hymenochaetales, Basidiomycota): two new species from Thailand based on morphological and molecular evidence. Mycol Prog; 2015; 14, pp. 1-8. [DOI: https://dx.doi.org/10.1007/s11557-015-1116-1]
Zhou, LW; Vlasák, J; Decock, C; Assefa, A; Stenlid, J; Abate, D; Wu, SH; Dai, YC. 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; 2016; 77, pp. 335-347. [DOI: https://dx.doi.org/10.1007/s13225-015-0335-8]
Zhou, LW; Vlasák, J; Qin, WM; Dai, YC. Global diversity and phylogeny of the Phellinus igniarius complex (Hymenochaetales, Basidiomycota) with the description of five new species. Mycologia; 2016; 108, pp. 192-204. [DOI: https://dx.doi.org/10.3852/15-099] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/26553776]
Zhou, YY; Zhang, W; Wu, LN; Zhang, J; Tan, HY; Chethana, KWT; Manawasinghe, IS; Liu, M; Li, XH; Hyde, KD; Yan, JY. Diversity of fungal communities associated with grapevine trunk diseases in China. Mycosphere; 2023; 14, pp. 1340-1435. [DOI: https://dx.doi.org/10.5943/mycosphere/14/1/15]
Zhu L, Xing JH, Cui BK (2017) Morphological characters and phylogenetic analysis reveal a new species of Sanghuangporus from China. Phytotaxa 311:270–276. https://doi.org/10.11646/phytotaxa.311.3.7
Zhu, L; Song, J; Zhou, JL; Si, J; Cui, BK. Species diversity, phylogeny, divergence time, and biogeography of the genus Sanghuangporus (Basidiomycota). Front Microbiol; 2019; 10, 812. [DOI: https://dx.doi.org/10.3389/fmicb.2019.00812] [PubMed: https://www.ncbi.nlm.nih.gov/pubmed/31057518][PubMedCentral: https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6478708]
Zimmerman, A. On some fungi observed on tropical cultivated plants III. Cent J Bacteriol Parasite Sci; 1902; 8, pp. 216-221.
Zulfiqar, R; Asghar, HS; Habib, K; Khalid, AN. Two new species of the genus Oxneriaria (lichenized Ascomycota: Megasporaceae) from Pakistan. Plant Syst Evol; 2023; 309, 2.1:CAS:528:DC%2BB3sXntVKitQ%3D%3D [DOI: https://dx.doi.org/10.1007/s00606-022-01836-w]
© The Author(s) under exclusive licence to Mushroom Research Foundation 2024.