Academia.eduAcademia.edu
Fungal Diversity https://doi.org/10.1007/s13225-020-00444-8 Taxonomic and phylogenetic contributions to fungi associated with the invasive weed Chromolaena odorata (Siam weed) Ausana Mapook1,2,3,4 · Kevin D. Hyde1,3,5 · Eric H. C. McKenzie6 · E. B. Gareth Jones7 · D. Jayarama Bhat8 · Rajesh Jeewon9 · Marc Stadler4 · Milan C. Samarakoon3 · Maitree Malaithong2 · Benjawan Tanunchai2 · François Buscot2,11 · Tesfaye Wubet2,10,11 · Witoon Purahong2 Received: 1 February 2020 / Accepted: 19 March 2020 © MUSHROOM RESEARCH FOUNDATION 2020 Abstract This article provides morphological descriptions and illustrations of microfungi associated with the invasive weed, Chromolaena odorata, which were mainly collected in northern Thailand. Seventy-seven taxa distributed in ten orders, 23 families (of which Neomassarinaceae is new), 12 new genera (Chromolaenicola, Chromolaenomyces, Longiappendispora, Pseudocapulatispora, Murichromolaenicola, Neoophiobolus, Paraleptospora, Pseudoroussoella, Pseudostaurosphaeria, Pseudothyridariella, Setoarthopyrenia, Xenoroussoella), 47 new species (Aplosporella chromolaenae, Arthrinium chromolaenae, Chromolaenicola chiangraiensis, C. lampangensis, C. nanensis, C. thailandensis, Chromolaenomyces appendiculatus, Diaporthe chromolaenae, Didymella chromolaenae, Dyfrolomyces chromolaenae, Leptospora chromolaenae, L. phraeana, Longiappendispora chromolaenae, Memnoniella chromolaenae, Montagnula chiangraiensis, M. chromolaenae, M. chromolaenicola, M. thailandica, Murichromolaenicola chiangraiensis, M. chromolaenae, Muyocopron chromolaenae, M. chromolaenicola, Neomassarina chromolaenae, Neoophiobolus chromolaenae, Neopyrenochaeta chiangraiensis, N. chromolaenae, N. thailandica, N. triseptatispora, Nigrograna chromolaenae, Nothophoma chromolaenae, Paraleptospora chromolaenae, P. chromolaenicola, Patellaria chromolaenae, Pseudocapulatispora longiappendiculata, Pseudoroussoella chromolaenae, Pseudostaurosphaeria chromolaenae, P. chromolaenicola, Pseudothyridariella chromolaenae, Pyrenochaetopsis chromolaenae, Rhytidhysteron chromolaenae, Setoarthopyrenia chromolaenae, Sphaeropsis chromolaenicola, Tremateia chiangraiensis, T. chromolaenae, T. thailandensis, Xenoroussoella triseptata, Yunnanensis chromolaenae), 12 new host records, three new taxonomic combinations (Chromolaenicola siamensis, Pseudoroussoella elaeicola, Pseudothyridariella mahakashae), and two reference specimens (Torula chromolaenae, T. fici) are described and illustrated. Unlike some other hosts, e.g. bamboo (Poaceae) and Pandanaceae, the dominant group of fungi on Siam weed were Dothideomycetes. Only 15 species previously recorded from northern Thailand were found in this study. Most of the taxa are likely to have jumped hosts from surrounding plants and are unlikely to be a specialist to Siam weed. Most fungal families found on Siam weed had divergence estimates with stem ages in the Cretaceous, which coincided with the expected origin of the host family (Asteraceae). This further indicates that the species have jumped hosts, as it is unlikely that the taxa on the alien Siam weed came from the Americas with its host. They may, however, have jumped from other Asteraceae hosts. In a preliminary screening 40 (65%) of the 62 species tested showed antimicrobial activity and thus, the fungi associated with C. odorata may be promising sources of novel bioactive compound discovery. We provide a checklist of fungi associated with C. odorata based on the USDA Systematic Mycology and Microbiology Laboratory (SMML) database, relevant literature and our study. In total, 130 taxa (116 identified and 14 unidentified species) are distributed in 20 orders, 48 families and 85 genera. Pseudocercospora is the most commonly encountered genus on Siam weed. Keywords 60 new taxa · Antimicrobial activity · Ascomycota · Checklist · Dothideomycetes · Evolutionary divergence times · Multi-gene phylogenetics · Sordariomycetes * Kevin D. Hyde kdhyde3@gmail.com Extended author information available on the last page of the article 13 Vol.:(0123456789) Fungal Diversity Introduction Invasive plants potentially affect ecosystem processes including decomposition and nutrient cycling as they may be more readily decomposed than native species and have a higher nitrogen concentration and lower carbon to nitrogen ratio (Liao et al. 2008; Bassett et al. 2010). They may also influence micro-environmental conditions and stimulate the activity of microbial decomposers (Březina et al. 2006; Heneghan et al. 2006; Tůma et al. 2009; Esperschütz et al. 2011, 2013). Chromolaena odorata (L.) King & Robinson is a serious invasive weed in the old-world tropics and subtropics (Holm et al. 1977; Gautier 1992; den Breeÿen et al. 2006). It is known by several common names including “Siam weed” and belongs to the flowering plant family Asteraceae (Jeffrey 2007). Siam weed is native to the Americas, where it occurs from southern Florida and Texas in the USA to northern Argentina (Gautier 1992). This important invasive weed has spread to tropical regions of Africa, Asia, and Oceania, impacting economically, ecologically and environmentally on forests (Holm et al. 1977; Moore 2004; Muniappan et al. 2005; Zachariades et al. 2013; Catarino et al. 2019). Two biotypes of Siam weed have been recognized, the South African biotype (SA), and the Asian/West African biotype (AWA). Based on comparison of both haplotype and multilocus genotypes of native and invasive populations, Trinidad and Tobago was proposed as the source location of the Siam weed Asian biotype (Yu et al. 2014), and the invasive species shows higher competitive ability under nutrient-rich conditions than in its native regions (Qin et al. 2013; Shao et al. 2018). The invasive weed also affects conservation and ecotourism, as it has spread to large areas of agricultural land, causing reduced cropland and crop yields, and reduced plant native species richness (Macdonald 1983; Cronk and Fuller 1995; Goodall and Erasmus 1996; Rose 1997; Matthews and Brand 2004). Siam weed has, however, also been considered as a nutrient sink with potential benefit to crops, as a source of organic matter and nutrients (Koutika and Rainey 2010). The weed can adapt to acidic soils better than some leguminous plants (Koutika and Rainey 2010) and has been identified as a hyper-accumulator of heavy metals, especially cadmium and zinc. These two harmful heavy metals play an important role in ecosystems, affecting the activity of soil microbial communities (Giller et al. 1998; Yao et al. 2000; Tanhan et al. 2007; Okoronkwo et al. 2014; Ruangdech et al. 2017). Siam weed also has numerous biological activities, such as antiviral (Pisutthanan et al. 2005; Vital and Rivera 2009; Atindehou et al. 2013; Stanley et al. 2014; Hanphakphoom and Krajangsang 2016), antibacterial (Suksamrarn et al. 2004; Johari et al. 2012; Kigigha and Zige 2013), 13 antifungal (Naidoo et al. 2011; Ngono et al. 2006; Vijayaraghavan et al. 2013), anthelmintic and antiprotozoal (Vital and Rivera 2009), antibiofilm (Yahya et al. 2014), antihepatotoxic (Asomugha et al. 2014), antioxidant (Akinmoladun et al. 2007; Boudjeko et al. 2015), anticancer (Kouamé et al. 2013; Adedapo et al. 2016), anticonvulsant or antiepileptic (Amazu et al. 2013; Kanase and Shaikh 2018), antidiabetic (Onkaramurthy et al. 2013; Aba et al. 2015) anti-diarrheal (Aba et al. 2015), anti-inflammatory (Pandith et al. 2013b; Hanh et al. 2011) and antiparasitic (Ezenyi et al. 2014).The weed is widely used as a folk medicine for wound healing treatments (Vaisakh and Pandey 2012; Pandith et al. 2013a; Sirinthipaporn and Jiraungkoorskul 2017) in India (Ayyanar and Ignacimuthu 2009; Thomas et al. 2014), Pakistan (Abbasi et al. 2010), Bangladesh (Rani and Mohammed 2012), and Ghana (Barku et al. 2014). Microbial endophytes and other microorganisms associated with Siam weed have also been reported for their biological properties. Buzugbe et al. (2018) found that an unidentified endophytic fungus, isolated from leaves of Siam weed, produced secondary metabolites with antimicrobial and antioxidant properties. Interestingly, microorganisms associated with weeds and medicinal plants are beneficial resources for increasing resistance of the host plants from unfavorable conditions and enhancing the growth of the host plant through their interactions (Köberl et al. 2013; Jia et al. 2016; Trognitz et al. 2016; Huang et al. 2018). For example, Trichoderma species, which are commonly soil inhabitants and can also be saprotrophs, mycoparasites, endophytes, as well as human pathogens (Walsh et al. 2004; Samuels 2006; Jaklitsch 2009; Kantarcioğlu et al. 2009; Chaverri and Samuels 2013), and have been reported to have antimicrobial and antiproliferative activities, as well as plant growth regulator activity (Chaverri and Samuels 2013; Leylaie and Zafari 2018; Zhang et al. 2018). Numerous Trichoderma species can also potentially protect their host plant from disease infection and have been used as biological control agents for plant protection (Harman and Kubicek 1998; Harman et al. 2004; Hyde et al. 2019b). Plant-associated microbes, especially microfungi associated with some selected groups of plants in Thailand and adjacent areas in China, have been studied to provide a hostfungus database for future studies and increase knowledge of fungal diversity, as well as new fungal discovery. For example, Dai et al. (2017) studied microfungi from bamboo, resulting in the discovery of 27 new species. Doilom et al. (2017) reported 28 fungal species on Tectona grandis, and Tibpromma et al. (2018) described 65 new species of microfungi on Pandanaceae. Although 68 fungal species have been recorded from Siam weed according to Farr and Rossman (2020; from https ://nt.ars-grin.gov/funga ldata bases /), few taxa have been reported from Thailand. For example, Tibpromma et al. (2017) described a new species, Fungal Diversity Hermatomyces chromolaenae and Li et al. (2017) described a new species, Torula chromolaenae with a new host record for T. fici from Siam weed in northern Thailand. Crous et al. (2018a) introduced a new genus, Neocochlearomyces, which was collected from leaves of Siam weed in Thailand. However, the diversity of fungi associated with Siam weed is still relatively unstudied and is likely to be much higher, as have been shown for other groups of fungi in northern Thailand (Hyde et al. 2018a, b). In this study, we provide morphological descriptions and illustrations of saprobic fungi associated with Chromolaena odorata, which were mainly collected in northern Thailand. Both multigene analyses and morphological comparison were used to confirm the current taxonomic placements of these fungi. The difference between fungi on Siam weed and other hosts will be explored for evidence of host-jumping. Results of preliminary screening for antimicrobial activity of fungi on Siam weed is provided. Furthermore, a checklist of fungi associated with C. odorata is provided based on the USDA Systematic Mycology and Microbiology Laboratory (SMML) database (Farr and Rossman 2020), together with relevant literature, and the author’s results from this study. Table of contents The numbers of taxa in this study are organized following the outline of Ascomycota (Wijayawardene et al. 2018) and updated from recent relevant literature. Phylum Ascomycota Caval.-Sm. Class Dothideomycetes sensu O.E. Erikss. & Winka Subclass Pleosporomycetidae C.L. Schoch et al. Hysteriales Lindau Hysteriaceae Chevall. 1. Rhytidhysteron bruguierae Dayarathne, in Dayarathne et al., Mycosphere 11(1): 20 (2020), new host record. 2. Rhytidhysteron chromolaenae Mapook & K.D. Hyde, sp. nov. Pleosporales Luttrell ex M.E. Barr Acrocalymmaceae Crous & Trakun. 3. Acrocalymma medicaginis Alcorn & J.A.G. Irwin, Trans. Br. mycol. Soc. 88(2): 163 (1987), new host record. Didymellaceae Gruyter et al. 4. Didymella chromolaenae Mapook & K.D. Hyde, sp. nov. 5. Nothophoma chromolaenae Mapook & K.D. Hyde, sp. nov. Didymosphaeriaceae Munk 6. Chromolaenicola Mapook & K.D. Hyde, gen. nov. 7. Chromolaenicola chiangraiensis Mapook & K.D. Hyde, sp. nov. 8. Chromolaenicola lampangensis Mapook & K.D. Hyde, sp. nov. 9. Chromolaenicola nanensis Mapook & K.D. Hyde, sp. nov. 10. Chromolaenicola siamensis (Jayasiri, E.B.G. Jones & K.D. Hyde) Mapook & K.D. Hyde, comb. nov. 11. Chromolaenicola thailandensis Mapook & K.D. Hyde, sp. nov. 12. Montagnula chiangraiensis Mapook & K.D. Hyde, sp. nov. 13. Montagnula chromolaenae Mapook & K.D. Hyde, sp. nov. 14. Montagnula chromolaenicola Mapook & K.D. Hyde, sp. nov. 15. Montagnula thailandica Mapook & K.D. Hyde, sp. nov. 16. Pseudopithomyces palmicola J.F. Li, Ariyaw. & K.D. Hyde, in Ariyawansa et al., Fungal Divers. 75: 41 (2015), new host record. 17. Tremateia chiangraiensis Mapook & K.D. Hyde, sp. nov. 18. Tremateia chromolaenae Mapook & K.D. Hyde, sp. nov. 19. Tremateia thailandensis Mapook & K.D. Hyde, sp. nov. Lophiostomataceae Sacc. 20. Flabellascoma minimum A. Hashim., K. Hiray. & Kaz. Tanaka, in Hashimoto et al., Stud Mycol. 90: 169 (2018), new host record. 21. Pseudocapulatispora Mapook & K.D. Hyde, gen. nov. 22. Pseudocapulatispora longiappendiculata Mapook & K.D. Hyde, sp. nov. Nigrogranaceae Jaklitsch & Voglmayr 23. Nigrograna chromolaenae Mapook & K.D. Hyde, sp. nov. Neomassarinaceae Mapook & K.D. Hyde, fam. nov. 24. Neomassarinaceae Mapook & K.D. Hyde, fam. nov. 13 Fungal Diversity 25. Neomassarina chromolaenae Mapook & K.D. Hyde, sp. nov. 26. Neomassarina thailandica Phook., Jayasiri & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 138 (2016), new host record. Pyrenochaetopsidaceae Valenzuela-Lopez et al. 47. Pyrenochaetopsis chromolaenae Mapook & K.D. Hyde, sp. nov. Roussoellaceae J.K. Liu et al. Neopyrenochaetaceae Valenzuela-Lopez et al. 27. Neopyrenochaeta chiangraiensis Mapook & K.D. Hyde, sp. nov. 28. Neopyrenochaeta chromolaenae Mapook & K.D. Hyde, sp. nov. 29. Neopyrenochaeta thailandica Mapook & K.D. Hyde, sp. nov. 30. Neopyrenochaeta triseptatispora Mapook & K.D. Hyde, sp. nov. Phaeosphaeriaceae M.E. Barr 31. Leptospora chromolaenae Mapook & K.D. Hyde, sp. nov. 32. Leptospora phraeana Mapook & K.D. Hyde, sp. nov. 33. Leptospora thailandica Phukhams. & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 100 (2016), new host record. 34. Murichromolaenicola Mapook & K.D. Hyde, gen. nov. 35. Murichromolaenicola chiangraiensis Mapook & K.D. Hyde, sp. nov. 36. Murichromolaenicola chromolaenae Mapook & K.D. Hyde, sp. nov. 37. Neoophiobolus Mapook & K.D. Hyde, gen. nov. 38. Neoophiobolus chromolaenae Mapook & K.D. Hyde, sp. nov. 39. Paraleptospora Mapook & K.D. Hyde, gen. nov. 40. Paraleptospora chromolaenae Mapook & K.D. Hyde, sp. nov. 41. Paraleptospora chromolaenicola Mapook & K.D. Hyde, sp. nov. 42. Pseudoophiosphaerella huishuiensis J.F. Zhang, J.K. Liu & Z.Y. Liu, in Zhang et al., Mycosphere 8(1): 207 (2019), new host record. 43. Pseudostaurosphaeria Mapook & K.D. Hyde, gen. nov. 44. Pseudostaurosphaeria chromolaenae Mapook & K.D. Hyde, sp. nov. 45. Pseudostaurosphaeria chromolaenicola Mapook & K.D. Hyde, sp. nov. 46. Yunnanensis chromolaenae Mapook & K.D. Hyde, sp. nov. 13 48. Pseudoroussoella Mapook & K.D. Hyde, gen. nov. 49. Pseudoroussoella chromolaenae Mapook & K.D. Hyde, sp. nov. 50. Pseudoroussoella elaeicola (Konta & K.D. Hyde) Mapook & K.D. Hyde, comb. nov. 51. Setoarthopyrenia Mapook & K.D. Hyde, gen. nov. 52. Setoarthopyrenia chromolaenae Mapook & K.D. Hyde, sp. nov. 53. Xenoroussoella Mapook & K.D. Hyde, gen. nov. 54. Xenoroussoella triseptata Mapook & K.D. Hyde, sp. nov. Thyridariaceae Q. Tian & K.D. Hyde 55. Chromolaenomyces Mapook & K.D. Hyde, gen. nov. 56. Chromolaenomyces appendiculatus Mapook & K.D. Hyde, sp. nov. 57. Pseudothyridariella Mapook & K.D. Hyde, gen. nov. 58. Pseudothyridariella chromolaenae Mapook & K.D. Hyde, sp. nov. 59. Pseudothyridariella mahakashae (Devadatha, V.V. Sarma, D.N. Wanas., K.D. Hyde & E.B.G. Jones) Mapook & K.D. Hyde, comb. nov. Torulaceae Corda 60. Torula chromolaenae Li, Phook., Mapook & K.D. Hyde, Mycol. Progr. 16(4): 454 (2017), reference specimen. 61. Torula fici Crous, IMA Fungus 6 (1): 192 (2015), reference specimen. 62. Torula polyseptata C.G. Lin & K.D. Hyde, in Hyde et al., Fungal Divers. 96: 71 (2019), new host record. Dothideomycetes orders incertae sedis Botryosphaeriales C.L. Schoch et al. Aplosporellaceae Slippers et al. 63. Aplosporella chromolaenae Mapook & K.D. Hyde, sp. nov. Fungal Diversity 64. Aplosporella hesperidica Speg., Anal. Soc. cient. argent. 13(1): 18 (1882), new host record. Botryosphaeriaceae Theiss. & P. Syd., Ann. Mycol. 16 (1–2): 16 (1918) 65. Dothiorella oblonga F.J.J. Van der Walt, Slippers & G.J. Marais, in Slippers et al., Persoonia 33: 163 (2014), new host record. 66. Sphaeropsis chromolaenicola Mapook & K.D. Hyde, sp. nov. Dyfrolomycetales K.L. Pang et al. Pleurotremataceae Walt. Watson (= Dyfrolomycetaceae K.D. Hyde et al.)* 67. Dyfrolomyces chromolaenae Mapook & K.D. Hyde, sp. nov. Muyocopronales Mapook et al. Muyocopronaceae K.D. Hyde Subclass Xylariomycetidae O.E. Erikss & Winka Amphisphaeriales D. Hawksw. & O.E. Erikss.* Apiosporaceae K.D. Hyde et al. 75. Arthrinium chromolaenae Mapook & K.D. Hyde, sp. nov. Xylariales Nannf. Cainiaceae J.C. Krug 76. Longiappendispora Mapook & K.D. Hyde, gen. nov. 77. Longiappendispora chromolaenae Mapook & K.D. Hyde, sp. nov. Materials and methods Collection, examination, and isolation of fungi 68. Muyocopron chromolaenae Mapook & K.D. Hyde, sp. nov. 69. Muyocopron chromolaenicola Mapook & K.D. Hyde, sp. nov. 70. Muyocopron lithocarpi Mapook, Boonmee & K.D. Hyde, Phytotaxa 265(3): 235 (2016), new host record Patellariales D. Hawksw. & O.E. Erikss. Patellariaceae Corda 71. Patellaria chromolaenae Mapook & K.D. Hyde, sp. nov. Class Sordariomycetes O.E. Erikss. & Winka Subclass Diaporthomycetidae Senan. et al. Diaporthales Nannf.* Diaporthaceae Höhn. ex Wehm. 72. Diaporthe chromolaenae Mapook & K.D. Hyde, sp. nov. Subclass Hypocreomycetidae O.E. Erikss. & Winka Hypocreales Lindau Hypocreaceae De Not. 73. Trichoderma guizhouense Q.R. Li, McKenzie & Yong Wang bis, in Li et al., Mycol. Progr. 12(2): 170 (2012) [2013], new host record. Stachybotryaceae Lombard & Crous 74. Memnoniella chromolaenae Mapook & K.D. Hyde, sp. nov. Fresh materials were collected from sampling sites in Thailand (Chiang Mai, Chiang Rai, Lampang, Mae Hong Son, Nan, Phetchaburi and Phrae Provinces) during 2015–2017. Dead aerial stems of Chromolaena odorata (Siam weed) were randomly collected from roadsides, abandoned fields or disturbed forests and examined using a Motic SMZ 168 Series microscope. A Carl Zeiss GmbH stereo microscope fitted with an AxioCam ERC 5S camera was used for image capture. Sections of ascomata were made by free hand. Fungal material was mounted in water, lactoglycerol, 5% KOH and/or Indian ink and photographed with a Nikon ECLIPSE 80i compound microscope fitted with a Canon EOS 550D digital camera, or a Nikon ECLIPSE Ni compound microscope fitted with a Canon EOS 600D and 750D digital camera. Measurements of microscopic characters were calculated using Tarosoft Image Frame Work program (IFW) version 0.97 and all photoplates were made using Adobe Photoshop CS6 version 13.1.2. Single spore isolations were obtained following the methods of Chomnunti et al. (2014). Spores were germinated on malt extract agar (MEA: 33.6 g/l sterile distilled water, Difco malt extract) and observed within 24 h using a Motic SMZ 168 Series microscope; germinated spores were transferred to new MEA media. Pure cultures were used for molecular studies. All specimens were kept in separate ziplock bags and envelopes with collection details and deposited in the herbaria of Mae Fah Luang University (Herb. MFLU) and Herbarium of Cryptogams, Kunming Institute 13 Fungal Diversity of Botany Academia Sinica (HKAS), China. Living cultures are also deposited in collection of Mae Fah Luang University (MFLUCC). Faces of fungi and Index Fungorum numbers were obtained as in Jayasiri et al. (2015) and Index Fungorum (2020). Preliminary screening of antimicrobial activity Preliminary screening of antimicrobial activity against different test organisms such as Bacillus subtilis (Gram-positive bacteria, DSM10), Escherichia coli (Gram-negative bacteria, DSM498), and Mucor plumbeus (filamentous fungus, MUCL 49355) were determined by the zone of inhibition using agar plug diffusion method (Balouiri et al. 2016). Penicillin and ciproloxacin were used as positive control for antibacterial screening, and nystatin was used as positive control for antifungal screening (El-Desoukey 2018). Sterile Mueller–Hinton agar media and yeast malt agar media (pH 6.3) were used for testing antibacterial activity and antifungal activity, respectively. Microbial suspensions were prepared by cell counting (6.7 × 105 cells/ml for bacteria and 5 × 104 cells/ml for filamentous fungi) and added to the sterile agar media prior to its solidification. Fungal mycelium plugs from our isolates were transferred to a solid media plate and incubated for 24 to 48 h at room temperature. Inhibition zones for antimicrobial activity was measured and compared with positive control. DNA extraction, PCR amplification and sequencing Genomic DNA was extracted using the ZR Soil Microbe DNA MiniPrep kit (Zymo Research, Irvine, CA, USA) following the manufacturer’s instructions. DNA quality and quantity were measured by spectrophotometric quantification with a NanoDrop ND-8000 V1.1.1 spectrophotometer (Thermo Fisher Scientific, Dreieich, Germany). DNA amplifications were performed by polymerase chain reaction (PCR) to amplify partial gene regions using primers shown in Table 1. PCR amplification and sequencing were carried out following Mapook et al. (2016a). The PCR products were sequenced with primers mentioned in Table 1 at SeqLab GmbH (Microsynth AG), Göttingen, Germany. GenBank based on recent publications. Forward and reverse sequences were assembled using ContigExpress (a component of Vector NTI Suite 6.0). Single gene sequences were aligned using MAFFT v.7 online program (https ://mafft .cbrc.jp/alignment/server/) (Katoh et al. 2019). The alignments were checked and uninformative gaps minimized manually where necessary in MEGA6 (Tamura et al. 2013). The align sequence datasets were combined using MEGA5 and MEGA6 (Tamura et al. 2011, 2013). The combined alignments in FASTA fomat were transform to PHYLIP formats by ALTER online program (http://www.sing-group .org/ALTER/) (Glez-Peña et al. 2010). Phylogenetic analysis of the combined aligned dataset was analyze based on maximum likelihood (ML) and Bayesian inference (BI). Maximum likelihood analysis was performed in RAxML (Stamatakis et al. 2008) implemented in raxmlGUI v.0.9b2 (Silvestro and Michalak 2010) with 1000 rapid bootstrap replicates using the GTR + GAMMA model of nucleotide substitution. Maximum likelihood bootstrap values (MLBP) equal or greater than 60% are given above each node. Posterior probabilities (PP) (Rannala and Yang 1996; Zhaxybayeva and Gogarten 2002) were determined by Markov chain Monte Carlo sampling (MCMC) in MrBayes v.3.0b4 (Huelsenbeck and Ronquist 2001). MrModeltest 2.2 (Nylander 2004) was used to perform the model of nucleotide substitution for each gene. Six simultaneous Markov chains were run for 5,000,000 generations or depending on individual settings for the fungal group, and trees were sampled every 100th generation. Tracer v1.6 (Rambaut et al. 2014) was used to examine the log-likelihood scores to decide extra runs and determine the stationary phase. The first 20% of trees were discarded as the burn-in phase. The remaining trees were used for calculating PP in the majority rule consensus tree. The run was stopped when the standard deviation of split frequencies was reached below 0.01 (Ariyawansa et al. 2013; Maharachchikumbura et al. 2015; Udayanga et al. 2015). Bayesian posterior probabilities (PP) equal to or greater than 0.95 are given above the nodes. Phylogenetic trees were drawn using FigTree 1.4.2 (Rambaut 2009) and edited by Microsoft Office PowerPoint 2013 and Adobe Illustrator CS6 (Adobe Systems, USA). Taxonomy Phylogenetic analysis The most closely related taxa for analyses were selected following nucleotide BLAST searches in GenBank (http:// www.ncbi.nlm.nih.gov/) as well as sequence data from 13 Phylum Ascomycota Caval.-Sm. Class Dothideomycetes sensu O.E. Erikss. & Winka Subclass Pleosporomycetidae C.L. Schoch et al. Fungal Diversity Table 1 Details of genes/loci with PCR primers and protocols Family or genus Class Dothideomycetes Hysteriaceae Didymosphaeriaceae Nigrogranaceae Phaeosphaeriaceae Acrocalymmaceae Lophiostomataceae Neomassarinaceae Roussoellaceae, Thyridariaceae Torulaceae, Muyocopronaceae, Patellariaceae Didymellaceae Neopyrenochaetaceae Pyrenochaetopsidaceae Aplosporellaceae Botryosphaeriaceae Pleurotremataceae Class Sordariomycetes Diaporthaceae Hypocreaceae Stachybotryaceae Apiosporaceae Cainiaceae Hysteriales Lindau Hysteriales was introduced by Lindau (1897b) with a monotypic family, Hysteriaceae. In this study, we follow the latest treatment and updated accounts of Hysteriales in Jayasiri et al. (2018) and Wijayawardene et al. (2018), with updated accounts of the genus Rhytidhysteron from recent relevant literature (Jayasiri et al. 2018; Dayarathne et al. 2020; Kumar et al. 2019). Gene/loci PCR primers (forward/reverse) References for primer LSU SSU ITS TEF1 LSU SSU ITS TEF1 RPB2 LROR/LR5 NS1/NS4 ITS5/ITS4 983F/2218R LROR/LR5 NS1/NS4 ITS5/ITS4 983F/2218R fRPB2-5f/fRPB2–7cR Vilgalys and Hester (1990) White et al. (1990) White et al. (1990) Rehner (2001) Vilgalys and Hester (1990) White et al. (1990) White et al. (1990) Rehner (2001) Liu et al. (1999) LSU ITS TEF1 RPB2 LSU ITS RPB2 LSU ITS TEF1 ITS TEF1 LSU SSU TEF1 LROR/LR5 ITS5/ITS4 983F/2218R fRPB2-5f/fRPB2–7cR LROR/LR5 ITS5/ITS4 fRPB2-5f/fRPB2–7cR LROR/LR5 ITS5/ITS4 983F/2218R ITS5/ITS4 983F/2218R LROR/LR5 NS1/NS4 983F/2218R Vilgalys and Hester (1990) White et al. (1990) Rehner (2001) Liu et al. (1999) Vilgalys and Hester (1990) White et al. (1990) Liu et al. (1999) Vilgalys and Hester (1990) White et al. (1990) Rehner (2001) White et al. (1990) Rehner (2001) Vilgalys and Hester (1990) White et al. (1990) Rehner (2001) ITS TEF1 RPB2 TEF1 LSU ITS TEF1 RPB2 LSU ITS TEF1 LSU ITS ITS5/ITS4 983F/2218R fRPB2-5f/fRPB2–7cR 983F/2218R LROR/LR5 ITS5/ITS4 983F/2218R fRPB2-5f/fRPB2–7cR LROR/LR5 ITS5/ITS4 983F/2218R LROR/LR5 ITS5/ITS4 White et al. (1990) Rehner (2001) Liu et al. (1999) Rehner (2001) Vilgalys and Hester (1990) White et al. (1990) Rehner (2001) Liu et al. (1999) Vilgalys and Hester (1990) White et al. (1990) Rehner (2001) Vilgalys and Hester (1990) White et al. (1990) Hysteriaceae Chevall. Hysteriaceae was introduced by Chevallier (1826) with Hysterium as the type genus. The family has been classified in several orders such as Pseudosphaeriales (Nannfeldt 1932; Gäumann 1949), Dothiorales (Müller and von Arx 1950; von Arx and Müller 1954), Dothideales (von Arx and Müller 1975) and presently, Hysteriales, which is closely related to the Pleosporales (Luttrell 1955; Kirk et al. 2008; Thambugala et al. 2016; Jayasiri et al. 2018; Wijayawardene 13 Fungal Diversity et al. 2018). Jayasiri et al. (2018) accepted 13 genera, which includes nine genera based on both morphology and phylogenetic analyses (Gloniopsis, Graphyllium, Hysterium, Hysterobrevium, Hysterodifractum, Oedohysterium, Ostreichnion, Psiloglonium, Rhytidhysteron), and four genera (Actidiographium, Gloniella, Hysterocarina, Hysteropycnis) based on morphology only. The divergence time estimates for this family are crown age of 149 Mya (90–213) in the late Jurassic and stem age of 219 Mya (161–282) in the late Triassic (Liu et al. 2017). Rhytidhysteron Speg. Rhytidhysteron was introduced by Spegazzini (1881a) to accommodate two species, R. brasiliense and R. viride. Subsequently, R. brasiliense was designated as the type species (Clements and Shear 1931). There are 21 epithets listed in Index Fungorum (2020). The genus has been classified within the family Patellariaceae (Kutorga and Hawksworth 1997; Eriksson 2006; Lumbsch and Huhndorf 2010a, b). However, recent studies based on multi-gene analyses have shown that the genus should be placed in Hysteriaceae Fig. 1 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, SSU, ITS and TEF1 sequence data. Thirty strains are included in the combined sequence analysis, which comprise 3568 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 9583.200312 is presented. The matrix had 646 distinct alignment patterns, with 33.57% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.245845, C = 0.235270, G = 0.278609, T = 0.240276; substitution rates: AC = 1.066264, AG = 2.632538, AT = 1.024795, CG = 0.715396, CT = 6.041010, GT = 1.000000; gamma distribution shape parameter α = 0.079339. 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 dark red bold and type species are in bold. Mytilinidion mytilinellum (EB 0386, CBS 303.34) are used as outgroup taxa 13 (Boehm et al. 2009a, b; de Almeida et al. 2014; Wijayawardene et al. 2014a; Jayasiri et al. 2018; Wijayawardene et al. 2018; Kumar et al. 2019) with seven species having available sequences in GenBank (R. hysterinum, R. mangrovei, R. neorufulum, R. opuntiae, R. rufulum, R. thailandicum, R. tectonae). Most species were found in Thailand (Thambugala et al. 2016; Doilom et al. 2017; Kumar et al. 2019) and isolated from diverse habitats, such as saprobic on decaying woody branches, stems or twigs in terrestrial habitats, as well as marine habitats from mangrove wood (Kumar et al. 2019). The genus has not been previously reported on C. odorata nor on the host family Asteraceae. Dayarathne et al. (2020) introduced a new species, R. bruguierae from submerged branches of Bruguiera sp. in Thailand. In this study, we introduce a new species, R. chromolaenae with a new host record for R. bruguierae, based on morphology and molecular data, together with descriptions and illustrations (Figs. 2, 3, 4). A phylogenetic tree based on combined LSU, SSU, ITS and TEF1 sequence data is presented in Fig. 1. Fungal Diversity Fig. 2 Rhytidhysteron bruguierae (new host record). a, b Appearance of ascomata on substrate. c Section through ascoma. d Exciple. e Pseudoparaphyses. f–i Asci. j–o Ascospores. Scale bars: a = 500 µm, b, c = 200 µm, d, f–i = 50 µm, e, j–o = 10 µm Rhytidhysteron bruguierae Dayarathne, Mycosphere 11(1): 20 (2020) Facesoffungi number: FoF 06154; Fig. 2 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 250–500 µm high, 500–1000 µm diam. ( x̄ = 360 × 720 µm, n = 10), hysterothecial, boat-shaped, superficial, with a carbonaceous rim, scattered, closed at first and opening at maturity, elongate, elliptic, dark brown to black with dark orange at the center, when dry folded at the margin. Exciple 45–60(–110) µm wide, comprising hyaline or pale brown to brown cells arranged in textura globulosa to textura angularis, continuous to the base (hypothecium). Hamathecium comprising 1.5–2.5 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses, 13 Fungal Diversity Fig. 3 Rhytidhysteron chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Exciple. e–g Pseudoparaphyses. h–k Asci. l–q Ascospores. Scale bars: a = 500 µm, b = 200 µm, c, h–k = 100 µm, d = 20 µm, e–g, l–q = 10 µm 13 Fungal Diversity Fig. 4 Culture characteristics on MEA: a Rhytidhysteron bruguierae (MFLUCC 17-1515). b Rhytidhysteron chromolaenae (MFLUCC 17-1516) slightly swollen and rounded at the apex, forming pale red to pinkish brown epithecium above the asci when mounted in lactoglycerol and slightly orange to yellowish orange epithecium above the asci when mounted in water. Asci (120–)130–140 × 10–15 µm ( x̄ = 130 × 11 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindrical, straight or slightly curved, with a short pedicel, apically rounded with an ocular chamber. Ascospores 18–22 × 7–9 µm ( x̄ = 20 × 8 µm, n = 30), overlapping, uniseriate, pale brown to brown, broadly fusiform, with upper part or second cell slightly wider, 1-septate when immature, becoming 3-euseptate when mature, slightly constricted at the central septum, straight or slightly curved, guttulate, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from second cell from apex. Colonies on MEA filamentous, mycelium velvety, fluffy, undulate to filamentous, white aerial hyphae at the surface and in reverse (Fig. 4a). Pre-screening for antimicrobial activity: Rhytidhysteron bruguierae (MFLUCC 17-1515) showed antimicrobial activity against Mucor plumbeus with an 18.5 mm inhibition zone, observable as partial inhibition, when compared with the positive control (17 mm), but no inhibition of Bacillus subtilis and Escherichia coli. Known hosts and distribution: On submerged branches of Bruguiera sp. (Rhizophoraceae) in Thailand (Dayarathne et al. 2020). Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 7 April 2017, A. Mapook (DP106, MFLU 20-0363); living culture MFLUCC 17-1515 (new host record); Phrae Province, Doi Pha Klong, on dead stems of C. odorata, 22 September 2016, A. Mapook (DPKP5, MFLU 20-0364); living culture MFLUCC 17-1511; Chiang Mai Province, Fah Hom Pok, on dead stems of C. odorata, 27 September 2016, A. Mapook (FHP4, MFLU 20-0365); living culture MFLUCC 17-1502; Mae Hong Son Province, Mae Yen, Pai, on dead stems of C. odorata, 25 June 2016, A. Mapook (MY8, MFLU 20-0366); living culture MFLUCC 17-1509. GenBank numbers: LSU: MN632455, MN632454, MN632453, MN632452, ITS: MN632460, MN632459, MN632458, MN632457, SSU: MN632466, MN632465, MN632464, MN632463, TEF1: MN635662, MN635661 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of our four strains with 100% similarity was Rhytidhysteron rufulum (strain 510A, EU020063), while the closest match with the LSU, SSU and TEF1 sequences with 99.34%, 99.78% and 96.67% similarity, respectively, was R. thailandicum (strain MFLUCC 14-0503). In the present phylogenetic analysis, MFLUCC 17-1515, MFLUCC 17-1511, MFLUCC 17-1502 and MFLUCC 17-1509 cluster with R. bruguierae (MFLUCC 18-0398) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 1). We therefore, identify our four strains as R. bruguierae based on phylogenetic analyses with morphological comparison (Table 2) and the four isolates are introduced here as a new host record from Chromolaena odorata collected in Thailand. Rhytidhysteron bruguierae is also closely related to R. thailandicum (Fig. 1). However, R. bruguierae differs from R. thailandicum in having slightly smaller asci [(120–)130–140 × 10–15 µm vs. 135–160 × 10.5–15 µm] and pale brown to brown, smaller ascospores (18–22 × 7–9 µm vs. 20–28(–31) × 7.5–12 µm) (Table 2). Rhytidhysteron chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557360, Facesoffungi number: FoF 07780; Fig. 3 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0367 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 275–290(–380) µm high × (685–)750–885 µm diam. ( x̄ = 305 × 780 µm, n = 5), hysterothecial, boat-shaped, superficial, with a carbonaceous rim, not perpendicularly striate, scattered, closed at first and opening at maturity, elongate, elliptic, dark brown to black with yellowish green on the margin, orange or dark brown to black at the center. Exciple (50–)60–85 µm wide, comprising pale brown or brown to dark brown cells, arranged in textura globulosa to textura angularis, continuous to the base (hypothecium). Hamathecium comprising (1.5–)2–3 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses, slightly swollen and rounded at the apex, forming pale orange to yellowish orange epithecium above the asci when mounted in water, orange to red epithecium above the asci when mounted in lactoglycerol and becoming slightly purple epithecium above the asci when mounted in 5% KOH. Asci 130–155 × (8–)11–14 µm ( x̄ = 145 × 12 µm, n = 10), 7–8-spored, bitunicate, fissitunicate, cylindrical, 13 Fungal Diversity straight or slightly curved, apically rounded. Ascospores 23–28 × 8–11 µm ( x̄ = 26.5 × 9 µm, n = 25), overlapping, uniseriate, broadly fusiform, widest at the center, hyaline to yellowish brown, 1-septate when immature, becoming brown to dark brown, 3-septate when mature, slightly constricted at the central septum, straight or slightly curved, guttulate, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from second cell from apex. Colonies on MEA irregular with slightly raised, undulate, white aerial hyphae spreading from the center of the colony, become smoke-gray to olivaceous brown at the surface and olivaceous brown with creamy white at the margin in reverse (Fig. 4b). Pre-screening for antimicrobial activity: Rhytidhysteron chromolaenae (MFLUCC 17-1516) showed antimicrobial activity against M. plumbeus with a 20 mm inhibition zone, observable as partial inhibition, when compared with the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 7 April 2017, A. Mapook (DP107, MFLU 20-0367, holotype); extype culture MFLUCC 17-1516. GenBank numbers: LSU: MN632456, ITS: MN632461, SSU: MN632467, TEF1: MN635663 Notes: Rhytidhysteron chromolaenae is similar to R. columbiense (CUVC 62421) in having ascomata with yellowish green margins and (1–)3-septate ascospores, but differs in having smaller ascomata (274–290(–380 ) × (685–)750–885 µm vs. 600–700 × 1200–1800 µm); smooth, not perpendicularly striate at margin, smaller asci (130–155 × (8–)11–14 µm vs. 175–190 × 14–18 µm) and smaller ascospores (23–28 × 8–11 µm vs. 38–52 × 13–18 µm) (Table 2). Rhytidhysteron columbiense was described from Colombia without molecular data (Soto-Medina and Lücking 2017). In the present phylogenetic analysis, R. chromolaenae forms a well-separated branch from other Rhytidhysteron species and is closely related to R. mangrovei, R. thailandicum and R. bruguierae (Fig. 1). In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence for MFLUCC 17-1516 is R. rufulum with 99.81% similarity to the strain B1a081-2-P30 (JQ388942), while the closest match with the LSU sequence was with R. neorufulum (strain MFLUCC 13-0216, NG_059649) with 98.90% similarity and closest matches with the TEF1 sequence were R. thailandicum (strain MFLUCC 14-0503, KU497490) with 95.90% similarity. Therefore, R. chromolaenae is described 13 as a new species based on phylogeny and morphological characters. Pleosporales Luttrell ex M.E. Barr Pleosporales was introduced by Luttrell (1955). The order is highly diverse with more than 75 families (Wijayawardene et al. 2018). We follow the latest treatment and updated accounts of Pleosporales in Liu et al. (2017) and Wijayawardene et al. (2018), together with recent relevant literature for updated accounts of each family. Acrocalymmaceae Crous & Trakun. Acrocalymmaceae was introduced by Trakunyingcharoen et al. (2014b) to accommodate the genus Acrocalymma. The divergence time estimates for this family are crown age of 23 Mya (8–44) in the Neogene period and stem age of 114 Mya (71–156) in the Cretaceous (Liu et al. 2017). Acrocalymma Alcorn & J.A.G. Irwin, Trans. Br. mycol. Soc. 88(2): 163 (1987) Acrocalymma is an ecologically diverse genus, containing plant pathogens, endophytes, and saprobes from terrestrial or freshwater as well as human superficial tissue (Alcorn and Irwin 1987; Zhang et al. 2012a, b; Trakunyingcharoen et al. 2014b; Valenzuela-Lopez et al. 2017; Jin et al. 2018; Jayasiri et al. 2019). The genus was introduced by Alcorn and Irwin (1987) with A. medicaginis as the type species. Jayasiri et al. (2019) introduced a new species, A. pterocarpi from a fallen pod of Pterocarpus indicus in Thailand. Seven epithets are listed in Index Fungorum (2020). In this study, the first record of Acrocalymma on Chromolaena odorata is introduced, based on morphology and molecular data with a description and illustrations (Fig. 6). A phylogenetic tree based on combined LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 5. Acrocalymma medicaginis Alcorn & J.A.G. Irwin, Trans. Br. mycol. Soc. 88(2): 163 (1987) Facesoffungi number: FoF 07099; Fig. 6 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 150–250 µm high × (115–)140–155 µm diam. ( x̄ = 195 × 140 µm, n = 5), immersed, appearing as black spots, coriaceous, solitary or scattered, subglobose to obpyriform, brown to dark brown. Ostiole long neck, papillate. Peridium 10–20 µm wide, 4–5 layers, comprising pale brown to brown cells of textura angularis. Hamathecium comprising 0.5–1.5 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 70–85 × 8–11 µm ( x̄ = 78 × 9 µm, n = 10), 8-spored, bitunicate, fissitunicate, Fungal Diversity Fig. 5 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU, TEF1 and RPB2 sequence data. Sixteen strains are included in the combined sequence analysis, which comprise 4699 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 11536.335405 is presented. The matrix had 503 distinct alignment patterns, with 54.55% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.242971, C = 0.232109, G = 0.270331, T = 0.254589; substitution rates: AC = 1.527101, AG = 2.317230, AT = 1.662072, CG = 1.025372, CT = 5.286531, GT = 1.000000; gamma distribution shape parameter α = 0.208825. 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 dark red bold and type species are in bold. Massarina eburnea (CBS 473.64) is used as outgroup taxon Table 2 Synopsis of Rhytidhysteron species with similar morphological features discussed in this study Species Asocomata (µm) Exciple (µm) R. columbiense 600–700 high × 1200– 60–90, textura angu(CUVC 62421) (1590)–1800 diam. laris R. bruguierae 410–520 high × 548– 148–162, textura (MFLUCC 18-0398) 570 diam. angularis R. bruguierae 250–500 high × 500– 45–60(–110), textura (MFLUCC 17-1515) 1000 diam. globulosa to textura angularis (50–)60–85, textura R. chromolaenae 274–290(–380) globulosa (MFLUCC 17-1516) high × (685–)750– 885 diam. 360–640 high × 53072–130, textura R. thailandicum 750 diam. angularis (MFLUCC 14-0503) 135–160 × 10.5–15 cylindric-clavate, straight or slightly curved, with a short pedicel, apically rounded, with an ocular chamber. Ascospores 15–21 × 3–5 µm (x̄ = 18 × 4 µm, n = 20), overlapping, 1–2-seriate, hyaline, oblong to broadly fusiform with Asci (µm) Ascospores (µm) References 175–190 × 14–18 38–(43.5)–52 × 13– (14.7)–18 14–26 × 6.2–9 (120–)130– 140 × 10–15 18–22 × 7–9 Soto-Medina and Lücking (2017) Dayarathne et al. (2020) This study 130–155 × (8–)11–14 23–28 × 8–11 This study 20–28(–31) × 7.5–12 Thambugala et al. (2016) 128–148 × 10–14 obtuse to slightly obtuse ends, uniseptate, straight or slightly curved, smooth, constricted at the septum, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. 13 Fungal Diversity Fig. 6 Acrocalymma medicaginis (new host record) a, b Appearance of ascomata on substrate. c Section through ascoma. d Peridium. e Pseudoparaphyses. f–i Asci. j–n Ascospores. o Ascospores surrounded by hyaline gelatinous sheath in Indian ink. Scale bars: a, b = 500 µm, c = 50 µm, d–i, o = 10 µm, j–n = 5 µm 13 Fungal Diversity Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature, germ tubes produced from both cells. Colonies on MEA irregular, mycelium slightly raised, moderately fluffy, filiform, white aerial hyphae at the surface, spreading from the center and dark brown to olivaceous-brown in reverse from the center with creamy white at rim (Fig. 7). Pre-screening for antimicrobial activity: Acrocalymma medicaginis (MFLUCC 17-1439) showed antimicrobial activity against B. subtilis with a 12 mm inhibition zone and against M. plumbeus with a 13 mm inhibition zone, observable as partial inhibition, when compared with the positive control (25 mm and 17 mm, respectively), but no inhibition of E. coli. Known hosts and distribution: Root and crown rotting of Medicago sativa (Fabaceae) in Australia (Alcorn and Irwin 1987; Trakunyingcharoen et al. 2014a, b) Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 8 July 2015, A. Mapook (DP5, MFLU 20-0296); living culture MFLUCC 17-1423 (new host record); 5 August 2015, A. Mapook (DP23, MFLU 20-0297); living culture MFLUCC 17-1439. GenBank numbers: LSU: MT214432, MT214433, ITS: MT214338, MT214339, SSU: MT214387, MT214388, TEF1: MT235797, MT235798 Notes: Multigene phylogenetic analyses (Fig. 5) show that two strains MFLUCC 17-1423, MFLUCC 17-1439 grouped within the Acrocalymma medicaginis clade. A BLASTn search of the LSU sequence data showed that the two strains are identical to A. medicaginis (strain CPC 24345, KP170718) with 100% similarity while the closest match with the ITS sequence of our two strains are A. medicaginis with 99.59% similarity to the strain CPC 24342 (KP170622). Fig. 7 Culture characteristic on MEA: Acrocalymma medicaginis (MFLUCC 17-1423) Therefore, we identify our isolates as A. medicaginis based on phylogenetic analyses. Only two species, A. walkeri and A. pterocarpi have been reported with a sexual morph, the other species being coelomycetous. In this study, we isolated the sexual morph of A. medicaginis from Chromolaena odorata collected in Thailand for the first time, and the isolates are also introduced here as a new host record. We did not obtain the asexual morph of our isolates in culture. Didymellaceae Gruyter et al. Didymellaceae contains numerous plant pathogenic, saprobic and endophytic species associated with a wide range of hosts, as well as some species of clinical or environmental origin. The family was introduced by De Gruyter et al. (2009) to accommodate Phoma and phoma-like genera, with Didymella as the type genus. Hyde et al. (2013) accepted 13 genera in the family based on both morphology and phylogenetic analyses (Ascochyta, Boeremia, Chaetasbolisia, Dactuliochaeta, Didymella, Epicoccum, Leptosphaerulina, Macroventuria, Microsphaeropsis, Monascostroma, Phoma, Piggotia, Pithomyces). Chen et al. (2015) introduced nine additional genera to the family (Allophoma, Calophoma, Heterophoma, Neoascochyta, Neodidymelliopsis, Nothophoma, Paraboeremia, Phomatodes, Xenodidymella) based on morphological observations with multi-locus phylogenetic analyses of ITS, LSU, RPB2 and TUB2 sequence data. In addition, Didymellocamarosporium, Didysimulans, Endocoryneum, Mixtura, Peyronellaea, Phaeomycocentrospora, Platychora, Pseudohendersonia and Stagonosporopsis were accepted in Outline of Ascomycota 2017 (Wijayawardene et al. 2018). Recently, Valenzuela-Lopez et al. (2018) introduced six new genera (Cumuliphoma, Ectophoma, Juxtiphoma, Remotididymella, Similiphoma, Vacuiphoma) of clinical or environmental origin and synonymized Peyronellaea under the genus Didymella. Divergence time estimates for this family are crown age of 63 Mya (35–97) in the Paleogene period and stem age of 115 Mya (84–149) in the Cretaceous (Liu et al. 2017). Soleimani et al. (2018) reported divergence time estimates with a focus on Didymellaceae and suggested that the family diverged from Aigialaceae at 86.7 Mya (53.9–155.4) in the Cretaceous and initial divergence of the family happened in the late Eocene. Didymella Sacc. ex D. Sacc. Didymella was introduced by Saccardo (1880) with D. exigua as the type species and contains some species which are important serious plant pathogens, also endophytes and saprobes on a wide range of substrates, as well as species of clinical or environmental origin (Aveskamp et al. 2010; Chen et al. 2015, 2017; Ahmadpour et al. 2017; Jayasiri et al. 2017; Valenzuela-Lopez et al. 2018). The genus was presented as monophyletic by Chen et al. (2015) and comprises 35 known species, with two unidentified species, 13 Fungal Diversity Fig. 8 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, TUB2 and RPB2 sequence data. Sixtytwo strains are included in the combined sequence analysis, which comprise 3028 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 14613.624205 is presented. The matrix had 624 distinct alignment patterns, with 13.98% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.240327, C = 0.243303, G = 0.275771, T = 0.240599; substitution rates: AC = 0.969862, AG = 4.713232, AT = 1.514683, CG = 0.711109, CT = 10.328441, GT = 1.000000; gamma distribution shape parameter α = 0.102562. 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 dark red bold and type species are in bold. Epicoccum poae (LC 8160) and Epicoccum nigrum (CBS 173.73) are used as outgroup taxa based on morphological observations and multi-locus phylogenetic analyses of ITS, LSU, RPB2 and TUB2 sequence data. Subsequently, Chen et al. (2017) introduced eleven new Didymella species while studying the distribution and biodiversity of Didymellaceae. Valenzuela-Lopez et al. (2018) described two new species D. brunneospora and D. keratinophila from flower-stalk and human superficial tissue, respectively. Jayasiri et al. (2019) introduced a new species, D. magnolia (MFLUCC 18-1560) collected from Magnolia grandiflora cone in China. In this study, D. chromolaenae 13 is introduced as a new species based on morphology and molecular data (Fig. 9). A phylogenetic tree based on combined LSU, ITS, TUB2 and RPB2 sequence data is presented in Fig. 8. Didymella chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557295, Facesoffungi number: FoF 07781; Fig. 9 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0309 Fungal Diversity Fig. 9 Didymella chromolaenae (holotype) a Appearance of ascomata on substrate. b Section through ascoma. c Peridium. d–f Immature and mature asci. g–k Ascospores. Scale bars: a = 500 µm, b, d–f = 50 µm, c = 20 µm, g–k = 10 µm Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 130–145 µm high × 145–160 µm diam. ( x̄ = 140 × 152 µm, n = 5), semi-immersed to superficial, appearing as small black spots, coriaceous, solitary or scattered, globose to irregular shape, brown to dark brown. Ostiole central. Peridium 20–30 µm wide, 2–5 layers of thin-walled, pale brown to hyaline cells of textura angularis. Hamathecium without pseudoparaphyses. Asci 80–85 × 25–35 µm ( x̄ = 83 × 29 µm, n = 5), 8-spored, bitunicate, cylindric-clavate to clavate, with a short pedicel, slightly curved, apically rounded with an ocular chamber. Ascospores 15–22 × 9–13 µm ( x̄ = 19 × 11 µm, n = 20), overlapping, irregularly arranged, yellow, oval to obovoid, with obtuse ends, widest at the upper cell and tapering towards the narrow-rounded ends, uniseptate, with small guttules, constricted at the septum, surrounded with mucilaginous sheath. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium flat, filiform, white at first, becoming gray to light grayish brown hyphae spreading at the surface, yellow white with grayish brown in reverse (Fig. 10). Pre-screening for antimicrobial activity: Didymella chromolaenae (MFLUCC 17-1459) showed no inhibition against E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Lampang Province, Chaehom, on dead stems of Chromolaena odorata, 24 September 2016, A. Mapook (JH5, MFLU 20-0309, holotype); ex-type culture MFLUCC 17-1459. GenBank numbers: LSU: MT214457, ITS: MT214363, SSU: MT214409, TEF1: MT235799 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Didymella chromolaenae (MFLUCC 17-1459, ex-holotype) is Didymella glomerata with 98.92% similarity to the strain FK2 (KC802087), while the closest match of the LSU sequence was identical with 99.78% to D. omnivirens (strain CBS 341.86, MH873653). Based on the present phylogenetic analysis (Fig. 8), D. chromolaenae groups in the clade comprising D. viburnicola (CBS 523.73), D. pteridis (CBS 379.96), D. ellipsoidea (CGMCC 3.18350) and D. macrostoma (CBS 223.96) but 13 Fungal Diversity Fig. 10 Culture characteristic on MEA: Didymella chromolaenae (MFLUCC 17-1459) low bootstrap support. However, based on morphological comparison, D. chromolaenae (MFLUCC 17-1459) resembles species of Didymella based on its immersed to semiimmersed ascomata with one-celled ascospores and distinct from D. macrostoma in having oval to obovoid, yellow ascospores surrounded with a mucilaginous sheath, while D. macrostoma has ellipsoid to obovoid, hyaline ascospores. A comparison of the ITS (+ 5.8S) gene region of D. chromolaenae with D. ellipsoidea and D. macrostoma reveals 9 base pair differences (1.85%) across 486 nucleotides. Therefore, D. chromolaenae is described as a new species. Nothophoma Q. Chen & L. Cai Nothophoma was introduced by Chen et al. (2015) with N. infossa as the type species, and four new combinations (N. anigozanthi, N. arachidis-hypogaeae, N. gossypiicola, N. quercina) based on morphology and phylogeny. Crous et al. (2016) introduced a new species, N. macrospora isolated from human respiratory tract. Abdel-Wahab (2017) introduced a new endophytic species, N. multilocularis isolated from the medicinal plant Rhazya stricta, and reported on its antimicrobial activity. Crous et al. (2017) introduced N. raii, isolated from soil and Valenzuela-Lopez et al. (2018) introduced N. variabilis which was also isolated from human respiratory tract. There are nine epithets listed in Index Fungorum (2020). Nothophoma chromolaenae is introduced as a new species based on morphology and molecular data, together with description and illustrations (Fig. 12). A phylogenetic tree based on combined LSU, ITS, TUB2 and RPB2 sequence data is presented in Fig. 11. 13 Nothophoma chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557346, Facesoffungi number: FoF 07782; Fig. 12 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0342 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 55–85 µm high × 85–100(–130) µm diam. (x̄ = 73 × 98 µm, n = 10), immersed to semi-immersed, appearing as small black spots, coriaceous, solitary or scattered, globose, dark brown to reddish brown. Ostiole central, papillate. Peridium (5–)10–20(–30) µm wide, 2–3 layers of pale brown to brown cells of textura angularis. Hamathecium without pseudoparaphyses. Asci 45–70 × 13–16 µm ( x̄ = 55 × 14.5 µm, n = 15), 8-spored, bitunicate, cylindricclavate, straight or slightly curved, with a short pedicel, apically round with an ocular chamber. Ascospores 7–17 × 4–6.5 µm ( x̄ = 10 × 5 µm, n = 15), overlapping, irregularly arranged, hyaline, broadly fusiform, aseptate, widest at the center and tapering towards narrow rounded ends, straight or slightly curved, rough surface with terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, filiform, dark brown aerial hyphae at the surface and dark brown in reverse (Fig. 13). Pre-screening for antimicrobial activity: Nothophoma chromolaenae (MFLUCC 17-1443) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Phetchaburi Province, Cha-am, Khao Nang Panthurat, on dead stems of Chromolaena odorata, 28 July 2015, A. Mapook (NPR6, MFLU 20-0342, holotype); ex-type culture MFLUCC 17-1443. GenBank numbers: LSU: MT214458, ITS: MT214364, SSU: MT214410 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Nothophoma chromolaenae (MFLUCC 17-1443, ex-holotype) is N. gossypiicola with 99.79% similarity to the strain UTHSC:DI16-294 (LT592943), while the closest match of the LSU sequence were identical with 98.72% to Macrophoma lageniformis (strain CBS 364.65, MH870253) and Nothophoma macrospora (strain UTHSC:DI16-294, LN907437). In the present phylogenetic analysis, N. chromolaenae groups in the clade comprising, N. gossypiicola, N. macrospora, N. multilocularis and N. raii (Fig. 11). Those species have been recorded as asexual morphs. We compared those asexual morphs (Table 3) and found that they are morphologically distinct. Nothophoma gossypiicola differs Fungal Diversity Fig. 11 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, TEF1 and RPB2 sequence data. Eleven strains are included in the combined sequence analysis, which comprise 2757 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 5700.668844 is presented. The matrix had 226 distinct alignment patterns, with 17.08% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.243238, C = 0.237511, G = 0.277502, T = 0.241749; substitution rates: AC = 0.970562, AG = 4.188061, AT = 1.216550, CG = 0.588503, CT = 14.649953, GT = 1.000000; gamma distribution shape parameter α = 0.020000. 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 dark red bold and type species are in bold. Phoma herbarum (CBS 615.75) is used as outgroup taxon from N. multilocularis in having smaller conidiomata (100–250 µm vs. 175–1500 µm), smaller conidiogenous cells (5–8 × 5–8 µm vs. 11–17 × 9–18 µm) and slightly smaller conidia (10–12.5 × 2.5–3.5 µm vs. 9–20 × 3–4(–5) µm), while N. multilocularis, N. gossypiicola and N. macrospora differs from N. raii in having wider conidia. However, our strain, N. chromolaenae was found as the sexual morph in nature and we could not obtain its asexual morph in culture. A comparison of the ITS (+5.8S) gene region of N. chromolaenae and N. raii reveals seven base pair differences (1.6%) across 434 nucleotides. Therefore, N. chromolaenae is described here as a new species based on morphology and phylogeny (Table 3). Didymosphaeriaceae Munk, Dansk bot. Ark. 15(no. 2): 128 (1953) Didymosphaeriaceae was introduced by Munk (1953) with Didymosphaeria as type genus. Ariyawansa et al. (2014b) provided an updated account of the family and accepted sixteen genera (Alloconiothyrium, Barria, Bimuria, Deniquelata, Didymocrea, Didymosphaeria, Julella, Kalmusia, Karstenula, Letendraea, Montagnula, Neokalmusia, Paraconiothyrium, Paraphaeosphaeria, Phaeodothis, Tremateia). They clarified the taxonomic relationships within the families Didymosphaeriaceae and Montagnulaceae, and synonymized Montagnulaceae under Didymosphaeriaceae based on the oldest name. Wijayawardene et al. (2014a, b) introduced two new asexual genera Paracamarosporium and Pseudocamarosporium. Subsequently, Crous et al. (2015b, c) introduced two new genera Verrucoconiothyrium and Xenocamarosporium. Ariyawansa et al. (2015) referred Austropleospora and Pseudopithomyces to the family. Wanasinghe et al. (2016) introduced Laburnicola and Paramassariosphaeria within the family and synonymized two Munkovalsaria species under the genus Montagnula. Thambugala et al. (2017b) introduced a new genus Kalmusibambusa which was collected from living culms of bamboo in Thailand. Jayasiri et al. (2019) introduced a new genus Cylindroaseptospora and Phookamsak et al. (2019) also added a new genus Vicosamyces in the family. Divergence 13 Fungal Diversity Fig. 12 Nothophoma chromolaenae (holotype) a Appearance of ascomata on substrate. b Section through ascoma. c Peridium. d–f Immature and mature asci. g–l Ascospores. Scale bars: a = 400 µm, b = 50 µm, c, d–f = 20 µm, g–l = 10 µm time estimates for this family are crown age of 72 Mya (48–101) and stem age of 109 Mya (83–139) during the Cretaceous (Liu et al. 2017). Chromolaenicola Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557279, Facesoffungi number: FoF 07783 Etymology: Named after the host genus, Chromolaena. Saprobic on dead stems. Sexual morph: Ascomata immersed to semi-immersed, solitary or scattered, appearing as small dark spots, coriaceous, globose to subglobose, brown to dark brown. Ostiolar neck protruding. Peridium 3–4 layers, comprising dark brown cells of textura angularis. Hamathecium composed of cylindrical to filiform, septate, branching pseudoparaphyses. Asci 6–8-spored, bitunicate, cylindrical, straight or slightly curved, apically rounded, pedicellate with an ocular chamber. Ascospores slightly overlapping, uni-seriate, initially hyaline, 1-septate when immature, becoming reddish brown to brown at maturity, ellipsoid to broadly fusiform, muriform, 3 transverse septate, and 1 vertical septum, constricted at the central septum, straight or slightly curved, without gelatinous sheath. Asexual morph: Conidiomata pycnidial, solitary, immersed 13 Fig. 13 Culture characteristic on MEA: Nothophoma chromolaenae (MFLUCC 17-1443) to semi-immersed, uni-loculate, globose to obpyriform, yellowish brown to brown, sometimes appearing as colonies on the host surface, superficial, scattered, gregarious, dark brown to black, not easy to remove from the host surface. Fungal Diversity Table 3 Morphological features of Nothophoma species discussed in this study Species Conidiomata (µm) Conidiogenous cells (µm) Conidia (µm) Hosts References N. gossypiicola (CBS. 377.67) N. macrospora (UTHSC: DI16-276) N. multilocularis (AUMC 12003) N. raii (MCC 1082) 100–250 5–8 × 5–8 10–12.5 × 2.5–3.5 De Gruyter (2002) 100–300 5–10 diam. 175–1500 11–17 × 9–18 (9–)10–15 × 2.5–3(– 3.5) 9–20 × 3–4(–5) 194.3–315.5 × 195.6– 411.3 – 11–14.5 × 1.5–2.5 Parasitic on Gossypium spp. Human clinical specimen Endophyte of Rhazya stricta. Soil from industrial area Crous et al. (2016) Abdel-Wahab (2017) Crous et al. (2017) Table 4 Synopsis of asexual morph of Chromolaenicola species with similar morphological features discussed in this study Species Conidiomata (μm) Conidiogenous cells (μm) Conidia (μm) C. chiangraiensis (MFLUCC 17-1493) – 3.5–6.5 × 1–2 9–14 × 6–9 C. lampangensis (MFLUCC 17-1462) 150–230 high × (155– – )170–270(–345) diam. C. siamensis (= Cylindroaseptospora siamensis, MFLUCC 17–2527) 110–165 high × 140– 190 diam. 6.5–7.4 × 3.2–4.7 Ostiole central, papillate. Pycnidial wall comprising 2–4 layers, pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells hyaline and unbranched, smooth, elongated, broadly filiform to ampulliform. Conidia oblong or oval to ellipsoid, globose to subglobose, hyaline to pale brown, aseptate when immature, becoming reddish brown to brown, 1-septate when mature, not constricted at the septum, thickwalled finely verruculose. Type species: Chromolaenicola nanensis Mapook & K.D. Hyde Notes: Multigene phylogenetic analyses based on combined LSU, ITS, SSU and TEF1 sequence data show that Chromolaenicola species form a monophyletic clade in Didymospheriaceae (0.92 BYPP) and basal to Cylindroaseptospora leucaenae (MFLUCC 17-2424) with low bootstrap support (Fig. 14). Chromolaenicola species differ from Cy. leucaenae in having oblong or oval to ellipsoid, globose to subglobose conidia, hyaline to pale brown and aseptate when immature, becoming dark and 1-septate at maturity, thick-walled, verruculose, not constricted at the septum, whereas, Cy. leucaenae has cylindrical conidia, hyaline, aseptate with smooth thin walls (Jayasiri et al. 2019). Hosts References This study Thailand, Chiang Rai Province on dead stems of Chromolaena odorata This study 12–15 × 4–6.5 Thailand, Lampang Province on dead stems of Chromolaena odorata Jayasiri et al. (2019) 7.2–9.4 × 5.4–6.5 Thailand, Lampang Province on decaying pod of Leucaena sp. (Fabaceae) Therefore, we introduce Chromolaenicola as a new genus to accommodate four new species with one new taxonomic combination, based on morphological comparison and phylogenetic analyses. Chromolaenicola chiangraiensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557280, Facesoffungi number: FoF 07784; Fig. 15 Etymology: Referring to the location where the specimen was collected, Chiang Rai Province, Thailand. Holotype: MFLU 20-0301 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Colonies superficial, covering the host, scattered, gregarious, dark brown to black, not easy to remove from the host surface. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 3.5–6.5 × 1–2 μm ( x̄ = 5 × 1.7 µm, n = 10), holoblastic, hyaline, branched, smooth, elongated, broadly filiform to ampulliform. Conidia 9–14 × 6–9 μm ( x̄ = 11 × 7.5 µm, n = 45), oval to ellipsoid, 1-septate, thick-walled, not constricted at the septum, reddish brown, verruculose. Culture characteristics: Conidia germinating on MEA within 24 h at room temperature and germ tubes produced 13 Fungal Diversity Fig. 14 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU and TEF1 sequence data. Ninetyfive strains are included in the combined sequence analysis, which comprise 3280 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 17325.094501 is presented. The matrix had 1124 distinct alignment patterns, with 34.85% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.241071, C = 0.243154, G = 0.273420, T = 0.242354; substitution rates: AC = 1.437040, AG = 2.295778, AT = 1.367419, CG = 0.976650, CT = 7.785933, GT = 1.000000; gamma distribution shape parameter α = 0.200569. 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 dark red bold and type species are in bold. Pleospora tarda (CBS 714.68), Pleospora herbarum (CBS 191.86) and Pleospora herbarum (IT 956) are used as outgroup taxa from the apex. Colonies on MEA circular or irregular, mycelium slightly flattened, filamentous, cultures white to creamy white on surface and pale brown to brown from the centre of the colony, olivaceous to olivaceous brown appearing as concentric rings pattern with white margin in reverse (Fig. 19a). 13 Pre-screening for antimicrobial activity: Chromolaenicola chiangraiensis (MFLUCC 17-1493) showed no inhibition against E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP92, MFLU 20-0301, holotype); ex-type culture MFLUCC 17-1493. Fungal Diversity Fig. 15 Chromolaenicola chiangraiensis (holotype) a, b Appearance of colonies on substrate. c–i Conidia with conidiogenous cells. Scale bars: a, b = 500 µm, c = 10 µm, d–i = 5 µm GenBank numbers: LSU: MN325005, ITS: MN325017 SSU: MN325011, TEF1: MN335650, RPB2: MN335655 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS, LSU and TEF1 sequences of Chromolaenicola chiangraiensis (MFLUCC 17-1493, ex-holotype) is Cylindroaseptospora siamensis (strain MFLUCC 17-2527) with 100% (MK347760), 99.88% (MK347976) and 99.32% (MK360048) similarity, respectively, while the closest match with the SSU sequence with 99.89% similarity was Didymosphaeria variabile (strain STE-U 6311, NG_064914). In the present phylogenetic analysis, Chromolaenicola chiangraiensis (MFLUCC 17-1493) is closely related to C. siamensis MFLUCC 17-2527 with high bootstrap support (100% ML and 1.00 BYPP, Fig. 14). Chromolaenicola chiangraiensis is similar to C. siamensis (= Cylindroaseptospora siamensis, MFLUCC 17–2527) in having 1-septate, verruculose conidia but differs in having smaller conidiogenous cells (3.5–6.5 × 1–2 μm vs. 6.5–7.4 × 3.2–4.7 μm) and larger conidia (9–14 × 6–9 μm vs. 7.2–9.4 × 5.4–6.5 μm) (Table 4). Therefore, we introduced a new species based on phylogeny and morphological comparison. Chromolaenicola lampangensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557281, Facesoffungi number: FoF 07785; Fig. 16 Etymology: Referring to the location where the specimen was collected, Lampang Province, Thailand. Holotype: MFLU 20-0302 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata 150–230 µm high × (155–)170–270(–345) µm diam. ( x̄ = 190 × 215 µm, n = 15), pycnidial, solitary, immersed to semi-immersed, uni-loculate, globose to obpyriform, yellowish brown to brown. Ostiole central, papillate. Pycnidial wall (5–)10–20 µm wide, comprising 2–4 layers of pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells holoblastic, ampulliform, hyaline, unbranched. Conidia 12–15 × 4–6.5 μm ( x̄ = 13.5 × 5 µm, n = 40), oblong to oval, hyaline to pale brown, aseptate when immature, becoming reddish brown to brown and 1-septate when mature, not constricted at septum, thick-walled, finely verruculose. Culture characteristics: Conidia germinating on MEA within 48 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular or irregular, mycelium slightly flattened, filamentous, cultures creamy white to yellow from the centre of the colony with white to pale green on surface and pale brown to brown from the 13 Fungal Diversity Table 5 Synopsis of sexual morph of Chromolaenicola species with similar morphological features discussed in this study Species Asocomata (µm) Peridium (µm) Asci (µm) Ascospores (µm) References C. nanensis (MFLUCC 17-1473) C. thailandensis (MFLUCC 17-1510) 210–230 × 200–220 (10–)15–20(–25) (90–)110–145 × 10–12.5 16–20 × 7.5–9 This study 145–225(–250) × (150– )175–240(–285) 10–20(–35) 90–160 × 10–14 16–24 × 9–11 This study Fig. 16 Chromolaenicola lampangensis (holotype) a, b Appearance of conidiomata on substrate. c Section through conidioma. d Ostiole. e Peridium. f–h Conidiogenous cells and developing conidia. i–m Conidia. Scale bars: a = 500 µm, b = 200 µm, c, d = 50 µm, e = 20 µm, h = 10 µm, f, g, i–m = 5 µm centre of the colony, olivaceous to olivaceous brown appearing as concentric rings pattern with white margin in reverse (Fig. 19b). Pre-screening for antimicrobial activity: Chromolaenicola lampangensis (MFLUCC 17-1462) showed antimicrobial activity against M. plumbeus with a 14 mm inhibition zone, observable as partial inhibition, when compared to 13 the positive control (16 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Lampang Province, Chaehom, on dead stems of Chromolaena odorata, 24 September 2016, A. Mapook (JH8, MFLU 20-0302, holotype); ex-type culture MFLUCC 17-1462. GenBank numbers: LSU: MN325004, ITS: MN325016, SSU: MN325010, TEF1: MN335649, RPB2: MN335654 Fungal Diversity Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Chromolaenicola lampangensis (MFLUCC 17-1462, ex-holotype) was Paraconiothyrium sp. with 98.75% similarity to the strain PNB15_1B1 (MH268018). The closest match with the LSU sequence with 99.02% similarity was Coniothyrium nitidae (strain CBS 119209, EU552112). The closest match with the TEF1 sequences with 98.97% similarity was Cylindroaseptospora siamensis (strain MFLUCC 17-2527, MK360048), while the closest match with the SSU sequence with 99.89% similarity was Paraconiothyrium thysanolaenae (strain MFLUCC 10-0550, NG_063570). In the present phylogenetic analysis, C. lampangensis (MFLUCC 17-1462) is closely related to C. nanensis with high bootstrap support (90% ML and 1.00 BYPP, Fig. 14). Chromolaenicola nanensis is introduced as a new species in this study and found as the sexual morph in nature; however, we failed to obtain the sexual morph of C. lampangensis. Our strain, C. lampangensis is similar to C. siamensis (= Cylindroaseptospora siamensis, MFLUCC 17–2527) and C. chiangraiensis (MFLUCC 17-1493) in having 1-septate with verruculose conidia but differs from C. siamensis in having larger conidiomata (150–230 × (155–)170–270(–345) μm vs. 110–165 × 140–190 μm) with larger conidia (12–15 × 4–6.5 μm vs. 7.2–9.4 × 5.4–6.5 μm), and differs from C. chiangraiensis in having narrower conidia (12–15 × 4–6.5 μm vs. 9–14 × 6–9 μm) (Table 4). A comparison of the RPB2 gene region of C. lampangensis and C. nanensis reveals 20 base pair differences (1.87%) across 1068 nucleotides. Therefore, we introduce a new species based on phylogeny and morphological comparison. Chromolaenicola nanensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557282, Facesoffungi number: FoF 07786; Fig. 17 Etymology: Referring to the location where the specimen was collected, Nan Province, Thailand. Holotype: MFLU 20-0304 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 210–230 µm high × 200–220 µm diam. ( x̄ = 216 × 211 µm, n = 5), immersed to semi-immersed, solitary or scattered, appearing as small dark spots, coriaceous, globose to subglobose, brown to dark brown. Ostiolar neck protruding. Peridium (10–)15–20(–25) µm wide, 3–4 layers, comprising dark brown cells of textura angularis. Hamathecium comprising 1.5–2.5(–3) µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci (90–)110–145 × 10–12.5 µm ( x̄ = 120 × 11 µm, n = 20), 6–8-spored, bitunicate, cylindrical, straight or slightly curved, apically rounded, pedicellate with an ocular chamber. Ascospores 16–20 × 7.5–9 µm ( x̄ = 18.5 × 8 µm, n = 45), slightly overlapping, uni-seriate, initially hyaline, 1-septate when immature, becoming reddish brown to brown at maturity, ellipsoid to broadly fusiform, muriform, and 3-transversely septate, with 1-vertical septum when mature, constricted at the central septum, straight or slightly curved, without gelatinous sheath. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly flattened, filamentous, cultures white to pale green on surface, creamy white to yellow in reverse from the centre of the colony, light green to olivaceous appearing as concentric rings with white margin (Fig. 19c). Pre-screening for antimicrobial activity: Chromolaenicola nanensis (MFLUCC 17-1473) showed antimicrobial activity against M. plumbeus with a 12 mm inhibition zone, observable as partial inhibition, when compared to the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Nan Province, Doi Phu Kha, on dead stems of Chromolaena odorata, 23 September 2016, A. Mapook (DPK5, MFLU 20-0304, holotype); ex-type culture MFLUCC 17-1473; (DPK11, MFLU 20-0303); living culture MFLUCC 17-1477. GenBank numbers: LSU: MN325003, MN325002, ITS: MN325015, MN325014, SSU: MN325009, MN325008, TEF1: MN335648, MN335647, RPB2: MN335653 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Chromolaenicola nanensis (MFLUCC 17-1473, ex-holotype) was Paraconiothyrium sp. with 98.75% similarity to the strain PNB15_1B1 (MH268018). The closest match with the LSU sequence with 99.01% similarity was Coniothyrium nitidae (strain CBS 119209, EU552112). The closest match with the TEF1 sequence with 98.68% similarity was Cylindroaseptospora siamensis (strain MFLUCC 17-2527, MK360048), while the closest match with the SSU sequence with 99.79% similarity was Paraconiothyrium thysanolaenae (MFLUCC 10-0550, NG_063570). Chromolaenicola nanensis (MFLUCC 17-1473) is found as sexual morph in nature and we could not obtain its asexual morph in culture. The strain is similar to C. thailandensis (MFLUCC 17-1510) but differs in having slightly smaller ascomata (210–230 × 200–220 µm vs. 145–2 25(–250) × (150–)175–240(–285) µm), slightly smaller asci [(90–)110–145 × 10–12.5 µm vs. 90–160 × 10–14 µm] and smaller ascospores (16–20 × 7.5–9 µm vs. 16–24 × 9–11 µm) (Table 5). In the present phylogenetic analysis, C. nanensis groups with C. lampangensis (MFLUCC 17-1462) which is found as an asexual morph in nature. However, they differ in culture characteristics on MEA (Fig. 19). A comparison of the RPB2 gene region of C. nanensis and C. lampangensis reveals 20 base pair differences (1.87%) across 1068 nucleotides. Therefore, C. nanensis is described as a new species based on phylogeny and morphological comparison. 13 Fungal Diversity Fig. 17 Chromolaenicola nanensis (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Immature and mature asci. k–q Ascospores. Scale bars: a, b = 200 µm, c, g–j = 50 µm, d, e = 20 µm, f, k–q = 10 µm Chromolaenicola siamensis (Jayasiri, E.B.G. Jones & K.D. Hyde) Mapook & K.D. Hyde, comb. nov. Index Fungorum number: IF557283, Facesoffungi number: FoF 07787 13 ≡ Cylindroaseptospora siamensis Jayasiri, E.B.G. Jones & K.D. Hyde, in Jayasiri et al., Mycosphere 10(1): 68 (2019) Holotype: THAILAND, Lampang Province on decaying pod of Leucaena sp. (Fabaceae), 18 August 2017, S.C. Fungal Diversity Fig. 18 Chromolaenicola thailandensis (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Immature and mature asci. k–p Ascospores. Scale bars: a, b = 500 µm, c, g–j = 50 µm, d, e = 20 µm, f, k–p = 10 µm Jayasiri, C 329 (MFLU 18–2147, holotype; KUN-HKAS 102427, isotype); ex-type culture MFLUCC 17–2527, KUMCC 18–0227. Morphological description: See Jayasiri et al. (2019) Notes: In our multigene phylogenetic study, Cylindroaseptospora siamensis (MFLUCC 17-2527) was in a clade separate from the type species, Cylindroaseptospora leucaenae (MFLUCC 17-2424) and clustered with Chromolaenicola chiangraiensis (MFLUCC 17-1493) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 14). The species also shares similar morphological characters with other Chromolaenicola spp., such as ampulliform conidiogenous cells and globose or subglobose to oval or ellipsoid conidia, hyaline to pale brown, aseptate when immature, becoming dark and 1-septate at maturity, thick-walled, not constricted at the septum, and verruculose, while Cy. leucaenae has cylindrical conidia that are hyaline, aseptate, smooth and thin-walled. Therefore, we transfer Cylindroaseptospora siamensis as Chromolaenicola siamensis. Chromolaenicola thailandensis Mapook & K.D. Hyde, sp. nov. 13 Fungal Diversity Index Fungorum number: IF557284, Facesoffungi number: FoF 07788; Fig. 18 Etymology: Named after Thailand, where the fungus was discovered. Holotype: MFLU 20-0306 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 145–225(–250) µm high × (150–)175–240(–285) µm diam. ( x̄ = 195 × 215 µm, n = 5), immersed, solitary or scattered, appearing as dark spots, coriaceous, globose, light brown to brown. Ostiole central. Peridium 10–20(–35) µm wide, 3–4 layers at side, thin-walled, inner layers comprising hyaline to pale brown cells of textura angularis, outer layers comprising brown cells of textura intricata. Hamathecium comprising (1.5–)2–2.5(–3) µm wide, cylindrical, septate, branching pseudoparaphyses. Asci 90–160 × 10–14 µm ( x̄ = 124 × 11.5 µm, n = 20), 8-spored, bitunicate, cylindrical, straight or slightly curved, apically rounded, pedicellate. Ascospores 16–24 × 9–11 µm ( x̄ = 21 × 10 µm, n = 35), uni-seriate, initially hyaline to pale brown, 1-septate when immature, becoming reddish brown to brown at maturity, ellipsoid to broadly fusiform, muriform, 3-transversely septate, with none or 1-vertical septum, constricted at the central septum, straight or slightly curved, without gelatinous sheath. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium flattened, curled, creamy white to pale yellow with white margin on surface, pale orange from the centre of the colony with white margin in reverse (Fig. 19d). Pre-screening for antimicrobial activity: Chromolaenicola thailandensis (MFLUCC 17-1510) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Mae Hong Son Province, Mae Yen, Pai, on dead stems of Chromolaena odorata, 25 June 2016, A. Mapook (MY11, MFLU 20-0306, holotype); ex-type culture MFLUCC 17-1510; Nan Province, Doi Phu Kha, on dead stems of C. odorata, 23 September 2016, A. Mapook (DPK7, MFLU 20-0305); living culture MFLUCC 17-1475. GenBank numbers: LSU: MN325006, MN325007, ITS: MN325018, MN325019, SSU: MN325012, MN325013, TEF1: MN335651, MN335652, RPB2: MN335656 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS and TEF sequences of Chromolaenicola thailandensis (MFLUCC 17-1510, ex-holotype) with 100% (MK347760) and 98.98% (MK360048) similarity, respectively, was Cylindroaseptospora siamensis strain MFLUCC 17-2527. The closest match with the LSU sequence with 99.19% similarity was Alloconiothyrium aptrootii (strain CBS 981.95, JX496235), while the closest match of the SSU sequence with 99.89% similarity was Paraconiothyrium thysanolaenae (strain MFLUCC 10-0550, NG_063570). In the present phylogenetic analysis, C. thailandensis (MFLUCC 17-1510) forms a sister taxon with other Chromolaenicola species (Fig. 14). The strain is found as a sexual morph in nature and we could not obtain its asexual morph in culture. Chromolaenicola thailandensis is similar to C. nanensis in having uni-seriate ascospores that are 1-septate when immature, becoming reddish brown to brown at maturity, ellipsoid to broadly fusiform, muriform, and 3-transversely septate, with 1-vertical septum when mature, constricted at the central septum, without gelatinous sheath, but differs in having slightly larger ascomata (145–225(–250) × (150–)175– 240(–285) µm vs. 210–230 × 200–220 µm), slightly larger asci (90–160 × 10–14 µm vs. (90–)110–145 × 10–12.5 µm) and slightly larger ascospores (16–24 × 9–11 µm vs. 16–20 × 7.5–9 µm) (Table 5). A comparison of the TEF1 gene region of C. thailandensis and C. nanensis reveals 12 base pair differences (1.5%) across 798 nucleotides. Therefore, C. thailandensis is described as a new species based on phylogeny and morphological comparison. Montagnula Berl., Icon. fung. (Abellini) 2: 68. 1896. Montagnula was introduced by Berlese (1896) to accommodate M. infernalis as the type species together with M. gigantean based on morphology. The genus comprises saprobes growing on dead wood, branches, stems, bark and Fig. 19 Culture characteristics on MEA: a Chromolaenicola chiangraiensis (MFLUCC 17-1493). b Chromolaenicola lampangensis (MFLUCC 17-1462). c Chromolaenicola nanensis (MFLUCC 17-1473). d Chromolaenicola thailandensis (MFLUCC 17-1510) 13 Fungal Diversity leaves, which were placed in the family Didymosphaeriaceae by Ariyawansa et al. (2014b). Liu et al. (2015) introduced a new species, M. graminicola and Hongsanan et al. (2015) introduced two new species M. bellevaliae and M. scabiosae which were collected in Italy. Wanasinghe et al. (2016) introduced a new species, M. saikhuensis from Thailand and synonymized two Munkovalsaria species (M. donacina and M. appendiculata) under the genus Montagnula. Hyde et al. (2016b) introduced M. cirsii from Italy, and M. jonesii was introduced as a new species by Tennakoon et al. (2016). Niranjan and Sarma (2018) introduced M. vakrabeejae as a new species from India, based on morphological comparison. Tibpromma et al. (2018) introduced M. krabiensis from Pandanaceae in Thailand. Presently 40 epithets are listed in Index Fungorum (2020). In this study, four new Montagnula species are introduced, based on morphology and molecular data, together with descriptions and illustrations (Figs. 21, 22, 23, 24). A phylogenetic tree based on combined LSU, ITS, SSU and TEF1 sequence data is presented in Fig. 20. Montagnula chiangraiensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557296, Facesoffungi number: FoF 07789; Fig. 21 Etymology: Referring to the location where the specimen was collected, Chiang Rai Province, Thailand. Holotype: MFLU 20-0322 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (130–)150–220 µm high × (170–)200–230 µm diam. ( x̄ = 172 × 207 µm, n = 10), immersed to erumpent, solitary, scattered, globose, coriaceous, reddish brown to brown. Ostiole papillate, protruding from substratum. Peridium (7–)10–20 µm wide, comprising 2–3-layers of thin-walled, reddish brown to light brown cells of textura angularis. Hamathecium comprising 1.5–3.5 µm wide, cylindrical to filiform, septate, branching Fig. 20 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU and TEF1 sequence data. Twenty strains are included in the combined sequence analysis, which comprise 4029 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 12550.757272 is presented. The matrix had 911 distinct alignment patterns, with 34.97% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.244653, C = 0.247314, G = 0.272188, T = 0.235846; substitution rates: AC = 1.302293, AG = 2.261934, AT = 1.509666, CG = 1.082059, CT = 5.323888, GT = 1.000000; gamma distribution shape parameter α = 0.176588. 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 dark red bold and type species are in bold. Pleospora herbarum (IT 956) is used as outgroup taxon 13 Fungal Diversity Fig. 21 Montagnula chiangraiensis (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Immature and mature asci. k–o Ascospores (immature and mature) with terminal appendages. p Ascospores in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c = 50 µm, d, g–j = 20 µm, e, k–p = 10 µm, f = 5 µm pseudoparaphyses. Asci (45–)60–75 × (6–)8–11 µm ( x̄ = 65 × 9.5 µm, n = 20), 8-spored, bitunicate, fissitunicate, cylindric-clavate, slightly curved, with a short, bulbous long pedicel, with small ocular chamber. Ascospores 11–15 × 4–6 µm ( x̄ = 12 × 5 µm, n = 25), overlapping 2–3-seriate, pale brown when immature and dark reddish brown to dark brown when mature, broadly fusiform to 13 ellipsoid, 1-septate, constricted at the septum, slightly widest at the upper cell and tapering towards ends, straight to slightly curved, with small guttules; sheath drawn out to form polar appendages 3–6 × 1–2.5 µm ( x̄ = 5 × 2 µm, n = 10), from both ends of the ascospores, straight or slightly curved. Asexual morph: Undetermined. Fungal Diversity Fig. 22 Montagnula chromolaenae (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores with terminal appendages. Scale bars: a, b = 200 µm, c = 50 µm, d, e, g–j = 20 µm, f, k–p = 10 µm Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA irregular, mycelium slightly flattened, undulate, cultures white to creamy white on surface, brown to dark brown in reverse from the centre of the colony, creamy white at margin with white mycelium appearing as small spots spreading around the colony (Fig. 25a). Pre-screening for antimicrobial activity: Montagnula chiangraiensis (MFLUCC 17-1420) showed antimicrobial activity against B. subtilis with an 8 mm inhibition zone and against M. plumbeus with an 11 mm inhibition zone, observable as partial inhibition, when compared to the positive control (26 mm and 17 mm, respectively), but no inhibition of E. coli. 13 Fungal Diversity Fig. 23 Montagnula chromolaenicola (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Immature and mature asci. k–p Ascospores. Scale bars: a, b = 500 µm, c = 100 µm, g–j = 50 µm, d, e = 20 µm, f, k–p = 5 µm Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 8 July 2015, A. Mapook, (DP1, MFLU 20-0322, holotype); extype culture MFLUCC 17-1420. GenBank numbers: LSU: MT214443, ITS: MT214349, SSU: MT214397 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Montagnula chiangraiensis (MFLUCC 17-1420, ex-holotype) with 98.91% similarity 13 was M. appendiculata (strain M-1, DQ435529). The closest match with the LSU sequence with 98.34% similarity was M. aloes (strain CBS 132531, NG_042676), while the closest match with the SSU sequence with 97.76% similarity was Montagnula sp. (strain MFLUCC 11-0576, KJ188099). In the present phylogenetic analysis, M. chiangraiensis is closely related to M. appendiculata with high bootstrap support (70% ML and 0.98 BYPP, Fig. 20). Montagnula chiangraiensis differs from M. appendiculata in its culture Fungal Diversity Fig. 24 Montagnula thailandica (holotype) a Appearance of ascomata on substrate. b Section through ascoma with ostiole. c Peridium. d Pseudoparaphyses. e–h Immature and mature asci. i–n Ascospores (immature and mature). Scale bars: a = 500 µm, b = 100 µm, e–h = 50 µm, c = 20 µm, i–n = 10 µm, d = 5 µm Fig. 25 Culture characteristics on MEA: a Montagnula chiangraiensis (MFLUCC 17-1420). b Montagnula chromolaenae (MFLUCC 17-1435). c Montagnula chromolaenicola (MFLUCC 17-1469). d Montagnula thailandica (MFLUCC 17-1508) characteristics on MEA. Montagnula chiangraiensis has white mycelium appearing as small spots spreading around the colony, which are not reported for M. appendiculata (Aptroot 2004). Moreover, M. chiangraiensis has dark reddish brown to dark brown ascospores, while M. appendiculata has yellowish brown to brown ascospores. However, we could not compare size of asci as this feature was not given for the holotype of M. appendiculata (Aptroot 2004) (Table 6). A comparison of the ITS (+5.8S) gene region of M. chiangraiensis and M. appendiculata reveals 14 base pair differences (1.53%) across 914 nucleotides. Therefore, 13 Fungal Diversity Table 6 Synopsis of Montagnula species with similar morphological features discussed in this study Species Asocomata (µm) Peridium (µm) Asci (µm) Ascospores (µm) Polar appendages (µm) References M. appendiculata (CBS 109027) M. chiangraiensis (MFLUCC 17-1420, DP1) M. chromolaenae (MFLUCC 17-1435, DP18) M. chromolaenicola (MFLUCC 17-1469, MY1) M. donacina (HVVV01) 100–200 diam. – Not reported 12–15 × 4–5 4–7 × 1.5–2.5 Aptroot (2004) (130–)150–220 high × (170– )200–230 diam. 170–175(–230) high × (140– )170–190 diam. 300–320 high × 215–310 diam. – (7–)10–20 (45–)60– 75 × (6–)8–11 11–15 × 4–6 3–6 × 1–2.5 This study 10–25 85–105 × 9–15 15–16.5 × 5–6 6–11 × (1.8–)2–3 This study 10–25 80–100 × 10–13 15–17 × 5–6.5 1-septate, without polar appendages This study – Not reported M. donacina 500 diam. – 90–100 × 12–13 M. graminicola (MFLUCC 13-0352) 37–117.22 diam. 14.9–16 (45–)50–132(– 137) × (6–)8– 13(–15) 10–20 70–100 × 10–12 (13.6–)14.8–15.2(– 1-septate, without Pitt et al. (2014) mucilaginous 17.3) × (6.6– sheath )7.5–7.7(–8.3) 12–15 × 4 1-septate, without Saccardo (1882) mucilaginous sheath Liu et al. (2015) 1-septate, sur(7.8–)9.8–13(– rounded by a 15) × (2.8–)3.8– mucilaginous 5.5(–6.5) sheath 12–16 × 4–6 1-septate, without Wanasinghe et al. (2016) mucilaginous sheath (12–)14–17 × 4.5– 1-septate, without This study 7.5 polar appendages 400–450 M. saikhuensis high × 400–500 (MFLUCC diam. 16-0315) (315–)405–415 M. thailandica high × 330–350 (MFLUCC diam. 17-1508, FHP16) (10–)15–20(–30) 80–100 × 9–15 M. chiangraiensis is described as a new species based on phylogeny and morphological comparison. Montagnula chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557297, Facesoffungi number: FoF 07790; Fig. 22 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0323 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 170–175(–230) µm high × (140–)170–190 µm diam. ( x̄ = 185 × 172 µm, n = 5), immersed, solitary or scattered, globose to subglobose, coriaceous, reddish brown to brown. Ostiole central, papillate. Peridium 10–25 µm wide, comprising of 1–2 layers of pale brown to yellowish brown cells of textura angularis. Hamathecium comprising (2.5–)3–4(–5) µm wide, cylindrical, septate, branching pseudoparaphyses. Asci 85–105 × 9–15 ( x̄ = 93 × 12 µm, n = 15), bitunicate, (4–)6–8-spored, cylindric-clavate, straight or slightly curved, long pedicel. Ascospores 15–16.5 × 5–6 ( x̄ = 16 × 5.5 µm, n = 15), overlapping bi-seriate, hyaline to pale brown when immature and yellowish brown to light brown when mature, broadly fusiform to ellipsoid, 1-septate, constricted 13 at the septum, slightly widest at the upper cell and tapering towards ends, straight to slightly curved; sheath drawn out to form polar appendages 6–11 µm long × (1.8–)2–3 µm wide ( x̄ = 8 × 2.5 µm, n = 20), from both ends of the ascospores, straight or slightly curved. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA irregular, mycelium slightly flattened, undulate, cultures white to creamy white on surface, brown to dark brown in reverse (Fig. 25b). Pre-screening for antimicrobial activity: Montagnula chromolaenae (MFLUCC 17-1435) showed antimicrobial activity against M. plumbeus with an 18 mm inhibition zone, observable as partial inhibition, when compared to the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP18, MFLU 20-0323, holotype); extype culture MFLUCC 17-1435. GenBank numbers: LSU: MT214444, ITS: MT214350, SSU: MT214398 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Montagnula chromolaenae Fungal Diversity (MFLUCC 17-1435, ex-holotype) with 97.60% similarity was M. appendiculata (strain M-1, DQ435529). The closest match with the LSU sequence with 99.09% similarity was Montagnula sp. (strain MFLUCC 11-0576, KJ188098), while the closest match of the SSU sequence with 99.16% similarity was Montagnula sp. (strain MFLUCC 11-0576, KJ188099). In the present phylogenetic analysis, M. chromolaenae is closely related to M. appendiculata and M. chiangraiensis (Fig. 20). However, M. chromolaenae differs from M. appendiculata in having larger ascomata (170–1 75(–230) × (140–)170–190 µm vs. 100–200 µm diam.) and larger ascospores (15–16.5 × 5–6 µm vs. 12–15 × 4–5 µm) with longer polar appendages (6–11 × (1.8–)2–3 µm vs. 4–7 × 1.5–2.5 µm) (Table 6). A comparison of the ITS (+5.8S) gene region of M. chromolaenae and M. appendiculata reveals 25 base pair differences (2.7%) across 914 nucleotides. Therefore, M. chromolaenae is described as a new species based on phylogeny and morphological comparison. Montagnula chromolaenicola Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557298, Facesoffungi number: FoF 07791; Fig. 23 Etymology: Name reflects the host genus Chromolaena, on which this species was growing. Holotype: MFLU 20-0324 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 300–320 µm high × 215–310 µm diam. (x̅ = 310 × 275 µm, n = 5), semi-immersed to erumpent, solitary or scattered, globose to obpyriform, coriaceous, brown to dark brown. Ostiole papillate, protruding from substratum. Peridium 10–25 µm wide, comprising several layers of thin-walled, pale brown to brown cells of textura angularis. Hamathecium comprising 1–2 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 80–100 × 10–13 µm ( x̄ = 90 × 12 µm, n = 10), 8-spored, bitunicate, elongate-clavate, slightly curved, long pedicel. Ascospores 15–17 × 5–6.5 µm ( x̄ = 15.5 × 6 µm, n = 15), overlapping 1–2-seriate, pale brown to yellowish brown when immature, becoming brown to dark brown when mature, broadly fusiform to ellipsoid, 1-septate, constricted at the septum, slightly wider upper cell and tapering towards ends, straight to slightly curved, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium slightly flattened, entire, cultures grayish green to light brown on surface and dark brown to black in reverse (Fig. 25c). Pre-screening for antimicrobial activity: Montagnula chromolaenicola (MFLUCC 17-1469) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Mae Hong Son Province, Mae Yen, Pai, on dead stems of Chromolaena odorata, 25 June 2016, A. Mapook (MY1, MFLU 20-0324, holotype); ex-type culture MFLUCC 17-1469. GenBank numbers: LSU: MT214445, ITS: MT214351, SSU: MT214399, TEF1: MT235773, RPB2: MT235809 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Montagnula chromolaenicola (MFLUCC 17-1469, ex-holotype) with 99.82% similarity was Aporospora terricola (strain 2711, EU272515). The closest match with the LSU sequence with 99.22% similarity was Munkovalsaria donacina (strain HVVV01, KJ628377). The closest match with the SSU sequence with 99.00% similarity was Didymocrea sadasivanii (strain CBS 438, DQ384066), while the closest match with the TEF1 sequence with 97.05% similarity was Montagnula sp. (strain UTHSC: DI16-251, LT797091). In the present phylogenetic analysis, M. chromolaenicola forms a separate branch and clusters with M. donacina (Fig. 20). However, M. chromolaenicola differs from M. donacina in having smaller ascomata (300–320 × 215–310 µm vs. 500 µm diam.) and slightly smaller asci (80–100 × 10–13 µm vs. 90–100 × 12–13 µm) with longer ascospores (15–17 µm × 5–6.5 vs. 12–15 × 4 µm) (Table 6). Therefore, M. chromolaenicola is described as a new species based on phylogeny and morphological comparison. Montagnula thailandica Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557299, Facesoffungi number: FoF 07792; Fig. 24 Etymology: The name reflects the country, where the specimen was collected, Thailand. Holotype: MFLU 20-0325 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (315–)405–415 µm high × 330–350 µm diam. ( x̄ = 380 × 340 µm, n = 5), immersed to erumpent, solitary or scattered, globose to obpyriform, coriaceous, brown to dark brown. Ostiole papillate, protruding from substratum. Peridium (10–)15–20(–30) µm wide, comprising several layers of thin-walled, pale brown to brown cells of textura angularis. Hamathecium comprising 1–2.5 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 80–100 × 9–15 µm ( x̄ = 90 × 11 µm, n = 15), 8-spored, bitunicate, fissitunicate, elongate-clavate, slightly curved, long pedicel. Ascospores (12–)14–17 × 4.5–7.5 µm ( x̄ = 15 × 5.5 µm, n = 25), overlapping 1–2-seriate, hyaline or pale brown to yellowish brown when immature and becoming brown to reddish brown when mature, broadly fusiform to ellipsoid, 1-septate, constricted at the septum, slightly wider upper cell and tapering towards ends, straight to slightly curved, without terminal appendages. Asexual morph: Undetermined. 13 Fungal Diversity Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium raised, velvety with moderately fluffy, filamentous at margin, cultures grey with white from the centre of the colony on surface and creamy white in reverse with pale greyish brown margin (Fig. 25d). Pre-screening for antimicrobial activity: Montagnula thailandica (MFLUCC 17-1508) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Mai Province, Fah Hom Pok, on dead stems of Chromolaena odorata, 27 September 2016, A. Mapook (FHP16, MFLU 20-0325, holotype); ex-type culture MFLUCC 17-1508. GenBank numbers: LSU: MT214446, ITS: MT214352, SSU: MT214400, TEF1: MT235774, RPB2: MT235810 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Montagnula thailandica (MFLUCC 17-1508, ex-holotype) with 99.29% similarity was Aporospora terricola (strain 2711, EU272515). The closest match with SSU sequences with 98.91% similarity was Didymocrea sadasivanii (strain CBS 438, DQ384066). The closest match with the LSU sequence with 99.22% similarity was Munkovalsaria donacina (strain HVVV01, KJ628377), while the closest match with the TEF1 sequence with 96.75% similarity was Paraconiothyrium cyclothyrioides (UTHSC: DI16-327, LT797124). In the present phylogenetic analysis, M. thailandica forms a separate branch and groups with Montagnula chromolaenicola, M. donacina, M. saikhuensis and M. graminicola (Fig. 20). Our strain is similar to those Montagnula species in having long pedicellate asci and broadly fusiform to ellipsoid, 1-septate ascospores, without terminal appendages. However, our strain differs from M. graminicola in having larger ascomata [(315–)405–415 × 330–350 µm vs. 37–117.22 µm diam.] and larger ascospores [(12–)14–17 × 4.5–7.5 µm vs. (7.8–)9.8–13(–15) × (2.8– )3.8–5.5(–6.5) µm], without gelatinous sheath, while M. graminicola has ascospores surrounded by a mucilaginous sheath and our strain also differs from M. saikhuensis in having smaller ascomata [(315–)405–415 × 330–350 µm vs. 400–450 × 400–500 µm) with slightly larger asci (80–100 × 9–15 µm vs. 70–100 × 10–12 µm) and slightly larger ascospores [(12–)14–17 × 4.5–7.5 µm vs. 12–16 × 4–6 µm] (Table 6). Montagnula thailandica also differs from M. chromolaenicola in having larger ascomata [(315–)405–415 × 330–350 µm vs. 300–320 × 215–310 µm], slightly wider asci (80–100 × 9–15 µm vs. 80–100 × 10–13 µm), with brown to reddish brown ascospores, while M. chromolaenicola has brown to dark brown ascospores. A comparison of the ITS (+5.8S) gene region of M. thailandica and M. chromolaenicola reveals 13 base pair differences (1.51%) across 858 13 nucleotides. Therefore, M. thailandica is described as a new species based on phylogeny and morphological comparison. Pseudopithomyces Ariyaw. & K.D. Hyde Pseudopithomyces was introduced by Ariyawansa et al. (2015) with P. chartarum as the type species, along with a new species, P. palmicola. Crous et al. (2016) introduced a new species, P. diversisporus from human toe nail with two new combinations (P. atro-olivaceus and P. karoo). Hyde et al. (2017) introduced a new species, P. kunmingensis from China. Wanasinghe et al. (2018) introduced P. rosae from Italy and Tibpromma et al. (2018) introduced P. pandanicola from Thailand, while Crous et al. (2018a) introduced P. angolensis from leaf spot of unknown host plant in Angola. Jayasiri et al. (2019) introduced a new species, P. entadae from pod of Entada phaseoloides in Thailand. We present a new host record for P. palmicola isolated from C. odorata, together with a description and illustrations (Figs. 27, 28). A phylogenetic tree based on combined ITS, LSU, RPB2 and GAPDH sequence data is presented in Fig. 26. Pseudopithomyces palmicola J.F. Li, Ariyaw. & K.D. Hyde, in Ariyawansa et al., Fungal Divers. 75: 41 (2015) Facesoffungi number: FoF 00939; Fig. 27 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Hyphomycetous. Colonies superficial or partly immersed on the host surface, scattered, gregarious, dark brown to black. Mycelium consisting of septate, branched, smooth, thin-walled, hyaline hyphae. Conidiophores reduced to conidiogenous cells. Conidiogenous cells (4–)6.5–11(–15) × (1.9–)2.5–4 µm ( x̄ = 9 × 3 µm, n = 15), holoblastic, monoblastic, terminal, hyaline, cylindrical. Conidia (6.5–)17–30 × (4.5–)8–25 µm ( x̄ = 23.5 × 16 µm, n = 50), globose or ellipsoid to amygdaloid, muriform, 2–3-transversely septate, with 1–2-vertical septa, slightly verruculose to echinulate, brown to dark brown, sometimes slightly constricted at the septa with dark bands at the septa. Culture characteristics: Conidia germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, velvety with moderately fluffy, entire, cultures pale brown to creamy brown on surface from the centre of the colony with white margin, and brown to dark brown in reverse from the centre of the colony with creamy white margin (Fig. 28). Pre-screening for antimicrobial activity: Pseudopithomyces palmicola (MFLUCC 17-1506) showed antimicrobial activity against M. plumbeus with a 10 mm inhibition zone, observable as partial inhibition, when compared to the positive control (18 mm), but no inhibition of B. subtilis and E. coli. Fungal Diversity Fig. 26 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS, LSU, RPB2 and GAPDH sequence data. Fortytwo strains are included in the combined sequence analysis, which comprise 3066 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 10725.898987 is presented. The matrix had 826 distinct alignment patterns, with 28.11% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.237796, C = 0.263846, G = 0.271148, T = 0.227210; substitution rates: AC = 1.281129, AG = 2.845304, AT = 1.125930, CG = 1.002622, CT = 6.844818, GT = 1.000000; gamma distribution shape parameter α = 0.149330. 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 dark red bold and type species are in bold. Laburnicola hawksworthii (MFLUCC 13-0602) and L. muriformis (MFLUCC 16-0290) are used as outgroup taxa Known hosts and distribution: On leaves lesions of Vitis vinifera (Vitaceae), Phaseolus vulgaris (Fabaceae), Poa annua (Poaceae) and Fragaria sp. (Rosaceae) in Italy (Liu et al. 2018); on dead leaves of Pandanus amaryllifolius (Pandanaceae) in Thailand (Tibpromma et al. 2018); on dead leaves of unidentified grass in China (Hyde et al. 2017b); on leaves of Acoelorrhaphe wrightii (Arecaceae) in Thailand (Ariyawansa et al. 2015). Material examined: THAILAND, Chiang Mai Province, Fah Hom Pok, on dead stems of C. odorata, 27 September 2016, A. Mapook (FHP10, MFLU 20-0355); living culture MFLUCC 17-1506 (new host record); Chiang Rai Province, Doi Pui, on dead stems of C. odorata, 2 February 2017, A. Mapook (DP95, MFLU 20-0354); living culture MFLUCC 17-1496. GenBank numbers: LSU: MT214447, MT214448, ITS: MT214353, MT214354, RPB2: MT235811, MT235812 13 Fungal Diversity Fig. 27 Pseudopithomyces palmicola (new host record) a, b Appearance of colonies on substrate. c–f Conidia and conidiophores. g–l Conidia. Scale bars: a = 200 µm, b = 100 µm, d–l = 10 µm, c = 5 µm Notes: A phylogenetic analyses showed that two strains MFLUCC 17-1506 and MFLUCC 17-1496 grouped within the Pseudopithomyces palmicola clade (Fig. 26). In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of our strains with 100% similarity was Pseudopithomyces palmicola (strain DTO 391-A6, MN788110). The closest match with the LSU sequence with 100% similarity was Pseudopithomyces sp. (strain C449, MK348017), while the closest match with the RPB2 sequence with 100% similarity was Pseudopithomyces palmicola (strain UC15, MH249026). We therefore, identify our isolates as P. palmicola based on phylogenetic analyses. Morphological characters also indicated that our strains belong to the genus Pseudopithomyces. In this study, we isolated P. palmicola from Chromolaena odorata collected in Thailand, and the isolates are introduced here as a new host record. We also treat P. pandanicola (MFLUCC 18-0116) and P. kunmingensis (MFLUCC 17-0314) under P. palmicola based on phylogeny with morphological comparison (Table 7). 13 Tremateia Kohlm., Volkm.-Kohlm. & O.E. Erikss., Bot. Mar. 38(2): 165 (1995) Tremateia was introduced by Kohlmeyer et al. (1995) with T. halophila as the type species. Hyde et al. (2016b) introduced two new species T. arundicola and T. guiyangensis which were collected from dead herbaceous stems, based on morphology and phylogeny support. Feng et al. (2019) introduced a new species, T. murispora from China. We introduce two new Tremateia species from C. odorata, together with descriptions and illustrations (Figs. 30, 31, 32). A phylogenetic tree based on combined SSU, LSU, ITS, TEF1 and RPB2 sequence data is presented in Fig. 29. Tremateia chiangraiensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557364, Facesoffungi numberx: FoF 07793; Fig. 30 Etymology: Referring to the location where the specimen was collected, Chiang Rai Province, Thailand. Holotype: MFLU 20-0374 Fungal Diversity Fig. 28 Culture characteristic on MEA: Pseudopithomyces palmicola (MFLUCC 17-1506) Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 220–250 µm high × 215–260(–280) µm diam. ( x̄ = 230 × 240 µm, n = 10), immersed, solitary or scattered, appearing as small dark spots, coriaceous, globose, reddish brown to brown. Ostiolar neck protruding. Peridium 10–20(–25) µm wide, 4–5 layers, inner layers comprising hyaline to pale brown cells of textura epidermoidea, outer layers comprising brown to dark brown cells of textura angularis. Hamathecium comprising (1.8–)2–3 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 90–120(–140) × 14–20 µm ( x̄ = 106.5 × 17.5 µm, n = 15), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, pedicellate. Ascospores 23–27.5 × 5–7 µm ( x̄ = 25 × 9 µm, n = 35), overlapping, 1–2 seriate, initially hyaline, 1-septate when immature, becoming golden-brown to brown at maturity, ellipsoid to broadly fusiform, muriform, 5–7-transversely septate, with 1 vertical septum, slightly constricted at the central septum, straight or slightly curved, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from several cells. Colonies on MEA circular, mycelium slightly raised, filamentous, cultures white on surface, white to creamy white in reverse (Fig. 33a). Pre-screening for antimicrobial activity: Tremateia chiangraiensis (MFLUCC 17-1428) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Huai Kang Pla waterfall, on dead stems of Chromolaena odorata, 23 June 2015, A. Mapook (HKP4, MFLU 20-0374, holotype); ex-type culture MFLUCC 17-1428, (HKP5, MFLU 20-0375); living culture MFLUCC 17-1429. GenBank numbers: LSU: MT214449, MT214450, ITS: MT214355, MT214356, SSU: MT214401, MT214402, TEF1: MT235775, MT235776, RPB2: MT235813, MT235814 Notes: In a BLASTn search of NCBI GenBank, the ITS sequence of Tremateia chiangraiensis (MFLUCC 17-1428, ex-holotype) with 97.66% similarity was T. guiyangensis (strain GZAAS01, KX274240). The LSU and SSU sequences with 97.17% (KX274248) and 97.51% (KX274254) similarity, respectively, were T. arundicola strain MFLUCC 16-1275, while the closest match with the TEF1 and RPB2 sequences with 97.90% (LT797094) and 90.02% (LT797014) similarity, respectively, were Kalmusia Table 7 Morphology of Pseudopithomyces species with similar morphological features discussed in this study Species Conidiogenous cells (µm) Conidia (µm) Conidia septation P. chartarum (MUCL 15905) P. chartarum (UTHSC 05-2460, UTHSC 03-2472) P. palmicola (MFLU 15-1474, holotype) P. palmicola (CBS 143933, CBS 143932, CBS 143935 and CBS 143934) P. palmicola (MFLUCC 17-1496 and MFLUCC 17-1496) P. palmicola (= P. pandanicola MFLUCC 18-0116) P. palmicola (= P. kunmingensis MFLUCC 17-0314) – – 18–29 × 10–17 12–33 × 9–19 3–4-transverse 1–2 vertical Ariyawansa et al. (2015) 3 transverse, 1–2 vertical da Cunha et al. (2014) 3.5–5.5 × 2.5–3.5 21.5–30.5 × 10–16.5 2–3 transverse, 1–3 vertical Ariyawansa et al. (2015) – 3 transverse, 2 vertical (11.5–)21.1(– 27.5) × (7.5–)11.7(– 16.5) (6.5–)17–30 × (4.5–)8–25 2–3 transverse 1–2 vertical (4–)6.5–11(–15) × (1.9– )2.5–4 References Liu et al. (2018) This study – 10–25 × 7–15 2–3 transverse, 1–2 vertical Tibpromma et al. (2018) 3–6 × 5–6 19–25 × 10–15 2–3 transverse, 1–2 vertical Hyde et al. (2017b) 13 Fungal Diversity Fig. 29 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS, LSU, SSU, TEF1 and RPB2 sequence data. Twelve strains are included in the combined sequence analysis, which comprise 4634 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 10566.353641 is presented. The matrix had 559 distinct alignment patterns, with 22.71% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.235911, C = 0.270697, G = 0.272162, T = 0.221230; substitution rates: AC = 1.376817, AG = 3.496002, AT = 1.093429, CG = 1.659156, CT = 6.548636, GT = 1.000000; gamma distribution shape parameter α = 0.147928. 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 dark red bold and type species are in bold. Bimuria novae-zelandiae (CBS 107.79) is used as outgroup taxon sp. strain UTHSC: DI16-256. In the present phylogenetic analysis, our two strains of T. chiangraiensis form a sister taxon with T. thailandensis (MFLUCC 17-1430) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 29). However, T. chiangraiensis differs from T. thailandensis in having larger ascomata (220–250 × 215–260(–280) µm vs. 200–230 × 150–175 µm) and slightly smaller asci (90–120(–140) × 14–20 µm vs. (80–)90–125 × 15–25 µm) and slightly smaller ascospores (23–27.5 × 5–7 µm vs. 20–26 × 5.5–9 µm) (Table 8). A comparison of the ITS (+5.8S) gene region of T. chiangraiensis and T. thailandensis reveals 17 base pair differences (3%) across 565 nucleotides. Therefore, T. chiangraiensis is described as a new species based on phylogeny and morphological comparison. Tremateia chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557365, Facesoffungi number: FoF 07794; Fig. 31 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. 13 Holotype: MFLU 20-0378 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 150–220 µm high × 170–200(–220) µm diam. ( x̄ = 183 × 191.5 µm, n = 10), immersed, solitary or scattered, appearing as small dark spots, coriaceous, globose, dark brown to brown. Ostiolar neck protruding. Peridium 10–15 µm wide, comprising 2–3 layers, brown to dark brown cells of textura angularis. Hamathecium comprising 2–3(–3.5) µm wide, broadly cylindrical, septate, branching pseudoparaphyses. Asci (65–)75–105 × (10–)15–20 µm ( x̄ = 89 × 15.5 µm, n = 15), 8-spored, bitunicate, fissitunicate, clavate to cylindric-clavate, straight or slightly curved, pedicellate, apically rounded. Ascospores 16.5–20 × 6.5–10 µm ( x̄ = 18 × 8 µm, n = 30), overlapping, 1–2-seriate, initially hyaline, 1-septate when immature, becoming golden-brown at maturity, oval to ellipsoidal, muriform, 3–4-transversely septate, with 1–2 vertical septa, upper part wider, slightly constricted at the central septum, straight or slightly curved, guttulate, surrounded by hyaline gelatinous sheath observed Fungal Diversity Fig. 30 Tremateia chiangraiensis (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores q Ascospore with gelatinous sheath in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c, d, g–j = 50 µm, e = 20 µm, k–q = 10 µm, f = 5 µm clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from several cells. Colonies on MEA circular, mycelium slightly raised, initially white, filamentous, becoming pale pinkish white on surface, brown to pale brown in reverse from the centre of the colony, white to creamy white at margin (Fig. 33b). Pre-screening for antimicrobial activity: All isolates of Tremateia chromolaenae showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP26, MFLU 20-0376); living culture MFLUCC 17-1442; Chiang Rai Province, Doi Pui, on dead stems of C. odorata, 8 July 2015, A. Mapook (DP8, MFLU 20-0377); living culture MFLUCC 17-1424; (DP9, MFLU 20-0378, holotype); ex-type culture MFLUCC 17-1425; Lampang Province, Chaehom, on dead stems of C. odorata, 13 Fungal Diversity Fig. 31 Tremateia chromolaenae (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores q Ascospore with gelatinous sheath in Indian ink. Scale bars: a, b = 500 µm, c, g–j = 50 µm, d = 20 µm, e, k–q = 10 µm, f = 5 µm 24 September 2016, A. Mapook (JH4, MFLU 20-0379); living culture MFLUCC 17-1458. GenBank numbers: LSU: MT214451, MT214452, MT214453, MT214454, ITS: MT214357, MT214358, MT214359, MT214360, SSU: MT214403, MT214404, MT214405, MT214406, TEF1: MT235777, MT235778, MT235779, MT235780, RPB2: MT235815, MT235816, MT235817, MT235818 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS and SSU sequences of Tremateia chromolaenae (MFLUCC 17-1442, ex-holotype) with 99.34% (KY038612) and 99.90% (KX274253) similarity, respectively, was Tremateia guiyangensis strain GZAAS01. The closest match with the LSU sequence with 99.88% similarity 13 was T. arundicola (strain MFLUCC 16-1275, KX274248), while the closest match with the TEF1 and RPB2 sequences with 97.58% (LT797094) and 92.31% (LT797014) similarity, respectively, was Kalmusia sp. strain UTHSC: DI16256. In the present phylogenetic analysis, four strains of T. chromolaenae form a separate clade and cluster with T. guiyangensis (GZAAS01) and T. arundicola (MFLUCC 16-1275), with high bootstrap support (Fig. 29). However, T. chromolaenae differs from T. guiyangensis and T. arundicola in having smaller ascomata (150–220 × 170–200(–220) µm vs. 130–280 × 190–400 µm and 200–300 × 250–350 µm), smaller asci [(65–)75–105 × (10–)15–20 µm vs. 152–160 × 21–27 µm and 170–200 × 15–20 µm] and smaller ascospores (16.5–20 × 6.5–10 µm vs. 20–28 × 9–12 µm Fungal Diversity Fig. 32 Tremateia thailandensis (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–o Ascospores p Ascospore with gelatinous sheath in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c, g–j = 50 µm, d = 20 µm, e, f, k–p = 10 µm Fig. 33 Culture characteristic on MEA: a Tremateia chiangraiensis (MFLUCC 17-1428). b Tremateia chromolaenae (MFLUCC 17-1424). c Tremateia thailandensis (MFLUCC 17-1430) 13 Fungal Diversity Table 8 Synopsis of Tremateia species with similar morphological features discussed in this study Species Ascomata (µm) Peridium (µm) Asci (µm) Ascospores (µm) Transverse septa References T. arundicola (MFLUCC 16-1275) T. chiangraiensis (MFLUCC 17-1428) T. chromolaenae (MFLUCC 17-1442) T. guiyangensis (GZAAS01) T. thailandensis (MFLUCC 17-1430) 200–300 high × 250– 350 diam. 220–250 high × 215– 260(–280) diam. 150–220 high × 170– 200(–220) diam. 130–280 high × 190– 400 diam. 200–230 high × 150– 175 diam. 10–20 170–200 × 15–20 20–30 × 6–9 3–6 Hyde et al. (2016b) 10–20(–25) 90–120(–140) × 14–20 23–27.5 × 5–7 5–7 This study 10–15 16.5–20 × 6.5–10 3–4 This study Up to 9–16 (65–)75–105 × (10– )15–20 152–160 × 21–27 20–28 × 9–12 3–5 Hyde et al. (2016b) 10–25 (80–)90–125 × 15–25 20–26 × 5.5–9 4–5 This study and 20–30 × 6–9 µm) (Table 8). A comparison of the ITS (+5.8S) gene region of T. chromolaenae and T. guiyangensis reveals 126 base pair differences (22.3%) across 565 nucleotides, and a comparison of the ITS (+5.8S) gene region of T. chromolaenae and T. arundicola reveals 136 base pair differences (22.9%) across 593 nucleotides. Therefore, T. chromolaenae is described as a new species based on phylogeny and morphological comparison. Tremateia thailandensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557366, Facesoffungi number: FoF 07795; Fig. 32 Etymology: Named after Thailand, where the fungus was first discovered. Holotype: MFLU 20-0380 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 200–230 µm high × 150–175 µm diam. ( x̄ = 220 × 166 µm, n = 5), immersed, solitary or scattered, coriaceous, subglobose to obpyriform, dark brown to brown. Ostiolar neck protruding. Peridium 10–25 µm wide, comprising 2–4 layers, brown to dark brown cells of textura angularis. Hamathecium comprising 2–3.5 µm wide, broadly cylindrical, septate, branching pseudoparaphyses. Asci (80–)90–125 × 15–25 µm ( x̄ = 103.5 × 18 µm, n = 30), 8-spored, bitunicate, fissitunicate, clavate to cylindric-clavate, straight or slightly curved, apically rounded, pedicellate. Ascospores 20–26 × 5.5–9 µm ( x̄ = 23 × 8 µm, n = 35), overlapping, 2 seriate, initially hyaline, 1-septate when immature, becoming golden-brown to brown at maturity, ellipsoid to broadly fusiform, muriform, 4–5-transversely septate, with 1 vertical septum, upper part wider and shorter, slightly constricted at the central septum, straight or slightly curved, guttulate, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes 13 produced from several cells. Colonies on MEA circular, mycelium slightly raised, filamentous, cultures white on surface, white to creamy white in reverse (Fig. 33c). Pre-screening for antimicrobial activity: Tremateia thailandensis (MFLUCC 17-1430) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Huai Kang Pla waterfall, on dead stems of Chromolaena odorata, 23 June 2015, A. Mapook (HKP6, MFLU 20-0380, holotype); ex-type culture MFLUCC 17-1430. GenBank numbers: LSU: MT214455, ITS: MT214361, SSU: MT214407, TEF1: MT235781, RPB2: MT235819 Notes: In a BLASTn search of NCBI GenBank, the ITS sequence of Tremateia thailandensis (MFLUCC 17-1430, ex-holotype) with 95.48% similarity was T. guiyangensis (strain GZAAS01, KX274240). The LSU and SSU sequences with 97.29% (KX274248) and 96.66% (KX274254) similarity, respectively, was T. arundicola strain MFLUCC 16-1275, while the closest match with the TEF1 and RPB2 sequences with 97.77% (LT797094) and 90.00% (LT797014) similarity, respectively, was Kalmusia sp. strain UTHSC: DI16-256. In the present phylogenetic analysis, T. thailandensis clusters with T. chiangraiensis with high bootstrap support (100% ML and 1.00 BYPP, Fig. 29). However, T. thailandensis differs from T. chiangraiensis in having smaller ascomata (200–230 × 150–175 µm vs. 220–250 × 215–260(–280) µm) and slightly larger asci [(80–)90–125 × 15–25 µm vs. 90–120(–140) × 14–20 µm] and ascospores (20–26 × 5.5–9 µm vs. 23–27.5 × 5–7 µm) (Table 8). A comparison of the ITS (+5.8S) gene region of T. thailandensis and T. chiangraiensis reveals 17 base pair differences (3%) across 565 nucleotides. Therefore, T. thailandensis is described as a new species based on phylogeny and morphological comparison. Lophiostomataceae Sacc. Lophiostomataceae was introduced by Saccardo (1883) with Lophiostoma as the type genus. The family contains Fungal Diversity saprobes found on woody plants from terrestrial, freshwater, and marine habitats, as well as on herbaceous twigs and pods (Thambugala et al. 2015; Devadatha et al. 2017; Hashimoto et al. 2018; Tennakoon et al. 2018; Jayasiri et al. 2019). Thambugala et al. (2015) accepted 16 genera in the family, which included eleven new genera (Alpestrisphaeria, Biappendiculispora, Capulatispora, Coelodictyosporium, Dimorphiopsis, Guttulispora, Lophiohelichrysum, Lophiopoacea, Lophiostoma, Neotrematosphaeria, Paucispora, Platystomum, Pseudolophiostoma, Pseudoplatystomum, Sigarispora, Vaginatispora). Wanasinghe et al. (2018) introduced a new genus Neopaucispora based on morphology and phylogeny, while Hashimoto et al. (2018) introduced seven new genera (Crassiclypeus, Flabellascoma, Lentistoma, Leptoparies, Neovaginatispora, Parapaucispora, Pseudopaucispora) based on morphological observations and phylogenetic analyses. Divergence time estimates for the family was stem age at 130 Mya (87–178) in the Cretaceous period (Liu et al. 2017). Fig. 34 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU, RPB2 and TEF1 sequence data. Seventytwo strains are included in the combined sequence analysis, which comprise 5256 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 30152.374563 is presented. The matrix had 1782 distinct alignment patterns, with 29.14% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.251190, C = 0.244578, G = 0.267209, T = 0.237023; substitution rates: AC = 1.507173, AG = 4.140389, AT = 1.272744, CG = 1.390194, CT = 8.158020, GT = 1.000000; gamma distribution shape parameter α = 0.167204. 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 dark red bold and type species are in bold. Teichospora rubriostiolata (TR 7) and T. trabicola (C 134) are used as outgroup taxa 13 Fungal Diversity Fig. 35 Flabellascoma minimum (new host record) a Appearance of ascomata on substrate. b Section through ascoma. c Ostiole. d Peridium. e Pseudoparaphyses. f–g Asci. h–k Ascospores. Scale bars: a = 200 µm, b = 50 µm, c, d, f, g = 20 µm, h–k = 10 µm, e = 5 µm Flabellascoma A. Hashim., K. Hiray. & Kaz. Tanaka (2018) Flabellascoma was introduced by Hashimoto et al. (2018) to accommodate two new species, F. cycadicola and F. minimum the type species. Jayasiri et al. (2019) introduced a new host record for F. minimum from fallen pods of Leucaena leucocephala in Thailand. In this study, we record F. minimum on Siam weed (Fig. 35). A phylogenetic tree based on combined LSU, ITS, SSU, RPB2 and TEF1 sequence data is presented in Fig. 34. 13 Flabellascoma minimum A. Hashim., K. Hiray. & Kaz. Tanaka, in Hashimoto et al., Studies in Mycology 90: 169 (2018) Facesoffungi number: FoF 05263; Fig. 35 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 210–220(–235) µm high × 165–190(–260) µm diam. ( x̄ = 221 × 193 µm, n = 5), immersed, appearing as black spots, coriaceous, solitary or scattered, globose to obpyriform, brown. ostiole long neck, carbonaceous, Fungal Diversity papillate, crest-like, elongated and laterally compressed, with hyaline periphyses. Peridium 15–20(–25) µm wide, comprising several layers, outer layers comprising brown to dark brown cells of textura prismatica, inner layers comprising light brown to hyaline cells of textura angularis. Hamathecium comprising 1–2(–2.5) µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci (50–)60–75(–85) × 7–10 µm ( x̄ = 66.5 × 8.5 µm, n = 15), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short pedicel, apically rounded, with an ocular chamber. Ascospores 4–16 × 4.5–6 µm ( x̄ = 15 × 5 µm, n = 20), overlapping, 1–2-seriate, hyaline, broadly fusiform with obtuse ends, uniseptate, straight or slightly curved, guttulate, constricted at the septum; sheath drawn out to form polar appendages, (4–)6–9 µm long × 1.5–2.3 µm wide ( x̄ = 7 × 2 µm, n = 30), from apex of ascospores, with a lateral pad-like structure. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature, at first becoming 4 celled and then germ tubes produced from both ends of the ascospores. Colonies on MEA filamentous, mycelium slightly raised, flattened, filiform, white aerial hyphae, spreading from the center, becoming grayish to light brown at the surface and dark olivaceous-brown to black in reverse from the center with creamy white at rim (Fig. 36). Pre-screening for antimicrobial activity: Flabellascoma minimum (MFLUCC 17-1474) showed antimicrobial activity against B. subtilis with a 15 mm inhibition zone and against M. plumbeus with a 17 mm inhibition zone, observable as partial inhibition, when compared to the positive control (26 mm and 17 mm, respectively), but no inhibition of E. coli. Known hosts and distribution: On petioles of Arenga engleri (Arecaceae), on pods of Bauhinia purpurea (Fabaceae) in Taiwan (Hashimoto et al. 2018); on pods of Fig. 36 Culture characteristic on MEA: Flabellascoma minimum (MFLUCC 17-1474) Leucaena leucocephala (Fabaceae) in Thailand (Jayasiri et al. 2019). Material examined: THAILAND, Nan Province, Doi Phu Kha, on dead stems of Chromolaena odorata, 23 September 2016, A. Mapook (DPK6, MFLU 20-0312); living culture MFLUCC 17-1474 (new host record). GenBank numbers: LSU: MT214461, ITS: MT214367, SSU: MT214414, TEF1: MT235782, RPB2: MT235820 Notes: A phylogenetic analyses show the strain MFLUCC 17-1474 grouped with Flabellascoma minimum (Fig. 34). In a BLASTn search of NCBI GenBank, the closest match of the ITS and RPB2 sequences of MFLUCC 17-1474 with 100% (LC312503) and 99.61% (LC312590) similarity was F. minimum strain KT 2013, while the closest match with the LSU, SSU and TEF1 sequences with 99.65% (MK347975), 99.72% (MK347865) and 98.78% (MK360054) similarity was F. minimum strain MFLUCC 18-0233. Therefore, we identify our isolate as F. minimum based on the phylogenetic analysis. Morphological characters also indicated that our strain is this species. We isolated F. minimum from Chromolaena odorata collected in Thailand, and the isolate is introduced here as a new host record. Pseudocapulatispora Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557285, Facesoffungi number: FoF 07796 Etymology: Referring to its similarity with Capulatispora. Saprobic on dead stems. Sexual morph: Ascomata immersed, coriaceous, solitary or scattered, ovoid, light brown to brown. Ostiole long neck, carbonaceous, papillate, crest-like, elongated and laterally compressed, with hyaline periphyses. Peridium comprising several layers, outer layers comprising light brown to brown cells of textura prismatica, inner layers comprising pale brown to hyaline cells of textura angularis. Hamathecium comprising cylindrical to filiform, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, fissitunicate, cylindric-clavate, with a short pedicel, apically rounded, with an ocular chamber. Ascospores overlapping, 1–2-seriate, hyaline, broadly fusiform, uniseptate, with a narrow sheath; sheath drawn out to form polar appendages from apex of ascospores. Asexual morph: Undetermined. Type species: Pseudocapulatispora longiappendiculata Mapook & K.D. Hyde Notes: A phylogenetic analyses based on combined dataset of LSU, ITS, SSU, RPB2 and TEF1 sequence data show that two strains of Pseudocapulatispora longiappendiculata form a monophyletic clade within the family Lophiostomataceae with high bootstrap support (83% ML and 0.99 BYPP, Fig. 34). Pseudocapulatispora species are also morphologically distinct in having a long sheath drawn out to form polar appendages from apex of ascospores with tips of the sheath 13 Fungal Diversity Fig. 37 Pseudocapulatispora longiappendiculata (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores. Scale bars: a = 500 µm, b = 200 µm, c, g–j, l–p = 50 µm, d, k = 20 µm, e = 10 µm, f = 5 µm capped. Therefore, we introduce Pseudocapulatispora as a new genus with a new species P. longiappendiculata, based on morphology and phylogeny. Pseudocapulatispora longiappendiculata Mapook & K.D. Hyde, sp. nov. 13 Index Fungorum number: IF557286, Facesoffungi number: FoF 07797; Fig. 37 Etymology: Referring to ascospores with long polar appendages. Holotype: MFLU 20-0347 Fungal Diversity Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (195–)250–265 µm high × 105–150 µm diam. ( x̄ = 240 × 135 µm, n = 5), immersed, appearing as black spots, coriaceous, solitary or scattered, ovoid, light brown to brown. Ostiole long neck, carbonaceous, papillate, crest-like, elongated and laterally compressed, with hyaline periphyses. Peridium 15–25 µm wide, comprising several layers, outer layers comprising light brown to brown cells of textura prismatica, inner layers comprising pale brown to hyaline cells of textura angularis. Hamathecium comprising (1–)2–3 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 75–120 × 14–20 µm ( x̄ = 100 × 15.5 µm, n = 18), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short pedicel, apically rounded, with an ocular chamber. Ascospores 24–29 × 6–9 µm ( x̄ = 26 × 7.5 µm, n = 20), overlapping, 1–2-seriate, hyaline, broadly fusiform, uniseptate, straight or slightly curved, guttulate, constricted at the septum, with a narrow sheath; sheath drawn out to form polar appendages, (5.5–)10–40 µm long × (2.5–)3–5(–9) µm wide ( x̄ = 25 × 4.5 µm, n = 30), from apex of the ascospores terminating in a glose drop at their tips. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature, at first becoming 4 celled and then germ tubes produced around the ascospores. Colonies on MEA, initial irregular, becoming filamentous in old culture, mycelium slightly raised, flattened, filiform, white aerial hyphae, spreading from the center, becoming olivaceous-brown at the surface and olivaceous-brown to black in reverse, MEA change to yellow (Fig. 38). Pre-screening for antimicrobial activity: Pseudocapulatispora longiappendiculata (MFLUCC 17-1452) showed antimicrobial activity against B. subtilis with a 7 mm inhibition zone and against M. plumbeus with a 23 mm inhibition zone, observable as partial inhibition, when compared to the Fig. 38 Culture characteristic on MEA: Pseudocapulatispora longiappendiculata (MFLUCC 17-1452) positive control (26 mm and 20 mm, respectively), but no inhibition of E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP45, MFLU 20-0347, holotype); extype culture MFLUCC 17-1452; Lampang Province, Chaehom, on dead stems of Chromolaena odorata, 24 September 2016, A. Mapook (JH3, MFLU 20-0348); living culture MFLUCC 17-1457. GenBank numbers: LSU: MT214462, MT214463, ITS: MT214368, MT214369, SSU: MT214415, MT214416, TEF1: MT235783, MT235784, RPB2: MT235821 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Pseudocapulatispora longiappendiculata (MFLUCC 17-1452, ex-holotype) with 88.99% similarity was Lophiostoma sp. (strain MA 4558, AJ972793). The closest match with the LSU and SSU sequences with 98.02% (AY016369) and 99.37% (AY016354) similarity, respectively, were Lophiostoma heterospora strain CBS 644.86, while the closest match with the TEF1 sequences with 96.04% similarity was Trematosphaeria sp. (LT797069). Nigrogranaceae Jaklitsch & Voglmayr Nigrogranaceae was introduced by Jaklitsch and Voglmayr (2016) to accommodate the monotypic genus Nigrograna, which is reported as ecologically diverse (Kolařík et al. 2017; Kolařík 2018) and contains 11 epithets (Index Fungorum 2019). The divergence time estimates for this family are crown age of 72 Mya (44–124) and stem age of 131 Mya (86–180) during Cretaceous period (Lui et al. 2017). Nigrograna Gruyter et al. Nigrograna is an ecologically diverse genus, comprising plant endophytes, saprobes, marine or estuarine and human mycetoma (Jaklitsch and Voglmayr 2016; Ahmed et al. 2018; Hyde et al. 2017b; Kolařík et al. 2017; Tibpromma et al. 2017; Kolařík 2018). The genus was described by de Gruyter et al. (2013) with the type species, N. mackinnonii. It has been considered as a synonym of Biatriospora by Ahmed et al. (2014), based on the sequence related to B. marina. Kolařík et al. (2017) described four new endophytic species of Biatriospora (B. antibiotica, B. carollii, B. peruviensis, B. yasuniana) from woody plants in temperate forests and in tropical regions based on morphology and multigene analyses. These species were synonymized under Nigrograna by Kolařík et al. (2018). Subsequently, Zhao et al. (2018) introduced a new species, N. locuta-pollinis from hive-stored pollen in China. In this study, we introduce a new species, Nigrograna chromolaenae, based on morphology and molecular data, and provide a description and illustrations (Fig. 40). A phylogenetic tree based on 13 Fungal Diversity Fig. 39 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU and TEF1 sequence data. Thirtytwo strains are included in the combined sequence analysis, which comprise 3190 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 7547.223829 is presented. The matrix had 431 distinct alignment patterns, with 29.75% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.245036, C = 0.242903, G = 0.269102, T = 0.242959; substitution rates: AC = 1.661081, AG = 2.575440, AT = 1.432192, CG = 0.774269, CT = 13.455559, GT = 1.000000; gamma distribution shape parameter α = 0.020000. 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 dark red bold and type species are in bold. Seriascoma didymospora (MFLUCC 11-0194) and S. didymospora (MFLUCC 11-0179) are used as outgroup taxa combined LSU, ITS, SSU and TEF1 sequence data is presented in Fig. 39. Nigrograna chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557345, Facesoffungi number: FoF 07297; Fig. 40 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0341 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 160–280 µm high × 115–130(–150) µm diam. ( x̄ = 205 × 128 µm, n = 5), immersed to erumpent, appearing as black spots, coriaceous, solitary or scattered, 13 subglobose to obpyriform, brown to dark brown. Ostiole long papillate, with hyaline periphyses. Peridium 15–25 µm wide, comprising several layers, thick-walled cells, outer layers comprising brown to dark brown cells of textura angularis, inner layers comprising light brown to hyaline cells of textura angularis. Hamathecium comprising 1–2(–2.5) µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 40–55(–65) × 7–10 µm ( x̄ = 50 × 8.5 µm, n = 15), 8-spored, bitunicate, fissitunicate, cylindricclavate, straight or slightly curved, with a short, bulbous pedicel, apically rounded. Ascospores 10.5–15 × 3.5–5 µm ( x̄ = 12 × 4 µm, n = 30), overlapping, 1–2-seriate, greyish brown to dark brown, broadly fusiform to inequilateral, Fungal Diversity Fig. 40 Nigrograna chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores. Scale bars: a = 500 µm, b = 200 µm, c = 50 µm, d, g–j = 20 µm, e = 10 µm, f, k–p = 5 µm with slightly obtuse ends, upper part or second cell slightly wider, 1-septate when immature, becoming 3-euseptate when mature, slightly constricted at the primary median septum, straight or slightly curved, guttulate, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from several cells. Colonies on MEA irregular, mycelium slightly raised, entire, white aerial hyphae at first, spreading from the center at the surface, becoming olivaceous-grey to brown and dark olivaceous-grey to black in reverse (Fig. 41). Pre-screening for antimicrobial activity: Nigrograna chromolaenae (MFLUCC 17-1437) showed antimicrobial activity against M. plumbeus with an 11 mm inhibition zone, observable as partial inhibition, when compared to the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 13 Fungal Diversity Fig. 41 Culture characteristic on MEA: Nigrograna chromolaenae (MFLUCC 17-1437) 2015, A. Mapook (DP21, MFLU 20-0341, holotype); extype culture MFLUCC 17-1437. GenBank numbers: LSU: MT214473, ITS: MT214379, TEF1: MT235801 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Nigrograna chromolaenae (MFLUCC 17-1437, ex-holotype) with 99.32% similarity was Nigrograna mackinnonii (strain L3396, KP263091), while the closest match with the LSU and TEF1 sequences with 100% (LN907384) and 100% (LT797087) similarity, respectively, was N. mackinnonii strain UTHSC: DI16241. In the present phylogenetic analysis, N. chromolaenae clusters with N. marina strain CY 1228 with low bootstrap support (Fig. 39). However, N. chromolaenae differs from N. marina in morphology, N. chromolaenae having broadly fusiform to inequilateral ascospores, with slightly obtuse ends, upper part or second cell slightly wider, 1-septate when immature, becoming 3-euseptate when mature, slightly constricted at the primary median septum, guttulate without terminal appendages, while N. marina has unusual ascospores that are dark brown, fusiform with hyaline, rounded swollen tips at both ends and 1–4 septa situated near the ends (Hyde et al. 2013). A comparison of the ITS (+5.8S) gene region of N. chromolaenae and N. marina reveals 15 base pair differences (1.77%) across 847 nucleotides. Therefore, N. chromolaenae is described as a new species based on phylogeny and morphological comparison. Neomassarinaceae Mapook & K.D. Hyde, fam. nov. Index Fungorum number: IF557341, Facesoffungi number: FoF 07798 Etymology: Referring to the type genus, Neomassarina Saprobic on dead leaf or stems of herbaceous plant. Sexual morph: Ascomata immersed or semi-immersed to erumpent, globose to subglobose or obpyriform, light brown to brown, coriaceous, solitary or scattered. Ostiole long neck, carbonaceous, papillate, crest-like, elongated and 13 laterally compressed, with or without hyaline periphyses. Peridium comprising 2–5 layers, pale brown to hyaline cells or dark brown to black cells of textura angularis. Hamathecium composed of cylindrical to filiform, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, fissitunicate, cylindrical to cylindric-clavate, with a short pedicel, apically rounded, with an ocular chamber. Ascospores overlapping, 1–2-seriate, hyaline to pale brown at maturity, fusiform, uniseptate, with a narrow sheath or surrounded by hyaline gelatinous sheath; sheath drawn out to form polar appendages from both ends of the ascospores, straight or slightly curved. Asexual morph: Undetermined. Type genus: Neomassarina Phook., Jayasiri & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 136 (2016) Notes: Based on a phylogenetic tree of combined LSU, ITS, SSU, TEF1 and RPB2 sequence data, Neomassarinaceae forms a distinct family in the clade comprising Sporormiaceae and Amorosiaceae (Fig. 42). Neomassarinaceae differs from Sporormiaceae in morphology. The new family has similarity with Amorosiaceae in its crestlike ostiole, cylindrical to cylindric-clavate asci with an ocular chamber and surrounded by a mucilaginous sheath (Thambugala et al. 2015; Tibpromma et al. 2017). However, they differ in crest-like, elongated and laterally compressed, carbonaceous ostiolar necks, with or without hyaline periphyses. Neomassarina Phook., Jayasiri & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 136 (2016) Neomassarina was introduced by Hyde et al. (2016b) with Neomassarina thailandica as the type species collected from Agave angustifolia (Asparagaceae) in Thailand and placed in Pleosporales genera incetae sedis based on phylogenetic analyses. Hyde et al. (2018a, b) included the genus in Sporormiaceae and introduced a new species, N. pandanicola from Pandanus sp. (Pandanaceae) in Thailand, based on both morphology and phylogenetic support. In this study, a new species, Neomassarina chromolaenae is introduced together with a new host record for N. thailandica, together with descriptions and illustrations (Figs. 43, 44). A phylogenetic tree based on combined LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 42. Neomassarina chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557342, Facesoffungi number: FoF 07799; Fig. 43 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0333 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 235–295 µm high × 125–155 µm diam. ( x̄ = 270 × 145 µm, n = 5), immersed, coriaceous, solitary Fungal Diversity Fig. 42 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU, TEF1 and RPB2 sequence data. Ninetyeight strains are included in the combined sequence analysis, which comprise 5181 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 56728.891914 is presented. The matrix had 2700 distinct alignment patterns, with 44.08% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.247521, C = 0.249316, G = 0.271783, T = 0.231379; substitution rates: AC = 1.589623, AG = 3.631327, AT = 1.592103, CG = 1.373577, CT = 7.761405, GT = 1.000000; gamma distribution shape parameter α = 0.271790. 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 dark red bold and type species are in bold. Melanomma pulvis-pyrius (CBS 124080) is used as outgroup taxon or scattered, globose or subglobose to obpyriform, light brown to brown. Ostiole long neck, carbonaceous, papillate, crest-like, elongated and laterally compressed, with hyaline periphyses. Peridium 15–20 µm wide, comprising 2–3-layers of pale brown to hyaline cells of textura angularis. Hamathecium comprising 1–2 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 70–100 × 12–13 µm ( x̄ = 88.5 × 12.5 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short pedicel, apically rounded, with an ocular chamber. Ascospores 20–30 × 4–6.5 µm ( x̄ = 24 × 5 µm, n = 25), overlapping, 1–2-seriate, hyaline, fusiform, 13 Fungal Diversity Fig. 43 Neomassarina chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–h Asci. i–n Ascospores with a gelatinous sheath forming an appendage apicaly. Scale bars: a, b = 500 µm, c = 100 µm, d = 20 µm, g, h = 50 µm, e, f, i–n = 10 µm uniseptate, straight or slightly curved, constricted at the septum, smooth, with a narrow sheath; sheath drawn out to form polar appendages 4.5–9.5 µm long × 2–4 µm wide ( x̄ = 6.5 × 3 µm, n = 30), from both ends of the ascospores, straight or slightly curved. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature, germ tubes produced from the apex. Colonies on MEA circular, mycelium slightly crateriform, flattened, filiform, white aerial hyphae at the surface and creamy brown to yellowish-brown at the center in reverse, white to yellow-white at the margin (Fig. 45a). 13 Pre-screening for antimicrobial activity: Neomassarina chromolaenae (MFLUCC 17-1480) showed antimicrobial activity against E. coli with an 11 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Lampang Province, Ngao, on dead stems of Chromolaena odorata, 21 September 2016, A. Mapook (LP1, MFLU 20-0333, holotype); extype culture MFLUCC 17-1480. GenBank numbers: LSU: MT214466, ITS: MT214372, SSU: MT214419, TEF1: MT235785, RPB2: MT235822 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS and TEF1 sequences for Fungal Diversity Fig. 44 Neomassarina thailandica (new host record) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k– n Ascospores. o–p Ascospores with gelatinous sheath in Indian ink. Scale bars: a = 200 µm, b = 100 µm, c = 50 µm, d, g–j = 20 µm, e, k–p = 10 µm, f = 5 µm Neomassarina chromolaenae (MFLUCC 17-1480, exholotype) is N. thailandica with 94.72% (NR_154244) and 96.30% (KX672163), respectively, similarity to the strain MFLU 11-0144, while the closest match with the LSU and SSU sequences were with N. pandanicola strain MFLUCC 16-0270 with 99.67% (MG298946) and 98.34% (MG298947) similarity. In the present phylogenetic analysis, N. chromolaenae is closely related to N. thailandica with high bootstrap support (100% ML and 1.00 BYPP, Fig. 42). However, N. chromolaenae differs from N. thailandica in having wider asci (70–100 × 12–13 µm vs. (70–)75–90(–93) × 7–8(–8.5) µm) and larger ascospores (20–30 × 4–6.5 µm vs. (17.5–)18–20 × 3–4(–5) µm) with a narrow sheath drawn out to form polar appendages (4.5–9.5 × 2–4 µm vs. 4.5–8 × 1.5–3.5 µm) (Table 9). A comparison of the ITS (+5.8S) gene region of N. chromolaenae and N. thailandica reveals 30 base pair differences (6%) across 495 nucleotides. Therefore, N. chromolaenae is described as a new species based on phylogeny with morphological comparison. Neomassarina thailandica Phook., Jayasiri & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 138 (2016) Facesoffungi number: FoF 02260; Fig. 44 13 Fungal Diversity Fig. 45 Culture characteristics on MEA: a Neomassarina chromolaenae (MFLUCC 17-1480). b Neomassarina thailandica (MFLUCC 17-1432) Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (90–)135–160 µm high × 95–130 µm diam. ( x̄ = 125 × 110 µm, n = 5), immersed, coriaceous, solitary or scattered, obpyriform, flattened base, light brown to brown. Ostiole long neck, carbonaceous, papillate, crest-like, elongated and laterally compressed, with hyaline periphyses. Peridium 5–15 µm wide, comprising 2 layers of pale brown to hyaline cells of textura angularis. Hamathecium comprising (1–)1.5–2.3 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 55–70 × 7–9 µm ( x̄ = 60 × 8 µm, n = 15), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short pedicel, apically rounded, with an ocular chamber. Ascospores 17.5–20 × 2.5–4.5 µm ( x̄ = 18.5 × 3.5 µm, n = 25), overlapping, 1–2-seriate, hyaline, fusiform, uniseptate, straight or slightly curved, guttulate, constricted at the septum, surrounded by hyaline gelatinous sheath; sheath drawn out to form polar appendages 4.5–8 µm long × 1.5–3.5 µm wide ( x̄ = 6 × 2.5 µm, n = 10), observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature, germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly crateriform, flattened, filiform, initial white aerial hyphae, becoming pale brown from the center at the surface and initial white to yellow-white at the margin, brown to dark brown at the center in reverse, becoming light brown to brown in old culture (Fig. 45b). Pre-screening for antimicrobial activity: Neomassarina thailandica (MFLUCC 17-1432) showed antimicrobial activity against M. plumbeus with a 30 mm inhibition zone, observable as partial inhibition, when compared to the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Known hosts and distribution: On dead bract-like leaves from flower stalk of Agave angustifolia (Asparagaceae) in Thailand (Hyde et al. 2016) Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP15, MFLU 20-0334); living culture MFLUCC 17-1432 (new host record). GenBank numbers: LSU: MT214467, ITS: MT214373, SSU: MT214420, TEF1: MT235786, RPB2: MT235823 Notes: In the present phylogenetic analysis, MFLUCC 17-1432 is closely related to N. thailandica with high bootstrap support (100% ML and 1.00 BYPP, Fig. 42). In a BLASTn search of NCBI GenBank, the closest match of ITS and TEF1 sequences showed that the strains is identical to N. thailandica with 100% (NR_154244) and 99.31% (KX672163), respectively, while the closest match with the LSU and SSU sequences were with N. pandanicola strain MFLUCC 16-0270 with 100% (MG298946) and 98.06% (MG298947) similarity. Therefore, we identify our isolates as N. thailandica based on phylogenetic analyses with morphological comparison (Table 9) and the isolates are introduced here as a new host record from Chromolaena odorata collected in Thailand. Neopyrenochaetaceae Valenzuela-Lopez et al. Neopyrenochaetaceae was introduced by ValenzuelaLopez et al. (2018) to accommodate the monotypic genus Neopyrenochaeta based on phylogeny. Neopyrenochaeta Valenzuela-Lopez et al. Neopyrenochaeta was introduced by Valenzuela-Lopez et al. (2018) for a new species N. fragariae. The genus Table 9 Synopsis of Neomassarina species with similar morphological features discussed in this study Species Ascomata (µm) Asci (µm) N. chromolaenae (MFLUCC 17-1480) N. pandanicola (MFLUCC 16-0270) N. thailandica (MFLUCC 17-1432) N. thailandica (MFLUCC 10-0552) 235–295 × 125–155 70–100 × 12–13 20–30 × 4–6.5 4.5–9.5 × 2–4 This study 140–160 × 130–180 50–80 × 6.5–10 11–20 × 3–5 3.5–6 long Hyde et al. (2018a, b) 4.5–8 × 1.5–3.5 This study 13 Ascospores (µm) Polar appendages (µm) References (90–)135–160 × 95–130 55–70 × 7–9 17.5–20 × 2.5–4.5 130–180 × 100–200 (17.5–)18–20 × 3–4(–5) Surrounded by a distinct mucilaginous sheath (70–)75–90(– 93) × 7–8(– 8.5) Hyde et al. (2016b) Fungal Diversity Fig. 46 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS and RPB2 sequence data. Fourty-nine strains are included in the combined sequence analysis, which comprise 2556 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 18819.011208 is presented. The matrix had 977 distinct alignment patterns, with 12.78% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.247993, C = 0.237502, G = 0.272601, T = 0.241905; substitution rates: AC = 1.771498, AG = 5.599945, AT = 1.899441, CG = 1.222461, CT = 10.018389, GT = 1.000000; gamma distribution shape parameter α = 0.168302. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given above or below the nodes. Newly generated sequences are in dark red bold and type species are in bold. Pleospora herbarum (CBS 191.86) and P. herbarum (IT956) are used as outgroup taxa belonging in Cucurbitariaceae (Aveskamp et al. 2010; de Gruyter et al. 2010, 2013; Doilom et al. 2013) based on phylogenetic analysis and available sequences from GenBank. The genus is mostly found as the asexual morph, and has been reported from diverse habitats. N. acicula was recorded from a waterpipe, N. fragariae from fruit (Fragaria ananassa), N. inflorescentiae from plants (Protea neriifolia), N. telephoni from a mobile phone (Crous et al. 2015d; Valenzuela-Lopez et al. 2018), and N. cercidis on pods of Cercis chinensis in China based on morphology and phylogeny (Jayasiri et al. 2019). Five epithets are listed in Index Fungorum (2020) with N. acicula as the type species. We introduce the sexual morphs of four new Neopyrenochaeta species, which were collected from dead aerial stems of Chromolaena odorata, based on morphology and molecular data, together with descriptions and illustrations (Figs. 47, 48, 49, 50). A phylogenetic tree based on combined LSU, ITS and RPB2 sequence data is presented in Fig. 46. Neopyrenochaeta chiangraiensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557311, Facesoffungi number FoF 07482; Fig. 47 13 Fungal Diversity Fig. 47 Neopyrenochaeta chiangraiensis (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Dark brown setae. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores. Scale bars: a = 500 µm, b = 200 µm, c, g–j = 20 µm, d, e, k–p = 10 µm, f = 5 µm Etymology: Referring to the location where the specimen was collected, Chiang Rai Province, Thailand. Holotype: MFLU 20-0337 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 100–110 µm high × 95–105 µm diam. ( x̄ = 105 × 100 µm, n = 5), superficial, appearing as small dark spots, coriaceous, solitary or scattered, globose to obpyriform, brown to dark brown. Ostiole short papillate, with numerous external brown setae. Peridium (3.5–)5–12 µm wide, comprising 1–2 layers of thin-walled, pale brown to 13 brown cells of textura angularis. Hamathecium comprising 1–2 µm wide, cylindrical, septate, branching pseudoparaphyses. Asci (45–)50–65 × 8–12 µm ( x̄ = 55 × 9.5 µm, n = 20), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short bulbous pedicel, apically rounded. Ascospores 14–18 × 3.5–5 µm ( x̄ = 16 × 4.5 µm, n = 15), overlapping, bi-seriate, light olivaceous-brown to pale yellowish brown, cylindric-fusiform to inequilateral, with slightly obtuse ends, widest at the second cell from the apex and tapering towards ends, 3-septate, straight or Fungal Diversity Fig. 48 Neopyrenochaeta chromolaenae (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Dark brown setae. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores. Scale bars: a = 500 µm, b = 100 µm, c, g–j = 20 µm, d, e, k–p = 10 µm, f = 5 µm slightly curved, small guttule, slightly constricted at the septa, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium flat, filamentous, white to light grey at first, become greyishbrown at the surface and greyish-brown to dark brown in reverse with age (Fig. 51a). Pre-screening for antimicrobial activity: Not tested. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP38, MFLU 20-0337, holotype); extype culture MFLUCC 17-1445. GenBank numbers: LSU: MT214468, ITS: MT214374 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Neopyrenochaeta chiangraiensis (MFLUCC 17-1445, ex-holotype) with 96.07% similarity was Setophaeosphaeria badalingensis (strain S139, MK304112), while the closest match with the LSU sequence with 96.04% similarity was Neopyrenochaeta cercidis (strain C136, MK347932). In the present phylogenetic analysis, N. chiangraiensis clusters with N. chromolaenae (MFLUCC 17-1446) with high bootstrap support (99% ML and 1.00 BYPP, Fig. 46). 13 Fungal Diversity Fig. 49 Neopyrenochaeta thailandica (holotype) a, b Appearance of superficial ascomata on substrate. c Ascoma d Section through ascoma. e Brown setae. f Peridium. g Pseudoparaphyses. h–k Asci. l–q Ascospores. Scale bars: a, b = 200 µm, c, d, h–k = 50 µm, e–f = 20 µm, g, l–q = 10 µm However, N. chiangraiensis differs from N. chromolaenae in having smaller asci [(45–)50–65 × 8–12 µm vs. (55.5–)60–80 × (14.5–)15–20 µm] and smaller ascospores (14–18 × 3.5–5 µm vs. 23–26 × 6–8 µm) that are light olivaceous-brown to pale yellowish brown with 3 septa, while N. chromolaenae has pale olivaceous-brown to smoke-grey, 5-septate ascospores (Table 10). A comparison of the ITS (+5.8S) gene region of N. chiangraiensis and N. chromolaenae reveals 23 base pair differences (4.4%) across 517 nucleotides. Therefore, N. chiangraiensis is described as a new species based on phylogeny and morphological comparison. Neopyrenochaeta chromolaenae Mapook & K.D. Hyde, sp. nov. 13 Index Fungorum number: IF557312, Facesoffungi number: FoF 07800; Fig. 48 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0338 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 95–110 µm high × 70–75(–145) µm diam. ( x̄ = 100 × 108 µm, n = 5), superficial, appearing as small dark spots, coriaceous, solitary or scattered, globose to obpyriform, brown to dark brown. Ostiole papillate, with numerous external brown setae. Peridium 5–15 µm wide, comprising 1–2 layers of thin-walled, pale brown to brown cells of textura angularis. Hamathecium comprising 1–2.5 µm wide, cylindrical, septate, branching, Fungal Diversity Fig. 50 Neopyrenochaeta triseptatispora (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Ostiole with brown setae. e Brown seta. f Peridium. g Pseudoparaphyses. h–k Immature and mature asci. l–q Ascospores. Scale bars: a = 200 µm, b = 100 µm, c, h–k = 20 µm, d, f = 10 µm, e, g, l–q = 5 µm pseudoparaphyses. Asci (55.5–)60–80 × (14.5–)15–20 µm ( x̄ = 70 × 16.5 µm, n = 20), 8-spored, bitunicate, fissitunicate, cylindric-clavate to broadly clavate, straight or slightly curved, with a short bulbous pedicel, apically rounded. Ascospores 23–26 × 6–8 µm ( x̄ = 24.5 × 7 µm, n = 25), overlapping 3–5-seriate, pale olivaceous-brown to smoke-grey, cylindric-fusiform, with slightly obtuse ends, widest at the center and tapering towards the rounded ends, 5-septate, straight or slightly curved, guttulate, constricted at the septa, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium flat, filamentous, white at first with brown from the center of the colony, become greyish to dark olivaceous at the surface and dark olivaceous in reverse (Fig. 51b). Pre-screening for antimicrobial activity: Neopyrenochaeta chromolaenae (MFLUCC 17-1446) showed antimicrobial activity against B. subtilis and E. coli (13 mm and 10 mm inhibition zone, respectively), when compared to the positive control (26 mm and 17 mm, respectively), but no inhibition of M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 13 Fungal Diversity Fig. 51 Culture characteristics on MEA: a Neopyrenochaeta chiangraiensis (MFLUCC 17-1445). b Neopyrenochaeta chromolaenae (MFLUCC 17-1446). c Neopyrenochaeta thailandica (MFLUCC 17-1461). d Neopyrenochaeta triseptatispora (MFLUCC 17-1436) 2015, A. Mapook (DP39, MFLU 20-0338, holotype); extype culture MFLUCC 17-1446. GenBank numbers: LSU: MT214469, ITS: MT214375, SSU: MT214421, TEF1: MT235787, RPB2: MT235824 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Neopyrenochaeta chromolaenae (MFLUCC 17-1446, ex-holotype) with 95.06% similarity was Setophaeosphaeria badalingensis (strain 41, MK311292). The closest match with the LSU sequence with 99.42% similarity was Neopyrenochaeta cercidis (strain C136, MK347932), while the closest match with the RPB2 sequence with 94.25% similarity was Setophaeosphaeria citricola (strain CBS 143179, MH108008). In the present phylogenetic analysis, N. chromolaenae clusters with N. chiangraiensis (MFLUCC 17-1445) with high bootstrap support (99% ML and 1.00 BYPP, Fig. 46). However, N. chromolaenae differs from N. chiangraiensis in having larger asci [(55.5–)60–80 × (14.5–)15–20 µm vs. (45–)50–65 × 8–12 µm] and ascospores (23–26 × 6–8 µm vs. 14–18 × 3.5–5 µm) that are pale olivaceous-brown to smokegrey with 5 septa, while N. chiangraiensis has light olivaceous-brown to pale yellowish brown, 3-septate ascospores (Table 10). A comparison of the ITS (+5.8S) gene region of N. chromolaenae and N. chiangraiensis reveals 23 base pair differences (4.4%) across 517 nucleotides. Therefore, N. chromolaenae is described as a new species based on phylogeny and morphological comparison. Neopyrenochaeta thailandica Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557313, Facesoffungi number: FoF 07801; Fig. 49 Etymology: The name reflects the country, where the specimen was collected, Thailand. Holotype: MFLU 20-0339 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 115–160 µm high × 110–150 µm diam. ( x̄ = 133 × 125 µm, n = 5), superficial, appearing as small dark spots, coriaceous, solitary or scattered, subglobose to obpyriform, reddish brown to brown. Ostiole short papillate, with numerous external reddish-brown setae. Peridium 5–12 µm wide, comprising 2–3-layers of thin-walled, pale brown to brown cells of textura angularis. Hamathecium comprising (1.5–)2–2.5 µm wide, cylindrical to broadly filiform, septate, branching pseudoparaphyses. Asci (55–)60–80 × 13–20 µm ( x̄ = 67.5 × 16.5 µm, Table 10 Morphological features of Neopyrenochaeta species discussed in this study Species Asocomata (µm) Peridium (µm) Asci (µm) Ascospores (µm) References N. chiangraiensis (MFLUCC 17-1445) 100–110 high × 95–105 diam. (3.5–)5–12 (45–)50–65 × 8–12 This study N. chromolaenae (MFLUCC 17-1446) 95–110 high × 70–75 (–145) diam. 5–15 (55.5–)60–80 × (14.5–)15–20 N. thailandica (MFLUCC 17-1461) 115–160 high × 110–150 diam. 5–12 (55–)60–80 × 13–20 N. triseptatispora (MFLUCC 17-1436) 75–110 high × 70–90 diam. 5–10(–15) 14–18 × 3.5–5; light olivaceous-brown to pale yellowish brown, 3-septate 23–26 × 6–8; pale olivaceous-brown to smoke-grey, 5-septate (21.5–)23–32 × 5–9; pale olivaceous-grey to slightly yellowish brown, 3–5-septate 15–20 × 4–7; hyaline, 3-septate 13 45–65 × 9.5–13 This study This study This study Fungal Diversity n = 15), 8-spored, bitunicate, fissitunicate, cylindricclavate to broadly clavate, straight or slightly curved, apically rounded, with a short, bulbous pedicel. Ascospores (21.5–)23–32 × 5–9 µm ( x̄ = 26 × 6.5 µm, n = 40), overlapping 1–3-seriate, pale olivaceous-grey to slightly yellowish brown, cylindrical to broadly fusiform, with slightly obtuse ends, widest at the center and tapering towards the rounded ends, 3–5-septate, straight or slightly curved, third cell from top slightly swollen, guttulate, slightly constricted at the septa, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium slightly raised, entire, greyish to dark olivaceous at the surface and dark olivaceous in reverse (Fig. 51c). Pre-screening for antimicrobial activity: Neopyrenochaeta thailandica (MFLUCC 17-1461) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Lampang Province, Chaehom, on dead stems of Chromolaena odorata, 24 September 2016, A. Mapook (JH7, MFLU 20-0339, holotype); ex-type culture MFLUCC 17-1461. GenBank numbers: LSU: MT214470, ITS: MT214376, SSU: MT214422, TEF1: MT235788, RPB2: MT235825 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Neopyrenochaeta thailandica (MFLUCC 17-1461, ex-holotype) with 98.96% similarity was Setophaeosphaeria hemerocallidis (strain A582, KX463035). The closest match with the LSU sequence with 99.12% similarity was N. cercidis (strain C136, MK347932), while the closest match with the RPB2 sequence with 88.88% similarity was Setophaeosphaeria citricola (strain CBS 143179, MH108008). In the present phylogenetic analysis, N. thailandica forms a sister taxon with N. chromolaenae (MFLUCC 17-1446), N. chiangraiensis (MFLUCC 17-1445) and N. triseptatispora (MFLUCC 17-1436) with bootstrap support (1.00 BYPP, Fig. 46). However, N. thailandica differs from N. chiangraiensis and N. triseptatispora in having larger asci [(55–)60–80 × 13–20 µm vs. (45–)50–65 × 8–12 µm and 45–65 × 9.5–13 µm] and larger ascospores [(21.5–)23–32 × 5–9 µm vs. 14–18 × 3.5–5 µm and 15–20 × 4–7 µm] that are 3–5-septate, while N. chiangraiensis and N. triseptatispora have 3-septate ascospores (Table 10). A comparison of the ITS (+5.8S) gene region of N. thailandica and N. triseptatispora reveals 66 base pair differences (12.15%) across 543 nucleotides. Therefore, N. thailandica is described as a new species based on phylogeny and morphological comparison. Neopyrenochaeta triseptatispora Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557314, Facesoffungi number: FoF 07802; Fig. 50 Etymology: The epithet “triseptatispora” refers to the 3-septate ascospores. Holotype: MFLU 20-0340 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 75–110 µm high × 70–90 µm diam. ( x̄ = 95 × 80 µm, n = 5), superficial, appearing as, small dark spots, coriaceous, solitary or scattered, globose to obpyriform, reddish brown to brown. Ostiole short papillate, with numerous external light brown to brown setae. Peridium 5–10(–15) µm wide, comprising 2–3 layers of thin-walled, hyaline or pale brown to brown cells of textura angularis. Hamathecium comprising 1–3 µm wide, cylindrical, septate, branching pseudoparaphyses. Asci 45–65 × 9.5–13 µm ( x̄ = 56 × 11 µm, n = 20), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short bulbous pedicel, apically rounded. Ascospores 15–20 × 4–7 µm ( x̄ = 18 × 6 µm, n = 40), overlapping, 2–3-seriate, hyaline, cylindric-fusiform to inequilateral, with obtuse ends, slightly widest at the second cell from the apex and tapering towards the rounded ends, 3-septate, straight or slightly curved, small guttule, constricted at the septa, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from several cells. Colonies on MEA circular, mycelium flat, filamentous, white at first, become greyish to dark olivaceous at the surface and dark olivaceous in reverse with age (Fig. 51d). Pre-screening for antimicrobial activity: Neopyrenochaeta triseptatispora (MFLUCC 17-1436) showed antimicrobial activity against B. subtilis with a 16 mm inhibition zone and against M. plumbeus with a 24 mm inhibition zone, observable as partial inhibition, when compared to the positive control (25 mm and 17 mm, respectively), but no inhibition of E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP20, MFLU 20-0340, holotype); extype culture MFLUCC 17-1436. GenBank numbers: LSU: MT214471, ITS: MT214377, SSU: MT214423, TEF1: MT235789, RPB2: MT235826 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Neopyrenochaeta triseptatispora (MFLUCC 17-1436, ex-holotype) with 99.62% similarity was Setophaeosphaeria citricola (strain CBS 143179, MH107916). The closest match with the LSU sequence with 98.97% similarity was N. cercidis (strain C136, MK347932), while the closest match with the RPB2 sequence with 93.75% similarity was Setophaeosphaeria citricola (strain CBS 143179, MH108008). In the present phylogenetic analysis, N. triseptatispora 13 Fungal Diversity is forms a sister taxon with N. chromolaenae (MFLUCC 17-1446) and N. chiangraiensis (MFLUCC 17-1445) with high bootstrap support (98% ML and 1.00 BYPP, Fig. 46). However, N. triseptatispora differs from N. chromolaenae in having smaller ascomata (75–110 × 70–90 µm vs. 95–110 × 70–75(–145) µm), smaller asci (45–65 × 9.5–13 µm vs. (55.5–)60–80 × (14.5–)15–20 µm) and smaller ascospores (15–20 × 4–7 µm vs. 23–26 × 6–8 µm) that are 3-septate, while N. chromolaenae has 5-septate ascospores (Table 10). N. triseptatispora is also similar to N. chiangraiensis in having 3-septate ascospores but differs in slightly larger ascomata (75–110 × 70–90 µm vs. 100–110 × 95–105 µm) and ascospores (15–20 × 4–7 µm vs. 14–18 × 3.5–5 µm). A comparison of the ITS (+5.8S) gene region of N. triseptatispora and N. chiangraiensis reveals 47 base pair differences (9.1%) across 515 nucleotides. Therefore, N. triseptatispora is described as a new species based on phylogeny and morphological comparison. Phaeosphaeriaceae M.E. Barr Fig. 52 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU, TEF1 and RPB2 sequence data. Thirteen strains are included in the combined sequence analysis, which comprise 4416 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 8333.195072 is presented. The matrix had 327 distinct alignment patterns, with 24.82% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.252190, C = 0.233838, G = 0.264447, T = 0.249525; substitution rates: AC = 1.127465, AG = 2.895211, AT = 2.022869, CG = 0.534488, CT = 8.448227, GT = 1.000000; gamma distribution shape parameter α = 0.020000. 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 dark red bold and type species are in bold. Phaeosphaeria oryzae (CBS 110110) is used as outgroup taxon 13 Phaeosphaeriaceae was introduced by Barr (1979) with Phaeosphaeria as type genus and monographed most recently by Phookamsak et al. (2014). Fifty-two genera were accepted in Wijayawardene et al. (2018). Subsequently, Yang et al. (2019) introduced a new genus, Neostagonosporella with Neostagonosporella sichuanensis as a new species. Zhang et al. (2019) introduced two new genera Hydeopsis and Pseudoophiosphaerella from Karst landforms in Guizhou Province, China, based on morphology and phylogeny. Recently, Phookamsak et al. (2019) introduced a new genus, Brunneomurispora from China. Divergence time estimates for this family are crown age of 75 Mya (46–102) and stem age of 99 Mya (73–129) in the Cretaceous (Liu et al. 2017). Leptospora Rabenh. Leptospora was introduced by Rabenhorst (1857) with L. rubella as the type species. Hyde et al. (2016b) reported a reference specimen of L. rubella from UK with two new species L. galii from Italy and L. thailandica from Thailand. Fungal Diversity Zhang et al. (2019) introduced a new species, L. hydei from China. We introduce two new Leptospora species with a new host record of L. thailandica from C. odorata, together with descriptions and illustrations (Figs. 53, 54, 55, 56). A phylogenetic tree based on combined LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 52. Leptospora chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557289, Facesoffungi number: FoF 07803; Fig. 53 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0313 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 200–260 µm high × 200–250 µm diam. ( x̄ = 230 × 220 µm, n = 10), semi-immersed to superficial, solitary or scattered, appearing as dark spots, coriaceous, globose to obpyriform, reddish brown or dark brown to black. Ostiolar neck protruding, with hyaline periphyseslike. Peridium (5–)10–20 µm wide, 2–3 layers, inner layers comprising of hyaline to pale brown cells of textura angularis, outer layers comprising of reddish brown to yellowish brown cells of textura epidermoidea. Hamathecium comprising 2–3.5 µm wide, cylindrical, septate, branching pseudoparaphyses, anastomosing above the asci. Asci 80–115 × 9.5–15 µm ( x̄ = 93 × 11 µm, n = 20), overlapping, 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, with a short, bulbous pedicel, ocular chamber visible when immature. Ascospores 60–95 × 2.5–3.5 µm ( x̄ = 77 × 3.2 µm, n = 25), fasciculate, in parallel, scolecosporous, hyaline to pale brown, filiform, (22–)27–28-septate, minute guttule in each cell, straight or slightly curved, with globose polar appendages, observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, moderately fluffy, entire to filamentous, cultures white to greyish-brown at the surface and initially creamy-white in reverse, becoming dark brown from the centre of the colony with creamy-white at the margin (Fig. 56a). Pre-screening for antimicrobial activity: Leptospora chromolaenae (MFLUCC 17-1421) showed antimicrobial activity against B. subtilis with a 13 mm inhibition zone and against M. plumbeus with a 20 mm inhibition zone, observable as partial inhibition, when compared to the positive control (26 mm and 17 mm, respectively), but no inhibition of E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 8 July 2015, A. Mapook (DP2, MFLU 20-0313, holotype); extype culture MFLUCC 17-1421; 5 August 2015, A. Mapook (DP41, MFLU 20-0314); living culture MFLUCC 17-1448; (DP44, MFLU 20-0315); living culture MFLUCC 17-1451; Lampang Province, Chaehom, on dead stems of C. odorata, 24 September 2016, A. Mapook (JH6, MFLU 20-0317); living culture MFLUCC 17-1460; (JH19, MFLU 20-0316); living culture MFLUCC 17-1501. GenBank numbers: LSU: MN994552, MN994553, MN994554, MN994555, MN994556, ITS: MN994575, MN994576, MN994577, MN994578, MN994579, SSU: MN994598, MN994599, MN994600, MN994601, MN994602, TEF1: MN998156, MN998157, MN998158, MN998159, MN998160 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Leptospora chromolaenae (MFLUCC 17-1421, ex-holotype) with 95.93% similarity was L. rubella (strain CAP17F, JQ936327). The closest match with the LSU sequence with 99.88% similarity was L. rubella (strain CPC 11006, DQ195792). The closest match with the SSU sequence with 99.91% similarity was Phaeosphaeria sp. (strain sn23-1, EU189215). The closest match with the TEF1 sequence with 96.04% similarity was L. hydei (strain GZCC 19-0004, MK523387), while the closest match with the RPB2 sequence with 77.11% similarity was Parastagonospora fusiformis (strain MFLUCC 13-0215, KX863711). In the present phylogenetic analysis, L. chromolaenae forms a monophyletic clade and clusters with the clade comprising, L. hydei, L. phraeana and L. thailandica with hight bootstrap support (89% ML and 1.00 BYPP, Fig. 52). However, L. chromolaenae differs from these species in having (22–)27–28-septate ascospores (Table 11). Therefore, L. chromolaenae is described as a new species based on phylogeny and morphological comparison. Leptospora phraeana Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557291, Facesoffungi number: FoF 07804; Fig. 54 Etymology: Referring to the location where the specimen was collected, Phrae Province, Thailand. Holotype: MFLU 20-0318 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (130–)180–235 µm high × 125–185 µm diam. ( x̄ = 200.5 × 155 µm, n = 10), semi-immersed, solitary or scattered, appearing as dark spots, coriaceous, globose to obpyriform, reddish brown to yellowish brown or brown. Ostiolar neck protruding, with hyaline periphyses-like cells. Peridium (10–)15–20 µm wide, 2–3 layers, comprising light brown to brown cells of textura angularis. Hamathecium comprising 2–3.5 µm wide, cylindrical, septate, branching pseudoparaphyses, anastomosing above the asci. Asci (60–)70–90 × 8–13 µm ( x̄ = 80 × 11 µm, n = 10), overlapping, 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, with a short pedicel, ocular chamber observed clearly when immature. 13 Fungal Diversity Fig. 53 Leptospora chromolaenae (holotype) a, b Appearance of superficial ascomata on substrate. c Section through an ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g, h Asci. i–l Ascospores with globose polar appendages. m Ascospores with polar appendages in Indian ink. Scale bars: a, b = 500 µm, c, g–m = 50 µm, d = 20 µm, e, f = 10 µm Ascospores 50–70 × 2–3 µm ( x̄ = 63 × 2.6 µm, n = 15), fasciculate, in parallel, scolecosporous, pale brown to yellowish brown, cylindrical to broadly filiform, 13-septate with minute guttule in each cell, straight or slightly curved, with polar appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, moderately fluffy, entire, cultures white at the surface and initially creamy-white to pale brown in reverse, becoming dark brown from the centre of the colony with creamy-white at the margin (Fig. 56b). 13 Pre-screening for antimicrobial activity: Leptospora phraeana (MFLUCC 17-1478) showed antimicrobial activity against M. plumbeus with a 22 mm inhibition zone, observable as partial inhibition, when compared to the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Phrae Province, Doi Pha Klong, on dead stems of Chromolaena odorata, 22 September 2016, A. Mapook (DPKP2, MFLU 20-0318, holotype); ex-type culture MFLUCC 17-1478. GenBank numbers: LSU: MN994557, ITS: MN994580, SSU: MN994603, TEF1: MN998161 Fungal Diversity Fig. 54 Leptospora phraeana (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g, h Asci. i–m Ascospores with polar appendages. Scale bars: a = 500 µm, b = 200 µm, c, g, h = 50 µm, d, e, i–m = 20 µm, f = 10 µm Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Leptospora phraeana (MFLUCC 17-1478, ex-holotype) with 96.26% similarity was L. rubella (strain CAP17F, JQ936327). The closest match with the LSU sequence with 99.65% similarity was L. rubella (strain CPC 11006, DQ195792). The closest match with the SSU sequence with 100% similarity was Phaeosphaeria sp. (strain sn23-1, EU189215). The closest match with the TEF1 sequence with 95.70% similarity was L. hydei (strain GZCC 19-0004, MK523387), while the closest match with the RPB2 sequence with 78.72% similarity was Neophaeosphaeria filamentosa (strain CBS 102202, GU371773). In the present phylogenetic analysis, L. phraeana forms a separate branch and clusters with L. thailandica with hight bootstrap support (85% ML and 1.00 BYPP, Fig. 52). However, L. phraeana differs from L. thailandica in having slightly smaller asci [(60–)70–90 × 8–13 µm vs. 68–114 × 7–13 µm] and slightly smaller ascospores (50–70 × 2–3 µm vs. 63–89 × 1.8–3.8 µm) that are 13-septate, while L. thailandica has (14–)20–22-septate ascospores (Table 11). A comparison of the ITS (+5.8S) gene region of L. phraeana and L. thailandica reveals 17 base pair differences (3.3%) across 506 nucleotides. Therefore, L. phraeana is described as a new species based on phylogeny and morphological comparison. 13 Fungal Diversity Fig. 55 Leptospora thailandica (new host record) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g, h Asci. i–m Ascospores with polar appendages. Scale bars: a = 200 µm, b = 100 µm, c, g, h = 50 µm, d, e, i–m = 20 µm, f = 5 µm Fig. 56 Culture characteristics on MEA: a Leptospora chromolaenae (MFLUCC 17-1421). b Leptospora phraeana (MFLUCC 17-1478). c Leptospora thailandica (MFLUCC 17-1468) Leptospora thailandica Phukhams. & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 100 (2016) Facesoffungi number: FoF 02381; Fig. 55 13 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 150–170 µm high × 120–160 µm diam. ( x̄ = 162 × 136 µm, n = 5), immersed to semi-immersed, This study 63–89 × 1.8–3.8 (60–)75–90(–100) × 2.5– 3.5 68–114 × 7–13 (65–)70–115 × 9.5–15 13–15-septate, with polar appendages Hyde et al. (2016b) This study 13-septate, with polar appendages (14–)20–22-septate 50–70 × 2–3 (130–)180–235 high × 125–185 diam. 188–207 high × 112–170 diam. 150–170 high × 120–160 diam. L. phraeana (MFLUCC 17-1478) L. thailandica (MFLUCC 16-0385) L. thailandica (MFLUCC 17-1468) (60–)70–90 × 8–13 Zhang et al. (2019) (104–)124–138(– 146) × 4–4.8 196–267 high × 168–187 diam. L. hydei (GZCC 19-0004) 147–165(–180) × 13–17(– 19.5) 11-septate This study (22–)27–28-septate, with polar appendages 60–95 × 2.5–3.5 80–115 × 9.5–15 (5–)10–20, 2–3 layers of textura angularis to textura epidermoidea 11–24, several layers of textura angularis to prismatica (10–)15–20, 2–3 layers of textura angularis (5–)10–24(–27), 5–7 layers of textura angularis 10–25, 2–4 layers of textura angularis 200–260 high × 200–250 diam. L. chromolaenae (MFLUCC 17-1448) Ascospores (µm) Asci (µm) Peridium (µm) Asocomata (µm) Species Table 11 Synopsis of Leptospora species with similar morphological features discussed in this study Septation of ascospores References Fungal Diversity solitary or scattered, appearing as dark spots, coriaceous, globose to obpyriform, reddish brown to brown. Ostiolar neck protruding. Peridium 10–25 µm wide, 2–4 layers, inner layers comprising hyaline cells of textura angularis, outer layers comprising light brown to brown cells of textura angularis. Hamathecium comprising 2–2.5 µm wide, oblong to cylindrical, septate, branching pseudoparaphyses, anastomosing above the asci. Asci (65–)70–115 × 9.5–15 µm ( x̄ = 89 × 13 µm, n = 15), overlapping, 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, with a short, bulbous pedicel, ocular chamber visible when immature. Ascospores (60–)75–90(–100) × 2.5–3.5 µm ( x̄ = 83 × 2.9 µm, n = 15), fasciculate, in parallel, scolecosporous, hyaline to pale brown, cylindrical to broadly filiform, 1-septate at the center of the ascospores when immature, 13–15-septate when mature, minute guttule in each cell, straight or slightly curved, with polar appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, entire, cultures white at the surface and initially creamy-white to pale brown in reverse, becoming creamy-brown to brown from the centre of the colony with creamy-white at the margin (Fig. 56c). Pre-screening for antimicrobial activity: Leptospora thailandica (MFLUCC 17-1468) showed antimicrobial activity against B. subtilis, E. coli and M. plumbeus (13 mm, 8 mm and 24 mm inhibition zone, respectively), observable as partial inhibition, when compared to the positive control (26 mm, 9 mm, and 17 mm, respectively). Known hosts and distribution: On dead branches of Duranta sp. (Verbenaceae) in Thailand (Hyde et al. 2016a, b) Material examined: THAILAND, Chiang Rai Province, Doi Mae Salong, on dead stems of Chromolaena odorata, 8 April 2017, A. Mapook (DMS5, MFLU 20-0319); living culture MFLUCC 17-1468 (new host record). GenBank numbers: LSU: MN994558, ITS: MN994581, SSU: MN994604, TEF1: MN998162 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of MFLUCC 17-1468 with 99.65% similarity was L. thailandica (strain MFLUCC 16-0385, NR_154133). The closest match with the LSU sequence with 99.40% similarity was L. hydei (strain GZCC 19-0004, MK522497). The closest match with the SSU sequence with 99.90% similarity was Phaeosphaeria sp. (strain sn23-1, EU189215). The closest match with the TEF1 sequence with 96.02% similarity was Populocrescentia ammophilae (strain MFLUCC 17-0665, MG829231), while the closest match with the RPB2 sequence with 80.18% similarity was Paraphoma fimeti (strain UTHSC: 13 Fungal Diversity Fig. 57 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU and TEF1 sequence data. One hundred sixty-seven strains are included in the combined sequence analysis, which comprise 3312 characters with gaps. The best scoring RAxML tree with a final likelihood value of − 33364.909121 is presented. The matrix had 1254 distinct alignment patterns, with 21.74% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.241300, C = 0.264151, G = 0.235962, T = 0.258586; substitution rates: AC = 0.989272, AG = 3.620674, AT = 1.917765, CG = 0.739088, CT = 4.661336, GT = 1.000000; gamma distribution shape parameter α = 0.156765. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given at the nodes. Newly generated sequences are in dark red bold and type species are in bold. Staurosphaeria rhamnicola (MFLUCC 17-0814) and S. rhamnicola (MFLUCC 17-0813) are used as outgroup taxa 13 Fungal Diversity Fig. 57 (continued) DI16-296, LT797032). In the present phylogenetic analysis, MFLUCC 17-1468 clusters with L. thailandica strain MFLUCC 16-0385 with hight bootstrap support (100% ML and 1.00 BYPP, Fig. 52). Therefore, we identify MFLUCC 17-1468 as Leptospora thailandica based on phylogenetic analyses with morphological comparison (Table 11) and the isolates are introduced here as a new host record from Chromolaena odorata collected in Thailand. Murichromolaenicola Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557338, Facesoffungi number: FoF 07805 Etymology: The generic epithet reflects the muriform ascospore, which was growing on the host genus Chromolaena. Saprobic on dead stems. Sexual morph: Ascomata semiimmersed to superficial, solitary or scattered, appearing as dark spots, coriaceous, globose to obpyriform, light brown to brown. Ostiolar neck protruding, with hyaline periphyses-like cells. Peridium comprising several layers of brown to dark 13 Fungal Diversity brown cells of textura angularis. Hamathecium composed of cylindrical, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, pedicellate. Ascospores overlapping, 1–2-seriate, initially hyaline to golden-yellow, 1-septate when immature, becoming golden-brown to brown at maturity, ellipsoid to broadly fusiform, muriform, 5–7-transversely septate, with 1–2-vertical septum, guttulate, slightly constricted at the central septum, straight or slightly curved, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Conidiomata pycnidial, solitary, immersed to semi-immersed in the host surface, uni-loculate, globose, yellowish brown to brown. Pycnidial wall comprising 4–5 layers of thick-walled, dense, yellowish brown to brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells phialidic, ampulliform to cylindrical, hyaline, unbranched. Conidia ellipsoid to broadly fusiform, muriform, 5–7-transversely septate, with 1–2-vertical septa, small guttule, not constricted at the septa, yellowish brown to brown, with polar appendages, observed clearly when mounted in Indian ink. Type species: Murichromolaenicola chromolaenae Mapook & K.D. Hyde Notes: A phylogenetic analyses based on combined dataset of LSU, ITS, SSU and TEF1 sequence data show that two new Murichromolaenicola species form a distinct lineage with Wojnowicia and Wojnowiciella with low bootstrap support (Fig. 57). Murichromolaenicola species are distinct morphologially in having ellipsoid to broadly fusiform, muriform ascospores, 5–7-transversely septate with 1–2-vertical septa, while Wojnowiciella has subcylindrical to ellipsoid conidia and Wojnowicia has fusiform to cylindrical conidia. Therefore, we introduce Murichromolaenicola as a new genus with two new species M. chiangraiensis and M. chromolaenae, based on morphology and phylogeny, together with a comparison of the TEF1 gene region. Murichromolaenicola chiangraiensis Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557339, Facesoffungi number: FoF 07806; Fig. 58 Etymology: Referring to the location where the specimen was collected, Chiang Rai Province, Thailand. Holotype: MFLU 20-0326 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 460–555 µm high × (350–)500–530 µm diam. ( x̄ = 515 × 475 µm, n = 5), semi-immersed to superficial, solitary or scattered, appearing as dark spots, coriaceous, globose to obpyriform, light brown to brown. Ostiolar neck protruding, with hyaline periphyses-like. Peridium 30–50 µm wide, comprising several layers of brown to dark brown cells of textura angularis. Hamathecium comprising 3–4 µm wide, cylindrical, septate, branching pseudoparaphyses. Asci 13 (80–)110–135 × (15–)17–20 µm ( x̄ = 117 × 18 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, pedicellate. Ascospores 25–35 × 10–13 µm ( x̄ = 29 × 11 µm, n = 20), overlapping, 1–2-seriate, initially hyaline to golden-yellow, 1-septate when immature, becoming golden-brown to brown at maturity, ellipsoid to broadly fusiform, muriform, 5–7-transversely septate, with 1–2-vertical septum, guttulate, slightly constricted at the central septum, straight or slightly curved, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from several cells. Colonies on MEA circular, mycelium crateriform, entire, cultures white with greyish from the centre of the colony at the surface, pale olivaceous-brown from the centre of the colony in reverse with creamy-white at the margin (Fig. 60a). Pre-screening for antimicrobial activity: Not tested. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 April 2017, A. Mapook (DP86, MFLU 20-0326, holotype); extype culture MFLUCC 17-1488. GenBank numbers: LSU: MN994559, ITS: MN994582, SSU: MN994605, TEF1: MN998163 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Murichromolaenicola chiangraiensis (MFLUCC 17-1488, ex-holotype) with 94.02% similarity was Wojnowicia dactylidicola (strain MFLUCC 13-0738, NR_154501). The closest match with the LSU and TEF1 sequences with 91.68% (KX430001) and 83.80% (KX430003) similarity, respectively, were W. italica strain MFLU 14-0732, while the closest match with the SSU sequence with 93.15% similarity was Pyrenochaeta corni (strain CBS 102828, GQ387548). In the present phylogenetic analysis, M. chiangraiensis clusters with M. chromolaenae (MFLUCC 17-1489) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 52). Although we could not compare the morphological characteristics of those species; M. chiangraiensis is found as sexual morph in nature and we could not obtain its asexual morph in culture, while M. chromolaenae is found as asexual morph in nature, they differ in culture characteristics on MEA (Fig. 60). A comparison of the TEF1 gene region of M. chiangraiensis and M. chromolaenae reveals 41 base pair differences (4.7%) across 872 nucleotides. Therefore, M. chiangraiensis is described here as a new species based on phylogeny. Murichromolaenicola chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557340, Facesoffungi number: FoF 07807; Fig. 59 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Fungal Diversity Fig. 58 Murichromolaenicola chiangraiensis (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Peridium. e Pseudoparaphyses. f–i Asci. j–n Ascospores. o Ascospores with gelatinous sheath in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c = 100 µm, f–i = 50 µm, d, j–o = 20 µm, e = 5 µm Holotype: MFLU 20-0327 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata 200–235 µm high × 195–230(–260) µm diam. ( x̄ = 228 × 226 µm, n = 5), pycnidial, solitary, immersed to semi-immersed in the host surface, uni-loculate, globose, yellowish brown to brown. Pycnidial wall 10–20 µm wide, comprising 4–5 layers of thick-walled, dense, yellowish brown to brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells phialidic, ampulliform to cylindrical, hyaline, unbranched. Conidia 14–25 × 6.5–11 μm ( x̄ = 20.5 × 8.5 µm, n = 50), ellipsoid to broadly fusiform, muriform, 5–7-transversely septate, with 1–2-vertical septa, small guttule, not constricted at the septa, yellowish brown to brown, with polar appendages, observed clearly when mounted in Indian ink. Culture characteristics: Conidia germinating on MEA within 24 h at room temperature and germ tubes produced 13 Fungal Diversity Fig. 59 Murichromolaenicola chromolaenae (holotype) a, b Appearance of conidiomata on substrate. c Section through conidioma. d Peridium. e–g Conidiogenous cells and developing conidia. h–j Conidia. k Conidia with polar appendages in Indian ink. Scale bars: a, b = 500 µm, c = 100 µm, d–k = 10 µm from both ends. Colonies on MEA circular, mycelium crateriform, undulate, cultures white with greyish from the centre of the colony at the surface, olivaceous from the centre of the colony in reverse with creamy-white at the margin (Fig. 60b). Pre-screening for antimicrobial activity: Murichromolaenicola chromolaenae (MFLUCC 17-1489) showed antimicrobial activity against E. coli with a 17 mm inhibition zone, when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 April 2017, A. Mapook (DP88, MFLU 20-0327, holotype); extype culture MFLUCC 17-1489. GenBank numbers: LSU: MN994560, ITS: MN994583, SSU: MN994606, TEF1: MN998164 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Murichromolaenicola chromolaenae (MFLUCC 17-1489, ex-holotype) with 90.65% similarity was Phaeosphaeria herpotrichoides (strain CBS 306.71, MH860137). The closest match with the LSU sequence with 98.90% similarity was Allophaeosphaeria muriformia (strain MFLUCC 13-0277, KX910089). The closest match with the SSU sequence with 99.82% similarity was Phaeosphaeria avenaria f. sp. triticae (strain ATCC 26370, EU189210), while the closest match with the TEF1 sequence with 99.82% similarity was Wojnowicia italica (strain MFLU 14-0732, KX430003). In the present phylogenetic analysis, M. chromolaenae clusters 13 Fig. 60 Culture characteristics on MEA: a Murichromolaenicola chiangraiensis (MFLUCC 17-1488). b Murichromolaenicola chromolaenae (MFLUCC 17-1489) with M. chiangraiensis (MFLUCC 17-1488) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 52). Although we could not compare the morphological characteristics of these species; M. chromolaenae is found as asexual morph in nature, while M. chiangraiensis is found as sexual morph in nature and we could not obtain its asexual morph in culture, but they differ in culture characteristics on MEA (Fig. 60). A comparison of the TEF1 gene region of M. chromolaenae and M. chiangraiensis reveals 41 base pair differences (4.7%) across 872 nucleotides. Therefore, M. chromolaenae is described as a new species based on phylogeny. Neoophiobolus Mapook & K.D. Hyde, gen. nov. Fungal Diversity Index Fungorum number: IF557343, Facesoffungi number: FoF 07808 Etymology: Etymology: Referring to its similarity with Ophiobolus. Saprobic on dead stems. Sexual morph: Ascomata immersed to semi-immersed, solitary, scattered, globose to subglobose, coriaceous, dark brown. Ostiole short papillate, with hyaline periphyses-like. Peridium comprising several layers of hyaline or pale brown to dark brown, pseudoparenchymatous cells, arranged in a textura angularis. Hamathecium composed of cylindrical to broadly filiform, septate, branching pseudoparaphyses, anastomosing above the asci. Asci overlapping, 8-spored, bitunicate, cylindrical to cylindric-clavate, straight or slightly curved, apically rounded, with ocular chamber visible when immature. Ascospores fasciculate, in parallel, scolecosporous, hyaline to pale yellow brown, filiform, 10–20-septate with minute guttule in each cell, slightly curved, constricted at the central septum where the spore separates into two parts, without polar appendages. Asexual morph: Undetermined. Type species: Neoophiobolus chromolaenae Mapook & K.D. Hyde Notes: Neoophiobolus is similar to the genus Ophiobolus in having scolecosporous ascospores with a swollen cell. However, phylogenetic analyses based on combined dataset of LSU, ITS, SSU and TEF1 sequence data show that N. chromolaenae is phylogenetically distant from Ophiobolus and forms a distinct lineage within the family Phaeosphaeriaceae (Fig. 52). Therefore, we introduce Neoophiobolus as a new genus to accommodate a new species N. chromolaenae. Neoophiobolus chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557344, Facesoffungi number: FoF 07809; Fig. 61 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0335 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 230–400 µm high × 200–350 µm diam. ( x̄ = 323.5 × 296 µm, n = 5), immersed to semi-immersed, solitary, scattered, globose to subglobose, coriaceous, dark brown. Ostiole short papillate, with hyaline periphyses-like. Peridium 15–30 µm wide, comprising several layers of hyaline or pale brown to dark brown, pseudoparenchymatous cells, arranged in a textura angularis. Hamathecium comprising 1.5–2.5 µm wide, cylindrical to broadly filiform, septate, branching pseudoparaphyses, anastomosing above the asci. Asci (90–)100–135 × 8–13 µm ( x̄ = 115 × 10 µm, n = 20), overlapping, 8-spored, bitunicate, cylindrical to cylindric-clavate, straight or slightly curved, apically rounded, with ocular chamber visible when immature. Ascospores 85–110 × 2.5–3.5 µm ( x̄ = 97 × 3 µm, n = 20), fasciculate, in parallel, scolecosporous, hyaline to pale yellow brown, filiform, 10–20-septate with minute guttule in each cell, slightly curved, constricted at the central septum where the spore separates into two parts, without polar appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, moderately fluffy, entire to filamentous, cultures white at the surface, creamy-white in reverse with brown to dark brown from the centre of the colony (Fig. 62). Pre-screening for antimicrobial activity: Neoophiobolus chromolaenae (MFLUCC 17-1467) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Mae Salong, on dead stems of Chromolaena odorata, 8 April 2017, A. Mapook (DMS4, MFLU 20-0335, holotype); ex-type culture MFLUCC 17-1467; Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP42, MFLU 20-0336); living culture MFLUCC 17-1449. GenBank numbers: LSU: MN994561, MN994562, ITS: MN994584, MN994585, SSU: MN994607, MN994608, TEF1: MN998165, MN998166 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS and TEF1 sequences of Neoophiobolus chromolaenae (MFLUCC 17-1467, ex-holotype) with 99.31% (LT796836) and 99.16% (LT797076) similarity, respectively, was Trematophoma sp. strain UTHSC: DI16-210. The closest match with the LSU sequence with 99.12% similarity was Paraphoma fimeti (strain CBS 127796, MH876144), while the closest match with the SSU sequences with 96.89% similarity was Ophiosphaerella narmari (strain ATCC 64688, KC848510). Paraleptospora Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557347, Facesoffungi number: FoF 07810 Etymology: Referring to its similarity with Leptospora. Saprobic on dead stems. Sexual morph: Ascomata immersed to semi-immersed, solitary or scattered, gregarious, coriaceous, globose or subglobose to ampulliform, brown to dark brown, appearing as dark spot with red area on host surface. Ostiole short papillate. Peridium several layers, comprising dense, thick-walled, reddish brown to dark brown, pseudoparenchymatous cells, arranged in textura angularis. Hamathecium composed of filiform or broadly filiform to cylindrical, septate, branching, pseudoparaphyses, anastomosing above the asci. Asci 8-spored, bitunicate, fissitunicate, cylindrical to cylindric-subclavate, straight or slightly curved, pedicellate, apically rounded, with an ocular chamber. Ascospores overlapping, 1–3-seriate, hyaline or pale yellow to yellowish brown, cylindric-fusiform, tapering towards narrow the rounded ends, (5–)6–7(–8)-septate, broader at the center and slightly constricted at septa, straight to slightly curved, guttulate, with or without polar appendages. Asexual morph: Undetermined. 13 Fungal Diversity Fig. 61 Neoophiobolus chromolaenae (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g, h Asci. i–m Ascospores. Scale bars: a = 500 µm, b = 200 µm, c = 100 µm, d, g, h = 50 µm, e, i–m = 20 µm, f = 5 µm Type species: Paraleptospora chromolaenae Mapook & K.D. Hyde Notes: Paraleptospora is similar to the genus Leptospora in staining the host surface red. However, phylogenetic analyses based on combined dataset of LSU, ITS, SSU and TEF1 sequence data show that Paraleptospora species are phylogenetically distant from Leptospora and form a sister lineage with Acericola, Jeremyomyces and Phaeosphaeriopsis (Fig. 52). Therefore, we introduce Paraleptospora as a new genus to accommodate two new species P. chromolaenae and P. chromolaenicola. Paraleptospora chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557348, Facesoffungi number: FoF 07811; Fig. 63 13 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0343 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 220–305 µm high × 210–340 µm diam. ( x̄ = 265 × 260 µm, n = 5), immersed to semi-immersed, solitary or scattered, gregarious, globose to subglobose, coriaceous, dark brown, appearing as dark spot with red area on host surface. Ostiole short papillate. Peridium 15–20 µm wide, several layers, comprising dense, thick-walled, reddish brown to dark brown, pseudoparenchymatous cells, arranged in textura angularis. Hamathecium comprising 1–2 µm wide, broadly filiform, septate, branching, pseudoparaphyses, anastomosing above the asci. Asci 80–120 × 7.5–11 µm ( x̄ = 100 × 9 µm, Fungal Diversity present phylogenetic analysis, P. chromolaenae clusters with P. chromolaenicola with high bootstrap support (100% ML and 1.00 BYPP, Fig. 52). However, P. chromolaenae differs from P. chromolaenicola in having slightly larger ascomata (220–305 × 210–340 µm vs. 140–310 × 120–300 µm) and slightly larger ascospores (30–40 × 3–4 µm vs. 28–38 × 3–4 µm) with 6–7-septa, and without polar appendages, while P. chromolaenicola has (5–)7–8-septate ascospores, and very few ascospores with polar appendages (Table 12). A comparison of the ITS (+5.8S) gene region of P. chromolaenae and P. chromolaenicola reveals 23 base pair differences (4.6%) across 503 nucleotides. Therefore, P. chromolaenae is described as a new species based on phylogeny and morphological comparison. Fig. 62 Culture characteristic on MEA: Neoophiobolus chromolaenae (MFLUCC 17-1467) n = 30), 8-spored, bitunicate, fissitunicate, cylindrical to cylindric-subclavate, slightly curved, pedicellate, apically rounded, with an ocular chamber. Ascospores 30–40 × 3–4 µm ( x̄ = 35 × 3.5 µm, n = 20), overlapping, 1–2-seriate, hyaline to pale yellow, cylindric-fusiform, tapering towards narrow the rounded ends, 6–7-septate, broader at the center and slightly constricted at septa, straight to slightly curved, guttulate, without polar appendages. Asexual morph: Undetermined. Pre-screening for antimicrobial activity: Paraleptospora chromolaenae (MFLUCC 17-1481) showed antimicrobial activity against E. coli with an 8 mm inhibition zone, when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium umbonate, cultures white at the surface, creamy-white in reverse (Fig. 65a). Material examined: THAILAND, Lampang Province, Ngao, on dead stems of Chromolaena odorata, 21 September 2016, A. Mapook (LP3, MFLU 20-0343, holotype); extype culture MFLUCC 17-1481. GenBank numbers: LSU: MN994563, ITS: MN994586, SSU: MN994609, TEF1: MN998167 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Paraleptospora chromolaenae (MFLUCC 17-1481, ex-holotype) with 91.36% similarity was Phaeosphaeriaceae sp. (strain MUT 4404, KC339239). The closest match with the LSU sequence with 99.34% similarity was Neostagonospora arrhenatheri (strain MFLUCC 15–0464, KX910091). The closest match with the SSU sequence with 99.71% similarity was Parastagonospora nodorum (strain LSNZN10, MH269310), while the closest match with the TEF1 sequence with 94.91% similarity was Yunnanensis phragmitis (strain MFLUCC 17-0365, MF683625). In the Paraleptospora chromolaenicola Mapook, Samarakoon & K.D. Hyde, sp. nov. Index Fungorum number: IF557349, Facesoffungi number: FoF 07812; Fig. 64 Etymology: Name reflects the host genus Chromolaena, on which this species was growing. Holotype: MFLU 18-0836 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 140–310 µm high × 120–300 µm diam., semi-immersed, solitary or scattered, gregarious, subglobose to ampulliform, coriaceous, brown to dark brown, appearing as dark spot with red area on host surface. Ostiole short papillate, flat top, red stained hyphae, with periphyses. Peridium 14–27.5 μm wide, several layers, comprising dense, thick-walled, brown to dark brown, pseudoparenchymatous cells, arranged in textura angularis. Hamathecium comprising 1–2 µm wide, filiform, septate, branching, pseudoparaphyses, anastomosing above the asci. Asci 100–120 × 8–11 µm ( x̄ = 107.5 × 9.4 μm, n = 30), 8-spored, bitunicate, fissitunicate, cylindrical to cylindric-clavate, slightly curved, pedicellate, apically rounded, with an ocular chamber. Ascospores 28–38 × 3–4 µm (x̄ = 33 × 3.5 μm, n = 30), overlapping, 2–3 seriate, hyaline or pale yellow to yellowish brown, cylindric-fusiform, tapering towards narrow the rounded ends, (5–)7–8-septate, broader at the center and slightly constricted at septa, straight to slightly curved, guttulate, very few ascospores with polar appendages, observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, moderately fluffy, entire to filamentous, cultures white at the surface, brown to dark brown from the centre of the colony with creamy-white at the margin in reverse (Fig. 65b). Pre-screening for antimicrobial activity: Paraleptospora chromolaenicola (MFLUCC 17-1450) showed antimicrobial activity against E. coli with a 12 mm inhibition zone, when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. 13 Fungal Diversity Fig. 63 Paraleptospora chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–q Ascospores. Scale bars: a, b = 500 µm, c = 100 µm, g–j = 50 µm, d, e, k–q = 20 µm, f = 5 µm Material examined: THAILAND, Chiang Mai Province, Mae Taeng, on dead stems of Chromolaena odorata, 1 September 2017, Milan C. Samarakoon, (SAMC016, MFLU 18-0836, holotype; HKAS 102326, isotype) ex-type culture MFLUCC 17-2670; Chiang Rai Province, Doi Pui, on dead 13 stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP43, MFLU 20-0344); living culture MFLUCC 17-1450. GenBank numbers: LSU: MN994564, MN994565, ITS: MN994587, MN994588, SSU: MN994610, MN994611, TEF1: MN998168, MN998169 Fungal Diversity Table 12 Synopsis of Paraleptospora species with similar morphological features discussed in this study Species Asocomata (µm) Peridium (µm) Asci (µm) Ascospores (µm) Septation of ascospores References P. chromolaenae (MFLUCC 17-1481) P. chromolaenicola (MFLUCC 17-1450) 220–305 high × 210–340 diam. 15–20 80–120 × 7.5–11 30–40 × 3–4 6–7-septate This study 140–310 high × 120–300 diam. 14–27.5 100–120 × 8–11 28–38 × 3–4 (5–)7–8-septate This study Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Paraleptospora chromolaenicola (MFLUCC 17-2670, ex-holotype) with 88.97% similarity was Phaeosphaeriaceae sp. (strain MUT 4404, KC339239). The closest match with the LSU sequence with 99.02% similarity was Loratospora aestuarii (strain CBS 117592, MH874575). The closest match with the SSU sequence with 100% similarity was Neosetophoma rosarum (strain MFLU 17-0308, NG_065146), while the closest match with the TEF1 sequence with 95.40% similarity was Neosetophoma sp. (strain UTHSC: DI16-337, LT797130). In the present phylogenetic analysis, P. chromolaenicola clusters with P. chromolaenae with high bootstrap support (100% ML and 1.00 BYPP, Fig. 52). However, P. chromolaenicola has (5–)7–8-septate ascospores, and very few ascospores with polar appendages, while P. chromolaenae differs from P. chromolaenicola in having slightly larger ascomata (220–305 × 210–340 µm vs. 140–310 × 120–300 µm) and slightly larger ascospores (30–40 × 3–4 µm vs. 28–38 × 3–4 µm), 6–7-septate, without polar appendages (Table 12). A comparison of the ITS (+5.8S) gene region of P. chromolaenicola and P. chromolaenae reveals 23 base pair differences (4.6%) across 503 nucleotides. Therefore, P. chromolaenicola is described as a new species based on phylogeny and morphological comparison. Pseudoophiosphaerella J.F. Zhang, J.K. Liu & Z.Y. Liu 2019 Pseudoophiosphaerella was introduced by Zhang et al. (2019) with P. huishuiensis as the type species from China, based on morphology and phylogeny. We introduce a new host record of Pseudoophiosphaerella huishuiensis based on morphology and molecular data, together with a description and illustrations (Fig. 66). A phylogenetic tree based on combined LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 52. Pseudoophiosphaerella huishuiensis J.F. Zhang, J.K. Liu & Z.Y. Liu, in Zhang, Liu, Jeewon, Wanasinghe & Liu, Mycosphere 8(1): 207 (2019) Facesoffungi number: FoF 05836; Fig. 66 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 275–350 µm high × 160–190 µm diam. (x̅ = 300 × 170 µm, n = 5), immersed to semi-immersed, solitary or scattered, appearing as dark spots, coriaceous, subglobose to obpyriform, dark brown to black with hyaline at side to basal. Ostiolar neck protruding, papillate, with hyaline periphyses-like. Peridium (15–)25–50 µm wide, comprising several layers, of hyaline to dark brown, pseudoparenchymatous cells, arranged in a textura angularis. Hamathecium comprising 1.5–3 µm wide, cylindrical, septate, branching pseudoparaphyses, anastomosing above the asci. Asci 140–175 × 7.5–10 µm ( x̄ = 150 × 8.5 µm, n = 20), overlapping, 8-spored, bitunicate, cylindrical to cylindricclavate, straight or slightly curved, apically rounded, with a small ocular chamber. Ascospores 130–170 × 2–2.5 µm ( x̄ = 147 × 2.3 µm, n = 25), fasciculate, arranged spirally in the centre, scolecosporous, hyaline when immature, becoming pale brown to yellowish brown when mature, filiform, 15–16-septate with minute guttule in each cell, slightly curved, not constricted at the septa, without polar appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from several cells. Colonies on MEA irregular, initially mycelium white with creamy-white from the centre of the colony at the surface, slightly raised, velvety with moderately fluffy, undulate to lobate, yellowish brown from the centre of the colony in reverse with creamy-white at the margin, becoming grayish-brown on surface in old cultures, dark brown to black in reverse (Fig. 67a, b). Pre-screening for antimicrobial activity: Pseudoophiosphaerella huishuiensis (MFLUCC 17-1453) showed antimicrobial activity against B. subtilis, E. coli and M. plumbeus (10 mm, 8 mm and 10 mm inhibition zone, respectively), observable as partial inhibition, when compared to the positive control (26 mm, 9 mm, and 17 mm, respectively). Known hosts and distribution: On dead culms of unidentified herbaceous plant in China (Zhang et al. 2019). Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 13 Fungal Diversity Fig. 64 Paraleptospora chromolaenicola (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–i Asci. j–m Ascospores. Scale bars: a = 1000 µm, c, d = 50 µm, e, g–m = 20 µm, f = 10 µm 2015, A. Mapook (DP46, MFLU 20-0350); living culture MFLUCC 17-1453 (new host record); (DP40, MFLU 20-0349); living culture MFLUCC 17-1447; Chiang Mai Province, Fah Hom Pok, on dead stems of C. odorata, 27 September 2016, A. Mapook (FHP1, MFLU 20-0351); living culture MFLUCC 17-1463; Mae Hong Son Province, Mae Yen, Pai, on dead stems of C. odorata, 25 June 2016, A. Mapook (MY5, MFLU 20-0352); living culture MFLUCC 17-1471. GenBank numbers: LSU: MN994566, MN994567, MN994568, MN994569, ITS: MN994589, MN994590, MN994591, MN994592, SSU: MN994612, MN994613, MN994614, MN994615, TEF1: MN998170, MN998171, MN998172, MN998173 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS and TEF1 sequences of 13 MFLUCC 17-1453 with 99.47% (MK522509) and 99.03% (MK523389) similarity, respectively, was Pseudoophiosphaerella huishuiensis strain HS-13. The closest match with the LSU sequence with 99.56% similarity was Dematiopleospora alliariae (strain MFLUCC 13-0070, KX494877), while the closest match with the SSU sequence with 99.46% similarity was Phaeosphaeriaceae sp. (strain SYPF 7948, MF588884). In the present phylogenetic analysis, MFLUCC 17-1453 clusters with Pseudoophiosphaerella huishuiensis strain HS-13 with hight bootstrap support (100% ML and 1.00 BYPP, Fig. 52). We therefore, identify MFLUCC 17-1453 as Pseudoophiosphaerella huishuiensis based on phylogenetic analyses with morphological comparison (Table 13), and the isolates are introduced here as a new host record from Chromolaena odorata collected in Thailand. Fungal Diversity Pseudostaurosphaeria Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557354, Facesoffungi number: FoF 07813 Etymology: Refer r ing to its similar ity with Staurosphaeria. Saprobic on dead stems. Sexual morph: Undetermined. Asexual morph: Conidiomata pycnidial, solitary, immersed or semi-immersed to superficial, uni-loculate, globose to subglobose, pale brown to brown or dark brown. Ostiole central, short papillate or apapillate. Pycnidial wall comprising 3–4 layers of thick-walled, hyaline or pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells holoblastic, ampulliform to cylindrical or oblong, hyaline, unbranched. Conidia globose or oblong to obovoid, 1–2-transverse septa, later developing 1–2-vertical septa, muriform, smoothwalled, slightly constricted at the septa, yellowish brown to brown or reddish brown, with single polar appendage from apex, observed clearly when mounted in Indian ink. Type species: Pseudostaurosphaeria chromolaenicola Mapook & K.D. Hyde Notes: Pseudostaurosphaeria is similar to the asexual morph of Staurosphaeria, which belongs to the family Coniothyriaceae (Wanasinghe et al. 2017), in having reddish brown conidia with a transverse septum, and later developing vertical septa dividing the conidium into four compartments. A phylogenetic analyses based on combined dataset of LSU, ITS, SSU and TEF1 sequence data show that Pseudostaurosphaeria species are phylogenetically distant from Staurosphaeria and form a sister lineage with Yunnanensis instead (Fig. 52). However, the asexual morph of Yunnanensis has been reported as having muriform conidia, ellipsoidal to obovoid with 3-transversely septate, and 1 vertical septum. Therefore, we introduce Pseudostaurosphaeria as a new genus to accommodate two new species Pseu. chromolaenae and Pseu. chromolaenicola. Pseudostaurosphaeria chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557355, Facesoffungi number: FoF 07814; Fig. 68 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0358 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata 110–145 µm high × 95–140 µm diam. ( x̄ = 130 × 118 µm, n = 10), pycnidial, solitary, semi-immersed to superficial, uni-loculate, subglobose, brown to dark brown, ostiolate, apapillate. Pycnidial wall 9–15 µm wide, comprising 3–4 layers of thick-walled, pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells holoblastic, ampulliform Fig. 65 Culture characteristics on MEA: a Paraleptospora chromolaenae (MFLUCC 17-1481), b Paraleptospora chromolaenicola (MFLUCC 17-1450) to oblong, hyaline, unbranched. Conidia 7–10 × 5–8.5 μm ( x̄ = 8.5 × 7 µm, n = 50), globose or oblong to obovoid, 1–2-transverse septa, later developing 1–2-vertical septum, muriform, smooth-walled, slightly constricted at the septa, yellowish brown to reddish brown, with single polar appendage from apex, observed clearly when mounted in Indian ink. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium umbonate, entire to filamentous, cultures white at the surface, yellowish brown to brown from the centre of the colony in reverse with creamy-white at the margin (Fig. 70a). Pre-screening for antimicrobial activity: Pseudostaurosphaeria chromolaenae (MFLUCC 17-1490) showed antimicrobial activity against B. subtilis with a 10 mm inhibition zone and against M. plumbeus with a 25 mm inhibition zone, observable as partial inhibition, when compared to the positive control (27 mm and 19 mm, respectively), but no inhibition of E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 April 2017, A. Mapook (DP89, MFLU 20-0358, holotype); extype culture MFLUCC 17-1490. GenBank numbers: LSU: MN994570, ITS: MN994593, SSU: MN994616, TEF1: MN998174 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Pseudostaurosphaeria chromolaenae (MFLUCC 17-1490, ex-holotype) with 89.07% similarity was Neosetophoma italica (strain FC-3846, LC206635). The closest match with the LSU sequence with 98.90% similarity was Phaeosphaeria sinensis (strain C454, MK348022). The closest match with the SSU sequence with 99.45% similarity was Phaeosphaeria avenaria f. sp. triticae (strain ATCC 26370, EU189210), while the closest match with the TEF1 sequence with 97.00% similarity was Paraphoma radicina (strain UTHSC: DI16-209, LT797075). In the present phylogenetic analysis, Pseu. chromolaenae clusters with Pseu. chromolaenicola with high bootstrap 13 Fungal Diversity Fig. 66 Pseudoophiosphaerella huishuiensis (new host record) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g, h Asci. i–m Ascospores. Scale bars: a, b = 500 µm, c, g–m = 50 µm, d, e = 20 µm, f = 5 µm support (100% ML and 1.00 BYPP, Fig. 52). However, Pseu. chromolaenae differs from Pseu. chromolaenicola in having larger conidiomata (110–145 × 95–140 μm vs. 105–120 × 85–120 μm) and larger conidia (7–10 × 5–8.5 μm vs. 6.5–9 × 5–6.5 μm), with globose or oblong to obovoid, 1–2-transverse septa, later developing 1–2-vertical septa, while Pseu. chromolaenicola has oblong to obovoid conidia with 1-transverse septum and later developing 1-vertical septum (Table 14). A comparison of the ITS (+5.8S) gene region of Pseu. chromolaenae and Pseu. chromolaenicola reveals 59 base pair differences (10.9%) across 539 13 nucleotides. Therefore, Pseu. chromolaenae is described as a new species based on phylogeny and morphological comparison. Pseudostaurosphaeria chromolaenicola Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557356, Facesoffungi number: FoF 07815; Fig. 69 Etymology: Name reflects the host genus Chromolaena, on which this species was growing. Holotype: MFLU 20-0360 Fungal Diversity Fig. 67 Culture characteristics on MEA: a, b Pseudoophiosphaerella huishuiensis (MFLUCC 17-1453) Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata 105–120 µm high × 85–120 µm diam. ( x̄ = 113 × 100 µm, n = 5), pycnidial, solitary, immersed to semi-immersed, uniloculate, globose to subglobose, pale brown to light brown. Ostiole central, short papillate. Pycnidial wall 5–15 µm wide, comprising 3–4 layers of thick-walled, hyaline or pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells holoblastic, ampulliform to cylindrical, hyaline and unbranched. Conidia 6.5–9 × 5–6.5 μm ( x̄ = 7.6 × 6 µm, n = 45), oblong to obovoid, 1-transverse septum, later developing 1-vertical septum, muriform, smooth-walled, slightly constricted at the septum, yellowish brown to brown with single polar appendage from apex, observed clearly when mounted in Indian ink. Culture characteristics: Conidia germinating on MEA within 48 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium umbonate, entire, cultures white at the surface, olivaceousbrown to dark brown from the centre of the colony in reverse with creamy-white at the margin (Fig. 70b). Pre-screening for antimicrobial activity: Pseudostaurosphaeria chromolaenicola (MFLUCC 17-1497) showed antimicrobial activity against M. plumbeus with an 18 mm inhibition zone, observable as partial inhibition, when compared to the positive control (18 mm), but no inhibition of E. coli and B. subtilis. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 April 2017, A. Mapook (DP90, MFLU 20-0359); living culture MFLUCC 17-1491; (DP96, MFLU 20-0360, holotype); extype culture MFLUCC 17-1497. GenBank numbers: LSU: MN994571, MN994572, ITS: MN994594, MN994595, SSU: MN994617, MN994618, TEF1: MN998175, MN998176 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Pseudostaurosphaeria chromolaenicola (MFLUCC 17-1497, ex-holotype) with 93.91% similarity was Neosetophoma sp. (strain CFE-31, MK614774). The closest match with the LSU sequence with 98.46% similarity was Phaeosphaeria sinensis (strain C454, MK348022). The closest match with the SSU sequence with 99.82% similarity was Phaeosphaeria avenaria f. sp. triticae (strain ATCC 26370, EU189210), while the closest match with the TEF1 sequence with 96.81% similarity was Yunnanensis phragmitis (strain MFLUCC 17-0365, MF683625). In the present phylogenetic analysis, Pseu. chromolaenicola clusters with Pseu. chromolaenae (MFLUCC 17-1490) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 52). However, Pseu. chromolaenicola differs from Pseu. chromolaenae in having oblong to obovoid conidia with 1-transverse septum and later developing 1-vertical septum, while Pseu. chromolaenae has larger conidiomata (110–145 × 95–140 μm vs. 105–120 × 85–120 μm) and larger conidia (7–10 × 5–8.5 μm vs. 6.5–9 × 5–6.5 μm), with globose or oblong to obovoid, 1–2-transverse septa, later developing 1–2-vertical septa (Table 14). A comparison of the ITS (+5.8S) gene region of Pseu. chromolaenicola and Pseu. chromolaenae reveals 59 base pair differences (10.9%) across 539 nucleotides. Therefore, Pseu. chromolaenicola is described as a new species based on phylogeny and morphological comparison. Yunnanensis Karun., Phook. & K.D. Hyde Yunnanensis was introduced by Karunarathna et al. (2017) with Yu. phragmitis as the type species, based on morphology and phylogeny. The species was collected on stems of Phragmites australis (Poaceae) from China. We Table 13 Morphological features of Pseudoophiosphaerella species discussed in this study Species Asocomata (µm) P. huishuiensis (MFLUCC 17-1463) P. huishuiensis (MFLUCC 19-0164) Peridium (µm) Asci (µm) Ascospores (µm) Septation of ascospores References (15–)25–50 275–350 high × 160–190 diam. 140–175 × 7.5–10 130–170 × 2–2.5 15–16-septate This study – 143–151.5 × 8–9.5 130.5–151.5 × 2.5–3.5 Multi-septate Zhang et al. (2019) 25–34 13 Fungal Diversity Fig. 68 Pseudostaurosphaeria chromolaenae (holotype) a, b Appearance of conidiomata on substrate. c Section through of conidioma. d Peridium. e–g Conidiogenous cells and developing conidia. h–m Micro- and macro-conidia with appendages n Conidia with appendages in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c = 50 µm, d, h = 10 µm, e–g, i–n = 5 µm introduce a new Yunnanensis species from C. odorata, based on morphology and molecular data, together with descriptions and illustrations (Fig. 71). A phylogenetic tree based on combined LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 52. Yunnanensis chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557293, Facesoffungi number: FoF 07816; Fig. 71 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0384 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 130–170(–210) µm high × 140–225 µm diam. ( x̄ = 167 × 175 µm, n = 10), semi-immersed to superficial, solitary or scattered, gregarious, appearing as dark, longitudinal spots, coriaceous, globose, brown to dark brown. Ostiole central. Peridium 10–15.5 µm wide, 2–3 layers, comprising pale brown to brown cells of textura angularis. Hamathecium comprising 1.5–3 µm wide, cylindrical to broadly filiform, septate, branching pseudoparaphyses. Asci 95–125(–155) × 19–28 µm ( x̄ = 112 × 23.5 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindric-clavate to obovoid, straight or slightly curved, apically rounded, pedicellate. Ascospores 20–28 × 10–12.5 µm ( x̄ = 24 × 11 µm, n = 20), overlapping, 1–2-seriate, initially hyaline to pale yellow, 1-septate when immature, becoming goldenbrown at maturity, ellipsoid to broadly fusiform, muriform, 4–6-transversely septate, with 1 vertical septum, constricted at the central septum, straight or slightly curved, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Table 14 Synopsis of Pseudostaurosphaeria species with similar morphological features discussed in this study Species Conidiomata (μm) Pycnidial wall (μm) Conidia (μm) Septate of conidia P. chromolaenae (MFLUCC 17-1490) P. chromolaenicola (MFLUCC 17-1497) 110–145 high × 95–140 diam. 9–15 7–10 × 5–8.5 105–120 high × 85–120 diam. 5–15 6.5–9 × 5–6.5 This study Globose or oblong to obovoid, 1–2-transverse septa, later developing 1–2-vertical septa This study Oblong to obovoid, 1-transverse septum, later developing 1-vertical septum 13 Reference Fungal Diversity Fig. 69 Pseudostaurosphaeria chromolaenicola (holotype) a, b Appearance of conidiomata on substrate. c Section through of conidioma. d Ostiole. e Peridium. f-i Conidiogenous cells and developing conidia. j-l Conidia with appendages m Conidia with appendages in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c, d = 20 µm, e-m = 10 µm Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from several cells. Colonies on MEA circular, mycelium crateriform, entire, cultures white at the surface, dark brown to olivaceous from the centre of the colony in reverse with creamy-white at the margin (Fig. 72). Pre-screening for antimicrobial activity: Yunnanensis chromolaenae (MFLUCC 17-1487) showed antimicrobial activity against E. coli with an 8 mm inhibition zone and against M. plumbeus with a 35 mm inhibition zone, observable as partial inhibition, when compared to the positive control (9 mm and 19 mm, respectively), but no inhibition of B. subtilis. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 April 2017, A. Mapook (DP85, MFLU 20-0384, holotype); extype culture MFLUCC 17-1487; (DP84, MFLU 20-0383); living culture MFLUCC 17-1486. GenBank numbers: LSU: MN994573, MN994574, ITS: MN994596, MN994597, SSU: MN994619, MN994620, TEF1: MN998177, MN998178 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS and TEF1 sequences of Yunnanensis chromolaenae (MFLUCC 17-1487, ex-holotype) with 93.24% (KY768867) and 97.84% (MF683624) similarity, respectively, was Yunnanensis phragmitis strain MFLUCC 17-0315. The closest match with the LSU sequence with 98.79% similarity was Tintelnotia destructans (strain CBS 127737, NG_058274), while the closest match with the SSU sequence with 99.82% similarity was Phaeosphaeria sp. (strain S93-48, EU189216). A phylogenetic analyses based on combined dataset of LSU, ITS, SSU and TEF1 sequence data show that Yu. chromolaenae forms a distinct lineage with Yu. phragmitis with high bootstrap support (99% ML and 1.00 BYPP, Fig. 52). Yu. chromolaenae is similar to Yu. phragmitis in having muriform spores, fusiform to ellipsoidal, surrounded by hyaline gelatinous sheath. However, Yu. 13 Fungal Diversity Fig. 70 Culture characteristics on MEA: a Pseudostaurosphaeria chromolaenae (MFLUCC 17-1490). b Pseudostaurosphaeria chromolaenicola (MFLUCC 17-1497) Fig. 71 Yunnanensis chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Peridium. e Pseudoparaphyses. f–i Asci. j–n Ascospores. o Ascospores with gelatinous sheath in Indian ink. Scale bars: a = 200 µm, b = 100 µm, c, f–i = 50 µm, d = 20 µm, j–o = 10 µm, e = 5 µm 13 chromolaenae differs from Yu. phragmitis in having larger semi-immersed to superficial ascomata, appearing as longitudinal spots on host surface and larger ascospores with 4–6-transversely septate, and 1 vertical septum, while Yu. phragmitis has smaller immersed to erumpent ascomata and smaller ascospores with 3-transverse septa, and 1 vertical septum in the second and third cells (Table 15). A comparison of the ITS (+5.8S) gene region of Yu. chromolaenae and Yu. phragmitis reveals 32 base pair differences (5.97%) across 536 nucleotides. Therefore, we introduce a new species based on morphology and phylogeny. Fungal Diversity Pyrenochaetopsidaceae Valenzuela-Lopez et al. Pyrenochaetopsidaceae was introduced by ValenzuelaLopez et al. (2018) to accommodate Pyrenochaetopsis as the type genus, plus two new genera, Neopyrenochaetopsis and Xenopyrenochaetopsis based on phylogenetic analyses. The family was described as an asexual morph with seven new species provided with genomic sequences. The genus has been isolated from plants, cyst of plant parasitic nematodes, as well as water and soil samples (Boerema et al. 2004; Crous et al. 2014; Papizadeh et al. 2017). ValenzuelaLopez et al. (2018) isolated most of the new species from human respiratory tract (bronchial wash and sputum sample), human superficial tissue (foot skin, sinusitis tissue, toe nail, ear lesion and dermatitis sample) and human deep tissue (blood sample). Pyrenochaetopsis Gruyter, Aveskamp & Verkley, in de Gruyter, et al., Mycologia 102(5): 1076 (2010) Pyrenochaetopsis was introduced by de Gruyter et al. (2010) with P. leptospora as the type species. The genus is ecologically diverse and had been accepted as a member of Cucurbitariaceae (Gruyter et al. 2010, 2013; Doilom et al. 2013; Hyde et al. 2013; Papizadeh et al. 2017; Wijayawardene et al. 2012, 2014a, Wijayawardene et al. 2017a, b, 2018). However, Valenzuela-Lopez et al. (2018) showed that the genus formed a distinct monophyletic clade outside the Cucurbitariaceae. Presently, 13 epithets are listed in Index Fungorum (2020). In this study, Pyrenochaetopsis chromolaenae is introduced based on morphology and molecular data, together with a description and illustrations (Fig. 74). A phylogenetic tree based on combined LSU, ITS and RPB2 sequence data is presented in Fig. 73. Pyrenochaetopsis chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557359, Facesoffungi number: FoF 07817; Fig. 74 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0362 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (60–)100–110 µm high × 75–125 µm diam. ( x̄ = 95 × 110 µm, n = 5), superficial, appearing as small dark spots, coriaceous, solitary or scattered, globose, brown to dark brown. Ostiole short papillate, with numerous external reddish brown setae. Peridium 5–10 µm wide, comprising 2-layers of thin-walled, pale brown to brown cells of textura angularis. Hamathecium comprising 1.5–2.5 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 45–90 × 9–12 µm ( x̄ = 65 × 10.5 µm, n = 15), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short, bulbous pedicel apically rounded. Ascospores 14–24 × 4–6 µm ( x̄ = 19 × 5 µm, n = 15), overlapping 2–3-seriate, hyaline to pale brown to slightly yellowish brown, cylindrical to broadly fusiform, with slightly obtuse ends, widest at the center and tapering towards rounded ends, 3–4-septate, straight or slightly curved, second cell from apex is slightly inflated, guttulate, constricted at the septa, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, entire, white aerial hyphae at the surface, spreading and dark brown to olivaceous in reverse (Fig. 75). Pre-screening for antimicrobial activity: Pyrenochaetopsis chromolaenae (MFLUCC 17-1440) showed antimicrobial activity against E. coli with a 9 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP24, MFLU 20-0362, holotype); extype culture MFLUCC 17-1440. GenBank numbers: LSU: MT214472, ITS: MT214378, SSU: MT214424, TEF1: MT235790, RPB2: MT235827 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Pyrenochaetopsis chromolaenae (MFLUCC 17-1440, ex-holotype) with 99.81% similarity was Py. microspora (strain CBS 102876, NR_160059). The closest match with the LSU sequence with 99.88% similarity was Py. leptospora (strain CBS 131.69, MH871006), while the closest match with the RPB2 sequence with 95.35% similarity was Pyrenochaetopsis sp. (strain UTHSC:DI16-275, LT593072). In the present phylogenetic analysis, Py. chromolaenae forms a separate clade and clusters with Py. setosissima (CBS 119739) and Py. americana (UTHSC:DI16-225), which were found as asexual morphs from culture, with bootstrap support (67% ML and 0.98 BYPP, Fig. 73). However, Py. chromolaenae was found as the sexual morph in nature and we could not obtain its asexual morph in culture. A comparison of the RPB2 gene region of Py. chromolaenae and Py. americana reveals 48 base pair differences (5.4%) across 884 nucleotides. Therefore, Py. chromolaenae is described as a new species based on phylogeny. Roussoellaceae Liu et al. Roussoellaceae was introduced to accommodate three genera Neoroussoella, Roussoella and Roussoellopsis which have trabeculate pseudoparaphyses (sensu Liew et al. 2000) and brown bicelled ascospores (Liu et al. 2014). Previously, the family has been treated as a synonym of Thyridariaceae by Jaklitsch and Voglmayr (2016). However, the latest updated accounts based on multigene phylogenetic analysis show that Roussoellaceae is well-resolved in Pleosporales 13 Fungal Diversity Fig. 72 Culture characteristic on MEA: Yunnanensis chromolaenae (MFLUCC 17-1487) (Dai et al. 2017; Tibpromma et al. 2017; Hyde et al. 2018a, b; Jayasiri et al. 2019; Jiang et al. 2019; Phookamsak et al. 2019). Wanasinghe et al. (2018) introduced two roussoellalike taxa, Pararoussoella and Pseudoneoconiothyrium in Thyridariaceae. However, Phookamsak et al. (2019) showed that these two genera clustered with Roussoella in Roussoellaceae, based on increased taxon sampling in the phylogenetic analysis. This finding was supported by Jayasiri et al. (2019) and Jiang et al. (2019). Divergence time estimates for this family are crown age of 62 Mya (34–91) in the Cenozoic Era (Paleogene period) and stem age of 77 Mya (44–110) in the late Cretaceous (Liu et al. 2017). Pseudoroussoella Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557351, Facesoffungi number: FoF 07818 Etymology: Referring to its similarity with Roussoella. Saprobic on dead stems or petiole. Sexual morph: Ascomata immersed to erumpent through the host surface, solitary, appearing as dark spots, coriaceous, globose to subglobose, dark brown to black, Ostiolar neck protruding. Peridium several layers, inner layers comprising of hyaline to light brown cells of textura epidermoidea to textura angularis, outer layers comprising of brown to dark brown cells of textura intricata. Hamathecium composed of cylindrical to filiform, septate, trabeculate pseudoparaphyses. Asci 8-spored, bitunicate, fissitunicate, cylindrical to clavate, straight or slightly curved, apically rounded, short pedicellate with small ocular chamber. Ascospores uniseriate, initially hyaline to pale brown, septate when immature, becoming yellowish brown to dark brown at maturity, oval to ellipsoid, 1-septate, constricted at the septum, straight or slightly curved, slightly widest at the upper cell and tapering towards obtuse ends, with a reticulate spore wall ornamentation, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Conidiomata pycnidial, solitary, superficial, uniloculate, globose to obpyriform, yellowish brown to brown. Ostiole central, papillate. Pycnidial wall comprising of 2–3 layers, hyaline or pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells annellidic, ampulliform to oblong, hyaline and unbranched. Conidia oblong to oval, pale brown to light brown when immature, becoming yellowish brown to reddish brown when mature, aseptate, thick-walled with guttulate. Type species: Pseudoroussoella elaeicola (Konta & K.D. Hyde) Mapook & K.D. Hyde Notes: A phylogenetic analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data show that two Pseudoroussoella species form a separate clade and group with Pseudoneoconiothyrium rosae and Roussoella euonymi which were found as asexual morphs, with bootstrap support (83% ML, Fig. 76). Asexual morph of Pseudoroussoella species differ from Pseudoneoconiothyrium rosae and R. euonymi in smaller oblong to oval conidia, while Pseudoneoconiothyrium rosae has globose to irregularly ellipsoid conidia (Table 16). Therefore, we introduce Pseudoroussoella as a new genus with two new species based on morphology and phylogeny. Pseudoroussoella chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557353, Facesoffungi number: FoF 07332; Fig. 77 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Table 15 Synopsis of Yunnanensis species with similar morphological features discussed in this study Species Asocomata (µm) Peridium (µm) Asci (µm) 130–170(–210) high × 140–225 10–15.5 Yu. chrodiam. molaenae (MFLUCC 17-1487) 100–130 high, 115–170 diam. 7–12 Yu. phragmitis (MFLUCC 17-0315) 13 Ascospores (µm) Septate of ascospores References 95–125(–155) × 19–28 20–28 × 10–12.5 4–6-transversely sep- This study tate, with 1-vertical septum 61–113 × 23–31 3-transversely septate, and 1-vertical septum 11–13 × 4–7 Karunarathna et al. (2017) Fungal Diversity Table 16 Synopsis of asexual morph of Pseudoroussoella species with similar morphological features discussed in this study Species Conidiomata (µm) Conidiogenous cells (µm) Roussoella euonymi (CBS 143426) 150–300 diam. 5–12 × 5–7 Pseudoneoconiothyrium rosae (MFLUCC 190–240 high × 120–170 diam. 4–7 × 3–7 15-0052) Pseudoroussoella chromolaenae 130–175(–230) high × 160–230 diam. – (MFLUCC 17-1492) Holotype: MFLU 20-0356 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata 130–175(–230) µm high × 160–230 µm diam. ( x̄ = 165 × 195 µm, n = 5), pycnidial, solitary, superficial, uni-loculate, globose to obpyriform, yellowish brown to brown. Ostiole central, papillate. Pycnidial wall 10–20 µm wide, comprising 2–3 layers of hyaline or pale brown to light brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells annellidic, ampulliform to oblong, hyaline and unbranched. Conidia 5.5–7 × 3.5–5 μm ( x̄ = 6.5 × 4.5 µm, n = 50), oblong to oval, pale brown to light brown when immature, becoming yellowish brown to reddish brown when mature, aseptate, thickwalled with guttule. Culture characteristics: Conidia germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, velvety with moderately fluffy, filamentous, cultures white with grayish-brown from the centre and near the margin of the colony on surface, olivaceous in reverse from the centre of the colony with white at margin (Fig. 79b). Pre-screening for antimicrobial activity: Pseudoroussoella chromolaenae (MFLUCC 17-1492) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP91, MFLU 20-0356, holotype); ex-type culture MFLUCC 17-1492. GenBank numbers: LSU: MT214439, ITS: MT214345, SSU: MT214393, TEF1: MT235769 Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Pseudoroussoella chromolaenae (MFLUCC 17-1492, ex-holotype) with 99.10% similarity was Roussoella sp. (strain MFLUCC 17-2059, MH744730). The closest match with the LSU sequence with 98.59% similarity was Roussoella hysterioides (strain CBS 125434, MH875155). The closest match with the SSU sequence with 91.77% similarity was Roussoella intermedia (strain KT 2303, AB524483), while the closest match with the TEF1 sequence with 96.49% similarity was Arthopyrenia Conidia (µm) References (6–)7(–8) × (4–)5–6 Crous et al. (2018b) 6–8 × 4–7 Wanasinghe et al. (2018) 5.5–7 × 3.5–5 This study sp. (strain UTHSC: DI16-334, LT797127). In the present phylogenetic analysis, P. chromolaenae clusters with P. elaeicola with high bootstrap support (100% ML and 1.00 BYPP, Fig. 76). A comparison of the TEF1 gene region of P. chromolaenae and P. elaeicola reveals 10 base pair differences (1.1%) across 899 nucleotides and the ITS (+5.8S) gene region reveals five base pair differences (1.1%) across 443 nucleotides. However, we could not compare the morphological characteristics of those species; P. chromolaenae is found as asexual morph in nature, while P. elaeicola is found as sexual morph in nature and we could not obtain its asexual morph in culture. Therefore, P. chromolaenae is described here as a new species based on phylogeny and culture characteristic comparison. Pseudoroussoella elaeicola (Konta & K.D. Hyde) Mapook & K.D. Hyde, comb. nov. Index Fungorum number: IF557352, Facesoffungi number: FoF 07819; Fig. 78 ≡ Roussoella elaeicola Konta & K.D. Hyde, in Phookamsak et al., Fungal Divers (2019) Holotype: THAILAND, Chiang Rai Province, on dead petiole of Elaeis guineensis (Arecaceae), 25 November 2014, S. Konta, HR02d (MFLU 15-0022), ex-type culture, MFLUCC 15-0276 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 225–475 µm high × 240–400 µm diam. ( x̄ = 365 × 325 µm, n = 5), immersed, solitary, appearing as dark spots, coriaceous, subglobose, dark brown to black. Ostiolar neck protruding. Peridium (20–)30–50 µm wide, several layers, inner layers comprising hyaline to light brown cells of textura epidermoidea, outer layers comprising brown to dark brown cells of textura intricata. Hamathecium comprising 1–2 µm wide, cylindrical to filiform, septate, trabeculate pseudoparaphyses. Asci (70–)95–135 × 6–8.5 µm ( x̄ = 108 × 7 µm, n = 10), 8-spored, bitunicate, cylindrical to clavate, straight or slightly curved, apically rounded, short pedicellate with small ocular chamber. Ascospores 10–14 × 4.5–6 µm ( x̄ = 12.5 × 5.5 µm, n = 25), uniseriate, initially hyaline to pale brown, septate when immature, becoming yellowish brown at maturity, oval to ellipsoid, 1-septate, constricted at the septum, straight or slightly curved, slightly 13 Fungal Diversity Fig. 73 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS and RPB2 sequence data. Fourty-nine strains are included in the combined sequence analysis, which comprise 2556 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 18819.011208 is presented. The matrix had 977 distinct alignment patterns, with 12.78% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.247993, C = 0.237502, G = 0.272601, T = 0.241905; substitution rates: AC = 1.771498, AG = 5.599945, AT = 1.899441, CG = 1.222461, CT = 10.018389, GT = 1.000000; gamma distribution shape parameter α = 0.168302. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given above or below the nodes. Newly generated sequences are in dark red bold and type species are in bold. Pleospora herbarum (CBS 191.86) and P. herbarum (IT956) are used as outgroup taxa widest at the upper cell and tapering towards obtuse ends, with a reticulate spore wall ornamentation, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, velvety with moderately fluffy, filamentous, cultures white, grayish-brown from the centre of the colony with white at margin on surface, brown to pale olivaceousbrown in reverse from the centre of the colony, white to creamy-white at margin (Fig. 79a). 13 Pre-screening for antimicrobial activity: Pseudoroussoella elaeicola (MFLUCC 17-1483) showed antimicrobial activity against E. coli with a 10 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Known hosts and distribution: On dead petiole of Elaeis guineensis (Arecaceae) in Thailand (Phookamsak et al. 2019). Material examined: THAILAND, Lampang Province, Ngao, on dead stems of Chromolaena odorata, 21 September 2016, A. Mapook (LP5, MFLU 20-0357); living culture MFLUCC 17-1483 (new host record). Fungal Diversity Fig. 74 Pyrenochaetopsis chromolaenae (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d Ostiole with brown setae. e Peridium. f Pseudoparaphyses. g–j Immature and mature asci. k–p Ascospores. Scale bars: a = 500 µm, b = 100 µm, c = 50 µm, g–j = 20 µm, d, e, k–p = 10 µm, f = 5 µm GenBank numbers: LSU: MT214442, ITS: MT214348, SSU: MT214396, TEF1: MT235772, RPB2: MT235808 Notes: A phylogenetic analyses show that the strain MFLUCC 17-1483 grouped with Pseudoroussoella elaeicola (= Roussoella elaeicola) (Fig. 76). In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of our strains with 99.80% similarity was Roussoella sp. (strain MFLUCC 17-2059, MH744730). The closest match with the LSU sequence with 98.47% similarity was Arthopyrenia salicis (strain MUT < ITA > :4879, KP671722). The closest match with the SSU sequence with 93.95% similarity was Roussoella intermedia (strain KT 2303, AB524483). The closest match with the TEF1 sequence with 96.06% similarity was Arthopyrenia sp. (strain UTHSC: DI16-334, LT797127), while the closest match with the RPB2 sequence with 90.92% similarity was R. euonymi (strain CBS 143426, MH108007). We therefore, identify our isolates as P. elaeicola based on phylogenetic analyses with morphological comparison (Table 17). In this study, we isolated P. elaeicola from Chromolaena odorata collected in Thailand, and the isolate is introduced here as a new host record. 13 Fungal Diversity Fig. 75 Culture characteristic on MEA: Pyrenochaetopsis chromolaenae (MFLUCC 17-1440) Setoarthopyrenia Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557361, Facesoffungi number: FoF 07820 Etymology: Referring to the Ostiolar neck with numerous external setae and its ascospores similarity with Arthopyrenia. Saprobic on dead stems. Sexual morph: Ascomata superficial, solitary, appearing as small spots, coriaceous, globose, brown to dark brown. Ostiolar neck protruding, with numerous external brown to dark brown setae. Peridium 2–3 layers, comprising brown to dark brown cells of textura angularis. Hamathecium composed of filiform to broadly filiform, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, cylindric-clavate, straight or slightly curved, apically rounded, short pedicellate. Ascospores bi-seriate, hyaline, ellipsoid to obovoid, 1-septate, guttulate, constricted at the septum, straight, without gelatinous sheath. Asexual morph: Undetermined. Type species: Setoarthopyrenia chromolaenae Mapook & K.D. Hyde Notes: A phylogenetic analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data show that Setoarthopyrenia chromolaenae forms a separate clade with Roussoella clade, with bootstrap support (83% ML and 0.99 BYPP, Fig. 76). Setoarthopyrenia differs from Roussoella in having a protruding ostiole neck, with numerous external brown to dark brown setae, bi-seriate ascospores that are hyaline, ellipsoid to obovoid, 1-septate without gelatinous sheath, while Roussoella species have uniseriate ascospores that are brown, fusiform-ellipsoidal, ornamented and surrounded by a wide mucilaginous sheath (Liu et al. 2014). Therefore, we introduce Setoarthopyrenia as a new genus with a new species, S. chromolaenae based on morphology and phylogeny. 13 Setoarthopyrenia chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557362, Facesoffungi number: FoF 07821; Fig. 80 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0368 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 65–105 µm high × 80–95 µm diam. (x̅ = 85 × 88 µm, n = 5), superficial, solitary, appearing as small spots, coriaceous, globose, brown to dark brown. Ostiolar neck protruding, with numerous external brown to dark brown setae. Peridium (4.5–)5–10 µm wide, 2–3 layers, comprising brown to dark brown cells of textura angularis. Hamathecium comprising 0.5–1.5 µm wide, filiform to broadly filiform, septate, branching pseudoparaphyses. Asci 40–60 × 12–15 µm (x̅ = 55 × 13.5 µm, n = 10), 8-spored, bitunicate, cylindric-clavate, straight or slightly curved, apically rounded, short pedicellate. Ascospores 14–17 × 6–8 µm (x̅ = 15 × 7 µm, n = 25), bi-seriate, hyaline, ellipsoid to obovoid, 1-septate, guttulate, constricted at the septum, straight, without gelatinous sheath. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, entire, cultures white, olivaceous-brown at margin on surface, dark brown to olivaceous-brown from the centre of the colony in reverse with creamy-white appearing as concentric ring pattern and dark brown to olivaceousbrown at margin (Fig. 81). Pre-screening for antimicrobial activity: Setoarthopyrenia chromolaenae (MFLUCC 17-1444) showed antimicrobial activity against B. subtilis with a 16 mm inhibition zone, when compared to the positive control (26 mm), but no inhibition of M. plumbeus and E. coli. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP37, MFLU 20-0368, holotype); extype culture MFLUCC 17-1444. GenBank numbers: LSU: MT214438, ITS: MT214344, SSU: MT214392, TEF1: MT235768, RPB2: MT235805 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS, LSU, TEF1 and RPB2 sequences of Setoarthopyrenia chromolaenae (MFLUCC 17-1444, exholotype) with 99.83% (LT796905), 100% (LN907505), 99.78% (LT797145) and 96.35% (LT797065) similarity, respectively, was Arthopyrenia sp. strain UTHSC: DI16-362, while the closest match of the SSU sequence with 99.81% similarity was Roussoella intermedia (strain CBS 170.96, KF443390). In the present phylogenetic analysis, S. chromolaenae (MFLUCC 17-1444) is closely related to Arthopyrenia sp. UTHSC: DI16-362 (Fig. 76). However, we could not compare the morphological characteristics of both strains as Fungal Diversity Fig. 76 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data. Eightyfour strains are included in the combined sequence analysis, which comprise 4416 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 35074.673502 is presented. The matrix had 1690 distinct alignment patterns, with 39.45% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.246392, C = 0.254324, G = 0.268438, T = 0.230846; substitution rates: AC = 1.657341, AG = 4.989387, AT = 2.205689, CG = 1.285998, CT = 10.615734, GT = 1.000000; gamma distribution shape parameter α = 0.486933. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given above or below the nodes. Newly generated sequences and new combination are in dark red bold and type species are in bold. Occultibambusa bambusae (MFLUCC 11-0394) and O. bambusae (MFLUCC 13-0855) are used as outgroup taxa the morphology of Arthopyrenia sp. (UTHSC: DI16-362) has not been reported. Therefore, we introduce a new species based on morphology and phylogeny. Index Fungorum number: IF557367, Facesoffungi number: FoF 07822 Etymology: Xeno = ξένος in Greek, distinct; Roussoella = roussoella-like Xenoroussoella Mapook & K.D. Hyde, gen. nov. 13 Fungal Diversity Table 17 Synopsis of sexual morph species with similar morphological features discussed in this study Species Ascomata (µm) Peridium (µm) Asci (µm) Ascospores (µm) References Pseudoroussoella elaeicola (= Roussoella elaeicola, MFLUCC 15-0276) P. elaeicola (MFLUCC 17-1483, LP5) Xenoroussoella triseptata (MFLUCC 17-1438, DP22) 315–410 high, 325–350 diam. 25–70 70–140 × 6–9 10–15 × 3–6, 1-septate 225–475 high × 240–400 diam. (155–)170–220 high × 185– 235 diam. (20–)30–50 (70–)95–135 × 6–8.5 10–14 × 4.5–6, 1-septate Phookamsak et al. (2019) This study 10–15 55–70 × 10.5–15 13.5–17 × 5–7.5, 3-septate This study Fig. 77 Pseudoroussoella chromolaenae (holotype) a, b Appearance of conidiomata on substrate. c Section through of conidioma. d Ostiole e Peridium. f–g Conidiogenous cells and developing conidia. h–i Conidia. Scale bars: a, b = 200 µm, c = 50 µm, d, e = 20 µm, h, i = 10 µm, f, g = 5 µm Saprobic on dead stems. Sexual morph: Ascomata immersed, solitary or scattered, appearing as orange spots, coriaceous, globose to subglobose, brown to dark brown. Ostiolar neck protruding. Peridium 2–4 layers, comprising hyaline to brown cells of textura angularis. Hamathecium composed of oblong to cylindrical, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, cylindric-clavate to clavate, straight or slightly curved, apically rounded, short pedicellate. Ascospores bi-seriate, initially hyaline, 1-septate when immature, becoming pale brown to dark brown at maturity, ellipsoid to obovoid, 3-septate, with irregular longitudinal striations, constricted at the central septum, straight or slightly curved, without gelatinous sheath. Asexual morph: Undetermined. 13 Type species: Xenoroussoella triseptata Mapook & K.D. Hyde Notes: A phylogenetic analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data show that Xenoroussoella triseptata forms a separate branch and groups with the clade comprising Pseudoroussoella species, Pseudoneoconiothyrium rosae and Roussoella euonymi, with high bootstrap support (99% ML and 1.00 BYPP, Fig. 76). Xenoroussoella differs from Pseudoroussoella in having smaller ascomata and asci, thin peridium, larger bi-seriate ascospores (Table 16) that are ellipsoid to obovoid, 3-septate with irregular longitudinal striations, without gelatinous sheath, while Pseudoroussoella has uniseriate ascospores that are oval to ellipsoid, 1-septate, with a reticulate spore wall ornamentation, surrounded by hyaline Fungal Diversity Fig. 78 Pseudoroussoella elaeicola (new host record) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–o Ascospores. p Ascospores with gelatinous sheath in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c = 100 µm, g–j = 50 µm, d, e = 20 µm, f = 10 µm, k–p = 5 µm gelatinous sheath. Therefore, we introduce Xenoroussoella as a new genus with a new species, Xe. triseptata based on morphology and phylogeny. Xenoroussoella triseptata Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557368, Facesoffungi number: FoF 07823; Fig. 82 Etymology: The specific epithet “triseptata” is based on the 3-septate ascospores Holotype: MFLU 20-0382 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata (155–)170–220 µm high × 185–235 µm diam. ( x̄ = 185 × 210 µm, n = 5), immersed, solitary or scattered, appearing as orange spots, coriaceous, globose to subglobose, brown to dark brown. Ostiolar neck protruding. Peridium 10–15 µm wide, comprising 2–4 layers of hyaline to brown cells of textura angularis. Hamathecium comprising 1.9–3 µm wide, oblong to cylindrical, septate, branching pseudoparaphyses. Asci 55–70 × 10.5–15 µm ( x̄ = 65 × 12 µm, n = 20), 8-spored, 13 Fungal Diversity phylogeny and their distinct morphology in having 3-septate ascospores, with irregular longitudinal striations. Fig. 79 Culture characteristics on MEA: a Pseudoroussoella elaeicola (MFLUCC 17-1483). b Pseudoroussoella chromolaenae (MFLUCC 17-1492) bitunicate, cylindric-clavate to clavate, straight or slightly curved, apically rounded, short pedicellate. Ascospores 13.5–17 × 5–7.5 µm ( x̄ = 15 × 6.5 µm, n = 30), bi-seriate, initially hyaline, 1-septate when immature, becoming pale brown to dark brown at maturity, ellipsoid to obovoid, 3-septate, with irregular longitudinal striations, constricted at the central septum, straight or slightly curved, without gelatinous sheath. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, undulate, cultures white, pale grayish brown from the centre of the colony with olivaceous to olivaceousbrown at margin on surface, creamy-white in reverse, olivaceous-brown to brown from the centre of the colony with olivaceous to olivaceous-brown at margin (Fig. 83). Pre-screening for antimicrobial activity: Xenoroussoella triseptata (MFLUCC 17-1438) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP22, MFLU 20-0382, holotype); extype culture MFLUCC 17-1438. GenBank numbers: LSU: MT214437, ITS: MT214343, SSU: MT214391, TEF1: MT235767, RPB2: MT235804 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS, LSU, TEF1 and RPB2 sequences of Xenoroussoella triseptata (MFLUCC 17-1438, ex-holotype) with 99.65% (LT796887), 99.88% (LN907477), 99.38% (LT797127) and 97.80% (LT797047) similarity, respectively, was Arthopyrenia sp. strain UTHSC: DI16-334, while the closest match of the SSU sequence with 93.39% similarity was Roussoella intermedia (strain KT 2303, AB524483). In the present phylogenetic analysis, Xe. triseptata (MFLUCC 17-1438) is closely related to Arthopyrenia sp. UTHSC: DI16-334 (Fig. 76). However, we could not compare the morphological characteristics of both strains as the morphology of Arthopyrenia sp. (UTHSC: DI16-334) has not been reported. Therefore, we introduce a new species based on 13 Thyridariaceae Q. Tian & K.D. Hyde Family Thyridariaceae was introduced by Hyde et al. (2013) to accommodate Thyridaria as its type genus. Jaklitsch and Voglmayr (2016) synonymized Roussoellaceae under Thyridariaceae with introduction of a new genus Parathyridaria in the family. However, Tibpromma et al. (2017) reinstated Roussoellaceae based on morphology with phylogenetic analysis and treated it as a well-resolved family in Pleosporales. Wanasinghe et al. (2018) introduced three new genera (Cycasicola, Pseudoneoconiothyrium, Pararoussoella), of which Pseudoneoconiothyrium and Pararoussoella are presently placed in Roussoellaceae. Devadatha et al. (2018) also introduced a new genus Thyridariella with two new species in Thyridariaceae. Phookamsak et al. (2019) introduced a new genus Liua based on morphological distinction and phylogenetic support. Divergence time estimates for this family are crown age of 15 Mya (3–38) in the Cenozoic period and stem age of 95 Mya (60–138) in the Cretaceous (Liu et al. 2017). Chromolaenomyces Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557333, Facesoffungi number: FoF 07824 Etymology: Named after the host genus Chromolaena, combined with “myces” for fungi. Saprobic on dead stems. Sexual morph: Ascomata immersed, coriaceous, solitary or scattered, globose to subglobose, light brown to brown. Ostiolar neck protruding. Peridium 2–3 layers, pale brown to dark brown cells of textura angularis. Hamathecium composed of cylindrical to filiform, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, fissitunicate, cylindrical, straight or slightly curved, apically rounded, with a short pedicel. Ascospores uni-seriate, irregular arrangement, hyaline, oval to broadly fusiform, uni-septate, constricted at the septum, smooth, with a narrow sheath; sheath drawn out to form polar appendages from both ends of the ascospores, straight or slightly curved. Asexual morph: Undetermined. Type species: Chromolaenomyces appendiculatus Mapook & K.D. Hyde Notes: Phylogenetic analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data show that Chromolaenomyces appendiculatus forms a separate branch and groups with Thyridariella mangrovei, with low bootstrap support (Fig. 84). However, Chromolaenomyces differs from Thyridariella in having cylindrical asci with a short pedicel and ascospores that are uni-seriate, arranged irregularly, oval to broadly fusiform, uni-septate with a narrow sheath, drawn out to form polar appendages from both ends of the ascospores, while Thyridariella has clavate asci Fungal Diversity Fig. 80 Setoarthopyrenia chromolaenae (holotype) a, b Appearance of superficial ascomata on substrate. c Section through ascoma. d dark brown setae. e Peridium. f Pseudoparaphyses. g–j Immature and mature asci. k–p Ascospores. Scale bars: a = 500 µm, b = 100 µm, c, g–j = 20 µm, d, k–p = 10 µm, e, f = 5 µm with a moderately long pedicel and ascospores that are uniseriate to bi-seriate, fusiform to ellipsoidal, muriform, and surrounded by hyaline gelatinous sheath (Devadatha et al. 2018). Therefore, we introduce Chromolaenomyces as a new genus with a new species, C. appendiculatus based on morphology and phylogeny. Chromolaenomyces appendiculatus Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557334, Facesoffungi number: FoF 07825; Fig. 85 Etymology: Referring to ascospores with sheath drawn out to form polar appendages Holotype: MFLU 20-0307 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 175–185 µm high × 140–165 µm diam. ( x̄ = 180 × 155 µm, n = 5), immersed, coriaceous, solitary or scattered, globose to subglobose, light brown to brown. Ostiolar neck protruding. Peridium 7–12(–15) µm wide, 2–3 layers, pale brown to dark brown cells of textura angularis. Hamathecium comprising 0.5–1.5 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci (75–)85–105 × 10–14 µm ( x̄ = 95 × 12 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindrical, straight or slightly curved, apically rounded, with a short pedicel. Ascospores 15–20.5 × 6–7.5 µm ( x̄ = 17.5 × 7 µm, n = 20), uni-seriate, irregular arrangement, hyaline, oval to broadly fusiform, uni-septate, constricted at the septum, smooth, with a 13 Fungal Diversity similarity was Cycasicola leucaenae (strain MFLUCC 17-0914, MK434900). In the present phylogenetic analysis, C. appendiculatus (MFLUCC 17-1455) is closely related to Thyridariella mangrovei (Fig. 84). However, morphological characteristics of both strains are distinct. Fig. 81 Culture characteristic on MEA: Setoarthopyrenia chromolaenae (MFLUCC 17-1444) narrow sheath; sheath drawn out to form polar appendages (4.5–)6.5–12(–16) µm long × 2–3.5 µm wide ( x̄ = 9 × 2.5 µm, n = 30), from both ends of the ascospores, straight or slightly curved. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature, germ tubes produced from both ends of the ascospores. Colonies on MEA circular to irregular, mycelium slightly raised, entire, cultures olivaceous-grey at the surface with white margin and olivaceous-brown at the center in reverse, white to yellowwhite at the margin (Fig. 86a). Pre-screening for antimicrobial activity: Chromolaenomyces appendiculatus (MFLUCC 17-1455) showed antimicrobial activity against B. subtilis with a 20 mm inhibition zone and against M. plumbeus with a 13 mm inhibition zone, observable as partial inhibition, when compared to the positive control (26 mm and 17 mm, respectively), but no inhibition of E. coli. Material examined: THAILAND, Lampang Province, Chaehom, on dead stems of Chromolaena odorata, 24 September 2016, A. Mapook (JH1, MFLU 20-0307, holotype); ex-type culture MFLUCC 17-1455. GenBank numbers: LSU: MT214440, ITS: MT214346, SSU: MT214394, TEF1: MT235770, RPB2: MT235806 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequences of Chromolaenomyces appendiculatus (MFLUCC 17-1455, ex-holotype) with 83.59% similarity was Roussoella sp. (strain IFM 64548, LC317761). The closest match of the LSU sequence with 99.20% similarity was Cycasicola goaensis (strain MFLU 17-0581, NG_059057). The closest match of the SSU sequence with 92.89% similarity was Parathyridaria percutanea (strain CBS 868.95, NG_062999). The closest match of the TEF1 sequence with 96.86% similarity was Cycasicola goaensis (strain MFLUCC 17-0754, MG829198), while the closest match of the RPB2 sequence with 87.21% 13 Pseudothyridariella Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557357, Facesoffungi number: FoF 07826 Etymology: Referring to its similarity with Thyridariella. Saprobic on dead stems or decaying wood. Sexual morph: Ascomata immersed to erumpent, solitary or scattered, appearing as dark spots, coriaceous, globose or subglobose to obpyriform, yellowish brown to brown. Ostiolar neck protruding. Peridium several layers, light brown or brown to yellowish brown cells of textura angularis. Hamathecium composed of cylindrical to filiform, septate, branching pseudoparaphyses. Asci 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, pedicellate at base, with an ocular chamber at apically rounded apex. Ascospores overlapping, 1–2 seriate, hyaline to greyish brown, 1-septate when immature, hyaline to brown or olivaceous-brown to dark brown at maturity, ellipsoid to broadly fusiform, muriform, with (3–)5–8-transverse septa, and 1 vertical septum, slightly constricted at the central septum, straight or slightly curved, surrounded by hyaline gelatinous sheath, with central septum observed clearly when mounted in Indian ink. Asexual morph: Undetermined. Type species: Pseudothyridariella chromolaenae Mapook & K.D. Hyde Notes: A phylogenetic analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data shows that Pseudothyridariella chromolaenae forms a sister clade with a clade comprising Liua, Cycasicola, Thyridariella and Chromolaenomyces in Thyridariaceae, with bootstrap support (100% ML and 1.00 BYPP, Fig. 84). Pseudothyridariella is similar to Thyridariella in having ellipsoid to broadly fusiform muriform ascospores surrounded by a hyaline gelatinous sheath, but differs in having ascospores with constriction at the central septum, while ascospores of Thyridariella are without constriction at the central septum (Devadatha et al. 2018). Therefore, we introduce Pseudothyridariella as a new genus with a new species, Pseudothyridariella chromolaenae based on morphology and phylogeny. Pseudothyridariella chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557358, Facesoffungi number: FoF 07827; Fig. 87 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0361 Fungal Diversity Fig. 82 Xenoroussoella triseptata (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores. Scale bars: a = 200 µm, b = 100 µm, c = 50 µm, d, g–j = 20 µm, e = 10 µm, f, k–p = 5 µm Saprobic on dead stems of Chromolaena odorata. Sexual morph:Ascomata 170–345 µm high × 95–220 µm diam. ( x̄ = 225 × 150 µm, n = 5), immersed, solitary or scattered, appearing as dark spots, coriaceous, obpyriform, yellowish brown to brown. Ostiolar neck protruding. Peridium (10–)15–20 µm wide, several layers, light brown to yellowish brown cells of textura angularis. Hamathecium comprising 1.5–2.5 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses. Asci 80–105(–135) × (14–)17–22 µm ( x̄ = 100 × 18.5 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, apically rounded, pedicellate. Ascospores 23–28 × 9–12.5 µm ( x̄ = 25 × 10.5 µm, n = 25), overlapping, 1–2 seriate, initially hyaline to greyish brown, 1-septate when immature, becoming brown or olivaceous-brown to dark brown at maturity, ellipsoid to broadly fusiform, muriform, with 5–8-transverse septa, and 1 vertical septum, slightly constricted at the central septum, straight or slightly curved, surrounded by hyaline gelatinous sheath observed clearly when mounted in Indian ink. Asexual morph: Undetermined. 13 Fungal Diversity closely related to Thyridariella mahakashae NFCCl 4215, which was treated as Pseudothyridariella mahakashae in this study (Fig. 84). However, P. chromolaenae differs from P. mahakashae in having smaller ascomata (170–345 × 95–220 µm vs. 250–550 × 195–500 µm) and asci (80–105(–135) × (14–)17–22 µm vs. 70–220 × 10–20 µm) with brown or olivaceous-brown to dark brown ascospores at maturity with 5–8-transverse septa, while P. mahakashae has hyaline ascospores with 3–6-transverse septa (Table 18). A comparison of the ITS (+ 5.8S) gene region of P. chromolaenae and P. mahakashaereveals 82 base pair differences (17%) across 482 nucleotides. Therefore, P. chromolaenae is described as a new species based on phylogeny and morphological comparison. Fig. 83 Culture characteristic on MEA: Xenoroussoella triseptata (MFLUCC 17-1438) Culture characteristics: Ascospores germinating on MEA within 48 h. at room temperature and germ tubes produced from several cells. Colonies on MEA circular, mycelium slightly raised, velvety with moderately fluffy, entire, cultures white at the surface, creamy-white in reverse with white at the margin (Fig. 88). Pre-screening for antimicrobial activity: Pseudothyridariella chromolaenae (MFLUCC 17-1472) showed antimicrobial activity against M. plumbeus with a 23 mm inhibition zone, observable as partial inhibition, when compared to the positive control (17 mm), but no inhibition of B. subtilis and E. coli. Material examined: THAILAND, Nan Province, Doi Phu Kha, on dead stems of Chromolaena odorata, 23 September 2016, A. Mapook (DPK1, MFLU 20-0361, holotype); ex-type living culture MFLUCC 17-1472. GenBank numbers: LSU: MT214441, ITS: MT214347, SSU: MT214395, TEF1: MT235771, RPB2: MT235807 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequences of Pseudothyridariella chromolaenae (MFLUCC 17-1472, ex-holotype) with 93.82% similarity was Thyridariella mahakashae (strain NFCCl 4215, MG020436). The closest match of the LSU sequence with 97.59% similarity was Roussoellaceae sp. (strain MUT 4884, KP671726). The closest match of SSU sequence with 95.95% similarity was Parathyridaria percutanea (strain CBS 868.95, NG_062999). The closest match of the TEF1 sequence with 96.58% similarity was Parathyridaria percutanea (strain UTHSC: DI16300, LT797113), while the closest match of the RPB2 sequence with 87.57% similarity was T. mangrovei (strain NFCCI 4213, MG020445). In the present phylogenetic analysis, P. chromolaenae (MFLUCC 17-1472) is found 13 Pseudothyridariella mahakashae (Devadatha, V.V. Sarma, D.N. Wanas., K.D. Hyde & E.B.G. Jones) Mapook & K.D. Hyde, comb. nov. Index Fungorum number: IF557369, Facesoffungi number: FoF 07828 ≡ Thyridariella mahakoshae Devadatha, V.V. Sarma, D.N. Wanas., K.D. Hyde & E.B.G. Jones, in Devadatha, Sarma, Jeewon, Wanasinghe, Hyde & Jones, Mycol. Progr.: https://doi.org/10.1007/s11557-018-1387-4, 8 (2018) Holotype: India, Tamil Nadu, Tiruvarur, Muthupet mangroves (10.4° N 79.5° E), on decaying wood of Avicennia marina (Acanthaceae), 29 March 2017, B. Devadatha (AMH-9933, holotype), ex-type culture, NFCCI-4215. Morphological description: See Devadatha et al. (2018) (Fig. 2). Notes: In our multigene phylogenetic study, Thyridariella mahakoshae (NFCCl 4215) was in a clade separate from the type species, Thyridariella mangrovei (NFCCI-4213) and clustered with Pseudothyridariella chromolaenae (MFLUCC 17-1472) with high bootstrap support (100% ML and 1.00 BYPP, Fig. 84). The species shares similar morphological characters with P. chromolaenae in having ellipsoid to broadly fusiform muriform ascospore, surrounded by a hyaline gelatinous sheath with constriction at the central septum, while T. mangrovei has ascospores surrounded by a wide gelatinous sheath in circle but not constricted at the central septum (Devadatha et al. 2018). Therefore, we transfer Thyridariella mahakoshae as Pseudothyridariella mahakashae based on phylogeny and morphological comparison. Torulaceae Corda Torulaceae was introduced by Corda (Sturm 1829) with Torula as the type genus. Crous et al. (2015a) provided molecular data for Torula species and also accepted the genus Dendryphion in the family. Su et al. (2016) introduced a new genus Neotorula from freshwater habitats and Li et al. (2016) introduced a new genus Sporidesmioides within the Fungal Diversity Fig. 84 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, TEF1, RPB2 and SSU sequence data. Eightyfour strains are included in the combined sequence analysis, which comprise 4416 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 35074.673502 is presented. The matrix had 1690 distinct alignment patterns, with 39.45% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.246392, C = 0.254324, G = 0.268438, T = 0.230846; substitution rates: AC = 1.657341, AG = 4.989387, AT = 2.205689, CG = 1.285998, CT = 10.615734, GT = 1.000000; gamma distribution shape parameter α = 0.486933. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given above or below the nodes. Newly generated sequences and new combination are in dark red bold and type species are in bold. Occultibambusa bambusae (MFLUCC 11-0394) and O. bambusae (MFLUCC 13-0855) are used as outgroup taxa family based on both morphology and phylogenetic analyses. Su et al. (2018) introduced a new genus Rostriconidium from freshwater habitats, and provided a morphological comparison and mutigene analyses. Currently, five genera, Torula, Dendryphion, Neotorula, Sporidesmioides and Rostriconidium are accepted in Torulaceae within the order Pleosporales (Li et al. 2016a, b, 2017; Su et al. 2016, 2018; Tibpromma et al. 2018). Divergence time estimates for this family are crown age of 15 Mya (4–34) in the Cenozoic Era 13 Fungal Diversity Fig. 85 Chromolaenomyces appendiculatus (holotype) a, b Appearance of immersed ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–p Ascospores. Scale bars: a, b = 100 µm, c, g–j = 50 µm, d, e = 20 µm, k–p = 10 µm, f = 5 µm (Neogene period) and stem age of 140 Mya (95–188) in the early Cretaceous (Liu et al. 2017). Torula Pers. Torula was introduced by Persoon (1794) with T. herbarum as the type species. Crous et al. (2015a) introduced three new species T. fici, T. hollandica and T. masonii based 13 morphology and phylogeny. Subsequently, Li et al. (2017) introduced four new species (T. chromolaenae, T. mackenziei, T. pluriseptata, T. chiangmaiensis) with two new host records for T. masonii and T. fici, based on DNA sequence analyses from nucleotides and protein genes; including two species T. fici and T. chromolaenae which were collected from Chromolaena odorata in Thailand. Hyde et al. (2019a) Fungal Diversity Fig. 86 Culture characteristic on MEA: Chromolaenomyces appendiculatus (MFLUCC 17-1455) introduced two new species T. polyseptata and T. breviconidiophora based on morphology and phylogeny. We present two reference specimens and a new species from C. odorata, together with descriptions and illustrations (Figs. 90, 91, 92). A phylogenetic tree based on combined LSU, SSU, TEF1, RPB2 and ITS sequence data is presented in Fig. 89. Torula chromolaenae Li, Phook., Mapook & K.D. Hyde, Mycol. Progr. 16(4): 454 (2017) Facesoffungi number: FoF 02713; Fig. 90 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Hyphomycetous. Colonies effuse on host, black, powdery. Mycelium partly immersed to superficial on the host surface, comprising septate, smooth, and pale brown to light brown hyphae. Conidiophores 4–4.7 × 4–5 µm ( x̄ = 4.5 × 4.6 µm, n = 5), macronematous, mononematous, solitary, arising from hypha, erect, pale brown to brown, smooth, thick-walled, subglobose, comprising 1-cell or reduced to conidiogenous cell. Conidiogenous cells 3–6 × 3.5–6 µm ( x̄ = 5 × 5 µm, n = 15), monoblastic, doliiform to subglobose, brown to dark brown, paler at apex, smooth to minutely verruculose, thickwalled. Conidia 7–18.5 × 4–8 µm ( x̄ = 13.5 × 6 µm, n = 30), phragmosporous, catenate, acrogenous, branched, oblong, elongated, 1–3-septate, slightly constricted with dark bands at the septa, dark brown to black, smooth to minutely verrucose, rounded at both ends and mostly with a dark coronate conidiogenous cell at the apex. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, entire, cultures greyish brown at the surface with white at the margin and pale pinkish brown from the centre of the colony in reverse with creamy-white at the margin (Fig. 93a). Pre-screening for antimicrobial activity: Torula chromolaenae (MFLUCC 17-1514) showed antimicrobial activity against B. subtilis, E. coli and M. plumbeus (9 mm, 8 mm and 18 mm inhibition zone, respectively), observable as partial inhibition, when compared to the positive control (26 mm, 9 mm, and 17 mm, respectively). Known hosts and distribution: On dead stems of Chromolaena odorata in Thailand (Li et al. 2017; this study); on dead leaf of Pandanus tectorius (Pandanaceae) in China (Tibpromma et al. 2018). Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP105, MFLU 20-0370); living culture MFLUCC 17-1514 (reference specimen); Chiang Mai Province, Fah hom pok, on dead stems of Chromolaena odorata, 27 September 2016, A. Mapook (FHP8, MFLU 20-0371); living culture MFLUCC 17-1504. GenBank numbers: LSU: MT214477, MT214478, ITS: MT214383, MT214384, SSU: MT214428, MT214429, TEF1: MT235792, MT235793, RPB2: MT235831, MT235832 Notes: A phylogenetic analysis shows that two strains MFLUCC 17-1514 and MFLUCC 17-1504 are grouped with Torula chromolaenae (Fig. 89). In a BLASTn search of NCBI GenBank, the closest match of the ITS, LSU, TEF1 and RPB2 sequences with 99.79% (MH275087), 100% (KY197860), 99.53% (KY197880) and 99.88% (KY197873) similarity, respectively, were identical to T. chromolaenae, while the closest match of the SSU sequence with 99.81% similarity was T. hollandica (strain CBS 220.69, KF443389). We therefore, identify our isolates as T. chromolaenae based on phylogenetic analyses and morphological comparison (Table 19). Moreover, the type and our strains are collected from the same host, Chromolaena odorata and from the same country as that of Li et al. (2017). Thus, we designate our strain as a reference specimen (sensu Ariyawansa et al. 2014a) for Torula chromolaenae. Torula fici Crous, IMA Fungus 6 (1): 192 (2015) Facesoffungi number: FoF 02712; Fig. 91 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Hyphomycetous. Colonies effuse on host, dark brown to black, powdery. Mycelium partly immersed to superficial on the host surface, comprising septate, smooth, and pale brown to light brown hyphae. Conidiophores 2.5–6.5 × 3.9–5.5 µm (x̅ = 5 × 4.5 µm, n = 15), macronematous, mononematous, solitary, arising from hypha, erect, pale brown to brown, comprising 1–2 cells, smooth, thick-walled, subglobose to oblong. Conidiogenous cells 4–7 × 5.5–7 µm ( x̄ = 6 × 6.25 µm, n = 15), mono- to polyblastic, doliiform to subglobose, dark brown to black, paler at apex, smooth, thick-walled. Conidia 10–25 × 5.5–8 µm ( x̄ = 18.5 × 7 µm, n = 20), 13 Fungal Diversity Fig. 87 Pseudothyridariella chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–j Asci. k–o Ascospores. p Ascospores with gelatinous sheath in Indian ink. Scale bars: a–c = 100 µm, g–j = 50 µm, d, e, p = 20 µm, f, k–o = 10 µm phragmosporous, catenate, acrogenous, branched, oblong, elongated, 1–5-septate, slightly constricted with dark bands at the septa, brown to dark brown, verrucose, rounded and mostly dark brown at the apex, straight or slightly curved. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced 13 from both ends. Colonies on MEA circular, mycelium slightly raised, entire, cultures white at the surface and olivaceous-brown from the centre of the colony in reverse with creamy-white at the margin (Fig. 93b). Fungal Diversity Fig. 88 Culture characteristic on MEA: Pseudothyridariella chromolaenae (MFLUCC 17-1472) Pre-screening for antimicrobial activity: Torula fici (MFLUCC 17-1494) showed antimicrobial activity against M. plumbeus with a 13 mm inhibition zone, observable as partial inhibition, when compared to the positive control (18 mm), but no inhibition of B. subtilis and E. coli. Known hosts and distribution: On dead stems of Chromolaena odorata in Thailand (Li et al. 2017; this study); on decaying cone of Magnolia grandiflora (Magnoliaceae) in China; on decaying fruit pericarp of Garcinia sp. (Clusiaceae) in Thailand (Jayasiri et al. 2019); on dead leaf of Pandanus sp. (Pandanaceae) in Thailand (Tibpromma et al. 2018); on submerged decaying wood in China (Su et al. 2018); on Ficus religiosa (Moraceae) in Cuba (Crous et al. 2015). Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP93, MFLU 20-0372); living culture MFLUCC 17-1494 (reference specimen). GenBank numbers: LSU: MT214479, ITS: MT214385, SSU: MT214430, TEF1: MT235794, RPB2: MT235833 Notes: Phylogenetic analyses show that the strain MFLUCC 17-1494 grouped within Torula fici clade (Fig. 89). In a BLASTn search of NCBI GenBank, the closest match of the ITS, LSU, SSU, TEF1 and RPB2 sequences with 99.80% (KF443409), 100% (MK348016), 99.81% (KF443388), 98.98% (MK360090) and 98.43% (MK434871) similarity, respectively was Torula fici. We, therefore, identify our isolates as T. fici based on phylogenetic analyses and morphological comparison (Table 19). Moreover, T. fici has been collected from the same host, Chromolaena odorata and from the same country as in study of Li et al. (2017). Thus, we designate our strain as a reference specimen (sensu Ariyawansa et al. 2014a) for Torula fici. Torula polyseptata C.G. Lin & K.D. Hyde, in Hyde et al., Fungal Divers. 96: 71 (2019) Facesoffungi number: FoF 05070; Fig. 92 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Hyphomycetous. Colonies effuse on host, dark brown to black, powdery. Mycelium partly immersed to superficial on the host surface, comprising septate, smooth, branched and pale brown to light brown hyphae. Conidiophores (4.5–)10–27 × 3.5–4.5 µm (x̅ = 15 × 4 µm, n = 10), micro- to macronematous, mononematous, solitary, arising from hypha, erect, pale brown to brown, comprising 1–3 cells, long, thick-walled, cylindrical to ampulliform. Conidiogenous cells 5.5–7.7 × 4.5–8 µm ( x̄ = 6 × 6.5 µm, n = 10), mono- to polyblastic, doliiform to subglobose, brown to dark brown, thick-walled. Conidia 15–55 × 6.5–9 µm ( x̄ = 31 × 8 µm, n = 15), phragmosporous, catenate, acrogenous, unbranched, oblong, elongated, 1–10-septate, constricted with dark bands at the septa, pale brown to yellowish brown, hyaline to light brown at apex when immature, becoming dark brown when mature, verrucose, rounded at both ends?, straight or slightly curved. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly flattened, curled, cultures greyish brown at the surface with white at the margin and pinkish brown from the centre of the colony in reverse with creamy-white at the margin, MEA change to yellow (Fig. 93c). Pre-screening for antimicrobial activity: Torula polyseptata (MFLUCC 17-1495) showed antimicrobial activity against M. plumbeus with a 12 mm inhibition zone, Table 18 Synopsis of Pseudothyridariella species with similar morphological features discussed in this study Species Ascomata (μm) Peridium (µm) Asci (µm) Ascospores (µm) References P. chromolaenae (MFLUCC 17-1472) 170–345 high × 95–220 diam. (10–)15–20 80–105(–135) × (14–)17– 22 P. mahakashae (= Thyridariella mahakashae NFCCl 4215) 250–550 high × 195–500 diam. 15–30 70–220 × 10–20 23–28 × 9–12.5, hyaline to This study greyish brown, 1-septate when immature, becoming brown or olivaceousbrown to dark brown at maturity, 5–8-transverse septa 17–27 × 5–12, hyaline, 3-6 Devadatha transverse septa et al. (2018) 13 Fungal Diversity Fig. 89 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, SSU, TEF1, RPB2 and ITS sequence data. Fourtysix strains are included in the combined sequence analysis, which comprise 4655 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 16753.701279 is presented. The matrix had 1163 distinct alignment patterns, with 37.38% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.245595, C = 0.259181, G = 0.270709, T = 0.224516; substitution rates: AC = 1.706173, AG = 2.694498, AT = 1.288686, CG = 1.108377, CT = 7.475019, GT = 1.000000; gamma distribution shape parameter α = 0.172063. 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 dark red bold and type species are in bold. Occultibambusa bambusae (MFLUCC 11-0394) and O. bambusae (MFLUCC 13-0855) are used as outgroup taxa observable as partial inhibition, when compared to the positive control (18 mm), but no inhibition of B. subtilis and E. coli. Known hosts and distribution: On submerged decaying wood in Thailand (Hyde et al. 2019a). Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP94, MFLU 20-0373); living culture MFLUCC 17-1495 (new host record). GenBank numbers: LSU: MT214476, ITS: MT214382, SSU: MT214427, TEF1: MT235791, RPB2: MT235830 13 Notes: Phylogenetic analyses show strain MFLUCC 17-1495 grouped with Torula polyseptata (Fig. 89). In a BLASTn search of NCBI GenBank, the closest match of the ITS sequences with 98.16% similarity was T. herbarum (strain MEF021, KT315411). The closest match of the LSU sequence with 100% similarity was T. chiangmaiensis (strain KUMCC 16-0039, KY197856). The closest match of the SSU and RPB2 sequences with 99.81% (KF443389) and 96.02% (KF443393) similarity, respectively, was T. hollandica strain CBS 220.69, while the closest match of the TEF1 sequence with 98.38% similarity was T. pluriseptata (strain MFLUCC 14-0437, KY197875). We therefore, identify our isolate as T. Fungal Diversity Fig. 90 Torula chromolaenae (reference specimen) a, b Appearance of colonies on substrate. c, d Conidiophores and Conidiogenous cells. e–g Conidia in chains. h–m Conidia. Scale bars: a = 500 µm, b = 100 µm, e–m = 10 µm, c, d = 5 µm polyseptata based on morphological comparison and phylogenetic analyses. The isolate is introduced here as a new host record from Chromolaena odorata collected in Thailand. Dothideomycetes orders incertae sedis Botryosphaeriales C.L. Schoch et al. Botryosphaeriales was introduced by Schoch et al. (2006). We follow the latest treatment and updated accounts of Botryosphaeriales by Phillips et al. (2019) with recent relevant literature for updated accounts of each family. Six families, Aplosporellaceae, Botryosphaeriaceae, Melanopsaceae, Phyllostictaceae, Planistromellaceae, and Saccharataceae were accepted in the order, based on morphology and molecular data (Phillips et al. 2019). Aplosporellaceae Slippers et al. Aplosporellaceae was introduced by Slippers et al. (2013) to accommodate the genera Bagnisiella and Aplosporella, with the latter as the type genus and suggested that the divergence time estimates for the order Botryosphaeriales originated in the Cretaceous period around 103 Mya (45–188). Sharma et al. (2017) introduced a new genus Alanomyces, the type of which was collected from soil in India. Subsequently, Liu et al. (2017) reported crown age of the order at 114 Mya (73–166) in the Cretaceous with estimates for stem age at 181 Mya (106–279) in the Jurassic. Phillips et al. (2019) reported crown age for Aplosporellaceae at 40 Mya in the Paleogene period with estimates for stem age at 94 Mya in the Cretaceous. Aplosporella Speg. Aplosporella was introduced by Spegazzini (1880), with A. chlorostroma as the type species. The genus comprises plant pathogens, endophytes and saprobes (Jami et al. 2014; Fan et al. 2015; Ekanayaka et al. 2016; Dou et al. 2017; Du et al. 2017; Zhu et al. 2018). Ekanayaka et al. (2016) introduced a new species, A. thailandica from Thailand. Subsequently, Dou et al. (2017) described a new species A. macropycnidia from Cerasus yedoensis in China and Du et al. (2017) introduced a new species, A. ginkgonis isolated 13 Fungal Diversity Fig. 91 Torula fici (reference specimen) a Appearance of colonies on substrate. b, c Conidiophores and conidiogenous cells. d–f Conidia in chains. g–i Conidia. Scale bars: a = 500 µm, d–f = 20 µm, g–i = 10 µm, b, c = 5 µm from symptomatic branches of Ginkgo biloba in China. Over 300 epithets are listed in Index Fungorum (2020). Only a few species have sequence data available in GenBank (Aplosporella africana, A. artocarpi, A. hesperidica, A. javeedii, A. ginkgonis, A. macropycnidia, A. papillata, A. prunicola, A. thailandica and A. yalgorensis). In this study, a new species, Aplosporella chromolaenae, and a new record A. hesperidica based on morphology and molecular data, together with descriptions and illustrations are introduced (Figs. 94, 95). A phylogenetic tree based on combined LSU, ITS and TEF1 sequence data is presented in Fig. 94. Aplosporella chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557287, Facesoffungi number: FoF 07829; Fig. 95 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0298 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiostromata 360–430 µm long, 685–780 µm wide ( x̄ = 400 × 735 µm, n = 10), superficial, coriaceous, gregarious, dark brown to black. Conidiomata 75–145 μm high × 80–160 µm diam. ( x̄ = 110.5 × 120.5 µm, n = 15), loculate; locules arranged in rows, clustered, gregarious, with 2–3 locules forming groups immersed in conidiostroma, globose to subglobose. Conidiomata wall comprising several layers of thick-walled, 13 dark brown to black cells of textura angularis. Hamathecium comprising 1.5–2.5 µm wide, cylindrical to filiform, septate, branching paraphyses. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 4.5–11 × 2.5–4 µm ( x̄ = 7 × 3.5 µm, n = 5), hyaline, holoblastic, oblong to ampuliform. Conidia(13–)16–20 × 8.5–12 µm ( x̄ = 17.5 × 9.8 µm, n = 25), hyaline when immature, becoming brown to dark brown when mature, aseptate, ellipsoid or oval to reniform, thick-walled, with finely verruculose wall observed clearly in mature conidia. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA, filamentous, mycelium slightly raised, moderately fluffy, white aerial hyphae at the surface, spreading from the center with white in reverse (Fig. 97a). Pre-screening for antimicrobial activity: Aplosporella chromolaenae (MFLUCC 17-1517) showed antimicrobial activity against E. coli with a 9 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 7 April 2017, A. Mapook (DP108, MFLU 20-0298, holotype); extype culture MFLUCC 17-1517. GenBank numbers: LSU: MT214434, ITS: MT214340, SSU: MT214389 Fungal Diversity Fig. 92 Torula polyseptata (new host record) a, b Appearance of colonies on substrate. c– g Conidiophores and conidiogenous cells. h–m Conidia. Scale bars: a = 500 µm, b = 200 µm, d = 20 µm, c, e–m = 10 µm Fig. 93 Culture characteristics on MEA: a Torula chromolaenae (MFLUCC 17-1514). b Torula fici (MFLUCC 17-1494). c Torula polyseptata (MFLUCC 17-1495) Notes: In a BLASTn search of NCBI GenBank, the closest match with the ITS sequence of Aplosporella chromolaenae (MFLUCC 17-1517, ex-holotype) with 100% similarity was Aplosporella artocarpi (strain CPC 22791, NR_154688), while the closest match with the LSU sequence with 99.56% similarity was Aplosporella sp. (strain ICMP 17587, EU931110). In the present phylogenetic analysis, A. chromolaenae clustered with A. artocarpi (CPC 22791), with 13 Fungal Diversity Table 19 Synopsis of Torula species with similar morphological features discussed in this study Species Conidiophores (µm) Conidiogenous cells (µm) Catenate conidia (µm) Septation References T. breviconidiophora (MFLU 18-1693) T. chiangmaiensis (MFLU 16-2815) T. chromolaenae (MFLUCC 17-1514) T. chromolaenae (MFLU 16-2819) T. fici (MFLUCC 17-1494) T. fici (MFLU 16-2817) T. polyseptata (MFLU 18-1694) T. polyseptata (MFLUCC 17-1495) 3.5–28 × 3–8 4–7 × 3–7 8–21 × 3.5–7 1–4-septate Hyde et al. (2019a) 4–12-septate Li et al. (2017) 1–3-septate This study 8–12.6(−16.1) × 4.5–5.2 3.4–6.5 × 4.8–7.6 4–4.7 × 4–5 3–6 × 3.5–6 (5.4–)25.5– 70(−86.5) × 5.6–7.8 7–18.5 × 4–8 5–6.3 × 3.5–4.6 (3.5–)4.5–5.1 × 4.8–5.6 12.1–16.5 × (3.6–)4.1–5 2–3-septate Li et al. (2017) 2.5–6.5 × 3.9–5.5 4–7 × 5.5–7 10–25 × 5.5–8 1–5-septate This study 9.4–12.5 × 3.7–4.5 10–40 × 3.5–8 5.4–8.5 × 5–7.4 4.5–8.5 × 4.5–8 12–20 × 4.6–6.6 10–40 × 3.5–7.5 2–4-septate 2–8-septate Li et al. (2017) Hyde et al. (2019a) (4.5–)10–27 × 3.5–4.5 5.5–7.7 × 4.5–8 15–55 × 6.5–9 1–10-septate This study bootstrap support (98% ML, Fig. 94). However, A. chromolaenae differs from A. artocarpi in having smaller conidiomata in conidiostromata (75–145 × 80–160 µm vs. (350– )540–550(–650) × (490–)540–600(–700) µm) and larger conidiogenous cells (3–5 × 2–4 µm vs. 4.5–11 × 2.5–4 µm) (Table 20). Therefore, A. chromolaenae is described as a new species based on phylogeny and morphological comparison. Aplosporella hesperidica Speg., Anal. Soc. cient. argent. 13(1): 18 (1882) Facesoffungi number: FoF 07830; Fig. 96 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata 365–440 µm high × 480–565 µm diam. ( x̄ = 398 × 516 µm, n = 5), semi-immersed to superficial, erumpent, uniloculate, coriaceous, solitary or gregarious, globose, dark-brown to black. Peridium 30–45 µm wide at side, comprising several layers of thick-walled, brown to dark brown cells of textura globulosa. Conidiophores reduced to conidiogenous cells. Conidiogenous cells hyaline, holoblastic, oblong or cylindrical to ampuliform. Conidia16–25(–30) × 9–15 µm ( x̄ = 19 × 11 µm, n = 25), brown to dark brown, aseptate, ellipsoid to oval, thick-walled, with finely verruculose wall observed clearly in mature conidia. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA, filamentous, mycelium slightly raised, velvety with fluffy, white aerial hyphae at the surface, spreading from the center with white in reverse (Fig. 97b). Pre-screening for antimicrobial activity: Aplosporella hesperidica (MFLUCC 17-1518) showed antimicrobial activity against E. coli with a 9 mm inhibition zone when 13 compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Known hosts and distribution: Citrus aurantium (Rutaceae) in India (Rao 1969); early stem-end rot of Citrus sinensis (Rutaceae) in Zimbabwe (Yang et al. 2017). Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 7 April 2017, A. Mapook (DP110, MFLU 20-0299); living culture MFLUCC 17-1518 (new host record). GenBank numbers: LSU: MT214435, ITS: MT214341, SSU: MT214390 Notes: A phylogenetic analysis shows strain MFLUCC 17-1518 grouped with Aplosporella hesperidica (Fig. 94). In a BLASTn search of NCBI GenBank, the closest match of the ITS and the LSU sequences of MFLUCC 17-1518 with 100% (JX681069) and 100% (KX464239) similarity, respectively, was A. hesperidica. We therefore, identify our isolates as A. hesperidica based on phylogenetic analyses and the isolates are introduced here as a new host record from Chromolaena odorata, collected in Thailand. Botryosphaeriaceae Theiss. & P. Syd., Annales Mycologici 16 (1–2): 16 (1918) Botryosphaeriaceae contains numerous plant pathogenic, saprobic and endophytic species associated with a wide range of hosts as well as opportunistic human pathogen (de Hoog et al. 2000; Slippers and Wingfield 2007; Phillips et al. 2013; Mehl et al. 2014; Trakunyingcharoen et al. 2014a; Doilom et al. 2015). The family was introduced by Theissen and Sydow (1918) with Botryosphaeria as the type genus. Divergence time estimates for this family is diverse around 94 Mya in the Cretaceous with crown age of 61 Mya in the Paleogene period (Phillips et al. 2019). Fungal Diversity Fig. 94 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS and TEF1 sequence data. Twenty-five strains are included in the combined sequence analysis, which comprise 1668 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 4044.627020 is presented. The matrix had 270 distinct alignment patterns, with 27.23% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.224904, C = 0.251747, G = 0.277253, T = 0.246096; substitution rates: AC = 2.555314, AG = 3.360132, AT = 1.917272, CG = 2.128299, CT = 6.642079, GT = 1.000000; gamma distribution shape parameter α = 0.149427. 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 dark red bold and type species are in bold. Saccharata proteae (CBS 115206) is used as outgroup taxon Dothiorella Sacc. Dothiorella was introduced by Saccardo (1880) with D. pyrenophora as the type species. The genus comprises plant pathogens, endophytes and saprobes of a wide range of hosts (Crous et al. 2006; Liu et al. 2012; Hyde et al. 2013; Phillips et al. 2013; Dissanayake et al. 2016; Váczy et al. 2018).Jayasiri et al. (2019) introduced a new species, D. lampangensis from unidentified decaying fruit in Thailand. Phookamsak et al. (2019) introduced a new species, D. acericola from dried twigs of Acer palmatum in China with record of D. sarmentorum from dead twigs of Platycladus orientalis in Russia based on both morphology and phylogenetic analyses. We present a new host record for D. dulcispinae isolated from C. odorata, together with a description and illustrations (Fig. 99). A phylogenetic tree based on combined ITS and TEF1 sequence data is presented in Fig. 98. Dothiorella oblonga F.J.J. Van der Walt, Slippers & G.J. Marais, in Slippers et al., Persoonia 33: 163 (2014) Facesoffungi number: FoF 07831; Fig. 99 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Conidiomata (110)180–280 µm high × 90–145 µm diam. ( x̄ = 189 × 115 µm, n = 5), semi-immersed, globose to obpyriform, coriaceous, solitary or scattered, appearing as dark spots. Ostiole papillate. Pycnidial wall 10–25 µm wide, 3–5 layers, brown to dark brown cells of textura angularis. Conidiophores reduced to conidiogenous cells. Conidiogenous cells 3.5–8 × 2.5–4.5(–6) µm ( x̄ = 6 × 3.5 µm, n = 20), hyaline, holoblastic, cylindrical. Conidia 17–25 × 8–12.5 µm ( x̄ = 22 × 10.5 µm, n = 30), hyaline to pale brown when immature, becoming drown to dark brown at maturity, ovoid or oblong to ellipsoidal, aseptate or uniseptate, constricted at the septum, moderately thick-walled with granular appearance, straight or slightly curved. Culture characteristics: Conidia germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium 13 Fungal Diversity Fig. 95 Aplosporella chromolaenae (holotype) a, b Appearance of stromata on substrate. c Section through stroma. d Peridium. e Hamathecium. f–h Conidiogenous cells and developing conidia. i–n Conidia. Scale bars: a, b = 500 µm, c = 100 µm, d = 50 µm, e–n = 10 µm raised, velvety with moderately fluffy, filiform, white at first, become olivaceous-grey with age (Fig. 100). Pre-screening for antimicrobial activity: Dothiorella oblonga (MFLUCC 17-1498) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Known hosts and distribution: Acacia mellifera (Fabaceae) in South Africa (Slippers et al. 2014) Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP97, MFLU 20-0310); living culture MFLUCC 17-1498 (new host record). GenBank numbers: LSU: MT214459, ITS: MT214365, SSU: MT214411 Notes: A phylogenetic analysis (Fig. 98) shows that MFLUCC 17-1498 grouped with Dothiorella oblonga, with bootstrap support (84% ML). In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence for MFLUCC 17-1498 is D. oblonga with 99.82% similarity to 13 the strain CBS 121765 (KF766163). We therefore, identify our isolates as D. oblonga based on phylogenetic analyses and morphological comparison (Table 21) and the isolates are introduced here as a new host record from Chromolaena odorata in Thailand. Sphaeropsis Sacc. Sphaeropsis was introduced by Saccardo (1880) with S. visci as the type species. More than 600 species epithets are listed in Index Fungorum (2020). Phookamsak et al. (2019) introduced a new host record of S. eucalypticola from Bauhinia purpurea in Thailand based on both morphology and phylogenetic analyses. In this study, a new Sphaeropsis species is introduced, based on morphology and molecular data, together with descriptions and illustrations (Fig. 102). A phylogenetic tree based on combined ITS, LSU, SSU, TEF1 and TUB2 sequence data is presented in Fig. 101. Fungal Diversity Fig. 96 Aplosporella hesperidica (new host record) a Appearance of conidiomata on substrate. b Section through conidioma. c Peridium. d–g Conidiogenous cells and developing conidia. h, i Conidia. Scale bars: a = 500 µm, b, c = 100 µm, g–i = 20 µm, d–f = 10 µm Fig. 97 Culture characteristics on MEA: a Aplosporella chromolaenae (MFLUCC 17-1517). b Aplosporella hesperidica (MFLUCC 17-1518) Sphaeropsis chromolaenicola Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557363, Facesoffungi number: FoF 07832; Fig. 102 Etymology: Name reflects the host genus Chromolaena, on which this species was growing. Holotype: MFLU 20-0369 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 200–235 µm high × 260–285 µm diam. ( x̄ = 220 × 275 µm, n = 5), immersed to erumpent through host tissue, globose, coriaceous, solitary or scattered, appearing as brown to dark brown spots. Ostiole central, short papillate. Peridium 20–50(–65)µm wide, 3–5 layers, reddish brown to dark brown cells of textura angularis. Hamathecium comprising 2.5–4 µm wide, oblong to cylindrical, septate pseudoparaphyses. Asci 90–130 × 34–45 µm ( x̄ = 112 × 37 µm, n = 10), 8-spored, bitunicate, fissitunicate, cylindric-clavate to clavate, with a short pedicel, straight to slightly curved, apically rounded with an ocular chamber. Ascospores 28.5–34 × 12–15 µm µm ( x̄ = 30 × 13.5 µm, n = 25), 2–3-seriate, overlapping in the ascus, hyaline to yellowish brown when immature, becoming brown to dark brown at maturity, ellipsoid to ovoid, aseptate, with granular appearance, widest at the center and tapering toward narrow ends, straight to slightly curved. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, mycelium velvety, moderately fluffy, filiform, smoke-grey to dark olivaceous at the surface and dark in reverse (Fig. 103). 13 Fungal Diversity Fig. 98 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS and TEF1 sequence data. Seventy-two strains are included in the combined sequence analysis, which comprise 888 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 4583.919557 is presented. The matrix had 382 distinct alignment patterns, with 14.43% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.212092, C = 0.286537, G = 0.252906, T = 0.248465; substitution rates: AC = 1.390409, AG = 2.605017, AT = 1.096995, CG = 1.336154, CT = 5.435452, GT = 1.000000; gamma distribution shape parameter α = 0.202628. 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 dark red bold and type species are in bold. Neofusicoccum luteum (CBS 110299) and N. luteum (CBS 110497) are used as outgroup taxa Pre-screening for antimicrobial activity: Sphaeropsis chromolaenicola (MFLUCC 17-1499) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP98, MFLU 20-0369, holotype); ex-type culture MFLUCC 17-1499. 13 GenBank numbers: LSU: MT214460, ITS: MT214366, SSU: MT214412 Notes: Multigene phylogenetic analyses (Fig. 101) show that our strain Sphaeropsis chromolaenicola (MFLUCC 17-1499, ex-holotype) clusters with S. eucalypticola clade, with high bootstrap support (92% ML and 0.99 BYPP). In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence for MFLUCC 17-1499 is S. eucalypticola with Fungal Diversity Table 20 Synopsis of Aplosporella species with similar morphological features discussed in this study Species A. artocarpi (CPC 22791) Conidiomata (µm) (350–)540–550(– 650) × (490–)540– 600(–700) A. chromolaenae (MFLUCC 17-1517) 75–145 × 80–160 A. hesperidica – A. hesperidica (MFLUCC 17-1518) 365–440 × 480–565 Conidiogenous cells (µm) Conidia (µm) Reference 3–5 × 2–4 (17–)18–21(–22) × (9–)10–11 Trakunyingcharoen et al. (2015) 4.5–11 × 2.5–4 – – (13–)16–20 × 8.5–12 22–25 × 9–11 16–25(–30) × 9–15 This study Spegazzini (1882) This study Fig. 99 Dothiorella oblonga (new host record) a, b Appearance of conidiomata on substrate. c Section through conidioma. d Peridium. e–i Conidiogenous cells and developing conidia. j, k Conidia. l Conidia with gelatinous sheath in Indian ink. Scale bars: a = 500 µm, b = 200 µm, c = 50 µm, d, e–l = 10 µm 100% similarity to the strain CBS 133993 (MH866075), while the closest match with the LSU sequence was with S. porosa (strain CBS 110574, DQ377895) with 99.67% similarity and closest matches with the SSU sequence were S. porosa (strain CBS 110496, NG_062740) and S. visci (strain CBS 100163, EU673177) with 99.82% similarity. Therefore, S. chromolaenicola is described as a new species based on phylogeny with morphological comparison. The new species, S. chromolaenicola is similar to S. eucalypticola in having cylindric-clavate or clavate asci with ellipsoidal to ovoid, aseptate ascospores, with wide center and tapering narrow ends, but S. chromolaenicola has wider asci (90–130 × 34–45 µm vs. 102 − 175 × 22–32 µm) and wider ascomata (200–235 × 260–285 µm), while S. eucalypticola has smaller ascomata in ascostromata (250–350 × 170 − 250 µm) (Table 22). Dyfrolomycetales K.L. Pang et al. Dyfrolomycetales was introduced by Pang et al. (2013) with a monotypic family Pleurotremataceae (= Dyfrolomycetaceae). We follow the latest treatment and updated accounts of Dyfrolomycetales in Pang et al. (2013), Norphanphoun et al. (2017), Zhang et al. (2017) and Hyde et al. (2018). Pleurotremataceae Walt. Watson (= Dyfrolomycetaceae K.D. Hyde et al.)* Pleurotremataceae was introduced by Watson (1929) and accepted with monotypic genus Pleurotrema within Chaetosphaeriales (Maharachchikumbura et al. 2015). The family was excluded from Sordariomycetes based on reexamination of the isotype specimen of Pleurotrema polysemum, which is morphologically similar to Saccardoella and 13 Fungal Diversity Fig. 100 Culture characteristic on MEA: Dothiorella oblonga (MFLUCC 17-1498) Dyfrolomyces in Dothideomycetes (Maharachchikumbura et al. 2016). Dyfrolomycetaceae was introduced by Pang et al. (2013) to accommodate the genus Dyfrolomyces with Fig. 101 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS, LSU, SSU, TEF1 and TUB2 sequence data. Twelve strains are included in the combined sequence analysis, which comprise 3285 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 6043.146013 is presented. The matrix had 256 distinct alignment patterns, with 17.97% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.239114, C = 0.251670, G = 0.275946, T = 0.233269; substitution rates: AC = 1.581937, AG = 2.385055, AT = 0.562202, CG = 1.369785, CT = 5.675544, GT = 1.000000; gamma distribution shape parameter α = 0.020000. 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 dark red bold and type species are in bold. Barriopsis fusca (CBS 174.26) and B. tectonae (MFLUCC 12-0381) are used as outgroup taxa 13 D. tiomanensis as the type species and three Saccardoella species, S. rhizophorae, S. mangrovei and S. marinospora were transferred to Dyfrolomyces. Currently, Dyfrolomycetaceae is a synonymy under Pleurotremataceae (Maharachchikumbura et al. 2016) with three accepted genera, Dyfrolomyces, Melomastia and Pleurotrema (Pang et al. 2013; Norphanphoun et al. 2017; Zhang et al. 2017; Wijayawardene et al. 2018). Divergence time estimate for this family and estimated for crown age are in the early Cenozoic Era at 55 Mya (38–74). The family shares a common ancestry with Palawaniaceae and Muyocopronaceae at 192 Mya (145–243) in the early Jurassic (Mapook et al. 2016c). Lui et al. (2017) estimated in the late Cretaceous at 76 Mya (38–125) for crown age and shares the common ancestry with Acrospermaceae in the Jurassic at 174 Mya (113–243). However, the families Palawaniaceae and Muyocopronaceae were not included in this analysis. Dyfrolomyces K.D. Hyde et al. Fungal Diversity Table 21 Synopsis of Dothiorella species with similar morphological features discussed in this study Species Conidiomata (µm) Conidiogenous cells (µm) Conidia (µm) References D. dulcispinae (PREM 60706) up to 200 wide – (14–)16 − 22(– 24) × (6–)7 − 10(–11) 17–25 × 8 − 12.5 Jami et al. (2012) D. oblonga (MFLUCC 17-1498) (110)180–280 high × 90–145 3.5–8 × 2.5–4.5(–6) diam. D. oblonga (PREM 59628) 840.5 high × up to 550 diam. (6 −)8.5 − 11.5( − 12.5) × (2.5–)4 − 4.5( − 5.5) D. thailandica (MFLUCC 11-0438) 400–800 wide, 200–250 2–5.5 × 1.5 − 4.5 high, 250–500 diam. Dyfrolomyces was introduced by Pang et al. (2013) with the type species, D. tiomanensis, collected from unidentified (18.5–)23.5 − 27(–28) × (10–)11.5–13(− 15) 15 − 20 × 6.5 − 8 This study Slippers et al. (2014) Liu et al. (2012) mangrove wood, and they transferred three Saccardoella species (S. rhizophorae, S. mangrovei and S. marinospora) Fig. 102 Sphaeropsis chromolaenicola (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–i Immature and mature asci. j–o Ascospores (immature and mature). Scale bars: a = 500 µm, b = 200 µm, c, g–i = 50 µm, d, e = 20 µm, f, j–o = 10 µm 13 Fungal Diversity Fig. 103 Culture characteristic on MEA: Sphaeropsis chromolaenicola (MFLUCC 17-1499) to Dyfrolomyces. Zhang et al. (2017) introduced two new species (D. thamplaensis and D. maolanensis) from terrestrial habitats based on morphology together with multigene analyses and transferred Saccardoella aquatica as D. aquatica, based on its similar morphological characters. Hyde et al. (2018) introduced a new species, D. phetchaburiensis from submerged wood of Rhizophora apiculata, based on combined LSU and SSU sequence data with morphological comparison. Eight epithets are listed in Index Fungorum (2020) and most species have been collected from marine habitats on mangrove wood, as well as in terrestrial (Pang et al. 2013; Zhang et al. 2017; Hyde et al. 2018). In this study, a new species Dyfrolomyces chromolaenae, is introduced, together with a description and illustrations, based on morphology and molecular data (Fig. 105). A phylogenetic tree based on combined LSU, SSU and TEF1 sequence data is presented in Fig. 104. Dyfrolomyces chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557290, Facesoffungi number: FoF 07833; Fig. 105 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0311 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 315–380 µm high × 315–400 µm diam. ( x̄ = 340 × 355 µm, n = 5), immersed to erumpent through host tissue, solitary or scattered, coriaceous to carbonaceous, without a subiculum. Clypeus extending outwards, thicker around the papilla. Ostiole central, papillate. Peridium 20–55 µm wide, comprising several layers of pale brown to brown cells of textura angularis. Hamathecium comprising 1.5–3.5 µm wide, cylindrical to broadly filiform, septate, branching pseudoparaphyses. Asci 135–160 × 7–8 µm ( x̄ 13 = 145 × 7.5 µm, n = 10), 8-spored, cylindrical, short pedicellate, straight or slightly curved, apically rounded, with an apical ring. Ascospores 29–35 × 4.5–6 µm ( x̄ = 32 × 5.5 µm, n = 15), uniseriate, hyaline, fusiform, tapering towards narrow ends, 1-septate when immature, becoming 9–11-septate when mature, straight or slightly curved, without terminal appendages. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium slightly raised, entire to slightly filamentous, white aerial hyphae at the surface, becoming greyish brown and dark brown in reverse (Fig. 106). Pre-screening for antimicrobial activity: Dyfrolomyces chromolaenae (MFLUCC 17-1434) showed antimicrobial activity against B. subtilis with a 12 mm inhibition zone, when compared to the positive control (26 mm), but no inhibition of E. coli and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP17, MFLU 20-0311, holotype); extype culture MFLUCC 17-1434. GenBank numbers: SSU: MT214413, TEF1: MT235800 Notes: In a BLASTn search of NCBI GenBank, the closest match of the SSU sequence of Dyfrolomyces chromolaenae (MFLUCC 17-1434, ex-holotype) is D. rhizophorae with 97.18% similarity to the strain JK 5349A (GU479766), while the closest match of the TEF1 sequence with 89.74% similarity was D. tiomanensis (strain NTOU3636 = MFLUCC13-0440, KC692157). In the present phylogenetic analysis, D. chromolaenae is found closely related to D. tiomanensis strain MFLUCC13-0440 (Fig. 104). However, D. chromolaenae differs from D. tiomanensis in having smaller ascomata (315–380 × 315–400 µm vs. 565–(615)–667 × 283–(374)–446 µm), smaller asci (135–160 × 7–8 µm vs. 316–(323)–333 × 12–(16)–17 µm) and smaller ascospores (29–35 × 4.5–6 µm vs. 69–(74)–82 × 9–(10)–11 µm) with 9–11-septa, while D. tiomanensis has 20–24-septate ascospores (Table 23). A comparison of the TEF1 gene region of D. chromolaenae and D. tiomanensis reveals 81 base pair differences (10%) across 801 nucleotides. Therefore, D. chromolaenae is described as a new species based on phylogeny and morphological comparison. Muyocopronales Mapook et al. Muyocopronales was introduced by Mapook et al. (2016b). The order comprises two families, Muyocopronaceae and Palawaniaceae, based on phylogenetic analyses and morphological comparison, together with their divergence times for additional evidence to recognize the status of higher level taxa. See Mapook et al. (2016b, c) for the details. Fungal Diversity Table 22 Synopsis of Sphaeropsis species with similar morphological features discussed in this study Species Ascomata (µm) Asci (µm) S. chromolaenicola (MFLUCC 17-1499) S. eucalypticola (MFLUCC 11-0579 = CBS 133993) S. eucalypticola (MFLUCC 12-0171) 200–235 high × 260–285 diam. 90–130 × 34–45 Ascospores (µm) References 28.5–34 × 12–15 This study – (90–)97 − 110(−125) × 28–30 27–35 × 11–14 Liu et al. (2012) Ascostromata, 250–350 high × 170 − 250 diam. 102 − 175 × 22–32 Phookamsak et al. (2019) 27–33 × 11–14 Muyocopronaceae K.D. Hyde Muyocopronaceae was introduced by Luttrell (1951) and was included in the order Hemisphaeriales. The family was reintroduced by Hyde et al. (2013) as a distinct family (Dothideomycetes family, incertae sedis) with monotypic genus Muyocopron. Mapook et al. (2016b) introduced the new order Muyocopronales to accommodate three new Muyocopron species which were collected in Thailand, based on phylogenetic analyses and morphological comparison. Crous et al. (2018a) introduced a new genus, Neocochlearomyces on leaves of Chromolaena odorata in Thailand and accepted six genera (Arxiella, Leptodiscella, Mycoleptodiscus, Muyocopron, Neocochlearomyces, Paramycoleptodiscus) in the family. Hernández-Restrepo et al. (2019) introduced Neomycoleptodiscus as a new genus similar to Mycoleptodiscus in Muyocopronaceae. Mapook et al. (2020) introduced a new genus Pseudopalawania in the family with their secondary metabolites production and biological activity. Divergence time estimates for this family are crown age of 52 Mya (38–66) in the Cenozoic Era, Paleogene and shares the most common ancestor with Palawaniaceae at 172 Mya (130–218) in Jurassic period (Mapook et al. 2016c). Samarakoon et al. Fig. 104 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, SSU and TEF1 sequence data. Thirteen strains are included in the combined sequence analysis, which comprise 2745 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 7474.968962 is presented. The matrix had 399 distinct alignment patterns, with 27.36% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.243143, C = 0.248312, G = 0.294370, T = 0.214174; substitution rates: AC = 1.048096, AG = 2.972507, AT = 1.232056, CG = 1.713483, CT = 11.686012, GT = 1.000000; gamma distribution shape parameter α = 0.159409. 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 dark red bold and type species are in bold. Acrospermum compressum (M151), A. gramineum (M152) and A. adeanum (M133) are used as outgroup taxa 13 Fungal Diversity Fig. 105 Dyfrolomyces chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Peridium. e Pseudoparaphyses. f–i Asci. j–o Ascospores. Scale bars: a = 500 µm, b = 200 µm, c = 100 µm, f–i = 50 µm, d = 20 µm, e, j–o = 10 µm (2019) recorded fossils justification of Muyocopron with two known fossil species M. mucoris and M. neyveliensis, which belong to the Cenozoic Era (59–5.33 Mya) and suggested 13 the minimum age at 54 Mya instead of 52 Mya for the future calibration point of molecular clock analyses. Fungal Diversity Fig. 106 Culture characteristic on MEA: Dyfrolomyces chromolaenae (MFLUCC 17-1434) Muyocopron Speg., Anales de la Sociedad Cientifica Argentina. 12: 113 (1881) Muyocopron was introduced by Spegazzini (1881b) with Mu. corrientinum as the type species, collected from dead leaves of Oncidium sp. in Argentina. The genus is mostly a saprobe common on the surface of dead aerial twigs, branches, stems or leaves of a wide variety of plants and reported from tropical and temperate regions (Taylor and Hyde 2003; Mapook et al. 2016b). Muyocopron was transferred to Botryosphaeriaceae (von Arx and Müller 1975) as well as Microthyriaceae (Saccardo 1883; Hawksworth et al. 1995; Lumbsch and Huhndorf 2007, 2010a, b; Kirk et al. 2008), based on morphology. Subsequently, the genus was excluded from Microthyriaceae and transferred to Dothideomycetes genera, incertae cedis (Wu et al. 2011a, b). Hyde et al. (2013) reintroduced Muyocorponaceae as a distinct family with two unidentified Muyocopron species based on morphological comparison, supported with molecular studies. Mapook et al. (2016b) introduced the new order Muyocopronales to accommodate three new Muyocopron species from northern Thailand (Mu. castanopsis, Mu. dipterocarpi and Mu. lithocarpi), based on a distinct lineage and morphological comparison with Dyfrolomycetales and Acrospermales. Tibpromma et al. (2016) also introduced a new species, Mu. garethjonesii collected from dead leaves of Pandanus sp. in China. Senwanna et al. (2019) reported a new host record, Mu. dipterocarpi with a new species, Mu. heveae from dried twig of Hevea brasiliensis in Thailand. Phookamsak et al. (2019) reported a new host record of Mu. lithocarpi from dead stems of herbaceous plant in China and Jayasiri et al. (2019) reported new host record of Mu. dipterocarpi from twigs and pod of Delonix regia from Thailand, as well as Mu. lithocarpi on leaves and wild pods. Hernández-Restrepo et al. (2019) introduced two new Muyocopron species associated with leaf spots (Mu. alcornii on Epidendrum sp. in Australia and Mu. zamiae on Zamia in USA), with four new combinations of Mycoleptodiscus laterale, My. coloratum, My. geniculatum and My. atromaculans, to the genus Muyocopron. In this study, Muyocopron chromolaenae is introduced as a new species and a new host record for Mu. lithocarpi from Thailand, together with description and illustrations, based on morphology and molecular data (Figs. 108, 109, 110). A phylogenetic tree based on combined LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 107. Muyocopron chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557309, Facesoffungi number: FoF 07834; Fig. 108 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0328 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 120–150(–160) µm high × (315–)370–425(–480) µm diam. ( x̄ = 135 × 390 µm, n = 10), superficial, coriaceous, solitary or scattered, appearing as circular, scattered, flattened, brown to dark brown spots, covering the host, without a subiculum, with a poorly developed basal layer and an irregular margin. Ostiole central. Peridium 10–40(–50) µm wide, widest at the sides, outer layers comprising dark brown to black pseudoparenchymatous, occluded cells of textura epidermoidea, inner layers comprising light brown to hyaline cells of textura epidermoidea. Hamathecium comprising 0.5–1.5 µm wide, cylindrical to filiform, septate pseudoparaphyses. Asci 40–55 × 10–20 µm ( x̄ = 50 × 16 µm, n = 10), 8-spored, bitunicate, saccate or broadly obpyriform, pedicellate, straight or slightly curved, with an ocular chamber. Ascospores 15–20 × 6–9 µm ( x̄ = 18 × 7.5 µm, n = 20), irregularly arranged, overlapping in the ascus, hyaline, broadly fusiform, aseptate, smooth, widest at the center and tapering towards ends, straight or slightly curved. Asexual morph: Undetermined. Table 23 Synopsis of Dyfrolomyces species with similar morphological features discussed in this study Species Ascomata (µm) Asci (µm) Ascospores (µm) D. chromolaenae (MFLUCC 17-1434) D. tiomanensis (MFLUCC13-0440 = NTOU3636) 315–380 × 315–400 135–160 × 7–8 29–35 × 4.5–6, 9–11-septate 565 − (615) − 667 × 283 316 − (323) − 333 × 1 69 − (74) − 82 × 9 − (10) − 11, − (374) − 446 2 − (16) − 17 20 − 24-septate References This study Pang et al. (2013) 13 Fungal Diversity Fig. 107 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU, TEF1 and RPB2 sequence data. Seventytwo strains are included in the combined sequence analysis, which comprise 5396 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 31179.160692 is presented. The matrix had 2314 distinct alignment patterns, with 61.16% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.237959, C = 0.253544, G = 0.286148, T = 0.222349; substitution rates: AC = 1.253242, AG = 2.577505, AT = 1.560994, CG = 1.058014, CT = 5.657031, GT = 1.000000; gamma distribution shape parameter α = 0.357213. 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 dark red bold and type species are in bold. Lophium mytilinum (AFTOL-ID 1609) and Mytilinidion rhenanum (CBS 135.45) are used as outgroup taxa Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA circular, initially mycelium white, slightly raised, velvety with moderately fluffy, filamentous, cultures grayish to light brown in old, flattened on surface, brown to dark brown in reverse from the centre of the colony, pale brown to white margin (Fig. 111a). 13 Pre-screening for antimicrobial activity: Muyocopron chromolaenae (MFLUCC 17-1513) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Phrae, Doi Pha Klong, on dead stems of Chromolaena odorata, 22 September 2016, A. Mapook (DPKP8, MFLU 20-0328, holotype); ex-type culture MFLUCC 17-1513. Fungal Diversity Fig. 108 Muyocopron chromolaenae (holotype) a, b Superficial ascomata on substrate. c, d Squash mounts showing ascomata walls. e Section of ascoma. f Peridium. g Pseudoparaphyses. h–k. Asci. l–q Unicellular ascospores. Scale bars: a = 500 µm, b = 200 µm, c, e = 50 µm, d, f, h–k = 20 µm, l–q = 10 µm, g = 5 µm GenBank numbers: LSU: MT137876, ITS: MT137777, SSU: MT137881, TEF1: MT136756, RPB2: MT136761 Notes: In a BLASTn search of NCBI GenBank, the closest match of the SSU sequence of Muyocopron chromolaenae (MFLUCC 17-1513, ex-holotype) was Mu. dipterocarpi with 96.69% similarity to the strain MFLUCC 14-1103(KU726969), while the closest match of the LSU sequence with 99.06% similarity was Muyocopron geniculatum (strain CBS 721.95, MH874185). On phylogenetic analysis, Mu. chromolaenae forms a separate branch and clusters with Mu. atromaculans, Mu. geniculatum and Mu. zamiae (Fig. 107). However, we could not compare the morphological characteristics of those species; Mu.chromolaenae is found as sexual morph in nature and we could not obtain its asexual morph in culture, while Mu. atromaculans, Mu. geniculatum and Mu. zamiae were found as asexual morphs in nature. A comparison of the ITS (+5.8S) gene region of Mu. chromolaenae and Mu. geniculatum reveals 37 base pair differences (6.6%) across 561 nucleotides. Therefore, Mu.chromolaenae is described here as a new species based on phylogeny. Muyocopron chromolaenicola Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557310, Facesoffungi number: FoF 07835; Fig. 109 Etymology: Name reflects the host genus Chromolaena, on which this species was growing. Holotype: MFLU 20-0329 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 80–135 µm high × 360–440 µm diam. ( x̄ = 106 × 396 µm, n = 10), superficial, coriaceous, solitary or scattered, appearing as circular, flattened, dark brown to black spots, without a subiculum, with a poorly developed basal layer and an irregular margin. Ostiole central. Peridium 20–35(–50)µm wide, widest at the sides, outer layers 13 Fungal Diversity Fig. 109 Muyocopron chromolaenicola (holotype) a, b Superficial ascomata on substrate. c, d Squash mounts showing ascomata walls. e Section of ascoma. f Peridium. g Pseudoparaphyses. h–k. Asci. l– q Unicellular ascospores. Scale bars: a = 1000 µm, b = 500 µm, c = 100 µm, e = 50 µm, f, h–k = 20 µm, d, l–q = 10 µm, g = 5 µm comprising dark brown to black pseudoparenchymatous, occluded cells of textura epidermoidea, inner layers comprising light brown cells of textura angularis. Hamathecium comprising (1.5–)2–3.5 µm wide, filiform to cylindrical, septate pseudoparaphyses. Asci 50–80 × 20–30 µm ( x̄ = 63 × 24 µm, n = 20), 8-spored, bitunicate, saccate or broadly obpyriform, pedicellate, straight or slightly curved, with a small ocular chamber. Ascospores 14.5–17 × 9–12 µm ( x̄ = 16 × 10.5 µm, n = 20), irregularly arranged, overlapping in the ascus, hyaline, oval to obovoid, with obtuse ends, aseptate, with granular appearance. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, slightly raised, initially mycelium white, filamentous, greyish white in old cultures at the margin, creamy brown in reverse from the centre of the colony, becoming brown to olivaceous brown with white margin (Fig. 111b). 13 Pre-screening for antimicrobial activity: Muyocopron chromolaenicola (MFLUCC 17-1470) showed no inhibition of E. coli, B. subtilis and M. plumbeus. Material examined: THAILAND, Mae Hong Son Province, Mae Yen, on dead stems of C. odorata, 25 June 2016, A. Mapook (MY3, MFLU 20-0329, holotype); ex-type culture MFLUCC 17-1470. GenBank numbers: LSU: MT137877, ITS: MT137778, SSU: MT137882, TEF1: MT136757 Notes: In a BLASTn search of NCBI GenBank, the closest match of the LSU sequence of Muyocopron chromolaenicola (MFLUCC 17-1470, ex-holotype) was Mu. lithocarpi with 99.76% similarity to the strain MFLUCC 16-0962 (MK348034), while the closest match of the SSU sequence with 99.62% similarity was Mu. lithocarpi (strain MFLUCC 14-1106, KU726970). In the present phylogenetic analysis, Mu. chromolaenicola forms a separate branch as a sister taxon to Mu. lithocarpi and Mu. heveae (Fig. 107). However, Mu. chromolaenicola differs from Mu. lithocarpi in having larger ascomata Fungal Diversity Fig. 110 Muyocopron lithocarpi (new host record) a, b Superficial ascomata on substrate. c, d Squash mounts showing ascomata walls. e Section of ascoma. f Peridium. g Pseudoparaphyses. h–k. Asci. l– q Unicellular ascospores. Scale bars: a = 500 µm, b = 200 µm, c = 100 µm, d–f = 50 µm, h–k = 20 µm, g, l–q = 10 µm (80–135 × 360–440 µm vs. 40–70(–120) × 220–320 µm) with longer asci (50–80 × 20–30 µm vs. 45–65 × (15–)23–28 µm), and differs from Mu. heveae in having smaller ascomata (80–135 × 360–440 µm vs. 65–180 × 180–620 µm) and smaller ascospores (14.5–17 × 9–12 µm vs. (19)22–32(37) × (8)10–14(15) µm) with shorter asci (50–80 × 20–30 µm vs. (57)77–116(130) × (19)22–31(36) µm) (Table 24). A comparison of the ITS (+5.8S) gene region of Mu. chromolaenicola and Mu. lithocarpi reveals 37 base pair differences (6.3%) across 584 nucleotides. Therefore, Mu. chromolaenicola is described as a new species based on phylogeny and morphological comparison. Muyocopron lithocarpi Mapook, Boonmee & K.D. Hyde, Phytotaxa 265(3): 235 (2016) Facesoffungi number: FoF 01890; Fig. 110 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 65–90.5 µm high × 225–330 µm diam. ( x̄ = 75 × 290 µm, n = 10), superficial, coriaceous, solitary or scattered with gregarious or confluent, appearing as circular, scattered, flattened, brown to dark brown spots, covering the host, without a subiculum, with a poorly developed basal layer and an irregular margin. Ostiole central. Peridium 15–25(–40) µm wide, widest at the sides, outer layers comprising dark brown to black pseudoparenchymatous, occluded cells of textura epidermoidea, inner layers comprising light brown cells of textura angularis. Hamathecium comprising 1–2 µm wide, cylindrical to filiform, septate pseudoparaphyses. Asci 49–77 × 17–26 µm ( x̄ = 60 × 22.5 µm, n = 25), 8-spored, bitunicate, saccate or broadly obpyriform, pedicellate, straight or slightly curved, with small ocular chamber. Ascospores 14–20 × 9–14 µm ( x̄ = 17 × 11 µm, n = 15), irregularly arranged, overlapping in the ascus, hyaline, oval to obovoid with obtuse ends, aseptate, with granular appearance. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA irregular, slightly raised, initially mycelium white, filamentous, greyish to light brown in old cultures, flattened on surface, brown to dark brown in reverse from the centre of the colony, pale brown to white margin (Fig. 111c). 13 Fungal Diversity Fig. 111 Culture characteristics on MEA: a Muyocopron chromolaenae (MFLUCC 17-1513). b Muyocopron chromolaenicola (MFLUCC 17-1470). c Muyocopron lithocarpi (MFLUCC 17-1500) Pre-screening for antimicrobial activity: In this study, all isolates of Muyocopron lithocarpi showed no inhibition of E. coli, B. subtilis and M. plumbeus. Known hosts and distribution: On dead leaves of Lithocarpus lucidus (Fagaceae) in Thailand (Mapook et al. 2016); on decaying pods of Peltophorum sp. (Fabaceae) in Thailand and on fallen pod of Cercis chinensis (Fabaceae) in China (Jayasiri et al. 2019); on dead stems of herbaceous plant in China (Phookamsak et al. 2019). Material examined: THAILAND, Chiang Rai Province, Doi Mae Salong, on dead stems of Chromolaena odorata, 17 May 2016, A. Mapook (DMS2, MFLU 20-0330); living culture MFLUCC 17-1466; Doi Pui, on dead stems of C. odorata, 2 February 2017, A. Mapook (DP99, MFLU 20-0331); living culture MFLUCC 17-1500 (new host record); Lampang Province, Chae Son, on dead stems of C. odorata, 24 September 2016, A. Mapook (JS1, MFLU 20-0332); living culture MFLUCC 17-1465. GenBank numbers: LSU: MT137878, MT137879, MT137880, ITS: MT137779, MT137780, MT137781, SSU: MT137883, MT137884, MT137885, TEF1: MT136758, MT136759, MT136760, RPB2: MT136762 Notes: Multigene phylogenetic analyses (Fig. 107) show that three strains MFLUCC 17-1466, MFLUCC 17-1500 and MFLUCC 17-1465 are grouped with Muyocopron lithocarpi clade. In a BLASTn search of NCBI GenBank, the closest match of the LSU and SSU sequences showed that the strains are identical to Mu. lithocarpi with 100% similarity. We therefore, identify our three isolates as Mu. lithocarpi based on phylogenetic analyses with morphological comparison (Table 24) and the isolates are introduced here as a new host record from Chromolaena odorata collected in Thailand. Patellariales D. Hawksw. & O.E. Erikss. Patellariales was introduced by Hawksworth and Eriksson (1986) with the monotypic family Patellariaceae. We follow the latest treatment and updated accounts of Patellariales in Hernández-Restrepo et al. (2016) and Pem et al. (2018). Patellariaceae Corda Patellariaceae was introduced by Corda (1838) with Patellaria as the type genus and contains species which are saprobes or weak parasites on wood and bark of various trees and shrubs. The family was classified in the order Patellariales and 21 genera were accepted in the outline classification of Ascomycota (Wijayawardene et al. 2018). Members of Patellariaceae have very few molecular studies. Five genera, Hysteropatella, Yuccamyces, Patellaria, Glyphium and Holmiella have been reported with available sequence in GenBank (Boehm et al. 2009b, 2015; Schoch et al. 2009; Hyde et al. 2013; Yacharoen et al. 2015; Pem et al. 2018). Divergence time estimates for this family are crown age of 164 Mya (72–283) in the Jurassic period and stem age of 311 Mya (244–407) in the Carboniferous period (Liu et al. 2017). Table 24 Synopsis of Muyocopron species with similar morphological features discussed in this study Species Ascomata (µm) Peridium (µm) Asci (µm) Mu. chromolaenicola (MFLUCC 17-1470) Mu. heveae (MFLUCC 17-0066) 80–135 × 360–440 20–35(–50) 65–180 × 180–620 12–41 Mu. lithocarpi (MFLUCC 65–90.5 × 225–330 15–25(–40) 17-1500) Mu. lithocarpi (MFLUCC 40–70(–120) × 220–320 10–20(–28) 14-1106) 13 50–80 × 20–30 Ascospores (µm) 14.5–17 × 9–12, oval to obovoid (19)22–32(37) × (8)10– (57)77– 116(130) × (19)22– 14(15), ellipsoidal or broad oblong to 31(36) obovoid 49–77 × 17–26 14–20 × 9–14, oval to obovoid 45–65 × (15–)23–28 13–18 × 9–11, oval to obovoid References This study Senwanna et al. (2019) This study Mapook et al. (2016b) Fungal Diversity Patellaria Fr., Systema Mycologicum 2: 158 (1822) Patellaria was introduced by Fries (1822) with Patellaria atrata as the type species. The genus has more than 500 species epithets (Index Fungorum 2019). However, only a few species have been explored with molecular data (Schoch et al. 2009; Hernández-Restrepo et al. 2016). HernándezRestrepo et al. (2016) described a new phoma-like fungus, Patellaria quercus based on 99% similarity of the LSU sequence with Patellaria cf. atrata. However, further collections are needed to resolve the asexual morphs linked to Patellariaceae. In this study, Patellaria chromolaenae is introduced as a new species based on morphology and molecular data support with a description and illustrations (Fig. 113). A phylogenetic tree based on LSU, ITS, SSU, TEF1 and RPB2 sequence data is presented in Fig. 112. Patellaria chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557350, Facesoffungi number: FoF 07836; Fig. 113 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0346 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 100–145 µm high × 415–625 µm diam. (x̅ = 120 × 525 µm, n = 5), apothecial, superficial, circular, flattened with a carbonaceous rim, scattered, brown to dark brown. Exciple (15–)20–30(–35) µm wide, outer layers comprising dark brown to black cells and inner layers of hyaline cells arranged in a textura prismatica, continuous to the base (hypothecium). Hamathecium comprising (1.5–)2–2.7 µm wide, cylindrical to filiform, septate, branching pseudoparaphyses, slightly swollen and rounded at the apex, forming magenta to dark purple epithecium Fig. 112 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, SSU, TEF1 and RPB2 sequence data. Twenty-three strains are included in the combined sequence analysis, which comprise 5609 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 19011.857286 is presented. The matrix had 1461 distinct alignment patterns, with 57.16% of undetermined characters or gaps. Estimated base frequen- cies were as follows: A = 0.247075, C = 0.242521, G = 0.279791, T = 0.230613; substitution rates: AC = 1.589017, AG = 2.732102, AT = 1.158430, CG = 1.354331, CT = 6.468048, GT = 1.000000; gamma distribution shape parameter α = 0.258762. 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 dark red bold and type species are in bold. Botryosphaeria dothidea (CBS 115476) is used as outgroup taxon 13 Fungal Diversity Fig. 113 Patellaria chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Exciple. e Pseudoparaphyses mounted in lactoglycerol. f Pseudoparaphyses mounted in water. g, h Asci mounted in lactoglycerol. i, j Asci mounted in water. k-p Ascospores. Scale bars: a, b = 500 µm, c = 100 µm, d, g–j = 20 µm, e, f, k–q = 10 µm above the asci when mounted in lactoglycerol and navy blue epithecium above the asci when mounted in water. Asci 46–60(–70) × 12–16 µm ( x̄ = 54 × 14 µm, n = 15), 8-spored, bitunicate, fissitunicate, cylindric-clavate, straight or slightly curved, with a short, bulbous pedicel, apically rounded. Ascospores 23–32 × 4–5.5 µm ( x̄ = 25.5 × 5 µm, n = 25), overlapping, 2–3-seriate, hyaline, allantoid to broadly fusiform, 3–5-septate, straight or slightly curved, guttulate, without terminal appendages. Asexual morph: Undetermined. 13 Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA irregular, mycelium slightly raised, undulate to lobate, brown with white aerial hyphae spreading from the center of the colony, becoming smoke-grey at the surface and dark yellowish brown to brown in reverse (Fig. 114). Pre-screening for antimicrobial activity: Patellaria chromolaenae (MFLUCC 17-1479) showed antimicrobial activity against E. coli with a 13 mm inhibition zone when Fungal Diversity compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Phrae Province, Mae Kam, on dead stems of Chromolaena odorata, 21 September 2016, A. Mapook (MK1, MFLU 20-0346, holotype); ex-type culture MFLUCC 17-1479; Lampang Province, Ngao, on dead stems of C. odorata, 21 September 2016, A. Mapook (LP4, MFLU 20-0345); living culture MFLUCC 17-1482. GenBank numbers: LSU: MT214474, MT214475, ITS: MT214380, MT214381, SSU: MT214425, MT214426, TEF1: MT235795, MT235796, RPB2: MT235828, MT235829 Notes: In a BLASTn search of NCBI GenBank, the closest match of the LSU sequence of Patellaria chromolaenae (MFLUCC 17-1479, ex-holotype) was Patellaria quercus with 90.42% similarity to the strain CPC 27232 (NG_059696). The closest match of the ITS sequence with 84.07% similarity was P. quercus (strain BHI-F768exna, MF161322). The closest match of the SSU sequence with 99.25% similarity was P. cf. atrata (strain BCC 28877, GU371837). The closest match of the TEF1 and RPB2 sequences with 91.25% (GU349038) and 88.43% (GU371726) similarity, respectively, was Patellaria atrata strain CBS 958.97. In the present phylogenetic analysis, P. chromolaenae forms a sister taxon to P. atrata (Fig. 112). However, P. chromolaenae differs from P. atrata in having smaller ascomata (100–145 × 415–625 µm vs. 675–1160 × 220–300 µm), smaller asci (46–60(–70) × 12–16 µm vs. 98–135 × 15–30 µm) and smaller ascospores (23–32 × 4–5.5 µm vs. 30–45 × 7–10 µm) that are 3–5-septate, while P. atrata has 5–11-septate ascospores (Table 25). A comparison of the ITS (+ 5.8S) gene region of P. chromolaenae and P. atrata reveals 135 base pair differences (21.5%) across 627 nucleotides. Fig. 114 Culture characteristic on MEA: Patellaria chromolaenae (MFLUCC 17-1479) Therefore, P. chromolaenae is described as a new species based on phylogeny and morphological comparison. Class Sordariomycetes O.E. Erikss. & Winka Subclass Diaporthomycetidae Senan. et al. Diaporthales Nannf.* Diaporthales was introduced by Nannfeldt (1932). We follow the latest treatments and updated accounts of Diaporthales in Maharachchikumbura et al. (2016), Senanayake et al. (2017), Hyde et al. (2019a), Le Dinh et al. (2019) and Hyde et al. (2020). Diaporthaceae Höhn. ex Wehm. Diaporthaceae was introduced by Höhnel (1917) with Diaporthe as type genus. The family comprises plant pathogens, endophytes or saprobes on terrestrial plants and rarely on submerged plants (Udayanga et al. 2011; Dai et al. 2014; Maharachchikumbura et al. 2016; Gao et al. 2017; Senanayake et al. 2017; Thambugala and Hyde 2018). Senanayake et al. (2018) and Hyde et al. (2020) accepted 15 genera in the family (Apioporthella, Apiosphaeria, Caudospora, Chaetoconis, Chiangraiomyces, Diaporthe, Hyaliappendispora, Leucodiaporthe, Mazzantia, Ophiodiaporthe, Paradiaporthe, Phaeocytostroma, Phaeodiaporthe, Pustulomyces, Stenocarpella). Divergence time estimates for this family have recently reported for crown age of fossil calibration at 61.15 Mya (29.7–89.9) and secondary calibration at 60.63 Mya (34.8–91.7) in the Paleogene period, and the family shares the most common ancestor with Cytosporaceae at 87.66 Mya for stem age of fossil calibration and 87.20 Mya for stem age of secondary calibration, in the Cretaceous (Guterres et al. 2018). Diaporthe Nitschke. Diaporthe was introduced by Nitschke (1870) with D. eres as the type species. The genus comprises important plant pathogens, endophytes or saprobes from a wide range of hosts (Udayanga et al. 2011; Gao et al. 2017; Wanasinghe et al. 2018; Le Dinh et al. 2019). More than 1000 epithets are listed in Index Fungorum (2020). Wanasinghe et al. (2018) introduced six taxa which comprised four new host records and two new species (D. rosae, D. rosicola) from Rosaceae. Yang et al. (2018a) introduced two new species D. sambucusii and D. schisandrae as Traditional Chinese Medicines in Northeast China. Yang et al. (2018b) introduced twelve new Diaporthe species associated with dieback plant diseases in China. Le Dinh et al. (2019) reported D. asparagi and D. unshiuensis, said to be pathogenic on Asparagus kiusianus in Japan. In this study, a new Diaporthe species is introduced, based on morphology and molecular data, together with descriptions and illustrations (Fig. 116). A phylogenetic tree based on combined ITS, TEF1 and TUB2 sequence data is presented in Fig. 115. Diaporthe chromolaenae Mapook & K.D. Hyde, sp. nov. 13 Fungal Diversity Fig. 115 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS, TEF1 and TUB2 sequence data. Sixtyfour strains are included in the combined sequence analysis, which comprise 2191 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 17612.889312 is presented. The matrix had 1188 distinct alignment patterns, with 31.17% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.222777, C = 0.316430, G = 0.233489, T = 0.227304; substitution rates: AC = 1.564927, AG = 4.654146, AT = 1.821542, CG = 1.281990, CT = 6.081044, GT = 1.000000; gamma distribution shape parameter α = 0.316506. 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 dark red bold and type species are in bold. Cytospora sacculus (CFCC 89626) is used as outgroup taxon Index Fungorum number: IF557294, Facesoffungi number: FoF 07837; Fig. 116 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0308 13 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 200–240 µm high × 76–95 µm diam. ( x̄ = 220 × 88 µm, n = 5), immersed to erumpent, appearing as raised, black spots, or black necks immerging through the host surface, coriaceous, solitary or scattered, subglobose to obpyriform, pale brown to brown. Fungal Diversity Table 25 Synopsis of Patellaria species with similar morphological features discussed in this study Species Ascomata (µm) P. atrata (IMI 32777) 675–1160 × 220–300 P.chromolaenae 100–145 × 415–625 (MFLUCC 17-1479) Exciple (µm) Asci (µm) Ascospores (µm) References 45–76 (15–)20–30(–35) 98–135 × 15–30 46–60(–70) × 12–16 30–45 × 7–10, 5–11-septate 23–32 × 4–5.5, 3–5-septate Yacharoen et al. (2015) This study Ostiole papillate, without periphyses. Peridium 7.5–15 µm wide, comprising 2–3 layers, pale brown to brown cells of textura angularis. Hamathecium with paraphyses not observed. Asci 30–48 × 7.5–15 µm (x̅ = 41.5 × 11 µm, n = 25), 8-spored, unitunicate, clavate to subclavate, straight to slightly curved, sessile, with a J- apical ring. Ascospores 9.5–11.5 × 3.5–4.5 µm ( x̄ = 10.5 × 4 µm, n = 25), overlapping, 1–2-seriate, hyaline, ellipsoidal, 1-septate, constricted at the septum, straight, guttulate when immature, smoothwalled. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from the apex. Colonies on MEA irregular, mycelium slightly raised, entire, white aerial hyphae, spreading from the center at the surface with white-yellow in reverse (Fig. 117). Pre-screening for antimicrobial activity: Not tested. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 5 August 2015, A. Mapook (DP21.2, MFLU 20-0308, holotype); extype culture MFLUCC 17-1422. GenBank numbers: LSU: MT214362, ITS: MT214456, SSU: MT214408 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Diaporthe chromolaenae (MFLUCC 17-1422, ex-holotype) with 100% similarity was Phomopsis asparagi strain HB3 (JQ613999) and HB2 (JQ613998). The closest match with the LSU sequence Fig. 116 Diaporthe chromolaenae (holotype) a Appearance of ascomata on substrate. b Section through ascoma. c Peridium. d–g Immature and mature asci. h–m Ascospores. Scale bars: a = 200 µm, b = 50 µm, c–g = 20 µm, h–m = 5 µm 13 Fungal Diversity Niessliaceae, Ophiocordycipitaceae, Sarocladiaceae, Stachybotryaceae, Tilachlidiaceae), based on morphology and molecular data (Hyde et al. 2020). We also follow the latest treatment and updated accounts of Trichoderma in Chaverri et al. (2015), Chen and Zhuang (2017), Qin and Zhuang (2017), Zhang and Zhuang (2017, 2018), Zhu et al. (2017), du Plessis et al. (2018) and Phookamsak et al. (2019). Fig. 117 Culture characteristic on MEA: Diaporthe chromolaenae (MFLUCC 17-1422) with 99.00% similarity was Diaporthe actinidiae (strain KFRD-8, KX609783), while the closest match with the SSU sequences with 99.64% similarity was D. amygdali (strain MUCC0101, AB454228). In the present phylogenetic analysis, D. chromolaenae is closely related to D. masirevicii strain LC6740 which is separate from the type of D. masirevicii strain BRIP 57892a (Fig. 115). However, we could not compare the morphological characteristics of both strains as the morphology of Diaporthe masirevicii strain LC6740 has not been reported. Moreover, morphological comparison of the related taxa, D. kongii (BRIP 54031) and D. masirevicii (BRIP 57892a) show that both strains differ in morphology (Table 26). Therefore, a new species is described here, based on phylogeny. Subclass Hypocreomycetidae O.E. Erikss. & Winka Hypocreales Lindau Hypocreales was introduced by Lindau (1897a). We follow the latest treatments and updated accounts of Hypocreales in Maharachchikumbura et al. (2016), and Hyde et al. (2019a, 2020). The order comprises 14 accepted families (Bionectriaceae, Calcarisporiaceae, Clavicipitaceae, Cocoonihabitaceae, Cordycipitaceae, Flammocladiellaceae, Hypocreaceae, Myrotheciomycetaceae, Nectriaceae, Hypocreaceae De Not. Hypocreaceae, introduced by De Notaris (1844), presently comprises 17 genera (Arachnocrea, Dialhypocrea, Escovopsioides, Escovopsis, Hypocreopsis, Hypomyces, Kiflimonium, Lichenobarya, Mycogone, Protocrea, Rogersonia, Sepedonium, Sphaerostilbella, Sporophagomyces, Stephanoma, Trichoderma, Verticimonosporium) and accepted in Hyde et al. (2020). Hyde et al. (2017a) reported the divergence time estimates for Hypocreomycetidae comprises Coronophorales, Clavicipitales, Falcocladiales, Hypocreales, Microascales and Torpedosporales at 171–241 Mya, during the mid Triassic to mid Jurassic period. Trichoderma Pers. Trichoderma was introduced by Persoon (1794) with T. viride as the type species. The genus contains many species which are commonly found from soil, plant materials, human and animal clinical specimens, as well as on macrofungi as hyperparasites (Klein and Eveleigh 1998; Rubini et al. 2005; Sandoval-Denis et al. 2014). Taxonomy of the T. harzianum species complex has been revised to include at least 14 species with nine new species based on multi loci (Chaverri et al. 2015). Bissett et al. (2015) listed 254 species with available representative sequences. Subsequently, several taxa have been introduced based on molecular and morphology studies. Qin and Zhuang (2017) introduced seven new species (T. angustum, T. crystalligenum, T. globoides, T. perviride, T. purpureum, T. tenue, T. viridulum) in the Harzianum and Strictipile clades. Chen and Zhuang (2017) introduced seven new species in the Viride clade, isolated from soil samples in China. Zhu et al. (2017) introduced two new species, T. fujianense and T. zonatum with green spores, while Zhang and Zhuang (2017) introduced four new species with hyaline ascospores from southwest China. Zhang and Zhuang (2018) introduced three new species, T. acremonioides, T. rugosum and T. subalni, and du Plessis et al. Table 26 Synopsis of Diaporthe species discussed in this study Species Conidiomata (µm) Conidiophores (µm) Conidiogenous cells (µm) Alpha conidia (µm) Beta conidia (µm) References D. kongii (BRIP 54031) D. masirevicii (BRIP 57892a) Up to 2000 diam. Up to 6 diam. 6–12 × 1.5–4 Up to 250 diam. 20–40 × 1.5–3.5 10–25 × 1.5–3.0 5.5–7(– 7.5) × 2–2.5(–3) (5.5–)6–7.5(– 8) × 2–3 Thompson et al. (2011) Thompson et al. (2015) 13 13–23 × 1–1.5 15–30 × 1.0–1.5 Fungal Diversity Fig. 118 Phylogram generated from maximum likelihood analysis based on combined dataset of TEF1 and RPB2 sequence data. Ninety-two strains are included in the combined sequence analysis, which comprise 2450 characters with gaps. The best scoring RAxML tree with a final likelihood value of − 15383.443045 is presented. The matrix had 1028 distinct alignment patterns, with 44.41% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.232835, C = 0.272110, G = 0.242876, T = 0.252179; substitution rates: AC = 0.975283, AG = 3.234057, AT = 0.990042, CG = 0.931920, CT = 4.895326, GT = 1.000000; gamma distribution shape parameter α = 0.323076. 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 dark red bold and type species are in bold. Nectria eustromatica (CBS 121896) and Nectria berolinensis (CBS 127382) are used as outgroup taxa (2018) described five new species (T. beinartii, T. caeruleimontis, T. chetii, T. restrictum, T. undulatum). Phookamsak et al. (2019) introduced three new species, T. koreanum, T. pinicola and T. rugulosum from Republic of Korea based on morphology and phylogeny. We describe a new isolate of T. guizhouense based on phylogeny, together with a description and illustrations (Fig. 119). A phylogenetic tree based on combined TEF1 and RPB2 sequence data is presented in Fig. 118. Trichoderma guizhouense Q.R. Li, McKenzie & Yong Wang bis, in Li et al., Mycol. Progr. 12(2): 170 (2012) [2013] Facesoffungi number: FoF 07838; Fig. 119 13 Fungal Diversity Fig. 119 Trichoderma guizhouense (new host record) a, b Appearance of ascoma on substrate. c, d Section through ascoma. e Peridium. f, g Asci. h–j Ascospores. Scale bars: a = 500 µm, b = 200 µm, c = 100 µm, d, f, g = 20 µm, e = 10 µm, h–j = 5 µm Saprobic on dead stems of Chromolaena odorata. Sexual morph: Stromata 375–555 µm long, 585–645 µm wide. ( x̄ = 480 × 615 µm, n = 5), superficial, solitary, scattered, coriaceous, olivaceous to dark green. Ascomata perithecial, 130–165(–185) μm high × 70–120 µm diam. ( x̄ = 155 × 90 µm, n = 10), arranged in rows, clustered, gregarious, with 5–6 perithecia forming groups immersed in stroma, irregular shape. Ostiole central. Peridium 5–10 µm wide, 4–5 layers, dense, comprising thick-walled, hyaline to pale brown cells of textura epidermoidea. Hamathecium with paraphyses not observed. Asci 50–80 × 3.5–5.5 µm ( x̄ = 65 × 4.5 µm, n = 15), appears to contain 16 13 part-ascospores, unitunicate, cylindrical, short pedicellate, straight or slightly curved, with inconspicuous apical ring. Ascospores 3–6.5 × 2.5–5.5 µm ( x̄ = 4.2 × 4 µm, n = 40), uniseriate, hyaline when immature, becoming unicellular, pale green to green when mature, globose or subglobose to oblong, verruculose with small guttule. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA, filamentous, mycelium slightly raised, velvety with fluffy, white aerial Fungal Diversity Fig. 120 Culture characteristic on MEA: Trichoderma guizhouense (MFLUCC 17-1512) hyphae at the surface, spreading from the center with white in reverse, filiform margin (Fig. 120). Pre-screening for antimicrobial activity: Trichoderma guizhouense (MFLUCC 17-1512) showed antimicrobial activity against E. coli with a 19 mm inhibition zone and against M. plumbeus with an 18 mm inhibition zone, observable as partial inhibition, when compared to the positive control (9 mm and 17 mm, respectively), but no inhibition of B. subtilis. Known hosts and distribution: Soil in China (Li et al. 2013); endophyte in stems of Ancistrocladus korupensis (Ancistrocladaceae) and Cola spp. (Malvaceae) in Cameroon (Chaverri et al. 2015). Material examined: THAILAND, Phrae Province, Doi Pha Klong, on dead stems of Chromolaena odorata, 22 September 2016, A. Mapook (DPKP6, MFLU 20-0381); living culture MFLUCC 17-1512 (new host record). GenBank numbers: LSU: MT214480, ITS: MT214386, SSU: MT214431, TEF1: MT235803 Notes: A phylogenetic analysis shows that the strain MFLUCC 17-1512 grouped within Trichoderma guizhouense (Fig. 118). A BLASTn search of the TEF1 sequence showed that the strain is identical to T. pollinicola with 99.78% similarity while the closest match of the RPB2 sequence with 95.29% similarity was T. lixii (strain GJS 00-18, FJ442750). We therefore, identify our isolates as T. guizhouense based on phylogenetic analyses and introduced here as a new host record from Chromolaena odorata collected in Thailand. In this study, the strain was found as the sexual morph in nature and we could not obtain its asexual morph in culture. Stachybotryaceae L. Lombard & Crous Stachybotryaceae was introduced by Crous et al. (2014) to accommodate three genera Myrothecium, Peethambara and Stachybotrys (type genus), which were earlier classified in the order Hypocreales genera incertae sedis. Subsequently, Lombard et al. (2016) resolved 33 genera in the family including nine known genera (Albosynnema, Alfaria, Didymostilbe, Myrothecium, Parasarcopodium, Peethambara, Septomyrothecium, Stachybotrys and Xepicula), and three reintroduced genera (Melanopsamma, Memnoniella, Virgatospora). In addition, 13 genera were newly described for those with myrothecium-like morphology (Albifimbria, Capitofimbria, Dimorphiseta, Gregatothecium, Inaequalispora, Myxospora, Neomyrothecium, Paramyrothecium, Parvothecium, Smaragdiniseta, Striaticonidium, Tangerinosporium, Xenomyrothecium) and eight new genera for those with stachybotrys-like morphology (Achroiostachys, Brevistachys, Cymostachys, Globobotrys, Grandibotrys, Kastanostachys, Sirastachys, Striatibotrys). Wijayawardene et al. (2018) accepted 36 genera in the family including three additional genera (Alfariacladiella, Koorchalomella, Xepiculopsis). Hyde et al. (2017a) recommended changes within the Sordariomycetes based on molecular clock evidence showing that Stachybotryaceae grouped with Nectriaceae and Niessliaceae which diverged from Clavicipitales at 157 Mya and the family was estimated in late Cretaceous for stem age at 88 Mya (50–130). Memnoniella Höhn. Memnoniella was introduced by Höhnel (1924) and has been considered as a synonym of Stachybotrys (Wang et al. 2015). However, Memnoniella was reintroduced based on morphology, with M. echinata (CBS 216.32) designated as ex-epitype strain, together with multigene analyses using cmdA, ITS, RPB2, TEF1 and TUB2 (Lombard et al. 2016). Nineteen epithets are listed in Index Fungorum (2020), with most species found as saprobic on decaying plant material. In addition, environmental samples (e.g. soil and air), as well as from indoor habitats, have also been reported, together with multigene sequence available in GenBank (Haugland and Heckman 1998; Lin et al. 2016; Lombard et al. 2016). We introduce a new species, Memnoniella chromolaenae based on morphological comparison (Table 27) together with a description and illustrations (Fig. 122). A phylogenetic tree based on combined LSU, ITS, TEF1, RPB2 and TUB2 sequence data is presented in Fig. 121. Memnoniella chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557292, Facesoffungi number: FoF 07839; Fig. 122 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype:*** Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Hyphomycetous. Colonies effuse on host, dark brown to black, powdery. Mycelium partly immersed to superficial on the host 13 Fungal Diversity Table 27 Synopsis of Memnoniella species with similar morphological features discussed in this study Species Conidiophores (µm) Conidiogenous cells (µm) Catenate conidia (µm) References M. chromolaenae (MFLUCC 17-1507) M. longistipitata (ATCC 22699) 40–85 × 2.5–4 (170–)260–460(– 610) × (3.2–)3.6–4.7(– 4.9) 85–180 × 6.5–12 5.5–7.5 × 3–4.5 9.7–10.2 × 4.7–5.5 3–4.5 × 3.5–4.8 5.8–8.5 × 6.3–8.3 This study Li et al. (2003) 9–14 × 4–6.4 8.5–12 × 4.5–7 Lin et al. (2016) M. oblongispora (MFLUCC 15-1074) surface, comprising septate, smooth, and hyaline hyphae. Conidiophores 40–85 × 2.5–4 µm ( x̄ = 56 × 3 µm, n = 10), wider at the base, macronematous, mononematous, solitary or in groups, arising from hypha, erect, unbranched, hyaline at the base becoming dark grey to black towards the apex, 2–3-septate, smooth to minutely verrucose, thickwalled, bearing 6 conidiogenous cells. Conidiogenous cells 5.5–7.5 × 3–4.5 µm ( x̄ = 6.5 × 3.8 µm, n = 15), phialidic, Fig. 121 Phylogram generated from maximum likelihood analysis based on combined dataset of LSU, ITS, TEF1, RPB2, TUB2 and cmdA sequence data. Twenty-nine strains are included in the combined sequence analysis, which comprise 3807 characters with gaps. Tree topology of the ML analysis was similar to the BYPP. The best scoring RAxML tree with a final likelihood value of − 14893.412643 is presented. The matrix had 1078 distinct alignment patterns, with 22.33% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.235371, C = 0.268803, G = 0.268291, T = 0.227535; substitution rates: AC = 1.011907, AG = 3.371254, AT = 1.029380, CG = 0.908817, CT = 7.410720, GT = 1.000000; gamma distribution shape parameter α = 0.174666. 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 dark red bold and type species are in bold. Peethambara sundara (CBS 646.77) is used as outgroup taxon 13 discrete, clustered at the apex of conidiophores, ellipsoid or obovoid to clavate or reniform, smooth, subhyaline. Conidia 3–4.5 × 3.5–4.8 µm ( x̄ = 3.5 × 4.5 µm, n = 30), catenate, acrogenous, globose to subglobose, aseptate, olivaceousgrey to black, smooth, thick-walled, formed in long chains. Culture characteristics: Conidia germinating on MEA within 24 h at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium Fungal Diversity crateriform, moderately fluffy, entire to filamentous, cultures white at the surface and pale brown to creamy-white in reverse appearing as concentric ring pattern with creamywhite at the margin (Fig. 123). Pre-screening for antimicrobial activity: Memnoniella chromolaenae (MFLUCC 17-1507) showed antimicrobial activity against E. coli with an 8 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Mai Province, Fah Hom Pok, on dead stems of Chromolaena odorata, 27 September 2016, A. Mapook (FHP11, MFLU 20-0321, holotype); ex-type culture MFLUCC 17-1507. GenBank numbers: LSU: MT214465, ITS: MT214371, SSU: MT214418 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Memnoniella chromolaenae (MFLUCC 17-1507, ex-holotype) was Mem. longistipitata with 99.83% similarity to the strain CBS 136197 (KU846155). The closest match with the LSU sequence with 99.55% similarity was Mem. echinata (strain CBS 216.32, MH866746), while the closest match with the SSU sequence with 99.91% similarity was Stachybotrys echinata (strain UAMH6594, AY489704). In the present phylogenetic analysis, Mem. chromolaenae is closely related to Mem. longistipitata and Mem. oblongispora (Fig. 121). A comparison of the ITS (+5.8S) gene region of Mem. chromolaenae and Mem. longistipitata reveals four base pair differences (0.73%) across 546 nucleotides, but we could not compare TEF1 gene region of those closely related taxa due to the lack of the TEF1 sequence. However, Mem. chromolaenae differs from Mem. longistipitata and Mem. oblongispora in having shorter conidiophore (40–85 × 2.5–4 µm vs. (170–)260–460(–610) × (3.2–)3.6–4. 7(–4.9) µm and 85–180 × 6.5–12 µm), smaller conidiogenous cells (5.5–7.5 × 3–4.5 µm vs. 9.7–10.2 × 4.7–5.5 µm and 9–14 × 4–6.4 µm) and smaller conidia (3–4.5 × 3.5–4.8 µm vs. 5.8–8.5 × 6.3–8.3 µm and 8.5–12 × 4.5–7 µm) (Table 27). Therefore, Mem. chromolaenae is described as a new species based on morphological comparison. Subclass Xylariomycetidae O.E. Erikss & Winka Amphisphaeriales D. Hawksw. & O.E. Erikss.* Amphisphaeriales was introduced by Eriksson and Hawksworth (1986). We follow the latest treatment and updated accounts of Amphisphaeriales in Hyde et al. (2020) with 14 accepted families (Amphisphaeriaceae, Apiosporaceae, Beltraniaceae, Clypeophysalosporaceae, Cylindriaceae, Hansfordiaceae, Hyponectriaceae, Iodosphaeriaceae, Melogrammataceae, Phlogicylindriaceae, Pseudomassariaceae, Fig. 122 Memnoniella chromolaenae (holotype) a, b Appearance of colonies on substrate. c-f Conidiophores, conidiogenous cells and conidia. g–h Conidia in chains. i Conidium. j Germinated conidia. Scale bars: a = 500 µm, b = 200 µm, c, d, g, j = 20 µm, e–f, h = 10 µm, i = 5 µm 13 Fungal Diversity Fig. 123 Culture characteristic on MEA: Memnoniella chromolaenae (MFLUCC 17-1507) Pseudotruncatellaceae, Sporocadaceae, Vialaeaceae). Furthermore, we follow the latest treatments and updated accounts of the genus Arthrinium in Crous and Groenewald (2013), Wang et al. (2018), Jiang et al. (2018), Pintos et al. (2019) and Hyde et al. (2020). Apiosporaceae K.D. Hyde et al. Apiosporaceae was introduced by Hyde et al. (1998) to accommodate the genera Appendicospora and Apiospora, which had been earlier assigned to Lasiosphaeriaceae by Barr (1990) based on morphology, with Apiospora as the type genus. Crous and Groenewald (2013) synonymized Apiospora, Pteroconium and Cordella under the genus Arthrinium. Six genera, Appendicospora, Arthrinium, Dictyoarthrinium, Endocalyx, Scyphospora and Spegazzinia were previously accepted in Outline of Ascomycota 2017 (Wijayawardene et al. 2018). However, Tanaka et al. (2015) and Jayasiri et al. (2019) placed Spegazzinia in Didymosphaeriaceae based on molecular data. Scyphospora was synonymized under the genus Arthrinium (Nag Raj 1974; Kirk 1986; Hyde et al. 2020), and Nigrospora was transferred to Apiosporaceae based on morphology and phylogeny (Wang et al. 2017). Recently, Hyde et al. (2020) placed Appendicospora, Arthrinium, Dictyoarthrinium, Endocalyx and Nigrospora in the family. Divergence time estimates for the family have been reported based on molecular clock evidence. The family was estimated for stem age at 69 Mya (50–130) in the late Cretaceous (Hyde et al. 2017a). Arthrinium Kunze, in Kunze & Schmidt, Mykologische Hefte (Leipzig) 1: 9 (1817) Arthrinium is an ecologically diverse genus found on various substrates and associated with various plants as endophytes, saprobes, and plant pathogens on some important ornamentals (Chen et al. 2014; Li et al. 2016), and also causes cutaneous infections in humans (Rai 1989; Zhao et al. 13 1990; de Hoog et al. 2000; Crous et al. 2013; Wang et al. 2018). The genus was introduced by Schmidt and Kunze (1817) with A. caricicola as the type species. Crous and Groenewald (2013) described eight new species (A. hydei, A. kogelbergense, A. malaysianum, A. ovatum, A. phragmites, A. pseudospegazzinii, A. pseudosinense, A. xenocordella) based on phylogeny and morphology. Wang et al. (2018) introduced eight new species (A. bambusae, A. camelliaesinensis, A. dichotomanthi, A. guizhouense, A. jiangxiense, A. obovatum, A. pseudoparenchymaticum, A. subroseum) from China and reported the major host plant families of Arthrinium species were Poaceae and Cyperaceae. Jiang et al. (2018) introduced two new species, A. qinlingense and A. gaoyouense from China. Subsequently, Pintos et al. (2019) introduced six new species (A. balearicum, A. descalsii, A. esporlense, A. ibericum, A. italicum, A. piptatheri) from Europe. We introduce a new species Arthrinium chromolaenae based on phylogeny and morphological comparison (Table 28) together with a description and illustrations (Fig. 125). A phylogenetic tree based on combined ITS, LSU, TEF1, and TUB2 sequence data is presented in Fig. 124. Arthrinium chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557288, Facesoffungi number: FoF 07840; Fig. 125 Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0300 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Undetermined. Asexual morph: Colonies on host surface, black, powdery. Mycelium consisting of smooth, hyaline, branched, septate, hyphae. Conidiophores reduced to conidiogenous cells, grouped together to form sporodochia. Conidiogenous cells basauxic, 6.5–12 µm high × 1–2 µm diam. ( x̄ = 8.5 × 1.5 µm, n = 5), aggregated in clusters on hyphae, hyaline, smooth, elongated, broadly filiform to ampulliform. Conidia 4–6 × 4.5–6.5 µm ( x̄ = 5 × 5.5 µm, n = 35), in surface view, irregular arrangement in pseudo-chains, pale brown to brown, smooth, globose to subglobose, lenticular in side view, with pale equatorial slit. Culture characteristics: Conidia germinating on MEA within 24 h at room temperature, germ tubes produced from conidia. Colonies on MEA circular, white, flat, spreading, with fluffy aerial mycelium, filiform, mycelia not tightly attached to the surface (Fig. 126). Pre-screening for antimicrobial activity: Arthrinium chromolaenae (MFLUCC 17-1505) showed antimicrobial activity against E. coli with an 8 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Fungal Diversity Table 28 Synopsis of Arthrinium species with similar morphological features discussed in this study Species Conidiophores (µm) Conidiogenous cells (µm) Conidia (µm) References A. chromolaenae (MFLUCC 17-1505) A. euphorbiae (IMI 110788) – 6.5–12 × 1–2 4–6 × 4.5–6.5 This study – – A. italicum (CBS 145138) A. thailandicum (MFLUCC 15–0202) A. vietnamense (IMI 99670= A. malaysianum) 10–50 × 1–3 5.5–11 × 3–4.5 (3–)4–7(–9) × (1.5–)2–3(–5) 11.5–39 × 2–3.5 (4–)4.7(–5.5) in surface view, Ellis (1965), Crous et al. (2013) (3–)3.2(–4) in side view 4–6 × 3–4 Pintos et al. (2019) 5–9 × 5–8 Dai et al. (2017) – 4–7 × 3–5 Material examined: THAILAND, Chiang Mai Province, Fah hom pok, on dead stems of Chromolaena odorata, 27 September 2016, A. Mapook (FHP9, MFLU 20-0300, holotype); ex-type culture MFLUCC 17-1505. GenBank numbers: LSU: MT214436, ITS: MT214342, TEF1: MT235802 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of A. chromolaenae (MFLUCC 17-1505, ex-holotype) was A. malaysianum with 97.38% similarity to the strain CBS 251.29 (KF144897). The closest match with the LSU sequence with 100% similarity was A. malaysianum (strain CBS 102053, NG_042780), while the closest match with the TEF sequence with 99.06% similarity was Arthrinium sp. (strain HKU41, LC439290). Crous et al. (2013) introduced A. malaysianum as a new species based on phylogeny and morphologically comparing with a closely related strain, A. euphorbiae. They mentioned that A. malaysianum differs from A. euphorbiae in having slightly longer conidia. Wang et al. (2017) demonstrated that the ex-type strain of A. vietnamensis and A. malaysianum are conspecific, supported by morphological comparison and proposed use of the older name, A. vietnamensis. Phylogenetic analysis showed that A. chromolaenae is closely related to the clade comprising, A. euphorbiae and A. vietnamense (= A. malaysianum) (Fig. 124). However, A. chromolaenae differs from A. vietnamense in having long and narrow conidiogenous cells (6.5–12 × 1–2 µm vs. 4–7 × 3–5 µm) (Table 28). A comparison of the ITS (+ 5.8S) gene region of A. chromolaenae and A. vietnamense reveals 16 base pair differences (2.62%) across 610 nucleotides. Therefore, A. chromolaenae is described as a new species based on phylogeny and morphological comparison. Xylariales Nannf. Xylariales was introduced by Nannfeldt (1932). We follow the latest treatment and updated accounts of Xylariales in Hyde et al. (2020) with 15 accepted families (Barrmaeliaceae, Cainiaceae, Clypeosphaeriaceae, Coniocessiaceae, Diatrypaceae, Graphostromataceae, Hansfordiaceae, Hypoxylaceae, Induratiaceae, Lopadostomataceae, 5–6 in surface view, 3–4 in side view Crous et al. (2013), Wang et al. (2017) Microdochiaceae, Polystigmataceae, Requienellaceae, Xylariaceae, Zygosporiaceae). Furthermore, we follow the latest treatments and updated accounts of the family Cainiaceae in Liu et al. (2015), Maharachchikumbura et al. (2015, 2016), Senanayake et al. (2015), Wijayawardene et al. (2017a, 2018) and Hyde et al. (2020). Cainiaceae J.C. Krug Cainiaceae was introduced by Krug (1977) to accommodate the generic type, Cainia, which is characterized by unique apical rings in the asci and longitudinal germ slit in the ascospores. Kang et al. (1999) revived the family and accepted five genera, Atrotorquata, Arecophila, Cainia, Ceriophora and Reticulosphaeria based on morphology. Subsequently, several molecular and morphological studies have confirmed the placement of these genera with five accepted genera in Cainiaceae (Amphibambusa, Arecophila, Atrotorquata, Cainia, Seynesia) (Jeewon et al. 2003; Smith et al. 2003; Liu et al. 2015; Maharachchikumbura et al. 2015, 2016; Senanayake et al. 2015; Wijayawardene et al. 2017a, 2018), while Reticulosphaeria was placed in Amphisphaeriales genera incertae sedis (Senanayake et al. 2015). The family has been placed within the Xylariales (Maharachchikumbura et al. 2015, 2016; Senanayake et al. 2015; Jaklitsch et al. 2016), while divergence time estimates suggested that the family share the most common ancestor with Xylariales, with a divergence time of ca 128 Mya (Hyde et al. 2017a). Hyde et al. (2020) introduced a new genus Alishanica from Taiwan in the family and placed six genera, Alishanica, Amphibambusa, Arecophila, Atrotorquata, Cainia and Seynesia in Cainiaceae within Xylariales. Longiappendispora Mapook & K.D. Hyde, gen. nov. Index Fungorum number: IF557336, Facesoffungi number: FoF 07841 Etymology: The generic epithet reflects the ascospores with long bristle-like polar appendages. Saprobic on dead stems. Sexual morph: Ascomata immersed beneath clypeus, unilocular, globose to subglobose, coriaceous, solitary or scattered, or sometimes 13 Fungal Diversity Fig. 124 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS, LSU, TEF1, and TUB2 sequence data. Eightysix strains are included in the combined sequence analysis, which comprise 3780 characters with gaps. The best scoring RAxML tree with a final likelihood value of − 25718.058495 is presented. The matrix had 1793 distinct alignment patterns, with 39.70% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.240172, C = 0.253238, G = 0.253613, T = 0.252978; substitution rates: AC = 1.178756, AG = 2.769752, AT = 1.160953, CG = 0.997632, CT = 4.711267, GT = 1.000000; gamma distribution shape parameter α = 0.246871. Bootstrap support values for ML equal to or greater than 60% and BYPP equal to or greater than 0.90 are given at the nodes. Newly generated sequences are in dark red bold and type species are in bold. Seiridium phylicae (CPC 19965) is used as outgroup taxon 13 Fungal Diversity Fig. 125 Arthrinium chromolaenae (holotype) a, b Appearance of colonies on substrate. c–g Conidia and conidiogenous cells. h–k Conidia. Scale bars: a = 500 µm, b = 100 µm, h = 20 µm, c–g, i–k = 5 µm Fig. 126 Culture characteristic on MEA: Arthrinium chromolaenae (MFLUCC 17-1505) gregarious. Ostiole central. Peridium comprising 2–3 layers, pale brown to hyaline cells of textura angularis. Hamathecium composed of cylindrical to broadly filiform, septate, guttulate, hyaline paraphyses, embedded in a gelatinous matrix. Asci 8-spored, unitunicate, cylindrical to broadly filiform, short-pedicellate, straight or slightly curved, with an apical ring, apically rounded. Ascospores uniseriate, hyaline when immature, becoming brown to dark brown when mature, 1-septate, ellipsoid to broadly fusiform, tapering towards narrow ends, constricted at the septum, straight or slightly curved, guttulate at both cells, with longitudinal striations and bristle-like polar appendages from both ends, without a gelatinous sheath. Asexual morph: Undetermined. Type species: Longiappendispora chromolaenae Mapook & K.D. Hyde Notes: A phylogenetic analysis based on combined dataset of the ITS and LSU sequence data show that Longiappendispora chromolaenae groups with a new genus, Alishanica miscanthii (FU 31025), with low bootstrap support in Cainiaceae and clusters with Cainia clade (0.93 BYPP, Fig. 127). However, Longiappendispora species differ from Alishanica by larger ascomata (500–580 × 360–505 µm vs. 277–272 × 288–285 µm), a thin peridium wall [(15–)20–30 µm vs. 52–60 µm], narrower asci (140–230 × 13–20 µm vs. 173–179 × 23–31 µm) and smaller ascospores (28.5–43 × 9.5–12 µm vs. 59–62 × 19–21 µm) with bristle-like polar appendages on both ends of the ascospores and without a gelatinous sheath (Table 29). Therefore, we introduce Longiappendispora as a new genus with a new species L. chromolaenae, based on morphological comparison with phylogenetic analyses. A comparison of the ITS (+5.8S) gene region of L. chromolaenae and Alishanica miscanthii reveals 78 base pair differences (17.9%) across 435 nucleotides. Longiappendispora chromolaenae Mapook & K.D. Hyde, sp. nov. Index Fungorum number: IF557337, Facesoffungi number: FoF 07842; Fig. 128 13 Fungal Diversity Fig. 127 Phylogram generated from maximum likelihood analysis based on combined dataset of ITS and LSU sequence data. Fourteen strains are included in the combined sequence analysis, which comprise 1506 characters with gaps. The best scoring RAxML tree with a final likelihood value of − 4717.509202 is presented. The matrix had 329 distinct alignment patterns, with 22.49% of undetermined characters or gaps. Estimated base frequencies were as follows: A = 0.255033, C = 0.233739, G = 0.280548, T = 0.230680; substitution rates: AC = 2.006227, AG = 2.877479, AT = 3.211990, CG = 2.743197, CT = 8.119404, GT = 1.000000; gamma distribution shape parameter α = 0.202825. 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 dark red bold and type species are in bold. Entosordaria perfidiosa (CBS 142773) is used as outgroup taxon Table 29 Morphological features of species in Cainiaceae discussed in this study Species Asocomata (µm) Peridium (µm) Asci (µm) Longiappendispora chromolaenae (MFLUCC 17-1485) 500–580 high × 360– 505 diam. (15–)20–30 Alishanica miscanthii 277–272 high × 288– (FU 31025) 285 diam. 13 52–60 Ascospores (µm) Morphology of ascospores References This study 140–230 × 13–20 28.5–43 × 9.5–12 With longitudinal striations and bristle-like polar appendages from both ends, without gelatinous sheath Hyde et al. (2020) 173–179 × 23–31 59–62 × 19–21 With striations, surrounded by a thick, hyaline mucilaginous sheath sub globose, parallel to the margin of the spore Fungal Diversity Fig. 128 Longiappendispora chromolaenae (holotype) a, b Appearance of ascomata on substrate. c Section through ascoma. d Ostiole. e Peridium. f Pseudoparaphyses. g–h Asci. i–o Ascospores. Scale bars: a = 500 µm, b, c = 200 µm, d, g–i = 50 µm, j–o = 20 µm, e, f = 5 µm Etymology: Name reflects the host genus Chromolaena, from which this species was isolated. Holotype: MFLU 20-0320 Saprobic on dead stems of Chromolaena odorata. Sexual morph: Ascomata 500–580 µm high × 360–505 µm diam. ( x̄ = 535 × 440 µm, n = 5), immersed beneath clypeus, unilocular, globose to subglobose, coriaceous, solitary or scattered, sometimes gregarious. Ostiole central. Peridium (15–)20–30 µm wide, comprising 2–3 layers, pale brown to hyaline cells of textura angularis. Hamathecium comprising 2–4 µm wide, cylindrical to broadly filiform, septate, guttulate, hyaline paraphyses, embedded in a gelatinous matrix. Asci 140–230 × 13–20 µm ( x̄ = 195 × 15.5 µm, n = 15), 8-spored, unitunicate, cylindrical to broadly filiform, short pedicellate, straight or slightly curved, with an apical ring, apically rounded. Ascospores 28.5–43 × 9.5–12 µm ( x̄ = 33 × 11 µm, n = 30), uniseriate, hyaline when immature, becoming brown to dark brown when mature, 1-septate, 13 Fungal Diversity name and fungal classification are used according to Index Fungorum (2020), outline of Ascomycota (Wijayawardene et al. 2018), Notes for genera: Ascomycota (Wijayawardene et al. 2017b) and recent relevant literature. Phylum Ascomycota Caval.-Sm. Class Dothideomycetes sensu O.E. Erikss & Winka Subclass Dothideomycetidae P.M. Kirk et al. Capnodiales Woron. Capnodiaceae (Sacc.) Höhn. ex Theiss. 1. Capnodium sp. Venezuela (Urtiaga 1986) Fig. 129 Culture characteristic on MEA: Longiappendispora chromolaenae (MFLUCC 17-1485) ellipsoid to broadly fusiform, tapering towards narrow ends, constricted at the septum, straight or slightly curved, guttulate at both cells, with longitudinal striations and bristlelike polar appendages from both ends, without a gelatinous sheath. Asexual morph: Undetermined. Culture characteristics: Ascospores germinating on MEA within 24 h. at room temperature and germ tubes produced from both ends. Colonies on MEA circular, mycelium raised and velvety with moderately fluffy, filamentous, white aerial hyphae at the surface and orange in reverse with white-yellow at the margin (Fig. 129). Pre-screening for antimicrobial activity: Longiappendispora chromolaenae (MFLUCC 17-1485) showed antimicrobial activity against E. coli with a 9 mm inhibition zone when compared to the positive control (9 mm), but no inhibition of B. subtilis and M. plumbeus. Material examined: THAILAND, Chiang Rai Province, Doi Pui, on dead stems of Chromolaena odorata, 2 February 2017, A. Mapook (DP83, MFLU 20-0320, holotype); ex-type culture MFLUCC 17-1485. GenBank numbers: LSU: MT214464, ITS: MT214370, SSU: MT214417 Notes: In a BLASTn search of NCBI GenBank, the closest match of the ITS sequence of Longiappendispora chromolaenae (MFLUCC 17-1485, ex-holotype) with 84.88% similarity was Cainia desmazieri (strain CBS 137.62, MH858124), while the closest match with the LSU sequence with 96.99% similarity was Cainia graminis (strain CBS 136.62, AF431949). Checklist of fungi associated with Chromolaena odorata (= Eupatorium odoratum) The checklist of fungi associated with C. odorata is based on the USDA Systematic Mycology and Microbiology Laboratory (SMML) database (Farr and Rossman 2020), together with relevant literature and the author’s studies. The current 13 Cladosporiaceae Nann. 2. Cladosporium sp. West Indies (Minter et al. 2001) 3. Cladosporium eupatorii Cif., Sydowia 8(1–6): 251 (1954) (= Hormodendrum eupatorii) Dominican Republic (Ciferri 1961; Barreto and Evans 1994) Mycosphaerellaceae Lindau 4. Cercospora sp. Cambodia (Litzenberger et al. 1994), Malaysia (Johnston 1960; Williams and Liu 1976), Nepal (Barreto and Evans 1994), Thailand (Puckdeedindan 1966; Nguanhom et al. 2015), Venezuela (Urtiaga 2004) 5. Micronematomyces caribensis (Crous & den Breeÿen) U. Braun, C. Nakash., Videira & Crous, in Videira et al., Stud. Mycol. 87: 337 (2017) Cuba, Jamaica (den Breeyen et al. 2006; Videira et al. 2017) 6. Micronematomyces chromolaenae (Crous & den Breeÿen) U. Braun, C. Nakash., Videira & Crous, in Videira et al. Stud. Mycol. 87: 337 (2017) Mexico (den Breeyen et al. 2006; Videira et al. 2017) 7. Passalora capsicicola (Vassiljevsky) U. Braun & F.O. Freire, Cryptog. Mycol. 23(4): 299 (2003) [2002] = Phaeoramularia capsicicola (Vassiljevsky) Deighton, in Ellis, Trans. Br. mycol. Soc. 67(1): 140 (1976) Jamaica, USA (Videira et al. 2017) 8. Passalora sp. Jamaica, USA (den Breeyen et al. 2006) 9. Phloeospora sp. Venezuela (Urtiaga 1986; Barreto and Evans 1994) 10. Pseudocercospora aciculina (Chupp) U. Braun & Crous, in Crous & Braun, CBS Diversity Ser. (Utrecht) 1: 42 (2003) Fungal Diversity 11. 12. 13. 14. 15. 16. 17. = Cercospora aciculina Chupp, Monograph of Cercospora: 118 (1954) Cambodia (Litzenberger el al. 1962), Nigeria (Fajola 1978; Barreto and Evans 1994) Pseudocercospora ageratoides (Ellis & Everh.) Y.L. Guo, Acta Mycol. Sin. 12(4): 271 (1993) = Cercospora ageratoides Ellis & Everh., J. Mycol. 5(2): 71 (1889) Cote d’Ivoire (Yen 1974) Pseudocercospora convoluta (Crous & den Breeÿen) U. Braun, C. Nakash., Videira & Crous, in Videira et al., Stud. Mycol. 87: 312 (2017) = Passalora convoluta Crous & den Breeÿen, in Breeÿen, Groenewald, Verkley & Crous, Fungal Divers. 23: 96 (2006) Costa Rica (den Breeyen et al. 2006; Videira et al. 2017) Pseudocercospora eupatoriella Crous & den Breeÿen, in Breeÿen, Groenewald, Verkley & Crous, Fungal Divers. 23: 101 (2006) Jamaica (den Breeyen et al. 2006; Guatimosim et al. 2016), USA (den Breeyen et al. 2006) Pseudocercospora eupatorii (Peck) U. Braun & R.F. Castañeda, Cryptog. bot. 2(2-3): 293 (1991) = Cercospora eupatorii Peck, Ann. Rep. N.Y. St. Mus. nat. Hist. 33: 29 (1883) [1880] Cambodia, Cote d’Ivoire, Cuba, Hawaii, India, Nepal, North America (Govindu et al. 1970; Yen 1974; Urtiaga 1986; Barreto and Evans 1994), Andaman Islands (Hosagoudar and Mathew 2000), India (Kamal 2010), Thailand (Meeboon et al. 2007) Pseudocercospora eupatorii-formosani U. Braun & Bagyan., Sydowia 51(1): 8 (1999) Bangladesh (Barreto and Evans 1994), Brunei Darussalam (Peregrine and Ahmad 1982; Barreto and Evans 1994; Braun and Sivapalan 1999), China (Roux et al. 1997; Liu and Guo 1998; Guo 1999a, b; Zhuang 2001), India (Barreto and Evans 1994), Laos (Phengsintham et al. 2010), Malaysia (Barreto and Evans 1994), Myanmar (Thaung 1984; Barreto and Evans 1994), Taiwan (Kirschner and Chen 2007), Thailand (Barreto and Evans 1994; Phengsintham et al. 2013), Venezuela (Urtiaga and Braun 2013) Pseudocercospora sp. Guatemala (Videira et al. 2017), Mexico (den Breeyen et al. 2006) Ragnhildiana perfoliati (Ellis & Everh.) U. Braun, C. Nakash., Videira & Crous, in Videira, Groenewald, Nakashima, Braun, Barreto, de Wit & Crous, Studies in Mycology 87: 345 (2017) = Mycovellosiella perfoliati (Ellis & Everh.) U. Braun, C. Nakash., Videira & Crous 1960 18. 19. 20. 21. 22. 23. 24. 25. = Passalora perfoliati (Ellis & Everh.) U. Braun & Crous, in Crous & Braun, CBS Diversity Ser. (Utrecht) 1: 314 (2003) Brazil (Barreto and Evans 1994), Laos (Videira et al. 2017), West Indies (Minter et al. 2001), Jamaica (den Breeyen et al. 2006), Taiwan (Kirschner and Chen 2007) Ramularia eupatorii J.M. Yen, Cahiers Pacif. 13: 272 (1969) (= Mycosphaerella eupatorii) Malaysia (Yen 1969; Barreto and Evans 1994) Ramularia tungurahuana Petr., Sydowia 4(1-6): 507 (1950) (= Mycosphaerella tungurahuana) Malaysia (Yen 1969) Septoria chromolaenae Crous & den Breeÿen, in Breeÿen et al., Fungal Divers 23: 102 (2006) Cuba (Verkley et al. 2013), Jamaica (den Breeyen et al. 2006) Septoria ekmaniana Petr. & Cif., Annls mycol. 30(3/4): 300 (1932) Dominican Republic (Ciferri 1961; Barreto and Evans 1994), Mexico (den Breeyen et al. 2006), West Indies (Minter et al. 2001) Septoria eupatorii Roberge ex Desm., Annls Sci. Nat., Bot., sér. 3 20: 90 (1853) Venezuela (Urtiaga 1986; Barreto and Evans 1994) Septoria sp. Guam (Russo 1985; Barreto and Evans 1994) Myriangiales Starbäck Myriangiaceae Nyl. Anhellia nigra (Viégas) Arx, Persoonia 2(4): 434 (1963) Rhytidhysteron bruguierae Dayarathne, in Dayarathne et al., Mycosphere 11(1): 20 (2020) = Tubercularia nigra F. Stevens, Annls mycol. 28(5/6): 371 (1930) Brazil (Barreto and Evans 1994), West Indies (Minter et al. 2001) Subclass Pleosporomycetidae C.L. Schoch et al. Hysteriales Lindau Hysteriaceae Chevall. Thailand (Dayarathne et al. 2020; This study) 26. Rhytidhysteron chromolaenae Mapook & K.D. Hyde Thailand (This study) Pleosporales Luttrell ex M.E. Barr Acrocalymmaceae Crous & Trakun. 13 Fungal Diversity 27. Acrocalymma medicaginis Alcorn & J.A.G. Irwin, Trans. Br. mycol. Soc. 88(2): 163 (1987) Thailand (This study) 44. Hermatomyces chromolaenae Jun F. Li, Mapook & K.D. Hyde, in Tibpromma et al., Fungal Divers 83: 47 (2017) Thailand (Tibpromma et al. 2017) Corynesporascaceae Sivan. Lophiostomataceae Sacc. 28. Corynespora cassiicola (Berk. & M.A. Curtis) C.T. Wei, Mycol. Pap. 34: 5 (1950) Palau (Dixon et al. 2009) Didymellaceae Gruyter et al. 29. Didymella chromolaenae Mapook & K.D. Hyde Thailand (This study) 30. Nothophoma chromolaenae Mapook & K.D. Hyde Thailand (This study) 45. Flabellascoma minimum A. Hashim., K. Hiray. & Kaz. Tanaka, in Hashimoto et al., Studies in Mycology 90: 169 (2018) Thailand (This study) 46. Pseudocapulatispora longiappendiculata Mapook & K.D. Hyde Thailand (This study) Melanommataceae G. Winter (= Pseudodidymellaceae A. Hashim. & Kaz. Tanaka) Didymosphaeriaceae Munk 31. Chromolaenicola chiangraiensis Mapook & K.D. Hyde Thailand (This study) 32. Chromolaenicola lampangensis Mapook & K.D. Hyde Thailand (This study) 33. Chromolaenicola nanensis Mapook & K.D. Hyde Thailand (This study) 34. Chromolaenicola thailandensis Mapook & K.D. Hyde Thailand (This study) 35. Didymosphaeria sp. India (Barreto and Evans 1994) 36. Montagnula chiangraiensis Mapook & K.D. Hyde Thailand (This study) 37. Montagnula chromolaenae Mapook & K.D. Hyde Thailand (This study) 38. Montagnula chromolaenicola Mapook & K.D. Hyde Thailand (This study) 39. Montagnula thailandica Mapook & K.D. Hyde Thailand (This study) 40. Pseudopithomyces palmicola J.F. Li, Ariyaw. & K.D. Hyde, in Ariyawansa et al., Fungal Divers. 75: 41 (2015) Thailand (This study) 41. Tremateia chiangraiensis Mapook & K.D. Hyde Thailand (This study) 42. Tremateia chromolaenae Mapook & K.D. Hyde Thailand (This study) 43. Tremateia thailandensis Mapook & K.D. Hyde Thailand (This study) 47. Byssosphaeria schiedermayeriana (Fuckel) M.E. Barr, Mycotaxon 20(1): 34 (1984) = Herpotrichia schiedermayeriana Fuckel, Jb. nassau. Ver. Naturk. 27-28: 27 (1874) [1873-74] Brazil (Barreto and Evans 1994) Nigrogranaceae Jaklitsch & Voglmayr 48. Nigrograna chromolaenae Mapook & K.D. Hyde Thailand (This study) Neomassarinaceae Mapook & K.D. Hyde 49. Neomassarina chromolaenae Mapook & K.D. Hyde Thailand (This study) 50. Neomassarina thailandica Phook., Jayasiri & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 138 (2016) Thailand (This study) Neopyrenochaetaceae Valenzuela-Lopez et al. 51. Neopyrenochaeta chiangraiensis Mapook & K.D. Hyde Thailand (This study) 52. Neopyrenochaeta chromolaenae Mapook & K.D. Hyde Thailand (This study) 53. Neopyrenochaeta thailandica Mapook & K.D. Hyde Thailand (This study) 54. Neopyrenochaeta triseptatispora Mapook & K.D. Hyde Thailand (This study) Hermatomycetaceae Locq. Phaeosphaeriaceae M.E. Barr 13 Fungal Diversity 55. Leptospora chromolaenae Mapook & K.D. Hyde Thailand (This study) 56. Leptospora phraeana Mapook & K.D. Hyde Thailand (This study) 57. Leptospora thailandica Phukhams. & K.D. Hyde, in Hyde et al., Fungal Divers. 80: 100 (2016) Thailand (This study) 58. Murichromolaenicola chiangraiensis Mapook & K.D. Hyde Thailand (This study) 59. Murichromolaenicola chromolaenae Mapook & K.D. Hyde Thailand (This study) 60. Neoophiobolus chromolaenae Mapook & K.D. Hyde Thailand (This study) 61. Ophiobolus ipohensis J.M. Yen, Cahiers Pacif. 13: 277 (1969) Malaysia (Yen 1969; Barreto and Evans 1994) 62. Ophiosphaerella eupatorii J.M. Yen, Cahiers Pacif. 13: 278 (1969) Malaysia (Yen 1969; Barreto and Evans 1994) 63. Paraleptospora chromolaenae Mapook & K.D. Hyde Thailand (This study) 64. Paraleptospora chromolaenicola Mapook & K.D. Hyde Thailand (This study) 65. Phaeosphaeria eupatoriicola J.M. Yen, Cahiers Pacif. 13: 279 (1969) Malaysia (Yen 1969; Barreto and Evans 1994) 66. Pseudoophiosphaerella huishuiensis J.F. Zhang, J.K. Liu & Z.Y. Liu, in Zhang, Liu, Jeewon, Wanasinghe & Liu, Mycosphere 8(1): 207 (2019) Thailand (This study) 67. Pseudostaurosphaeria chromolaenae Mapook & K.D. Hyde Thailand (This study) 68. Pseudostaurosphaeria chromolaenicola Mapook & K.D. Hyde Thailand (This study) 69. Setophoma chromolaenae Quaedvl., Verkley, R.W. Barreto & Crous, Studies in Mycology 75: 373 (2013) Brazil (Quaedvlieg et al. 2013) 70. Yunnanensis chromolaenae Mapook & K.D. Hyde Thailand (This study) Pleosporaceae Nitschke 71. Alternaria zinniae M.B. Ellis, Mycol. Pap. 131: 22 (1972) Brazil (Barreto and Evans 1994) Pleosporales genera incertae sedis 72. Scolecobasidium eupatorii Y.L. Guo, Acta Mycol. Sin. 12(4): 271 (1993) China (Zhuang 2001) Pyrenochaetopsidaceae Valenzuela-Lopez et al. 73. Pyrenochaetopsis chromolaenae Mapook & K.D. Hyde Thailand (This study) Roussoellaceae J.K. Liu et al. 74. Pseudoroussoella chromolaenae Mapook & K.D. Hyde Thailand (This study) 75. Pseudoroussoella elaeicola (Konta & K.D. Hyde) Mapook & K.D. Hyde Thailand (This study) 76. Setoarthopyrenia chromolaenae Mapook & K.D. Hyde Thailand (This study) 77. Xenoroussoella triseptata Mapook & K.D. Hyde Thailand (This study) Thyridariaceae Q. Tian & K.D. Hyde 78. Chromolaenomyces appendiculatus Mapook & K.D. Hyde Thailand (This study) 79. Pseudothyridariella chromolaenae Mapook & K.D. Hyde Thailand (This study) Torulaceae Corda 80. Torula chromolaenae JF. Li, Phook., Mapook & K.D. Hyde, in Li et al., Mycol. Progr. 16(4): 454 (2017) Thailand (Li et al. 2017; This study) 81. Torula fici Crous, in Crous et al., IMA Fungus 6(1): 192 (2015) Thailand (Li et al. 2017; This study) 82. Torula polyseptata C.G. Lin & K.D. Hyde, in Hyde et al., Fungal Divers. 96: 71 (2019) Thailand (Hyde et al. 2019a; This study) Dothideomycetes orders incertae sedis Botryosphaeriales C.L. Schoch et al. Aplosporellaceae Slippers et al. 83. Aplosporella chromolaenae Mapook & K.D. Hyde Thailand (This study) 84. Aplosporella hesperidica Speg., Anal. Soc. cient. argent. 13(1): 18 (1882) 13 Fungal Diversity Thailand (This study) Botryosphaeriaceae Theiss. & P. Syd., Annales Mycologici 16 (1-2): 16 (1918) 85. Dothiorella oblonga F.J.J. Van der Walt, Slippers & G.J. Marais, in Slippers et al., Persoonia 33: 163 (2014) Thailand (This study) 86. Sphaeropsis chromolaenicola Mapook & K.D. Hyde Thailand (This study) 94. Patellaria chromolaenae Mapook & K.D. Hyde Thailand (This study) Dothideomycetes families incertae sedis Dysrhynchisceae Boonmee & K.D. Hyde* 95. Dysrhynchis citricola Bat. & Peres, Nova Hedwigia 2(4): 469 (1960) Dominican Republic (Ciferri 1961; Barreto and Evans 1994) Dothideomycetes genera incertae sedis Phyllostictaceae Fr. 87. Asteromella eupatoriicola (Kabát & Bubák) H. Ruppr. 1958 = Phyllosticta eupatoriicola Kabát & Bubák, Hedwigia 46: 288 (1907) Malaysia, South America (Barreto and Evans 1994), Puerto Rico (Stevenson 1975), Virgin Islands (Stevenson 1975), West Indies (Minter et al. 2001) 88. Phyllosticta eupatorii Punith., Mycol. Pap. 136: 21 (1974) Sri Lanka (Barreto and Evans 1994) Dyfrolomycetales K.L. Pang et al. Pleurotremataceae Walt. Watson (= Dyfrolomycetaceae K.D. Hyde et al.)* 89. Dyfrolomyces chromolaenae Mapook & K.D. Hyde Thailand (This study) Muyocopronales Mapook et al. Muyocopronaceae K.D. Hyde 90. Muyocopron chromolaenae Mapook & K.D. Hyde Thailand (This study) 91. Muyocopron chromolaenicola Mapook & K.D. Hyde Thailand (This study) 92. Muyocopron lithocarpi Mapook, Boonmee & K.D. Hyde, Phytotaxa 265(3): 235 (2016), new host record Thailand (This study) 93. Neocochlearomyces chromolaenae Pinruan, Sommai, Suetrong, J.Z. Groenew. & Crous, in Crous et al., Persoonia 41: 381 (2018) Thailand (Crous et al. 2018) Patellariales D. Hawksw. & O.E. Erikss. Patellariaceae Corda 13 96. Ampullifera sp. West Indies (Minter et al. 2001) 97. Melioliphila melioloides (Speg.) Piroz., Kew Bull. 31(3): 596 (1977) West Indies (Minter et al. 2001) 98. Phaeoramularia eupatorii-odorati (J.M. Yen) X.J. Liu & Y.L. Guo, Acta phytopath. sin. 12(no. 4): 7 (1982) = Mycovellosiella eupatorii-odorati (J.M. Yen) J.M. Yen, Bull. trimest. Soc. mycol. Fr. 97(3): 131 (1981) China (Liu and Guo 1988; Morgan-Jones 1997; Roux et al. 1997; Zhuang 2001; Guo and Liu 2003), Malaysia (Yen JM 1968; Barreto and Evans 1994) Class Eurotiomycetes Tehler ex O.E. Eriksson & K. Winka Subclass Chaetothyriomycetidae Doweld Chaetothyriales M.E. Barr Chaetothyriaceae Hansf. ex M.E. Barr 99. Chaetothyrium dominicanum Cif., Sydowia 10(1-6): 133 (1957) [1956] = Sphaerochaetia dominicana (Cif.) Bat. & Cif., Beih. Sydowia 3: 27 (1962) Dominican Republic (Ciferri 1961; Batista and Ciferri 1962; Barreto and Evans 1994), on living leaf of C. odorata (= E. odoratum) Class Leotiomycetes O.E. Erikss. & Winka Helotiales Nannf Ploettnerulaceae Kirschst 100. Pyrenopeziza sp. (= Cylindrosporium sp.) Cambodia (Barreto and Evans 1994) Class Sordariomycetes O.E. Erikss. & Winka Subclass Diaporthomycetidae Senan. et al. Diaporthales Nannf.* Diaporthaceae Höhn. ex Wehm. Fungal Diversity 101. Diaporthe chromolaenae Mapook & K.D. Hyde Thailand (This study) 102. Diaporthe eupatoriicola Petr., Annls mycol. 20(3/4): 147 (1922) Sri Lanka (Barreto and Evans 1994) Diaporthomycetidae families incertae sedis Sporidesmiaceae Fr. 103. Sporidesmium sp. West Indies (Minter et al. 2001) Subclass Hypocreomycetidae O.E. Erikss. & Winka Glomerellales Chadef. ex Réblová et al. Glomerellaceae Locq. ex Seifert & W. Gams 104. Colletotrichum gloeosporioides (Penz.) Penz. & Sacc., Atti Inst. Veneto Sci. lett., ed Arti, Sér. 6 2: 670 (1884) Cuba (Urtiaga 1986; Barreto and Evans 1994), Sri Lanka (Barreto and Evans 1994), Venezuela (Urtiaga 2004) 105. Colletotrichum sp. : Cambodia - (Barreto and Evans 1994), Venezuela - (Urtiaga 2004) Hypocreales Lindau Cordycipitaceae Kreisel ex G.M. Sung et al. 106. Akanthomyces lecanii (Zimm.) Spatafora, Kepler & B. Shrestha, in Kepler et al. 2017 = Verticillium lecanii (Zimm.) Viégas, Revista Inst. Café São Paulo 14: 754 (1939) West Indies (Minter et al. 2001) Hypocreaceae De Not. 107. Acrostalagmus albus Preuss, Linnaea 24: 126 (1851) South America (Viegas 1961) 108. Acrostalagmus aphidum Oudem., Beih. Botan. Centralbl., Abt. B 11: 537 (1902) Puerto Rico, Virgin Islands (Stevenson 1975); West Indies (Minter et al. 2001) 109. Trichoderma guizhouense Q.R. Li, McKenzie & Yong Wang bis, in Li et al., Mycol. Progr. 12(2): 170 (2012) [2013] Thailand (This study) 111. Fusarium fujikuroi Nirenberg, Mitt. biol. BundAnst. Ld- u. Forstw. 169: 32 (1976) = Fusarium moniliforme J. Sheld., Nebraska Agric. Exp. Stat. Rep. 17: 23 (1904) Nigeria (Richardson 1990) 112. Fusarium pallidoroseum (Cooke) Sacc., Syll. fung. (Abellini) 4: 720 (1886) Nigeria (Richardson 1990) 113. Fusarium solani (Mart.) Sacc., Michelia 2(no. 7): 296 (1881) Nigeria (Richardson 1990) Stachybotryaceae Lombard & Crous 114. Memnoniella chromolaenae Mapook & K.D. Hyde Thailand (This study) Subclass Sordariomycetidae O.E. Erikss & Winka (= Meliolomycetidae P.M. Kirk & K.D. Hyde)* Meliolales Gäum. ex D. Hawksw. & O.E. Erikss. Meliolaceae G.W. Martin ex Hansf. 115. Appendiculella sororcula (Speg.) Hansf., Beih. Sydowia 2: 615 (1961) = Irene sororcula (Speg.) F. Stevens, Annls mycol. 25(5/6): 423 (1927) Dominican Republic (Ciferri 1961), Puerto Rico, Venezuela (Hansford 1949), Trinidad and Tobago (Baker and Dale 1951) 116. Meliola sp. India (Barreto and Evans 1994) 117. Ophiociliomyces bauhiniae Bat. & I.H. Lima, Anais Soc. Biol. Pernambuco 13(2): 30 (1955) Brazil (Barreto and Evans 1994), Cambodia - (Litzenberger et al. 1994) Subclass Xylariomycetidae O.E. Erikss & Winka Amphisphaeriales D. Hawksw. & O.E. Erikss.* Apiosporaceae K.D. Hyde et al. 118. Arthrinium chromolaenae Mapook & K.D. Hyde Thailand (This study) Xylariales Nannf. Cainiaceae J.C. Krug Nectriaceae Tul. & C. Tul. 119. Longiappendispora chromolaenae Mapook & K.D. Hyde Thailand (This study) 110. Fusarium culmorum (Wm.G. Sm.) Sacc., Syll. fung. (Abellini) 10: 726 (1892) Nigeria (Richardson 1990) Sordariomycetes genera incertae sedis 13 Fungal Diversity 120. Pleurophragmium capense (Thüm.) S. Hughes, Can. J. Bot. 36: 796 (1958) = Spiropes capensis (Thüm.) M.B. Ellis, Mycol. Pap. 114: 5 (1968) West Indies (Minter et al. 2001) Ascomycota unclassified Ascomycota genera incertae sedis 121. Alysidiopsis yunnanensis Y.L. Guo & X.J. Liu, Acta Mycol. Sin. 11(3): 213 (1992) China (Zhuang 2001), on leaves of C. odorata (= E. odoratum) 122. Hansfordia pulvinata (Berk. & M.A. Curtis) S. Hughes 1958 = Dicyma pulvinata (Berk. & M.A. Curtis) Arx 1981 West Indies (Minter et al. 2001) 123. Redbia trichomambusta R.W. Barreto, in Barreto & Evans, Mycol. Res. 98(10): 1111 (1994) Brazil (Barreto and Evans 1994) 124. Redbia sp. West Indies (Minter et al. 2001) Phylum Basidiomycota R.T. Moore Class Agaricomycetes Doweld Cantharellales Gäum. Ceratobasidiaceae G.W. Martin Colombia (Chardon and Toro 1930; Kern et al. 1933b), Venezuela (Chardon and Toro 1934; Kern et al. 1934) Pucciniales genera incertae sedis 129. Uredo bullula F. Kern, Mycologia 20(2): 77 (1928) Dominican Republic (Kern 1930; Kern et al. 1933a; Ciferri 1961; Barreto and Evans 1994), West Indies (Minter et al. 2001) Pucciniosiraceae Cummins & Y. Hirats. 130. Cionothrix praelonga (G. Winter) Arthur, N. Amer. Fl. (New York) 7(2): 124 (1907) = Cronartium praelongum G. Winter, Hedwigia 26: 24 (1887) Dominican Republic, Mexico, Trinidad and Tobago, Venezuela (Arthur JC 1922; Baker and Dale 1951; Barreto and Evans 1994), Colombia (Buritica and Pardo Cardona 1996; Pardo Cardona 1998), Costa Rica (Arthur 1918a; Berndt 2004) Guatemala (Arthur 1918b), Mexico (Gallegos and Cummins 1981), Panama (Piepenbring 2006), West Indies (Minter et al. 2001) Discussion 125. Rhizoctonia solani J.G. Kühn, Ann. Sper. agr., N.S.: 224 (1858) = Thanatephorus cucumeris (A.B. Frank) Donk, Reinwardtia 3: 376 (1956) Taiwan (Sawada 1931) Class Pucciniomycetes Bauer et al. Pucciniales Clem. & Shear Coleosporiaceae Dietel 126. Coleosporium eupatorii Arthur, Bull. Torrey bot. Club 33: 31 (1906) China (Tai 1979; Barreto and Evans 1994; Zhuang 2005) 127. Coleosporium steviae Arthur, Bot. Gaz. 40: 197 (1905) China (Barreto and Evans 1994), Mexico (Gallegos and Cummins 1981) Cronartiaceae Dietel 128. Cronartium praelongum G. Winter, Hedwigia 26: 24 (1887) 13 Taxonomic and phylogeny of fungi associated with Chromolaena odorata Chromolaena odorata (Siam weed) grows well in sunny, open areas (abandoned fields, disturbed forests, pastures, roadsides), as well as partial shade, but does not grow in shade (Zachariades et al. 2009). In this study, we describe the saprobic fungi associated with C. odorata, which were mainly collected in northern Thailand. Our study included morphological comparison and multigene analyses. A total of 253 fungal specimens were collected. Eighty-eight fungal collections were successfully isolated into culture and identified to 77 taxa distributed in ten orders and 23 families. These included one new family, 11 new genera, 43 new species, 11 new host records, three new combinations, and two reference specimens belonging to Dothideomycetes. Pleosporales was the dominant order with most taxa in the families Didymosphaeriaceae and Phaeosphaeriaceae (Figs. 130, 131, 132). One new genus, four new species, and one new host record of taxa belonging to Sordariomycetes were also identified, with Hypocreales being the dominant order (Figs. 130, 133, 134). The result indicates that Siam weed harbors a high fungal diversity, especially Fungal Diversity Fig. 130 Treatment of fungal taxa from C. odorata Fig. 131 Families in each order of Dothideomycetes associated with C. odorata of Dothideomycetes (Schoch et al. 2009; Hyde et al. 2013; Jayasiri et al. 2019). Numerous studies have reported the capability of Dothideomycetes to survive extreme environmental conditions, such as extreme temperatures, severe droughts, and solar radiation (Ohm et al. 2012; Murgia et al. 2018). Murgia et al. (2018) studied the biodiversity of fungi in hot desert sands and reported that the most abundant fungal Phyla identified in all the samples was Ascomycota and Dothideomycetes were the dominant group. This group therefore plays an important role in the desert ecosystem, as well as severe conditions on rock surfaces (Ruibal et al. 2009). In this study, Dothideomycetes were the dominant group of saprobic fungi on Siam weed in northern Thailand. Collections were mostly from roadsides, abandoned fields, and disturbed forests in sunny and open areas. However, only 35% (88/253) of the saprobic fungi we collected from Siam weed were successfully cultured. Another 65% of taxa contained dormant spores which did not germinate, and some collections were partially or completely discharged with dry or empty ascomata. This may be due to unfavorable nutritional and environmental factors, such as temperature, humidity, as well as UV light, all of which may reduce the capability of spore germination (Gottlieb 1950; Ayerst 1969; Osman and 13 Fungal Diversity Fig. 132 Numbers of taxa in each family of Dothideomycetes associated with C. odorata Fig. 133 Families in each order of Sordariomycetes associated with C. odorata 13 Fungal Diversity Fig. 134 Numbers of taxa in each family of Sordariomycetes associated with C. odorata Valadon 1981; Fourtouni et al. 1998; Mitakakis et al. 2003; Dagno et al. 2011; Wenneker et al. 2013). Forty percent (25/62) of the saprobic taxa identified in this study had spores with appendages or mucilaginous sheaths. Such propagules have also been reported in terrestrial ascomycetes, and frequently occur in aquatic fungi, especially marine species, for spore dispersal and attachment to substrata (Hyde and Jones 1988; Hyde and Goh 2003; Jones 2006; Hyde et al. 2016a; Jones et al. 2019). Moreover, 77.5% (48/62) of the taxa had spore pigmentation, which is divided into 55% (34/62) of dark pigmentation and 22.5% (14/62) with light pigmentation, while 22.5% (14/62) of the spores were hyaline. We assume that appendages, mucilaginous sheaths, and pigmentation of the spores serve to protect the propagules from drought and UV light, as well as other unfavorable conditions. In addition, saprobic fungi associated with C. odorata in our study were mainly sexual morphs, which is related to sexual reproduction. This is likely to result in evolution of species for survival in adverse environmental competition and better competition with other species. Evolution occurs due to genetic complementation provided by the combination of different parents’ genomes that has benefits to repair DNA damage and increase genetic variability with the evolution rate of sexual species (Maynard Smith 1978; Colegrave 2012; Wallen and Perlin 2018). A comparison of fungi on Siam weed and other hosts Of the 62 species identified in this study as saprobes on Siam weed, only 15 are previously known, while the other 47 species are new to science. Of these 15 species, we found that 13 species are previously known from unrelated hosts, which comprised eight known species reported from Thailand, while five known species have been reported from elsewhere, and the other two species (Torula chromolaenae and Torula fici) are previously known from Siam weed (Table 30). The previously known reported species were mainly found in northern Thailand, such as Leptospora thailandica, Muyocopron lithocarpi, Pseudopithomyces palmicola, Pseudoroussoella elaeicola, Torula polyseptata, Neomassarina thailandica and Flabellascoma minimum, while some have been reported in the adjacent area, China (Table 30). Furthermore, we found some known species occurred in different habitats. Rhytidhysteron bruguierae was collected from submerged branches of Bruguiera sp. in a marine habitat and also from terrestrial habitats on a dead stem of Siam weed in this study. The strain from submerged habitats was collected from Phetchaburi Province in western Thailand (Dayarathne et al. 2020), while our strains from terrestrial habitats were collected from several provinces in 13 Fungal Diversity northern Thailand (Chiang Rai, Phrae, Chiang Mai and Mae Hong Son). Why are the fungi on Siam weed and other hosts so different? Some studies have focused on microfungi associated with selected groups of plants in both Thailand and China and from different climates. Doilom et al. (2017) showed that Dothideomycetes were a dominant group on Tectona grandis (teak; Lamiaceae) in northern Thailand with collecting sites in six provinces (Phayao, Chiang Mai, Chiang Rai, Sukhothai, Phrae, Uttaradit), while Dai et al. (2017) found a similar number of Dothideomycetes and Sordariomycetes on bamboo (Poaceae) from northeastern Thailand (Mukdahan province), northern Thailand (Phayao, Chiang Rai, Chiang Mai, Uttaradit and Phitsanulok province), southern Thailand (Phangnga and Krabi province), and southern China. Tibpromma et al. (2018) reported on the fungi on Pandanaceae from northern Thailand (Chiang Rai and Chiang Mai provinces), eastern Thailand (Chonburi Province), western Thailand (Prachuabkhirikhan Province), and southern Thailand (Krabi, Phangnga, Nakhonsithummarat, Phatthalung and Phuket provinces), as well as in southern China. In general, the fungal communities differed, depending on the host, with low overlap. This indicates that if we study other unrelated hosts, we should expect to find a high, unique diversity (Hyde et al. 2018a, b, 2019a, b). In this study, fungi on Siam weed were mainly collected in northern Thailand (Chiang Mai, Chiang Rai, Lampang, Mae Hong Son, Nan, and Phrae provinces), as well as in western Thailand (Phetchaburi Province). We found three overlapping species (Pseudopithomyces palmicola, Torula chromolaenae and T. fici) which were previously reported from Pandanus in Tibpromma et al. (2018); however, we could not find any overlapping species with teak (Doilom et al. 2017) and bamboo (Dai et al. 2017). Although collecting sites are in the same provinces, fungi are likely to differ on the unrelated hosts. In addition, we also compared existing species to microfungi on Tamarix (Tamaricaceae) from Italy and Russia in Thambugala et al. (2017a), on grass (Poaceae) from China, Italy, Russia and Thailand in Thambugala et al. (2017b), as well as Rosaceae, which were collected from Thailand, China, England, India, Italy, Russia, Sweden, and Wales in Wanasinghe et al. (2018); however, there were no overlapping species. This indicates that fungi may be typically different in each country, and each study site, and on each unrelated host. It has been postulated that the differences in the dominant taxa occurring on unrelated plant species may be due to factors, such as moisture content, temperature, pH, nutrients, organic content in each substrate, decay period, fungal adaptation, competitive interactions among the fungal group, as well as plant and soil chemistry, 13 which are likely to influence diversity, species, and the community of fungi (Venugopal et al. 2016, 2017). This was supported by Boonyuen et al. (2014) where high humidity influenced the number of fungal species compared to other sites with low humidity. Therefore, in areas with similar or relatively similar environmental conditions we should expect to find numerous overlapping species. However, this is not the case, as shown with studies on communities on bamboo, Pandanaceae, and teak and thus host dependence is probably causing speciation to be more strongly shaped rather than climate (Ezard et al. 2011). The saprobic fungal communities on bamboo, grass, Pandanaceae, Rosaceae and teak have been shown to have very few overlapping species. Thus herein, we studied the fungi on Siam weed and also found few overlapping species with those on the aforementioned hosts. The reason for the uniqueness of taxa on any host may be due to them having endophytic lifestyles (Promputtha et al. 2007). Endophytes are likely to be a host-specific and close relationship with their particular host (Wong and Hyde 2001; Chauhan et al. 2019). However, most endophytes recovered to date are not unique to their hosts (Guo et al. 2001; Wang et al. 2005; Huang et al. 2009; Sun et al. 2011), but this is probably because techniques isolate only fast growing taxa (Guo et al. 2000, 2001; Promputtha et al. 2007). Life modes of fungi on Siam weed Fungi are able to survive in various substrates in varied habitats (Cai et al. 2006; Kodsueb et al. 2016), and they can also change their lifestyle. Some endophytic fungi have parasitic or pathogenic modes and this is thought to occur when environmental and physiological conditions become suitable, such as the host becoming stressed (Strobel 2018). Endophytes may also be mutualistic and protect the host plant from diseases or herbivores (Carroll 1988; Fróhlich et al. 2000; Chaverri and Samuels 2013). Several studies have postulated that endophytes which live within plants are the initial colonizers of dead leaves and twigs from the plant and they can be saprotrophic when the host senescence or when the plant part dies (Promputtha et al. 2007, 2010; Purahong and Hyde 2011; Sun et al. 2011; Jeewon et al. 2013). Some of the previously known species were found to have saprotrophic lifestyles in our study; however, they have also been reported to occupy other life modes in plants (Table 30). For example, three fungal species found as saprobes on Siam weed, have been reported as pathogenic on various hosts. For example, Acrocalymma medicaginis causes crown rot disease of Medicago sativa (Fabaceae) in Australia (Alcorn and Irwin 1987; Trakunyingcharoen et al. 2014), Aplosporella hesperidica causes early stem-end rot of Citrus sinensis (Rutaceae) in Zimbabwe (Yang et al. 2017), and Pseudopithomyces palmicola causes chlorotic and/or Fungal Diversity Table 30 Species found on Siam weed with their previous known host Species Host Host family Origin Lifestyle Reference Acrocalymma medicaginis Crown rot of Medicago sativa Fabaceae Australia Pathogen Asteraceae Chiang Rai; Thailand Saprobe (Alcorn and Irwin 1987), (Trakunyingcharoen et al. 2014) This study Rutaceae Rutaceae India Zimbabwe Saprobe Pathogen Rao (1969) Yang et al. (2017) Asteraceae Chiang Rai; Thailand Saprobe This study Fabaceae Asteraceae Namibia; South Africa Chiang Rai; Thailand Uncertain Saprobe Slippers et al. (2014) This study Arecaceae Taiwan Saprobe Hashimoto et al. (2018) Fabaceae Taiwan Saprobe Hashimoto et al. (2018) Fabaceae Lampang; Thailand Saprobe Jayasiri et al. (2019) Asteraceae Nan; Thailand Saprobe This study Verbenaceae Chiang Rai; Thailand Saprobe Hyde et al. (2016a, b) Asteraceae Chiang Rai; Thailand Saprobe This study Fagaceae Chiang Rai; Thailand Saprobe Fabaceae Chiang Rai; Thailand Saprobe Mapook et al. (2016a, b, c) Jayasiri et al. (2019) Fabaceae Guizhou; China Saprobe Jayasiri et al. (2019) Unknown Yunnan; China Saprobe Phookamsak et al. (2019) Asteraceae Saprobe This study Asparagaceae Chiang Rai, Lampang; Thailand Chiang Mai; Thailand Saprobe Hyde et al. (2016a, b) Asteraceae Chiang Rai; Thailand Saprobe This study Unknown China Saprobe Zhang et al. (2019) Asteraceae Saprobe This study Vitaceae Chiang Rai, Chiang Mai, Mae Hong Son; Thailand Italy Fabaceae Unknown Weak Pathogen Liu et al. (2018) Poaceae Unknown Weak Pathogen Liu et al. (2018) Rosaceae Unknown Weak Pathogen Liu et al. (2018) Dead stems of Chromolaena odorata Aplosporella hesperidica Citrus aurantium Early stem-end rot of Citrus sinensis Dead stems of C. odorata Dothiorella oblonga Acacia mellifera Dead stems of C. odorata Flabellascoma minimum On petioles of Arenga engleri On pods of Bauhinia purpurea On pods of Leucaena leucocephala Dead stems of C. odorata Leptospora thailandica On dead branches of Duranta sp. Dead stems of C. odorata Muyocopron lithocarpi On dead leaves of Lithocarpus lucidus On decaying pods of Peltophorum sp. On fallen pod of Cercis chinensis On dead stems of herbaceous plant Dead stems of C. odorata Neomassarina thailanOn dead bract-like leaves dica from flower stalk of Agave angustifolia Dead stems of C. odorata Pseudoophiosphaerella On dead culms of unihuishuiensis dentified herbaceous plant Dead stems of C. odorata Pseudopithomyces palmicola On leaves lesions of Vitis vinifera On leaves lesions of Phaseolus vulgaris On leaves lesions of Poa annua On leaves lesions of Fragaria sp. Weak Pathogen Liu et al. (2018) 13 Fungal Diversity Table 30 (continued) Species Pseudoroussoella elaeicola Rhytidhysteron bruguierae Torula chromolaenae Torula fici Torula polyseptata Trichoderma guizhouense Host Host family Origin Lifestyle Reference On dead leaves of Pandanus amaryllifolius On dead leaves of unidentified grass species On leaves of Acoelorrhaphe wrightii Dead stems of C. odorata On dead petiole of Elaeis guineensis Dead stems of C. odorata Submerged branches of Bruguiera sp. Dead stems of C. odorata Pandanaceae Chiang Rai; Thailand Saprobe Tibpromma et al. (2018) Unknown Yunnan; China Saprobe Hyde et al. (2017b) Arecaceae Chiang Rai; Thailand Saprobe Ariyawansa et al. (2015) Asteraceae This study Arecaceae Chiang Rai, Chiang Mai; Saprobe Thailand Chiang Rai; Thailand Saprobe Asteraceae Lampang; Thailand Saprobe This study Rhizophoraceae Phetchaburi; Thailand Saprobe Dayarathne et al. (2020) Asteraceae This study Dead stem of Chromolaena odorata Dead leaf of Pandanus tectorius Dead stems of C. odorata On dead stems of Chromolaena odorata On decaying cone of Magnolia grandiflora On decaying fruit pericarp of Garcinia sp. On dead leaf of Pandanus sp. On submerged decaying wood Ficus religiosa Dead stems of C. odorata On submerged decaying wood Dead stems of C. odorata Soil Asteraceae Chiang Rai, Chiang Mai, Saprobe Mae Hong Son, Phrae; Thailand Chiang Mai; Thailand Saprobe Pandanaceae Yunnan; China Tibpromma et al. (2018) Asteraceae This study Asteraceae Chiang Rai, Chiang Mai; Saprobe Thailand Chiang Rai; Thailand Saprobe Magnoliaceae Yunnan; China Saprobe Jayasiri et al. (2019) Clusiaceae Ranong; Thailand Saprobe Jayasiri et al. (2019) Pandanaceae Chiang Mai; Thailand Saprobe Tibpromma et al. (2018) Unknown China Saprobe Su et al. (2018) Moraceae Asteraceae Cuba Chiang Rai; Thailand Unknown Saprobe Crous et al. (2015a) This study Unknown Chiang Rai; Thailand Saprobe Hyde et al. (2019a) Asteraceae Chiang Rai; Thailand Saprobe This study – China Saprobe Li et al. (2013) Endophyte in stems of Ancistrocladus korupensis Endophyte in stems of Cola spp. Dead stems of C. odorata Ancistrocladaceae Cameroon Endophyte Chaverri et al. (2015) Malvaceae Cameroon Endophyte Chaverri et al. (2015) Asteraceae Phrae; Thailand Saprobe This study necrotic symptoms on leaves of Vitis vinifera (Vitaceae), Phaseolus vulgaris (Fabaceae), Fragaria sp. (Rosaceae) and Poa annua (Poaceae) in Italy (Liu et al. 2018). In addition, Trichoderma guizhouense was initially found in soil polluted 13 Saprobe Phookamsak et al. (2019) Li et al. (2017) Li et al. (2017) by heavy metals in China, and has also been reported as endophytic in stems of Ancistrocladus korupensis (Ancistrocladaceae) and Cola spp. (Malvaceae) in Cameroon (Li et al. 2013; Chaverri et al. 2015), while in this study it was Fungal Diversity reported as a saprobic on Siam weed. This indicates that some of the fungi on Siam weed are likely to have the ability to be endophytes or plant pathogen in other hosts, while most of the previously known species had saprobic lifestyles on Siam weed. Where did the fungi on Siam weed come from? Co-introduction and host-jumping are hypotheses generally used to explain the fungi occurring in invasive plants (Shipunov et al. 2008; Burgess et al. 2016; Truter et al. 2017). In co-introduction, the plant colonizes a new invasive range accompanied by associated microorganisms from the native range. In the case of fungi, however, this could only occur if the new invasive range was adjacent to native range and spores could be wind-dispersed or the fungi are seed-borne. In host-jumping, the invasive plants colonizes a new invasive range without any associated fungi from the native range. It then obtains newly associated fungi from surrounding plants in the invasive range. In the case of Siam weed, whose origins are the Americas, it is likely that cointroduction also occurred, as the weed was dispersed by seed (Scott et al. 1998; Setter and Campbell 2002; Zachariades et al. 2009). Esuruoso (1971) reported seed-borne fungi from Siam weed, were found as important pathogens of the various food crops in Nigeria. However, there are no further studies on seed-borne fungi of Siam weed and the distances involved should prohibit dispersal by wind. Therefore, we suggest that Siam weed gained its associated fungi via hostjumping rather than co-introduction. A few associated fungi may have been dispersed by seeds when it colonized a new area (co-introduction hypothesis), but probably most newly associated fungi were acquired from the surrounding hosts (host-jumping hypothesis). In this study, the previously known species were mostly reported from other hosts in Thailand and elsewhere, while there is no report in closely related plants, such as other Asteraceae. Although we have little evidence, we suggest that the saprobic fungi found on Siam weed in this study, are likely to have jumped host from unrelated adjacent plants, rather than arrived by co-introduction. Alternatively, they may occur on other Asteraceae hosts in Thailand, but there have been very few studies of the fungal diversity of this host. Did the fungi on Siam weed jump hosts? The term host-jumping to unrelated or distantly hosts was discussed in Roy (2001) and seems to have occurred here. There is some evidence that the fungi on Siam weed jumped from other hosts. Muyocopron lithocarpi was previously found in Thailand from the host family Fabaceae on Lithocarpus lucidus and Peltophorum sp. (Mapook et al. 2016b, Jayasiri et al. 2019) and Cercis chinensis in China, while it was found on Siam weed in this study. It is likely that Muyocopron lithocarpi jumped from Fabaceae or some other host to Siam weed (Asteraceae). Muyocopron dipterocarpi was introduced as saprobe from dried twigs of Dipterocarpus tuberculatus (Dipterocarpaceae) in northern Thailand, and did not produce specialized infection structures (Mapook et al. 2016b). The taxon was subsequently found on a dried twig of a rubber tree, Hevea brasiliensis (Euphorbiaceae). The species produces appressoria from germinating ascospores, indicating an endophytic lifestyle (Konta et al. 2016; Senwanna et al. 2019). This species, is therefore, likely to have jumped host from some other local host, as both Siam weed and rubber are not native plants (Senwanna et al. 2019). Muyocopron heveae was also collected from rubber trees, and produced hyaline appressoria (Senwanna et al. 2019). This suggests that Muyocopron species are likely to have the ability to be endophytic or pathogenic and have jumped to unrelated hosts. Most pathogens (e.g. Colletotrichum species) produce appressoria, so that they can invade the host and some species are thought to be host-specific (Sutton 1968; Zhou and Hyde 2001; Nesher et al. 2008; Yan et al. 2018). Many saprobes also produce appressoria (Konta et al. 2016; Phukhamsakda et al. 2016; Senwanna et al. 2019; Hyde et al. 2020), so that they can also infect the host. Like pathogens, the endophytes will have built up a relationship with the host over an evolutionary timescale so that they can infect the host without being killed and also without causing symptoms. For example, the palm and Pandanaceae genera Linocarpon and Oxydothis, the majority of which are saprobes, have been shown to produce appressoria, indicating that they colonize living plants (Konta et al. 2016). The saprobic species, Sparticola junci, grows on Spartium junceum (Fabaceae) and also produce appressoria, which suggests that also have endophytic lifestyles (Phukhamsakda et al. 2016). Fungi have the ability to exhibit more than one lifestyle in a different host. Although they are the same species, the ability to produce appressoria is important for living host colonization. Appressoria formation is related to surface signals with cell division and cell wall modifications in the early germination stages of pathogenic conditions, followed by cAMP with other additional pathways, while the appressoria formation is not needed for a saprobic lifestyle and its germination pathways is a reversed process (Barhoom and Sharon 2004). There was no evidence for appressoria formation in this study, as all taxa germinated without forming specialized infection structures. Most of the taxa on Siam weed are therefore likely to have jumped host from other plants and are unlikely to be a specialist to this weed. 13 Fungal Diversity Evolution of Asteraceae Asteraceae, a large and widespread flowering plant (Angiospermae) family which accommodates Chromolaena odorata (Jeffrey 2007) evolved in the late Cretaceous at 85.9 Mya (82.3-91.5) with a stem age at 88–89 Mya. This dating is based on the oldest fossils of pollen grains (Barreda et al. 2015). According to the study of Barreda et al. (2015), the majority of the subfamilies in the Asteraceae clade diverged during the warmest Earth’s temperature, which started during the late Paleocene to early Eocene Epoch in the Cenozoic Era. Climatic change may have been associated with the diversification of Asteraceae and some Asteraceae members probably survived across the Cretaceous-Tertiary (K-T) mass extinction event (approximately 66 Mya), due to their adaptation and tolerance to extreme environmental, as well as ecological conditions (McFadyen 1996; Chauhan and Johnson 2008; Koutika and Rainey 2010; Barreda et al. 2015). For example, Siam weed has been reported as adapted for acidic soils (Koutika and Rainey 2010) and can tolerate harmful heavy metals, such as cadmium and zinc (Agunbiade and Fawale 2009; Okoronkwo et al. 2014; Ruangdech et al. 2017), as well as improving soil contaminated by used oil (Anolifo and Vwioko 2001; Atagana 2011; Ikhajiagbe and Akindolor 2016), which may also affect the microbial community, especially soil microorganisms. Does evolution drive the diversity of saprobic fungi associated with Chromolaena odorata? The divergence time estimates of each fungal family identified in this study, were obtained from previous studies (Mapook et al. 2016c; Hyde et al. 2017a; Hongsanan et al. 2017; Liu et al. 2017; Guterres et al. 2018; Soleimani et al. 2018; Phillips et al. 2019; Samarakoon et al. 2019). The divergence times for 19 of the 23 families are recorded in Tables 31 and 32. Divergence time estimates for stem ages show that one family diverged in the Carboniferous (5%), one in the Triassic (5%), two in the Jurassic (10%), and 15 (79%) in the Cretaceous Period (Fig. 135), while the results of divergence time estimates for crown age show that the fungal family associated with C. odorata in this study mostly diverged in the Paleogene period (Fig. 136). Although most fungal families coincided with the expected origin of the host family (Asteraceae), all species identified are probably not specific to Siam weed. Perhaps these taxa evolved with other Asteraceae present in Thailand. Four fungal families found on Siam weed, e.g. Pleurotremataceae (174 Mya) and Muyocopronaceae (172 Mya), evolved earlier than Asteraceae in the Jurassic, Hysteriaceae (219 Mya) evolved in the Triassic, and Patellariaceae (311 Mya) evolved in the Carboniferous (Table 31). Asteraceae evolved 85.9 Mya in the late Cretaceous. Fungi that evolved 13 before the Asteraceae may have jumped from hosts or have evolved with the ancestors of the family. The fungal families that evolved earlier than Asteraceae produce ascomata with heavy pigmentation in the thick peridium wall, which may protect them from drought and UV light and appear to have wide host ranges (Gueidan et al. 2015; Ekanayaka et al. 2017; Hernández-Restrepo et al. 2019; Jayasiri et al. 2019). Antimicrobial activity of fungi on Siam weed In this study, we provide evidence of the antimicrobial activity of fungi associated with Chromolaena odorata (Table 33). Forty (65%) out of 62 species showed potential antimicrobial activity, with 12 species showing antimicrobial activity against Mucor plumbeus (filamentous fungus), 12 species against Escherichia coli (Gram-negative bacterium), eight species against both M. plumbeus and Bacillus subtilis (Gram-positive bacterium), three species against bacteria and filamentous fungus, two species against only B. subtilis, as well as two species against both E. coli and M. plumbeus, and one species against both B. subtilis and E. coli. In contrast, 18 (29%) of the 62 species did not show antimicrobial activity and four species were not tested (Figs. 137, 138). This infers that more than half of the fungal species associated with C. odorata have antimicrobial potential and could be promising sources for the discovery of novel bioactive compounds. Among the previously known species in this study, 12 of the 15 lack previously reported antimicrobial activity. For example, our study showed for the first time that Pseudopithomyces palmicola, Neomassarina thailandica, Torula fici, T. polyseptata and Rhytidhysteron bruguierae showed antimicrobial activity against M. plumbeus and Leptospora thailandica, Pseudoophiosphaerella huishuiensis and Torula chromolaenae showed antimicrobial activity against all tested organisms. Furthermore, Acrocalymma medicaginis and Flabellascoma minimum showed antimicrobial activity against B. subtilis and M. plumbeus. Aplosporella hesperidica and Pseudoroussoella elaeicola showed antimicrobial activity against E. coli, while Dothiorella oblonga and Muyocopron lithocarpi showed no antimicrobial activity. Trichoderma guizhouense, which had previously been reported with antifungal activity against Fusarium oxysporum f. sp. cubense causing banana wilt disease (Zhang et al. 2016), also showed antimicrobial activity against E. coli and M. plumbeus in this study. However, research on novel bioactive compounds and other chemical discoveries from these new fungal taxa is in its infancy. Fungal Diversity Table 31 Divergence time estimates of the Dothideomycetes families recorded in this study Subclass Order Pleosporomycetidae Hysteriales Dothideomycetes orders incertae sedis Family Divergence times (crown age) Hysteriaceae 149 Mya (90–213), late Jurassic Pleosporales Acrocalymmaceae 23 Mya (8–44), the Neogene period Didymellaceae 63 Mya (35–97), Paleogene period Didymosphaeriaceae 72 Mya (48–101), Cretaceous Lophiostomataceae – Nigrogranaceae 72 Mya (44–124), Cretaceous Neopyrenochaetaceae – Phaeosphaeriaceae 75 Mya (46–102), Cretaceous Pyrenochaetopsidaceae – Roussoellaceae 62 Mya (34–91), Cenozoic Era (Paleogene period) Thyridariaceae 15 Mya (3–38), Cenozoic Era (Neogene period) Torulaceae 15 Mya (4–34), Cenozoic Era (Neogene period) Neomassarinaceae – Botryosphaeriales Aplosporellaceae 40 Mya, Cenozoic Era, (Paleogene period) Botryosphaeriaceae Dyfrolomycetales Pleurotremataceae Muyocopronales Muyocopronaceae Patellariales Patellariaceae 61 Mya, Cenozoic Era, (Paleogene period) 76 Mya (38–125), late Cretaceous 52 Mya (38–66), Cenozoic Era, (Paleogene period) 164 Mya (72–283), Jurassic Divergence times (stem age) Reference 219 Mya (161–282), late Triassic 114 Mya (71–156), Cretaceous 115 Mya (84–149), Cretaceous 109 Mya (83–139), Cretaceous – 131 Mya (86–180), Cretaceous – 99 Mya (73–129), Cretaceous – 77 Mya (44–110), late Cretaceous Liu et al. (2017) 95 Mya (60–138), Cretaceous 140 Mya (95–188), early Cretaceous – 94 Mya, Cretaceous Liu et al. (2017) – Phillips et al. (2019) 94 Mya, Cretaceous Phillips et al. (2019) 174 Mya (113–243), Jurassic 172 Mya (130–218), Jurassic Lui et al. (2017) 311 Mya (244–407), Carboniferous period Liu et al. (2017) Liu et al. (2017) Liu et al. (2017) Liu et al. (2017) – Liu et al. (2017) – Liu et al. (2017) – Liu et al. (2017) Liu et al. (2017) Mapook et al. (2016c) Table 32 Divergence time estimates of the Sordariomycetes families recorded in this study Subclass Order Diaporthomycetidae Diaporthales Family Diaporthaceae Divergence times (crown age) 61.15 Mya (29.7–89.9), Cenozoic Era, (Paleogene period) Hypocreomycetidae Hypocreales Hypocreaceae Cenozoic Era, (Paleogene period) Stachybotryaceae Cenozoic Era, (Paleogene period) Xylariomycetidae Amphisphaeriales Apiosporaceae Cenozoic Era, (Paleogene period) Xylariales Cainiaceae Cenozoic Era, (Paleogene period) Divergence times (stem age) Reference 87.66 Mya, Cretaceous Guterres et al. (2018) Cretaceous Hyde et al. (2017a) 88 Mya (50–130), Cretaceous 69 Mya (50–130), late Cretaceous 128 Mya, Cretaceous Hyde et al. (2017a) Hyde et al. (2017a) Hyde et al. (2017a) 13 Fungal Diversity Fig. 135 Fungal families recorded on Siam weed with stem age divergence estimates Checklist of fungi associated with Chromolaena odorata There are 68 fungal species previously reported from Siam weed, based on the USDA Systematic Mycology and Microbiology Laboratory (SMML) database (Farr and Rossman 2020), together with relevant literature. These 68 species were recorded from Asia (45), Americas (37), West Indies (36), Africa (7), and Oceania (1). Several taxa are known from more than one location, while only seven species were recorded from Thailand (Cercospora sp., Hermatomyces chromolaenae, Neocochlearomyces chromolaenae, Pseudocercospora eupatorii, P. eupatorii-formosani, Torula chromolaenae, and T. fici). Fig. 136 Fungal families recorded on Siam weed with crown age divergence estimates 13 The 130 taxa recorded from C. odorata, including the 62 species listed in this study are distributed in 20 orders, 48 families and 85 genera and are divided among (i) Ascomycota: four classes, 18 orders, 43 families, 80 genera and 110 species identified with 13 unidentified species (Fig. 139), and (ii) Basidiomycota: two classes, two orders, five families, five genera, with six species identified (Fig. 140). Siam weed harbors a higher diversity, especially of Dothideomycetes with 75% of fungal taxa (98/130) from 30 families with Pleosporales being the dominant order (Figs. 139, 141) and 15% (20/130) of fungal taxa are belonging to Sordariomycetes with Hypocreales is dominant orders (Fig. 142). The most commonly reported genus on Siam weed is Pseudocercospora with six known species including Pseudocercospora aciculina, P. ageratoides, P. convoluta, P. Fungal Diversity Table 33 Preliminary screening of Siam weed fungi for antimicrobial activity against Bacillus subtilis, Escherichia coli and Mucor plumbeus with diameter of inhibition zone Species name B. subtilis (mm) E. coli (mm) M. plumbeus (mm) Acrocalymma medicaginis Aplosporella chromolaenae Aplosporella hesperidica Arthrinium chromolaenae Chromolaenicola chiangraiensis Chromolaenicola lampangensis Chromolaenicola nanensis Chromolaenicola thailandensis Chromolaenomyces appendiculatus Diaporthe chromolaenae Didymella chromolaenae Dothiorella oblonga Dyfrolomyces chromolaenae Flabellascoma minimum Leptospora chromolaenae Leptospora phraeana Leptospora thailandica Longiappendispora chromolaenae Memnoniella chromolaenae Montagnula chiangraiensis Montagnula chromolaenae Montagnula chromolaenicola Montagnula thailandica Murichromolaenicola chiangraiensis Murichromolaenicola chromolaenae Muyocopron chromolaenae Muyocopron chromolaenicola Muyocopron lithocarpi Neomassarina chromolaenae Neomassarina thailandica Neoophiobolus chromolaenae Neopyrenochaeta chiangraiensis Neopyrenochaeta chromolaenae Neopyrenochaeta thailandica Neopyrenochaeta triseptatispora Nigrograna chromolaenae Nothophoma chromolaenae Paraleptospora chromolaenae Paraleptospora chromolaenicola Patellaria chromolaenae Pseudocapulatispora longiappendiculata Pseudoophiosphaerella huishuiensis Pseudopithomyces palmicola Pseudoroussoella chromolaenae Pseudoroussoella elaeicola Pseudostaurosphaeria chromolaenae Pseudostaurosphaeria chromolaenicola Pseudothyridariella chromolaenae Pyrenochaetopsis chromolaenae 12 – – – No inhibition – – No inhibition 20 – No inhibition No inhibition 12 15 13 – 13 – – 8 – No inhibition No inhibition – – No inhibition No inhibition No inhibition – – No inhibition – 13 No inhibition 16 – No inhibition – – – 7 10 – No inhibition – 10 – – – – 9 9 8 13 – – – – – 14 12 – 13 – – – – 8 9 8 – – – 17 20 22 24 – – 11 18 17 – 11 – – 30 10 – – – 24 11 8 12 13 – 8 – – – – 23 10 10 10 – – – 9 – 25 18 23 – 13 Fungal Diversity Table 33 (continued) Fig. 137 Fungal species on Siam weed with potential for antimicrobial activity Fig. 138 Numbers of fungal species on Siam weed with potential antimicrobial activity against: G + Gram-positive bacteria, G − Gram-negative bacteria, F filamentous fungus, A active with all microbial 13 Species name B. subtilis (mm) E. coli (mm) M. plumbeus (mm) Rhytidhysteron bruguierae Rhytidhysteron chromolaenae Setoarthopyrenia chromolaenae Sphaeropsis chromolaenicola Torula chromolaenae Torula fici Torula polyseptata Tremateia chiangraiensis Tremateia chromolaenae Tremateia thailandensis Trichoderma guizhouense Xenoroussoella triseptata Yunnanensis chromolaenae – – 16 No inhibition 9 – – No inhibition No inhibition No inhibition – No inhibition – – – – 18.5 20 – 8 – – 18 13 12 19 18 8 35 Fungal Diversity Fig. 139 Checklist of fungi associated with C. odorata in each class of Ascomycota Fig. 140 Checklist of fungi associated with C. odorata in each class of Basidiomycota 13 Fungal Diversity Fig. 141 Fungal taxa associated with C. odorata in each order of Dothideomycetes Fig. 142 Fungal taxa associated with C. odorata in each order of Sordariomycetes eupatoriella, P. eupatorii, P. eupatorii-formosani, and Pseudocercospora sp. The genus is generally found as plant pathogens from leaf lesions, and belongs to the family Mycosphaerellaceae (Den Breeyen et al. 2006). Most previous studies on Siam weed focused on fungal pathogens from symptomatic leaves, to assess the potential for biological control (Barreto and Evans 1994), while our study focused on saprobic fungi on dead stems. Thus, we only found two overlapping species, Torula chromolaenae, and T. fici, which have been previously recorded from dead stems of Siam weed in Thailand as saprotrophs. The use of culture-independent approaches based on Next Generation Sequencing 13 techniques may further increase understanding of the Siam weed mycobiome and community structure, together with corresponding environmental factors and functional groups. Acknowledgements The Sanger sequencing cost was funded by personal research budgets of W. Purahong and T. Wubet from the UFZHelmholtz Centre for Environmental Research. Ausana Mapook was financially supported by Research and Researchers for Industry Program (RRI) PHD57I0012, Thailand and the German Academic Exchange Service (DAAD) for a joint TRF-DAAD (PPP 2017–2018) academic exchange grant to Kevin D. Hyde and Marc Stadler. Kevin D. Hyde thanks to the 2019 high-end foreign expert introduction plan to Kunming Institute of Botany (granted by the Ministry of Science and Technology of the People’s Republic of China, Grant Number Fungal Diversity G20190139006), Thailand Research grants entitled Biodiversity, phylogeny and role of fungal endophytes on above parts of Rhizophora apiculata and Nypa fruticans (Grant No: RSA5980068), the future of specialist fungi in a changing climate: baseline data for generalist and specialist fungi associated with ants, Rhododendron species and Dracaena species (Grant No: DBG6080013), Impact of climate change on fungal diversity and biogeography in the Greater Mekong Subregion (Grant No: RDG6130001). Ausana Mapook also thanks to Shaun Pennycook, Katalee Jariyavidyanont, Dhanushka N. Wanasinghe, Ishani D. Goonasekara, Chayanard Phukhamsakda, Saowaluck Tibpromma, Sirinapa Konta and Chanokned Senwanna for their valuable suggestions and help. References Aba PE, Joshua PE, Ezeonuogu FC et al (2015) Possible anti-diarrhoeal potential of ethanol leaf extract of Chromolaena odorata in castor oil-induced rats. J Complement Integr Med 12:301–306 Abbasi AM, Khan MA, Ahmad M et al (2010) Ethnobotanical study of wound healing herbs among the tribal communities in Northern Himalaya Ranges district Abbottabad, Pakistan. Pak J Bot 42:3747–3753 Abdel-Wahab F (2017) Natural products of Nothophoma multilocularis sp. nov. an endophyte of the medicinal plant Rhazya stricta. Mycosphere 8:1185–1200 Adedapo AA, Oyagbemi AA, Fagbohun OA et al (2016) Evaluation of the anticancer properties of the methanol leaf extract of Chromolaena odorata on HT29 lung cancer cell line. FASEB J 30:1193.6 Agunbiade FO, Fawale AT (2009) Use of Siam weed biomarker in assessing heavy metal contaminations in traffic and solid waste polluted areas. Int J Environ Sci Technol 6:267–276 Ahmadpour SA, Farokhinejad R, Mehrabi-Koushki M (2017) Further characterization and pathogenicity of Didymella microchlamydospora causing stem necrosis of Morus nigra in Iran. Mycosphere 8:835–852 Ahmed SA, van de Sande WWJ, Stevens DA et al (2014) Revision of agents of black-grain eumycetoma in the order Pleosporales. Persoonia 33:141–154 Ahmed SA, González GM, Tirado-Sánchez A et al (2018) Nigrograna mackinnonii, Not Trematosphaeria grisea (syn., Madurella grisea), is the main agent of black grain eumycetoma in Latin America. J Clin Microbiol 56:e01723 Akinmoladun AC, Ibukun EO, Dan Ologe IA (2007) Phytochemical constituents and antioxidant properties of extracts from the leaves of Chromolaena odorata. Sci Res Essay 2:191–194 Alcorn JL, Irwin JAG (1987) Acrocalymma medicaginis gen. et sp. nov. causing root and crown rot of Medicago sativa in Australia. Trans Br Mycol Soc 88:163–167 Amazu LU, Omoregie P, Ajuwo AO, Ifezulike CC (2013) Anticonvulsant potency of the leaf extract of Chromolaena odorata in rats. Unique Res J Med Med Sci 1:64–69 Anolifo GO, Vwioko DE (2001) Tolerance of Chromolaena odorata (L) K. and R. grown in soil contaminated with spent lubricating oil. J Trop Biosci 1:20–24 Aptroot A (2004) Two new ascomycetes with long gelatinous appendages collected from monocots in the tropics. Stud Mycol 50:307–311 Ariyawansa HA, Kang J-C, Alias SA et al (2013) Towards a natural classification of Dothideomycetes: the genera Dermatodothella, Dothideopsella, Grandigallia, Hysteropeltella and Gloeodiscus (Dothideomycetes incertae sedis). Phytotaxa 147:35–47 Ariyawansa HA, Hawksworth DL, Hyde KD et al (2014a) Epitypification and neotypification: guidelines with appropriate and inappropriate examples. Fungal Divers 69:57–91 Ariyawansa HA, Tanaka K, Thambugala KM et al (2014b) A molecular phylogenetic reappraisal of the Didymosphaeriaceae (= Montagnulaceae). Fungal Divers 68:69–104 Ariyawansa HA, Hyde KD, Jayasiri SC et al (2015) Fungal diversity notes 111–252—taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 75:27–274 Arthur JC (1918a) Uredinales of Costa Rica based on collections by E.W.D., Holway. Mycologia 10:111–154 Arthur JC (1918b) Uredinales of Guatemala based on collections by E.W.D. Holway. I. Introduction, Coleosporiaceae and Uredinaceae. Am J Bot 5:325–336 Arthur JC (1922) Uredinales collected by R. Thaxter and J.B. Rorer in Trinidad. Bot Gaz 73:58–69 Asomugha RH, Okafor PN, Ijeh II et al (2014) Hepatic effects of aqueous extract of Chromolaena odorata in male Wistar albino rats. Pharmacol Online 1:127–136 Atagana HI (2011) The potential of Chromolaena odorata (L.) to decontaminate used engine oil impacted soil under greenhouse conditions. Int J Phytoremediation 13:627–641 Atindehou M, Lagnika L, Guérold B et al (2013) Isolation and identification of two antibacterial agents from Chromolaena odorata L. activity against four diarrheal strains. Adv Microbiol 3:115–121 Aveskamp MM, de Gruyter J, Woudenberg JHC et al (2010) Highlights of the Didymellaceae: a polyphasic approach to characterise Phoma and related pleosporalean genera. Stud Mycol 65:1–60 Ayerst G (1969) The effects of moisture and temperature on growth and spore germination in some fungi. J Stored Prod Res 5:127–141 Ayyanar M, Ignacimuthu S (2009) Herbal medicines for wound healing among tribal people in Southern India: ethnobotanical and scientific evidences. Int J Appl Res Nat Prod 2:29–42 Baker RED, Dale WT (1951) Fungi of Trinidad and Tobago. Mycol Pap 33:1–123 Balouiri M, Sadiki M, Ibnsouda SK (2016) Methods for in vitro evaluating antimicrobial activity: a review. J Pharm Anal 6:71–79 Barhoom S, Sharon A (2004) cAMP regulation of “pathogenic” and “saprophytic” fungal spore germination. Fungal Genet Biol FG B 41:317–326 Barku VY, Boahen YO, Dali GA (2014) Ethnobotanical study of wound healing plants in Kpando traditional area, Ghana. Int J Phytomed 6:564–572 Barr ME (1979) A classification of Loculoascomycetes. Mycologia 71:935–957 Barr ME (1990) Prodromus to nonlichenized, pyrenomycetous members of class Hymenoascomycetes. Mycotaxon 39:98–100 Barreda VD, Palazzesi L, Tellería MC et al (2015) Early evolution of the angiosperm clade Asteraceae in the Cretaceous of Antarctica. Proc Natl Acad Sci USA 112:10989–10994 Barreto RW, Evans HC (1994) The mycobiota of the weed Chromolaena odorata in southern Brazil with particular reference to fungal pathogens for biological control. Mycol Res 98:1107–1116 Bassett IE, Beggs JR, Paynter Q (2010) Decomposition dynamics of invasive alligator weed compared with native sedges in a Northland lake. New Zeal J Ecol 34:324–331 Batista AC, Ciferri R (1962) The Chaetothyriales. Beih. Sydowia 3:1–129 Berlese AN (1896) Icones fungorum. Pyrenomycetes 2:1–216 Berndt R (2004) A checklist of Costa Rican rust fungi. In Agerer R, Piepenbring M, Blanz P (eds) Frontiers in basidiomycote mycology, pp 185–236 Bissett J, Gams W, Jaklitsch W, Samuels GJ (2015) Accepted Trichoderma names in the year 2015. IMA Fungus 6:263–295 Boehm EWA, Mugambi GK, Miller AN et al (2009a) A molecular phylogenetic reappraisal of the Hysteriaceae, Mytilinidiaceae and Gloniaceae (Pleosporomycetidae, Dothideomycetes) with keys to world species. Stud Mycol 64:49–83 13 Fungal Diversity Boehm EWA, Schoch CL, Spatafora JW (2009b) On the evolution of the Hysteriaceae and Mytilinidiaceae (Pleosporomycetidae, Dothideomycetes, Ascomycota) using four nuclear genes. Mycol Res 113:461–479 Boehm EWA, Marson G, Mathiassen GH et al (2015) An overview of the genus Glyphium and its phylogenetic placement in Patellariales. Mycologia 107:607–618 Boerema GH, de Gruyer J, Noordeloos ME et al (2004) Phoma identification manual. Differentiation of specific and infra-specific taxa in culture. CABI Publishing, Wallingford Boonyuen N, Manoch L, Chamswarng C et al (2014) Fungal occurrence on sugarcane filter cake and bagasse isolated from sugar refineries in Thailand. Thai J Agric Sci 47:77–86 Boudjeko T, Megnekou R, Woguia AL et al (2015) Antioxidant and immunomodulatory properties of polysaccharides from Allanblackia floribunda Oliv stem bark and Chromolaena odorata (L.) King and H.E. Robins leaves. BMC Res Notes 8:759 Braun U, Sivapalan A (1999) Cercosporoid hyphomycetes from Brunei. Fungal Diver 3:1–27 Březina S, Koubek T, Münzbergová Z, Herben T (2006) Ecological benefits of integration of Calamagrostis epigejos ramets under field conditions. Flora Morphol Distrib Funct Ecol Plants 201:461–467 Burgess TI, Crous CJ, Slippers B et al (2016) Tree invasions and biosecurity: eco-evolutionary dynamics of hitchhiking fungi. AoB Plants 8:plw076 Buritica P, Pardo Cardona VM (1996) Flora Uredineana Colombiana. Rev Acad Colomb Cienc 20:183–236 Buzugbe HS, Eze PM, Chukwunwejim CR et al (2018) Investigation of secondary metabolites of an endophytic fungus isolated from the leaves of Chromolaena odorata for possible antimicrobial and antioxidant activities. Pharm Chem J 5:72–77 Cai L, Ji KF, Hyde KD (2006) Variation between freshwater and terrestrial fungal communities on decaying bamboo culms. Antonie Van Leeuwenhoek 89:293–301 Carroll G (1988) Fungal endophytes in stems and leaves: from latent pathogen to mutualistic symbiont. Ecology 69:2–9 Catarino L, Indjai B, Duarte MC, Monteiro F (2019) Chromolaena odorata invasion in Guinea-Bissau (West Africa): first records and trends of expansion. BioInvasions Rec 8:190–198 Chardon CE, Toro RA (1930) Mycological explorations of Colombia. J Dept Agric Porto Rico 14:195–369 Chardon CE, Toro RA (1934) Mycological Explorations of Venezuela. Monogr Univ Puerto Rico B 2:1–351 Chauhan BS, Johnson DE (2008) Germination ecology of two troublesome Asteraceae species of rainfed rice: Siam Weed (Chromolaena odorata) and Coat Buttons (Tridax procumbens). Weed Sci 56:567–573 Chauhan NM, Gutama AD, Aysa A (2019) Endophytic fungal diversity isolated from different agro-ecosystem of Enset (Ensete ventericosum) in Gedeo zone, SNNPRS, Ethiopia. BMC Microbiol 19:172 Chaverri P, Samuels GJ (2013) Evolution of habitat preference and nutrition mode in a cosmopolitan fungal genus with evidence of interkingdom host jumps and major shifts in ecology. Evolution 67:2823–2837 Chaverri P, Branco-Rocha F, Jaklitsch W et al (2015) Systematics of the Trichoderma harzianum species complex and the re-identification of commercial biocontrol strains. Mycologia 107:558–590 Chen K, Zhuang WY (2017) Seven soil-inhabiting new species of the genus Trichoderma in the Viride clade. Phytotaxa 312:28–46 Chen K, Wu XQ, Huang MX, Han YY (2014) First report of brown culm streak of Phyllostachys praecox caused by Arthrinium arundinis in Nanjing, China. Plant Dis 98:1274 Chen Q, Jiang JR, Zhang GZ et al (2015) Resolving the Phoma enigma. Stud Mycol 82:137–217 13 Chen Q, Hou LW, Duan WJ et al (2017) Didymellaceae revisited. Stud Mycol 87:105–159 Chevallier FF (1826) Flore générale des environs de Paris, vol I. Ferra LibrairieEditeur, Paris Chomnunti P, Hongsanan S, Hudson BA et al (2014) The sooty moulds. Fungal Divers 66:1–36 Ciferri R (1961) Mycoflora domingensis integrata. Quaderno 19:1–539 Clements FE, Shear CL (1931) The genera of fungi. HW Wilson Co. Publ., Minneapolis Colegrave N (2012) The evolutionary success of sex. EMBO Rep 13:774–778 Corda ACJ (1838) Patellariaceae in Icones Fungorum hucusque cognitorum 2:1–43 Cronk Q, Fuller J (1995) Plant invaders: the threat to natural ecosystems. Chapman and Hall Publishing Co, London Crous PW, Groenewald JZ (2013) A phylogenetic re-evaluation of Arthrinium. IMA Fungus 4:133–154 Crous PW, Slippers B, Wingfield MJ et al (2006) Phylogenetic lineages in the Botryosphaeriaceae. Stud Mycol 55:235–253 Crous PW, Shivas RG, Quaedvlieg W et al (2014) Fungal Planet description sheets: 214–280. Persoonia 32:184–306 Crous PW, Carris LM, Giraldo A et al (2015a) 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 6:163–198 Crous PW, Schumacher RK, Wingfield MJ et al (2015b) Fungal systematics and evolution: FUSE 1. Sydowia 67:81–118 Crous PW, Wingfield MJ, Guarro J et al (2015c) Fungal planet description sheets: 320–370. Persoonia 34:167–266 Crous PW, Wingfield MJ, Le Roux JJ et al (2015d) Fungal planet description sheets: 371–399. Persoonia 35:264–327 Crous PW, Wingfield MJ, Burgess TI et al (2016) Fungal planet description sheets: 469–557. Persoonia 37:218–403 Crous PW, Wingfield MJ, Burgess TI et al (2017) Fungal Planet description sheets: 625–715. Persoonia 39:270–467 Crous PW, Luangsa-ard JJ, Wingfield MJ et al (2018a) Fungal Planet description sheets: 785–867. Persoonia 41:238–417 Crous PW, Schumacher RK, Wingfield MJ et al (2018b) New and interesting fungi. 1. Fungal Syst Evol 1:169–215 da Cunha KC, Sutton DA, Gené J et al (2014) Pithomyces species (Montagnulaceae) from clinical specimens: identification and antifungal susceptibility profiles. Med Mycol 52:748–757 Dagno K, Lahlali R, Diourté M, Jijakli MH (2011) Effect of temperature and water activity on spore germination and mycelial growth of three fungal biocontrol agents against water hyacinth (Eichhornia crassipes). J Appl Microbiol 110:521–528 Dai DQ, Wijayawardene NN, Bhat DJ et al (2014) Pustulomyces gen. nov. accommodated in Diaporthaceae, Diaporthales, as revealed by morphology and molecular analyses. Cryptogam Mycol 35:63–72 Dai DQ, Phookamsak R, Wijayawardene NN et al (2017) Bambusicolous fungi. Fungal Divers 82:1–105 Dayarathne MC, Jones EBG, Maharachchikumbura SSN et al (2020) Morpho-molecular characterization of microfungi associated with marine based habitats. Mycosphere 11:1–188 de Almeida DAC, Gusmão LFP, Miller AN (2014) A new genus and three new species of hysteriaceous ascomycetes from the semiarid region of Brazil. Phytotaxa 176:298–308 de Gruyter J (2002) Contributions towards a monograph of Phoma (Coelomycetes)—IX Section Macrospora. Persoonia 18:85–102 de Gruyter J, Aveskamp MM, Woudenberg JHC et al (2009) Molecular phylogeny of Phoma and allied anamorph genera: towards a reclassification of the Phoma complex. Mycol Res 113:508–519 Fungal Diversity de Gruyter J, Woudenberg JHC, Aveskamp MM et al (2010) Systematic reappraisal of species in Phoma section Paraphoma, Pyrenochaeta and Pleurophoma. Mycologia 102:1066–1081 de Gruyter J, Woudenberg JHC, Aveskamp MM et al (2013) Redisposition of phoma-like anamorphs in Pleosporales. Stud Mycol 75:1–36 de Hoog GS, Guarro J, Gene J et al (2000) Atlas of Clinical Fungi. Centraalbureau voor Schimmelcultures, 2nd edn, Utrecht, pp 1–1160 De Notaris G (1844) Osservazione su alcuni generi e specie della tribu dei Pirenomiceti sferiacei G. Bot Ital 2:38–55 den Breeÿen A, Groenewald JZ, Verkley GJ, Crous PW (2006) Morphological and molecular characterisation of Mycosphaerellaceae associated with the invasive weed, Chromolaena odorata. Fungal Divers 23:89–110 Devadatha B, Sarma VV, Wanasinghe DN et al (2017) Introducing the new Indian mangrove species, Vaginatispora microarmatispora (Lophiostomataceae) based on morphology and multigene phylogenetic analysis. Phytotaxa 329:139–149 Devadatha B, Sarma VV, Jeewon R et al (2018) Thyridariella, a novel marine fungal genus from India: morphological characterization and phylogeny inferred from multigene DNA sequence analyses. Mycol Progress 17:791–804 Dissanayake AJ, Camporesi E, Hyde KD et al (2016) Dothiorella species associated with woody hosts in Italy. Mycosphere 7:51–63 Dixon LJ, Schlub RL, Pernezny K, Datnoff LE (2009) Host specialization and phylogenetic diversity of Corynespora cassiicola. Phytopathology 99:1015–1027 Doilom M, Liu JK, Jaklitsch WM et al (2013) An outline of the family Cucurbitariaceae. Sydowia 65:167–192 Doilom M, Shuttleworth LA, Roux J et al (2015) Botryosphaeriaceae associated with Tectona grandis (teak) in Northern Thailand. Phytotaxa 233:1–26 Doilom M, Dissanayake AJ, Wanasinghe DN et al (2017) Microfungi on Tectona grandis (teak) in Northern Thailand. Fungal Divers 82:107–182 Dou ZP, Lu M, Wu JR et al (2017) A new species and interesting records of Aplosporella from China. Sydowia 69:1–7 du Plessis IL, Druzhinina IS, Atanasova L et al (2018) The diversity of Trichoderma species from soil in South Africa, with five new additions. Mycologia 110:559–583 Du Z, Fan XL, Yang Q et al (2017) Aplosporella ginkgonis (Aplosporellaceae, Botryosphaeriales), a new species isolated from Ginkgo biloba in China. Mycosphere 8:1246–1252 Ekanayaka AH, Dissanayake AJ, Jayasiri SC et al (2016) Aplosporella thailandica; a novel species revealing the sexual-asexual connection in Aplosporellaceae (Botryosphaeriales). Mycosphere 7:440–447 Ekanayaka AH, Ariyawansa HA, Hyde KD et al (2017) DISCOMYCETES: the apothecial representatives of the phylum Ascomycota. Fungal Divers 87:237–298 El-Desoukey RMA (2018) Phytochemical and antimicrobial activity of Acacia Ehrenbergiana hayne (Salam) as a grazing herb against some animal pathogens. Adv Anim Vet Sci 6:246–251 Ellis MB (1965) Dematiaceous Hyphomycetes. VI. Mycol Pap 103:1–46 Eriksson OE (2006) Outline of Ascomycota. Myconet 12:1–88 Eriksson OE, Hawksworth DL (1986) An alphabetical list of the generic names of ascomycetes. CAB Int Mycol Inst, Wallingford Esperschütz J, Welzl G, Schreiner K et al (2011) Incorporation of carbon from decomposing litter of two pioneer plant species into microbial communities of the detritusphere. FEMS Microbiol Lett 320:48–55 Esperschütz J, Zimmermann C, Dümig A et al (2013) Dynamics of microbial communities during decomposition of litter from pioneering plants in initial soil ecosystems. Biogeosciences 10:5115–5124 Esuruoso OF (1971) Seed-borne fungi of the Siam weed, Eupatorium odoratum in Nigeria. PANS Pest Artic News Summ 17:458–460 Ezard THG, Aze T, Pearson PN, Purvis A (2011) Interplay between changing climate and species’ ecology drives macroevolutionary dynamics. Science 332:349–351 Ezenyi IC, Salawu OA, Kulkarni R, Emeje M (2014) Antiplasmodial activity-aided isolation and identification of quercetin-4’-methyl ether in Chromolaena odorata leaf fraction with high activity against chloroquine-resistant Plasmodium falciparum. Parasitol Res 113:4415–4422 Fajola AO (1978) Cultural studies in Cercospora taxonomy: I. Interrelationships between some species from Nigeria. Nova Hedwigia 29:912–921 Fan XL, Yang Q, Cao B et al (2015) New record of Aplosporella javeedii on five hosts in China based on multi-gene analysis and morphology. Mycotaxon 130:749–756 Farr DF, Rossman AY (2020) Fungal databases, U.S. National Fungus Collections, ARS, USDA. https://nt.ars-grin.gov/fungaldatabases /. Accessed 17 March 2020 Feng Y, Zhang S-N, Liu Z-Y (2019) Tremateia murispora sp. nov. (Didymosphaeriaceae, Pleosporales) from Guizhou, China. Phytotaxa 416:79–87 Fourtouni A, Manetas Y, Christias C (1998) Effects of UV-B radiation on growth, pigmentation, and spore production in the phytopathogenic fungus Alternaria solani. Can J Bot 76:2093–2099 Fries EM (1822) Systema. Mycologicum 2:1–275 Fróhlich J, Hyde KD, Petrini O (2000) Endophytic fungi associated with palms. Mycol Res 104:1202–1212 Gallegos ML, Cummins GB (1981) Uredinales (royas) de Mexico, vol 2. Instituto Nacional de Investigaciones Agricoles, Culiacan Gao Y, Liu F, Duan W et al (2017) Diaporthe is paraphyletic. IMA Fungus 8:153–187 Gäumann EA (1949) Die Pilze, Grundzüge ihrer Entwicklungsgeschichte und Morphologie. Birkhäuser, Basel Gautier L (1992) Taxonomy and distribution of a tropical weed, Chromolaena odorata (L.) R. King and H. Robinson. Candollea 47:645–662 Giller KE, Witter E, Mcgrath SP (1998) Toxicity of heavy metals to microorganisms and microbial processes in agricultural soils: a review. Soil Biol Biochem 30:1389–1414 Glez-Peña D, Gómez-Blanco D, Reboiro-Jato M et al (2010) ALTER: program-oriented conversion of DNA and protein alignments. Nucleic Acids Res 38:W14–W18 Goodall JM, Erasmus DJ (1996) Review of the status and integrated control of the invasive alien weed, Chromolaena odorata in South Africa. Agric Ecosyst Environ 56:151–164 Gottlieb D (1950) The physiology of spore germination in fungi. Bot Rev 16:229–257 Govindu HC, Thirumalachar MJ, Nag Raj TR (1970) Notes on some Indian Cercosporae—XII. Sydowia 24:297–301 Guatimosim E, Schwartsburd PB, Barreto RW, Crous PW (2016) Novel fungi from an ancient niche: cercosporoid and related sexual morphs on ferns. Persoonia 37:106–141 Gueidan C, Hill DJ, Miadlikowska J, Lutzoni F (2015) 4 Pezizomycotina: Lecanoromycetes. In: McLaughlin DJ, Spatafora JW (eds) Systematics and evolution: part B. Springer, Berlin, pp 89–120 Guo Yl (1999a) Fungal flora of tropical Guangxi, China: hyphomycetes I. Mycotaxon 72:349–358 Guo Yl (1999b) Imperfect fungi in the tropical areas of China II. Some species of Pseudocercospora from Guangdong province. Mycosystema 18:130–134 Guo YL, Liu XJ (2003) Flora Fungorum Sinicorum. Mycovellosiella, Passalora, Phaeoramularia, vol 20. Science Press, Beijing, 189 pp 13 Fungal Diversity Guo LD, Hyde KD, Liew ECY (2000) Identification of endophytic fungi from Livistona chinensis based on morphology and rDNA sequences. New Phytol 147:617–630 Guo LD, Hyde KD, Liew ECY (2001) Detection and taxonomic placement of endophytic fungi within frond tissues of Livistona chinensis based on rdna sequences. Mol Phylogenet Evol 20:1–13 Guterres DC, Galvão-Elias S, de Souza BCP et al (2018) Taxonomy, phylogeny, and divergence time estimation for Apiosphaeria guaranitica, a Neotropical parasite on bignoniaceous hosts. Mycologia 110:526–545 Hanh TTH, Hang DTT, Van Minh C, Dat NT (2011) Anti-inflammatory effects of fatty acids isolated from Chromolaena odorata. Asian Pac J Trop Med 4:760–763 Hanphakphoom S, Krajangsang S (2016) Antimicrobial activity of Chromolaena odorata extracts against bacterial human skin infections. Mod Appl Sci 10:159–171 Hansford CG (1949) Tropical fungi—III. New species and revisions. Proc Linn Soc Lond 160:116–153 Harman GE, Kubicek CP (1998) Trichoderma and gliocladium, vol 2. In: Enzymes, biological control and commercial applications. Taylor and Francis, London Harman GE, Howell CR, Viterbo A et al (2004) Trichoderma species—opportunistic, avirulent plant symbionts. Nat Rev Microbiol 2:43–56 Hashimoto A, Hirayama K, Takahashi H et al (2018) Resolving the Lophiostoma bipolare complex: generic delimitations within Lophiostomataceae. Stud Mycol 90:161–189 Haugland RA, Heckman JL (1998) Identification of putative sequence specific PCR primers for detection of the toxigenic fungal species Stachybotrys chartarum. Mol Cell Probes 12:387–396 Hawksworth DL, Eriksson OE (1986) The names of accepted orders of ascomycetes. Syst Ascomycetum 5:175–184 Hawksworth DL, Kirk PM, Sutton BC et al (1995) Ainsworth & Bisby’s dictionary of the fungi, vol 8. CABI, Wallingford Heneghan L, Fatemi F, Umek L et al (2006) The invasive shrub European buckthorn (Rhamnus cathartica, L.) alters soil properties in Midwestern U.S. woodlands. Appl Soil Ecol 32:142–148 Hernández-Restrepo M, Schumacher RK, Wingfield MJ et al (2016) Fungal systematics and evolution: FUSE 2. Sydowia 68:193–230 Hernández-Restrepo M, Bezerra JDP, Tan YP et al (2019) Re-evaluation of Mycoleptodiscus species and morphologically similar fungi. Persoonia 42:205–227 Holm LG, Plucknett DL, Pancho JV, Herberger PD (1977) The world’s worst weeds: distribution and biology. University Press of Hawaii, Honolulu Hongsanan S, Hyde KD, Bahkali AH et al (2015) Fungal biodiversity profiles 11–20. Cryptogam Mycol 36:355–380 Hongsanan S, Maharachchikumbura SSN, Hyde KD et al (2017) An updated phylogeny of Sordariomycetes based on phylogenetic and molecular clock evidence. Fungal Divers 84:25–41 Hosagoudar VB, Mathew S (2000) A preliminary report on the mycoflora of the Andaman & Nicobar Islands, India. J Econ Taxon Bot 24:631–640 Huang WY, Cai YZ, Surveswaran S et al (2009) Molecular phylogenetic identification of endophytic fungi isolated from three Artemisia species. Fungal Divers 36:69–88 Huang W, Long C, Lam E (2018) Roles of plant-associated microbiota in Traditional Herbal Medicine. Trends Plant Sci 23:559–562 Huelsenbeck JP, Ronquist F (2001) MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17:754–755 Hyde KD, Goh TK (2003) Adaptations for dispersal in filamentous freshwater fungi. Fungal Divers Res Ser 10:231–258 Hyde KD, Jones EBG (1988) Marine mangrove fungi. Mar Ecol 9:15–33 13 Hyde KD, Fröhlich J, Taylor JE (1998) Fungi from palms. XXXVI. Reflections on unitunicate ascomycetes with apiospores. Sydowia 50:21–80 Hyde KD, Jones EBG, Liu JK et al (2013) Families of Dothideomycetes. Fungal Divers 63:1–313 Hyde KD, Fryar S, Tian Q et al (2016a) Lignicolous freshwater fungi along a north–south latitudinal gradient in the Asian/Australian region; can we predict the impact of global warming on biodiversity and function? Fungal Ecol 19:190–200 Hyde KD, Hongsanan S, Jeewon R et al (2016b) Fungal diversity notes 367–490: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 80:1–270 Hyde KD, Maharachchikumbura SSN, Hongsanan S et al (2017a) The ranking of fungi: a tribute to David L. Hawksworth on his 70th birthday. Fungal Divers 84:1–23 Hyde KD, Norphanphoun C, Abreu VP et al (2017b) Fungal diversity notes 603–708: taxonomic and phylogenetic notes on genera and species. Fungal Divers 87:1–235 Hyde KD, Chaiwan N, Norphanphoun C et al (2018a) Mycosphere notes 169–224. Mycosphere 9:271–430 Hyde KD, Norphanphoun C, Chen J et al (2018b) Thailand’s amazing diversity: up to 96% of fungi in northern Thailand may be novel. Fungal Divers 93:215–239 Hyde KD, Tennakoon DS, Jeewon R et al (2019a) Fungal diversity notes 1036–1150: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers 96:1–242 Hyde KD, Xu J, Rapior S et al (2019b) The amazing potential of fungi: 50 ways we can exploit fungi industrially. Fungal Divers 97:1–136 Hyde KD, Norphanphoun C, Maharachchikumbura SSN et al (2020) Refined families of Sordariomycetes. Mycosphere 11:305–1059 Ikhajiagbe B, Akindolor A (2016) Comparative effects of pretreatment of stem cuttings of Chromolaena odorata (Siam weed) with sodium azide and hydroxylamide on the survival and phyoremediative performance in an oil-polluted soil. Niger J Biotechnol 31:27–39 Jaklitsch WM (2009) European species of Hypocrea Part I. The greenspored species. Stud Mycol 63:1–91 Jaklitsch WM, Voglmayr H (2016) Hidden diversity in Thyridaria and a new circumscription of the Thyridariaceae. Stud Mycol 85:35–64 Jaklitsch WM, Gardiennet A, Voglmayr H (2016) Resolution of morphology-based taxonomic delusions: Acrocordiella, Basiseptospora, Blogiascospora, Clypeosphaeria, Hymenopleella, Lepteutypa, Pseudapiospora, Requienella, Seiridium and Strickeria. Persoonia 37:82–105 Jami F, Slippers B, Wingfield MJ, Gryzenhout M (2012) Five new species of the Botryosphaeriaceae from Acacia karroo in South Africa. Cryptogam Mycol 33:245–266 Jami F, Slippers B, Wingfield MJ et al (2014) Botryosphaeriaceae species overlap on four unrelated, native South African hosts. Fungal Biol 118:168–179 Jayasiri S, Hyde KD, Jones EBG et al (2017) Taxonomy and multigene phylogenetic evaluation of novel species in Boeremia and Epicoccum with new records of Ascochyta and Didymella (Didymellaceae). Mycosphere 8:1080–1101 Jayasiri SC, Hyde KD, Jones EBG et al (2018) Taxonomic novelties of hysteriform Dothideomycetes. Mycosphere 9:803–837 Jayasiri SC, Hyde KD, Jones EBG et al (2019) Diversity, morphology and molecular phylogeny of Dothideomycetes on decaying wild seed pods and fruits. Mycosphere 10:1–186 Jeewon R, Liew ECY, Hyde KD (2003) Molecular systematics of the Amphisphaeriaceae based on cladistic analyses of partial LSU rDNA gene sequences. Mycol Res 107:1392–1402 Jeewon R, Ittoo J, Mahadeb D et al (2013) DNA Based identification and phylogenetic characterisation of endophytic and saprobic Fungal Diversity fungi from Antidesma madagascariense, a medicinal plant in Mauritius. J Mycol 1–10 Jeffrey C (2007) Introduction with key to tribes. In: Kadereit JW, Jeffrey C (eds) the families and genera of vascular plants, vol 8. Flowering plants. Eudicots. Asterales. Springer, Berlin, pp 61–87 Jia M, Chen L, Xin HL et al (2016) A friendly relationship between endophytic fungi and medicinal plants: a systematic review. Front Microbiol 7:1 Jiang N, Li J, Tian CM (2018) Arthrinium species associated with bamboo and reed plants in China. Fungal Syst Evol 2:1–9 Jiang HB, Hyde KD, Jayawardena RS et al (2019) Taxonomic and phylogenetic characterizations reveal two new species and two new records of Roussoella (Roussoellaceae, Pleosporales) from Yunnan, China. Mycol Progress 18:577–591 Jin HQ, Liu HB, Xie YY et al (2018) Effect of the dark septate endophytic fungus Acrocalymma vagum on heavy metal content in tobacco leaves. Symbiosis 74:89–95 Johari SA, Kiong LS, Mohtar M et al (2012) Efflux inhibitory activity of flavonoids from Chromolaena odorata against selected methicillin-resistant Staphylococcus aureus (MRSA) isolates. Afr J Microbiol Res 6:5631–5635 Johnston A (1960) A supplement to a host list of plant diseases in Malaya. Mycol Pap 77:1–30 Jones EBG (2006) Form and function of fungal spore appendages. Mycoscience 47:167 Jones EBG, Pang KL, Abdel-Wahab MA et al (2019) An online resource for marine fungi. Fungal Divers 96:347–433 Kamal (2010) Cercosporoid fungi of India. Bishen Singh Mahendra Pal Singh, Dehra Dun Kanase V, Shaikh S (2018) A pharmacognostic and pharmacological review on Chromolaena odorata (Siam weed). Asian J Pharm Clin Res 11:34–38 Kang JC, Hyde KD, Kong RYC (1999) Studies on Amphisphaeriales: the Cainiaceae. Mycol Res 103:1621–1627 Kantarcioğlu AS, Celkan T, Yücel A et al (2009) Fatal Trichoderma harzianum infection in a leukemic pediatric patient. Med Mycol 47:207–215 Karunarathna A, Papizadeh M, Senanayake IC et al (2017) Novel fungal species of Phaeosphaeriaceae with an asexual/sexual morph connection. Mycosphere 8:1818–1834. Katoh K, Rozewicki J, Yamada KD (2019) MAFFT online service: multiple sequence alignment, interactive sequence choice and visualization. Brief Bioinform 20:1160–1166 Kern FD (1930) Fungi of Santo Domingo—III, Uredinales. Mycologia 22:111–117 Kern FD, Ciferri R, Thurston-Jr HW (1933a) The rust-flora of the Dominican Republic. Ann Mycol 31:1–40 Kern FD, Thurston HW Jr, Whetzel HH (1933b) Annotated index of the rusts of Colombia. Mycologia 25:448–503 Kern FD, Thurston HW Jr, Whetzel HH (1934) Uredinales. Monogr Univ Puerto Rico B 2:262–303 Kigigha LT, Zige DV (2013) Activity of Chromolaena odorata on enteric and superficial etiologic bacterial agents. Am J Res Commun 1:266–276 Kirk PM (1986) New or interesting microfungi. XV. Miscellaneous hyphomycetes from the British Isles. Trans Br Mycol Soc 86:409–428 Kirk PM, Cannon PF, Minter DW, Stalpers JA (2008) Ainsworth & Bisby’s dictionary of the fungi, 10th edn. CABI, Wallingford Kirschner R, Chen CJ (2007) Foliicolous hyphomycetes from Taiwan. Fungal Divers 26:219–239 Klein D, Eveleigh DE (1998) Ecology of Trichoderma. In: Kubicek CP, Harman GE (eds) Trichoderma and Gliocladium. Taylor & Francis, London, pp 57–74 Köberl M, Schmidt R, Ramadan EM et al (2013) The microbiome of medicinal plants: diversity and importance for plant growth, quality and health. Front Microbiol 4:400 Kodsueb R, Lumyong S, McKenzie EHC et al (2016) Relationships between terrestrial and freshwater lignicolous fungi. Fungal Ecol 19:155–168 Kohlmeyer J, Volkmann-Kohlmeyer B, Eriksson OE (1995) Fungi on Juncus roemerianus. New dictyosporous ascomycetes. Bot Mar 38:165–174 Kolařík M (2018) New taxonomic combinations in endophytic representatives of the genus Nigrograna. Czech Mycol 70:123–126 Kolařík M, Spakowicz DJ, Gazis R et al (2017) Biatriospora (Ascomycota: Pleosporales) is an ecologically diverse genus including facultative marine fungi and endophytes with biotechnological potential. Plant Syst Evol 303:35–50 Konta S, Hongsanan S, Tibpromma S et al (2016) An advance in the endophyte story: Oxydothidaceae fam. nov. with six new species of Oxydothis. Mycosphere 7:1425–1446 Kouamé PBK, Jacques C, Bedi G et al (2013) Phytochemicals isolated from leaves of Chromolaena odorata: impact on viability and clonogenicity of cancer cell lines. Phytother Res 27:835–840 Koutika LS, Rainey HJ (2010) Chromolaena odorata in different ecosystems: weed or fallow plant? Appl Ecol Environ Res 8:131–142 Krug JC (1977) The genus Cainia and a new family, Cainiaceae. Sydowia 30:122–133 Kumar V, Cheewangkoon R, Thambugala KM et al (2019) Rhytidhysteron mangrovei (Hysteriaceae), a new species from mangroves in Phetchaburi Province, Thailand. Phytotaxa 266:250–260 Kutorga E, Hawksworth DL (1997) A re-assessment of the genera referred to the family Patellariaceae (Ascomycota). Systema Ascomycetum 15:1–110 Le Dinh T, Zaw M, Matsumoto M (2019) Diaporthe species complex occurring on Asparagus kiusianus in Japan. J Plant Pathol 101:161–167 Leylaie S, Zafari D (2018) Antiproliferative and antimicrobial activities of secondary metabolites and phylogenetic study of endophytic Trichoderma species from Vinca Plants. Front Microbiol 9:1484 Li D, Yang CS, Haugland RA et al (2003) A new species of Memnoniella. Mycotaxon 85:253–257 Li QR, Tan P, Jiang YL et al (2013) A novel Trichoderma species isolated from soil in Guizhou, T. guizhouense. Mycol Prog 12:167–172 Li BJ, Liu PQ, Jiang Y, Weng QY, Chen QH (2016a) First report of culm rot caused by Arthrinium phaeospermum on Phyllostachys viridis in China. Plant Dis 100:1013 Li JF, Bhat DJ, Phookamsak R et al (2016b) Sporidesmioides thailandica gen. et sp. nov. (Dothideomycetes) from northern Thailand. Mycol Progress 15:1169–1178 Li JF, Phookamsak R, Jeewon R et al (2017) Molecular taxonomy and morphological characterization reveal new species and new host records of Torula species (Torulaceae, Pleosporales). Mycol Progress 16:447–461 Liao C, Peng R, Luo Y et al (2008) Altered ecosystem carbon and nitrogen cycles by plant invasion: a meta-analysis. New Phytol 177:706–714 Liew ECY, Aptroot A, Hyde KD (2000) Phylogenetic significance of the pseudoparaphyses in Loculoascomycete taxonomy. Mol Phylogeny Evol 16:392–402 Lin C, McKenzie EHC, Bhat DJ et al (2016) Stachybotrys-like taxa from karst areas and a checklist of stachybotrys-like species from Thailand. Mycosphere 7:1273–1291 Lindau G (1897a) Hypocreales. In: Englier HA, Prantl KAE (eds) Die Naturlichen Pflanzenfamilien 1. Verlag W, Engelman, pp 343–372 Lindau G (1897b) Hysteriineae. In: Engler & Prantl, Naturliche Pflanzenfamilien. I. Teil, I. Abteilung. 1:265–278 13 Fungal Diversity Litzenberger SC, Farr ML, Lip HT (1962) A preliminary list of cambodian plant diseases. USAID Division of Agriculture & Natural Resources, Phnom-Penh Liu XJ, Guo YL (1988) Studies on the genus Mycovellosiella of China. Mycosystema 1:241–268 Liu XJ, Yl Guo (1998) Flora Fungorum Sinicorum. Pseudocercospora, vol 9. Science Press, Beijing Liu YJ, Whelen S, Hall BD (1999) Phylogenetic relationships among ascomycetes: evidence from an RNA polymerase II subunit. Mol Biol Evol 16:1799–1808 Liu JK, Phookamsak R, Doilom M et al (2012) Towards a natural classification of Botryosphaeriales. Fungal Divers 57:149–210 Liu JK, Phookamsak R, Dai DQ et al (2014) Roussoellaceae, a new pleosporalean family to accommodate the genera Neoroussoella gen. nov. Roussoella and Roussoellopsis. Phytotaxa 181:1–33 Liu JK, Hyde KD, Jones EBG et al (2015) Fungal diversity notes 1–110: taxonomic and phylogenetic contributions to fungal species. Fungal Divers 72:1–197 Liu JK, Hyde KD, Jeewon R et al (2017) Ranking higher taxa using divergence times: a case study in Dothideomycetes. Fungal Divers 84:75–99 Liu LN, Razaq A, Atri NS et al (2018) Fungal systematics and evolution: FUSE 4. Sydowia 70:211–286 Lombard L, Houbraken J, Decock C et al (2016) Generic hyper-diversity in Stachybotriaceae. Persoonia 36:156–246 Lumbsch HT, Huhndorf SM (2007) Outline of Ascomycota—2007. Myconet 13:1–99 Lumbsch HT, Huhndorf SM (2010a) Myconet volume 14. Part one. Outline of ascomycota—2009. Part two. Notes on ascomycete systematics. Nos. 4751–5113. Fieldiana Life Earth Sci 16:1–65 Lumbsch HT, Huhndorf SM (2010b) Outline of Ascomycota—2009. Fieldiana Life Earth Sci 1:1–60 Luttrell ES (1951) Taxonomy of pyrenomycetes. University of Missouri Studies 24:1–120 Luttrell ES (1955) The ascostromatic ascomycetes. Mycologia 47:511–532 Macdonald IAW (1983) Alien trees, shrubs and creepers invading indigenous vegetation in the Hluhluwe-Umfolozi Game Reserve Complex in Natal. Bothalia 14:949–959 Maharachchikumbura SSN, Hyde KD, Jones EBG et al (2015) Towards a natural classification and backbone tree for Sordariomycetes. Fungal Divers 72:199–301 Maharachchikumbura SSN, Hyde KD, Jones EBG et al (2016) Families of Sordariomycetes. Fungal Divers 79:1–317 Mapook A, Boonmee S, Ariyawansa HA et al (2016a) Taxonomic and phylogenetic placement of Nodulosphaeria. Mycol Prog 15:34 Mapook A, Hyde KD, Dai D-Q et al (2016b) Muyocopronales, ord. nov., (Dothideomycetes, Ascomycota) and a reappraisal of Muyocopron species from northern Thailand. Phytotaxa 265:225–237 Mapook A, Hyde KD, Hongsanan S et al (2016c) Palawaniaceae fam. nov., a new family (Dothideomycetes, Ascomycota) to accommodate Palawania species and their evolutionary time estimates. Mycosphere 7:1732–1745 Mapook A, Macabeo APG, Thongbai B, et al (2020) Polyketidederived secondary metabolites from a Dothideomycetes fungus, Pseudopalawania siamensis gen. et sp. nov., (Muyocopronales) with antimicrobial and cytotoxic activities. Biomolecules 10:569 Matthews S, Brand K (2004) Africa invaded: the growing danger of invaded alien species. The global Invasive species Programme (GISP) Secretariat. Cape Town, South Africa. www.gisp.org/ downloadpubs/gipsAfrica.pdf Maynard Smith J (1978) The evolution of sex. Cambridge University Press, Cambridge McFadyen R (1996) Potential distribution of Chromolaena odorata (Siam weed) in Australia, Africa and Oceania. Agric Ecosyst Environ 59:89–96 13 Meeboon J, Hidayat I, To-anun C (2007) An annotated list of cercosporoid fungi in Northern Thailand. J Agric Technol 3:51–63 Mehl JWM, Slippers B, Roux J, Wingfield MJ (2014) Botryosphaeriaceae associated with die-back of Schizolobium parahyba trees in South Africa and Ecuador. Forest Pathol 44:396–408 Minter DW, Rodríguez-Hernández M, Mena Portales J (2001) Fungi of the Caribbean: an annotated checklist. PDMS Publishing Mitakakis TZ, O’meara TJ, Tovey ER (2003) The effect of sunlight on allergen release from spores of the fungus Alternaria. Grana 42:43–46 Moore AB (2004) Alien invasive species: impacts on forests and forestry. Forest Health and Biosecurity Working Papers FAO, Rome Morgan-Jones G (1997) Notes on hyphomycetes. LXXIII. Redescription of Phaeoramularia coalescens, Phaeoramularia eupatoriiodorati and Phaeoramularia pruni. Mycotaxon 61:363–373 Müller E, von Arx JA (1950) Einige aspekte zur systematik pseudosphärialer ascomyceten. Ber Schw Bot Ges 60:329–397 Muniappan R, Reddy GVP, Lai PY (2005) Distribution and biological control of Chromolaena odorata. In: Inderjit A (ed) Invasive plants: ecological and agricultural aspects. Birkhäuser, Basel, pp 223–233 Munk A (1953) The system of the Pyrenomycetes: a contribution to a natural classification of the group Sphacriales sensu Lindau. Dansk Bot Arkiv 15:1–163 Murgia M, Fiamma M, Barac A et al (2018) Biodiversity of fungi in hot desert sands. Microbiol Open 8:e00595 Nag Raj TR (1974) Icones generum coelomycetum VI. Univ Waterloo Biol Ser 13:1–41 Naidoo KK, Coopoosamy RM, Naidoo G (2011) Screening of Chromolaeana odorata (L.) King and Robinson for antibacterial and antifungal properties. Med Plants Res 5:4859–4862 Nannfeldt JA (1932) Studien über die Morphologie und Systematik der nicht-lichenisierten, inoperkulaten Discomyceten. Nova Acta Regiae Soc Sci Uppsaliensis Ser 6:1–368 Nesher I, Barhoom S, Sharon A (2008) Cell cycle and cell death are not necessary for appressorium formation and plant infection in the fungal plant pathogen Colletotrichum gloeosporioides. BMC Biol 6:9 Ngono NA, Ebelle Etame R, Ndifor F et al (2006) Antifungal activity of Chromolaena odorata (L.). King and Robinson (Asteraceae) of Cameroon. Chemotherapy 52:103–106 Nguanhom J, Cheewangkoon R, Groenewald JZ et al (2015) Taxonomy and phylogeny of Cercospora spp. from Northern Thailand. Phytotaxa 233:27–48 Niranjan M, Sarma VV (2018) Twelve new species of Ascomycetous fungi from Andaman Islands, India. Kavaka 50:84–97 Nitschke T (1870) Pyrenomycetes Germanici 2, 161–320. Eduard Trewendt, Wroclaw Norphanphoun C, Jeewon R, Mckenzie EHC et al (2017) Taxonomic position of Melomastia italica sp. nov. and phylogenetic reappraisal of Dyfrolomycetales. Cryptogam. Mycology 38:507–526 Nylander JAA (2004) MrModeltest 2.0. Program distributed by the author. Evolutionary Biology Centre, Uppsala University Ohm RA, Feau N, Henrissat B et al (2012) Diverse Lifestyles and strategies of plant pathogenesis encoded in the genomes of eighteen dothideomycetes fungi. PLoS Pathog 8:e1003037 Okoronkwo A, Aiyesanmi A, Odiyi A et al (2014) Bioaccumulation of cadmium in Siam (Chromolaena odorata) and Node (Synedrella nodiflora) weeds: impact of ethylene diamine tetraacetic acid (EDTA) on uptake. Environ Nat Resour Res 4:p39 Onkaramurthy M, Veerapur VP, Thippeswamy BS et al (2013) Antidiabetic and anti-cataract effects of Chromolaena odorata Linn., in streptozotocin-induced diabetic rats. J Ethnopharmacol 145:363–372 Fungal Diversity Osman M, Valadon LRG (1981) Effect of light (especially near-U.V.) on spore germination and ultrastructure of Verticillium agaricinum. Trans Br Mycol Soc 77:187–189 Pandith H, Zhang X, Liggett J et al (2013a) Hemostatic and woundhealing properties of Chromolaena odorata leaf extract. ISRN Dermatol 2013:168269 Pandith H, Zhang X, Thongpraditchote S et al (2013b) Effect of Siam weed extract and its bioactive component scutellarein tetramethyl ether on anti-inflammatory activity through NF-κB pathway. J Ethnopharmacol 147:434–441 Pang KL, Hyde KD, Alias SA et al (2013) Dyfrolomycetaceae, a new family in the Dothideomycetes, Ascomycota. Cryptogam. Mycology 34:223–232 Papizadeh M, Soudi MR, Amini L et al (2017) Pyrenochaetopsis tabarestanensis (Cucurbitariaceae, Pleosporales), a new species isolated from rice farms in north Iran. Phytotaxa 297:15–28 Pardo Cardona VM (1998) Distribucion de las especies colombianas de Uredinales segun los grupos taxonomicos de sus hospederos. Rev Fac Nal Agric Medellin 51:285–319 Pem D, Gafforov Y, Jeewon R et al (2018) Multigene phylogeny coupled with morphological characterization reveal two new species of Holmiella and taxonomic insights within Patellariaceae. Cryptogam. Mycology 39(2):193–209 Peregrine WTH, Ahmad KB (1982) Brunei: A first annotated list of plant diseases and associated organisms. Phytopathol Pap 27:1–87 Persoon CH (1794) Neuer Versuch einer systematischen Einteilung der Schwämme. N Mag Die Bot Ihrem Ganzen Umfange 1: 63–128 Phengsintham P, Chukeatirote E, Abdelsalam KA, Hyde KD, Braun U (2010) Cercospora and allied genera from Laos 2. Cryptogam. Mycology 31:161–181 Phengsintham P, Braun U, McKenzie EHC, Chukeatirote E, Cai L, Hyde KD (2013) Monograph of Cercosporoid fungi from Thailand. Pl Pathol Quarantine 3:67–138 Phillips AJL, Alves A, Abdollahzadeh J et al (2013) The Botryosphaeriaceae: genera and species known from culture. Stud Mycol 76:51–167 Phillips AJL, Hyde KD, Alves A, Liu JK (2019) Families in Botryosphaeriales: a phylogenetic, morphological and evolutionary perspective. Fungal Divers 94:1–22 Phookamsak R, Liu JK, McKenzie EHC et al (2014) Revision of Phaeosphaeriaceae. Fungal Divers 68:159–238 Phookamsak R, Hyde KD, Jeewon R et al (2019) Fungal diversity notes 929–1035: taxonomic and phylogenetic contributions on genera and species of fungi. Fungal Divers 95:1–273 Phukhamsakda C, Ariyawansa HA, Phillips AJL et al (2016) Additions to Sporormiaceae: Introducing two novel genera, Sparticola and Forliomyces, from Spartium. Cryptogam. Mycology 37:75–97 Piepenbring M (2006) Checklist of fungi in Panama. Preliminary version. Puente Biol 1:1–190 Pintos Á, Alvarado P, Planas J, Jarling R (2019) Six new species of Arthrinium from Europe and notes about A. caricicola and other species found in Carex spp. hosts. MycoKeys 49:15–48 Pisutthanan N, Liawruangrath B, Liawruangrath S et al (2005) Constituents of the essential oil from aerial parts of Chromolaena odorata from Thailand. Nat Prod Res 20:636–640 Pitt W, Úrbez-Torres JR, Trouillas FP (2014) Munkovalsaria donacina from grapevines and Desert Ash in Australia. Mycosphere 5:656–661 Promputtha I, Lumyong S, Dhanasekaran V et al (2007) A phylogenetic evaluation of whether endophytes become saprotrophs at host senescence. Microb Ecol 53:579–590 Promputtha I, Hyde KD, McKenzie EHC et al (2010) Can leaf degrading enzymes provide evidence that endophytic fungi becoming saprobes? Fungal Divers 41:89–99 Puckdeedindan P (1966) A supplementary host list of plant diseases in Thailand. Technical Bulletin Department of Agriculture, Bangkok 7, pp 1–23 Purahong W, Hyde KD (2011) Effects of fungal endophytes on grass and non-grass litter decomposition rates. Fungal Divers 47:1–7 Qin WT, Zhuang WY (2017) Seven new species of Trichoderma (Hypocreales) in the Harzianum and Strictipile clades. Phytotaxa 305:121–139 Qin RM, Zheng YL, Valiente-Banuet A et al (2013) The evolution of increased competitive ability, innate competitive advantages, and novel biochemical weapons act in concert for a tropical invader. New Phytol 197:979–988 Quaedvlieg W, Verkley GJM, Shin HD, Barreto RW, Alfenas AC, Swart WJ, Groenewald JZ, Crous PW (2013) Sizing up Septoria. Stud Mycol 75:307–390 Rabenhorst (1857) Ein Notizblatt für kryptogamische Studien. Hedwigia 1:116 Rai MK (1989) Mycosis in man due to Arthrinium phaeospermum var. indicum. First case report. Mycoses 32:472–475 Rambaut A (2009) FigTree v1.4: tree figure drawing tool. http://tree. bio.ed.ac.uk/software/figtree/ Rambaut A, Suchard MA, Drummond AJ (2014) Tracer v 1.6, available from: http://tree.bio.ed.ac.uk/software/tracer/ Rani CA, Mohammed R (2012) Ethnobotanical study of wound healing plants among the folk medicinal practitioners of several districts in Bangladesh. Am Eur J Sustain Agric 6:371–377 Rannala B, Yang Z (1996) Probability distribution of molecular evolutionary trees: a new method of phylogenetic inference. J Mol Evol 43:304–311 Rao VG (1969) Fungi on Citrus from India. Sydowia 23:215–224 Rehner S (2001) Primers for elongation factor 1-α (EF1-α). http://ocid. NACSE.ORG/research/deephyphae/EF1primer.pdf Richardson MJ (1990) An annotated list of seed-borne diseases, 4th edn. International Seed Testing Association, Zurich Rose S (1997) Influence of suburban edges on invasion of Pittosporum undulatum into the bush land of northern Sydney, Australia. Aust J Ecol 22:89–99 Roux J, Wingfield MJ, Morris MJ (1997) Botryosphaeria dothidea, a pathogen of Acacia mearnsii in South Africa. S Afr J Sci 93:xii Roy BA (2001) Patterns of association between crucifers and their flower-mimic pathogens: host jumps are more common than coevolution or cospeciation. Evolution 55:41–53 Ruangdech T, Wongphatcharachai M, Staley C et al (2017) Influence of heavy metals on rhizosphere microbial communities of Siam weed (Chromolaena odorata L.) using a 16S rRNA gene amplicon sequencing approach. Agric Nat Resour 51:137–141 Rubini MR, Silva-Ribeiro RT, Pomella AW et al (2005) Diversity of endophytic fungal community of cacao (Theobroma cacao L.) and biological control of Crinipellis perniciosa, causal agent of witches’ broom disease. Int J Biol Sci 1:24–33 Ruibal C, Gueidan C, Selbmann L et al (2009) Phylogeny of rockinhabiting fungi related to Dothideomycetes. Stud Mycol 64:123–133 Russo VM (1985) Leaf spot disease of Chromolaena odorata caused by Septoria sp. in Guam. Pl Dis 69:1101 Saccardo PA (1880) Conspectus genera fungorum Italiae inferiorum nempe ad Sphaeropsideas, Melanconieas et Hyphomyceteas pertinentium systemate sporologico dispositorum. Michelia 2:1–38 Saccardo PA (1882) Sylloge Fungorum omnium hucusque cognitorum 1:715 Saccardo PA (1883) Sylloge Fungorum omnium hucusque cognitorum 2:658–661 Samarakoon MC, Hyde KD, Hongsanan S et al (2019) Divergence time calibrations for ancient lineages of Ascomycota classification based on a modern review of estimations. Fungal Divers 96:285–346 13 Fungal Diversity Samuels GJ (2006) Trichoderma: systematics, the sexual state, and ecology. Phytopathology 96:195–206 Sandoval-Denis M, Sutton DA, Cano-Lira JF et al (2014) Phylogeny of the clinically relevant species of the emerging fungus Trichoderma and their antifungal susceptibilities. J Clin Microbiol 52:2112–2125 Sawada K (1931) List of fungi found in Formosa. Taiwan Central Institute, Taipei Schoch CL, Shoemaker RA, Seifert KA et al (2006) A multigene phylogeny of the Dothideomycetes using four nuclear loci. Mycologia 98:1041–1052 Schoch CL, Crous PW, Groenewald J et al (2009) A class wide phylogenetic assessment of Dothideomycetes. Stud Mycol 64:1–15 Scott LJ, Lange CL, Graham GC, Yeates DK (1998) Genetic diversity and origin of siam weed (Chromolaena odorata) in Australia. Weed Technol 12:27–31 Senanayake IC, Maharachchikumbura SSN, Hyde KD et al (2015) Towards unraveling relationships in Xylariomycetidae (Sordariomycetes). Fungal Divers 73:73–144 Senanayake IC, Crous PW, Groenewald JZ et al (2017) Families of Diaporthales based on morphological and phylogenetic evidence. Stud Mycol 86:217–296 Senanayake IC, Jeewon R, Chomnunti P et al (2018) Taxonomic circumscription of Diaporthales based on multigene phylogeny and morphology. Fungal Divers 93:241–443 Senwanna C, Hongsanan S, Phookamsak R et al (2019) Muyocopron heveae sp. nov. and M. dipterocarpi appears to have host-jumped to rubber. Mycol Progress 18:741–752 Setter MJ, Campbell SD (2002) Impact of foliar herbicides on germination and viability of Siam weed (Chromolaena odorata) seeds located on plants at the time of application. Plant Prot Q 17:155–157 Shao X, Li Q, Lin L, He T (2018) On the origin and genetic variability of the two invasive biotypes of Chromolaena odorata. Biol Invasions 20:2033–2046 Sharma R, Kulkarni G, Sonawane MS (2017) Alanomyces, a new genus of Aplosporellaceae based on four loci phylogeny. Phytotaxa 297:168–175 Shipunov A, Newcombe G, Raghavendra AKH, Anderson CL (2008) Hidden diversity of endophytic fungi in an invasive plant. Am J Bot 95:1096–1108 Silvestro D, Michalak I (2010) raxmlGUI: a graphical front-end for RAxML Program and documentation. http://www.sourc eforg e.com Sirinthipaporn A, Jiraungkoorskul W (2017) Wound healing property review of Siam weed, Chromolaena odorata. Pharmacogn Rev 11:35 Slippers B, Wingfield MJ (2007) Botryosphaeriaceae as endophytes and latent pathogens of woody plants: diversity, ecology and impact. Fungal Biol Rev 21:90–106 Slippers B, Boissin E, Phillips AJL et al (2013) Phylogenetic lineages in the Botryosphaeriales: a systematic and evolutionary framework. Stud Mycol 76:31–49 Slippers B, Roux J, Wingfield MJ et al (2014) Confronting the constraints of morphological taxonomy in the Botryosphaeriales. Persoonia Mol Phylogeny Evol Fungi 33:155–168 Smith GJD, Liew ECY, Hyde KD (2003) The Xylariales: a monophyletic order containing 7 families. Fungal Divers 13:185–218 Soleimani P, Soleimani MJ, Hosseini S (2018) Phylogenetic relationship and evolution of Neodidymelliopsis isolates collected from Iran. Mycosphere 9:1235–1255 Soto-Medina EA, Lücking R (2017) A new species of Rhytidhysteron (Ascomycota: Patellariaceae) from Colombia, with a provisional working key to known species in the world. Rev Acad Colomb Cienc Exactas Físicas Nat 41:59 13 Spegazzini C (1880) Fungi argentini. Pugillus tertius (Continuacion). Anal Soc Cient Argent 10:145–168 Spegazzini C (1881a) Fungi argentini additis nonnullis brasiliensibus montevideensibusque. Pugillus quartus (Continuacion). Anal Soc Cient Argent 12:174–189 Spegazzini C (1881b) Fungi argentini additis nonnullis brasiliensibus montevideensibusque. Pugillus quartus (Continuacion). Anal Soc Cient Argent 12:97–117 Stamatakis A, Hoover P, Rougemont J (2008) A rapid bootstrap algorithm for the RAxML web-servers. Syst Biol 75:758–771 Stanley MC, Ifeanyi OE, Nwakaego CC, Esther IO (2014) Antimicrobial effects of Chromolaena odorata on some human pathogens. Int J Curr Microbiol Appl Sci 3:1006–1012 Stevenson JA (1975) Fungi of Puerto Rico and the American Virgin Islands. Contr Reed Herb 23:743 Strobel G (2018) The emergence of endophytic microbes and their biological promise. J Fungi 4:57 Sturm J (1829) Deutschlands flora, Abt. III. Die Pilze Deutschlands 2:1–136 Su H, Hyde KD, Maharachchikumbura SSN et al (2016) The families Distoseptisporaceae fam. nov., Kirschsteiniotheliaceae, Sporormiaceae and Torulaceae, with new species from freshwater in Yunnan Province, China. Fungal Divers 80:375–409 Su XJ, Luo ZL, Jeewon R et al (2018) Morphology and multigene phylogeny reveal new genus and species of Torulaceae from freshwater habitats in northwestern Yunnan, China. Mycol Progress 17:531–545 Suksamrarn A, Chotipong A, Suavansri T et al (2004) Antimycobacterial activity and cytotoxicity of flavonoids from the flowers of Chromolaena odorata. Arch Pharm Res 27:507–511 Sun X, Guo L-D, Hyde KD (2011) Community composition of endophytic fungi in Acer truncatum and their role in decomposition. Fungal Divers 47:85–95 Sutton BC (1968) The appressoria of Colletotrichum graminicola and C. falcatum. Can J Bot 46:873–876 Tai FL (1979) Sylloge Fungorum Sinicorum. Science Press, Peking Tamura K, Peterson D, Peterson N et al (2011) MEGA5: molecular evolutionary genetics analysis using maximum likelihood, evolutionary distance, and maximum parsimony methods. Mol Biol Evol 28:2731–2739 Tamura K, Stecher G, Peterson D et al (2013) MEGA6: molecular evolutionary genetics analysis version 6.0. Mol Biol Evol 30:2725–2729 Tanaka K, Hirayama K, Yonezawa H et al (2015) Revision of the Massarineae (Pleosporales, Dothideomycetes). Stud Mycol 82:75–136 Tanhan P, Kruatrachue M, Pokethitiyook P, Chaiyarat R (2007) Uptake and accumulation of cadmium, lead and zinc by Siam weed [Chromolaena odorata (L.) King & Robinson]. Chemosphere 68:323–329 Taylor JE, Hyde KD (2003) Microfungi of tropical and temperate palms. Fungal Divers Res Ser 12:1–459 Tennakoon DS, Hyde KD, Wanasinghe DN et al (2016) Taxonomy and phylogenetic appraisal of Montagnula jonesii sp. nov. (Didymosphaeriaceae, Pleosporales). Mycosphere 7:1346–1356 Tennakoon DS, Kuo CH, Jeewon R et al (2018) Saprobic Lophiostomataceae (Dothideomycetes): Pseudolophiostoma mangiferae sp. nov. and Neovaginatispora fuckelii, a new record from Mangifera indica. Phytotaxa 364:157–171 Thambugala KM, Hyde KD (2018) Additions to the genus Massariothea in Diaporthaceae. Mycol Progress 17:1139–1147 Thambugala KM, Hyde KD, Tanaka K et al (2015) Towards a natural classification and backbone tree for Lophiostomataceae, Floricolaceae, and Amorosiaceae fam. nov. Fungal Divers 74:199–266 Fungal Diversity Thambugala KM, Hyde KD, Eungwanichayapant PD et al (2016) Additions to the genus Rhytidhysteron in Hysteriaceae. Cryptogam. Mycology 37:99–116 Thambugala KM, Daranagama DA, Phillips AJL et al (2017a) Microfungi on Tamarix. Fungal Divers 82:239–306 Thambugala KM, Wanasinghe DN, Phillips AJL et al (2017b) Mycosphere notes 1–50: Grass (Poaceae) inhabiting Dothideomycetes. Mycosphere 8:697–796 Thaung MM (1984) Some fungi of Cercospora complex from Burma. Mycotaxon 19:425–452 Theissen F, Sydow H (1918) Vorentwürfe zu den Pseudosphaeriales. Ann Mycol 16:1–34 Thomas B, Arumugam R, Veerasamy A, Ramamoorthy S (2014) Ethnomedicinal plants used for the treatment of cuts and wounds by Kuruma tribes, Wayanadu districts of Kerala, India. Asian Pac J Trop Biomed 4:88–91 Thompson SM, Tan YP, Young AJ et al (2011) Stem cankers on sunflower (Helianthus annuus) in Australia reveal a complex of pathogenic Diaporthe (Phomopsis) species. Persoonia 27:80–89 Thompson SM, Tan YP, Shivas RG et al (2015) Green and brown bridges between weeds and crops reveal novel Diaporthe species in Australia. Persoonia 35:39–49 Tibpromma S, McKenzie EHC, Karunarathna SC et al (2016) Muyocopron garethjonesii sp. nov. (Muyocopronales, Dothideomycetes) on Pandanus sp. Mycosphere 7:1480–1489 Tibpromma S, Hyde KD, Jeewon R et al (2017) Fungal diversity notes 491–602: taxonomic and phylogenetic contributions to fungal taxa. Fungal Divers 83:1–261 Tibpromma S, Hyde KD, McKenzie EHC et al (2018) Fungal diversity notes 840–928: micro-fungi associated with Pandanaceae. Fungal Divers 93:1–160 Trakunyingcharoen T, Cheewangkoon R, To-anun C et al (2014a) Botryosphaeriaceae associated with diseases of mango (Mangifera indica). Australas Plant Pathol 43:425–438 Trakunyingcharoen T, Lombard L, Groenewald JZ et al (2014b) Mycoparasitic species of Sphaerellopsis, and allied lichenicolous and other genera. IMA Fungus 5:391–414 Trakunyingcharoen T, Lombard L, Groenewald JZ et al (2015) Caulicolous Botryosphaeriales from Thailand. Persoonia 34:87–99 Trognitz F, Hackl E, Widhalm S, Sessitsch A (2016) The role of plantmicrobiome interactions in weed establishment and control. FEMS Microbiol Ecol 92:1–15 Truter M, Přikrylová I, Weyl OLF, Smit NJ (2017) Co-introduction of ancyrocephalid monogeneans on their invasive host, the largemouth bass, Micropterus salmoides (Lacepéde, 1802) in South Africa. Int J Parasitol Parasit Wildl 6:420–429 Tůma I, Holub P, Fiala K (2009) Soil nutrient heterogeneity and competitive ability of three grass species (Festuca ovina, Arrhenatherum elatius and Calamagrostis epigejos) in experimental conditions. Biologia 64:694–704 Udayanga D, Liu X, McKenzie EHC et al (2011) The genus Phomopsis: biology, applications, species concepts and names of common phytopathogens. Fungal Divers 50:189 Udayanga D, Castlebury LA, Rossman AY et al (2015) The Diaporthe sojae species complex: phylogenetic re-assessment of pathogens associated with soybean, cucurbits and other field crops. Fungal Biol 119:383–407 Urtiaga R (1986) Indice de enfermedades en plantas de Venezuela y Cuba. Impresos en Impresos Nuevo Siglo. S.R.L., Barquisimeto, Venezuela Urtiaga R (2004) Host index of plant diseases and disorders from Venezuela—Addendum. Lara, Venezuela Urtiaga R, Braun U (2013) New species and new records of cercosporoid hyphomycetes from Cuba and Venezuela (Part 2). Mycosphere 4:174–214 Váczy KZ, Németh MZ, Csikós A et al (2018) Dothiorella omnivora isolated from grapevine with trunk disease symptoms in Hungary. Eur J Plant Pathol 150:817–824 Vaisakh MN, Pandey A (2012) The invasive weed with healing properties: a review on Chromolaena odorata. Int J Pharm Sci Res 3:80–83 Valenzuela-Lopez N, Sutton DA, Cano-Lira JF et al (2017) Coelomycetous fungi in the clinical setting: morphological convergence and cryptic diversity. J Clin Microbiol 55:552–567 Valenzuela-Lopez N, Cano-Lira JF, Guarro J et al (2018) Coelomycetous Dothideomycetes with emphasis on the families Cucurbitariaceae and Didymellaceae. Stud Mycol 90:1–69 Venugopal P, Junninen K, Linnakoski R et al (2016) Climate and wood quality have decayer-specific effects on fungal wood decomposition. For Ecol Manag 360:341–351 Venugopal P, Junninen K, Edman M, Kouki J (2017) Assemblage composition of fungal wood-decay species has a major influence on how climate and wood quality modify decomposition. FEMS Microbiol Ecol 93:1–8 Verkley GJM, Quaedvlieg W, Shin HD, Crous PW (2013) A new approach to species delimitation in Septoria. Stud Mycol 75:213–305 Videira SIR, Groenewald JZ, Nakashima C et al (2017) Mycosphaerellaceae—chaos or clarity? Stud Mycol 87:257–421 Viegas AP (1961) Indice de Fungos da America do SuI. Instituto Agronomico de Campinas, Campinas Vijayaraghavan K, Mohamed Ali S, Maruthi R (2013) Studies on phytochemical screening and antioxidant activity of Chromolaena odorata and Annona squamosal. Int J Innov Res Sci Eng Technol 2:7315–7321 Vilgalys R, Hester M (1990) Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. J Bacteriol 172:4239–4246 Vital PG, Rivera WL (2009) Antimicrobial activity and cytotoxicity of Chromolaena odorata (L. f.) King and Robinson and Uncaria perrottetii (A. Rich) Merr. extracts. J Med Plants Res 3:511–518 von Arx JA, Müller E (1954) Die Gattungen der amerosporen Pyrenomyceten. Beitr Kryptogamenfl Schweiz 11:1–434 von Arx JA, Müller E (1975) A re-evaluation of the bitunicate Ascomycetes with keys to families and genera. Stud Mycol Baarn 9:1–159 von Höhnel F (1917) System der Diaportheen. Ber Dtsch Bot Gesellschaft 35:631–638 von Höhnel F (1924) Studien über Hyphomyzeten. Zent Bakteriol Parasitenkd Abt 2:1–26 Wallen RM, Perlin MH (2018) An overview of the function and maintenance of sexual reproduction in dikaryotic fungi. Front Microbiol 9:503 Walsh TJ, Groll A, Hiemenz J et al (2004) Infections due to emerging and uncommon medically important fungal pathogens. Clin Microbiol Infect 10(Suppl 1):48–66 Wanasinghe DN, Jones EBG, Camporesi E et al (2016) Taxonomy and phylogeny of Laburnicola gen. nov. and Paramassariosphaeria gen. nov. (Didymosphaeriaceae, Massarineae, Pleosporales). Fungal Biol 120:1354–1373 Wanasinghe DN, Hyde KD, Jeewon R et al (2017) Phylogenetic revision of Camarosporium (Pleosporineae, Dothideomycetes) and allied genera. Stud Mycol 87:207–256 Wanasinghe DN, Phukhamsakda C, Hyde KD et al (2018) Fungal diversity notes 709–839: taxonomic and phylogenetic contributions to fungal taxa with an emphasis on fungi on Rosaceae. Fungal Divers 89:1–236 Wang Y, Guo LD, Hyde KD (2005) Taxonomic placement of sterile morphotypes of endophytic fungi from Pinus tabulaeformis (Pinaceae) in northeast China based on rDNA sequences. Fungal Divers 20:235–260 13 Fungal Diversity Wang Y, Hyde KD, McKenzie EHC et al (2015) Overview of Stachybotrys (Memnoniella) and current species status. Fungal Divers 71:17–83 Wang M, Liu F, Crous PW, Cai L (2017) Phylogenetic reassessment of Nigrospora: Ubiquitous endophytes, plant and human pathogens. Persoonia 39:118–142 Wang M, Tan XM, Liu F, Cai L (2018) Eight new Arthrinium species from China. MycoKeys 34:1–24 Watson W (1929) The classification of lichens. II. New Phytol 28:85–116 Wenneker M, Joosten NN, Luckerhoff LLP (2013) Use of (pulsed) UV-C light to control spore germination and mycelial growth of storage diseases causing fungi, and effect on control of storage rot in apples and pears. In: Ioratti C, Altindisli F, Borve J, Escudero-Colomar L, Lucchi A, Molinari F (eds) IOBC/wprs bulletin, pp 389–393 White TJ, Bruns T, Lee S, Taylor J (1990) Amplification and direct se- quencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (eds) PCR protocols: a guide to methods and applications. Academic, San Diego, pp 315–322 Wijayawardene N, McKenzie EHC, Hyde KD (2012) Towards incorporating anamorphic fungi in a natural classification—checklist and notes for 2011. Mycosphere 3:157–228 Wijayawardene NN, Crous PW, Kirk PM et al (2014a) Naming and outline of Dothideomycetes-2014 including proposals for the protection or suppression of generic names. Fungal Divers 69:1–55 Wijayawardene NN, Hyde KD, Bhat DJ et al (2014b) Camarosporiumlike species are polyphyletic in pleosporales; introducing Paracamarosporium and Pseudocamarosporium gen. nov. in Montagnulaceae. Cryptogam. Mycology 35:177–198 Wijayawardene N, Hyde KD, Tibpromma S et al (2017a) Towards incorporating asexual fungi in a natural classification: checklist and notes 2012–2016. Mycosphere 8:1457–1555 Wijayawardene NN, Hyde KD, Rajeshkumar KC et al (2017b) Notes for genera: Ascomycota. Fungal Divers 86:1–594 Wijayawardene NN, Hyde KD, Lumbsch HT et al (2018) Outline of Ascomycota: 2017. Fungal Divers 88:167–263 Williams TH, Liu PSW (1976) A host list of plant diseases in Sabah, Malaysia. Phytopathol Pap 19:1–67 Wong MKM, Hyde KD (2001) Diversity of fungi on six species of Gramineae and one species of Cyperaceae in Hong Kong. Mycol Res 105:1485–1491 Wu HX, Jaklitsch WM, Voglmayr H et al (2011a) Epitypification, morphology and phylogeny of Tothia fuscella. Mycotaxon 118:203–211 Wu HX, Schoch CL, Boonmee S et al (2011b) A reappraisal of Microthyriaceae. Fungal Divers 51:189–248 Yacharoen S, Tian Q, Chomnunti P et al (2015) Patellariaceae revisited. Mycosphere 6:290–326 Yahya MFZR, Ibrahim MSA, Zawawi WHAWM, Hamid UMA (2014) Biofilm killing effects of Chromolaena odorata extracts against Pseudomonas aeruginosa. Res J Phytochem 8:64–73 Yan Y, Yuan Q, Tang J et al (2018) Colletotrichum higginsianum as a model for understanding host–pathogen interactions: a review. Int J Mol Sci 19:2142 Yang T, Groenewald JZ, Cheewangkoon R et al (2017) Families, genera, and species of Botryosphaeriales. Fungal Biol 121:322–346 Yang Q, Du Z, Tian CM (2018a) Phylogeny and morphology reveal two new species of Diaporthe from Traditional Chinese Medicine in Northeast China. Phytotaxa 336:159–170 Yang Q, Fan X-L, Guarnaccia V, Tian C-M (2018b) High diversity of Diaporthe species associated with dieback diseases in China, with twelve new species described. MycoKeys 39:97–149 Yang CL, Xu XL, Wanasinghe DN et al (2019) Neostagonosporella sichuanensis gen. et sp. nov. (Phaeosphaeriaceae, Pleosporales) 13 on Phyllostachys heteroclada (Poaceae) from Sichuan Province, China. MycoKeys 46:119–150 Yao HY, He ZL, Wilson MJ, Campbell CD (2000) Microbial biomass and community structure in a sequence of soils with increasing fertility and changing land use. Microb Ecol 40:223–237 Yen JM (1968) Etude sur les champignons parasites du sud-est Asiatique X. Sixieme note sur les Cercospora de malaisie (1). Bull Soc Mycol France 84:5–18 Yen JM (1969) Etude sur les champignons parasites du sud-est asiatique XI. Quelques especes de Sphaeriales de Malaisie. Cah Pac 13:271–280 Yen JM (1974) Les Cercospora de Cote D’Ivoire. Bull Soc Mycol France 90:307–324 Yu X, He T, Zhao J, Li Q (2014) Invasion genetics of Chromolaena odorata (Asteraceae): extremely low diversity across Asia. Biol Invasions 16:2351–2366 Zachariades C, Day M, Muniappan R, Reddy G (2009) Chromolaena odorata (L.) King and Robinson (Asteraceae). In: Muniappan R, Reddy G, Raman A (eds) Biological control of tropical weeds using arthropods. Cambridge University Press, Cambridge, pp 130–162 Zachariades C, Janse Van Rensburg S, Witt A (2013) Recent spread and new records of Chromolaena odorata in Africa. In: Zachariades C, Strathie LW, Day MD, Muniappan R (eds) Proceedings of the Eighth International Workshop on Biological Control and Management of Chromolaena odorata and other Eupatorieae, 1–2 November 2010. Nairobi, Kenya. ARC-PPRI, Pretoria, pp 20–27 Zhang YB, Zhuang WY (2017) Four new species of Trichoderma with hyaline ascospores from southwest China. Mycosphere 8:1914–1929 Zhang YB, Zhuang WY (2018) New species of Trichoderma in the Harzianum, Longibrachiatum and Viride clades. Phytotaxa 379:131–142 Zhang H, Hyde KD, Mckenzie EHC et al (2012a) Sequence data reveals phylogenetic affinities of Acrocalymma aquatica sp. nov., Aquasubmersa mircensis gen. et sp. nov. and Clohesyomyces aquaticus (freshwater coelomycetes). Cryptogam. Mycology 33:333–346 Zhang Y, Crous PW, Schoch CL et al (2012b) Pleosporales. Fungal Divers 53:1–221 Zhang J, Bayram Akcapinar G, Atanasova L et al (2016) The neutral metallopeptidase NMP1 of Trichoderma guizhouense is required for mycotrophy and self-defence. Environ Microbiol 18:580–597 Zhang JF, Liu JK, Hyde KD et al (2017) Two new species of Dyfrolomyces (Dyfrolomycetaceae, Dothideomycetes) from karst landforms. Phytotaxa 313:267–277 Zhang S, Xu B, Zhang J, Gan Y (2018) Identification of the antifungal activity of Trichoderma longibrachiatum T6 and assessment of bioactive substances in controlling phytopathgens. Pestic Biochem Physiol 147:59–66 Zhang JF, Liu JK, Jeewon R et al (2019) Fungi from Asian Karst formations III. Molecular and morphological characterization reveal new taxa in Phaeosphaeriaceae. Mycosphere 10:202–220 Zhao YM, Deng CR, Chen X (1990) Arthrinium phaeospermum causing dermatomycosis, a new record of China. Acta Mycol Sin 9:232–235 Zhao YZ, Zhang ZF, Cai L et al (2018) Four new filamentous fungal species from newly-collected and hivestored bee pollen. Mycosphere 9:1089–1116 Zhaxybayeva O, Gogarten JP (2002) Bootstrap, Bayesian probability and maximum likelihood mapping: exploring new tools for comparative genome analyses. BMC Genom 3:4 Zhou D, Hyde KD (2001) Host-specificity, host-exclusivity, and hostrecurrence in saprobic fungi. Mycol Res 105:1449–1457 Fungal Diversity Zhu ZX, Xu HX, Zhuang WY, Li Y (2017) Two new green-spored species of Trichoderma (Sordariomycetes, Ascomycota) and their phylogenetic positions. MycoKeys 26:61–75 Zhu HY, Tian CM, Fan XL (2018) Studies of botryosphaerialean fungi associated with canker and dieback of tree hosts in Dongling Mountain of China. Phytotaxa 348:63–76 Zhuang WY (2001) Higher fungi of tropical China. Mycotaxon Ltd, New York Zhuang WY (2005) Fungi of northwestern China. Mycotaxon Ltd, Ithaca Affiliations Ausana Mapook1,2,3,4 · Kevin D. Hyde1,3,5 · Eric H. C. McKenzie6 · E. B. Gareth Jones7 · D. Jayarama Bhat8 · Rajesh Jeewon9 · Marc Stadler4 · Milan C. Samarakoon3 · Maitree Malaithong2 · Benjawan Tanunchai2 · François Buscot2,11 · Tesfaye Wubet2,10,11 · Witoon Purahong2 1 6 Manaaki Whenua-Landcare Research, Private Bag 92170, Auckland, New Zealand 7 Department of Soil Ecology, UFZ-Helmholtz Centre for Environmental Research, Theodor-Lieser-Str. 4, 06120 Halle (Saale), Germany Department of Botany and Microbiology, College of Science, King Saud University, P.O Box 2455, Riyadh 11451, Kingdom of Saudi Arabia 8 3 No. 128/1-J, Azad Co-Op Housing Society, Curca, Goa Velha 403108, India Center of Excellence in Fungal Research, Mae Fah Luang University, Chiang Rai 57100, Thailand 9 4 Department of Health Sciences, Faculty of Science, University of Mauritius, Reduit 80837, Mauritius Department Microbial Drugs, Helmholtz Centre for Infection Research, and German Centre for Infection Research (DZIF), Partner Site Hannover-Braunschweig, Inhoffenstrasse 7, 38124 Brunswick, Germany 10 Present Address: Department of Community Ecology, UFZ-Helmholtz Centre for Environmental Research, Theodor-Lieser-Str. 4, 06120 Halle (Saale), Germany 5 Institute of Plant Health, Zhongkai University of Agriculture and Engineering, Haizhu District, Guangzhou 510225, People’s Republic of China 11 German Centre for Integrative Biodiversity Research (iDiv), Halle-Jena-Leipzig, Deutscher Platz 5e, 04103 Leipzig, Germany 2 Key Laboratory for Plant Biodiversity and Biogeography of East Asia (KLPB), Kunming Institute of Botany, Chinese Academy of Science Kunming, Kunming 650201, Yunnan, People’s Republic of China 13