Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Karst caves are distinctly characterised by darkness, low to moderate temperatures, high humidity, and scarcity of organic matter. During the years of 2014-2015, we explored the mycobiota in two unnamed Karst caves in Guizhou province, China, and obtained 563 fungal strains via the dilution plate method. Preliminary ITS analyses of these strains suggested that they belonged to 246 species in 116 genera, while 23.5 % were not identified to species level. Among these species, 85.8 % (211 species) belonged to Ascomycota; 7.3 % (18 species) belonged to Basidiomycota; 6.9 % (17 species) belonged to Mucoromycotina. The majority of these species have been previously known from other environments, mostly from plants or animals as pathogens, endophytes or via a mycorrhizal association. We also found that 59 % of these species were discovered for the first time from Karst caves, including 20 new species that are described in this paper. The phylogenetic tree based on LSU sequences revealed 20 new species were distributed in six different orders. In addition, ITS or multi-locus sequences were employed to infer the phylogenetic relationships of new taxa with closely related allies. We conclude that Karst caves encompass a high fungal diversity, including a number of previously unknown species. Novel species described include: Amphichorda guana, Auxarthronopsis guizhouensis, Biscogniauxia petrensis, Cladorrhinum globisporum, Collariella quadrum, Gymnoascus exasperatus, Humicola limonisporum, Metapochonia variabilis, Microascus anfractus, Microascus globulosus, Microdochium chrysanthemoides, Paracremonium variiforme, Pectinotrichum chinense, Phaeosphaeria fusispora, Ramophialophora globispora, Ramophialophora petraea, Scopulariopsis crassa, Simplicillium calcicola, Volutella aeria, and Wardomycopsis longicatenata.

Free full text 


Logo of persooniaLink to Publisher's site
Persoonia. 2017 Dec; 39: 1–31.
Published online 2017 May 29. https://doi.org/10.3767/persoonia.2017.39.01
PMCID: PMC5832949
PMID: 29503468

Culturable mycobiota from Karst caves in China, with descriptions of 20 new species

Abstract

Karst caves are distinctly characterised by darkness, low to moderate temperatures, high humidity, and scarcity of organic matter. During the years of 2014–2015, we explored the mycobiota in two unnamed Karst caves in Guizhou province, China, and obtained 563 fungal strains via the dilution plate method. Preliminary ITS analyses of these strains suggested that they belonged to 246 species in 116 genera, while 23.5 % were not identified to species level. Among these species, 85.8 % (211 species) belonged to Ascomycota; 7.3 % (18 species) belonged to Basidiomycota; 6.9 % (17 species) belonged to Mucoromycotina. The majority of these species have been previously known from other environments, mostly from plants or animals as pathogens, endophytes or via a mycorrhizal association. We also found that 59 % of these species were discovered for the first time from Karst caves, including 20 new species that are described in this paper. The phylogenetic tree based on LSU sequences revealed 20 new species were distributed in six different orders. In addition, ITS or multi-locus sequences were employed to infer the phylogenetic relationships of new taxa with closely related allies. We conclude that Karst caves encompass a high fungal diversity, including a number of previously unknown species. Novel species described include: Amphichorda guana, Auxarthronopsis guizhouensis, Biscogniauxia petrensis, Cladorrhinum globisporum, Collariella quadrum, Gymnoascus exasperatus, Humicola limonisporum, Metapochonia variabilis, Microascus anfractus, Microascus globulosus, Microdochium chrysanthemoides, Paracremonium variiforme, Pectinotrichum chinense, Phaeosphaeria fusispora, Ramophialophora globispora, Ramophialophora petraea, Scopulariopsis crassa, Simplicillium calcicola, Volutella aeria, and Wardomycopsis longicatenata.

Keywords: diversity, ITS DNA barcodes, morphology, systematics, troglobitic fungi

INTRODUCTION

Caves differ from the land surface habitats in their darkness, low to moderate temperatures, high humidity, and scarcity of organic matter (Gabriel & Northup 2013). Environmental conditions of the caves may be affected by a variety of factors such as the movement of water (streams or water seeps), air currents, visitors and chemolithoautotrophy (Hose et al. 2000, Barton & Jurado 2007, Gabriel & Northup 2013, Ortiz et al. 2014). All these factors may contribute to the microbial flora in caves (Ogórek et al. 2013). Fungi are an important part of cave microbiota, because they play an important role in the feeding strategies of cave fauna (Nováková 2009).

The earliest study of fungi in caves was published by Humboldt in 1794 as described in Dobat (1967), but unfortunately these data currently provide very little information (Vanderwolf et al. 2013). Lagarde (1913) studied several caves in Europe and described a new species, Ombrophila speluncarum, which has been believed to be a true troglobitic species. Studies on cave fungi during 1950s–1980s were mostly about animal pathogens, e.g., Histoplasma capsulatum (Ajello et al. 1960a, b, Al-Doory & Rhoades 1968, Di Salvo et al. 1969, Zamora 1977), Trichophyton mentagrophytes and other dermatophytes (Lurie & Borok 1955, Lurie & Way 1957, Kajihiro 1965).

Recent studies demonstrated that caves encompass a high diversity of fungi. In the research on Lechuguilla Cave, New Mexico, species from nine genera were isolated and Aspergillus and Penicillium were found to be the most common genera (Cunningham et al. 1995). Nagai et al. (1998) investigated the alkalophilic and alkali-tolerant fungi in two limestone caves in Japan, and obtained 52 species. Koilraj et al. (1999) investigated six different caves in India and obtained 35 sporulating fungi belonging to 18 genera and seven sterile fungi. Nováková (2009) isolated 195 species belonging to 73 genera from caves in Slovakia, including 92 species from bat droppings and guano. In total, more than 1 000 species of fungi in 528 genera have been documented from caves and mines worldwide by 2012 (Vanderwolf et al. 2013).

Fungal diversity in Karst caves has rarely been documented. Yunnan-Guizhou Plateau, located in Southwest China, is the largest and most complex developing Karst topography in the world (Zhou et al. 2007). During the past two years, fungal communities from two caves were investigated based on a culture dependent method. Samples of air, water, rock, soil, and organic litter were collected and used for isolation of fungi. Strains were identified based on morphological characters and phylogenetic affinities. Novel species are described, illustrated, and compared with similar species.

MATERIAL AND METHODS

Sampling sites

Suiyang county is located in Guizhou province, China, with a typical subtropical monsoon climate. The annual mean temperature is 13.5 °C, and the annual rainfall is 1116–1350 mm (Jiang et al. 2012).

Two unnamed Karst caves in Suiyang, herein named as Cave 1 (N28°12′629″ E107°13′639″) and Cave 2 (N28°12′599″ E107°13′661″), are located at the edge of Kuankuoshui National Natural Reserve. Both caves are zonal and horizontal and have one entrance hiding in the forest on a hillside (Fig. 1). The two caves are 500 m apart from each other, thus might belong to the same cave system with a subterranean river connection (Fig. 1). Two bat roosts were found in Cave 1, the first one was about 50 m deep and the second one was at the end of the cave (about 390 m deep). Only one bat roost was found in Cave 2 (about 50 m deep). Other animals were also found in both caves, such as pellucid tadpoles, small pellucid snails, and spiders (Fig. 1).

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g001.jpg

Visited caves. a. Entrance to Cave 1; b. stalactite; c. pool at the end of Cave 1; d. sampled rocks; e. soil sediment; f. pellucid tadpole; g. bat guano colonised by fungal mycelia; h–i. faeces from unknown animals.

The elevation of Cave 1 is 908 m, the length is 400 m, the humidity is 75–80 %, and the temperature is 21–22 °C. The elevation of Cave 2 is 930 m, the length is 750 m, the humidity is 75–85 %, and the temperature is 20–23 °C.

Sample collection

Samples of air, rock, soil, and water were collected along the two caves and preserved at 4 °C before transfer to the laboratory. From the entrance of the caves, each sampling site was c. 100 m distant from the next.

Air samples were collected using the Koch sedimentation method (Borda et al. 2004, Kuzmina et al. 2012). Three Petri dishes that contained 2 % potato-dextrose agar (PDA, Difco) were exposed to the atmosphere in the cave for 15 min at each sampling site, then sealed with Parafilm and placed in zip-locked plastic bags. Seeping, stream, and pool water was collected for 10 mL per sample, respectively, and kept in 15 mL sterile centrifuge tubes. Ten grams of soil were collected at shallow depth (1–5.0 cm) after removing the surface layer (c. 1 cm) from three sites of each location. Rock samples were collected and packed in zip-locked plastic bags according to Ruibal et al. (2005). At each sample site, five pieces of rock in different orientations were collected. Rocks that were apparently being colonised by fungi were also chipped off and collected along the caves. Organic litter, when discovered, were collected, as well as bat droppings, guano, animal dung, carcasses, and plant debris.

Isolation

Fungi were isolated following a modified dilution plate method (Zhang et al. 2015b). One gram of each soil and organic litter sample (1 mL for each water sample) was suspended in 9 mL sterile water in a 15 mL sterile centrifuge tube. The tubes were shaken with Vortex vibration meter thoroughly. The suspension was then diluted to a series of concentrations, i.e., 10−1, 10−2, 10−3, 10−4, 10−5, and 10−6. Diluted concentration of 10−3 and 10−4 appeared to be most convenient for colony pickup in the isolating process from organic litters, while that for water samples and soil samples were 10−1 and 10−2, respectively. Two hundred microliters suspensions from each concentration were spread onto PDA containing ampicillin (50 μg/mL) and streptomycin (50 μg/mL) with three replicates.

Rock samples were treated following the protocol of Ruibal et al. (2005) with some modifications. Firstly, the rock surface was washed with 95 % ethanol to eliminate the contamination from dust and airborne spores, and washed once with sterile water containing 0.1 % of Tween 20. The small pieces of rocks were then ground into powder using a mortar and pestle. Suspensions were made by adding sterilised water to the concentration of 10−1. Three different volumes of the rock powder suspension, i.e., 300, 500, and 1000 μL, were respectively placed onto three PDA plates supplemented with ampicillin (50 μg/mL) and streptomycin (50 μg/mL) (Ruibal et al. 2005, Selbmann et al. 2005, Collado et al. 2007).

All the plates were incubated at room temperature (23–25 °C) for 3–4 wk, and from which the single colonies were picked up and inoculated onto new PDA plates every 2 d. All fungal strains were stored at 4 °C for further studies.

Molecular analyses

Total genomic DNAs were extracted following a modified protocol of Doyle (1987). The large subunit (LSU) rDNA, the internal transcribed spacer regions and intervening 5.8S nrRNA gene (ITS), the translation elongation factor 1-alpha (EF-1α), β-tubulin (TUB), and RNA polymerase II second largest subunit (RPB2) regions were amplified using primer pairs LR0R/LR5 (Vilgalys & Hester 1990), ITS1/ITS4 (White et al. 1990), 983F/2218R (Rehner & Buckley 2005), Bt2a/Bt2b (Glass & Donaldson 1995), and RPB2-5F2/fRPB2-7cR (Liu et al. 1999, Sung et al. 2007), respectively. Amplification reactions were performed in a 25 μL reaction volume including 2.5 μL 10× PCR Buffer (Dingguo, Beijing, China), 2 mM MgCl2, 50 μM dNTPs, 0.1 μM of each forward and reverse primer, 0.5 U Taq DNA polymerase and 1–10 ng genomic DNA in amplifier (Dongsheng, EDC-810, China). PCR parameters were as follows: 94 °C for 10 min, followed by 35 cycles of denaturation at 94 °C for 30 s, annealing at a suitable temperature for 30 s, extension at 72 °C for 30 s and a final elongation step at 72 °C for 10 min. Annealing temperature for each gene were 54 °C for ITS, 51 °C for LSU, and 57 °C for EF-1α and TUB. Sequencing reactions were performed with the same primer pairs used for amplification by OmegaGenetcis Company Limited, Beijing, China.

All obtained strains were BLASTn searched in NCBI and assigned to potential genera and species. The strains whose ITS sequences had closest similarities below 97 % were recognised as potential new species and further identified through morphological characterisation and multi-locus phylogenetic analyses.

To reveal the order placements of new species described in this paper, an LSU tree was constructed. To reveal the phylogenetic relationships and taxonomic distinctions of novel species, analyses were performed based on ITS, LSU, and genetic markers recommended in recent publications, such as TUB and EF1-α. All the sequences were aligned using MAFFT (http://www.ebi.ac.uk/Tools/msa/mafft/) (Katoh & Toh 2010) and edited manually using MEGA v. 6 (Tamura et al. 2013). Individual alignments were then concatenated and used to construct the phylogenetic tree. Ambiguously aligned regions were excluded from the analysis.

Bayesian inference (BI) and Maximum Likelihood (ML) methods were used to construct the phylogenetic trees. For Bayesian analysis, the best fit model of evolution was estimated by jModelTest v. 2.1.7 (Guindon & Gascuel 2003, Darriba et al. 2012). Posterior probabilities (PP) (Rannala & Yang 1996, Zhaxybayeva & Gogarten 2002) were determined by Markov Chain Monte Carlo sampling (MCMC) in MrBayes v. 3.2.1 (Huelsenbeck & Ronquist 2001), using the estimated model of evolution. Six simultaneous Markov chains were run for 1 000 000 generations and trees were sampled every 100th generation (resulting 10 000 total trees). The first 2 000 trees, which represented the burn-in phase of the analyses, were discarded and the remaining 8 000 trees were used to calculate posterior probabilities (PP) in the majority rule consensus tree. The ML analyses were implemented using RAxML v. 7.0.3 (Stamatakis 2006) with 1 000 replicates under the GTR-GAMMA model. The robustness of branches was assessed by bootstrap analysis with 1 000 replicates. Trees were visualized in TreeView (Page 1996). All the sequences generated were deposited in GenBank (Table 1), the multi-locus alignment and tree in TreeBASE (submission no: 19811), and typifications in MycoBank (Crous et al. 2004).

Table 1

Strain numbers and sequence accession numbers of new species.

Species nameStrain number1Sequence accession number
ITSLSUTUBEF1-αRPB2
Amphichorda guanaCGMCC3.17908 TKU746665KU746711KU746757KX855211
CGMCC3.17909KU746666KU746712KU746758KX855212
Auxarthronopsis guizhouensisCGMCC3.17910 TKU746668KU746714KU746759KX855213
CGMCC3.17911KU746667KU746713KU746760KX855214
Biscogniauxia petrensisCGMCC3.17912 TKU746669KU746715KU746761KX855215
CGMCC3.17913KU746670KU746716KU746762KX855216
CGMCC3.17949KU746671KU746717KU746763KX855217
Cladorrhinum globisporumCGMCC3.17921 TKU746680KU746726KU746771KX855226
CGMCC3.17922KU746679KU746725KU746772KX855225
Collariella quadrumCGMCC3.17917 TKU746675KU746721KU746767KX855221KY575870
CGMCC3.17918KU746676KU746722KU746768KX855222KY575871
CGMCC3.17919KU746677KU746723KU746769KX855223KY575872
CGMCC3.17920KU746678KU746724KU746770KX855224KY575873
Gymnoascus exasperatusCGMCC3.17923 TKU746682KU746728KU746773KX855227
CGMCC3.17924KU746681KU746727KU746774KX855228
Humicola limonisporumCGMCC3.17914 TKU746672KU746718KU746764KX855218KY575867
CGMCC3.17915KU746673KU746719KU746765KX855219KY575868
CGMCC3.17916KU746674KU746720KU746766KX855220KY575869
Metapochonia variabilisCGMCC3.17925 TKU746684KU746730KU746775KX855229
CGMCC3.17926KU746683KU746729KU746776KX855230
Microascus anfractusCGMCC3.17950 TKU746686KU746732KU746777KX855231
CGMCC3.17951KU746685KU746731KU746778KX855232
Microascus globulosusCGMCC3.17927 TKU746688KU746734KU746779KX855233
CGMCC3.17928KU746687KU746733KU746780KX855234
Microdochium chrysanthemoidesCGMCC3.17929 TKU746690KU746736KU746781KX855235
CGMCC3.17930KU746689KU746735KU746782KX855236
Paracremonium variiformeCGMCC3.17931 TKU746691KU746737KU746783KX855237
CGMCC3.17932KU746692KU746738KU746784KX855238
CGMCC3.17933KU746693KU746739KU746785KX855239
CGMCC3.17934KU746694KU746740KU746786KX855240
Pectinotrichum chinenseCGMCC3.17935 TKU746695KU746741KU746787KX855241
CGMCC3.17936KU746696KU746742KU746788KX855242
Phaeosphaeria fusisporaCGMCC3.17937 TKU746698KU746744KU746789KX855243
CGMCC3.17938KU746697KU746743KU746790KX855244
Ramophialophora globisporaCGMCC3.17939 TKU746700KU746746KU746791KX855245
CGMCC3.17940KU746699KU746745KU746792KX855246
Ramophialophora petraeaCGMCC3.17952 TKU746702KU746748KU746793KX855247
CGMCC3.17953KU746701KU746747KU746794KX855248
Scopulariopsis crassaCGMCC3.17941 TKU746704KU746750KU746795KX855249
CGMCC3.17942KU746703KU746749KU746796KX855250
Simplicillium calcicolaCGMCC3.17943 TKU746706KU746752KU746797KX855252
CGMCC3.17944KU746705KU746751KU746798KX855251
Volutella aeriaCGMCC3.17945 TKU746708KU746754KU746799KX855253
CGMCC3.17946KU746707KU746753KU746800KX855254
Wardomycopsis longicatenataCGMCC3.17947 TKU746710KU746756KU746801KX855255
CGMCC3.17948KU746709KU746755KU746802KX855256

1 Ex-type strains are indicated with T.

Morphological studies

Strains of potentially new species were transferred to new plates of PDA and synthetic nutrient-poor agar (SNA; Nirenberg 1976) and were incubated at room temperature (23–25 °C). Colony characters and pigment production on PDA and SNA were examined after 10 d. Growth rates were measured after 7 d, while slow growing strains were measured after 10 d or even 8 wk. Cultures were examined periodically for the development of reproductive structures. Photomicrographs were taken using a Nikon 80i microscope with differential interference contrast. Measurements for each structure were made according to methods described by Liu et al. (2012). The dry cultures were deposited in the Herbarium of Microbiology, Academia Sinica (HMAS), while living cultures were deposited in the China General Microbiological Culture Collection Center (CGMCC) and LC Culture Collection (personal culture collection held in the laboratory of Dr Lei Cai).

RESULTS

In this study, 85 samples from Cave 1, and 115 samples from Cave 2 were collected and 563 fungal strains were isolated. These strains belong to 116 genera, and 246 species by employing a BLASTn search in GenBank using the ITS sequences. Among these species, 85.8 % (i.e., 211 species, 489 strains) belong to 98 genera of Ascomycota; 7.3 % (i.e., 18 species, 29 strains) belong to 13 genera of Basidiomycota; 6.9 % (i.e., 17 species, 45 strains) belong to four genera of Mucoromycotina (Table 2). The most common genera included: Aspergillus (7.7 %), Penicillium (7.7 %), Chaetomium (3.3 %), Mortierella (3.7 %), Trichoderma (3.7 %), Phoma (2.8 %), Mucor (2.4 %), Arthrinium (2.4 %), Xylaria (2.0 %), and Fusarium (2.0 %) (Table 3). The most common species include Cephalotrichum verrucisporum (3.9 %), Aspergillus (As.) thesauricus (3.4 %), As. versicolor (3.2 %), Mortierella alpina (2.7 %), Eutypella scoparia (2.5 %), Chaetomium (Ch.) trigonosporum (2.3 %), Ch. nigricolor (1.6 %), Clonostachys rosea (1.6 %), Mortierella sp. 3 (1.3 %), Alternaria tenuissima (1.2 %), Amphichorda felina (1.2 %), Bionectria ochroleuca (1.2 %), As. candidus (1.1 %), Isaria fumosorosea (1.1 %), Lecanicillium fusisporum (1.1 %), and Trichocladium asperum (1.1 %). Among the obtained strains, 220 strains were isolated from Cave 1, belonging to 133 species in 75 genera, and 343 strains were isolated from Cave 2, belonging to 167 species in 83 genera. Forty-four genera were commonly isolated from both caves, and 30 and 39 genera were only obtained from Cave 1 and Cave 2, respectively. For the substrates of isolation, 200 strains from organic litter belong to 98 species in 61 genera; 143 strains from soil belong to 89 species in 52 genera; 96 strains from air belong to 64 species in 47 genera; 73 strains from rock belong to 54 species in 37 genera; and 51 strains from water belong to 39 species in 30 genera. In our study, 28 of the 116 genera and 111 of the 188 identified species (59.0 %) were reported for the first time from caves (Vanderwolf et al. 2013).

Table 2

An overview of fungal species isolated from the cave samples.

Fungal species1Cave 1
Cave 2
AirWaterSoilRockOrganic litterAirWaterSoilRockOrganic litter
Ascomycota
 Acremonium nepalense1
 A. persicinum1
 A. sp. 121
 A. sp. 21
 Acrodontium crateriforme1
 Acrostalagmus luteoalbus1
 Alternaria alternata11
 Al. tamaricis1
 Al. tenuissima3112
 Amphichorda felina7
 Am. guana2
 Arthrinium arundinis11
 Ar. malaysianum1
 Ar. marii1
 Ar. phaeospermum11
 Ar. sacchari1
 Ar. sp.111
 Arthroderma curreyi1
 Art. quadrifidum2
 Arthropsis hispanica1
 Aspergillus candidus114
 As. cavernicola121
 As. creber1
 As. flavus11
 As. fumigatus4
 As. niger11
 As. niveoglaucus1
 As. pragensis1
 As. reptans1
 As. ruber11
 As. sp.11
 As. spelunceus11
 As. tennesseensis1
 As. thesauricus266122
 As. tubingensis1
 As. ustus1
 As. versicolor11214
 As. wentii1
 Auxarthron thaxteri1
 Auxarthronopsis guizhouensis2
 Bionectria ochroleuca115
 Biscogniauxia petrensis3
 B. sp.1
 Calcarisporium sp.3
 Capnodium sp.1
 Cephalotrichum verrucisporum245119
 Ceratobasidium sp.1
 Chaetomidium arxii1
 Chaetomium bostrychodes13
 C. crispatum1
 C. globosum1
 C. murorum1
 C. nigricolor113211
 C. sp.1
 C. trigonosporum31117
 C. udagawae1
 Chalara holubovae1
 Chloridium sp.13
 Chrysosporium pseudomerdarium1
 Ch. sp. 12
 Ch. sp. 21
 Cladorrhinum globisporum2
 Cl. sp.1
 Cladosporium cladosporioides112
 Clad. sphaerospermum112
 Clonostachys rosea112113
 Clon. sp.1
 Collariella quadrum211
 Colletotrichum gloeosporioides11
 Coll. karstii1
 Coll. sp.2
 Corynespora sp.1
 Cylindrocarpon olidun112
 Cyl. sp.1
 Diaporthe phoenicicola11
 Doratomyces columnaris1
 D. nanus11
 D. sp.11
 Epicoccum nigrum1
 Eutypella scoparia132251
 Fusarium graminearum1
 F. merismoides1
 F. solani21
 F. sp.11
 F. verticillioides1
 Geotrichum candidum2
 Gibberella moniliformis1
 G. pulicaris1
 G. zeae2
 Gliomastix murorum11
 Gymnoascus exasperatus2
 Gymn. reesii2
 Gyrothrix sp.2
 Humicola limonisporum111
 Hypocrea citrina1
 Hypoxylon perforatum2
 Ilyonectria robusta1
 I. sp.1
 Isaria farinosa1
 Is. fumosorosea1122
 Is. tenuipes1
 Kernia sp.5
 Lecanicillium fusisporum6
 Leptosphaeria sp.1111
 Lophiostoma corticola1
 L. sp. 11
 L. sp. 21
 Massarina sp.1
 Metapochonia bulbilosa1
 M. rubescens1
 M. variabilis2
 Metarhizium anisopliae311
 Met. guizhouense1
 Microascus anfractus2
 Mic. chartarus1
 Mic. globulosus2
 Microdochium chrysanthemoides2
 Microsphaeropsis arundinis1
 Myriodontium keratinophilum1112
 Myrothecium sp.1
 Nectria haematococca111
 Nemania bipapillata1
 N. diffusa11
 Neonectria discophora1
 Neurospora intermedia1
 Paecilomyces fumosoroseus1121
 P. lilacinus1111
 P. sp.1
 Paracremonium variiforme31
 Paraphoma radicina1
 Pectinotrichum chinense11
 Penicillium buchwaldii1
 Pen. camemberti11
 Pen. chrysogenum2
 Pen. coprophilum1
 Pen. expansum11
 Pen. fellutanum1
 Pen. glabrum1
 Pen. herquei1
 Pen. inflatum1
 Pen. janthinellum112
 Pen. lividum1
 Pen. malachiteum1
 Pen. minioluteum2
 Pen. pancosmium11
 Pen. parvulum1
 Pen. pinophilum1
 Pen. sp.1
 Pen. thomii1
 Pen. urticae1
 Pestalotiopsis guepinii11
 Pest. microspora1
 Phaeoacremonium iranianum1
 Phaeosphaeria fusispora2
 Phialemonium sp.1
 Phoma herbarum1
 Ph. insulana1
 Ph. macrostoma1
 Ph. radicina1
 Ph. senecionis1
 Ph. sp. 11
 Ph. sp. 221
 Phomopsis sp.1
 Plectosphaerella cucumerina11
 Pl. sp.1
 Pleosporales sp.2–.
 Preussia aemulans1
 Protocrea farinosa1
 Pseudallescheria fimeti11
 Purpureocillium lilacinum11
 Ramophialophora globispora2
 R. petraea2
 Scopulariopsis crassa2
 S. sp.1
 Scutellinia sp.1
 Simplicillium calcicola2
 Stachybotrys chartarum1
 St. longispora1
 Staphylotrichum boninense1
 Staph. coccosporum1
 Staph. sp.1
 Stephanonectria keithii13
 Talaromyces sp.2
 Thielavia sp.113
 Togninia argentinensis1
 T. viticola1
 Tolypocladium cylindrosporum2
 Torula caligans3
 Tor. herbarum2111
 Trichocladium asperum2211
 Tr. sp.1
 Trichoderma atroviride1
 Trich. citrinoviride1
 Trich. hamatum112
 Trich. koningiopsis1
 Trich. lixii1
 Trich. longibrachiatum1
 Trich. rossicum1
 Trich. sp. 11
 Trich. sp. 2111
 Trichosporon akiyoshidainum1
 Trichos. laibachii11
 Trichos. rubrum1
 Verticillium sp.1
 Virgaria nigra1
 Volutella aeria2
 Wardomycopsis longicatenata2
 Xylaria arbuscula2
 X. schweinitzii1
 X. sp. 11
 X. sp. 22111
Basidiomycota
 Clitopilus kamaka2
 Coprinellus radians1
 Datronia mollis2
 Ganoderma gibbosum1
 Hyphodermella rosae1
 H. sp.1
 Peniophora cinerea1
 Penioph. limitata11
 Penioph. sp.111
 Periconia sp.1
 Phanerochaete sordida11
 Psathyrella candolleana11
 Ps. cf. gracilis1
 Rigidoporus sp.111
 Rig. vinctus11
 Schizophyllum commune1
 Tinctoporellus epimiltinus1
 Trametes versicolor11
Mucoromycotina
 Mortierella alpina321414
 Mort. horticola1
 Mort. hyalina1
 Mort. indohii2
 Mort. minutissima2
 Mort. sp. 11
 Mort. sp. 21
 Mort. sp. 311233
 Mucor flavus1
 Muc. hiemalis1
 Muc. moelleri1
 Muc. racemosus11
 Muc. sp. 11
 Muc. sp. 21
 Muc. sp. 31
 Rhizomucor variabilis11
 Rhizopus oryzae11

1 Names in bold signify new species described in this study.

Table 3

Most common genera (≥ 5 species) obtained from karst caves.

GenusSpeciesStrains
Aspergillus1977
Arthrinium610
Chaetomium831
Fusarium58
Mortierella934
Mucor67
Penicillium1927
Phoma79
Trichoderma914
Xylaria59

The LSU phylogenetic tree (Fig. 2) showed that our 20 new species (marked with bold font) scattered in six different orders, i.e., Hypocreales, Microascales, Onygenales, Pleosporales, Sordariales, and Xylariales. Trees presenting the phylogenetic relationships and taxonomic distinction of each species have been deposited in MycoBank. Significant ML bootstrap values (≥ 70 %) and Bayesian posterior probabilities (≥ 75 %) are shown in the phylogenetic tree.

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g002.jpg

Maximum likelihood (ML) tree based on LSU sequences showing the order placements of new species described in this study. ML bootstrap values (≥ 70 %) and Bayesian posterior probability (≥ 75 %) are indicated along branches (ML/PP). The tree is rooted with Sarcoscypha coccinea (FF176859). Novel species are indicated in bold font and the orders are shown on the right side of the figure.

Taxonomy

Amphichorda guana Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818245; Fig. 3

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g003.jpg

Amphichorda guana (from ex-holotype CGMCC 3.17908). a. A. guana on bat guano; b–c. upper and reverse views of cultures on PDA and SNA 10 d after inoculation; d. synnemata; e–f. conidiophores and conidia; g. conidia. — Scale bars: e–g = 10 μm.

Etymology. Referring to the material (bat guano) from which this fungus was isolated.

Colonies on PDA attaining 14–18 mm diam after 14 d, dense, slightly convex, fimbriate, white to yellowish, with a white margin. Yellowish exudate usually appeared on old colonies. Reverse white at first, slowly becoming pale yellow. Colonies on SNA 13–21 mm diam after 14 d, margin entire, white, mycelia sparse. Reverse white.

Vegetative hyphae hyaline, septate, smooth-walled, 1.5–3.5 μm diam, sometimes swollen, up to 7 μm diam. Synnemata arising in the centre part of the colony on PDA, up to 15 mm high and 1–3 mm wide, white, cylindrical, tomentose, occasionally branched at the apex. Conidiophores arising laterally from hyphae, straight or slightly curved. Conidiogenous cells mostly borne on conidiophores, occasionally in simple whorls on lateral branches from the mycelia, fusiform or ellipsoidal, straight or irregularly bent, 7–10 × 2–3 μm. Conidia holoblastic, solitary or clumped, hyaline, smooth, broadly ellipsoid to subglobose, unicellular, 4.5–5.5 × 3.5–5 μm (mean = 5.0 ± 0.3 × 3.9 ± 0.3 μm, n = 20). Chlamydospores not observed.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, on bat guano, 19 July 2014, X. Zhou (HMAS 246919 holotype designated here, ex-type living culture CGMCC 3.17908 = LC5815); ibid., CGMCC 3.17909 = LC5819.

Notes — This species should be classified in the genus Amphichorda because of its long white synnemata and flask-shaped conidiogenous cells from which 1-celled, terminal holoblastic conidia are produced (De Hoog 1972). Amphichorda was established by Fries (1825) and currently comprises only one species, A. felina (syn. Beauveria felina) (Seifert et al. 2011). Amphichorda guana can be distinguished from A. felina by its larger conidia (4.5–5.5 × 3.5–4.5 μm vs 3.5–4.0 × 2.5–3 μm). Amphichorda is similar to Beauveria in morphology, while Beauveria is differentiated by its elongate conidiogenous cells with apical denticulate rachis (Rehner et al. 2011, Chen et al. 2013) and the phylogeny based on ITS sequences showed that species of Amphichorda clustered in a distinct clade distant from Beauveria (phylogenetic tree deposited in MycoBank: MB 818245).

Auxarthronopsis guizhouensis Z.F. Zhang & L. Cai, sp. nov. — MycoBank MB818246; Fig. 4

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g004.jpg

Auxarthronopsis guizhouensis (from ex-holotype CGMCC 3.17910). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c–g. arthroconidia and aleurioconidia (d–g. arthroconidia and aleurioconidia in cotton blue). — Scale bars: c–g = 10 μm.

Etymology. Referring to the province where this fungus was collected.

Colonies on PDA 25–31 mm diam after 21 d, felty to cottony, flat, margin entire or dentate, white to yellow-brown. Reverse pale yellow to brown. Colonies on SNA 17–24 mm diam after 21 d, colourless, mycelia extremely scarce. Reverse colourless.

Vegetative hyphae hyaline, septate, branched, smooth-walled, 1–3 μm wide, sometimes swollen, up to 8.0 μm wide. Fertile hyphae hyaline, occasionally branched. Conidia abundant, most arthric, intercalary or few terminal, hyaline, unicellular, solitary, subglobose, ellipsoidal, cylindrical or pyriform, 3.5–9.5 × 2–4.5 μm (mean = 6.1 ± 1.5 × 3.2 ± 0.5 μm, n = 30), frequently separated by 1–3 autolytic connective cells, smooth- and thick-walled; some aleurioconidial, subglobose or ellipsoidal, sessile or extremely short stalked, 3.5–7 × 2–4.5 μm (mean = 4.9 ± 0.8 × 3.1 ± 0.6 μm, n = 20).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, from air, 19 July 2014, Z.F. Zhang (HMAS 246920 holotype designated here, ex-type living culture CGMCC 3.17910 = LC5705); ibid., CGMCC 3.17911= LC6219.

Notes — The genus Auxarthronopsis was established by Sharma et al. (2013) and currently comprises only one species, A. bandhavgarhensis. Our isolates are characterised by the intercalary or terminal, hyaline, solitary and aseptate arthric conidia, separated by autolytic connective cells, which is in good agreement with the morphological circumscription of Auxarthronopsis (Sharma et al. 2013). Based on the BLASTn research, the closest hit using ITS sequence is the type of A. bandhavgarhensis, NFCCI 2185 (HQ164436, identity = 86 %). Although A. bandhavgarhensis was not sufficiently described, we roughly measured the conidiogenous cells using f. 3 of Sharma et al. (2013) according to the provided scale bars. We concluded that conidiogenous cells of A. bandhavgarhensis were significantly longer (absent or up to 10 μm) than A. guizhouensis (absent or shorter than 2 μm). In addition, arthroconidia in A. guizhouensis are more abundant than A. bandhavgarhensis, while the aleurioconidia are less abundant.

Biscogniauxia petrensis Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818247; Fig. 5

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g005.jpg

Biscogniauxia petrensis (from ex-holotype CGMCC 3.17912). a–b. Upper and reverse views of cultures on PDA and SNA 14 d after inoculation; c. exudate; d. conidiomata under stereomicroscope; e–g. conidiophores and conidia; h. conidia. — Scale bars: e–h = 10 μm.

Etymology. Referring to the material it was isolated from, rock.

Colonies on PDA attaining 80–85 mm diam within 10 d, cottony to woolly, whitish to light pink, aerial mycelia abundant. Red droplets secretions appeared within 2 wk. Reverse yellowish to red-brown. Colonies on SNA attaining 80–85 mm diam within 10 d, fascicular, cottony to woolly, pink-white. Reverse pink-white.

Vegetative hyphae hyaline to brown, septate, branched, thin-walled aerial mycelia abundant. Conidiophores hyaline to slightly yellowish, rough-walled, composed of main axis, 3–4.5 μm diam, and sometimes one or more major branches, with conidiogenous cells arising terminally or laterally. Conidiogenous cells hyaline, swollen at the apex and with conidial secession scars, thin- and rough-walled, cylindrical to oblong, 7–13 × 3–4.5 μm. Conidia holoblastic, unicellular, hyaline, smooth, ovoid to clavate, 4.5–7.5 × 2.5–4.5 μm (mean = 5.7 ± 0.6 × 3.3 ± 0.4 μm, n = 35), with obtuse tip and acute truncated base.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, rock, 19 July 2014, Z.F. Zhang (HMAS 246921 holotype designated here, ex-type living culture CGMCC 3.17912 = LC5697); ibid., CGMCC 3.17913 = LC5698; ibid., CGMCC 3.17949 = LC5751.

Notes — Morphological characteristics of this species fit well with the generic concept of Biscogniauxia, i.e., coarse, warty and brown conidiophores, swollen conidiogenous areas with conidial secession scars and holoblastically produced conidia (Ju et al. 1998). Three strains formed a distinct clade within the genus Biscogniauxia, and are closely related to B. capnodes (strain no.: CM-AT-015) (phylogenetic tree deposited in MycoBank: MB818247). While B. petrensis differs from B. capnodes in the slightly wider conidia and the hyaline to slightly yellowish conidiophores (yellowish to brownish in B. capnodes) (Ju et al. 1998).

Cladorrhinum globisporum Z.F. Zhang & L. Cai, sp. nov. — MycoBank MB818250; Fig. 6

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g006.jpg

Cladorrhinum globisporum (from ex-holotype CGMCC 3.17921). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c. sporodochium on SNA; d–h. phialides; i. conidia; j. coiled mycelia. — Scale bars: d–j = 10 μm.

Etymology. Referring to its globose conidia.

Colonies on PDA 59–71 mm diam after 10 d, cottony to fluffy, flat to slightly plicated, margin entire, pale grey to dark grey. Reverse dark grey to dark olivaceous. Colonies on SNA 60–65 mm diam after 10 d, margin entire, white, aerial mycelia sparse. Reverse white.

Vegetative hyphae hyaline to pale olivaceous, branched, septate, thin-walled, slightly rough, 1.5–4 μm diam, sometimes swollen, occasionally coiled. Microsclerotia not observed. Sporodochium forming on SNA within 50 d or longer, margin round, yellowish, scattered over entire colony. Fertile hyphae 2–3.5 μm wide, tufted, lateral branched, bearing terminal, lateral phialides or lateral phialidic openings of intercalary conidiogenous cells. Terminal and lateral phialides flask-shape, hyaline to pale olivaceous, straight or slightly sinuous, 8–18(–21) × 2–3.5 μm; intercalary conidiogenous openings with short flaring collarette, 2.5–5.0 × 1.5–2.5 μm. Conidia enteroblastic, hyaline, smooth- and thin-walled, globose, 2–3 μm diam (mean = 2.4 ± 0.3 μm, n = 40), guttulate, aggregated in slimy head.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, water, 19 July 2014, Q. Chen (HMAS 246924 holotype designated here, ex-type living culture CGMCC 3.17921 = LC5415); ibid., CGMCC 3.17922 = LC5370.

Notes — Our isolates are morphologically similar to the cladorrhinum-like anamorph of Cercophora and Podospora. The BLASTn search also showed that the closest hits using ITS sequence of these isolates are the sequences of Cercophora, Cladorrhinum, and Podospora species. Cladorrhinum and Phialophora are anamorph typified genera related to Cercophora and Podospora (Miller & Huhndorf 2005, Madrid et al. 2011, Kruys et al. 2015). Previous studies showed that these genera are morphologically similar but polyphyletic (Cai et al. 2005, 2006, Miller & Huhndorf 2005, Madrid et al. 2011, Kruys et al. 2015). Cladorrhinum can be distinguished from Phialophora by the relative abundance of intercalary phialides and the phylogenetic affinities (Mouchacca & Gams 1993, Madrid et al. 2011).

Phylogenetic analysis based on ITS and LSU showed that Cl. globisporum clustered in the clade of Cladorrhinum (phylogenetic tree deposited in MycoBank: MB818250) and the morphology fitted well to genus Cladorrhinum, which is characterised by the relative abundance of intercalary vs terminal phialides, the pigmentation of mycelia, and the conidial shape (Mouchacca & Gams 1993, Madrid et al. 2011). Cladorrhinum globisporum is morphologically similar to several species of Cladorrhinum producing globose to dacryoid conidia, e.g., Cl. bulbillosum, Cl. flexuosum, Cl. foecundissimum, Cl. samala, and the anamorphs of Cercophora striata and Podospora fimiseda (Madrid et al. 2011). However, Cl. globisporum does not produce blackish microsclerotia in culture, which are present in cultures of Cl. bulbillosum, and the anamorph of Ce. striata. Cladorrhinum globisporum differs from Cl. samala in the absence of dark, thick-walled setiform hyphae; from Cl. flexuosum in the rather regular conidiophores, which are strongly flexuous in Cl. flexuosum; from Cl. foecundissimum in the longer terminal and lateral phialides (8–18 μm vs 5.0–11 μm) and the globose conidia, which are dacryoid to almost globose in Cl. foecundissimum; from the asexual morph of P. fimiseda in its smaller conidia (2–3 μm vs 3–4 μm).

Collariella quadrum Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818249; Fig. 7

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g007.jpg

Collariella quadrum (from ex-holotype CGMCC 3.17917). a–b. Upper and reverse views of cultures on PDA and SNA 18 d after inoculation; c, e. ascomata; d. masses of ascospores; f–g. outer surface of peridium; h–i. ascomatal hairs; j–k. asci; l. ascospores. — Scale bars: c, e = 100 μm; h = 50 μm; f–g, i–l = 10 μm.

Etymology. Referring to the shape of its ascospores.

Colonies on PDA attaining 34–37 mm diam after 10 d, felty, margin entire, white to pale yellow. Reverse pale cream-yellow. Colonies on SNA attaining 40–42 mm diam after 10 d, flat, aerial mycelia sparse, white. Reverse white.

Vegetative hyphae hyaline, septate, branched, smooth-walled, 2–6 μm diam. Ascomata black, grey-green, subglobose, oval to fusiform, 250–500 μm high, 200–280 μm diam, with rounded base, ostiolate, neck unconspicuous. Peridium brown, comprised of textura angularis, arranged in a petaloid pattern. Ascomatal hair 260–450 μm long, 4–7 μm diam at base, tapering, unbranched, septate, straight at first, then straight below, spirally and loosely coiled upper, brown, verrucose. Asci fasciculate, clavate, eight-spored, long-stalked, 30–60 × 9.5–13 μm. Ascospores exuded as elongated cirrhi, biseriately arranged, pale olivaceous, square-pillow-shaped, quadrangular in frontal view, with obtuse angle, 4.5–5.5 × 4.5–5.5 × 4–4.5 μm (mean = 5.0 ± 0.1 × 4.9 ± 0.1 × 4.2 ± 0.2 μm, n = 30).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, soil, 19 July 2014, X. Zhou (HMAS 246923 holotype designated here, ex-type living culture CGMCC 3.17917 = LC5446); ibid., CGMCC 3.17918 = LC5693; ibid., CGMCC 3.17919 = LC5781; ibid., CGMCC 3.17920 = LC5782.

Notes — The genus Collariella was recently established to accommodate several species previously accommodated in Chaetomium, based on both phylogenetic and morphological data (Wang et al. 2016). Collariella quadrum is morphologically and phylogenetically most closely related to C. quadrangulata (phylogenetic tree deposited in MycoBank: MB818248). However, C. quadrum differs from C. quadrangulatum in producing smaller ascospores (4.5–5.5 × 4.5–5.5 × 4.0–4.5 μm vs 6.5–7.5 × 6–7 × 4–5 μm) exuded as elongated cirrhi (Fig. 7d).

Gymnoascus exasperatus Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818251; Fig. 8

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g008.jpg

Gymnoascus exasperatus (from ex-holotype CGMCC 3.17923). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c. fertile mycelial structure; d, f. terminal, lateral and intercalary conidia; e. racquet hyphae; g. conidia. — Scale bars: d = 20 μm; e–g = 10 μm.

Etymology. Referring to the texture of its conidial wall, rough.

Colonies on PDA attaining 17–21 mm diam after 10 d, felty, flat, plicate, margin entire, white to pale pink. Reverse plicate, white to pale pink. Colonies on SNA attaining 16–18 mm diam after 10 d, felty, annular, margin entire, white to pale yellow, aerial mycelia sparse. Reverse white to pale yellow.

Vegetative hyphae pale yellow to yellow, septate, branched, smooth or slightly rough, sometimes fascicular, 1.5–5 μm diam; racquet hyphae present, ‘racquet’ 10 μm wide. Fertile mycelia usually gathered into a special, superficial structure, where conidia borne mostly. Conidia terminal, lateral or intercalary, sessile or borne on short protrusions or side branches, solitary, frequently separated by one hyphal cell, subhyaline to pale brown, rough- and thin-walled; terminal and lateral conidia subglobose, obovoid to ellipsoidal, intercalary conidia cylindrical, 5–8 × 4–6 μm (mean = 6.4 ± 0.7 × 4.8 ± 0.5 μm, n = 30), with one or two sides truncated bases.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, bat guano, 19 July 2014, Z.F. Zhang (HMAS 246925 holotype designated here, ex-type living culture CGMCC 3.17923 = LC5640); ibid., CGMCC 3.17924 = LC6217.

Notes — Gymnoascus exasperatus was isolated from bat guano in the cave. Phylogenetically, it forms a distinct clade sister to G. reessii (CBS 410.72), the type species of Gymnoascus (Fig. 2, phylogenetic tree deposited in MycoBank: MB818251). Gymnoascus exasperates is unique in the genus as it is only known from the asexual morph. The sexual morph of G. exasperatus was not observed despite repeated attempts using OA, PDA, and SNA media, as well as human hair and nails as inducer (Orr & Kuehn 1972).

Humicola limonisporum Z.F. Zhang & L. Cai, sp. nov. — MycoBank MB818248; Fig. 9

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g009.jpg

Humicola limonisporum (from ex-holotype CGMCC 3.17914). a–b. Upper and reverse views of cultures on PDA and SNA 20 d after inoculation; c. ascomata; d. outer surface of peridium; e. ascomatal hairs; f–g. asci; h. ascospores; i. aleurioconidia. — Scale bars: c = 100 μm; f = 50 μm; d–e, g–i = 10 μm.

Etymology. Referring to its limoniform ascospores.

Colonies on PDA attaining 53–56 mm diam after 21 d, slightly raised near the centre, floccose, yellowish to yellow-brown. Reverse pale yellow to brown. Colonies on SNA attaining 69–72 mm diam after 21 d, flat, white, aerial mycelia sparse. Reverse white.

Vegetative hyphae hyaline, septate, branched, smooth-walled, 2.0–5.5 μm diam. Ascomata appeared after about 40 d, scattered over the colonies, brown to black-brown, oval to subglobose, 180–270 μm high, 120–200 μm diam, with rounded base, ostiolate, neck unconspicuous. Peridium brown, textura irregularis. Rhizoids well developed, septate, pale brown, 2.5–5.5 μm wide. Ascomatal hair 220–720 μm long, 2.0–5.0 μm diam at base, tapering, unbranched, septate, straight at base, spirally and loosely coiled upper, yellow to brown, verrucose. Asci fasciculate, clavate, eight-spored, long-stalked, 45–87 × 14.5–21 μm. Ascospores dark brown, thick-walled, limoniform to subglobose, 7.5–10.5 × 5.5–9.5 μm (mean = 9.1 ± 0.6 × 7.3 ± 1.1 μm, n = 50), with an apical germ pore. Aleurioconidia subhyaline to pale brown, subglobose to globose, solitary, unicellular, 8.5–13.5 μm diam (mean = 10.9 ± 1.1 μm, n = 45).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′599″ E107°13′661″, soil, 19 July 2014, X. Zhou (HMAS 246922 holotype designated here, ex-type living culture CGMCC 3.17914 = LC5610); ibid., CGMCC 3.17915= LC5707; ibid., CGMCC 3.17916 = LC5708.

Notes — Traditionally morphologically defined Chaetomium was heterogeneous and thus has been recently revised (Wang et al. 2016). Most previously reported Chaetomium species do not produce an asexual morph, but few species have a humicola-like anamorph. A distinct clade including the type of Humicola, i.e., H. fuscoatra (Traaen 1914), was recognised by Wang et al. (2016). In this study, H. limonisporum clustered in this clade, closely related to H. fuscoatra (the type species of Humicola) and H. olivacea (phylogenetic tree deposited in MycoBank: MB818248). Humicola limonisporum differs from H. olivacea in producing white colonies on SNA (grey olivaceous for H. olivacea); from H. fuscoatra in producing subglobose to globose aleurioconidia (subglobose to obpyriform for H. fuscoatra). The genus Humicola remains polyphyletic however, as most of the currently known species do not cluster with the type of the genus (Wang et al. 2016). Morphologically, the asexual morph of H. limonisporum is comparable to H. globosa. However, it differs in having yellow-brown colonies and hyaline somatic hyphae on PDA. In contrast, colonies of H. globosa on PDA are dark green to black and the somatic hyphae are predominantly brown.

Metapochonia variabilis Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818252; Fig. 10

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g010.jpg

Metapochonia variabilis (from ex-holotype CGMCC 3.17925). a–b. Upper and reverse views of cultures on PDA (10 d) and SNA (21 d); c, e. phialides; d. conidia aggregated in slimy head; f. conidia. — Scale bars: c–f = 10 μm.

Etymology. Referring to its various conidial shapes.

Colonies on PDA attaining 22–26 mm diam after 10 d, pulvinate, compact, sometimes plicated, white to pale brown. Reverse yellow-brown to brown. Colonies on SNA attaining 17–28 mm diam after 10 d, flat, margin entire, white to yellowish, aerial mycelia sparse. Reverse white to yellowish.

Vegetative hyphae hyaline, smooth-walled, branched, septate. Conidiophores arising from prostrate aerial hyphae, straight, hyaline, with 1–4 phialides per node lateral or in whorls of 3–6 phialides terminal, c. 2 μm diam. Phialides arising from aerial hyphae or conidiophore, slender, awl-shaped phialides, hyaline, 16–26.5 μm long, 1.5–2.5 μm diam at base, tapering toward the tip. Conidia single or aggregated in slimy heads, 1-celled, hyaline, falcate, fusiform, pyriform, ellipse to subglobose, or some other irregularly shapes, 3–6(–8) × 2–3.5 μm (mean = 4.9 ± 1.1 × 2.4 ± 0.3 μm, n = 30). Dictyochlamydospores not observed. Crystals absent.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, soil, 19 July 2014, X. Zhou (HMAS 246926 holotype designated here, ex-type living culture CGMCC 3.17925 = LC5717); ibid., CGMCC 3.17926 = LC6221.

Notes — Metapochonia was established by Kepler et al. (2014) and currently comprises five species. Metapochonia species are verticillium-like, producing conidia on slender, awl-shaped phialides that may be whorled or singular, and most species are also known to produce stalked, thick-walled, and multicellular dictyochlamydospores (Gams & Zare 2001, Zare & Gams 2007, Kepler et al. 2014). Based on the phylogenetic analysis of ITS sequences, our isolates clustered together with other Metapochonia species but formed a distinct clade with high support value (phylogenetic tree deposited in MycoBank: MB818252). Morphologically, M. variabilis differs from the closely related species M. bulbillosa in producing wider conidia (2–3.5 μm vs 1.2–2 μm); from M. gonioides in producing different shaped, larger conidia (3–6(–8) × 2–3.5 μm vs 1.8–2.5 μm diam).

Microascus anfractus Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818253; Fig. 11

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g011.jpg

Microascus anfractus (from ex-holotype CGMCC 3.17950). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c, e. coiled mycelia; d–f. conidiogenous cells and conidia; g. conidia; h–i. swollen mycelia. — Scale bars: c–h = 10 μm; i = 20 μm.

Etymology. Referring to its coiled mycelia.

Colonies on PDA 10–13 mm diam after 10 d, felty, compact, margin entire to undulate, plicated, low convex, pink to salmon. Reverse salmon. Colonies on SNA 9–12 mm diam after 10 d, compact, margin entire to undulate, white to yellowish, aerial mycelia sparse. Reverse white to yellowish.

Vegetative hyphae hyaline to pale brown, septate, branched, smooth-walled, 1–2.5 μm diam, sometimes coiled. The tip of semi-immerged mycelia sometimes swollen, globose, up to 20 μm diam. Conidiogenous cells borne laterally on aerial hyphae, solitary, hyaline, smooth-walled, lageniform, ampulliform or pyriform, straight or slightly curved, 7–13.5(–28.5) × 2–3.5 μm. Conidia formed in chains, hyaline, smooth- and thin-walled, ellipsoidal, fusiform to globose, 3.5–6 × 3.5–5 μm (mean = 5.0 ± 0.6 × 4.1 ± 0.5 μm, n = 30).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, plant debris, 19 July 2014, Z.F. Zhang (HMAS 246927 holotype designated here, ex-type living culture CGMCC 3.17950 = LC5843); ibid., CGMCC 3.17951 = LC6224.

Notes — Phylogenetically, M. anfractus nested within the Microascus clade based on ITS, LSU, TUB, and EF1-α sequences (phylogenetic tree deposited in MycoBank: MB818253) and its morphological characteristics fit well to this genus, i.e., ampulliform or lageniform conidiogenous cells and smooth- and thin-walled or finely rough- and thick-walled conidia (Sandoval-Denis et al. 2016).

Microascus anfractus is morphologically and phylogenetically comparable to three species in Microascus, i.e., M. campaniformis, M. cirrosus, and M. croci. The conidiogenous cells of M. anfractus arise from hyphae singly, while those of M. cirrosus arise from conidiophores and are arranged in whorls of three to five. Microascus anfractus can be distinguished from M. campaniformis and M. croci in its wider conidia (3.5–5 μm vs 2.5–3.5 μm for M. campaniformis and 2.5–3.5 μm for M. croci).

Microascus globulosus Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818254; Fig. 12

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g012.jpg

Microascus globulosus (from ex-holotype CGMCC 3.17927). a–b. Upper and reverse views of cultures on PDA (21 d) and SNA (42 d); c–f. conidiogenous cells and conidia in cotton blue; g–h. conidia. — Scale bars: c–h = 10 μm.

Etymology. Referring to its globose conidia.

Colonies on PDA attaining 29–35 mm diam after 15 d, felty, compact, plicated, umbonate, pink, salmon to rusty red, with white margin, aerial mycelia sparse. Reverse tawny to brown. Colonies on SNA attaining 17–19 mm after 15 d, flocculent, margin radially striate with lobate edge, conspicuously radial gaps, pale grey, aerial mycelia sparse. Reverse pale grey.

Vegetative hyphae hyaline, septate, branched, thin- and smooth-walled, 1.2–2.8 μm diam. Conidiogenous cells lateral or terminal on aerial hyphae, solitary, hyaline, smooth-walled, cylindrical, ampulliform or irregular shaped, erect or curved, constricted at base, occasionally branched, 5.5–24 × 1–3 μm. Conidia formed in chains, globose to subglobose, thick- and smooth-walled, hyaline to pale brown, 6.5–9 μm diam (mean = 7.5 ± 0.9 μm, n = 40).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, bat guano, 19 July 2014, Z.F. Zhang (HMAS 246928 holotype designated here, ex-type living culture CGMCC 3.17927 = LC5820); ibid., CGMCC 3.17928 = LC6223.

Notes — Microascus globulosus is phylogenetically closely related to M. chartarus based on ITS, LSU, TUB, and EF1-α sequences (phylogenetic tree deposited in MycoBank: MB818254). However, the conidia of M. chartarus are ovate, green-brown, and often with a pointed end, as compared to the globose to subglobose, hyaline to pale brown conidia of M. globulosus.

Microdochium chrysanthemoides Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818255; Fig. 13

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g013.jpg

Microdochium chrysanthemoides (from ex-holotype CGMCC 3.17929). a–b. Upper and reverse views of cultures on PDA (10 d) and SNA (14 d); c. sporodochia semi-submerged in agar (black arrow); d–e. sporodochia; f–g. conidiogenous cells from sporodochia; h–i. conidia forming on conidiophore directly (stained with cotton blue); j–k. conidia. — Scale bars: d–k = 10 μm.

Etymology. Referring to the shape of its sporodochium, chrysanthemum-like (Fig. 13d).

Colonies on PDA attaining 40–46 mm diam after 10 d, felty, compact, erose or dentate, white initially, then becoming yellowish with age. Exudate occasionally appeared on old sporodochia. Reverse yellowish to orange, due to the soluble pigment secreted. Colonies on SNA 55–57 mm diam after 10 d, entire, white, aerial mycelia sparse. Exudate absent. Reverse white.

Vegetative hyphae hyaline, abundant, branched, septate, thin-walled. Conidiophores aggregated in a sporodochium, or borne directly from the hyphae, hyaline, unbranched. Sporodochia appeared within 7 d or longer, yellowish to salmon, semi-submerged. Conidiophores borne from hyphae straight or slightly curved, aseptate, 21–58 × 1–1.5 μm, with conidia formed terminal or lateral, solitary or aggregated in a mass. Conidiogenous cells holoblastic, solitary, hyaline, apical, simple, ampulliform, lageniform, cylindrical to ellipsoidal, straight or bent, 5–12 × 3.0–4.5 μm; denticles not observed. Conidia aseptate, ellipsoid or allantoid, straight or curved, obtuse, guttulate in mature conidia, 4.5–7 × 2–3 μm (mean = 5.5 ± 0.7 × 2.7 ± 0.3 μm, n = 35).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, air, 19 July 2014, Z.F. Zhang (HMAS 246929 holotype designated here, ex-type living culture CGMCC 3.17929 = LC5363); ibid., CGMCC 3.17930 = LC5466.

Notes — Microdochium has been regarded as the asexual morph of Monographella (Zhang et al. 2015a, Hernández-Restrepo et al. 2016). Although these two genera were described in the same journal in 1924 (Petrak 1924, Sydow 1924), Microdochium has more species and has been more frequently used in literature (Hernández-Restrepo et al. 2016), and thus the name should be protected with the implementation of ‘one fungus one name’ approach (Hawksworth et al. 2011). Microdochium is characterised by its verticillate conidiophores, holoblastic, discrete, small papillate conoid conidiogenous cells and solitary, fusiform to subfalcate, hyaline conidia (Hernández-Restrepo et al. 2016). Microdochium chrysanthemoides is phylogenetically closely allied to M. neoqueenslandicum (CBS 445.95 and CBS 108926) and formed a distinct clade (phylogenetic tree deposited in MycoBank: MB818255). Morphologically, conidia of M. chrysanthemoides are ellipsoid or falcate, straight or curved, and guttulate, while that in M. neoqueenslandicum are consistently lunate, allantoid, curved and non-guttullate (Hernández-Restrepo et al. 2016).

Paracremonium variiforme Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818264; Fig. 14

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g014.jpg

Paracremonium variiforme (from ex-holotype CGMCC 3.17931). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c. sporulation on PDA under stereomicroscope; d–e. conidiophores, phialides and conidia; f. phialides and conidia in cotton blue; g. conidia. — Scale bars: d–e = 20 μm; f–g = 10 μm.

Etymology. Refers to the various shapes of the conidia.

Colonies on PDA attaining 43–49 mm diam after 21 d, flat, margin entire, milk-white to yellow-white, aerial mycelia sparse. Reverse milk-white to yellow-white. Colonies on SNA attaining 43–46 mm diam after 21 d, margin entire, white. Reverse white.

Vegetative hyphae hyaline, smooth- and thin-walled, septate, branched, 1.5–9.5 μm diam, inconspicuously swollen at the hyphal septa. Sporulation abundant, mostly phalacrogenous, varying to nematogenous. Conidiophores erect, simple or mostly branched, septate, bearing whorls of 2–4 conidiogenous cells. Conidiogenous cells terminal or lateral, straight, acicular or elongate-ampulliform, tapering towards apex, hyaline, 18–41 × 2–3.5 μm, with prominent periclinal thickening and inconspicuous collarette, 1–1.5 μm diam. Conidia unicellular, hyaline, clavate, ovoid or elliptical, thick- and smooth-walled, with slightly apiculate base, 9–14.5 × 4–6 μm (mean = 11.1 ± 1.3 × 4.9 ± 0.6 μm, n = 40). Chlamydospores not observed.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, water, 19 July 2014, Z.F. Zhang, L. Cai, Q. Chen & X. Zhou (HMAS 246930 holotype designated here, ex-type living culture CGMCC 3.17931 = LC5806); ibid., CGMCC 3.17932 = LC5809; ibid., CGMCC 3.17933 = LC5832; ibid., CGMCC 3.17934 = LC5837.

Notes — Phylogenetic analysis based on ITS, LSU and TUB sequences showed that our isolates clustered within the genus Paracremonium and formed a distinct clade (MB818264). According to the description of Lombard et al. (2015), Paracremonium is distinguished from other acremonium-like genera by the formation of sterile coils from which conidiophores radiate with inconspicuously swollen septa in the hyphae. However, sterile coils were not observed in our isolates. Paracremonium variiforme is phylogenetically most closely related to P. inflatum and P. contagium (phylogenetic tree deposited in MycoBank: MB818264), but could be easily differentiated from them by its branched conidiophores.

Pectinotrichum chinense Z.F. Zhang & L. Cai, sp. nov. — Myco-Bank MB818256; Fig. 15

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g015.jpg

Pectinotrichum chinense (from ex-holotype CGMCC 3.17935). a–b. Upper and reverse views of cultures on PDA and SNA 20 d after inoculation; c. aerial mycelia; d–e. conidiophores and conidiogenous cells; f. conidia. — Scale bars: d–f = 10 μm.

Etymology. Referring to the country where the fungus was firstly discovered.

Colonies on PDA 27–29 mm diam after 10 d, fluffy, flat, margin entire, white to yellow, powdery. Yellow pigment secreted. Reverse yellow to white from centre to margin. Colonies on SNA 21–22 mm diam after 10 d, yellowish, aerial mycelia extremely sparse. Yellow pigment secreted. Reverse pale yellow.

Vegetative hyphae hyaline, 2–3.2 μm diam. Conidiophores absent or reduced to conidiogenous cells, formed on aerial hyphae, hyaline, straight or slightly curved, 20–45 × 1.5–3 μm or longer, irregularly branched, sometimes two or three times branched, with conidia formed on branches terminal or lateral. Conidia borne on conidiophores or seldom on aerial hyphae directly or with a short stem, 1-celled, hyaline, smooth-walled, cylindrical, oval or pyriform, base rounded or with inconspicuous scars, 3.5–7.5 × 1–2.5 μm (mean = 4.4 ± 1.1 × 1.7 ± 0.3 μm, n = 25).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, soil, 19 July 2014, X. Zhou (HMAS 246931 holotype designated here, ex-type living culture CGMCC 3.17935 = LC5811); ibid., CGMCC 3.17936 = LC5824.

Notes — Pectinotrichum was established by Varsavsky & Orr (1971) and currently contains only one species, P. llanense, which was reported as a keratinophilic fungus and usually isolated via the hair-bait technique. Based on a BLASTn search the closest hit using ITS sequence of P. chinense is that from the ex-type strain of P. llanense CBS 882.71 (NR119467, identity = 96 %), and the other BLAST results are all below 90 % identity. Pectinotrichum chinense differs from P. llanense in its narrower, sometimes cylindrical conidia (3.5–7.5 × 1–2.5 μm, mean = 4.4 ± 1.1 × 1.7 ± 0.3 μm, vs 4–6.5 × 2–3 μm) (Van Oorschot 1980).

Phaeosphaeria fusispora Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818257; Fig. 16

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g016.jpg

Phaeosphaeria fusispora (from ex-holotype CGMCC 3.17937). a–b. Upper and reverse views of cultures on PDA and SNA 28 d after inoculation; c. exudates; d. section of ascomata; e–i. asci and ascospores. — Scale bars: d, f = 50 μm; e, g = 20 μm; h–i = 10 μm.

Etymology. Referring to its fusiform ascospores.

Colonies on PDA attaining 36–48 mm diam after 3 wk, felty, flat, margin entire, white to olivaceous from edge to centre. Yellow-brown to black brown secretions exuded. Reverse grey-green, with pale yellow margin. Colonies on SNA attaining 27–29 mm diam after 3 wk, felty, margin entire, white to light brown, aerial mycelia sparse. Reverse white to olivaceous, with an olivaceous circle.

Vegetative hyphae hyaline to brown, septate, branched, smooth-walled. Ascomata uniloculate, scattered, immersed, globose, glabrous, 120–225 μm high, 150–250 μm diam, with an unconspicuous beak at the apex. Peridia 15–30 μm, composed of 3–6 layers of polygonal pseudoparenchymatic cells. Asci hyaline to pale brown, cylindrical, clavate to long fusiform, 8-spored, bitunicate, 60–110 × 8–15 μm, with short stipes. Ascospores hyaline to pale brown, smooth-walled, fusiform, slightly curved, guttulate, 3-septate, occasionally 4-septate, slightly constricted at septum, L/W = 6.0, 25–40 × 4–6 μm (mean = 30.3 ± 3.5 × 5.0 ± 0.4 μm, n = 30).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, air, 19 July 2014, Z.F. Zhang (HMAS 246932 holotype designated here, ex-type living culture CGMCC 3.17937 = LC5367); ibid., CGMCC 3.17938 = LC6215.

Notes — This species should be classified in Phaeosphaeria because of its typical characters: uniloculate, scattered, immersed, globose, and glabrous ascomata with a beak; 8-spored, bitunicate asci; fusiform, 3-septate ascospores with weak constriction (Shoemaker & Babcock 1989, Quaedvlieg et al. 2013). Although Phaeosphaeriopsis (Ps.) was similar to Phaeosphaeria (Pa.) and had a high identity of LSU sequence with our isolates, the ascospores of Phaeosphaeriopsis are cylindrical rather than narrowly fusiform (Câmara et al. 2003, Quaedvlieg et al. 2013) and our isolates are phylogenetically allied to Phaeosphaeria (phylogenetic tree deposited in MycoBank: MB818257). Phaeosphaeria fusispora is morphologically similar to Pa. franklinensis, Pa. juncicola, and Pa. juncinella. While Pa. fusispora differs from Pa. franklinensis and Pa. juncicola in the longer asci (60–110 μm vs 45–70 μm for Pa. franklinensis and 45–60 μm for Pa. juncicola); it differs from Pa. juncinella in having shorter asci (60–110 μm vs 100–140 μm) and constricted septa. Phylogenetically, the new species clustered apart from morphologically similar species.

Ramophialophora globispora Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818258; Fig. 17

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g017.jpg

Ramophialophora globispora (from ex-holotype CGMCC 3.17939). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c. conidial beam on SNA under stereomicroscope; d. conidial beam; e–f. rough-walled conidiophores; g. conidia. — Scale bars: d = 20 μm; e–g = 10 μm.

Etymology. Referring to its globose conidia.

Colonies on PDA 43–52 mm diam after 3 wk, felty to cottony, flat, margin entire, grey-white to brown-grey. Reverse yellow-green to black. Colonies on SNA 51–58 mm diam after 3 wk, cottony, margin entire, taupe, aerial mycelia sparse. Reverse taupe.

Vegetative hyphae hyaline to pale yellow, branched, septate, smooth-walled. Conidiophores arise from prostrate aerial hyphae solitary, erect, straight, septate, unbranched, hyaline to pale yellow, thin- and rough-walled, 42–200 × 2–3 μm, globose at the apex, bearing tufty and terminally conidiogenous cells. Phialides often penicillate, hyaline, smooth-walled, elliptical, with inconspicuous apical collarette, 5–11 × 2.5–3.5 μm. Conidia enteroblastic, hyaline, globose, thin- and smooth-walled, 2–3 μm diam (mean = 2.3 ± 0.2 μm, n = 30), in chains, long macroscopic conidial beam formed with the aggregation of conidial chains.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 1, N28°12′629″ E107°13′639″, plant debris, 19 July 2014, Z.F. Zhang (HMAS 246933 holotype designated here, ex-type living culture CGMCC 3.17939 = LC5696); ibid., CGMCC 3.17940 = LC6218.

Notes — Ramophialophora was established by Calduch et al. (2004) to accommodate several species traditionally classified in Phialophora but bear phylogenetic affinity to Sordariales. Currently the genus includes only two species. Multilocus phylogenetic analysis based on ITS, LSU, and TUB sequences showed that R. globispora and R. petraea clustered with Ramophialophora, and several Cercophora and Podospora species (phylogenetic tree deposited in MycoBank: MB818258). Morphologically, R. globispora and R. petraea are well allied to Ramophialophora. Ramophialophora globispora can easily be distinguished from R. humicola and R. vesiculosa by the unbranched conidiophores, discrete conidiogenous cells, and spherical conidia without protuberant basal hila. The ex-type strain of Phialophora cyclaminis, CBS 166.42, also clustered in Sordariales, thus might need to be transferred to Ramophialophora. Ramophialophora globispora can easily be distinguished from P. cyclaminis by its unbranched and rough-walled conidiophores vs the sparsely branched and smooth-walled conidiophores in P. cyclaminis.

Ramophialophora petraea Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818259; Fig. 18

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g018.jpg

Ramophialophora petraea (from ex-holotype CGMCC 3.17952). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c–f. phialides and conidia; g. conidia. — Scale bars: c–g = 10 μm.

Etymology. Referring to the sample where this species was first isolated from.

Colonies on PDA attaining 39–41 mm diam after 21 d, flat, margin entire, grey-green at the centre and white to the margin. Reverse pale brown to white. Colonies on SNA attaining 30–34 mm diam after 21 d, margin entire, white, aerial mycelia sparse. Reverse white.

Vegetative hyphae hyaline to pale green, septate, branched, thin- and smooth-walled, 2–3 μm diam, sometimes swollen. Conidiophores reduced to the conidiogenous cells, or one supporting cell. Phialides not abundant, arising laterally or terminally from aerial mycelia, or from the supporting cells of the conidiophores, solitary, ampulliform or sometimes irregular, straight or curved, slightly constricted at the base and gradually tapering toward the apex, 7–15 × 2–4 μm, with one or occasionally two conspicuous collarettes. Conidia aggregated in small slimy heads, enteroblastic, globose, smooth, hyaline, 1.5–3 μm diam (mean = 2.2 ± 0.3 μm, n = 30), sometimes with a basal hilum.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, rock, 19 July 2014, Z.F. Zhang (HMAS 246934 holotype designated here, ex-type living culture CGMCC 3.17952 = LC5789); ibid., CGMCC 3.17953 = LC6222.

Notes — This fungus is phylogenetically allied to Ramophialophora (MB818259), which was established by Calduch et al. (2004), and the morphological features were also similar. Ramophialophora petraea differs from the known species in the genus in its phialides which arise laterally or terminally from aerial mycelia or the supporting cells, and with 1 or 2 conspicuous collarettes.

Scopulariopsis crassa Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818260; Fig. 19

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g019.jpg

Scopulariopsis crassa (from ex-holotype CGMCC 3.17941). a–b. Upper and reverse views of cultures on PDA and SNA 21 d after inoculation; c. conidiomata; d–e. conidiophores; f–g. conidia. — Scale bars: d–g = 10 μm.

Etymology. Referring to its thick-walled conidia.

Colonies on PDA 23–35 mm diam after 10 d, felty, flat, margin fimbriate, pale brown, aerial mycelia sparse. Reverse pale brown. Colonies on SNA 28–32 mm diam after 10 d, flat, margin fimbriate, pale yellow. Reverse pale yellow.

Vegetative hyphae hyaline to pale brown, septate, branched, smooth- and thick-walled. Conidiophores arise from prostrate hyphae or aggregated in conidiomata, erect, straight or slightly curved, branched, septate, smooth-walled, hyaline to pale brown, 2–3.5 μm diam. Conidiogenous cells borne on aerial hyphae or conidiophores in whorls of 1–3, annellidic, hyaline, cylindrical, or some irregular shapes, slightly curved, occasionally septate, 15–43 × 3–6 μm. Conidia in chains, hyaline to pale brown, thick-walled, smooth or finely verrucose, globose or subglobose, 5–10.5 × 5–8.5 μm (mean = 7.8 ± 1.2 × 6.6 ± 1.1 μm, n = 40), with truncated bases.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, soil, 19 July 2014, X. Zhou (HMAS 246935 holotype designated here, ex-type living culture CGMCC 3.17941 = LC5847); ibid., CGMCC 3.17942 = LC6225.

Notes — Scopulariopsis crassa is phylogenetically closely related to S. asperula and S. candida based on the analysis of ITS, LSU, TUB, and EF1-α sequences (phylogenetic tree deposited in MycoBank: MB818260). However, S. crassa can be differentiated from these two species in producing longer conidiogenous cells (15–43 μm for S. crassa vs 5–27 μm for S. asperula, 5–16 μm for S. candida).

Simplicillium calcicola Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818261; Fig. 20

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g020.jpg

Simplicillium calcicola (from ex-holotype CGMCC 3.17943). a–b. Upper and reverse views of cultures on PDA (10 d) and SNA (21 d); c. phialides on synnemata; d–f. phialides; g–h. microconidia and macroconidia. — Scale bars: c–h = 10 μm.

Etymology. Referring to the substrate it was isolated from, calcaire.

Colonies on PDA attaining 34–38 mm diam after 10 d, cottony, compact, margin entire, white. Yellow pigment produced with aging. Reverse pale yellow to yellow. Colonies on SNA attaining 33–38 mm diam after 10 d, fluffy, margin entire, white. Reverse white.

Vegetative hyphae hyaline, aseptate, unbranched, smooth-walled. Phialides arise from prostrate hyphae or synnemata, solitary or up to 2–3 in whorls, straight or a little curved, tapering towards the apex, without basal septum, 14–38 × 1–2 μm. Conidia variable in size and shape, 1-celled, smooth-walled; microconidia globose, oval or ellipsoidal, 2–3.5 × 1–1.5 μm (mean = 2.5 ± 0.3 × 1.4 ± 0.1 μm, n = 20), macroconidia fusiform, 4.5–8 × 1–2 μm (mean = 5.8 ± 0.9 × 1.4 ± 0.3 μm, n = 20). Octahedral crystals absent.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, rock, 19 July 2014, X. Zhou (HMAS 246936 holotype designated here, ex-type living culture CGMCC 3.17943 = LC5586); ibid., CGMCC 3.17944 = LC5371.

Notes — Simplicillium is characterised by predominantly solitary phialides, conidial masses either in globose slimy heads, short chains, or formed in sympodial succession (Zare & Gams 2001, Nonaka et al. 2013). Simplicillium lamellicola is similar to S. calcicola in producing both microconidia and macroconidia. However, the octahedral crystals of S. calcicola are absent and its marcoconidia are wider than those of S. lamellicola (1–2 μm vs 0.8–1.2 μm).

Volutella aeria Z.F. Zhang & L. Cai, sp. nov. — MycoBank MB818262; Fig. 21

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g021.jpg

Volutella aeria (from ex-holotype CGMCC 3.17945). a–b. Upper and reverse views of cultures on PDA and SNA 14 d after inoculation; c–d. sporodochia under stereomicroscope and microscope; e–f. conidiophores and phialides from sporodochia; g. conidia from sporodochia; h–i. conidiophores and phialides from aerial mycelia; j. conidia from aerial mycelia. — Scale bars: d = 100 μm; e–g, j = 10 μm; h = 50 μm; i = 20 μm.

Etymology. Referring to the sample where this species was first isolated from.

Colonies on PDA attaining 37–43 mm diam after 14 d, ulotrichy, margin slightly undulate, white to pale brown. Reverse plicated, yellowish to brown. Colonies on SNA attaining 46–54 mm diam after 14 d, margin erose, white, aerial mycelia sparse. Reverse white.

Vegetative hyphae hyaline to brown, septate, branched, thin- and smooth-walled. Setae hyaline, aseptate, thick-walled, tapering to end, 300–600 μm long, 2.5–4 μm wide at base, swollen terminally or intermediately. Sporodochia sessile, globose, cream yellow, 100–220 μm diam, 150–280 μm high, with several marginal setae. Conidiophores hyaline, cylindrical, branched 1–3 times, 1.5–2.5 μm wide. Conidiogenous cells hyaline, cylindrical, 9–15 × 1.8–2.5 μm, gathered into a dense parallel layer. Conidia forming slimy heads on sporodochium, 1-celled, hyaline, bacillary, 5.5–8 × 1.5–2 μm (mean = 6.6 ± 0.6 × 1.7 ± 0.1 μm, n = 35). Verticillium-like synasexual morph present. Conidiophores on aerial hyphae hyaline, branched, septate, the axis 2–3 μm, producing 1–5 phialides per node laterally or in whorls of 3–6 phialides terminally. Phialides hyaline, aseptate, slender, tapering to end, 16–34 μm long, 1.5–2.5 μm wide at base. Conidia hyaline, smooth, cylindrical, with obtuse ends, solitary, 5.5–11.5 × 2–3.5 μm (mean = 7.5 ± 1.5 × 2.7 ± 0.4 μm, n = 40).

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, air, 19 July 2014, Z.F. Zhang (HMAS 246937 holotype deposited here, ex-type living culture CGMCC 3.17945 = LC5434); ibid., CGMCC 3.17946 = LC6216.

Notes — Volutella is characterised by discoid sporodochia with marginal setae, simple to verticillate conidiophores, compact and phialidic conidiogenous cells, and 1-celled, ovoid to oblong conidia; synasexual morph present in some species and with two or more whorls of conidiogenous cells (Gräfenhan et al. 2011, Luo & Zhuang 2012, Lombard et al. 2015). Only four species in the genus were known to produce sporodochia and verticillium-like synasexual morphs, i.e., V. asiana, V. ciliata, V. consors, and the new species V. aeria described in this study. Volutella aeria differs from V. consors in producing longer setae (300–600 μm vs 250–260 μm) and the absence of flaring collarette on sporodochia; from V. ciliata in its longer conidia (5.5–11.5 μm vs 3–5.5 μm); from V. asiana in producing conidiophores with whorls of phialides on aerial mycelia, while conidiophores in V. asiana are simple with a single phialide.

Wardomycopsis longicatenata Z.F. Zhang, F. Liu & L. Cai, sp. nov. — MycoBank MB818263; Fig. 22

An external file that holds a picture, illustration, etc.
Object name is per-39-1-g022.jpg

Wardomycopsis longicatenata (from ex-holotype CGMCC 3.17947). a–b. Upper and reverse views of cultures on PDA and SNA 8 wk after inoculation; c. immersed ascoma; d. ascoma; e. peridium; f. asci; g. ascospore; h–k. conidiophores and conidiogenous cells; l–m. conidia. — Scale bars: d = 100 μm; e, l = 20 μm; f–k, m = 10 μm.

Etymology. Referring to its long conidial chains.

Colonies on PDA 40–45 mm diam after 8 wk, felty, slightly raised, margin entire, yellow-green to dark grey. Reverse yellow-green to dark green. Colonies on SNA 45–51 mm diam after 8 wk, compact, aerial mycelia sparse, margin fimbriate, white to dark green. Reverse dark green.

Vegetative hyphae hyaline to pale brown, septate, branched, thin- and smooth-walled. Ascomata dark brown to black, immersed, globose or subglobose, 200–330 μm diam, 210–310 μm high, with unconspicuous ostiole. Peridia of textura angularis, olive green, appendages lacking. Asci ovate, globose or subglobose, 8-spored, 7–10.5 × 6–7.5 μm. Ascospores triangular to lunate, hyaline to pale red-brown, 3–4.5 × 2–3 μm (mean = 3.8 ± 0.3 × 2.7 ± 0.2 μm, n = 25). Conidiophores arising from hyphae, straight or flexuous, septate, occasionally branched one or two times, smooth, hyaline to pale brown. Conidiogenous cells solitary on aerial hyphae, or in whorls of 2–3 on apex of conidiophores, hyaline to pale brown, ampulliform, cylindrical, slightly curved, 3–6(–8.5) × 1.5–2.5 μm. Conidia in a long chain, brown, thick-walled, ellipsoidal, 4–6 × 2–2.5 μm (mean = 5.4 ± 0.4 × 2.1 ± 0.1 μm, n = 30), with truncated base and median longitudinal germ slit.

Specimens examined. China, Guizhou, Kuankuoshui National Natural Reserve, unnamed Karst Cave 2, N28°12′599″ E107°13′661″, air, 19 July 2014, Z.F. Zhang (HMAS 246938 holotype designated here, ex-type living culture CGMCC 3.17947 = LC5709); ibid., CGMCC 3.17948 = LC6226.

Notes — Wardomycopsis was established to accommodate asexual morphs of Microascus (Udagawa & Furuya 1978), and characterised by dark, globose, thick-walled conidia with germ slits that form short chains on annellidic conidiogenous cells (Silvera-Simón et al. 2008). However, recent phylogenetic study based on the ITS and LSU sequences suggested that Wardomycopsis and Microascus are both monophyletic but distinct from each other (Sandoval-Denis et al. 2016). Wardomycopsis longicatenata clustered within Wardomycopsis and formed a distinct clade with high support value based on the ITS, LSU, TUB, and EF1-α sequence analysis (phylogenetic tree deposited in MycoBank: MB818263). Currently, this genus has four species. Morphologically W. longicatenata should be compared to W. humicola which produces similar ellipsoidal conidia. While they can easily be distinguished from one other by the different shapes of conidiogenous cells, which is ampulliform in W. longicatenata but ovoid to subglobose in W. humicola.

DISCUSSION

This study significantly improved our understanding of the mycobiota in caves and our data further suggested that fungal communities among different caves are largely different from each other (59 % of identified species in this study were reported for the first time from caves). The total number of species of Ascomycota was much higher than that of Basidiomycota, which may be explained by the lack of large, nutrient rich substrates such as plant debris or dung in the caves.

Our data also suggested that the majority of fungi documented from caves originated from the outside environment, in agreement to that of Vanderwolf et al. (2013). The most common species revealed in this study are very similar to that listed by Vanderwolf et al. (2013), which are commonly found in the above-ground environments. All genera recorded in this study are known from the other environments, and most species (83 %) have also been reported from other environments.

Kuzmina et al. (2012) suggested that cave systems might be a good harbour for the development and preservation of allochthonous microorganisms, including pathogenic species. Many species we obtained in this study were known as plant endophytic or pathogenic species. They may originate from outside environments; dispersed through water or air flow; and remain alive in caves. For example, Diaporthe phoenicicola isolated from soil in Cave 2, is a species known to cause fungal scleral keratitis in humans (Gajjar et al. 2011). Fusarium graminearum isolated from the air in Cave 1, is a plant pathogen which causes head blight of wheat (Bai & Shaner 2004). Pestalotiopsis guepinii, isolated from water on Cave 1, was reported to cause azalea petal blight in Argentina (Rivera & Wright 2000). Many species of Colletotrichum, Cylindrocarpon, and Phoma complexes investigated in this study are also well-known plant pathogenic fungi.

The distribution of microbial colonies in caves appeared to be largely determined by the bio-receptivity and susceptibility of the host materials, and the internal micro-environmental conditions, especially water availability and the mobilisation of nutrients in favour of the predominant mass and energy fluxes (Cuezva et al. 2009, Jurado et al. 2009). In our study, most fungi (99 species belonging to 59 genera) were isolated from organic litter compared to other samples, followed by soil, air, rock, and water samples. Therefore, oligotrophy might be a major limitation of fungal colonisation in caves and higher fungal diversity would be discovered on samples with higher organic carbon concentration in caves (Bastian et al. 2010, Jurado et al. 2010, Kuzmina et al. 2012).

Whether obligate troglobitic fungi exist in caves is a very interesting question, but it needs further investigation. There are many species that have been exclusively isolated from caves (e.g., Aspergillus baeticus, As. spelunceus, As. thesauricus, Chrysosporium (Ch.) chiropterorum, Ch. speluncarum, Microascus caviariformis, Mucor troglophilus, Ochroconis anomala, Ochroconis lascauxensis, Ombrophila speluncarum, Trichosporon (Tr.) akiyoshidainum, Tr. cavernicola, Tr. chiropterorum) (Vanderwolf et al. 2013). Several of these species were also obtained in this study, such as As. thesauricus and Tr. akiyoshidainum. Currently we could not conclude if the 20 new species described in this study are obligate troglobitic fungi, or opportunistic colonisers that have dispersed from outside environments.

Several species obtained from this study may be potentially highly valuable. We obtained seven strains of Amphichorda felina (syn. Beauveria felina, Isaria felina), which is a widely known species in producing insecticidal cyclodepsipeptide (Baute et al. 1981, Langenfeld et al. 2011, Seifert et al. 2011). It would be interesting to investigate whether our new species Amphichorda guana generates insecticidal activity. Another example is Trichoderma hamatum isolated from soils from Cave 1 and Cave 2, a species that has been used as biocontrol agents against fungal diseases of plants (Harman 2006). Trichoderma longibrachiatum, a xylanase producing species, was isolated from soils in Cave 2 (Felse & Panda 1999).

In summary, our investigation of the culturable mycobiota in caves revealed a high fungal diversity, including a number of new species scattered in different families and orders. Fungal communities in different caves are largely different from each other but most of the identified species have been reported from other environments, and the outside environments appear to be the major source of fungal flora in caves. Although a number of new species were discovered in this study, this and previous studies on cave fungi did not find any new genera or families, indicating a lack of independently evolved fungal lineage in caves. This is possibly a reflection of the fact that the geographic history of caves (mostly shorter than two million years, Zhang et al. 2000) on earth is relatively short for fungal evolution and speciation. Future study should incorporate culture-independent methods to better reveal the overall picture of fungal diversity and community compositions in caves.

Acknowledgments

This study was financially supported by the Project for Fundamental Research on Science and Technology, Ministry of Science and Technology of China (2014FY120100). Zhi-Feng Zhang acknowledges QYZDB-SSW-SMC044 for supporting his studentship. Jia-Rui Jiang and Qian Chen are thanked for help with sample collection. We also thank Prof. dr Pedro W. Crous who provided valuable and constructive suggestions to this study.

REFERENCES

  • Ajello L, Briceño-Maaz T, Campins H, et al. 1960a. Isolation of Histoplasma capsulatum from an oil bird (Steatornis caripensis) cave in Venezuela. Mycopathologia et Mycologia Applicata 12: 199–206. [Abstract] [Google Scholar]
  • Ajello L, Manson-Bahr PEC, Moore JC. 1960b. Amboni caves, Tanganyika, a new endemic area for Histoplasma capsulatum. American Journal of Tropical Medicine and Hygiene 9: 633–638. [Abstract] [Google Scholar]
  • Al-Doory Y, Rhoades ER. 1968. Isolation of Histoplasma capsulatum from a Texas cave. Mycopathologia et Mycologia Applicata 35: 201–207. [Abstract] [Google Scholar]
  • Bai G, Shaner G. 2004. Management and resistance in wheat and barley to Fusarium head blight. Annual Review of Phytopathology 42: 135–161. [Abstract] [Google Scholar]
  • Barton HA, Jurado V. 2007. What’s up down there? Microbial diversity in caves. Microbe 2: 132–138. [Google Scholar]
  • Bastian F, Jurado V, Novakova A, et al. 2010. The microbiology of Lascaux Cave. Microbiology 156: 644–652. [Abstract] [Google Scholar]
  • Baute R, Deffieux G, Merlet D, et al. 1981. New insecticidal cyclodepsipeptides from the fungus Isaria felina. I. Production, isolation and insecticidal properties of isariins B. C and D. The Journal of Antibiotics 34: 1261–1265. [Abstract] [Google Scholar]
  • Borda D, Borda C, Tămaş T. 2004. Bats, climate, and air microorganisms in a Romanian cave. Mammalia 68: 337–343. [Google Scholar]
  • Cai L, Jeewon R, Hyde KD. 2005. Phylogenetic evaluation and taxonomic revision of Schizothecium based on ribosomal DNA and protein coding genes. Fungal Diversity 19: 1–21. [Google Scholar]
  • Cai L, Jeewon R, Hyde KD. 2006. Molecular systematics of Zopfiella and allied genera: evidence from multi-gene sequence analyses. Mycological Research 110: 359–368. [Abstract] [Google Scholar]
  • Calduch M, Gené J, Stchigel AM, et al. 2004. Ramophialophora, a new anamorphic genus of Sordariales. Studies in Mycology 50: 83–88. [Google Scholar]
  • Câmara MP, Ramaley AW, Castlebury LA, et al. 2003. Neophaeosphaeria and Phaeosphaeriopsis, segregates of Paraphaeosphaeria. Mycological Research 107: 516–522. [Abstract] [Google Scholar]
  • Chen MJ, Huang B, Li ZZ, et al. 2013. Morphological and genetic characterisation of Beauveria sinensis sp. nov. from China. Mycotaxon 124: 301–308. [Google Scholar]
  • Collado J, Platas G, Paulus B, et al. 2007. High-throughput culturing of fungi from plant litter by a dilution-to-extinction technique. FEMS Microbiology Ecology 60: 521–533. [Abstract] [Google Scholar]
  • Crous PW, Gams W, Stalpers JA, et al. 2004. MycoBank: an online initiative to launch mycology into the 21st century. Studies in Mycology 50: 19–22. [Google Scholar]
  • Cuezva S, Sanchez-Moral S, Saiz-Jimenez S, et al. 2009. Microbial communities and associated mineral fabrics in Altamira Cave, Spain. International Journal of Speleology 38: 83–92. [Google Scholar]
  • Cunningham KI, Northup DE, Pollastro RM, et al. 1995. Bacteria, fungi and bioKarst in Lechuguilla Cave, Carlsbad Caverns National Park, New Mexico. Environmental Geology 25: 2–8. [Google Scholar]
  • Darriba D, Taboada GL, Doallo R, et al. 2012. jModelTest 2: more models, new heuristics and parallel computing. Nature Methods 9: 772. [Europe PMC free article] [Abstract] [Google Scholar]
  • De Hoog GS. 1972. The genera Beauveria, Isaria, Tritirachium and Acrodontium gen. nov. Studies in Mycology 1: 1–41. [Google Scholar]
  • Di Salvo AF, Ajello L, Palmer JW, Jr, et al. 1969. Isolation of Histoplasma capsulatum from Arizona bats. American Journal of Epidemiology 89: 606–614. [Abstract] [Google Scholar]
  • Dobat K. 1967. Ein bisher unveroffentlichtes botanisches manuskript Alexander von Humboldts: Plantae subterranae Europ. 1794 cum Iconibus. Akademie Der Wissenschaften und der Literatur 6: 16–19. [Google Scholar]
  • Doyle JJ. 1987. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochemical Bulletin 19: 11–15. [Google Scholar]
  • Felse PA, Panda T. 1999. Self-directing optimization of parameters for extra-cellular chitinase production by Trichoderma harzianum in batch mode. Process Biochemistry 34: 563–566. [Google Scholar]
  • Fries EM. 1825. Systema orbis vegetabilis. Lund. [Google Scholar]
  • Gabriel CR, Northup DE. 2013. Microbial ecology: caves as an extreme habitat. In: Cheeptham N. (ed), Cave microbiomes: a novel resource for drug discovery: 85–108. Springer Press, Germany. [Google Scholar]
  • Gajjar DU, Pal AK, Parmar TJ, et al. 2011. Fungal scleral keratitis caused by Phomopsis phoenicicola. Journal of Clinical Microbiology 49: 2365–2368. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gams W, Zare R. 2001. A revision of Verticillium sect. Prostrata. III. Generic classification. Nova Hedwigia 72: 329–337. [Google Scholar]
  • Glass NL, Donaldson GC. 1995. Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous Ascomycetes. Applied and Environmental Microbiology 61: 1323–1330. [Europe PMC free article] [Abstract] [Google Scholar]
  • Gräfenhan T, Schroers HJ, Nirenberg HI, et al. 2011. An overview of the taxonomy, phylogeny, and typification of nectriaceous fungi in Cosmospora, Acremonium, Fusarium, Stilbella, and Volutella. Studies in Mycology 68: 79–113. [Europe PMC free article] [Abstract] [Google Scholar]
  • Guindon S, Gascuel O. 2003. A simple, fast and accurate method to estimate large phylogenies by maximum-likelihood. Systematic Biology 52: 696–704. [Abstract] [Google Scholar]
  • Harman GE. 2006. Overview of mechanisms and uses of Trichoderma spp. Phytopathology 96: 190–194. [Abstract] [Google Scholar]
  • Hawksworth DL, Crous PW, Redhead SA, et al. 2011. The Amsterdam declaration on fungal nomenclature. IMA Fungus 2: 105–112; Mycotaxon 116: 491–500. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hernández-Restrepo M, Groenewald JZ, Crous PW. 2016. Taxonomic and phylogenetic re-evaluation of Microdochium, Monographella and Idriella. Persoonia 36: 57–82. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hose LD, Palmer AN, Palmer MV, et al. 2000. Microbiology and geochemistry in a hydrogen-sulphide-rich Karst environment. Chemical Geology 169: 399–423. [Google Scholar]
  • Huelsenbeck JP, Ronquist F. 2001. MrBayes: Bayesian inference of phylogenetic trees. Bioinformatics 17: 754–755. [Abstract] [Google Scholar]
  • Jiang XY, He YQ, Shen CC, et al. 2012. Stalagmite-inferred Holocene precipitation in northern Guizhou Province, China, and asynchronous termination of the Climatic Optimum in the Asian monsoon territory. Chinese Science Bulletin 57: 795–801. [Google Scholar]
  • Ju YM, Rogers JD, San Martin F, et al. 1998. The genus Biscogniauxia. Mycotaxon 66: 1–98. [Google Scholar]
  • Jurado V, Fernandez-Cortes A, Cuezva S, et al. 2009. The fungal colonisation of rock-art caves: experimental evidence. Naturwissenschaften 96: 1027–1034. [Abstract] [Google Scholar]
  • Jurado V, Porca E, Cuezva S. 2010. Fungal outbreak in a show cave. Science of the Total Environment 408: 3632–3638. [Abstract] [Google Scholar]
  • Kajihiro ES. 1965. Occurrence of dermatophytes in fresh bat guano. Applied Microbiology 13: 720–724. [Europe PMC free article] [Abstract] [Google Scholar]
  • Katoh K, Toh H. 2010. Parallelization of the MAFFT multiple sequence alignment program. Bioinformatics 26: 1899–1900. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kepler RM, Humber RA, Bischoff JF, et al. 2014. Clarification of generic and species boundaries for Metarhizium and related fungi through multigene phylogenetics. Mycologia 106: 811–829. [Abstract] [Google Scholar]
  • Koilraj AJ, Marimuthu G, Natarajan K, et al. 1999. Fungal diversity inside caves of southern India. Current Science 77: 1081–1084. [Google Scholar]
  • Kruys Å, Huhndorf SM, Miller AN. 2015. Coprophilous contributions to the phylogeny of Lasiosphaeriaceae and allied taxa within Sordariales (Ascomycota, Fungi). Fungal Diversity 70: 101–113. [Google Scholar]
  • Kuzmina LY, Galimzianova NF, Abdullin SR, et al. 2012. Microbiota of the Kinderlinskaya Cave (South Urals, Russia). Microbiology 81: 251–258. [Google Scholar]
  • Lagarde J. 1913. Biospeologica in champignons. Archives de Zoologie Experimentale et Generale 53: 277–307. [Google Scholar]
  • Langenfeld A, Blond A, Gueye S, et al. 2011. Insecticidal cyclodepsipeptides from Beauveria felina. Journal of Natural Products 74: 825–830. [Abstract] [Google Scholar]
  • Liu F, Hu DM, Cai L. 2012. Conlarium duplumascospora gen. et. sp. nov. and Jobellisia guangdongensis sp. nov. from freshwater habitats in China. Mycologia 104: 1178–1186. [Abstract] [Google Scholar]
  • Liu YJ, Whelen S, Hall BD. 1999. Phylogenetic relationships among ascomycetes: evidence from an RNA polymerase II subunit. Molecular Biology and Evolution 16: 1799–1808. [Abstract] [Google Scholar]
  • Lombard L, Van der Merwe NA, Groenewald JZ, et al. 2015. Generic concepts in Nectriaceae. Studies in Mycology 80: 189–245. [Europe PMC free article] [Abstract] [Google Scholar]
  • Luo J, Zhuang WY. 2012. Volutellonectria (Ascomycota, Fungi), a new genus with Volutella anamorphs. Phytotaxa 44: 1–10. [Google Scholar]
  • Lurie HI, Borok R. 1955. Trichophyton mentagrophytes isolated from the soil of caves. Mycologia 47: 506–510. [Google Scholar]
  • Lurie HI, Way M. 1957. The isolation of dermatophytes from the atmosphere of caves. Mycologia 49: 178–180. [Google Scholar]
  • Madrid H, Cano J, Gené J, et al. 2011. Two new species of Cladorrhinum. Mycologia 103: 795–805. [Abstract] [Google Scholar]
  • Miller AN, Huhndorf SM. 2005. Multi-gene phylogenies indicate ascomal wall morphology is a better predictor of phylogenetic relationships than ascospore morphology in the Sordariales (Ascomycota, Fungi). Molecular Phylogenetics and Evolution 35: 60–75. [Abstract] [Google Scholar]
  • Mouchacca J, Gams W. 1993. The hyphomycete genus Cladorrhinum and its teleomorph connections. Mycotaxon 48: 415–440. [Google Scholar]
  • Nagai K, Suzuki K, Okada G. 1998. Studies on the distribution of alkalophilic and alkali-tolerant soil fungi II: Fungal flora in two limestone caves in Japan. Mycoscience 39: 293–298. [Google Scholar]
  • Nirenberg HI. 1976. Untersuchungen über die morphologische und biologische Differenzierung in der Fusarium-Section Liseola. Mitteilungen der Biologischen Bundesanstalt für Land- und Forstwirtschaft 169: 1–117. [Google Scholar]
  • Nonaka K, Kaifuchi S, Ōmura S, et al. 2013. Five new Simplicillium species (Cordycipitaceae) from soils in Tokyo, Japan. Mycoscience 54: 42–53. [Google Scholar]
  • Nováková A. 2009. Microscopic fungi isolated from the Domica Cave system (Slovak Karst National Park, Slovakia). A review. International Journal of Speleology 38: 71–82. [Google Scholar]
  • Ogórek R, Lejman A, Matkowski K. 2013. Fungi isolated from Niedźwiedzia Cave in Kletno (Lower Silesia, Poland). International Journal of Speleology 42: 161–166. [Google Scholar]
  • Orr GF, Kuehn HH. 1972. Notes on Gymnoascaceae. II. Some Gymnoascaceae and keratinophilic fungi from Utah. Mycologia 64: 55–72. [Abstract] [Google Scholar]
  • Ortiz M, Legatzki A, Neilson JW, et al. 2014. Making a living while starving in the dark: metagenomic insights into the energy dynamics of a carbonate cave. The ISME Journal 8: 478–491. [Europe PMC free article] [Abstract] [Google Scholar]
  • Page RDM 1996. TreeView: an application to display phylogenetic trees on personal computers. Computer Applications in the Biosciences 12: 357–358. [Abstract] [Google Scholar]
  • Petrak F. 1924. Mykologische Notizen. VII. Annales Mycologici 22: 1–182. [Google Scholar]
  • Quaedvlieg W, Verkley GJM, Shin HD, et al. 2013. Sizing up Septoria. Studies in Mycology 75: 307–390. [Europe PMC free article] [Abstract] [Google Scholar]
  • Rannala B, Yang Z. 1996. Probability distribution of molecular evolutionary trees: a new method of phylogenetic inference. Journal of Molecular Evolution 43: 304–311. [Abstract] [Google Scholar]
  • Rehner SA, Buckley E. 2005. A Beauveria phylogeny inferred from nuclear ITS and EF1-α sequences: evidence for cryptic diversification and links to Cordyceps teleomorphs. Mycologia 97: 84–98. [Abstract] [Google Scholar]
  • Rehner SA, Minnis AM, Sung GH, et al. 2011. Phylogeny and systematics of the anamorphic, entomopathogenic genus Beauveria. Mycologia 103: 1055–1073. [Abstract] [Google Scholar]
  • Rivera MC, Wright ER. 2000. First report of Azalea Petal Blight caused by Pestalotiopsis guepini in Argentina. Plant Disease 84: 100. [Abstract] [Google Scholar]
  • Ruibal C, Platas G, Bills GF. 2005. Isolation and characterization of melanized fungi from limestone formations in Mallorca. Mycological Progress 4: 23–38. [Google Scholar]
  • Sandoval-Denis M, Gené J, Sutton DA, et al. 2016. Redefining Microascus, Scopulariopsis and allied genera. Persoonia 36: 1–36. [Europe PMC free article] [Abstract] [Google Scholar]
  • Seifert KA, Morgan-Jones G, Gams W, et al. 2011. The genera of Hyphomycetes. Utrecht, CBS-KNAW Fungal Biodiversity Centre. [Google Scholar]
  • Selbmann L, De Hoog GS, Mazzaglia A, et al. 2005. Fungi at the edge of life: cryptoendolithic black fungi from Antarctic desert. Studies in Mycology 51: 1–32. [Google Scholar]
  • Sharma R, Gräser Y, Singh SK. 2013. Auxarthronopsis, a new genus of Onygenales isolated from the vicinity of Bandhavgarh National Park, India. IMA Fungus 4: 89–102. [Europe PMC free article] [Abstract] [Google Scholar]
  • Shoemaker RA, Babcock CE. 1989. Phaeosphaeria. Canadian Journal of Botany 67: 1500–1599. [Google Scholar]
  • Silvera-Simón C, Gené J, Cano J, et al. 2008. Wardomycopsis litoralis, a new soil-borne hyphomycete from Spain. Mycotaxon 105: 195–202. [Google Scholar]
  • Stamatakis A. 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688–2690. [Abstract] [Google Scholar]
  • Sung GH, Sung JM, Hywel-Jones NL, et al. 2007. A multi-gene phylogeny of Clavicipitaceae (Ascomycota, Fungi): identification of localized incongruence using a combinational bootstrap approach. Molecular Phylogenetics and Evolution 44: 1204–1223. [Abstract] [Google Scholar]
  • Sydow H. 1924. Mycotheca germanica. Fasc. XLII–XLV (No. 2051–2250). Annales Mycologici 22: 257–268. [Google Scholar]
  • Tamura K, Stecher G, Peterson D, et al. 2013. MEGA6: molecular evolutionary genetics analysis version 6.0. Molecular Biology and Evolution 30: 2725–2729. [Europe PMC free article] [Abstract] [Google Scholar]
  • Traaen AE. 1914. Untersuchungen über Bodenpilze aus Norwegen. Nytt Magazin for Naturvidenskapene 52: 20–121. [Google Scholar]
  • Udagawa SI, Furuya K. 1978. New species of Microascus and its peculiar conidial state. Mycotaxon 7: 91–96. [Google Scholar]
  • Van Oorschot CAN. 1980. A revision of Chrysosporium and allied genera. Studies in Mycology 20: 1–89. [Google Scholar]
  • Vanderwolf KJ, Malloch D, McAlpine DF, et al. 2013. A world review of fungi, yeasts, and slime molds in caves. International Journal of Speleology 42: 77–96. [Google Scholar]
  • Varsavsky E, Orr GF. 1971. A new genus of the gymnoascaceae. Mycopathologia et Mycologia Applicata 43: 229–234. [Abstract] [Google Scholar]
  • Vilgalys R, Hester M. 1990. Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. Journal of Bacteriology 172: 4238–4246. [Europe PMC free article] [Abstract] [Google Scholar]
  • Wang XW, Houbraken J, Groenewald JZ, et al. 2016. Diversity and taxonomy of Chaetomium and chaetomium-like fungi from indoor environments. Studies in Mycology 84: 145–224. [Europe PMC free article] [Abstract] [Google Scholar]
  • White TJ, Bruns T, Lee S, et al. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. Innis MA, Gelfand DH, Sninsky JJ, et al. (eds), PCR protocols: a guide to methods and applications: 315–322. Academic Press Inc., New York, USA. [Google Scholar]
  • Zamora JRC. 1977. Isolation of Histoplasma capsulatum from the air in the Aguas Buenas caves, Aguas Buenas, Puerto Rico. Mycopathologia 60: 163–165. [Abstract] [Google Scholar]
  • Zare R, Gams W. 2001. A revision of Verticillium section Prostrata. IV. The genera Lecanicillium and Simplicillium gen. nov. Nova Hedwigia 73: 1–50. [Google Scholar]
  • Zare R, Gams W. 2007. Pochonia globispora sp. nov. Nova Hedwigia 84: 421–428. [Google Scholar]
  • Zhang ML, Lin YS, Ran JC, et al. 2000. The characteristics of Karst cave development in Libo, Guizhou. Carsologica Sinica 19: 13–20. [Google Scholar]
  • Zhang W, Nan Z, Tian P, et al. 2015a. Microdochium paspali, a new species causing seashore paspalum disease in southern China. Mycologia 107: 80–89. [Abstract] [Google Scholar]
  • Zhang Y, Liu F, Wu W, et al. 2015b. A phylogenetic assessment and taxonomic revision of the thermotolerant hyphomycete genera Acrophialophora and Taifanglania. Mycologia 107: 768–779. [Abstract] [Google Scholar]
  • Zhaxybayeva O, Gogarten JP. 2002. Bootstrap, Bayesian probability and maximum likelihood mapping: exploring new tools for comparative genome analyses. Genomics 3: 1–15. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhou J, Gu Y, Zou C, et al. 2007. Phylogenetic diversity of bacteria in an earth-cave in Guizhou Province, Southwest of China. The Journal of Microbiology 45: 105–112. [Abstract] [Google Scholar]

Articles from Persoonia : Molecular Phylogeny and Evolution of Fungi are provided here courtesy of Naturalis Biodiversity Center & Centraalbureau voor Schimmelcultures

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/33988600
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/33988600

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.3767/persoonia.2017.39.01

Supporting
Mentioning
Contrasting
15
115
0

Article citations


Go to all (60) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.