Next Article in Journal
Toxicological Assessment of Pure Lolitrem B and Ryegrass Seed Infected with the AR37 Endophyte Using Mice
Next Article in Special Issue
Additions to The Knowledge of Tubakia (Tubakiaceae, Diaporthales) in China
Previous Article in Journal
Photodynamic Inactivation Effectively Eradicates Candida auris Biofilm despite Its Interference with the Upregulation of CDR1 and MDR1 Efflux Genes
Previous Article in Special Issue
Description of Four Novel Species in Pleosporales Associated with Coffee in Yunnan, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Morpho-Phylogenetic Evidence Reveals New Species of Fuscosporellaceae and Savoryellaceae from Freshwater Habitats in Guizhou Province, China

1
School of Life Science and Technology, Center for Informational Biology, University of Electronic Science and Technology of China, Chengdu 611731, China
2
School of Pharmacy, Guizhou University of Traditional Chinese Medicine, Guiyang 550025, China
3
Department of Entomology and Plant Pathology, Faculty of Agriculture, Chiang Mai University, Chiang Mai 50200, Thailand
4
Center of Excellence in Fungal Research, Mae Fah Luang University, Chiang Rai 57100, Thailand
*
Authors to whom correspondence should be addressed.
J. Fungi 2022, 8(11), 1138; https://doi.org/10.3390/jof8111138
Submission received: 16 September 2022 / Revised: 21 October 2022 / Accepted: 25 October 2022 / Published: 28 October 2022
(This article belongs to the Special Issue Polyphasic Identification of Fungi 2.0)

Abstract

:
During a survey of freshwater fungi in Guizhou Province, China, six hyphomycetous collections were founded on decaying wood from freshwater habitats. These taxa were characterized and identified based on morphology, phylogeny, and culture characteristics. Phylogenetic analysis of combined LSU, SSU, ITS, RPB2 and TEF1α sequence data indicated that our six isolates formed three distinct lineages and were distributed within Fuscosporellaceae and Savoryellaceae. They can be organized as three new species: Fuscosporella guizhouensis, Mucispora aquatica and Neoascotaiwania guizhouensis. Fuscosporella guizhouensis and Neoascotaiwania guizhouensis have sporodochial conidiomata, micronematous conidiophores and dark brown conidia. The former possesses irregularly ellipsoidal conidia with apical appendages, while the latter has fusiform to obovoid conidia. Mucispora aquatica is characterized by macronematous conidiophores, elongating percurrently and dark brown, narrowly obovoid conidia. The detailed, illustrated descriptions and notes for each new taxon are provided, and the species of Fuscosporella is reported for the first time in China.

1. Introduction

Freshwater fungi were defined as ‘‘fungi that the whole or part of their life cycle rely on freshwater’’ [1]. They are a diverse and heterogeneous group comprising different species and play an essential role in the organic carbon cycle of aquatic ecosystems [2,3]. Freshwater habitats include lentic and lotic water [1,2] and artificial habitats [4,5]. Calabon et al. [6] recently gave a comprehensive review of the freshwater fungal numbers and listed 3870 freshwater fungal species. Two hundred and ninety-eight novel species have been described in China and Thailand from 2015 to 2020 [6]. It is worth noting that the application of molecular techniques coupled with traditional morphology has significantly improved fungal identification and classification, especially the phylogenetic relationships of freshwater taxa.
The monotypic order Fuscosporellales was introduced by Yang et al. [7], with Fuscosporellaceae as the type family, based on phylogenetic analyses, and six genera were assigned, viz. Bactrodesmiastrum, Fuscosporella (type genus), Mucispora, Parafuscosporella, Plagiascoma, and Pseudoascotaiwania [7]. Plagiascoma and Pseudoascotaiwania are known for their sexual morphs, which have immersed to semi-immersed, dark brown to black ascomata, unitunicate, cylindrical to cylindric-fusiform, stipitate, 8-spored asci with a non-amyloid apical ring, and uniseriate, hyaline or light brown, fusiform, septate ascospores [7,8]. Asexual genera Bactrodesmiastrum, Fuscosporella, and Parafuscosporella share the features of having sporodochial conidiomata, semi-macronematous to macronematous, hyaline to brown, smooth-walled conidiophores, monoblastic, integrated, hyaline to dark brown conidiogenous cells and ellipsoidal, obovoid to pyriform, brown to dark brown, septate conidia [7,9,10]. In comparison, Mucispora is distinct in having macronematous, mononematous, solitary, erect, brown conidiophores, usually elongating percurrently, and ellipsoidal to obovoid conidia, sometimes with a hyaline mucilaginous sheath [7,11].
Boonyuen et al. [12] established Savoryellales to accommodate Ascotaiwania, Canalisporium (=Ascothailandia), and Savoryella based on multi-gene analyses (LSU, SSU, 5.8S rDNA, RPB1, RPB2 and TEF1α). They are distributed in freshwater, brackish, marine and terrestrial habitats, and Savoryellaceae was later formally introduced by Jaklitsch and Réblová [13]. Subsequently, Hernández-Restrepo et al. [14] introduced a bactrodesmium-like genus Neoascotaiwania, and Luo et al. [15] added a monotypic and monodictys-like genus Dematiosporium in Savoryellaceae. Réblová et al. [16] assessed the systematic placement of several Bactrodesmium species within Savoryellaceae. Sexual morphs of Savoryellales have non-stromatic, immersed, semi-immersed to superficial, dark, coriaceous ascomata, clavate to cylindrical unitunicate asci with a non-amyloid apical ring, ellipsoid to fusiform, transversely septate ascospores with hyaline polar cells and brown middle cells. Asexual morphs in Savoryellales are characterized by semi-macronematous conidiophores, monoblastic conidiogenous cells and transversely septate or dictyoseptate conidia [4,8,12,17,18,19]. Fuscosporellales and Savoryellales were initially placed in Hypocreomycetidae (Sordariomycetes) [7,12], whereafter, based on the phylogenetic and molecular clock analyses, they were referred to as a new subclass of Savoryellomycetidae (Sordariomycetes) along with Conioscyphales and Pleurotheciales by Hongsanan et al. [20].
Six isolates were obtained from submerged decaying wood during the survey of freshwater fungi in Guizhou Province, China. This study aims to describe these new findings and contribute to fungal diversity in China. Morphological comparison coupled with multi-gene phylogeny was carried out to determine the classification of these new collections. As a result, three new species are introduced, and the establishment of these new taxa is justified by morphology and phylogenetic evidence.

2. Materials and Methods

2.1. Collection and Examination of Specimens

Specimens of submerged decaying wood were collected from a freshwater stream in Guizhou Province, China, in February 2021. Samples were brought to the laboratory in plastic bags and incubated in plastic boxes lined with moistened tissue paper at room temperature for one week. Morphological observations were made using a Motic SMZ (Stereoscopic Zoom Microscope) 168 Series dissecting microscope (Motic, Xiamen, China) for fungal structures on a natural substrate. The fruiting bodies were collected using a syringe needle and transferred to a drop of tap water on a clean slide. The features were examined and photographed using a Nikon ECLIPSE Ni-U compound microscope fitted with a Nikon DS-Ri2 digital camera. Measurements were made with the Tarosoft Image Frame Work v. 0.9.7 software following the procedures outlined by Liu et al. [21], and images used for photo plates were processed with Adobe Photoshop CC 2018 software (Adobe Systems, San Jose, CA, USA). Single-spore isolations were made on potato dextrose agar (PDA) or water agar (WA) and later transferred onto new PDA plates following the methods described in Senanayake et al. [22]. Incubation and cultural growth were observed at 25 °C.
Herbarium specimens were deposited in the Herbarium of Cryptogams, Kunming Institute of Botany Academia Sinica (HKAS), Kunming, China, and Herbarium, University of Electronic Science and Technology (HUEST), Chengdu, China. The pure cultures obtained in this study were deposited in the China General Microbiological Culture Collection Center (CGMCC) in Beijing, China, and the University of Electronic Science and Technology Culture Collection (UESTCC), Chengdu, China. The new taxa were registered in MycoBank (2022).

2.2. DNA Extraction, PCR Amplification and Sequencing

Isolates grew in PDA medium at 25 °C for one month. Fungal mycelia were scraped off and transferred to 1.5 mL microcentrifuge tubes using a sterilized lancet for genomic DNA extraction. A Tsingke Fungus Genomic DNA Extraction Kit (Tsingke Biotech, Shanghai, China) was used to extract DNA following the manufacturer’s instructions. Five gene regions were amplified by Polymerase Chain Reaction (PCR). The nuclear large subunit rDNA (28S, LSU), nuclear small subunit rDNA (18S, SSU), internal transcribed spacer (ITS), RNA polymerase second-largest subunit (RPB2) and translation elongation factor 1-alpha (TEF1α) were selected for the study. The primers used were LR0R/LR5 for LSU [23], NS1/NS4 for SSU [24], ITS5/ITS4 for ITS [24], fRPB2-5F and fRPB2-7cR for RPB2 [25] and TEF1-983F/TEF1-2218R for TEF1α [26]. The amplifications were performed in a 25 μL reaction volume containing 9.5 μL of ddH2O, 12.5 μL of 2 × Taq PCR Master Mix with blue dye (Sangon Biotech, Shanghai, China), 1 μL of DNA template and 1 μL of each primer. The amplification condition for ITS, LSU, SSU and TEF1α consisted of initial denaturation at 94 °C for 3 min, followed by 40 cycles of 45 s at 94 °C, 50 s at 55 °C and 1 min at 72 °C, and a final extension period of 10 min at 72 °C. The amplification condition for the RPB2 gene consisted of initial denaturation at 95 °C for 5 min; followed by 37 cycles of 15 s at 95 °C, 50 s at 56 °C and 2 min at 72 °C, and a final extension period of 10 min at 72 °C. The PCR product purification and sequencing were performed at Beijing Tsingke Biotechnology (Chengdu) Co., Ltd., Chengdu, China.

2.3. Phylogenetic Analyses

In this study, the taxa included in the phylogenetic analyses were selected and obtained from previous studies and GenBank (Table 1), with a total of 50 taxa, including four orders, namely, Conioscyphales, Fuscosporellales, Pleurotheciales and Savoryellales. Tolypocladium capitatum (OSC 110991) and T. japonicum (OSC 71233) (Hypocreales) were selected as outgroup taxa. Single-gene alignments were made in MAFFT v. 7 (http://mafft.cbrc.jp/alignment/server/ (accessed on 7 May 2022)) [27] and checked visually using AliView [28]. The alignments were trimmed using trimAl v 1.2 [29] with minimal coverage (-cons) = 0.8 and gap threshold (-gt) = 0.6. Five single-gene alignments were combined using SequenceMatrix 1.7.8 [30]. Maximum likelihood (ML), Bayesian inference (BI) and maximum parsimony (MP) analyses were employed to assess phylogenetic relationships as detailed in Dissanayake et al. [31].
ML analyses were performed with RAxML-HPC v.8 on XSEDE (8.2.12) [32,33] through the CIPRES Science Gateway V. 3.3 (https://www.phylo.org/portal2/login!input.action (accessed on 18 May 2022)) [34]. The tree search included 1000 non-parametric bootstrap replicates; the best scoring tree was selected among suboptimal trees from each run by comparing likelihood scores under the GTRGAMMA substitution model. The resulting replicates were plotted on to the best scoring tree obtained previously. ML bootstrap values equal to or greater than 75% were marked near each node.
BI was performed in MrBayes 3.2.6 [35]. The program MrModeltest 2 v. 2.3 [36] was used to determine the best nucleotide substitution model for each data partition. The GTR + I+G substitution model was decided for all LSU, SSU, ITS, RPB2 and TEF1α genes. Posterior probabilities (PP) [37] were determined by Markov chain Monte Carlo sampling (MCMC). Six simultaneous Markov chains were run for 10 million generations, and trees were sampled every 1000th generation. The first 25% of the saved trees, representing the burn-in phase of the analysis, were discarded. The remaining trees were used for calculating posterior probabilities in the majority rule consensus tree [38]. PP values equal to or greater than 0.95 were marked near each node.
MP analyses with the heuristic search were performed in PAUP v. 4.0 b10 [39]. The gaps in the alignment were treated as missing characters, and all characters were unordered. Maxtrees were unlimited, branches of zero length were collapsed, and all multiple, equally parsimonious trees were saved. Clade stability was assessed using a bootstrap (BT) analysis with 1000 replicates, each with 10 replicates of random stepwise addition of taxa [40]. MP bootstrap values equal to or greater than 75% were marked near each node.
Phylogenetic trees were printed with Fig Tree v. 1.4.4 (http://tree.bio.ed.ac.uk/software/figtree/ accessed on 18 July 2022)) and the layout was created in Adobe Illustrator CS6 software (Adobe Systems, Inc., San Jose, CA, USA). The sequences generated in this study were deposited in GenBank (Table 1).
Table 1. Taxa used in the phylogenetic analyses and the corresponding GenBank accession numbers.
Table 1. Taxa used in the phylogenetic analyses and the corresponding GenBank accession numbers.
TaxonSourceGenBank Accession NumberReferences
LSUSSUITSRPB2TEF1α
Ascotaiwania latericollaICMP 22739 TMN699407MN699390MN704312[16]
Ascotaiwania lignicolaNIL 00006HQ446365HQ446285HQ446342HQ446308[12]
Bactrodesmiastrum obovatumFMR 6482 TFR870266FR870264[41]
Bactrodesmiastrum pyriformeFMR 10747 TFR870265FR870263[41]
Bactrodesmiastrum pyriformeFMR 11931HE646637HE646636[41]
Bactrodesmiastrum monilioidesFMR 10756KF771879KF771878[10]
Bactrodesmium leptopusCBS 144542MN699423MN699374MN699388MN704297MN704321[16]
Bactrodesmium obovatumCBS 144407MN699426MN699377MN699397MN704299MN704324[16]
Canalisporium elegansSS 00895GQ390271GQ390256HQ446425HQ446311[12]
Canalisporium caribenseSS 03683GQ390269GQ390254[12]
Canalisporium grenadoidiaBCC 20507 TGQ390267GQ390252GQ390282HQ446420HQ446309[12]
Conioscypha hoehneliiFMR 11592 TKY853497HF937348KY853437[14]
Conioscypha japonicaCBS 387.84 TAY484514JQ437438JQ429259[42,43]
Conioscypha lignicolaCBS 335.93 TAY484513JQ437439JQ429260[42,43]
Conioscypha variaCBS 113653AY484512AY484511JQ429261[42,43]
Dematiosporium aquaticumMFLU 18-1641MK835855MN194029MN200286[15]
Fuscosporella aquaticaMFLUCC 16-0859MG388209MG388212[44]
Fuscosporella guizhouensisCGMCC 3.20884TOP376725OP376721OP376715OP367755OP367761This study
Fuscosporella guizhouensisUESTCC 22.0017OP376729OP376720OP376727OP367756OP367762This study
Fuscosporella pyriformisMFLUCC 16-0570 TKX550896KX550900MG388217KX576872[7]
Mucispora aquaticaCGMCC 3.20882TOP376717OP376726OP376713OP367752OP367757This study
Mucispora aquaticaUESTCC 22.0018OP376716OP376718OP376712OP367758This study
Mucispora infundibulataMFLUCC 16-0866 T MH457139MH457171MH457174[11]
Mucispora obscuriseptataMFLUCC 15-0618 TKX550892KX550897MG388218KX576870[7]
Mucispora phangngaensisMFLUCC 16-0865MG388210MG388213[44]
Neoascotaiwania fusiformisMFLUCC 15-0621 TKX550893MG388215KX576871[7]
Neoascotaiwania fusiformisMFLUCC 15-0625KX550894KX550898MG388216[7]
Neoascotaiwania guizhouensisCGMCC 3.20883TOP376731OP376719OP376728OP367753OP367759This study
Neoascotaiwania guizhouensisUESTCC 22.0019OP718560OP376730OP367754OP367760This study
Neoascotaiwania limneticaCBS 126576KY853513KT278689KY853452MN704308MN704331[8,14,16]
Neoascotaiwania limneticaCBS 126792KY853514KT278690KY853453MN704309MN704332[8,14,16]
Neoascotaiwania terrestrisCBS 144402MN699434MN699386MN699405MN704310MN704333[16]
Neoascotaiwania terrestrisCBS 142291 TKY853515KY853547KY853454[14,16]
ParafuscosporellamoniliformisMFLUCC 15-0626 TKX550895KX550899MG388219[7]
Parafuscosporella mucosaMFLUCC 16-0571 TMG388211MG388214[7]
Parafuscosporella pyriformisKUMCC 19-0008MN512340MN513031[45]
Parafuscosporella garethiiFF00725.01 TKX958430KX958429KX958432[46]
Parafuscosporella aquaticaKUMCC 19-0211 TMN512343MN513034[45]
Phaeoisaria aquaticaMFLUCC 16-1298 TMF399254MF399237MF401406[47]
Phaeoisaria fasciculataCBS 127885 TKT278705KT278693KT278719KT278741[8]
Plagiascoma frondosumCBS 139031 TKT278713KT278701KT278749[8]
Pleurotheciella erumpensCBS 142447 TMN699435MN699387MN699406MN704311MN704334[8]
Pleurotheciella guttulataKUMCC 15-0296 TMF399257MF399223MF399240MF401409[47]
Pleurothecium aquaticumMFLUCC 17-1331 TMF399263MF399245[47]
Pleurothecium floriformeMFLUCC 15-1163 TKY697277KY697279KY697281[48]
Pseudoascotaiwania persooniiA57 14C TAY094190[49]
Savoryella lignicolaNF 00204HQ446378HQ446300HQ446357HQ446334[12]
Savoryella nypaeMFLUCC 18-1570MK543210MK543237MK543219MK542516[50]
Tolypocladium capitatumOSC 71233AY489721AY489689DQ522421AY489615[51,52]
Tolypocladium japonicumOSC 110991DQ518761DQ522547DQ522428DQ522330[52]
Remarks: The superscript T denotes ex-type isolates. “−” denotes the sequence is unavailable. The newly generated sequences and new species are indicated in bold. Abbreviations: BCC: BIOTEC Culture Collection, Bangkok, Thailand; CBS: CBS−KNAW Fungal Biodiversity Centre, Utrecht, The Netherlands; CGMCC: China General Microbiological Culture Collection Center, Institute of Microbiology, Chinese Academy of Sciences, Beijing, China; FMR: Facultat de Medicina i Ciencies de la Salut, Reus, Spain; ICMP: International Collection of Microorganisms from Plants, Auckland, New Zealand; ILLS: University of Illinois Fungus Collection, Illinois, America; KUMCC: Kunming Institute of Botany Culture Collection, Kunming, China; MFLU: Mae Fah Luang University Herbarium Collection, Chiang Rai, Thailand; MFLUCC: Mae Fah Luang University Culture Collection, Chiang Rai, Thailand; OSC: Oregon State University Herbarium, Oregon, America; UESTCC: University of Electronic Science and Technology Culture Collection, Chengdu, China; Isolates with the prefix NF, NIL and SS, SAT are from the BIOTEC Culture Collection (BCC).

3. Phylogenetic Results

Five gene loci, LSU, SSU, ITS, RPB2, and TEF1α, were used to determine the phylogenetic placement of the new collections. The concatenated matrix was comprised of 50 taxa with a total of 4822 characters (LSU: 1–942 bp, SSU: 943–2168 bp, ITS: 2169–2794 bp, RPB2: 2795–3871 bp, TEF1α: 3872–4822 bp) including gaps. Single-gene analyses were carried out to compare the topologies and clade stabilities, respectively. The results showed that ML, MP and Bayesian inference (BI) were similar in topology without significant conflictions, and these results agree with previous studies [7,16,53]. The best scoring RAxML tree (−ln = −38 991.137) is shown in Figure 1.
In the phylogenetic analyses (Figure 1), isolates of Fuscosporella guizhouensis (CGMCC 3.20884 and UESTCC 22.0017) and Mucispora aquatica (CGMCC 3.20882 and UESTCC 22.0018) were distributed in Fuscosporellales. Two strains of Neoascotaiwania guizhouensis (CGMCC 3.20883 and UESTCC 22.0017) belonged to Savoryellales. Fuscosporella guizhouensis clustered together with F. aquatica (MFLUCC 16-0859) and F. pyriformis (MFLUCC 16-0570) and formed a strongly supported monophyletic clade representing the genus Fuscosporella (100% MLBS/1.00 PP/100% MPBS). Mucispora aquatica nested within the Mucispora clade and grouped with M. infundibulata (MFLU 18-0142) and M. obscuriseptata (MFLUCC 15-0618) without significant support. Neoascotaiwania guizhouensis clustered together with Neoascotaiwania taxa and was sister to N. terrestris (CBS 142,291 and CBS 144402).

4. Taxonomy

Fuscosporella guizhouensis H.Z. Du and Jian K. Liu, sp. nov., Figure 2.
MycoBank number: MB 845466.
Etymology: Referring to the location where the fungus was collected, Guizhou, China.
Holotype: HKAS 122794.
Saprobic on decaying wood in freshwater habitat. Sexual morph: Undetermined. Asexual morph: Colonies on natural substrate sporodochial, scattered, black, clustered on substrates. Mycelium partly immersed, partly superficial. Conidiophores micronematous, indistinct, branched, hyaline, smooth-walled. Conidiogenous cells monoblastic, integrated, terminal, globose, subglobose, ellipsoidal or clavate, hyaline to pale brown, 15−26 × 7−15 μm ( x ¯ = 20 × 11 μm, n = 20). Conidia solitary, acrogenous, ellipsoidal, hyaline when immature, dark brown to black when mature, smooth, (28.5−)42−60 × 24−34 μm ( x ¯ = 49.5 × 29 μm, n = 30), with obvious apical appendages, globose to ellipsoidal, or irregular shaped, connected in series.
Culture characteristics: Conidia germinated on WA within 24 h, and germ tubes produced from basal cell. Colonies growing on PDA reached 12−15 mm in diameter after one month at 25 °C, obverse olive to greyish green or dark greyish green in the inner, and light greyish green in the outer ring from above, reverse dark greyish green. Mycelium in culture up to 1–3 μm wide, subhyaline to brown, septate, branched. Conidiophores and conidiogenous cells indistinct. Chlamydospores are apparent in culture, globose to ellipsoidal or irregular shaped, hyaline at the beginning, becoming brown to black with ages, 10−21 × 5−19 μm ( x ¯ = 16 × 12 μm, n = 30).
Material examined: China, Guizhou Province, Guiyang City, Wudang District, Xiangsihe scenic spot, undisturbed forests with freshwater habitats, 26°26′51″ N, 106°37′53″ E, on decaying wood submerged in a freshwater stream, 22 February 2021, H.Z. Du, S99 (HKAS 122794, holotype); ex-holotype living culture CGMCC 3.20884; ibid., HUEST 22.0017, isotype, ex-isotype living culture UESTCC 22.0017.
Notes: Fuscosporella guizhouensis resembles F. pyriformis in forming sporodochial colonies and dark brown, smooth conidia. However, F. guizhouensis has larger conidiogenous cells (15−26 × 7−15 µm vs. 7.5−23 × 3.5−9 μm) and conidia (42−60 × 24−34 µm vs. 23.5−36 × 14−21 μm) [7]. The conidia of F. guizhouensis are irregular ellipsoidal, while F. pyriformis has obovoid to pyriform conidia. In addition, F. guizhouensis is distinguished by the hyaline apical appendages, which are absent in F. pyriformis and F. aquatica [44]. Fuscosporella guizhouensis can be distinguished from F. aquatica (17/859 in LSU, 60/524 in ITS) and from F. pyriformis (13/814 in LSU, 58/587 in ITS and 52/1024 in RPB2). Therefore, Fuscosporella guizhouensis is introduced as a new species, and this is the first Fuscosporella species reported from China.
Mucispora aquatica H.Z. Du and Jian K. Liu, sp. nov., Figure 3 and Figure 4.
MycoBank number: MB 845473.
Etymology: Referring to the aquatic habitat of this fungus.
Holotype: HKAS 122795.
Saprobic on decaying wood in freshwater habitat. Sexual morph: Undetermined. Asexual morph: Colonies on natural substrate effuse, glistening, black. Mycelium partly immersed, partly superficial, consisting of septate, smooth, hyaline to pale-brown hyphae, (1.5−) 2−4 (−6) um wide. Conidiophores macronematous, mononematous, solitary, erect, smooth, mid brown, paler towards the apex, straight or broadly curved, 2−8-septate, (41−)68−128 × 5−7.5 µm ( x ¯ = 104 × 6 µm, n = 20), with 1–2 percurrent proliferations. Conidiogenous cells monoblastic, integrated, terminal, cylindrical, pale brown to brown, 5−13 × 5−8 µm ( x ¯ = 10 × 6 µm, n = 20). Conidia acrogenous, ellipsoidal or obovoid, rarely pyriform, rounded at the apex and truncate at the base, smooth, dark brown to black, 34−43 µm ( x ¯ = 37 µm, n = 30) long, 17.5−23 µm ( x ¯ = 20 µm, n = 30) wide at broadest, 5.5−8 µm ( x ¯ = 7 µm, n = 30) wide at the base, septate with dark bands, becoming invisible when mature.
Culture characteristics: Conidia germinated on WA within 24 h, and germ tubes were produced from basal cell. Colonies growing on PDA reached 10−12 mm in diameter after one month at 25 °C, with light greyish green and dense mycelia on the surface, center elevated, reverse light grey. After one month, the diameter did not increase significantly. Mycelium subhyaline to pale brown, 2.5−4 µm wide in culture. Conidiophores light brown to brown, 14.5−37 × 4−6 µm ( x ¯ = 24 × 5 µm, n = 20). Conidiogenous cells integrated, subhyaline to pale brown, 5−7.5 × 5−8 µm ( x ¯ = 6 × 7 µm, n = 20). Conidia pale brown to black, 1−4-septate, mostly 2-septate, globose to obovoid, rounded at the apex and truncate at the base, smooth, constricted at the septa, 33−39 µm ( x ¯ = 36 µm, n = 30) long × 19−23 µm ( x ¯ = 21 µm, n = 30) wide at broadest, 5.5−9 um ( x ¯ = 7 µm, n = 30) wide at base.
Material examined: CHINA, Guizhou Province, Guiyang City, Wudang District, Xiangsihe scenic spot, undisturbed forests with freshwater habitats, 26°26′51″ N, 106°37′53″ E, on decaying wood submerged in a freshwater stream, 22 February 2021, H.Z. Du, S95 (HKAS 122795, holotype); ex-holotype living culture CGMCC 3.20882; ibid., HUEST 22.0018, isotype, ex-isotype living culture UESTCC 22.0018.
Notes: Mucispora aquatica resembles M. obscuriseptata, M. phangngaensis and M. infundibulate in forming scattered, dark brown to black colonies, macronematous, mononematous, solitary, erect, smooth conidiophores and acrogenous, ellipsoidal to obovoid conidia. However, M. aquatica is distinguished from M. obscuriseptata by the absence of conidial sheath [7]. The conidiophores of M. aquatica (68−128 µm) are smaller than those of M. obscuriseptata (80−170 µm) and M. phangngaensis (170−305 µm) [44] but larger than M. infundibulata (50−60 µm). Mucispora infundibulata is unique in its inflated cupulate conidiogenous cells [11]. In addition, Mucispora aquatica can be distinguished from M. infundibulata (30/836 in LSU, 70/606 in ITS); from M. phangngaensis (27/844 in LSU, 58/575 in ITS); and from M. obscuriseptata (35/861 in LSU, 69/605 in ITS and 60/879 in RPB2). Phylogenetic analysis (Figure 1) showed that Mucispora aquatica has a close phylogenetic relationship with M. infundibulata and M. obscuriseptata, but it can be recognized as a distinctly phylogenetic species. Therefore, we introduced Mucispora aquatica as a new species based on morphology and phylogeny.
Neoascotaiwania guizhouensis H.Z. Du and Jian K. Liu, sp. nov., Figure 5.
MycoBank number: MB 845474.
Etymology: Referring to the location where the fungus was collected, Guizhou Province, China.
Holotype: HKAS 122796.
Saprobic on decaying wood in freshwater habitat. Sexual morph: Undetermined. Asexual morph: Colonies on natural substrate sporodochial, glistening, black, clustered on substrates. Mycelium partly immersed, partly superficial. Conidiophores micronematous, mononematous, hyaline to pale brown, smooth, thin-walled. Conidiogenous cells monoblastic, cylindrical, hyaline to pale brown. Conidia solitary, ellipsoidal, pyriform to obovoid, broadly rounded or cuneate at the apex, 3−6 septate, pale brown when young, becoming dark brown to black when mature, paler at the basal cell, 49−62(−68) × 29−36(−39) μm ( x ¯ = 56 × 32 μm, n = 30).
Culture characteristics: Conidia germinated on WA within 24 h, and germ tubes produced from basal cell. Colonies growing on PDA reached 8−10 mm in diameter after one month at 25 °C, with white and dense mycelium on the surface, the center greyish green, reverse greyish brown and with a dark greyish brown ring in the middle. After one month, the diameter did not increase significantly. Mycelium hyaline to brown, septate, branched, 2−4 μm ( x ¯ = 3 μm, n = 30) wide, Chlamydospores are apparent, hyaline at the beginning, becoming brown or dark brown, 8−13 × 6−10 ( x ¯ = 11 × 9 μm, n = 30).
Material examined: CHINA, Guizhou Province, Guiyang City, Wudang District, Xiangsihe scenic spot, undisturbed forests with freshwater habitats, 26°26′51″ N, 106°37′53″ E, on decaying wood submerged in a freshwater stream, 22 February 2021, H.Z. Du, S95-2 (HKAS 122796, holotype); ex-holotype living culture CGMCC 3.20883; ibid., HUEST 22.0019, isotype, ex-isotype living culture UESTCC 22.0019.
Notes: Neoascotaiwania guizhouensis resembles N. limnetica and N. terrestris in forming dark, effuse colonies consisting of single, dark brown, transversely septate conidia. However, N. guizhouensis has larger conidia (49–68 × 29–39 μm) than those of N. limnetica (23–39 × 14.5–18.5 μm) and N. terrestris (25.5−44.5 × 13–22 μm) [8,14,16]. Furthermore, N. guizhouensis differs from N. fusiformis by its micronematous conidiophores, while the latter has macronematous conidiophores [7]. Additionally, Neoascotaiwania guizhouensis can be distinguished from N. terrestris (35/1017 in SSU, 14/554 in ITS, 13/1069 in RPB2 and 18/938 in TEF1α); from N. limnetica (13/862 in LSU, 36/561 in ITS, 36/845 in RPB2 and 28/884 in TEF1α); and from N. fusiformis (16/860 in LSU, 34/594 in ITS and 35/817 in RPB2). In our phylogenetic tree (Figure 1), Neoascotaiwania guizhouensis was sister to N. terrestris, but they are distinguishable in morphology and phylogeny. Therefore, we introduced Neoascotaiwania guizhouensis as a new species.

5. Discussion

The phylogenetic analyses based on the combined gene regions (LSU, SSU, ITS, RPB2 and TEF1α) placed three new species, Fuscosporella guizhouensis, Mucispora aquatica and Neoascotaiwania guizhouensis, in Fuscosporellaceae and Savoryellaceae (Savoryellomycetidae, Sordariomycetes) and are described in asexual stages without known sexual morphs. Species in Fuscosporella and Mucispora are reported from freshwater habitats in Thailand and China [7,11,44,54]; they may be exclusive in freshwater habitats. In this study, we provide the first record of Fuscosporella in China. Neoascotaiwania taxa are widely distributed in France, Spain and Thailand [7,8,14]. Neoascotaiwania guizhouensis, N. fusiformis, and N. limnetica are also found on decaying submerged wood in freshwater habitats [7,8,14,16,55], while N. terrestris was isolated from soil [14], which indicates that they are widely distributed and not limited by the growth environment.
The sexual morph of Neoascotaiwania differs from Ascotaiwania in having cylindrical asci with a thinner, non-amyloid and discoid apical ring, different septate ascospores and bactrodesmium-like asexual morph [14]. Besides, Ascotaiwania has monodictys-like [56], monotosporella-like [57,58] and trichocladium-like [56] asexual morphs. Dayarathne et al. [59] synonymized Neoascotaiwania under Ascotaiwania based on similar morphology and multi-gene phylogeny analysis. However, recent studies showed that Neoascotaiwania and Ascotaiwania were not congeneric [16,60]. We follow this treatment and treat Ascotaiwania and Neoascotaiwania as distinct genera.
Multi-locus phylogenetic analysis has been crucial for delimiting the novel fungi [61]. The use of multi-gene datasets to infer phylogenetic relationships has dramatically improved the resolution, especially when protein genes are combined with other genes, and the solution substantially increased [62,63]. For Fuscosporella and Mucispora, ITS, LSU and SSU rDNA datasets are available for all the species [7,11,44,54]. However, for the protein genes, only two species had the RPB2 sequence (unverified), and no TEF1α dataset. Therefore, the problem of low similarity occurred after the blastn search without a corresponding sequence in the same genus for alignment. This study provides the RPB2 and TEF1α sequences of Fuscosporella and Mucispora, which make up for the lack of protein genes in these two genera.

Author Contributions

Conceptualization, J.-K.L., H.-Z.D. and N.-G.L.; methodology, H.-Z.D. and J.Y.; formal analysis, H.-Z.D. and N.-G.L.; resources, H.-Z.D.; data curation, H.-Z.D.; writing—original draft preparation, H.-Z.D., J.Y. and N.-G.L.; writing—review and editing, H.-Z.D., N.-G.L., J.Y. and J.-K.L.; supervision, J.-K.L. and R.C.; project administration, J.-K.L.; funding acquisition, J.-K.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the Joint Fund of the National Natural Science Foundation of China and the Karst Science Research Center of Guizhou province (Grant No. U1812401).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The sequences data were submitted to GenBank.

Acknowledgments

H.-Z.D. thanks Na Wu for her assistance on morphological work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Thomas, K. Australian freshwater fungi. In Fungi of Australia; Introductory Volume to the Fungi (Part 2); Grgurinovic, C.A., Ed.; Australian Biological Resources Study: Canberra, ACT, Australia, 1996; Volume 1B, pp. 1–27. [Google Scholar]
  2. Wong, M.K.M.; Goh, T.-K.; Hodgkiss, I.J.; Hyde, K.D.; Ranghoo, V.M.; Tsui, C.K.M.; Ho, W.-H.; Wong, W.S.W.; Yuen, T.-K. Role of fungi in freshwater ecosystems. Biodivers. Conserv. 1998, 7, 1187–1206. [Google Scholar] [CrossRef]
  3. Benstead, J.P.; Rosemond, A.D.; Cross, W.F.; Wallace, J.B.; Eggert, S.L.; Suberkropp, K.; Gulis, V.; Greenwood, J.L.; Tant, C.J. Nutrient enrichment alters storage and fluxes of detritus in a headwater stream ecosystem. Ecology 2009, 90, 2556–2566. [Google Scholar] [CrossRef] [PubMed]
  4. Gareth Jones, E.B.; Eaton, R.A. Savoryella lignicola gen. et sp.nov. from water-cooling towers. Trans. Br. Mycol. Soc. 1969, 52, 161-IN114. [Google Scholar] [CrossRef]
  5. Udaiyan, K. Some interesting fungi from the industrial water cooling towers of Madras. II. J. Econ. Taxon. Bot. 1991, 15, 649–665. [Google Scholar]
  6. Calabon, M.S.; Hyde, K.D.; Jones, E.B.G.; Luo, Z.-L.; Dong, W.; Hurdeal, V.G.; Gentekaki, E.; Rossi, W.; Leonardi, M.; Thiyagaraja, V.; et al. Freshwater fungal numbers. Fungal Divers. 2022, 114, 3–235. [Google Scholar] [CrossRef]
  7. Yang, J.; Maharachchikumbura, S.S.N.; Bhat, D.J.; Hyde, K.D.; McKenzie, E.H.C.; Jones, E.B.G.; Al-Sadi, A.M.; Lumyong, S. Fuscosporellales, a new order of aquatic and terrestrial hypocreomycetidae (sordariomycetes). Cryptogam. Mycol. 2016, 37, 449–475. [Google Scholar] [CrossRef]
  8. Réblová, M.; Seifert, K.A.; Fournier, J.; Štěpánek, V. Newly recognized lineages of perithecial ascomycetes: The new orders conioscyphales and pleurotheciales. Persoonia 2016, 37, 57–81. [Google Scholar] [CrossRef] [Green Version]
  9. Věra, H.-J. Bactrodesmiastrum, a new genus of lignicolous hyphomycetes. Folia Geobot. Phytotaxon. 1984, 19, 103–106. [Google Scholar]
  10. Hernández-Restrepo, M.; Castañeda-Ruiz, R.F.; Guarro, J.; Gené, J.; Mena-Portales, J. Emendation of the genus Bactrodesmiastrum (Sordariomycetes) and description of Bactrodesmiastrum monilioides sp novfrom plant debris in Spain. Mycol Prog. 2015, 14, 48–54. [Google Scholar] [CrossRef]
  11. Hyde, K.D.; Norphanphoun, C.; Maharachchikumbura, S.S.N.; Bhat, D.J.; Jones, E.B.G.; Bundhun, D.; Chen, Y.J.; Bao, D.F.; Boonmee, S.; Calabon, M.S.; et al. Refined families of Sordariomycetes. Mycosphere 2020, 11, 305–1059. [Google Scholar] [CrossRef]
  12. Boonyuen, N.; Chuaseeharonnachai, C.; Suetrong, S.; Sri-Indrasutdhi, V.; Sivichai, S.; Jones, E.B.; Pang, K.L. Savoryellales (Hypocreomycetidae, Sordariomycetes): A novel lineage of aquatic ascomycetes inferred from multiple-gene phylogenies of the genera Ascotaiwania, Ascothailandia, and Savoryella. Mycologia 2011, 103, 1351–1371. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Jaklitsch, W.M.; Réblová, M. Savoryellaceae Jaklitsch & Réblová. Index Fungorum 2015, 209, 1. [Google Scholar]
  14. Hernández-Restrepo, M.; Gené, J.; Castañeda-Ruiz, R.F.; Mena-Portales, J.; Crous, P.W.; Guarro, J. Phylogeny of saprobic microfungi from Southern Europe. Stud. Mycol. 2017, 86, 53–97. [Google Scholar] [CrossRef] [PubMed]
  15. Luo, Z.-L.; Hyde, K.D.; Liu, J.-K.; Maharachchikumbura, S.S.N.; Jeewon, R.; Bao, D.-F.; Bhat, D.J.; Lin, C.-G.; Li, W.-L.; Yang, J.; et al. Freshwater sordariomycetes. Fungal Divers. 2019, 99, 451–660. [Google Scholar] [CrossRef] [Green Version]
  16. Réblová, M.; Hernández-Restrepo, M.; Fournier, J.; Nekvindová, J. New insights into the systematics of Bactrodesmium and its allies and introducing new genera, species and morphological patterns in the Pleurotheciales and Savoryellales (Sordariomycetes). Stud. Mycol. 2020, 95, 415–466. [Google Scholar] [CrossRef]
  17. Jones, E.B.G.; Hyde, K.D. Taxonomic studies on savoryella jones et eaton (Ascomycotina). Bot. Mar. 1992, 35, 83–91. [Google Scholar] [CrossRef]
  18. Jones, E.B.G.; Sakayaroj, J.; Suetrong, S.; Somrithipol, S.; Pang, K.L. Classification of marine ascomycota, anamorphic taxa and basidiomycota. Fungal Divers. 2009, 35, 187. [Google Scholar]
  19. Sri-indrasutdhi, V.; Boonyuen, N.; Suetrong, S.; Chuaseeharonnachai, C.; Sivichai, S.; Jones, E.B.G. Wood-inhabiting freshwater fungi from Thailand: Ascothailandia grenadoidia gen. et sp. nov., Canalisporium grenadoidia sp. nov. with a key to Canalisporium species (Sordariomycetes, Ascomycota). Mycoscience 2010, 51, 411–420. [Google Scholar] [CrossRef]
  20. Hongsanan, S.; Maharachchikumbura, S.S.N.; Hyde, K.D.; Samarakoon, M.C.; Jeewon, R.; Zhao, Q.; Al-Sadi, A.M.; Bahkali, A.H. An updated phylogeny of Sordariomycetes based on phylogenetic and molecular clock evidence. Fungal Divers. 2017, 84, 25–41. [Google Scholar] [CrossRef]
  21. Liu, J.K.; Chomnunti, P.; Cai, L.; Phookamsak, R.; Chukeatirote, R.; Jones, E.B.G.; Moslem, M.; Hyde, K.D. Phylogeny and morphology of Neodeightonia palmicola sp. nov. from palms. Sydowia 2010, 62, 261–276. [Google Scholar]
  22. Senanayake, I.; Calabon, M.S. Morphological approaches in studying fungi: Collection, examination, isolation, sporulation and preservation. Mycosphere 2020, 11, 2678–2754. [Google Scholar] [CrossRef]
  23. Vilgalys, R.; Hester, M. Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. J. Bacteriol. 1990, 172, 4238–4246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. White, T.J.; Bruns, T.; Lee, S.; Taylor, J. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In PCR Protocols: A Guide to Methods and Applications; Innis, M.A., Gelfand, D.H., Shinsky, J.J., White, T.J., Eds.; Academic Press, Inc.: New York, NY, USA, 1990; pp. 315–322. [Google Scholar] [CrossRef]
  25. Liu, Y.J.; Whelen, S.; Hall, B.D. Phylogenetic relationships among ascomycetes evidence from an RNA polymerse II subunit. Mol. Biol. Evol. 1999, 16, 1799–1808. [Google Scholar] [CrossRef] [Green Version]
  26. Rehner, S.A.; Buckley, E. A beauveria phylogeny inferred from nuclear ITS and EF1-a sequences evidence for cryptic diversification and links to Cordyceps teleomorphs. Mycologia 2005, 97, 84–98. [Google Scholar] [CrossRef]
  27. Katoh, K.; Standley, D.M. MAFFT multiple sequence alignment software version 7: Improvements in performance and usability. Mol. Biol. Evol. 2013, 30, 772–780. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. Larsson, A. AliView: A fast and lightweight alignment viewer and editor for large datasets. Bioinformatics 2014, 30, 3276–3278. [Google Scholar] [CrossRef] [Green Version]
  29. Capella-Gutierrez, S.; Silla-Martinez, J.M.; Gabaldon, T. trimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 2009, 25, 1972–1973. [Google Scholar] [CrossRef] [Green Version]
  30. Vaidya, G.; Lohman, D.J.; Meier, R. SequenceMatrix concatenation software for the fast assembly of multi gene datasets with character set and codon information. Cladistics 2011, 27, 171–180. [Google Scholar] [CrossRef]
  31. Dissanayake, A.J.; Bhunjun, C.S.; Maharachchikumbura, S.S.N.; Liu, J.K. Applied aspects of methods to infer phylogenetic relationships amongst fungi. Mycosphere 2020, 11, 2652–2676. [Google Scholar] [CrossRef]
  32. Stamatakis, A. RAxML-VI-HPC: Maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 2006, 22, 2688–2690. [Google Scholar] [CrossRef] [Green Version]
  33. Stamatakis, A.; Hoover, P.; Rougemont, J. A rapid bootstrap algorithm for the RAxML web servers. Syst. Biol. 2008, 57, 758–771. [Google Scholar] [CrossRef] [PubMed]
  34. Miller, M.A.; Pfeiffer, W.; Schwartz, T. Creating the CIPRES science gateway for inference of large phylogenetic trees. In Proceedings of the Gateway Computing Environments Workshop (GCE), New Orleans, LA, USA, 14 November 2010; Volume 14, pp. 1–8. [Google Scholar] [CrossRef] [Green Version]
  35. Ronquist, F.; Teslenko, M.; van der Mark, P.; Ayres, D.L.; Darling, A.; Hohna, S.; Larget, B.; Liu, L.; Suchard, M.A.; Huelsenbeck, J.P. MrBayes 3.2: Efficient bayesian phylogenetic inference and model choice across a large model space. Syst. Biol. 2012, 61, 539–542. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Nylander, J. MrModeltest2 v. 2.3 (Program for Selecting DNA Substitution Models Using PAUP*); Evolutionary Biology Centre: Uppsala, Sweden, 2008. [Google Scholar]
  37. Rannala, B.; Yang, Z. Probability distribution of molecular evolutionary trees a new method of phylogenetic inference. J. Mol. Evol. 1996, 43, 304–311. [Google Scholar] [CrossRef] [PubMed]
  38. Larget, B.; Simon, D.L. Markov chain monte carlo algorithms for the bayesian analysis of phylogenetic trees. Mol. Biol. Evol. 1999, 16, 750–759. [Google Scholar] [CrossRef] [Green Version]
  39. Swofford, D.L. PAUP*. Phylogenetic Analysis Using Parsimony (*and Other Methods), 4th ed.; Sinauer Associates: Sunderland, MA, USA, 2003. [Google Scholar]
  40. Hillis, D.M.; Bull, J.J. An empirical test of bootstrapping as a method for assessing confidence in phylogenetic analysis. Syst. Biol. 1993, 42, 182–192. [Google Scholar] [CrossRef]
  41. Hernández-Restrepo, M.; Mena-Portales, J.; Gené, J.; Cano, J.; Guaarro, J. New bactrodesmiastrum and bactrodesmium from decaying wood in Spain. Mycologia 2013, 105, 172–180. [Google Scholar] [CrossRef] [PubMed]
  42. Réblová, M.; Seifert, K.A. Conioscyphascus, a new ascomycetous genus for holomorphs with conioscypha anamorphs. Stud. Mycol. 2004, 50, 95–108. [Google Scholar]
  43. Réblová, M.; Seifert, K.A.; Fournier, J.; Stepánek, V. Phylogenetic classification of pleurothecium and pleurotheciella gen. nov. and its dactylaria-like anamorph (Sordariomycetes) based on nuclear ribosomal and protein-coding genes. Mycologia 2012, 104, 1299–1314. [Google Scholar] [CrossRef]
  44. Yang, J.; Liu, J.K.; Hyde, K.D.; Jones, E.B.G.; Liu, Z.Y. Two new species in Fuscosporellaceae from freshwater habitats in Thailand. Mycosphere 2017, 8, 1893–1903. [Google Scholar] [CrossRef]
  45. Yang, H.; Dong, W.; Yu, X.D.; Bhat, D.J.; Boonmee, S.; Zhang, H. Four freshwater dematiaceous hyphomycetes in sordariomycetes with two new species of parafuscosporella. Phytotaxa 2020, 441, 19–34. [Google Scholar] [CrossRef]
  46. Boonyuen, N.; Chuaseeharonnachai, C.; Suetrong, S.; Sujinda, S.; Somrithipol, S. Parafuscosporella garethii sp. nov. (Fuscosporellales) from a rivulet in a community-based northern forest, in Thailand. Mycosphere 2016, 7, 1265–1272. [Google Scholar] [CrossRef]
  47. Luo, Z.-L.; Hyde, K.D.; Bhat, D.J.; Jeewon, R.; Maharachchikumbura, S.S.N.; Bao, D.-F.; Li, W.-L.; Su, X.-J.; Yang, X.-Y.; Su, H.-Y. Morphological and molecular taxonomy of novel species pleurotheciaceae from freshwater habitats in Yunnan, China. Mycol. Prog. 2018, 17, 511–530. [Google Scholar] [CrossRef]
  48. Hyde, K.D.; Norphanphoun, C.; Abreu, V.P.; Bazzicalupo, A.; Thilini Chethana, K.W.; Clericuzio, M.; Dayarathne, M.C.; Dissanayake, A.J.; Ekanayaka, A.H.; He, M.-Q.; et al. Fungal diversity notes 603–708: Taxonomic and phylogenetic notes on genera and species. Fungal Divers. 2017, 87, 1–235. [Google Scholar] [CrossRef]
  49. Campbell, J.; Shearer, C.A. Annulusmagnus and ascitendus, two new genera in the annulatascaceae. Mycologia 2004, 96, 822–833. [Google Scholar] [CrossRef] [PubMed]
  50. Zhang, S.-N.; Abdel-Wahab, M.A.; Jones, E.B.G.; Hyde, K.D.; Liu, J.-K. Additions to the genus savoryella (savoryellaceae), with the asexual morphs savoryella nypae comb. nov. and S. sarushimana sp. nov. Phytotaxa 2019, 408, 195–207. [Google Scholar] [CrossRef]
  51. Castlebury, L.A.; Rossman, A.Y.; Sung, G.-H.; Hyten, A.S.; Spatafora, J.W. Multigene phylogeny reveals new lineage for Stachybotrys chartarum, the indoor air fungus. Mycol. Res. 2004, 108, 864–872. [Google Scholar] [CrossRef] [Green Version]
  52. Spatafora, J.W.; Sung, G.H.; Sung, J.M.; Hywel-Jones, N.L.; White, J.F., Jr. Phylogenetic evidence for an animal pathogen origin of ergot and the grass endophytes. Mol. Ecol. 2007, 16, 1701–1711. [Google Scholar] [CrossRef]
  53. Boonyuen, N.; Chuaseeharonnachai, C.; Nuankaew, S.; Kwantong, P.; Pornputtapong, N.; Suwannarach, N.; Jones, E.B.G.; Somrithipol, S. Novelties in fuscosporellaceae (fuscosporellales): Two new parafuscosporella from thailand revealed by morphology and phylogenetic analyses. Diversity 2021, 13, 517. [Google Scholar] [CrossRef]
  54. Wijayawardene, N.N.; Dissanayake, L.S.; Dai, D.-Q.; Li, Q.-R.; Xiao, Y.; Wen, T.-C.; Karunarathna, S.C.; Wu, H.-X.; Zhang, H.; Tibpromma, S.; et al. Yunnan–Guizhou Plateau: A mycological hotspot. Phytotaxa 2021, 523, 1–31. [Google Scholar] [CrossRef]
  55. Chang, H.S.; Hsieh, S.Y.; Jones, E.B.G.; Read, S.J.; Moss, S.T. New freshwater species of ascotaiwania and savoryella from Taiwan. Mycol. Res. 1998, 102, 709–718. [Google Scholar] [CrossRef]
  56. Chang, H.-s. Trichocladium anamorph of ascotaiwania hsilio and monodictys-like anamorphic states of ascotaiwania lignicola. Fung. Sci. 2001, 16, 35–38. [Google Scholar]
  57. Ranghoo, V.M.; Hyde, K.D. Ascomycetes from freshwater habitats: Ascolacicola aquatica gen. et sp. nov. and a new species of ascotaiwania from wood submerged in a reservoir in Hong Kong. Mycologia 1998, 90, 1055–1062. [Google Scholar] [CrossRef] [Green Version]
  58. Sivichai, S.; HyweI-Jones, N.; Jones, E.B.G. Lignicolous freshwater Ascomycota from Thailand: 1. Ascotaiwania sawada and its anamorph state monotosporella. Mycoscience 1998, 39, 307–311. [Google Scholar] [CrossRef]
  59. Dayarathne, M.C.; Maharachchikumbura, S.S.N.; Jones, E.B.G.; Dong, W.; Devadatha, B.; Yang, J.; Ekanayaka, A.H.; De Silva, W.; Sarma, V.V.; Al-Sadi, A.M.; et al. Phylogenetic revision of savoryellaceae and evidence for its ranking as a subclass. Front. Microbiol. 2019, 10, 840. [Google Scholar] [CrossRef] [Green Version]
  60. Dong, W.; Jeewon, R.; Hyde, K.D.; Yang, E.-F.; Zhang, H.; Yu, X.; Wang, G.; Suwannarach, N.; Doilom, M.; Dong, Z. Five novel taxa from freshwater habitats and new taxonomic insights of pleurotheciales and savoryellomycetidae. J. Fungi 2021, 7, 711. [Google Scholar] [CrossRef]
  61. Torres-Garcia, D.; García, D.; Cano-Lira, J.F.; Gené, J. Two novel genera, neostemphylium and scleromyces (pleosporaceae) from freshwater sediments and their global biogeography. J. Fungi 2022, 8, 868. [Google Scholar] [CrossRef]
  62. Tang, A.M.; Jeewon, R.; Hyde, K.D. Phylogenetic utility of protein (RPB2, beta-tubulin) and ribosomal (LSU, SSU) gene sequences in the systematics of sordariomycetes (Ascomycota, Fungi). Antonie Van Leeuwenhoek 2007, 91, 327–349. [Google Scholar] [CrossRef]
  63. Hsieh, H.-M.; Ju, Y.-M.; Rogers, J.D. Molecular phylogeny of hypoxylon and closely related genera. Mycologia 2005, 97, 844–865. [Google Scholar] [CrossRef]
Figure 1. Phylogenetic tree based on the combined LSU, SSU, ITS, RPB2 and TEF1α sequences constructed by maximum likelihood (RAxML) of selected members of Savoryellomycetidae (Sordariomycetes). Thickened branches indicate branch support with MLBS = 100%, PP = 1 and MPBS = 100%. Branch support for ML and MP greater than 75% and BI greater than 0.95 are marked above or below branches as MLBS/PP/MPBS. The abbreviation T indicates the ex-type strain. Species’ names and culture collections in bold are newly collected taxa. The tree was rooted with Tolypocladium capitatum (OSC 71233) and T. japonicum (OSC 110991).
Figure 1. Phylogenetic tree based on the combined LSU, SSU, ITS, RPB2 and TEF1α sequences constructed by maximum likelihood (RAxML) of selected members of Savoryellomycetidae (Sordariomycetes). Thickened branches indicate branch support with MLBS = 100%, PP = 1 and MPBS = 100%. Branch support for ML and MP greater than 75% and BI greater than 0.95 are marked above or below branches as MLBS/PP/MPBS. The abbreviation T indicates the ex-type strain. Species’ names and culture collections in bold are newly collected taxa. The tree was rooted with Tolypocladium capitatum (OSC 71233) and T. japonicum (OSC 110991).
Jof 08 01138 g001
Figure 2. Fuscosporella guizhouensis (HKAS 122794, holotype). (a,b) Colony on submerged wood. (c,d) Conidiophores with conidia. (eh) Conidia with apical appendages. (ip) Conidiogenous cells and conidia. (q) Germinated conidium. (r,s) Colony on PDA (r from above, s from below). Scale bars: (c) = 30 µm, (d) = 40 µm, (ep) = 30 µm, (q) = 40 µm.
Figure 2. Fuscosporella guizhouensis (HKAS 122794, holotype). (a,b) Colony on submerged wood. (c,d) Conidiophores with conidia. (eh) Conidia with apical appendages. (ip) Conidiogenous cells and conidia. (q) Germinated conidium. (r,s) Colony on PDA (r from above, s from below). Scale bars: (c) = 30 µm, (d) = 40 µm, (ep) = 30 µm, (q) = 40 µm.
Jof 08 01138 g002
Figure 3. Mucispora aquatica (HKAS 122795, holotype). (a,b) Colonies on submerged wood. (ch) Conidiophores with conidia. (i,j) Conidiogenous cells. (k) Mycelium. (lo) Conidia. (p) Germinated conidium. (q,r) Colony on PDA ((q) from above, (r) from below). Scale bars: (ch) = 40 µm, (ip) = 20 µm.
Figure 3. Mucispora aquatica (HKAS 122795, holotype). (a,b) Colonies on submerged wood. (ch) Conidiophores with conidia. (i,j) Conidiogenous cells. (k) Mycelium. (lo) Conidia. (p) Germinated conidium. (q,r) Colony on PDA ((q) from above, (r) from below). Scale bars: (ch) = 40 µm, (ip) = 20 µm.
Jof 08 01138 g003
Figure 4. Reproduced asexual morph of Mucispora aquatica (CGMCC 3.20882, ex-holotype) on PDA medium. (a,b) Colonies on PDA. (ce) Hyphae and conidiophores with conidia. (fo) Conidiogenous cells and conidia. Scale bars: (c) = 40 µm, (d) = 50 µm, (eo) = 20 µm.
Figure 4. Reproduced asexual morph of Mucispora aquatica (CGMCC 3.20882, ex-holotype) on PDA medium. (a,b) Colonies on PDA. (ce) Hyphae and conidiophores with conidia. (fo) Conidiogenous cells and conidia. Scale bars: (c) = 40 µm, (d) = 50 µm, (eo) = 20 µm.
Jof 08 01138 g004
Figure 5. Neoascotaiwania guizhouensis (HKAS 122796, holotype). (ac) Colonies on submerged wood. (dh) Conidiophores with conidia. (ip) Conidia. (q) Germinated conidium. (r,s) Colony on PDA ((r) from above, s from below). Scale bars: (d) = 40 µm, (ep) = 20 µm, (q) = 40 µm.
Figure 5. Neoascotaiwania guizhouensis (HKAS 122796, holotype). (ac) Colonies on submerged wood. (dh) Conidiophores with conidia. (ip) Conidia. (q) Germinated conidium. (r,s) Colony on PDA ((r) from above, s from below). Scale bars: (d) = 40 µm, (ep) = 20 µm, (q) = 40 µm.
Jof 08 01138 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Du, H.-Z.; Yang, J.; Liu, N.-G.; Cheewangkoon, R.; Liu, J.-K. Morpho-Phylogenetic Evidence Reveals New Species of Fuscosporellaceae and Savoryellaceae from Freshwater Habitats in Guizhou Province, China. J. Fungi 2022, 8, 1138. https://doi.org/10.3390/jof8111138

AMA Style

Du H-Z, Yang J, Liu N-G, Cheewangkoon R, Liu J-K. Morpho-Phylogenetic Evidence Reveals New Species of Fuscosporellaceae and Savoryellaceae from Freshwater Habitats in Guizhou Province, China. Journal of Fungi. 2022; 8(11):1138. https://doi.org/10.3390/jof8111138

Chicago/Turabian Style

Du, Hong-Zhi, Jing Yang, Ning-Guo Liu, Ratchadawan Cheewangkoon, and Jian-Kui Liu. 2022. "Morpho-Phylogenetic Evidence Reveals New Species of Fuscosporellaceae and Savoryellaceae from Freshwater Habitats in Guizhou Province, China" Journal of Fungi 8, no. 11: 1138. https://doi.org/10.3390/jof8111138

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop