Nine New Records of Ascomycetes from Different Niches in Korea

Research Article
Monmi Pangging1Thuong Thuong Thi Nguyen1Hyang Burm Lee1*

Abstract

We isolated nine fungal strains from different environmental materials collected from different locations during a survey of fungal diversity in Korea. Using molecular phylogenetic analyses and morphological characteristics, nine previously undescribed strains were identified and assigned to the species Collariella robusta, Fusicolla acetilerea, Hongkongmyces pedis, Hongkongmyces snookiorum, Mariannaea fusiformis, Metarhizium pemphigi, Pallidocercospora crystallina, Scopulariopsis candida, and Volutella citrinella. Diverse environmental samples may thus be a promising source for isolating and investigating novel fungal species, thus sampling efforts should be increased in future studies. This study also reports identification of some rare fungal species belonging to the genera Hongkongmyces and Pallidocercospora from Korea.

Keyword



INTRODUCTION

The Ascomycetes is the largest phylum encompassing more than 33,000 species and a vast number of undescribed fungi. Information on the diversity and ecology of Korean fungal species belonging to the Ascomycetes (especially Dothideomycetes and Sordariomycetes) is still lacking. Dothideomycetes, previously known as Loculoascomycetes [1,2] is the largest group of ascomycetes with more than 23 orders, 119 families, 1,261 genera, and 19,000 species [3]. This group includes ecologically diverse fungi often found as plant pathogens, as saprobes degrading cellulose and other complex carbohydrates in dead or partially digested plant matters, in leaf litter or dung, as endophytes or epiphytes on living plants, some as coprophilous species, and a few as lichen-forming fungi [4]. They occur in almost every part of the world ranging from terrestrial to freshwater and marine habitats. Members of Dothideomycetes are characterized morphologically by ascolocular ascoma development and bitunicate, fissitunicate asci [1,5-8]. As regards to Sordariomycetes, this class is the second largest and comprises of 6 subclasses, 28 orders, 90 families, and 1,344 genera [9,10]. Species of Sordariomycetes have a cosmopolitan distribution, accommodates mostly terrestrial taxa as endophytes, saprobes, epiphytes, plant pathogens, fungicolous, lichenized or lichenicolous fungi and aquatic taxa as pathogens of arthropods and mammals [9-13]. Members of Sordariomycetes are mainly characterized by non-lichenized, perithecial ascomata and inoperculate unitunicate or non-fissitunicate asci [14,15]. Dothideomycetes and Sordariomycetes species have the capacity to produce an extensive array of secondary metabolites across the fungal kingdom [16,17]. Some of the secondary metabolites act as potential biocontrol agents and also exploited for various medicinal and industrial uses [18-21] whereas some are phytotoxic or mycotoxic [22].

The fungal taxa belonging to Dothideomycetes and Sordariomycetes have been documented well in Thailand, UK, China, India, and Japan [23,24-26]. The knowledge on the ecology of the species belonging to Dothideomycetes and Sordariomycetes are limited in Korea despite their role in ecosystem health, global carbon cycling as saprotrophs and degraders of plant biomass and metabolic profiles [27-29]. Therefore, the objective of this study is to characterize the unrecorded nine fungal species in Korea using both morphological and molecular analyses: C. robusta, F. acetilereae, H. pedis, H. snookiorum, M. fusiformis, Met. pemphigi, P. crystallina, S. candida and V. citrinella.

MATERIALS AND METHODS

Fungal strains isolation

The information related to sample sources, location and sampling sites are listed in Table 1. A dilution plating method was used for the isolation of fungal strains from freshwater, soil, and seawater. For isolating fungi from plant samples, a previously described methods was employed [30]. Pure isolates were transferred to new fresh potato dextrose Agar (PDA; Difco, Sparks, MD, USA) and were incubated for 4-5 days at 25℃. For stock storage, pure isolates were maintained in PDA slant tubes and 20% glycerol at -80℃ in the Environmental Microbiology Laboratory Fungarium, Chonnam National University, Gwangju, Korea as CNUFC BCSM3, CNUFC F7-24-5, CNUFC HRG1, CNUFC HRW1-12, CNUFC YJS7, CNUFC AS1-26, CNUFC PLTFB118, CNUFC KU1-1, and CNUFC DYR1-2. CNUFC BCSM3, CNUFC AS1-26, CNUFC PLTFB118, CNUFC KU1-1, CNUFC F7-24-5, CNUFC HRG1, CNUFC HRW1-12, CNUFC YJS7, and CNUFC DYR1-2, were also deposited at the Collection of National Institute of Biological Resources (NIBR) Incheon, Korea as NIBRFG0000503047, IMYKFGC000000029, NIBRFG0000503054, NIBRFGC000508425 and Nakdonggang National Institute of Biological Resources (NNIBR), Sangju, Korea as NNIBRFG9312, NNIBRFG31611, NNIBRFG31612, NNIBRFG9315 and NNIBRFG9334, respectively.

Morphological analysis

All the isolated strains were grown on PDA, malt extract agar (MEA; malt extract, 20 g; agar, 20 g; distilled water, 1 L), oatmeal agar (OA; oatmeal flakes, 30 g; agar, 20 g; distilled water, 1 L), and cornmeal agar (cornmeal 30 g; agar, 15 g; distilled water, 1 L) at 25℃ for morphological characterization. Samples were examined using an Olympus BX51 microscope with DIC optics (Olympus, Tokyo, Japan) by mounting in a lactophenol solution (Junsei Chemical Co., Ltd., Tokyo, Japan).

Table 1. The isolates information used in this study.http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Table_KJOM_49_03_01_T1.png

DNA extraction, PCR and sequencing

Fungal genomic DNA was extracted using a Solg Genomic DNA preparation Kit (Solgent Co., Ltd., Daejeon, Korea). Amplification of fragments of the gene regions of the large subunit of the nuclear ribosomal DNA (LSU), the internal transcribed spacers (ITS) and beta-tubulin (TUB) were performed using the following primer pairs LR0R/LR5 [31], ITS5/ITS4 [32], ITS5/LR7 [31,32], ITS1/NL4 [32,33] and T10/Bt2b [34]. PCR amplification was performed according to the conditions described by Hongsanan et al. [8], Luo et al. [12], and Sandoval-Denis et al. [35]. After amplification, PCR products were purified using an Accuprep PCR Purification Kit (Bioneer Corp., Daejeon, Korea). Sequencing was performed using the same primers as those used for amplification with an ABI PRISM 3730XL Genetic Analyzer (Applied Biosystems, Foster City, CA, USA).

Phylogenetic analysis

A homology search of DNA sequences was performed using the BLASTn algorithm of the National Center for Biotechnology Information (https://www.ncbi.nlm.nih.gov). Each strains sequences were aligned with reference sequences downloaded from GenBank using Clustal X version 2.1 [36] and edited manually using Bioedit version 7.2.6.0 [37]. Maximum likelihood (ML) phylogenies were constructed using MEGA version 7 [38].

ITS sequences of strains CNUFC BCSM3, CNUFC F7-24-5, CNUFC YJS7, CNUFC PLTFB118, CNUFC KU1-1, and CNUFC DYR1-2 were deposited in the GenBank database under accession numbers MZ435755, MZ435756, MZ435757, MZ435758, MZ435759 and MZ435760, respectively. LSU sequences of strains CNUFC BCSM3, CNUFC F7-24-5, CNUFC HRG1, CNUFC HRW1-12, CNUFC YJS7, CNUFC PLTFB118, and CNUFC DYR1-2 were deposited in the GenBank database under accession numbers MZ436990, MZ436991, MZ436992, MZ436993, MZ436994, OK021601 and MZ436995, respectively. TUB sequences of strains CNUFC AS1-26 and CNUFC KU1-1 were deposited in the GenBank database under accession numbers MZ450120 and MZ450121, respectively.

RESULTS

Phylogenetic analysis

According to the BLASTn, ITS sequences of CNUFC BCSM3, CNUFC F7-24-5, CNUFC YJS7, CNUFC PLTFB118, CNUFC KU1-1, and CNUFC DYR1-2 showed similarities of 100% (508/508 bp), 99.9% (1,045/1,046 bp), 98.7% (1,002/1,015 bp), 99.3% (1,052/1,060 bp), 99.8% (567/568 bp), and 100% (508/508 bp) with those of C. robusta CBS 508.84 (MH861773), F. acetilerea RF3 (LC333211), M. elegans NRRL 62999 (KM056317), P. crystallina CBS 681.95 (EU167579), S. candida CBS 389.52 (KX924012) and V. citrinella VC-101 (MK357063), respectively. Similarly, BLASTn using LSU sequences of CNUFC BCSM3, CNUFC F7-24-5, CNUFC HRG1, CNUFC HRW1-12, CNUFC YJS7, CNUFC PLTFB118, and CNUFC DYR1-2, showed similarities of 99.1% (541/546 bp), 99.3% (554/558 bp), 99.6% (512/514 bp), 99.5% (582/585 bp), 99.8% (550/551 bp), 99.7% (565/567 bp), and 99.8% (566/567 bp) with those of C. robusta CBS 508.84 (MH873473), F. siamensis MFLUCC 17-2577 (MT215550), H. pedis HKU35 (NG_056287), H. snookiorum ILLS00125755 (MH161189), M. fusiformis LC1701 (KX986140), P. crystallina ZJUM 2 (KP895884) and V. citrinella DAOM 226716 (HQ843770), respectively. BLASTn using TUB sequence of CNUFC AS1-26 and CNUFC KU1-1, showed similarities of 100% (614/614 bp) and 99.60% (493/495 bp) with Met. pemphigi ARSEF 7491 (KJ398591) and S. candida CBS 205.27 (KX924444), respectively. Phylogram generated from the Maximum Likelihood analysis for ITS, LSU, and TUB revealed that the strains, CNUFC BCSM3, CNUFC F7-24-5, CNUFC HRG1, CNUFC HRW1-12, CNUFC YJS7, CNUFC AS1-26, CNUFC PLTFB118, CNUFC KU1-1 and CNUFC DYR1-2 were identical to C. robusta, F. acetilerea, H. pedis, H. snookiorum, M. fusiformis, Met. pemphigi, P. crystallina, S. candida and V. citrinella, respectively (Figs. 1-8).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F1.png

Fig. 1. Phylogenetic tree based on the maximum likelihood analyses of the combined internal transcribed spacers (ITS) and large subunit (LSU) rDNA sequences of Collariella robusta CNUFC BCSM3. Microascus trigonosporus CBS 218.31 was used as outgroup. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type.

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F2.png

Fig. 2. Phylogenetic tree based on the maximum likelihood analysis of the combined internal transcribed spacers (ITS) and large subunit (LSU) rDNA sequences of Fusicolla acetilerea CNUFC F7-24-5. Microcera larvarum CBS 738.79 and M. coccophila CBS 310.34 were used as outgroups. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type.

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F3.png

Fig. 3. Phylogenetic tree based on the maximum likelihood analysis of large subunit (LSU) rDNA sequences of Hongkongmyces pedis CNUFC HRG1 and Hongkongmyces snookiorum CNUFC HRW1-12. Salsuginea ramicola KT2597.2 was used as outgroup. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type.

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F4.png

Fig. 4. Phylogenetic tree based on the maximum likelihood analysis of the combined internal transcribed spacers (ITS) and large subunit (LSU) rDNA sequences of Mariannaea fusiformis CNUFC YJS7. Stachybotrys chartarum CBS 129.13 was used as outgroup. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F5.png

Fig. 5. Phylogenetic tree based on the maximum likelihood analysis of beta-tubulin (TUB) sequences of Metarhizium pemphigi CNUFC AS1-26. Metapochonia rubescens CBS 464.88 and Metapochonia microbactrospora CBS 101433 were used as outgroups. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type.

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F6.png

Fig. 6. Phylogenetic tree based on the maximum likelihood analysis of the combined internal transcribed spacers (ITS) and large subunit (LSU) rDNA sequences of Pallidocercospora crystallina CNUFC PLTFB118. Mycosphaerella madeirae CBS 112895 was used as outgroup. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type, HT=ex-holotype.

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F7.png

Fig. 7. Phylogenetic tree based on the maximum likelihood analysis of internal transcribed spacers (ITS) rDNA and beta-tubulin (TUB) sequences of Scopulariopsis candida CNUFC KU1-1. Pseudoscopulariopsis schumacheri CBS 435.86 was used as outgroup. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type.

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F8.png

Fig. 8. Phylogenetic tree based on the maximum likelihood analysis of the combined internal transcribed spacers (ITS) and large subunit (LSU) rDNA sequences of Volutella citrinella CNUFC DYR1-2. Wardomycopsis longicatenata CGMCC 3.17947, Microascus anfractus CGMCC 3.17950, and M. globulosus CGMCC 3.17927 were used as outgroups. Bootstrap support values of ≥50% from 1,000 replicates are indicated at the nodes. The bar indicates the number of substitutions per position. T=type, HT=ex-holotype.

Taxonomy

Collariella robusta (L.M. Ames) X. Wei Wang & Samson, Studies in Mycology 84: 217 (2016) (Fig. 9).

≡Chaetomium robustum L.M. Ames, A monograph of the Chaetomiaceae: 35 (1963).

≡Chaetomium caprinum Bainier, Bulletin de la Société Mycologique de France 25 (3): 223 (1910).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F9.png

Fig. 9. Morphology of Collariella robusta CNUFC BCSM3. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA) (A-C: Obverse view, D-F: Reverse view). (G, H) Mature ascomata. (I) Asci. (J) Ascospores (scale bars: G, H=20 μm, I, J=10 μm).

Description: Colonies on PDA plane, regular, cottony, white, reverse white, measuring 24-28 mm in diameter after 7 d at 25℃. Colonies on MEA plane, regular, transparent white, reverse transparent white, measuring 23-26 mm after 7 d at 25℃. Colonies on OA regular, white, reverse pale white, measuring 23-27 mm after 7 d at 25℃.

Micromorphology: Ascomata were ampulliform, elongated and measuring 260-410 × 160-303 μm. Asci were club-shaped, 8-spored and measuring 32-37 × 8-11 μm. Ascospores were limoniform and measuring 5.9-6.5 × 4.7-5.6 μm.

Remarks: Species of Collariella is morphologically characterized by the production of a dark collar-like apex around the ostiolar pore of the ascomata [39]. CNUFC BCSM3 shares similar morphological features to C. robusta [39]. Also, the phylogeny based on combined ITS and LSU rDNA placed CNUFC BCSM3 with C. robusta CBS 551.83 (T). Collariella robusta was reported to occur in litter and woodlot soil from Jamaica and in rabbit pellets from the United Kingdom [39,40]. The species was isolated from seawater providing further information regarding the ecological characteristics of this species.

Fusicolla acetilerea (Tubaki, C. Booth & T. Harada) Gräfenhan & Seifert, Studies in Mycology 68: 100 (2011) (Fig. 10).

≡Fusarium merismoides var. acetilereum Tubaki, C. Booth & T. Harada, Transactions of the British Mycological Society 66 (2): 355 (1976).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F10.png

Fig. 10. Morphology of Collariella robusta CNUFC BCSM3. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA) (A-C: Obverse view, D-F: Reverse view). (G, H) Mature ascomata. (I) Asci. (J) Ascospores (scale bars: G, H=20 μm, I, J=10 μm).

Description: Colonies on PDA floccose, regular, pale yellow orange, reverse pale yellowish brown, measuring 22-24 mm after 7 d at 25℃. Colonies on MEA plane, regular, transparent white, reverse pale white, measuring 20-23 mm after 7 d at 25℃. Colonies on OA plane, regular, white mycelium to pale orange toward center, reverse white mycelium to pale orange, measuring 23-25 mm after 7 d at 25℃.

Micromorphology: Macroconidia were long, slender, falcate, with 3-4 transverse septa and measuring 40-58 × 2.5-4 μm. Chlamydospore was not observed. No perithecia were produced.

Remarks: The genus Fusicolla is characterized by superficial, nonstromatic, globose to sub-globose, pale yellow to pale orange ascomata, one-septate, ellipsoidal, hyaline and straight or slightly curved ascospores [41,42]. Fusicolla acetilerea was described with phialidic, straight or curved with a round apex and a foot cell macroconidia mostly with 3(-4) transverse septa when mature, measuring (30-)35-40 × 3.5(-4.0) μm [43]. However, CNUFC F7-24-5 isolate can be differentiated from previous described F. acetilerea by producing relatively longer macroconidia size. Fusicolla acetilerea was found to occur mostly in soil habitats, on decaying wood, and on the beetle Xyleborinus saxesenii [41,44,45]. CNUFC F7-24-5 is reported to occur in a freshwater environment for the first time.

Hongkongmyces pedis C.C. Tsang, J.F.W. Chan, N.J. Trendell-Smith, A.H.Y. Ngan, I.W.H. Ling, S.K.P. Lau & P.C.Y. Woo, Medical Mycology 52 (7): 740 (2014) (Fig. 11).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F11.png

Fig. 11. Morphology of Hongkongmyces pedis CNUFC HRG1. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on cornmeal agar (CMA). (A-C: Obverse view, D-F: Reverse view). (G) Dark hyphae (observed under stereomicroscope). (H, I) Only sterile mycelia consisted of gray and septate hyphae consists of round and spore-like bodies shown by arrow (scale bars=20 μm).

Description: Colonies on PDA greyish green, slow growing, reverse slight yellow pigments, measuring 21-23 mm after 14 d at 25℃. Colonies on MEA slightly irregular, grey, velvety, reverse dark olive, measuring 26-27 mm after 14 d at 25℃. Colonies on CMA slightly regular, dark olive grey, reverse olive, measuring 23-26 mm after 14 d at 25℃.

Micromorphology:Only sterile mycelia were produced after 3 months of incubation on all three media. On PDA, hyphae become darkly pigmented and round, spore-like bodies occasionally present along the hyphae. Hyphae were septate, narrow, and branched. Mycelia were sterile. No fruiting bodies or conidia were produced.

Remarks: The genus Hongkongmyces was erected with H. pedis HKU35 (T) isolated from biopsy tissues of a patient’s infected foot as type species [46]. Morphologically H. pedis HKU35 (T) was characterized by forming grey colonies on media, composed of grey, narrow, septate, branched hyphae with acute angles, sterile mycelia with no fruiting bodies or sporulating structures. CNUFC HRG1 is morphologically similar to H. pedis HKU35 (T) with producing sterile mycelia and narrow, septate, branched hyphae. Similarly, CNUFC HRG1 clustered within the same clade with H. pedis HKU35 (T) in the phylogeny based on LSU rDNA.

Hongkongmyces snookiorum Raudabaugh, Iturr. & A.N. Mill., Persoonia 40: 289 (Fig. 12).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F12.png

Fig. 12. Morphology of Hongkongmyces pedis CNUFC HRG1. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on cornmeal agar (CMA). (A-C: Obverse view, D-F: Reverse view). (G) Dark hyphae (observed under stereomicroscope). (H, I) Only sterile mycelia consisted of gray and septate hyphae consists of round and spore-like bodies shown by arrow (scale bars=20 μm).

Description: Colonies on PDA regular, dull greyish green, reverse dark olive green, measuring 23-25 mm after 14 d at 25℃. Colonies on MEA regular, hyaline, olive, reverse olive grey, measuring 26-28 mm after 14 d at 25℃. Colonies on CMA regular, pale olive, reverse olive grey, measuring 24-27 mm after 14 d at 25℃.

Micromorphology: Conidiogenous cells were discrete, hyaline, phialidic, subulate and tightly aggregated. Conidia were hyaline, ellipsoid to ovoid and measuring 4.3-5.3 × 3.2-4 μm.

Remarks: Hongkongmyces snookirorum is coelomycetous with globose to ampulliform pycnidia, hyaline, subulate to ampulliform conidiogenous cells with sympodial proliferations, hyaline, ellipsoid to ovoid conidia [47]. The species was also reported from freshwater and appeared to be identical with that previously isolated from detritus submerged in a freshwater fen [47].

Mariannaea fusiformis D.M. Hu & L. Cai, Mycological Progress 16: 278 (2017) (Fig. 13).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F13.png

Fig. 13. Morphology of Mariannaea fusiformis CNUFC YJS7. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA). (A-C: Obverse view, D-F: Reverse view). (G-I) Conidiophores and verticillate branches of phialides. (J) Conidia (scale bars: G, I, J=20 μm, H= 50 μm).

Description: Colonies on PDA floccose, zonate, white to purple and yellowish at the center, reverse purple and yellowish at the center, measuring 22-23 mm after 7 d at 25℃. Colonies on MEA plane, regular, pale purple, reverse pale purple, measuring 26-28 mm after 7 d at 25℃. Colonies on OA plane, regular, purple, light cottony mycelium, reverse pale purple, measuring 24-26 mm after 7 d at 25℃.

Micromorphology: Conidiophores were erect, septate, cylindrical, macronematous, mononematous and measuring 550-800 μm long. Phialides were hyaline, smooth-walled, ampulliform and measuring 16.9-35.8 × 4-5 μm. Conidia were fusiform to subglobose, hyaline, smooth, and aseptate, produced in imbricate chains and measuring 4.9-7 × 3-4 μm. Chlamydospores were hyaline, thick-walled, intercalary or terminal and measuring 6.6-9.2 × 4.9-6.5 μm.

Remarks: The genus Mariannaea is characterized by flask-shaped phialides, distinctly stalked and one- or two-celled conidia forming imbricate or straight chains [48]. Mariannaea fusiformis CGMCC 3.17272 (T) isolated from submerged wood is characterized by its purple colonies on PDA, fusiform to subglobose conidia [49]. CNUFC YJS7 also shows purple colonies cultured on PDA, MEA and OA media and fusiform to subglobose conidia. CNUFC YJS7 is reported from freshwater. The species was isolated from freshwater showing that the species are associated with freshwater habitat.

Metarhizium pemphigi (Driver & Milner) Kepler, S.A. Rehner & Humber, Mycologia 106: 824 (2014) (Fig. 14).

≡Metarhizium flavoviride var. pemphigi Driver & Milner, Mycological Research 104 (2): 144 (2000).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F14.png

Fig. 14. Morphology of Metarhizium pemphigi CNUFC AS1-26. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA). (A-C: Obverse view, D-F: Reverse view). (G, H) Basipetal chains comprising conidia and conidiophores forming a sporulating layer as conidial columns. (I) Conidia (scale bars: G=10 μm, H, I=20 μm).

Description: Colonies on PDA floccose, cottony white, reverse pale white, measuring 24-26 mm after 7 d at 25℃. Colonies on MEA plane, regular, pale orange, white cottony towards the center, reverse pale orange, measuring 18-20 mm after 7 d at 25℃. Colonies on OA plane, regular, white to light green, reverse pale orange, measuring 25-27 mm after 7 d at 25℃.

Micromorphology: Conidiophores formed a sporulating layer as conidial columns; basipetal chains were formed in conidiation. Conidia were uninucleate and ovoid to cylindrical and measuring 4.8-7.4 × 2.0-2.3 μm.

Remarks: The genus Metarhizium is characterized by the production of conidia in long chains, dense phialides in parallel arrangement, and aggregated conidiophores with repeated, verticillate branching [50,51]. Metarhizium pemphigi is characterized by ovoid to elongate conidia, measuring 5.4 (±0.47)×2.4 (±0.43) μm, occasionally <9 μm long, borne in chains on cylindrical phialides and light green spore [52]. Metarhizium pemphigi was previously isolated from forest or wood soil, Pemphigus treherni, other Hemiptera, and Melanotus cribricollis [52-55]. Our isolate CNUFC AS1-26 was isolated from soil environment.

Pallidocercospora crystallina (Crous & M.J. Wingf.) Crous & M.J. Wingf., Studies in Mycology 75: 74 (2012) (Fig. 15).

≡Pseudocercospora crystallina Crous & M.J. Wingf., Mycologia 88 (3): 451 (1996).

≡Mycosphaerella crystallina Crous & M.J. Wingf., Mycologia 88 (3): 451 (1996).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F15.png

Fig. 15. Morphology of Pallidocercospora crystallina CNUFC PLTFB118. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA). (A-C: Obverse view, D-F: Reverse view). (G) Plant sample. (H, I) Dark hyphae (observed under stereomicroscope). (J, K) Hyphae (scale bars=20 μm).

Description: Colonies on PDA velvety, olive green, radially concentrated toward the center, reverse iron-grey, measuring 23-26 mm after 7 d at 25℃. Colonies on MEA regular, plane, peanut brown, reverse olive green, measuring 22-24 mm after 7 d at 25℃. Colonies on CMA regular, olive green, reverse dark olive, measuring 19-21 mm after 7 d at 25℃.

Micromorphology: Filamentous septate hyphae showed pale brown, thickened walls; no sporulation were observed.

Remarks: The genus Pallidocercospora was introduced to include Cercospora-like species however not congeneric with Cercospora and is typified by P. heimii Crous (termed Pseudocercospora heimii Crous) [56]. Previously, Pallidocercospora was believed to contain Mycosphaerella heimii complex species based on culture characteristics, with pale brown cercosporoid conidia, and segregated from Pseudocercospora species mainly by the distinct phylogenetic positions of its members. CNUFC PLTFB118 shows similar morphology to P. crystallina reported to cause infections in humans [57]. Similarly, Huang et al. [58] also reported that P. crystallina isolated from spot leaves of citrus cannot be induced to sporulate on artificial media. To the best of our knowledge, this is the first report of P. crystallina occurring as an endophyte in Korea.

Scopulariopsis candida Vuill., Bulletin de la Société Mycologique de France 27: 143 (1911) (Fig. 16).

≡Monilia candida Guég: 271 (1899).

≡Chrysosporium keratinophilum var. denticola C. Moreau, Mycopathologia et Mycologia Applicata 37 (1): 37 (1969).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F16.png

Fig. 16. Morphology of Scopulariopsis candida CNUFC KU1-1. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA). (A-C: Obverse view, D-F: Reverse view). (G-I) Conidiophores, annellides and conidia. (J) Conidia (scale bars=20 μm).

Description: Colonies on PDA irregular, white to cream-colored, reverse pale orange, measuring 24-28 mm after 7 d at 25℃. Colonies on MEA irregular, cream-colored, pale cream-colored, measuring 22-25 mm after 7 d at 25℃. Colonies on OA plane, regular, cottony white towards the center, reverse pale white, measuring 21-24 mm after 7 d at 25℃.

Micromorphology: Conidiophores were abundant. Conidia were subglobose to broadly ovate, hyaline smooth and measuring 4.3-8.3 × 4.4-7.3 μm.

Remarks: The genus Scopulariopsis is characterized by annellidic conidiogenesis with mostly thick-walled, basally truncate conidia arranged in long, dry chains and colony color ranging from white to brown or black [59]. Scopulariopsis candida is characterized by subglobose to broadly ovate, hyaline, smooth-walled conidia, irregularly ellipsoidal asci, broadly lunate to reniform ascospores and white colonies [35]. CNUFC KU1-1 differs from S. candida [35] by cream color colonies and no sexual morphology is observed. Previously S. candida has been reported from environmental samples (air, dust, and soil), predominantly in the Northern Hemisphere, especially Europe and North America; also isolated from clinical samples, mainly from superficial tissue of humans and animals suffering from onychomycosis [35,60,61]. This is the first report on the isolation of the species from seawater environment.

Volutella citrinella (Cooke &Massee) Seifert, Studies in Mycology 68: 110 (2011) (Fig. 17).

≡Stilbella aciculosa (Ellis & Everh.) Seifert, Studies in Mycology 37: 44 (1985).

≡Stilbum aciculosum Ellis & Everh., Journal of Mycology 1 (12): 153 (1885).

≡Botryonipha aciculosa (Ellis & Everh.) Kuntze, Revisio generum plantarum 2: 845 (1891).

≡Stilbella bulbicola Henn., Hedwigia 44: 176 (1905).

-Stilbum bulbicola (Henn.) M.A. Litv., Opredelitel' Mikroskopicheskikh Pochvennykh Gribov: 196 (1967).

http://dam.zipot.com:8080/sites/KJOM/images/N0320490301_image/Fig_KJOM_49_03_01_F17.png

Fig. 17. Morphology of Volutella citrinella CNUFC DYR1-2. (A, D) Colonies on potato dextrose agar (PDA). (B, E) Colonies on malt extract agar (MEA). (C, F) Colonies on oatmeal agar (OA). (A-C: Obverse view, D-F: Reverse view). (G, H) Determinate synnemata developed in culture (observed under stereomicroscope). (I, J) Conidiophores. (K) Conidia (scale bars=20 μm).

-Stilbum bulbicola (Henn.) Sacc., Sylloge Fungorum 18: 633 (1906).

-Tubercularia bulbicola (Henn.) Sacc. (?).

≡Stilbella flavescens Estey, Transactions of the British Mycological Society 68 (1): 122 (1977).

≡Stilbum pallidulum Penz. & Sacc., Malpighia 15: 250 (1902).

≡Stilbum citrinellum Cooke & Massee, Grevillea 16 (79): 81 (1888).

-Volutella citrinella (Cooke & Massee) Seifert, Studies in Mycology 68: 110 (2011).

Description: Colonies on PDA floccose, zonate, white to creamy pink, reverse white to pink, measuring 21-25 mm after 7 d at 25℃. Colonies on MEA plane, regular, pale white, reverse pale white, measuring 25-26 mm after 7 d at 25℃. Colonies on OA plane, plane, regular, pale white to light creamy pink towards the center, reverse pale white, measuring 25-28 mm after 7 d at 25℃.

Micromorphology: Hyphae of synnema stipes thick-walled may appear seta-like when diverging from the synnema and measuring 1.3-3.0 μm wide with cell walls. Conidia were smooth and subglobose elliptical and measuring 4.1-9.3 × 2.2-2.6 μm.

Remarks: The genus Volutella includes species characterized by discoid sporodochia with marginal setae, compact and phialidic conidiogenous cells, simple to verticillate conidiophores, one-celled, ovoid to oblong conidia [19,62,63]. Volutella citrinella has been previously isolated from the debris of Solanum tuberosum and soil samples [9,41,64]. This is the first report on the occurrence of the species in a freshwater environment in Korea.

DISCUSSION

Phylogeny based on combined ITS-LSU rDNA and individual LSU rDNA sequences have placed seven isolates, CNUFC BCSM3, CNUFC F7-24-5, CNUFC HRG1, CNUFC HRW1-12, CNUFC YJS7, CNUFC PLTFB118, and CNUFC DYR1-2 with their respective type species of C. robusta, F. acetilerea, H. pedis, H. snookiorum, M. fusiformis, P. crystallina and V. citrinella, respectively. Since ITS do not aid in proper identification of genus Metarhizium and Scopulariopsis, TUB sequences for CNUFC AS1-26 and CNUFC KU1-1 were amplified [35,53,65]. Metarhizium pemphigi F1-72 (T) and FI-1101 isolated from root aphids lack the TUB gene [52], therefore Met. pemphigi isolates ARSEF 6569 and ARSEF 7491 of TUB gene have been used in the phylogeny [53,65]. Combined ITS-TUB phylogeny placed CNUFC KU1-1 with S. candida MUCL 40743 (T) [66].

Out of ten accepted Collariella species, only one species of Collariella carteri has been reported from Korea [67-70]. Species of Collariella were reported to produce a variety of metabolites, including chaetoquadrin E, cochliodinol B, cochliodones 1-3, prenisatin, and SB236049/SB236050/SB238569 [39,69,71]. Fusicolla comprises of 18 species, however only 14 of these are represented by rDNA sequences in GenBank [23,72-74]. Fusicolla merismoides was isolated from brackish water in Korea [75]. Fusicolla acetilerea was found to degrade triacrylonitrile (1,3,6-hexanetricarbonitrile) as a nitrogen source and 4-N-trimethylamino-1-butanol as a sole source of carbon and nitrogen [76,77]. There is no detailed study on the metabolites produced by four species described in Hongkongmyces [78]. According to the Index Fungorum 2021 (www.indexfungorum.org), the genus Mariannaea comprises of 23 species but only 18 species have GenBank records [49,79-81]. So far, three species have been reported from Korea [82]. Mariannaea species are known to produce extracellular enzymes such as amylases, β-glucosidases, proteases, and cellulases and thus can degrade cellobiase, starch and xylan [83,84]. Mariannaea elegans found to produce mariannamides A and B, new cyclic octapeptides and successfully mariannamides A showing antimicrobial activities against E. coli and C. neoformans [85]. Two species of Metarhizium have been reported from Korea from the accepted 34 species [86-88]. In particular, Metarhizium species have been exploited for the control of a wide range of arthropod pests such as plague locusts in Africa [89] as well as mosquito vectors of malaria [90]. Moreover Met. pemphigi is a promising biological control agent of ticks [91]. Eight Pallidocercospora species are accepted in this genus reported particularly as plant pathogens [92]. In total, 77 species are accepted in this genus [35,59] and 7 new species were introduced recently [69,93,94]. Scopulariopsis brevicaulis was found in Korea as a clinical isolate after cosmetic face surgery [95]. Scopulariopsis species known to cause deterioration of building materials [96,97] and accumulate various elements converting these into toxic volatiles [98,99] which makes them an important group to study in indoor environments. About 147 records of Volutella are listed in Index Fungorum 2021. with eight species having available sequences in GenBank [69,73,100]. Volutella ciliata was isolated from crop soil in Korea [101]. Volutella citrinella may be a candidate to produce new terpendole congeners N-P (1-3) and voluhemins that act as inhibitors of sterol-O-acyltransferase [64,102].

Results of this study clearly showed that there is relatively diverse group of species belonging to Dothideomycetes and Sordariomycetes associated with freshwater, soil, seawater, and plant samples from Korea. DNA sequences, especially those of the ITS and LSU rDNA regions have proven to be reliable for identifying species belonging to the Dothideomycetes and Sordariomycetes and some additional markers have also been introduced. The current study thus contributes to the knowledge on the fungal diversity from Dothideomycetes and Sordariomycetes in Korea with respect to their cosmopolitan distributions and environmental relevance. Future research is expected to cover the identification of important biochemical components and their activities and more genomes of Dothideomycetes and Sordariomycetes fungi become easily accessible and available.

ACKNOWLEDGMENTS

This study was financially supported by Chonnam National University [Grant number: 2017-2827]. This work was in part supported by the Graduate Program for the Undiscovered Taxa of Korea, the Project on Survey and Discovery of Indigenous Fungal Species of Korea funded by NIBR, and the Project on Discovery of Fungi from Freshwater and Collection of Fungarium, funded by NNIBR of the Ministry of Environment.

References

1 1. Nannfeldt JA. Studien uber die morphologie und systematik der nicht–lichenisierten, inoperkulaten Discomyceten. Nova Acta Regiae Soc Sci Upsal IV 1932;8:1-368.  

2 2. Liu YJ, Hall DB. Body plan evolution of ascomycetes, as inferred from an RNA polymerase II phylogeny. Proc Natl Acad Sci USA 2004;101:4507-12.  

3 3. Wijayawardene NN, Hyde KD, Rajeshkumar KC, Hawksworth DL, Madrid H, Kirk PM, Braun U, Singh RV, Crous PW, Kukwa M, et al. Notes for genera: Ascomycota. Fungal Divers 2017;86:1-594.  

4 4. Schoch CL, Crous PW, Groenewald JZ, Boehm EWA, Burgess TI, Gruyter JD, de Hoog GS, Dixon LJ, Grube M, Gueidan C, et al. A class-wide phylogenetic assessment of Dothideomycetes. Stud Mycol 2009;64:1-15.  

5 5. Luttrell ES. The ascostromatic Ascomycetes. Mycologia 1955;47:511-32.  

6 6. Eriksson OE. The families of bitunicate ascomycetes. Nordic J Bot 1981;1:800.  

7 7. Barr ME, Huhndorf SM. Loculoascomycetes. In: McLaughlin DJ, McLaughlin EG, Lemke PA editors. The mycota VII, part A. Systematics and evolution. Berlin: Springer Verlag; 2001. p. 283-305.  

8 8. Hongsanan S, Hyde KD, Phookamsak R, Wanasinghe DN, McKenzie EHC, Sarma VV, Lücking R, Boonmee S, Bhat JD, Liu NG, et al. Refined families of Dothideomycetes: Orders and families incertae sedis in Dothideomycetes. Fungal Divers 2020;105:17-318.  

9 9. Maharachchikumbura SSN, Hyde KD, Jones EBG, McKenzie EHC, Bhat JD, Dayarathne MC, Huang SK, Norphanphoun C, Senanayake IC, Perera RH, et al. Families of Sordariomycetes. Fungal Divers 2016;79:1-317.  

10 10. Hongsanan S, Maharachchikumbura SSN, Hyde KD, Samarakoon MC, Jeewon R, Zhao Q, Al-Sadi AM, Bahkali AH. An updated phylogeny of Sordariomycetes based on phylogenetic and molecular clock evidence. Fungal Divers 2017;84:25-41.  

11 11. Tang AMC, Jeewon R, Hyde KD. Phylogenetic utility of protein (RPB2, β-tubulin) and ribosomal (LSU, SSU) gene sequences in the systematics of Sordariomycetes (Ascomycota, Fungi). Antonie van Leeuwenhoek 2007;91:327-49.  

12 12. Luo ZL, Hyde KD, Bhat DJ, Jeewon R, Maharachchikumbura SSN, Bao DF, Li WL, Su XJ, Yang XY, Su HY. Morphological and molecular taxonomy of novel species Pleurotheciaceae from freshwater habitats in Yunnan, China. Mycol Prog 2018;17:511-30.  

13 13. Yang J, Liu NG, Liu JK, Hyde KD, Jones EBG, Liu ZY. Phylogenetic placement of Cryptophiale, Cryptophialoidea, Nawawia, Neonawawia gen. nov. and Phialosporostilbe. Mycosphere 2018;9:1132-50.  

14 14. Zhang N, Castlebury LA, Miller AN, Huhndorf SM, Schoch CL, Seifert KA, Rossman AY, Rogers JD, Kohlmeyer J, Volkmann-Kohlmeyer B, et al. An overview of the systematics of the Sordariomycetes based on four gene phylogeny. Mycologia 2006;98:1076-87.  

15 15. Kirk PM, Cannon PF, Minter DW, Stalpers JA. Dictionary of the fungi, 10th edn. Wallingford: CABI; 2008.  

16 16. Spatafora JW, Bushley KE. Phylogenomics and evolution of secondary metabolism in plant associated fungi. Curr Opin Plant Biol 2015;26:37-44.  

17 17. Akimitsu K, Tsuge T, Kodama M, Yamamoto M, Otani H. Alternaria host-selective toxins: Determinant factors of plant disease. J Plant Pathol 2014;80:109-22.  

18 18. Kaewchai S, Soytong K, Hyde KD. Mycofungicides and fungal biofertilizers. Fungal Divers 2009;38:25-50.  

19 19. Semenova EF, Shpichka AI, Moiseeva IY. About essential oils biotechnology on the base of microbial synthesis. Pharmaceutical Sci 2012;4:29-31.  

20 20. Xu J, Yang X, Lin Q. Chemistry and biology of Pestalotiopsis derived natural products. Fungal Divers 2014;66:37-68.  

21 21. Hyde KD, Xu J, Rapior S, Jeewon R, Lumyong S, Niego AGT, Abeywickrama PD, Aluthmuhandiram JVS, Brahamanage RS, Brooks S, et al. The amazing potential of fungi: 50 ways we can exploit fungi industrially. Fungal Divers 2019;31:1-36.  

22 22. Stergiopoulos I, Collemare J, Mehrabi R, De Wit PJGM. Phytotoxic secondary metabolites and peptides produced by plant pathogenic Dothideomycete fungi. FEMS Microbiol Rev 2013;37:67-93.  

23 23. Jones EBG, Devadatha B, Abdel-Wahab MA, Dayarathne MC, Zhang SN, Hyde KD, Liu JK, Bahkali AH, Sarma VV, Tibell S, et al. Phylogeny of new marine Dothideomycetes and Sordariomycetes from mangroves and deep-sea sediments. Botanica Marina 2020;63:155-81.  

24 24. Ramakrishnan D, Tiwari MK, Manoharan G, Sairam T, Thangamani R, Lee JK, Marimuthu J. Molecular characterization of two alkyresorcylic acid synthases from Sordariomycetes fungi. Enzyme Microbiol Technol 2018;115:16-22.  

25 25. Hyde KD, Norphanphoun C, Maharachchikumbura SSN, Bhat DJ, Jones EBG, Bundhun D, Chen YJ, Bao DF, Boonmee S, Calabon MS, et al. Refined families of Sordariomycetes. Mycosphere 2020;11:305-1059.  

26 26. Jayasiri SC, Hyde KD, Jones EBG, McKenzie EHC, Jeewon R, Phillips AJL, Bhat DJ, Wanasinghe DN, Liu JK, Lu YZ, et al. Diversity, morphology and molecular phylogeny of Dothideomycetes on decaying wild seed pods and fruits. Mycosphere 2019;10:1-186.  

27 27. Das K, Lee SY, Jung HY. Morphology and phylogeny of two novel species within the class Dothideomycetes collected from soil in Korea. Mycobiol 2021;49:15-23.  

28 28. Goh J, Mun HY, Jeon YJ, Chung N, Park YW, Park S, Hwang H, Cheon W. First report of six Sordariomycetes fungi isolated from plant litter in freshwater ecosystems of Korea. Kor J Mycol 2020;48:103-16.  

29 29. Nam B, Lee JS, Lee HB, Choi YJ. Pezizomycotina (Ascomycota) fungi isolated from freshwater environments of Korea: Cladorrhinum australe, Curvularia muehlenbeckiae, Curvularia pseudobrachyspora, and Diaporthe longicolla. Kor J Mycol 2020;48:29-38.  

30 30. Eo JK, Park E, Choe HN. Dermea piceina (Dermateaceae): An unrecorded endophytic fungus of isolated from Abies koreana. Kor J Mycol 2020;48:485-9.  

31 31. Vilgalys R, Hester M. Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. J Bacteriol 1990;172:4238-46.  

32 32. White TJ, Bruns TD, Lee S, Taylor JW. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. PCR Protocols: A guide to methods and applications. In: Innis MA, Gelfand DH editors. London: Academic Press; 1990. p. 315-22.  

33 33. O’Donnell K. Fusarium and its near relatives. Wallingford: UK; 1993.  

34 34. Glass NL, Donaldson GC. Development of primer sets designed for use with the PCR to amplify conserved genes from filamentous ascomycetes. Appl Environ Microbiol 1995;61:1323-30.  

35 35. Sandoval-Denis M, Gené J, Sutton DA, Cano-Lira JF, De Hoog GS, Decock CA, Wiederhold NP, Guarro J. Redefining Microascus, Scopulariopsis and allied genera. Persoonia 2016;36:1-36.  

36 36. Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins DG. The CLUSTAL_X windows interface: Flexible strategies for multiple sequence alignment aided by quality analysis tools. Nucleic Acids Res 1997;25:4876-82.  

37 37. Hall TA. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symp Ser 1999;41:95-8.  

38 38. Kumar S, Stecher G, Tamura K. MEGA7: Molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol 2016;33:1870-4.  

39 39. Wang XW, Houbraken J, Groenewald JZ, Meijer M, Andersen B, Nielsen KF, Crous PW, Samson RA. Diversity and taxonomy of Chaetomium and Chaetomium-like fungi from indoor environments. Stud Mycol 2016;84:145-224.  

40 40. von Arx JA, Guarro J, Figueras MJ. The ascomycete genus Chaetomium. Beih Nova Hedwigia 1986;84:1-162.  

41 41. Gräfenhan T, Schroers HJ, Nirenberg HI, Seifert KA. An overview of the taxonomy, phylogeny, and typification of nectriaceous fungi in Cosmospora, Acremonium, Fusarium, Stilbella, and Volutella. Stud Mycol 2011;68:79-113.  

42 42. Lechat C, Rossman A. A new species of Fusicolla (Hypocreales), F. ossicola, from Belgium. Ascomycete org 2017;9:225-8.  

43 43. Tubaki K, Booth C, Harada T. A new variety of Fusarium merismoides. Trans Br Mycol Soc 1976;66:355-66.  

44 44. Ding S, Hu H, Gu JD. Fungi colonizing wood sticks of Chinese fir incubated in subtropical urban soil growing with Ficus microcarpa trees. Int J Env Sci Technol 2015;12:3781-90.  

45 45. Biedermann PHW, Klepzig KD, Taborsky M, Six DL. Abundance and dynamics of filamentous fungi in the complex ambrosia gardens of the primitively eusocial beetle Xyleborinus saxesenii Ratzeburg (Coleoptera: Curculionidae, Scolytinae). FEMS Microbiol Ecol 2013;83:711-23.  

46 46. Tsang CC, Chan JFW, Trendell-Smith NJ, Ngan AHY, Ling IWH, Lau SKP, Woo PCY. Subcutaneous phaeohyphomycosis in a patient with IgG4-related sclerosing disease caused by a novel ascomycete, Hongkongmyces pedis gen. et sp. nov.: First report of human infection associated with the family Lindgomycetaceae. Med Mycol 2014;52:736-47.  

47 47. Crous PW, Wingfield MJ, Burgess TI, Hardy GESt, Gené J, Guarro J, Baseia IG, Garcia D, Gusmao LFP, Souza-Motta CM, et al. Fungal Planet description sheets: 716-784. Persoonia 2018;40:240-393.  

48 48. Samson RA. Paecilomyces and some allied hyphomycetes. Stud Mycol 1974;6:1-119.  

49 49. Hu DM, Wang M, Cai L. Phylogenetic assessment and taxonomic revision of Mariannaea. Mycol Prog 2017;16:271-83.  

50 50. Sorokin N. Rastitelnye parazity cheloveka i zhivotnykn’ kak’ prichina zaraznykn’ boleznei (Plant parasites causing infectious diseases of man and animals). Vyp. II.Pervoe prilozenie k Voenno-Meditsinskomu Zhurnalu za 1883 g (First supplement to J Military Med 1883). St. Petersburg: Izdanie glavnogo Voenno-Meditsinskago Upraveleneia; 1883. p. 168-98.  

51 51. Tulloch M. The genus Metarhizium. Trans Br Mycol Soc 1976;66:407-11.  

52 52. Driver F, Milner RJ, Trueman JWH. A taxonomic revision of Metarhizium based on a phylogenetic analysis of rDNA sequence data. Mycol Res 2000;104:134-50.  

53 53. Nishi O, Hasegawa K, Iiyama K, Yasunaga-Aoki C, Shimizu S. Phylogenetic analysis of Metarhizium spp. isolated from soil in Japan. App Entomol Zool 2011;46:301-9.  

54 54. Chen ZH, Zhang YG, Yang XN, Chen K, Liu Q, Xu L. A new fungus Metarhizium gaoligongense from China. Int J Agric Biol 2018;20:2271-6.  

55 55. Montalva C, Collier K, Rocha LFN, Inglis PW, Lopes RB, Luz C, Humber RA. A natural fungal infection of a sylvatic cockroach with Metarhizium blattodeae sp. nov., a member of the M. flavoviride species complex. Fungal Biol 2016;120:655-65.  

56 56. Crous PW, Braun U, Hunter GC, Wingfield MJ, Verkley GJM, Shin HD, Nakashima C, Groenewald JZ. Phylogenetic lineages in Pseudocercospora. Stud Mycol 2013;75:37-114.  

57 57. Guo Y, Zhu Z, Gao J, Zhang C, Zhang X, Dang E, Li W, Qiao H, Liao W, Wang G, et al. The phytopathogenic fungus Pallidocercospora crystallina-caused localized subcutaneous phaeohyphomycosis in a patient with a homozygous missense CARD9 mutation. J Clin Immunol 2019;39:713-25.  

58 58. Huang F, Groenewald JZ, Zhu L, Crous PW, Li H. Cercosporoid diseases of citrus. Mycologia 2016;107:1151-71.  

59 59. Samson RA, Houbraken J, Thrane U, Frisvad JC, Andersen B. Food and indoor fungi. CBS Laboratory Manual Series 2. Utrecht: CBS-KNAW Fungal Biodiversity Centre; 2010.  

60 60. de Hoog GS, Guarro J, Gene J, Figueras MJ. Atlas of clinical fungi. CD-ROM version 3.1. Utrecht: CBS-KNAW Fungal Biodiversity Centre; 2011.  

61 61. Sandoval-Denis M, Sutton DA, Fothergill AW, Cano-Lira J, Gené J, Decock CA, de Hoog GS, Guarro J. Scopulariopsis, a poorly known opportunistic fungus: spectrum of species in clinical samples and in vitro responses to antifungal drugs. J Clin Microbiol 2013;51:3937-43.  

62 62. Fries EM. Systema Mycologicum. E. Moritz, Greifswald, Germany 1832;3:261-524.  

63 63. Lombard L, van der Merwe NA, Groenewald JZ, Crous PW. Generic concepts in Nectriaceae. Stud Mycol 2015;80:189-245.  

64 64. Ohshiro T, Morita H, Nur EAA, Hosoda K, Uchida R, Tomoda H. Voluhemins, new inhibitors of sterol-O-acyltransferase, produced by Volutella citrinella BF-0440. J Antibiot 2020;73:748-55.  

65 65. Kepler RM, Humber RA, Bischoff JF, Rehner SA. Clarification of generic and species boundaries for Metarhizium and related fungi through multigene phylogenetics. Mycologia 2014;106:811-29.  

66 66. Woudenberg JHC, Meijer M, Houbraken J, Samson RA. Scopulariopsis and Scopulariopsis-like species from indoor envionments. Stud Mycol 2017;88:1-35.  

67 67. Aghyl H, Mehrabi-Koushki M, Esfandiari M. Chaetomium iranicum and Collariella capillicompacta spp. nov. and notes to new hosts of Amesia species in Iran. Sydowia 2020;73:21-30.  

68 68. Crous PW, Wingfield MJ, Burgess TI, Carnegie AJ, Hardy GEStJ, Smith D, Summerell BA, Cano-Lira JF, Guarro J, Houbraken J, et al. Fungal planet description sheets: 625-715. Persoonia 2017;39:270-467.  

69 69. Zhang ZF, Liu F, Zhou X, Liu XZ, Liu SJ, Cai L. Culturable mycobiota from Karst caves in China, with descriptions of 20 new species. Persoonia 2017;39:1-31.  

70 70. Nguyen TTT, Lee SH, Jeon SJ, Lee HB. First records of rare ascomycete fungi, Acrostalagmus luteoalbus, Bartalinia robillardoides, and Collariella carteri from freshwater samples in Korea. Mycobiology 2019;47:1-11.  

71 71. Crous PW, Wingfield MJ, Burgess TI, Carnegie AJ, Hardy GEST, Smith D, Summerell BA, Cano-Lira JF, Guarro J, Houbraken J, et al. Fungal planet description sheets: 625-715. Persoonia 2017;39:270-467.  

72 72. Dayarathne MC, Jones EBG, Maharachchikumbura SSN, Devadatha B, Sarma VV, Khongphinitbunjong K, Chomnunti P, Hyde KD. Morphomolecular characterization of microfungi associated with marine based habitats. Mycosphere 2020;11:1-188.  

73 73. Perera RH, Hyde KD, Maharachchikumbura SSN, Jones EBG, McKenzie EHC, Stadler M, Lee HB, Samarakoon MC, Ekanayaka AH, Camporesi E, et al. Fungi on wild seeds and fruits. Mycosphere 2020;11:2108-480.  

74 74. Luo ZL, Hyde KD, Liu JK, Maharachchikumbura SSN, Jeewon R, Bao DF, Bhat DJ, Lin CG, Li WL, Yang J, et al. Freshwater Sordariomycetes. Fungal Divers 2019;99:451-660.  

75 75. Jeon YJ, Goh J, Mun HY. Diversity of fungi in brackish water in Korea. Kor J Mycol 2020;48:457-73.  

76 76. Fujimitsu H, Taniyama Y, Tajima S, Ahmed IAM, Arima J, Mori N. Purification and characterization of 4-N-trimethylamino-1-butanol dehydrogenase from Fusarium merismoides var. acetilereum. Bio Biotech Biochem 2016;80:1753-58.  

77 77. Asano Y, Ando S, Tani Y, Yamada H, Ueno T. Fungal degradation of Triacrylonitrile. Agric Biol Chem 1981;45:57-62.  

78 78. Dong W, Wang B, Hyde KD, McKenzie EHC, Raja HA, Tanaka K, Abdel-Wahab MA, Abdel-Aziz FA, Doilom M, Phookamsak R, et al. Freshwater Dothideomycetes. Fungal Divers 2020;105:319-575.  

79 79. Hyde KD, Dong Y, Phookamsak R, Jeewon R, Bhat DJ, Jones EBG, Liu NG, Abeywickrama PD, Mapook A, Wei D, et al. Fungal diversity notes 1151-1276: Taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Divers 2020;100:5-277.  

80 80. Cai L, Kurniawati E, Hyde KD. Morphological and molecular characterization of Mariannaea aquaticola sp. nov. collected from freshwater habitats. Mycol Prog 2010;9:337-43.  

81 81. Crous PW, Carnegie AJ, Wingfield MJ, Sharma R, Mughini G, Noordeloos ME, Santini A, Shouche YS, Bezerra JDP, Dima B, et al. Fungal planet description sheets: 868-950. Persoonia 2019;42:291-473.  

82 82. Nguyen TTT, Pangging M, Lee HB. Three unrecorded fungal species from fecal and freshwater samples in Korea. Kor J Mycol 2017;45:304-18.  

83 83. Tang L, Hyun MW, Yun YH, Suh DY, Kim SH, Sung GH, Choi HK. Mariannaea samuelsii isolated from a bark beetle-infested elm tree in Korea. Mycobiol 2012;40:94-9.  

84 84. Tang L, Hyun MW, Yun YH, Suh DY, Kim SH, Sung GH. New record of Mariannaea elegans var. elegans in Korea. Mycobiol 2012;40:14-9.  

85 85. Ishiuchi K, Hirose D, Kondo T, Watanabe K, Terasaka K, Makino T. Mariannamides A and B, new cyclic octapeptides isolated from Mariannaea elegans NBRC102301. Bioorganic Med Chem Lett 2020;30:126946.  

86 86. Mongkolsamrit S, Khonsanit A, Thanakitpipattana D, Tasanathai K, Noisripoom W, Lamlertthon S, Himaman W, Houbraken J, Samson RA, Luangsa-ard J. Revisiting Metarhizium and the description of new species from Thailand. Stud Mycol 2020;95:171-251.  

87 87. Lee HW, Nguyen TTT, Mun HY, Lee H, Kim C, Lee HB. Confirmation of two undescribed fungal species from Dokdo of Korea based on current classification system using multiloci. Mycobiol 2015;43:392-401.  

88 88. Kim WG, Seok SJ, Weon HY, Lee KH, Lee CJ, Kim YS. Isolation and identification of entomopathogenic fungi collected from mountains and islands in Korea. Kor J Mycol 2010;38:99-104.  

89 89. FAO Media Centre. Red locust disaster in eastern Africa prevented: Biopesticides being used on a large scale. 24th June 2009, Rome. Available from .  

90 90. Scholte E-J, Ng'habi K, Kihonda J, Takken W, Paaijmans K, Abdulla S, Killeen GF, Knols BGJ. An entomopathogenic fungus for control of adult African malaria mosquitoes. Sci 2005;308:1641-2.  

91 91. Lorenz SC, Humbert P, Patel AV. Chitin increases drying survival of encapsulated Metarhizum pemphigi blastospores for Ixodes ricinus control. Ticks Tick Borne Dis 2020;11:101537.  

92 92. Crous PW, Wingfield MJ, Guarro J, Cheewangkoon R, Bank MVD, Swart WJ, Stchigel AM, Cano-Lira JF, Roux J, Madrid H, et al. Fungal planet description sheets: 154-213. Persoonia 2013;31:188-296.  

93 93. Li XL, Ojaghian MR, Zhang JZ, Zhu SJ. A new species of Scopulariopsis and its synergistic effect on pathogenicity of Verticillium dahliae on cotton plants. Microbiol Res 2017;201:12-20. 

94 94. Jagielski T, Sandoval-Denis M, Yu J, Yao L, Bakula Z, Kalita J, Skóra M, Krzyściak P, de Hoog GS, Guarro J, et al. Molecular taxonomy of Scopulariopsis-like fungi with description of new clinical and environmental species. Fungal Biol 2016;120:586-602.  

95 95. Oh BJ, Chae MJ, Cho D, Kee SJ, Shin MG, Shin JH, Suh SP, Ryang DW. Infection with Scopulariopsis brevicaulis after cosmetic surgery of the face. Korean J Lab Med 2006;26:32-5.  

96 96. Gutarowska B. Moulds in biodeterioration of technical materials. Folia Biologica et Oecologica 2014;10:27-39.  

97 97. Lavin P, de Saravia SG, Guiamet P. Scopulariopsis sp. and Fusarium sp. in the documentary heritage: evaluation of their biodeterioration ability and antifungal effect of two essential oils. Microbiol Ecol 2016;71:628-33.  

98 98. Cheng CN, Focht DD. Production of arsine and methylarsines in soil and in culture. Appl Env Microbiol 1979;38:494-8.  

99 99. Boriová K, Čerňanský S, Matúš P, Bujdoš M, Simonovičová A. Bioaccumulation and biovolatilization of various elements using filamentous fungus Scopulariopsis brevicaulis. Lett Appl Microbiol 2014;59:217-23.  

100 100. Tibpromma S, Hyde KD, McKenzie EH, Bhat DJ, Phillips AJL, Wanasinghe DN, Samarakoon MC, Jayawardena RS, Dissanayake AJ, Tennakoon DS, et al. Fungal diversity notes 840-928: Micro-fungi associated with Pandanaceae. Fungal Divers 2018;93:1-160.  

101 101. Babu AG, Kim SW, Yadav DR, Adhikari M, Kim C, Lee HB, Lee YS. A new record of Volutella ciliata isolated from crop field soil in Korea. Mycobiology 2015;43:71-4.  

102 102. Nur EAA, Kobayashi K, Amagai A, Ohshiro T, Tomoda H. New terpendole congeners, inhibitors of sterol-O-acyltransferase, produced by Volutella citrinella BF-0440. Molecules 2020;25:3079.