Skip to main content
Studies in Mycology logoLink to Studies in Mycology
. 2024 Sep 18;109:273–321. doi: 10.3114/sim.2024.109.04

New Mucorales from opposite ends of the world

TTT Nguyen 1,#, ALCM de A Santiago 2,#, JE Hallsworth 3,#, TRL Cordeiro 2, K Voigt 4, PM Kirk 5, PW Crous 6, MAM Júnior 7, C Elsztein 7, HB Lee 1,*
PMCID: PMC11663423  PMID: 39717656

Abstract

The Mucorales is a group of ancient fungi with global distribution. In the current study we accessed mucoralean fungi isolated from two countries on opposite sides of the Earth and in different hemispheres: South Korea and Brazil. Mucorales isolates were obtained from freshwater, soil, invertebrates, and fruit seeds and identified using phenotypic techniques combined with the DNA sequence data. These analyses revealed 15 new species including one that we affiliated to a newly proposed genus, Neofennellomyces. Names proposed for these 15 new species are Absidia cheongyangensis, A. fluvii, A. kunryangriensis, A. paracylindrospora, A. tarda, A. variiprojecta, A. variispora, Backusella varians, Mucor albicolonia, M. aurantiacus, M. cryophilus, M. glutinatus, M. paraorantomantidis, M. timomeni, and Neofennellomyces jeongsukae. Of these new species, 12 were isolated from South Korea: A. cheongyangensis, A. fluvii, A. kunryangriensis, A. paracylindrospora, B. varians, M. albicolonia, M. aurantiacus, M. cryophilus, M. glutinatus, M. paraorantomantidis, M. timomeni, and N. jeongsukae, and three from Brazil: A. tarda, A. variiprojecta, and A. variispora. Niche specificity of these fungi is discussed including newly recorded invertebrate hosts and a new geographic distribution for species of Backusella, Circinella, Cunninghamella, and Mucor. Given these findings, we provide an inventory of Mucorales.

Taxonomic novelties: New genus: Neofennellomyces Hyang B. Lee & T.T.T. Nguyen. New species: Absidia cheongyangensis Hyang B. Lee & T.T.T. Nguyen, Absidia fluvii Hyang B. Lee, A.L. Santiago, P.M. Kirk, K. Voigt & T.T.T. Nguyen, Absidia kunryangriensis Hyang B. Lee & T.T.T. Nguyen, Absidia paracylindrospora Hyang B. Lee & T.T.T. Nguyen, Absidia tarda T.R.L. Cordeiro, Hyang B. Lee & A.L. Santiago, Absidia variiprojecta T.R.L. Cordeiro & A.L. Santiago, Absidia variispora T.R.L. Cordeiro & A.L. Santiago, Backusella varians Hyang B. Lee & T.T.T. Nguyen, Mucor aurantiacus Hyang B. Lee & T.T.T. Nguyen, Mucor cryophilus Hyang B. Lee & T.T.T. Nguyen, Mucor albicolonia Hyang B. Lee & T.T.T. Nguyen, Mucor glutinatus Hyang B. Lee & T.T.T. Nguyen, Mucor paraorantomantidis Hyang B. Lee & T.T.T. Nguyen, Mucor timomeni Hyang B. Lee & T.T.T. Nguyen, Neofennellomyces jeongsukae Hyang B. Lee & T.T.T. Nguyen.

Citation: Nguyen TTT, de A. Santiago ALCM, Hallsworth JE, Cordeiro TRL, Voigt K, Kirk PM, Crous PW, Júnior MAM, Elsztein C, Lee HB (2024). New Mucorales from opposite ends of the world. Studies in Mycology 109: 273–321. doi: 10.3114/sim.2024.109.04

Keywords: Biogeography, habitat and ecology, Mucorales taxonomy, Mucoromycota, Neofennellomyces, new taxa, niche specificity

INTRODUCTION

About 155 000 fungal species have been described to date although estimates of fungal diversity suggest that the actual number is an order-of-magnitude higher (Hawksworth & Lücking 2017, Niskanen et al. 2023). Of these, more than 300 species are in the ancient lineage Mucorales. This order is the largest in the Mucoromycota, with 16 families and 55 genera (Walther et al. 2019, Wijayawardene et al. 2022, Zhao et al. 2023).

Members of Mucorales are mostly saprobes from soil (Benny et al. 2016), although several species are parasites of plants, fungi and animals or endophytes (Benny et al. 2008, Walther et al. 2013, Rashmi et al. 2019, Santiago et al. 2023). Mucorales contains several species that have been reported as agents of mucormycosis in humans, especially in immunocompromised patients (Ribes et al. 2000, Walther et al. 2019). For example, Lichtheimia spp. cause rhino-orbital-cerebral, cutaneous, pulmonary, and disseminated infections (He et al. 2022, Mamali et al. 2022). Rhizopus is an opportunistic agent of human and animal disease and the main cause of rhino-orbital-cerebral infections. They also cause pulmonary or cutaneous mucormycoses (Gomes et al. 2011). Recently, Cunninghamella arunalokei has been reported to cause disease in an immunocompetent individual from India (Hallur et al. 2021). Besides their negative impact on human activities, several species in Mucorales are also used in the production of fermented foods such as tempeh and ragi (Ogawa et al. 2004, Nout & Aidoo 2010, Dolatabadi et al. 2016), or in biotechnology to produce lactic, fumaric, malic, and other organic acids, ethanol, carotenoids, and some hydrolytic enzymes (Lampila et al. 1985, Abe et al. 2003, Luo et al. 2020). Several species also proved to be effective biocontrol agents against plant pathogens. Mucor hiemalis was able to colonize inflorescence and reduce colonization of Thielaviopsis paradoxa – a pathogen that causes inflorescence brown rot (Ziedan et al. 2013). Mucor moelleri shows antagonistic activities against Athelia rolfsii and Colletotrichum gloeosporioides (Nartey et al. 2022). Chlorflavonin obtained from the endophytic fungus M. irregularis exhibited an antibacterial activity against Mycobacterium tuberculosis (Deshmukh et al. 2022).

Mucorales fungi form branched hyphae, which are coenocytic when young (aseptate and with multiple nuclei), sometimes becoming septate with age. Asexual reproduction occurs by spores formed in the sporangia, sporangiola, merosporangia, or chlamydospores, and less commonly by arthrospores and yeast-like cells (Benny et al. 2014, 2016). Sexual reproduction is by zygospores formed after the fusion of two gametangia. The shape and size of these structures, sometimes along with the maximum growth temperature and mating experiments, have been traditionally used by taxonomists for the identification and classification of Mucorales species (Schipper 1984, Zheng & Chen 2001, Zheng et al. 2007, 2017). Although many Mucoralean species can be delimited only by morphology, there are relatively few morphological features that are taxonomically relevant in these fungi (Hoffmann et al. 2013, Walther et al. 2013, Wagner et al. 2020). In addition, intraspecific morphological variability is high within this order and size and shape of some structures vary according to the experimental or environmental conditions, such as temperature, water activity, colony age, and culture medium or substrate (Schipper 1978). Since polyphyly was revealed within Mucoralean fungi (O’Donnell et al. 2001, Voigt & Wöstemeyer 2001), several groups within this order have been revised using molecular phylogenetic analyses, e.g. Lentamyces (Hoffmann & Voigt 2009), Lichtheimia (Alastruey-Izquierdo et al. 2010), and Mucor (Wagner et al. 2020).

Studies on the diversity of Mucorales fungi have been carried out in several countries, but knowledge about the ecology and distribution of these fungi is still scarce. Even though these fungi are considered cosmopolitan, this cannot be confirmed for many species in this group (Voigt et al. 2021). Because a few studies have been carried out specifically focusing on knowledge about the richness of these mucoraceous fungi in soil and other substrates (Boedijn 1959, Mirza et al. 1979, Urquhart et al. 2021) much of the knowledge about these fungi is restricted to fragmented reports in some countries and information provided by culture collections and clinical studies (Walther et al. 2019). In Brazil specific surveys on the occurrence of Mucorales are spatially and temporally fragmented with reports concentrated in the Pernambuco and São Paulo states, mostly from soil and some from herbivore dung samples (Upadhyay, 1967, 1699, 1970, Trufem 1981a, b, c, Trufem & Viriato 1985, Lima et al. 2016a, 2020a, c, Santiago et al. 2011, 2013c, Melo et al. 2020). No Mucorales diversity survey in South Korea has been published to date. Most researchers focus on the isolation of new species and new records from different types of substrates. Most records were from Cheongyang, which is located in the Chungnam province, South Korea (Nguyen et al. 2021, 2023, Nguyen & Lee 2022).

Since 2015, a high number of new species of Mucorales have been described, mainly those belonging to the Absidia, Backusella and Mucor (Hurdeal et al. 2021, 2022, Nguyen et al. 2021, Urquhart et al. 2021, Zong et al. 2021, Freitas et al. 2022, Lima et al. 2022, Nguyen & Lee 2022, Zhao et al. 2022, Cordeiro et al. 2023), with a smaller number of Cunninghamella, Gongronella and Rhizopus species (Guo et al. 2015, Li et al. 2016, Freitas et al. 2020, Hallur et al. 2021, Suwannarach et al. 2021). However, these reports were concentrated in a few research laboratories located in Australia, Brazil, China, South Korea and Thailand, and therefore the species richness of these fungi in other countries has not been recently accessed.

Data on the distribution of species of this order are fragmented and the Mucorales are little-studied by taxonomists (Voigt et al. 2021). Furthermore, relatively little is known about their abundance, biogeography, and niche specificity. The overarching goal of the present study was to access Mucorales fungi from two countries, South Korea and Brazil, that are on opposite sides of the planet. Furthermore, South Korea is in the Northern Hemisphere and Brazil in the Southern Hemisphere. The specific aims were to: isolate Mucorales fungi from freshwater, soil, invertebrates, and fruit seeds; identify isolates from different niches using a polyphasic approach combining morphological, physiological and molecular data; describe and illustrate new species of this order from South Korea and Brazil; access the similarity between the South Korean and Brazilian Mucorales communities; and provide a new inventory of species in the Mucorales.

MATERIALS AND METHODS

Sampling and isolations

Freshwater, soil, invertebrate, and cherry seed samples were collected from South Korea and Brazil as shown in Table 1. Fungal isolations from freshwater, soil, and invertebrates were carried out as previously detailed (Nguyen et al. 2017, 2021, Cordeiro et al. 2020). For soil and water samples, 1 g of soil and 1 mL of freshwater were mixed with 9 mL of sterile distilled H2O. Then, serial dilutions of the mixtures were performed (from 10−1 to 10−4 for soil samples and 10−1 for freshwater samples). For each dilution, a 100 μL aliquot was dispensed onto potato dextrose agar (PDA; Becton, Dickinson and Co., Sparks, MD) and malt extract agar (MEA; Becton, Dickinson and Co.) amended with neomycin (50 mg/L). For fruit seeds, hyphal tips or spore were transferred to potato dextrose agar (PDA) media using the tips of heat-stretched capillary tubes. To isolate the fungal strains from invertebrate samples, cadavers were broken up into small pieces and placed on PDA, and MEA media. The plates were incubated at 25 °C in the dark for 7 d. Type specimens were deposited at Chonnam National University (CNUFC) Fungarium, Gwangju, South Korea, or at the URM Culture Collection of the Universidade Federal de Pernambuco, Brazil as inactive dried cultures. Ex-type living cultures were deposited at the Environmental Microbiology Laboratory Fungarium, Chonnam National University (CNUFC), Gwangju, South Korea, or at the URM Culture Collection at the Universidade Federal de Pernambuco, Recife, Brazil.

Table 1.

Fungal strains isolated during this study.

Species name Strain no. Habitat pH Sampling date Location collected
Absidia cheongyangensis CNUFC CY2203 Leaf litter in rainwater 6.33 25 Jul. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia cheongyangensis CNUFC CY2401 Soil 7.05 18 Mar. 2024 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia fluvii CNUFC CY2240 Rainwater in a rubber bucket 6.31 20 Jun. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia fluvii CNUFC CY2241 Rainwater in a rubber bucket 6.31 20 Jun. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia fluvii CNUFC CY2315 Soil 6.33 5 Dec. 2023 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia kunryangriensis CNUFC CY2230 Rainwater 6.25 13 Jun. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia paracylindrospora CNUFC L2207 Soil under Rhododendrons 5.64 28 Oct. 2022 Gohado island, Mokpo Province, South Korea
Absidia pararepens CNUFC CY2245 Rainwater in a rubber bucket 6.12 16 Aug. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia pararepens CNUFC CY0158 Scolopendra sp. 18 Jun. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Absidia tarda URM 8412 Soil from an upland forest 5.85 18 Feb. 2019 Saloá, Pernambuco state, Brazil
Absidia variiprojecta URM 8620 Soil from an upland forest 5.85 18 Feb. 2019 Saloá, Pernambuco state, Brazil
Absidia variispora URM 8720 Soil from an upland forest 5.75 11 May 2019 Saloá, Pernambuco state, Brazil
Backusella obliqua CNUFC S805 Apis sp. 25 Sep. 2021 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Backusella varians CNUFC CY2201 Decaying leaves in soil 6.5 4 Jul. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Backusella varians CNUFC CY2202 Decaying leaves in soil 6.5 4 Jul. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Backusella varians CNUFC CY2404 Scolopendra sp. 1 Apr. 2024 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Backusella varians CNUFC CY2408 Scolopendra sp. 1 Apr. 2024 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Circinella umbellata CNUFC CY2220 Tiphia sp. 20 Jun. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Cunninghamella intermedia CNUFC IL021 Pholcus phalangioides 3 Oct. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor albicolonia CNUFC CY2311 Soil 6.83 5 Dec. 2023 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor albicolonia CNUFC CY2027 Nephila sp. 18 May 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor albicolonia CNUFC CY2028 Nephila sp. 18 May 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor aurantiacus CNUFC CY030 Protaetia orientalis 13 Jun. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor aurantiacus CNUFC CY031 Protaetia orientalis 13 Jun. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor cryophilus CNUFC CHS1 Cherry seeds 30 Mar. 2021 Chonnam National University, Jeonnam Province, South Korea
Mucor cryophilus CNUFC CHS2 Cherry seeds 30 Mar. 2021 Chonnam National University, Jeonnam Province, South Korea
Mucor glutinatus CNUFC CY2012 Soil on the skin surface of toad (Bufo gargarizans) 13 Jun. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor glutinatus CNUFC CY2016 Soil on the skin surface of toad (Bufo gargarizans) 13 Jun. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor paraorantomantidis CNUFC CY205 Isyndus obscurus 5 Aug. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor paraorantomantidis CNUFC CY206 Isyndus obscurus 5 Aug. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor saturninus CNUFC IO1 Isyndus obscurus 3 Oct. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Mucor timomeni CNUFC CY2118 Soil 7.1 8 Feb. 2021 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea.
Mucor timomeni CNUFC CY701 Timomenus komarovi 18 Apr. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea.
Neofennellomyces jeongsukae CNUFC CY2001 Euborellia sp. 20 Oct. 2020 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea
Neofennellomyces jeongsukae CNUFC CY2204 Porcellio scaber 4 Jul. 2022 Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, South Korea

Morphological and physiological studies

Pure cultures were grown in triplicate on PDA, MEA, and synthetic mucor agar (SMA) as previously detailed (Benny 2008, Nguyen et al. 2021). Incubation was carried out at 15, 20, 25, 30, 35, 37, 40, and 45 °C in the dark for 7 d and colony diameters were measured every 24 h. The Tmax was determined by growing 27 strains on MEA, PDA and SMA at temperatures one degree higher than the last temperature with growth. For morphological identification, mycelial fragments were removed from the cultures, placed on microscope slides in the presence of lactic acid (60 % v/v), and observed under a light microscope using a differential interference contrast microscope (Olympus BX53, Tokyo, Japan). For each sample, at least 50 sporangia, columellae, and sporangiospores were measured.

DNA extraction, PCR, cloning, and sequencing

Total genomic DNA was extracted from fresh mycelia grown on PDA at 25 °C for 3–5 d using a modified cetyltrimethylammonium bromide (CTAB) protocol, as described by Oliveira et al. (2016), along with the SolgTM Genomic DNA Preparation Kit (Solgent Co. Ltd., Daejeon, South Korea) according to the manufacturer’s protocol (Nguyen et al. 2021). The ITS region of rDNA was amplified using forward primers ITS1 (White et al. 1990) or V9G (de Hoog & van den Ende 1998) and reverse primers ITS4 (White et al. 1990) or LS266 (Masclaux et al. 1995); the LSU ribosomal DNA was amplified using forward primers LR0R (Bunyard et al. 1994) or LR1 (van Tuinen et al. 1998) and reverse primers LR5 (Vilgalys & Hester 1990) or LSU2 (Santiago et al. 2014); the mini-chromosome maintenance complex component 7 (mcm7) was amplified using forward primer Mcm7-709f and reverse primer Mcm7-1348rev (Schmitt et al. 2009); the gene encoding the largest subunit of RNA polymerase II (rpb1) was amplified using forward primers RPB1-f1 (Wagner et al. 2020) or RPB1-F1843 and reverse primer RPB1-R3096 (Houbraken & Samson 2011); a partial arginosuccinate lyase gene fragment (argA) was amplified using forward primer AP52 and reverse primer AP53 (Urquhart et al. 2021); the translation elongation factor 1-α (EF-1α) was amplified using forward primer MEF-1 and reverse primer MEF-4 (O’Donnell et al. 2001); actin (act-1) was amplified using forward primer Act-1 and reverse primer Act-4r (Voigt & Wöstemeyer 2001). The PCR amplifications were performed as described by Alastruey-Izquierdo et al. (2010), Santiago et al. (2014), Wagner et al. (2020), and Nguyen et al. (2021). The PCR amplicons were purified and sequenced directly or cloned with a pGEM-T Easy Vector (Promega, Madison, WI, USA) for ITS (Absidia, Backsuella, and Mucor) and mcm7 sequences (Neofennellomyces).

Phylogenetic analyses

The sequences of all strains obtained in this study were checked in BioEdit v. 7.0.5.3 (Hall et al. 1999) for quality. The SeqMan v. 7.0 program was used to assemble and edit the raw sequences. The sequences were used to search for similar taxa via the Basic Local Alignment Search Tool (BLAST) in the National Center for Biotechnology (NCBI), and the most closely related taxa were obtained for the phylogenetic analyses. Sequences of each locus were aligned using MAFFT v. 7 with default parameters (http://mafft.cbrc.jp/alignment/server) (Katoh et al. 2017) and then confirmed manually in MEGA v. 7 (Kumar et al. 2016). The alignments of the ITS region of Absidia, Backusella and Cunninghamella was trimmed with trimAl under default parameters (Capella-Gutiérrez et al. 2009). The most suitable substitution model was determined using the jModelTest v. 2.1.10 software (Guindon & Gascuel 2003, Darriba et al. 2012). The data were converted from a FASTA format to NEXUS formats using the online tool Alignment Transformation Environment (https://sing.ei.uvigo.es/ALTER/) (Glez-Peña et al. 2010). Phylogenetic reconstructions by maximum likelihood (ML) were carried out using Randomized Axelerated Maximum Likelihood for high-performance computing (RAxML-HPC2) on XSEDE on the online CIPRES Portal (https://www.phylo.org/portal2) with a default general time reversible (GTR) substitution matrix and 1 000 rapid bootstraps. Bayesian inference (BI) analysis was performed using MrBayes v. 3.2.2 (Ronquist et al. 2012). Four Markov chain Monte Carlo (MCMC) chains were run from a random starting tree for five million generations, and trees were sampled every 100th generation. The first 25 % of the trees were removed as burn-in, and the remaining trees were used to calculate posterior probabilities. The trees were visualized using FigTree v. 1.3.1 (Rambaut 2009). Support values were provided at the branches [ML bootstrap support (BS) and BI posterior probability (PP)]. The GenBank accession numbers of the sequences of the strains used in this study are available in Supplementary Table S1.

Data collection

Information on Mucorales species recorded in South Korea and Brazil was collected from articles and book published as of June 2024. Species names are based on the most recent lists of accepted species in Walther et al. (2013), Wagner at al. (2019), and subsequently published names were verified in the Index Fungorum and MycoBank databases.

Similarity between the South Korean and Brazilian Mucorales communities

The similarity of the Mucorales communities between South Korea and Brazil was accessed based on Sørensen (1978).

RESULTS

Phylogenetic analyses

In this study, six combined phylogenetic analyses were performed to determine the placement of our new taxa and new records within the Cunninghamellaceae, Backusellaceae, Lichtheimiaceae, and Mucoraceae as follows:

Cunninghamellaceae

Absidia

The phylogenetic tree of Absidia was determined using a combined sequence dataset of three loci (ITS, LSU, and act-1). The aligned dataset contained 2 247 characters, including gaps (ITS: 1–605, LSU: 606–1 318 and act-1: 1 319–2 247). The best RAxML tree with a final likelihood value of -26575.475712 is presented (Fig. 1). The matrix had 1 256 distinct alignment patterns, with 35.66 % undetermined characters or gaps. The estimated base frequencies were as follows: A = 0.256935, C = 0.209782, G = 0.234648, T = 0.298635, with substitution rates AC = 1.065432, AG = 2.934662, AT = 1.602805, CG = 0.700770, CT = 4.903722, GT = 1.000000; gamma distribution shape parameter α = 0.327694; Tree-Length: 6.613812. For BI analysis, TVM+I+G, TrN+I+G, and HKY+I+G were selected as the best-fit model by BIC in jModelTest v. 2.1.10 for ITS, LSU, and act-1, respectively. Phylogenetic analyses of the combined ITS, LSU, and act-1 sequence dataset (Fig. 1) showed that A. fluvii and A. paracylindrospora clustered close to but separated from A. globospora and A. cylindrospora var. cylindrospora, respectively. Absidia variiprojecta was placed in a subclade sister to one formed by A. kunryangriensis, A. koreana, A. zonata, A. aguabelensis and A. longissima. Absidia variispora was closely related to A. tarda, A. cheongyangensis, and A. variicolumellata. The strains CNUFC CY2245 and CNUFC CY0158 clustered with the known species A. pararepens. Based on the evidence of phylogeny and morphological characteristics, seven new species are poposed, along with a new record for A. pararepens in South Korea.

Fig. 1.

Fig. 1

Phylogenetic relationship of Absidia species based on the combined ITS, LSU, and act-1 sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages (left) and Bayesian posterior probabilities (right). Bootstrap values > 70 % and Bayesian posterior probabilities > 0.95 are shown. Cunninghamella phaeospora CBS 692.68 and C. vesiculosa CBS 989.96 were used as outgroups. Ex-type, ex-isotype and ex-neotype strains are marked with T, IT, and NT, respectively. Newly generated sequences are in bold blue.

Cunninghamella

The phylogenetic tree of Cunninghamella was determined using a combined sequence dataset of three loci (ITS, LSU and EF-). The aligned dataset contained 1 905 characters, including gaps (ITS: 1–620, LSU: 621–1 338 and EF-1α: 1 339–1 905). The best RAxML tree with a final likelihood value of -12539.260820 is presented (Fig. 2). The matrix had 799 distinct alignment patterns, with 29.16 % undetermined characters or gaps. The estimated base frequencies were as follows: A = 0.283470, C = 0.184264, G = 0.216597, T = 0.315669, with substitution rates AC = 1.011874, AG = 2.876363, AT = 2.139887, CG = 0.496412, CT = 4.975235, GT = 1.000000; gamma distribution shape parameter α = 0.287404; Tree-Length = 2.204513. For BI analysis, TVM+G, TIM3+G, and TIM2ef+G were selected as the best-fit model by BIC in jModelTest v. 2.1.10 for ITS, LSU, and EF-1α, respectively. The multigene phylogenetic analysis showed that our strain CNUFC IL021 clustered with other strains of C. intermedia (Fig. 2). Therefore, we identify our strain as C. intermedia.

Fig. 2.

Fig. 2

Phylogenetic relationship of Cunninghamella species based on the combined ITS, LSU and EF-1α sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages (left) and Bayesian posterior probabilities (right). Bootstrap values > 70 % and Bayesian posterior probabilities > 0.95 are shown. Absidia californica CBS 126.68 was used as outgroup. Ex-type, ex-isotype, and ex-neotype strains are marked with T, IT, and NT, respectively. Newly generated sequences are in bold blue.

Backusellaceae

Backusella

The phylogenetic tree of Backusella was determined by analysis of concatenated sequence datasets of four loci (ITS, LSU, rpb1 and argA). The aligned dataset contained 2 926 characters, including gaps (ITS: 1–521, LSU: 522–1 155, rpb1: 1 156–2 184, and argA: 2 185–2 926). The best RAxML tree with a final likelihood value of -19815.414823 is presented (Fig. 3). The matrix had 1 168 distinct alignment patterns, with 28.04 % undetermined characters or gaps. The estimated base frequencies were as follows: A = 0.261981, C = 0.203740, G = 0.244590, T = 0.289689, with substitution rates AC = 1.181711, AG = 3.855317, AT = 1.349560, CG = 0.750793, CT = 6.249831, GT = 1.000000; gamma distribution shape parameter α = 0.252728; Tree-Length = 2.432248. For BI analysis, TIM3+I+G, TPM1uf+I+G, TIM3+G, and TrN+G were selected as the best-fit model by BIC in jModelTest v. 2.1.10 for ITS, LSU, rpb1 and argA, respectively. The multigene phylogenetic analysis showed that our strain CNUFC S805 clustered with the type species B. obliqua (Fig. 3). Therefore, we identify our strain as B. obliqua and are new record in South Korea. Other strains CNUFC CY2201, CNUFC CY2202, CNUFC CY2404, CNUFC CY2408 grouped in a single well-supported clade. As a result, strains CNUFC CY2201, CNUFC CY2202, CNUFC CY2404, and CNUFC CY2408 are recognized as a novel species.

Fig. 3.

Fig. 3

Phylogenetic relationship of Backusella species based on the combined ITS, LSU, rpb1, and argA sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages (left) and Bayesian posterior probabilities (right). Bootstrap values > 70 % and Bayesian posterior probabilities > 0.95 are shown. Utharomyces epallocaulus CBS 329.73 was used as outgroup. Ex-type, ex-epitype, ex-lectotype, and ex-neotype strains are marked with T, ET, LT, and NT, respectively. Newly generated sequences are in bold blue.

Lichtheimiaceae

Circinella and Neofennellomyces

The phylogenetic tree of Circinella and Neofennellomyces were determined by analysis of concatenated sequence datasets of three loci (ITS, LSU, and mcm7). The aligned dataset contained 2 243 characters, including gaps (ITS: 1–807, LSU: 808–1 490, and mcm7: 1 491–2 243). The best RAxML tree with a final likelihood value of -9848.906747 is presented (Fig. 4). The matrix had 845 distinct alignment patterns, with 40.86 % of undetermined characters or gaps. The estimated base frequencies were as follows: A = 0.292076, C = 0.170684, G = 0.220701, T = 0.316539, with substitution rates AC = 0.976467, AG = 3.018463, AT = 1.213000, CG = 0.635986, CT = 5.250357, GT = 1.000000; gamma distribution shape parameter α = 0.364538; Tree-Length = 1.509782. For BI analysis, HKY+G, TrN+I+G, and TIM3+G were selected as the best-fit model by BIC in jModelTest v. 2.1.10 for ITS, LSU and mcm7, respectively. The multigene phylogenetic analysis showed that our strain CNUFC CY2220 clustered with other strain of C. umbellata (Fig. 4). The combined analysis showed, with high statistical support, that the Neofennellomyces jeongsukae strains CNUFC CY2001, CNUFC CY2204 formed a well-supported clade. Neofennellomyces jeongsukae was embedded among the clade of Circinella spp. and Phascolomyces articulosus (Fig. 4). In the ITS, and mcm7 phylogenetic tree, this species has a basal position to a clade containing species of Circinella (Supplementary Figs. S2, S4). However, analysis of the LSU shows that this species belongs to a clade containing species of Circinella (Supplementary Fig. S3). Based on the present phylogenetic position inferred by ITS, LSU, and mcm7 analyses, and also with careful morphological comparison with related species, we are therefore confident to describe Neofennellomyces as a new genus with type species, Neofennellomyces jeongsukae.

Fig. 4.

Fig. 4

Phylogenetic relationship of Neofennellomyces species and related taxa based on the combined ITS, LSU, and mcm7 sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages (left) and Bayesian posterior probabilities (right). Bootstrap values > 70 % and Bayesian posterior probabilities > 0.95 are shown. Fennellomyces heterothallicus CBS 290.86, F. heterothallicus CBS 292.86, F. linderi CBS 158.54, Thamnostylum nigricans CBS 690.76, T. piriforme CBS 316.66, and T. repens CBS 692.76 were used as the outgroups. Ex-type and ex-neotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.

Mucoraceae

Mucor

The phylogenetic tree of Mucor in the M. amphibiorum group and related taxa was determined by analysis of concatenated sequence datasets of three loci (ITS, LSU, and rpb1). The aligned dataset contained 2 302 characters, including gaps (ITS: 1–558, LSU: 559–1 279, and rpb1: 1 280–2 302. The best RAxML tree with a final likelihood value of -20227.882688 is presented (Fig. 5). The matrix had 1 104 distinct alignment patterns, with a 34.27 % of undetermined characters or gaps. The estimated base frequencies were as follows: A = 0.296850, C = 0.171962, G = 0.222792, T = 0.308397, with substitution rates AC = 1.045037, AG = 4.479357, AT = 1.684739, CG = 0.755279, CT = 6.930357, GT = 1.000000; gamma distribution shape parameter α = 0.254180; Tree-Length = 3.869121. For BI analysis, TPM1uf+I+G, TPM3uf+I+G, and HKY+G were selected as the best-fit model by BIC in jModelTest v. 2.1.10 for ITS, LSU and rpb1, respectively. The multigene phylogenetic analysis showed that M. albicolonia, M. aurantiacus, and M. paraorantomantidis were closely related to M. gigasporus, M. azygosporus, and M. orantomantidis, respectively. Mucor timomeni was sister clade to M. chiangraiensis, and M. nederlandicus (Fig. 5).

Fig. 5.

Fig. 5

Phylogenetic relationship of Mucor species in Mucor amphibiorum group and related taxa based on the combined ITS, LSU, and rpb1 sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages (left) and Bayesian posterior probabilities (right). Bootstrap values > 70 % and Bayesian posterior probabilities > 0.95 are shown. Backusella oblongispora CBS 569.70 was used as outgroup. Ex-type, ex-epitype, ex-holotype, ex-isotype, and ex-neotype strains are marked with T, ET, HT, IT, and NT, respectively. Newly generated sequences are in bold blue.

The phylogenetic tree of Mucor in the M. mucedo group, M. flavus group, M. hiemalis group and related taxa is based on a concatenated dataset of ITS, LSU, and rpb1. The aligned dataset contained 2 361 characters, including gaps (ITS: 1–627, LSU: 628–1 335, and rpb1: 1 336–2 361. The best RAxML tree with a final likelihood value of -15350.792940 is presented (Fig. 6). The matrix had 881 distinct alignment patterns, with a 28.15 % of undetermined characters or gaps. The estimated base frequencies were as follows: A = 0.291918, C = 0.170713, G = 0.214965, T = 0.322403, with substitution rates AC = 0.922284, AG = 3.863225, AT = 1.603871, CG = 0.618744, CT = 5.674739, GT = 1.000000; gamma distribution shape parameter α = 0.186268; Tree-Length = 2.227159. For BI analysis, HKY+I+G, TIM3+I+G, and TPM2uf+I+G were selected as the best-fit model by BIC in jModelTest v. 2.1.10 for ITS, LSU and rpb1, respectively. The multigene phylogenetic analysis (Fig. 6) indicated that M. cryophilus formed an independent clade. Mucor glutinatus was closely related to M. silvaticus. Strain CNUFC IO1 clustered with other strains of M. saturninus (CBS 974. 68, CBS 599.78, CBS 598.78). Therefore, we identify our strain as a new record for M. saturninus from South Korea.

Fig. 6.

Fig. 6

Phylogenetic relationship of Mucor species in M. mucedo group, M. flavus group, M. hiemalis group and related taxa based on a concatenated dataset of ITS, LSU, and rpb1 sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages (left) and Bayesian posterior probabilities (right). Bootstrap values > 70 % and Bayesian posterior probabilities > 0.95 are shown. Mucor amphibiorum CBS 763.74 was used as outgroup. Ex-type, ex-epitype, ex-holotype, and ex-neotype strains are marked with T, ET, HT, and NT, respectively. Newly generated sequences are in bold blue.

Taxonomy

Mucorales Dumort.

Cunninghamellaceae Naumov ex R.K. Benj.

Absidia Tiegh., Ann. Sci. Nat., Bot. 4: 350. 1878.

Absidia fluvii Hyang B. Lee, A.L. Santiago, P.M. Kirk, K. Voigt & T.T.T. Nguyen, sp. nov. Index Fungorum 900607. Fig. 7.

Fig. 7.

Fig. 7

Absidia fluvii (CNUFC CY2240). A, B. Colony on SMA (A obverse view, B reverse view). C–F. Branched sporangiophore with columellae (observed under a dissecting microscope). G, H. Mature sporangia. I. Two sporangiophores with columellae showing one projection. J. Columella without projection. K. Sporangiospores. Scale bars: G–J = 20 μm; K = 10 μm.

Etymology: Refers to the freshwater from which the species was first isolated.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from rainwater, 20 Jun. 2022, H.B. Lee & J.S. Kim (holotype CNUFC HT2280, ex-type culture CNUFC CY2240, GenBank numbers: ITS = PP844891, LSU = PP852703, act-1 = PP893201).

Description: Colony rapidly growing, white to greyish olive, reaching 71 mm diam after 4 d on SMA at 25 °C; reverse light olive grey, wavy zonate. Rhizoids branching and finger-like. Sporangiophores 85–1 130 × 6–13.5(–15) μm, arising from stolons, hyaline to dark brown towards the sporangium, singly or two to six in a whorl, unbranched and sympodially branched up to three, always with a septum 15–23 μm below sporangium; sporangiophore showing swellings formed below the sporangium, 17–26 μm diam. Sporangia subglobose and pyriform, hyaline to light brown, smooth-walled, 27.5–50.5(–55.5) × 26–47.5(–53.5) μm. Columellae subglobose to oval, hemispherical, hyaline to brown, 17–36(–40) × 11–29(–31) μm, smooth-walled; collar present or indistinct, hyaline to brown. Projection up to 4 μm in length, needle or nipple-like, not all columellae with projection. Apophysis dark brown, 15–21.5 μm diam. Sporangiospores hyaline, mostly globose, some subglobose, 2.5–3.5 μm diam, smooth-walled. Chlamydospores and zygospores absent. Sporangia on PDA and MEA (25–45 × 24–41 μm) are slightly smaller than those on SMA. Branches are usually less on PDA and MEA.

Culture characteristics: On PDA, the colonies attained 69 mm diam after 4 d at 25 °C. On MEA, the colonies attained 70 mm diam after 4 d at 25 °C. Maximum growth temperature was 32 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, from rainwater sample, 20 Jun. 2022, H.B. Lee & J.S. Kim, culture CNUFC CY2241; Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from soil, 5 Dec. 2023, H.B. Lee, culture CNUFC CY2315.

Notes: Absidia fluvii is phylogenetically related to A. globospora. Absidia fluvii differs from A. globospora by having smaller sporangiospores and longer sporangiophores. Absidia fluvii produces sporangiophores singly or in whorls (2–6), whereas A. globospora produces sporangiophores up to five in whorls (Zong et al. 2021).

Absidia cheongyangensis Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900608. Fig. 8.

Fig. 8.

Fig. 8

Absidia cheongyangensis (CNUFC CY2203). A, D. Colony on PDA. B, E. Colony on MEA. C, F. Colony on SMA (A–C obverse view, D–F reverse view). G. Unbranched sporangiophore with sporangium. H. Mature sporangium. I. Columella. J. Sporangiospores. Scale bars: G–I = 20 μm; J = 10 μm.

Etymology: Refers to the isolation location, Cheongyang (South Korea) from where the species was first isolated.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from a leaf litter in rainwater, 25 Jul. 2022, H.B. Lee & J.S. Kim (holotype CNUFC HT2284, ex-type culture CNUFC CY2203, GenBank numbers: ITS = PP844904 (c1), LSU = PP852788, act-1 = PP893196).

Description: Colony slow growing, white to greyish brown, reaching 57 mm diam after 4 d on SMA at 25 °C; reverse grey, wavy zonate. Sporangiophores (54–)62–375 × 3–6 μm, arising from stolons, hyaline to brown, more coloured towards the sporangium, singly to four in a whorl, but mostly two–three per whorl; a septum is present 11.5–18 μm below the sporangium. Rhizoids about 6 μm diam, hyaline, unbranched and branched. Sporangia globose to pyriform, multi-spored, deliquescing and apophysate, 23–43.5 × 22–41.5 μm, hyaline to yellow-brown. Columellae spherical or hemispherical, 14.5–28 × 12.5–27 μm, smooth, often with a projection, hyaline; projections 2.5–4.5 × 1.5–2.5 μm, needle or nipple-like, or cylindrical rounded above; collar present or indistinct, hyaline. Sporangiospores oval to short cylindrical, 2.5–3.5(–4) × 2–2.5 μm, smooth-walled. Chlamydospores and zygospores absent. The size and shape of sporangia, columellae, and sporangiosphores on PDA and MEA are similar to those on SMA. Good sporulation was observed on SMA followed by MEA and PDA.

Culture characteristics: On PDA, the colonies attained 59 mm diam after 4 d at 25 °C. On MEA, the colonies attained 60 mm diam after 4 d at 25 °C. Maximum growth temperature was 33 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, from soil sample, 18 Mar. 2024, H.B. Lee & J.S. Kim, culture CNUFC CY2401.

Notes: Absidia cheongyangensis is a sister taxon to A. variicolumellata and A. tarda. Absidia cheongyangensis produces sporangiospores in whorls of up to four, whereas A. variicolumellata produces sporangiophores in whorls of up to eight. The sporangiospores of A. variicolumellata are larger than those A. cheongyangensis [3.5–7.2 × 2.5–4.5 μm vs. 2.5–3.5(–4) × 2–2.5]. Absidia tarda produces swellings along the sporangiophores, stolons and hyphae, which are not observed in A. cheongyangensis.

Absidia kunryangriensis Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900609. Fig. 9.

Fig. 9.

Fig. 9

Absidia kunryangriensis (CNUFC CY2230). A, D. Colony on PDA. B, E. Colony on MEA. C, F. Colony on SMA (A–C obverse view, D–F reverse view). G, H. Young and mature sporangium. I. Columella, collar and projection. J. Sporangiospores. Scale bars: G–I = 20 μm; J = 10 μm.

Etymology: Refers to the isolation location, Kunryang-ri (South Korea) from where the species was first isolated.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from rainwater, 13 Jun. 2022, H.B. Lee & J.S. Kim (holotype CNUFC HT2278, ex-type culture CNUFC CY2230, GenBank numbers: ITS = PP844905 (c1); PP844906 (c3), LSU = PP956882, act-1 = PP893198).

Description: Colony slow growing, floccose, white to light grey, margin irregular, reaching 59 mm diam after 4 d on SMA at 25 °C; reverse cream white. Sporangiophores (38–)45–400 × 4–7 μm, erect, arising from stolons, mostly unbranched, some branches may re-branch up to two times, hyaline to light brown, and more coloured towards the sporangium, singly to five in a whorl, always with a septum 10–19.5 μm below sporangium. Rhizoids hyaline, unbranched. Sporangia subglobose, subpyriform to pyriform, multi-spored, hyaline to light yellow, smooth-walled, 29–43.5 × 27.5–36 μm, with funnel-shaped apophysis. Columellae hemispherical in the large terminal sporangia, or subglobose to hemispherical in lateral sporangia, 10.5–23(–26) × 8.5–21(–25) μm, smooth-walled, hyaline to brown; collar present or indistinct, hyaline to bright brown. A single projection is present on the upper surface of the columellae, short spine to rounded cylinder, 1.5–4 × 1.5–2 μm. Sporangiospores hyaline, cylindrical 3–4(–4.5) × 2–2.5 μm, smooth-walled. Chlamydospores and zygospores absent. On MEA and PDA, sporangia are slightly smaller (MEA: up to 39.5 μm diam; PDA: up to 38 μm diam) than on SMA. Projection on PDA (up to 6.5 × 3 μm) are larger than those on SMA and MEA (up to 4 × 1.5 μm). Good sporulation was observed on PDA followed by SMA and MEA.

Culture characteristics: On PDA, the colonies attained 66 mm diam after 4 d at 25 °C. On MEA, the colonies attained 65 mm diam after 4 d at 25 °C. Slow growth was observed at 35 °C on MEA, PDA and SMA. Maximum growth temperature was 36 °C MEA, PDA and SMA (Supplementary Table S2).

Notes: Absidia kunryangriensis is phylogenetically related to A. koreana. Absidia kunryangriensis differs from A. koreana by having larger sporangia (29–43.5 × 27.5–36 μm vs. 19.33–23.64 × 21.06–26.35 μm) and larger columellae [10.5–23(–26) × 8.5–21(–25) μm vs. 10.9–16.96 × 11.46–18.89 μm] (Ariyawansa et al. 2015).

Absidia paracylindrospora Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900707. Fig. 10.

Fig. 10.

Fig. 10

Absidia paracylindrospora (CNUFC L2207). A, D. Colony on PDA. B, E. Colony on MEA. C, F. Colony on SMA (A–C obverse view, D–F reverse view). G. Unbranched sporangiophore with sporangium. H. Mature sporangium. I. Sporangiospores. Scale bars: G, H = 20 μm; I = 10 μm.

Etymology: Refers to the phylogenetic proximity to Absidia cylindrospora.

Typus: South Korea, Mokpo Province, Gohado Island, from soil, 28 Oct. 2022, H.B. Lee (holotype CNUFC HT2289, ex-type culture CNUFC L2207, GenBank numbers: ITS = PP844907 (c1), PP844908 (c2), LSU = PP956883, act-1 = PP893204).

Description: Colony slow growing, white to light brown, reaching 44 mm diam after 4 d on SMA at 25 °C; reverse white, wavy zonate. Sporangiophores (30–)40–210 × 4–6 μm, arising from stolons, unbranched, hyaline, singly to four in a whorl, a septum present 14–23 μm below the sporangium. Sporangia pyriform, multi-spored, 22–31 × 20–29 μm, some smaller 14–16 × 12.5–14.5 μm, hyaline to pale yellow. Columellae mostly hemispherical, some subglobose, hyaline, 9.5–17 μm diam. Collar present or indistinct, hyaline. Sporangiospores cylindrical, 3–4 × 2–2.5 μm, smooth-walled. Chlamydospores and zygospores absent.

Culture characteristics: On PDA, the colonies attained 51 mm diam after 4 d at 25 °C. On MEA, the colonies attained 48 mm diam after 4 d at 25 °C. Maximum growth temperature was 32 °C on PDA, while 31 °C on MEA and SMA (Supplementary Table S2).

Notes: In the phylogenetic tree (Fig. 1), A. paracylindrospora clustered close to but separated from A. cylindrospora var. cylindrospora. Absidia paracylindrospora produces smaller columellae and sporangiospores than A. cylindrospora var. cylindrospora (Columellae: 9.5–17 μm diam vs. 8.5–26 μm diam; Sporangiospores: 3–4 × 2–2.5 μm vs. 3.3–5.5 × 2.2–3.5 μm). Absidia cylindrospora var. cylindrospora produces zygospores, whereas A. paracylindrospora does not.

Absidia pararepens Jurjević et al., Persoonia 44: 351. 2020. MycoBank MB 834983. Fig. 11.

Fig. 11.

Fig. 11

Absidia pararepens (CNUFC CY2245). A, D. Colony on PDA. B, E. Colony on MEA. C, F. Colony on SMA (A–C obverse view, D–F reverse view). G, H. Mature sporangia and columella. I. Columella, collar and projection. J. Sporangiospores. Scale bars: G–I = 20 μm; J = 10 μm.

Description: Colony slow growing, first white, becoming greyish brown, reaching 46 mm diam after 4 d on SMA at 25 °C; reverse white to light grey. Rhizoids finger-like, short and long, unbranched. Sporangiophores erect, (22–)28–134 × 3.5–6 μm, arising from stolons, unbranched, hyaline to brown, singly to six in a whorl, a septum present 7–14.5 μm below the sporangium. Sporangia hyaline to greyish brown, oval to elliptical, or pyriform, 13–29 × 11.5–22 μm. Columellae hyaline to brown, mostly hemispherical, 7–15 × 5.5–11 μm, some subglobose, 8.5–19 × 6–15 μm, smooth-walled. Collar present. Projections 4–6.5 μm in length, hyaline, needle-shaped, bulbous. Sporangiospores globose, 2.5–3.5 μm, smooth-walled. Chlamydospores present in the aerial mycelia. Zygospores absent.

Culture characteristics: On PDA, the colonies attained 47 mm diam after 4 d at 25 °C. On MEA, the colonies attained 50 mm diam after 4 d at 25 °C. Maximum growth temperature was 32 °C on PDA, MEA and SMA (Supplementary Table S2).

Material examined: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from rainwater, 16 Aug. 2022, H.B. Lee & J.S. Kim, culture CNUFC CY2245; Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from Scolopendra sp., 18 Jun. 2020, H.B. Lee & J.S. Kim, culture CNUFC CY0158.

Notes: Our isolates were placed within the clade of ex-type strain of A. pararepens (Fig. 1). Sporangiospores of our isolate are globose, whereas sporangiospores of the ex-type strain of A. pararepens are sub-globose to globose and oval. Absidia pararepens produces larger sporangiospores than our isolates [(3.3–)3.5–5(–9) × (3.3–) 3.5–6 μm vs. 2.5–3.5 μm] (Crous et al. 2020).

Absidia tarda T.R.L. Cordeiro, Hyang B. Lee & A.L. Santiago, sp. nov. Index Fungorum 900615. Fig. 12.

Fig. 12.

Fig. 12

Absidia tarda (URM 8412). A. Colony verse (left) and reverse (right) on MEA. B, C. Once branched sporangiophore with sporangia. D, E. Unbranched sporangiophore with one swelling and sporangium (D) and columella (E). F. Unbranched sporangiophore with columella showing one projection. G. Swollen hyphae resembling giant cells. H. Rhizoid. I. Sporangiospores. Scale bars: B–H = 25 μm; I = 10 μm.

Etymology: Relating to the slow development of asexual reproductive structures.

Typus: Brazil, Saloá, from soil, 18 Feb. 2019, T.R.L. Cordeiro (holotype URM 8412h, ex-type culture URM 8412, GenBank numbers: ITS = PP844911 (c2), PP844912 (c3), LSU = PP956884, act-1 = PP893199).

Description: Colonies on MEA at 25 °C, slow growing, initially white, turning light brown, with 4.5 cm in height and 9 cm diam after 5 d of incubation; reverse cream. Rhizoids present, short or long, branched or unbranched. Sterile mycelium abundant. Stolons light brown, wall slightly encrusted. Sporangiophores erect, arising from stolons, usually unbranched, single or in whorls of up to three, up to 5 μm in width; when branched they are usually monopodially branched, rarely sympodially branched; some swellings may be observed along the sporangiophores, stolons and hypha, some resembling giant cells. One septum was observed near the apophysis. Sporangia greyish brown, globose, (12–)15–30 μm diam, apophysate, wall smooth and deliquescent. Apophysis brownish, long or short, cup-shaped, (5–)7–12(–15) × (5–)7–15 μm. Columellae hyaline, hemispherical, (7–)9.5–15(–20) μm, or subglobose to conical, (5–)10–15 × (7–)12–20 μm, smooth-walled; collar visible. One projection on columellae filiform, some bulbous, occasionally triangle-shaped, up to 4 × 2.5 μm is observed. Some columellae may not show one projection on their surface. Sporangiospores hyaline, mostly cylindrical constricted in the centre, some short cylindrical (2.5–)4.5–7.5(–9) × 2.5–5(–7) μm, smooth-walled. Chlamydospores and zygospores not observed.

Culture characteristics: On PDA, the colonies attained 6 cm diam after 5 d at 25 °C. On SMA, the colonies attained 83 mm diam after 5 d at 25 °C. Maximum growth temperature was 36 °C on MEA, PDA, and SMA (Supplementary Table S2).

Notes: In our phylogenetic tree (Fig. 1), A. tarda was placed in a subclade closer to A. cheongyangensis and A. variicolumellata. Morphologically, A. tarda can be distinguished from both species based on swellings formed along the sporangiophores, stolons and hyphae. Other distinctive feature is that A. tarda takes 7 d or more to form reproductive structures at 25 °C on both MEA and PDA.

Absidia variiprojecta T.R.L. Cordeiro & A.L. Santiago, sp. nov. Index Fungorum 900614. Fig. 13.

Fig. 13.

Fig. 13

Absidia variiprojecta (URM 8620). A. Colony verse (left) and reverse (right) on MEA. B. Sporangiophore with sporangium. C. Branched sporangiophore with columellae. D. Sporangiophore with columella. E–G. Sporangiophore with columella showing a projection. H. Rhizoid. I. Sporangiospores. Scale bars: B–H = 25 μm; I = 10 μm.

Etymology: Reference to the varied-shaped projections observed on columellae.

Typus: Brazil, Saloá, from soil, 18 Feb. 2019, T.R.L. Cordeiro (holotype URM 8620h, ex-type culture URM 8620, GenBank numbers: ITS = PP844913 (c1), PP844914 (c8), LSU = PP956885, act-1 = PP893200).

Description: Colonies on MEA at 25 °C, slow growing, initially white, turning light to dark brown, with 9 cm diam after 5 d of incubation; reverse cream. Rhizoids short or long, branched or unbranched. Stolons brown, wall slightly encrusted. Sporangiophores erect, arising from stolons, usually unbranched, single or in whorls of up to five, up to 5 μm in width; when branched they are usually monopodially branched, rarely sympodially branched. One septum is observed near the apophysis; some sterile branches may arise from the main sporangiophore. Sporangia brown, pyriform, up to 15–35 × (15–)25–35 μm, apophysate, wall smooth and deliquescent. Apophysis brownish, long or short, cup-shaped, (2.5–)5–12 × (4–)7–15(–20) μm. Columellae hyaline, subglobose, (5–)7–20, or subglobose to fig-shaped, (5–)7.5–15 × (7–)10–20 μm, smooth-walled; one projection on columellae generally needle- or triangle-shaped, some bulbous, very rarely feather-shaped, up to 1.5 × 5 μm is observed. Some columellae may not show one projection on their surface. Sporangiospores hyaline, mostly cylindrical slightly constrict in the centre, some elliptical, (2.5–)4.5–7.5(–9) × 2.5–4(–6) μm, some subglobose and globose, (2–)3.5–6(–7.5) μm diam, smooth-walled. Chlamydospores present, globose or ovoid, up to 15 × 10 μm. Zygospores not observed.

Culture characteristics: On PDA, the colonies attained 9 cm diam after 5 d at 25 °C. On SMA, the colonies attained 75 mm diam after 5 d at 25 °C. Maximum growth temperature was 32 °C on MEA and PDA, while 33 °C on SMA (Supplementary Table S2).

Notes: Absidia variiprojecta was placed in a subclade sister to one formed by A. kunryangriensis, A. koreana, A. zonata, A. aguabelensis and A. longissima (Fig. 1). However, among those species, only A. variiprojecta forms columellae that are subglobose to fig-shaped with projections triangular or rarely feather-shaped, and sporangiospores mostly cylindrical, slightly constrict in the centre, some elliptical, subglobose and globose.

Absidia variispora T.R.L. Cordeiro & A.L. Santiago, sp. nov. Index Fungorum 900616. Fig. 14.

Fig. 14.

Fig. 14

Absidia variispora (URM 8720). A. Colony obverse (left) and reverse (right) on MEA. B. Unbranched sporangiophore with one swelling and sporangium. C–E. Branched sporangiophore with sporangia and columellae. F. Sporangiophore with columella and one projection. G. Sporangiophore with a sweeling and columella. H. Rhizoid. I. Sporangiospores. Scale bars: B–H = 25 μm; I = 10 μm.

Etymology: Reference to the varied-shaped sporangiospores formed.

Typus: Brazil, Saloá, from soil sample, 11 May 2019, T.R.L. Cordeiro (holotype URM 8720h, ex-type culture URM 8720, GenBank numbers: ITS = PP844915 (c1), PP844916 (c2), LSU = PP956886, act-1 = PP893205).

Description: Colonies on MEA at 25 °C, slow growing, initially white, turning pale to light grey, with 9 cm diam after 5 d of incubation; marginal colonies are frequent. Rhizoids short or long, branched or unbranched. Stolons light brown, wall slightly encrusted. Sporangiophores erect, arising from stolons, usually unbranched, single or in whorls of up to six, up to 5 μm in width; when branched they usually form monopodial, rarely sympodial branches; one swelling may be present below the sporangium. One septum is observed near the apophysis, two septa are rarely present. Sporangia brown, globose (15–)20–35(–40) μm diam, apophysate, wall smooth and deliquescent. Apophysis brownish, long or short, cup-shaped, (2–)4.5–9.5(–12) × (1.5–)5.5–15 μm. Columellae hyaline, globose or subglobose to conical, (4.5–)7–15(–20) × (5.5–)7–17(–20) μm, smooth-walled; collar visible. One projection on columellae generally needle-, kidney-, or triangular-shaped, up to 2.5 × 7 μm is observed. Rarely a second projection is observed on columellae. Some columellae may not show any projections on their surface. Sporangiospores hyaline, mostly cylindrical and elliptical, (2.5–)3–6(–7.5) × 2.5–4(–5) μm, globose or subglobose, (2–)3.5–6(–7.5) μm diam, smooth-walled. Chlamydospores absent. Zygospores not observed.

Culture characteristics: On PDA, the colonies attained 9 cm diam after 4 d at 25 °C. On SMA, the colonies attained 48 mm diam after 4 d at 25 °C. At 5 °C, the strain grew slowly, reaching 20 mm diam after 20 d. Colonies initially white, becoming grey after 7 d. Sporulation excellent on all media. Maximum growth temperature was 32 °C on MEA, PDA, and SMA (Supplementary Table S2).

Notes: In our phylogenetic tree (Fig. 1), A. variispora is phylogenetically close to A. tarda, A. cheongyangensis and A. variicolumellata. However, varied-shaped sporangiospores are only formed in A. variispora, which are mostly cylindrical and elliptical, globose or subglobose, whereas sporangiospores of other species are oval to short cylindrical (A. cheongyangensis), cylindrical constricted in the centre, some short cylindrical (A. tarda) and cylindrical and slightly constricted in the centre (A. variicolumellata; Freitas et al. 2022). The formation of columellae with triangular and kidney-shaped projections is another intrinsic feature of A. variispora not observed in those other three species mentioned above.

Cunninghamella Matr., Ann. Mycol. 1: 47. 1903.

Cunninghamella intermedia K.B. Deshp. & Mantri, Mycopathol. Mycol. Appl. 28: 343. 1966. MycoBank MB 329433. Fig. 15.

Fig. 15.

Fig. 15

Cunninghamella intermedia (CNUFC IL021). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G. Sporangiola on branched sporophore (observed under compound microscope). H. Development of sporangiola on vesicles. I–K. Vesicles bearing sporangiola. L. Sporangiola. Scale bars = 20 μm.

Description: Colonies grew fast on SMA, attaining a diameter of 54 mm after 3 d at 25 °C. Colonies initially white, then gradually becoming grey; reverse greyish-white. Rhizoids frequent, simple or finger-like. Sporophores arising from stolons or from aerial hyphae, 5.5–11.5 μm diam, simple or rebranching into secondary branches, (14–)39–224.5 μm long. Septa in sporangiophore sometimes present at its base. Vesicles of main axes of the sporophores light grey, globose to subglobose, 17.5–35(–38.5) μm diam. Lateral vesicles light grey, smooth-walled, globose to subglobose, 14.5–25(–29) μm diam. Sporangiola smooth-walled, mostly globose to subglobose, (9–)11.5–13.5(–15.5) μm diam, hyaline to brownish to dark brown. Chlamydospores not seen. Zygospores were not observed. The shape and size of the vesicles on MEA, PDA and SMA were similar. Sporangiola on PDA and SMA were similar, but slightly larger on MEA (up to 17.5 μm).

Culture characteristics: On PDA, the colonies attained 45 mm diam after 3 d at 25 °C. On MEA, the colonies attained 47 mm diam after 3 d at 25 °C. Optimal growth was observed around 30–37 °C, slow growth was observed at 10 °C. Maximum growth temperature was 45 °C on MEA, PDA and SMA (Supplementary Table S2).

Material examined: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from the leg of Pholcus phalangioides (house ghost spider), 3 Oct. 2020, H.B. Lee, culture CNUFC IL021.

Notes: Morphologically, the isolated strain presents similar features with C. intermedia as previously described (Zheng & Chen 2001), such as sporangiola size and culture characteristics. However, the terminal vesicle sizes reported in the literature (19–32 μm) (Zheng & Chen 2001) are slightly smaller than those of our isolate.

Backusellaceae K. Voigt & P.M. Kirk

Backusella Hesselt. & J.J. Ellis, Mycologia 61: 863. 1969.

Backusella obliqua C.L. Lima et al., J. Fungi 8: 1038. 2022. MycoBank MB 559793. Fig. 16.

Fig. 16.

Fig. 16

Backusella obliqua (CNUFC S805). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G. Sporangium. H–J. Multi-spored sporangiola borne on circinate branches. K, L. Columellae from sporangia. M. Columellae of multi-spored sporangiola. N. Sporangiospores. Scale bars = 20 μm.

Description: Colonies on SMA develop rapidly, reaching 70 mm diam after 4 d of incubation at 25 °C. Colonies white to light grey; reverse light grey. Sporangiophores arising from substrate, simple, monopodial, or sympodially branched, curved when young and erect at maturity, 5.5–11.5 μm wide. Sporangia globose to subglobose, pale brown, multi-spored, 35–61.5 μm diam. Columellae of sporangia globose or ovoid, 18.5–34.5 × 17.5–30.5 μm with small collar. Multi-spored sporangiola abundant, globose and subglobose, 15–27.5 μm diam, wall spinulose, containing (2–)3–12(–18) μm sporangiospores. Columellae of sporangiola hemispherical, 9.5–22 μm, smooth-walled. Uni-spored sporangiola globose, 13.5–25.5 μm diam, wall spinulose. Sporangiospores of sporangia and multi-spored sporangiola were similar, mostly subglobose, 10.5–13.5 × 9.5–11.5 μm. Chlamydospores or zygospores were not observed.

Culture characteristics: On PDA, the colonies attained 60 mm diam after 5 d at 25 °C. On MEA, the colonies attained 68 mm diam after 5 d at 25 °C. Maximum growth temperature was 32 °C on MEA, PDA and SMA (Supplementary Table S2).

Material examined: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from Apis sp., 25 Sep. 2021, H.B. Lee, culture CNUFC S805.

Notes: In the combined phylogenetic tree, our strain clustered with the ex-type strain of B. obliqua. Our isolate resembles B. obliqua in shape and size of sporangia and sporangiospores. Some columellae may have one side more swollen than the other and some are arranged obliquely on the sporangiophores presented in the ex-type isolate of B. obliqua, which were also observed in our isolate.

Backusella varians Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900708. Fig. 17.

Fig. 17.

Fig. 17

Backusella varians (CNUFC CY2201). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G. Sporangiophores with sporangia. H. Branched sporangiophore with sporangium. I, J. Sporangia. K. Multi-spored sporangiola. L. Columella of multi-spored sporangiola. M. Sporangiospores. Scale bars: G = 200 μm; H = 100 μm; I–M = 20 μm.

Etymology: Refers to the species producing sporangiospores of various shapes.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, decaying leaves in soil, 4 Jul. 2022, H.B. Lee & J.S. Kim (holotype CNUFC HT2281, ex-type culture CNUFC CY2201, GenBank numbers: ITS = PP844920 (c1), PP844921 (c3), PP844922 (c5), LSU = PP851422, rpb1 = PP860560, argA = PP860551).

Description: Colonies on SMA white, reaching 53 mm diameter after 4 d of incubation at 25 °C; reverse light grey. Sporangiophores arising from substrate mycelia, curved when young and erect at maturity, hyaline or with greenish grey granular content, slightly constriction below sporangium, 7–12 μm wide. Sporangia globose to subglobose, greyish yellow, 45–65(–72.5) μm diam. Columellae of sporangia globose, subglobose, oval, some ellipsoid 22–36.5(–40) × 20–33(–36) μm with small collar, hyaline to greyish. Multi-spored sporangiola arising from pedicels, globose, 15–38.5 μm diam, wall spinulose, pale yellow. Columellae from sporangiola hemispherical, conical, subglobose, up to 17.5 μm long, hyaline to greyish, smooth-walled. Collar present. Unispored sporangiola globose, 17–27.5 μm diam, minutely spinulose. Sporangiospores of sporangia and multi-spored sporangiola were similar, hyaline to greyish, globose, subglobose, oval and irregular, 10–14 × 9–12.5 μm, containing small greenish granules, smooth-walled. Rhizoids well branched. Chlamydospores present. Zygospores and giant cells not observed. On PDA, sporangia [39–63.5 (–70.5) μm diam] were slightly smaller than those on SMA. The shape and size of sporangiospores were similar on SMA, PDA, and MEA.

Culture characteristics: On PDA, the colonies attained 51 mm diam after 4 d at 25 °C. On MEA, the colonies attained 61 mm diam after 4 d at 25 °C. Slow growth was observed at 30 °C on MEA, PDA, and SMA. Maximum growth temperature was 32 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, decaying leaves in soil, 4 Jul. 2022, H.B. Lee & J.S. Kim, culture CNUFC CY2202; from Scolopendra sp., 1 Apr. 2024, H.B. Lee & J.S. Kim, cultures CNUFC CY2404 and CNUFC CY2408.

Notes: Our phylogenetic analyses showed that B. varians forms a separate lineage, phylogenetically related to B. azygospora, B. brasiliensis, B. liffmaniae, B. mclennaniae, and B. psychrophila (Fig. 3). Backusella varians produces larger sporangia than B. liffmaniae (26.5–55.2 × 26–54.1 μm), B. mclennaniae (25.6–39.6 × 24.2–39.1 μm), and B. psychrophila (28.1–43.5 × 28.1–40.5 μm) (Urquhart et al. 2021). Backusella brasiliensis produces abundant giant cells, whereas giant cells have not observed in B. varians (Lima et al. 2022).

Lichtheimiaceae Kerst. Hoffmann, Walther & K. Voigt

Circinella Tiegh. & G. Le Monn., Ann. Sci. Nat., Bot. 17: 298. 1873.

Circinella umbellata Tiegh. & G. Le Monn., Ann. Sci. Nat. Bot. 17: 300. 1873. MycoBank MB 196867. Fig. 18.

Fig. 18.

Fig. 18

Circinella umbellata (CNUFC CY2220). A, D. Colony on SMA. B, E. Colony on MEA. C, F. Colony on PDA (A–C obverse view, D–F reverse view). G–L. Umbels of sporangia. M. Mature sporangium. N. Columella with sporangial wall attached and sporangiospores. O. Sporangiospores. Scale bars: G–K = 250 μm; L, M = 50 μm; N, O = 20 μm.

Description: Colonies on SMA attaining 32 mm diameter after 5 d at 25 °C, at first white, becoming golden brown after 14 d; reverse dark brown. Rhizoids absent. Sporangiophores simple or branched, erect, or circinate, raising from the substrate or aerial mycelia, branches 7.5–11.5(–14) μm diam, each branch with a single sporangium or more often terminating into an umbel of 4–14 sporangia. Sporangia hyaline to dark brown, multi-spored, globose to subglobose, with deliquescent wall, (40.5–)50–93.5(–108) μm diam. Columellae ovoid, oblong, or slightly pyriform, (21–)31.5–47.5(–50) × (20–)27–44.5 μm, brownish or light greyish brown. Sporangiospores globose to ovoid 7–8.5 × 6.5–8 μm, smooth-walled, brownish grey. Zygospores not observed. On PDA, sporangiospores (up to 9.5 μm diam) and columellae (up to 54 × 47.5 μm) were slightly larger than those on MEA and SMA. On MEA, sporangia (up to 75 μm diam) and columellae (up to 47 × 37 μm) were slightly smaller than those on PDA and SMA.

Culture characteristics: On PDA, the colonies attained 42 mm diam after 5 d at 25 °C. On MEA, the colonies attained 55 mm diam after 5 d at 25 °C. Maximum growth temperature was 35 °C on MEA and PDA, while 34 °C on SMA (Supplementary Table S2).

Material examined: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from Tiphia sp., 20 Jun. 2022, H.B. Lee, culture CNUFC CY2220.

Notes: This species can be morphologically differentiated from other species in the genus by forming sporangiophores terminating in an umbel of 2–12 sporangia. According to Crous et al. (2020) and our Fig. 4 C. umbellata is phylogenetically closer to C. lampensis. The main morphological difference between both species is because the latter forms sporangiophores with an umbel of 2–6 sporangia. This is the first report of this species in South Korea.

Fig. 19.

Fig. 19

Neofennellomyces jeongsukae (CNUFC CY2001). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G, H. Branched sporangiophore. I. Pedicellate sporangia arranged along the sporangiophore. J, K. Terminal sporangium. L. Pedicellate sporangium. M–R. Typical columellae. S, T. Columella and sporangiospores from pedicellate sporangia. U. Rhizoid. V. Sporangiospores. Scale bars = 20 μm.

Neofennellomyces Hyang B. Lee & T.T.T. Nguyen, gen. nov. Index Fungorum 900613.

Etymology: Refers to its morphological similarity to Fennellomyces.

Description: Sporangiophores erect, arising directly from the substrate mycelium, simple or branched sympodially, producing pedicellate sporangia. Sporangiospores elliptical to broadly elliptical or ovoid.

Type species: Neofennellomyces jeongsukae Hyang B. Lee & T.T.T Nguyen

Notes: Neofennellomyces differs from Circinella by forming sporangia born on erect sporangiophores, besides sporangia born on pedicellate branches. It differs from Fennellomyces by not presenting a swelling distally immediately below a large terminal sporangium, which is observed in the latter genus.

Neofennellomyces jeongsukae Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum number 900612. Fig. 19.

Etymology: In honour of the collector of the type specimen, Ms. Jeong Suk Kim.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from the antenna of the earwig (Euborellia sp.), 20 Oct. 2020, J.S. Kim (holotype CNUFC HT2004, ex-type culture CNUFC CY2001, GenBank numbers: ITS = PP858084, LSU = PP858090, mcm7 = PP960588 (c1), PP960589 (c2)).

Description: Colonies on SMA attaining 42 mm diam after 5 d at 25 °C, at first white, turning to moderate orange yellow; reverse strong brown, irregularly zonate. Rhizoids present, finger-like, branched. Sporangiophores erect, simple or branched sympodially up to three times or rarely four times, (5–)6–13 μm diam, arising directly from the substrate mycelium, hyaline to brownish towards columella, smooth, sometimes with one or two septa below sporangia, and terminated by sporangia. Pedicellate sporangia are found to be irregularly arranged along the sporangiophores after 5 d, singly or only two arising from the same point. Sporangia borne erect, globose, subglobose to pyriform, (32–)39–77 × (31–)35–74.5 μm. Columellae variable in shape, subglobose, obovoid, oblong, 33.5–47(–50.5) × 23.5–32(–44) μm, and pyriform, or pyriform broadened at the base, 42.5–63.5 × 34.5–48 μm, with one or more projections at the apex on the smaller columella, wart-like, needle-shaped, conical, or irregular. Collar present. Pedicellate sporangia, globose to subglobose, 19.5–39.5 μm diam. Columellae of pedicellate sporangia subglobose to hemispherical, 15–25.5 μm diam, smooth, brownish. Sporangiospores of terminal sporangia narrowly elliptical, 5–8.5 × 3–4.5 μm, and broadly elliptical in pedicellate sporangia (6.5–)7.5–10 × (3.5–)4–5.5 μm. Chlamydospores and zygospores were not observed.

Culture characteristics: Optimal growth was observed around 25 °C, slow growth was observed at 10 °C and 35 °C. Maximum growth temperature 36 °C (without sporulation) on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, from Porcellio scaber, 4 Jul. 2022, H.B. Lee & J.S. Kim, culture CNUFC CY2204.

Notes: This species is characterized by the variable shape of its columellae, projections, terminated by sporangia borne erectly. Pedicellate sporangia irregularly arranged along the sporangiophores. The species grew well on SMA but displayed restricted growth and very rare aerial mycelia on PDA and MEA.

Mucoraceae Fr.

Mucor Fresen., Beitr. Mykol. 1: 7. 1850.

Mucor albicolonia Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900709. Fig. 20.

Fig. 20.

Fig. 20

Mucor albicolonia (CNUFC CY2027). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G, H. Sporangiophores with sporangia. I–K. Mature sporangium with columella and sporangiospores. L, M. Columellae. N. Sporangiospores. Scale bars: G, H = 500 μm; I–N = 20 μm.

Etymology: Refers to the white colonies of this species.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, Nephila sp., 18 May 2020, H.B. Lee & J.S. Kim (holotype CNUFC HT2002, ex-type culture CNUFC CY2027, GenBank numbers: ITS = PP844894, LSU = PP851426).

Description: Colony on MEA at 20 °C, white, high, slow growing, reaching 37 mm diam after 3 d of incubation. Sporangiophores arising directly from the substrate, simple, rarely single-branched, straight, 7–18.5 μm wide. Sporangia globose, wall slightly echinulate, rapidly deliquescing, (30–)40–83 μm diam. Columellae mostly subglobose, cylindrical with a truncated base, ellipsoidal or pyriform, (21–)24.5–39.5 × (18.5–)21–33.5 μm, hyaline to greyish. Collar present. Sporangiospores hyaline to greyish, oval to subglobose and irregular, 11–17.5 × 9.5–15 μm, smooth-walled. No chlamydospores or zygospores were observed.

Culture characteristics: On PDA, the colonies attained 34 mm diam after 3 d at 20 °C. On SMA, the colonies attained 13 mm diam after 3 d at 20 °C. Maximum growth temperature was 33 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional materials examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, Nephila sp., 18 May 2020, H.B. Lee & J.S. Kim, culture CNUFC CY2028; Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from soil, 5 Dec. 2023, H.B. Lee, culture CNUFC CY2311.

Notes: Mucor albicolonia is phylogenetically related to M. gigasporus. Mucor albicolonia differs from M. gigasporus by having a variable shape of columellae (subglobose, cylindrical with a truncated base, ellipsoidal or pyriform), and shorter sporangiospores (11–17.5 × 9.5–15 μm vs. 11.4–30.6 × 9.9–15.3 μm) (Chen & Zheng 1986).

Mucor aurantiacus Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900761. Fig. 21.

Fig. 21.

Fig. 21

Mucor aurantiacus (CNUFC CY030). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G. Sporangiophores with sporangia. H, I. Mature sporangia. J–M. Columellae. N. Sporangiospores. O. Chlamydospores. Scale bars: G = 500 μm; H–O = 20 μm.

Etymology: Refers to its orange colour on all media.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from Protaetia orientalis, 13 Jun. 2020, H.B. Lee & J.S. Kim (holotype CNUFC HT2001, ex-type culture CNUFC CY030, GenBank numbers: ITS = PP844897, LSU = PP851429, rpb1 = PP893222).

Description: Colonies on MEA at 25 °C, first white, becoming orange, reaching 60 mm diam after 5 d of inoculation; reverse concolourous. Sporangiophores erect, unbranched, 8.5–17 μm diam. Sporangia globose to subglobose, 85–146(–173) × 81–144(–166) μm, rapidly deliquescent, orange. Columellae usually globose or subglobose, or slightly obovoid, hyaline to orange, smooth-walled, 55–118 × 51–115 μm. Collar present. Sporangiospores smooth, hyaline, mostly ellipsoidal, some globose or oval, 4.5–7 × 3.5–5.5 μm. Chlamydospores abundant, formed on substrate hyphae, in chains, mostly subglobose, cylindrical to irregular, intercalary or terminal. Zygospores not observed.

Culture characteristics: On PDA, the colonies attained a diameter of 56 mm after 5 d at 25 °C. On SMA, the colonies attained 64 mm diam after 5 d at 25 °C. Maximum growth temperature was 41 °C on MEA and PDA, while 40 °C on SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, from Protaetia orientalis, 13 Jun. 2020, H.B. Lee, culture CNUFC CY031.

Notes: In the combined phylogeny (Fig. 5), M. aurantiacus clustered as a sister taxon to M. azygosporus. Mucor aurantiacus differs from M. azygosporus by producing smaller sporangiospores [4.5–7 × 3.5–5.5 μm vs. 5–13(–19) × 4–12 μm], and varied in shape columellae (globose to subglobose, and or slightly obovoid). Furthermore, M. azygosporus produces abundant zygospores, while M. aurantiacus does not produce any (Benjamin & Mehrotra 1963).

Mucor cryophilus Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900611. Fig. 22.

Fig. 22.

Fig. 22

Mucor cryophilus (CNUFC CHS1). A. Colony on SMA. B. Colony on PDA. C. Colony on MEA. D. Sporangiophores with sporangia. E–G. Young and mature sporangia. H–J. Typical columellae. K. Sporangiospores. L. Sporangiospores germination. Scale bars: D = 500 μm; E–L = 20 μm.

Etymology: Refers to the cold-loving character of the species.

Typus: South Korea, Jeonnam Province, Chonnam National University, from cherry seeds, 30 Mar. 2021, H.B. Lee & T.T.T. Nguyen (holotype CNUFC HT2155, ex-type culture CNUFC CHS1, GenBank numbers: ITS = PP844923 (c1), PP844924 (c2), PP844925 (c4), LSU = PP852708, rpb1 = PP886119).

Description: Colonies on MEA at 20 °C, fast growing, at first white, becoming grey, reaching 76 mm diam after 3 d of inoculation; reverse uncoloured. Sporangiophores erect, 10–35.5 μm diam, with a wider base and a slight constriction next to the sporangium, simple or single-branched. Sporangia globose or subglobose, hyaline to greyish-yellow, deliquescent-walled, (28–)45–116(–135) × (27.5–)43–110(–132) μm. Columellae hyaline to greyish brown, various shapes, subglobose, some ellipsoidal, or oblong to pyriform, 35.5–68 × 34–44.5 μm, with a distinct collar. Sporangiospores globose, subglobose, ovoid, 7–12 × 6–9 μm, or irregular, 9–16(–19) × 7–11.5 μm, sometimes slightly flattened at one side, hyaline to greyish brown. Rhizoids present. Zygospores not observed.

Culture characteristics: On PDA, the colonies attained 74 mm diam after 3 d at 20 °C. On SMA, the colonies attained 82 mm diam after 3 d at 20 °C. Poor development seen at 25 °C; optimum growth and good sporulation seen at 15–20 °C; slow, poor development observed at 5 °C. Maximum growth temperature was 28 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Jeonnam Province, Chonnam National University, from cherry seeds, 30 Mar. 2021, H.B. Lee & T.T.T Nguyen, culture CNUFC CHS2.

Notes: Mucor cryophilus is characterized by restricted growth on MEA, PDA and SMA incubated at 25 °C, and variable columellae (subglobose, some ellipsoidal, or oblong to pyriform).

Mucor glutinatus Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900710. Fig. 23.

Fig. 23.

Fig. 23

Mucor glutinatus (CNUFC CY2012). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G, H. Sporangiophores with sporangia. I. Young sporangium. J, K. Sporangiophore with columella and sporangiospores. L, M. Columellae. N. Sporangiospores. Scale bars: G = 500 μm; H = 250 μm; I–N = 20 μm.

Etymology: Refers to tightly agglutinated conidia.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, soil on the skin surface of toad (Bufo gargarizans), 13 Jun. 2020, H.B. Lee & J.S. Kim (holotype CNUFC HT2027, ex-type culture CNUFC CY2012, GenBank numbers: ITS = PP844899, LSU = PP852710, rpb1 = PP886121).

Description: Colony on MEA at 20 °C, white, gradually becoming white cotton, reaching 34 mm diam after 3 d of incubation. Sporangiophores arising directly from the substrate, simple, or single-branched, straight, 5.5–12 μm wide. Sporangia globose, wall slightly echinulate, rapidly deliquescent, 29–71 μm diam. Columellae subglobose, oval, smooth-walled, hyaline to greyish, with or without yellowish or greenish grey granular contents, sometimes flattened on one side, (17.5–)20.5–34.5 × (15.5–)18–32.5 μm. Collar present. Sporangiospores hyaline to greyish, globose, (12–)14–18.5 μm diam, smooth-walled. No chlamydospores or zygosporangia were observed.

Culture characteristics: On PDA, the colonies attained 32 mm diam after 3 d at 20 °C. On SMA, the colonies attained 15 mm diam after 3 d at 20 °C. Slow growth was observed at 30 °C on MEA, PDA, and SMA. Maximum growth temperature was 33 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, soil on the skin surface of toad (Bufo gargarizans), 13 Jun. 2020, H.B. Lee & J.S. Kim, culture CNUFC CY2016.

Notes: Mucor glutinatus is characterized by producing globose sporangiospores, subglobose, oval or sometimes flattened on one side columellae.

Mucor paraorantomantidis Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 901134. Fig. 24.

Fig. 24.

Fig. 24

Mucor paraorantomantidis (CNUFC CY205). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G. Sporangiophores with sporangia. H, I. Sporangia. J–L. Columellae. M. Sporangiospores. N. Chlamydospores. Scale bars: G = 250 μm; H–N = 20 μm.

Etymology: Refers to the phylogenetic proximity to Mucor orantomantidis.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, Isyndus obscurus, 5 Aug. 2020, H.B. Lee & J.S. Kim (holotype CNUFC HT2010, ex-type culture CNUFC CY205, GenBank numbers: ITS = PP844901, LSU = PP851431, rpb1 = PP893224).

Description: Colony on MEA at 25 °C, light yellow, reaching 52 mm diam after 4 d of incubation; reverse light yellow. Sporangiophores erect, arising directly from the substrate, unbranched, with a wider base and a slight constriction next to the sporangium, 6–11.5 μm wide. Sporangia globose to subglobose, (33–)48–73.5 × (32.5–)46–72.5 μm. Columellae globose to subglobose, ellipsoid, 17.5–30 × 15.5–26.5 μm, hyaline, smooth, with or without yellowish granular contents. Collar present. Sporangiospores hyaline to greenish grey, elliptical to fusiform and some elliptical flattened at one side, (5.5–)7–11.5 × 3–5 μm. Chlamydospores common in substrate hyphae, terminal or intercalary, single or in long chains, hyaline, globose to subglobose, ellipsoidal, cylindrical, or irregular. Zygospores not observed.

Culture characteristics: On PDA, the colonies attained 44 mm diam after 4 d at 25 °C. On SMA, the colonies attained 32 mm diam after 4 d at 25 °C. Maximum growth temperature was 35 °C on SMA, while 34 °C on MEA and PDA (SupplementaryTable S2).

Material examined: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from Isyndus obscurus, 5 Aug. 2020, H.B. Lee & J.S. Kim, culture CNUFC CY206.

Notes: In the combined phylogeny (Fig. 5), M. paraorantomantidis forms a sister taxon with M. orantomantidis. Mucor paraorantomantidis differs from M. orantomantidis based on its combination of the sizes of sporangia, columellae, and sporangiospores. Mucor paraorantomantidis has narrower spoangiospores than M. orantomantidis [(5.5–)7–11.5 × 3–5 μm vs. 8–11.5 × 4.5–6 μm]. In addition, zygospores are not formed in M. paraorantomantidis.

Mucor saturninus Hagem, Ann. Mycol. 8: 265. 1910. MycoBank MB 191108. Fig. 25.

Fig. 25.

Fig. 25

Mucor saturninus (CNUFC IO1). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G. Sporangiophores with sporangia. H. Monopodial branching of sporangiophores with sporangia. I, K. Branched sporangiophore with columella and sporangium. J. Mature sporangium. L. Columella. M. Sporangiospores. Scale bars = 20 μm.

Description: Colony on MEA at 20 °C, fast growing, light brownish grey, reaching 61 mm diam after 4 d of incubation; reverse yellowish white. Sporangiophores growing directly from the substrate, simple or with short or long sympodial branches, 10–21.5 μm diam, sometimes with one or several septa. Sporangia globose to subglobose, yellow when young, then brownish, (50.5–)56.5–113(–127) μm diam, walls slightly echinulate; a few sporangia with 23–35 μm diam were observed on the short monopodial branches. Columellae variable in shape, ellipsoid, obovoid, subglobose, (22–)27–49.5 × 23.5–51 μm, or cylindrical-ellipsoidal (44.5–55 × 50–76.5 μm). Sporangiospores mostly ellipsoidal, sometimes irregular, 7.5–10 × 4.5–7 μm. Chlamydospores and zygospores not observed. On PDA, sporangia (up to 100 μm diam) were slightly smaller than those on SMA and MEA. Sporangiospores on SMA (up to 12 × 8 μm) were slightly larger than those on MEA and PDA.

Culture characteristics: On PDA, the colonies attained 60 mm diam after 4 d at 20 °C. On SMA, the colonies attained 64 mm diam after 4 d at 20 °C. Optimal growth was observed around 15–20 °C. Slow growth was observed at 5 °C. Maximum growth temperature was 27 °C on MEA, PDA and SMA (Supplementary Table S2).

Material examined: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, from Isyndus obscurus, 3 Oct. 2020, H.B. Lee & J.S. Kim, culture CNUFC IO1.

Notes: The isolate CNUFC IO1 was placed within the clade of M. saturninus (Fig. 6). However, CNUFC IO1 has smaller sporangia compared to ex-neotype strain of M. saturninus [(50.5–)56.5–113(–127) μm vs. up to 175 μm]. Sporangiospores of CNUFC IO1 are larger than those of the ex-neotype strain [(7.5–10 × 4.5–7 μm) vs. (4.5–8 × 3–5.4 μm); Schipper 1975].

Mucor timomeni Hyang B. Lee & T.T.T. Nguyen, sp. nov. Index Fungorum 900610. Fig. 26.

Fig. 26.

Fig. 26

Mucor timomeni (CNUFC CY701). A, D. Colony on SMA. B, E. Colony on PDA. C, F. Colony on MEA (A–C obverse view, D–F reverse view). G, H. Young and mature sporangia. I–L. Typical columellae. M. Sporangiospores. N. Chlamydospores. Scale bars: G–L, N = 20 μm; M = 10 μm.

Etymology: Refers to the host Timomenus komarovi, from which the species was first isolated.

Typus: South Korea, Kunryang-ri, Cheongyang-eup, Cheongyang, Chungnam Province, Timomenus komarovi, 18 Apr. 2020, H.B. Lee & J.S. Kim (holotype CNUFC HT2013, ex-type culture CNUFC CY701, GenBank numbers: ITS = PP844926 (c1), PP844927 (c2), LSU = PP892773, rpb1 = PP893226).

Description: Colonies on MEA at 25 °C, fast growing, at first white, becoming mouse grey, reaching 67 mm diam after 4 d of incubation; reverse mouse grey. Sporangiophores erect, 3.5–9 μm diam, simple or branched; one swelling is sometimes seen on the sporangiophore below sporangia. Some branches bear a sterile sporangium. Sporangia globose, 20.5–45.5(–50) μm diam, first slightly yellow, then becoming light brown, rapidly deliquescing. Columellae highly variable in shape, subglobose, oval, sometimes flattened at one side, 15.5–27.5 × 14–22.5 μm; one septum is sometimes found below the columella. Collar distinct. Sporangiospores ellipsoid, sometimes flattened on one side, 4.5–6 × 2.5–3 μm. Chlamydospores abundantly borne substrate hyphae in chains, globose, subglobose, doliform, or irregular, 10.5–20.5 μm. Zygospores not observed. The shape and size of the sporangia and columellae on MEA, PDA and SMA were similar. Sporangiospores on SMA were slightly larger (4.5–6.5 × 2.5–3.5 μm) than those on MEA and PDA.

Culture characteristics: On PDA, the colonies attained a diam of 61 mm after 4 d at 25 °C. On SMA, the colonies attained a diam of 69 mm after 4 d at 25 °C. Slow growth was observed at 30 and 10 °C on MEA, PDA and SMA. Maximum growth temperature was 32 °C on MEA, PDA and SMA (Supplementary Table S2).

Additional material examined: South Korea, Kunryang-ri, Cheongyangeup, Cheongyang, Chungnam Province, from soil sample, 8 Feb. 2021, H.B Lee & J.S. Kim, culture CNUFC CY2118.

Notes: Mucor timomeni is phylogenetically closely related to M. chiangraiensis, and M. nederlandicus (Fig. 5). However, M. timomeni differs from these species by producing variable columellae, sometimes flattened on one side. Mucor nederlandicus produces sporangia up to 70 μm diam (Dade 1937), while those of M. timomeni are up to 50 μm diam. The sporangiospores and columellae of M. timomeni are slightly smaller than those of M. chiangraiensis, but the sporangia of the new species are slightly larger than those of M. chiangraiensis (Hurdeal et al. 2021).

Species of Mucorales in South Korea and Brazil

We have accessed all the Mucorales fungi reported in South Korea and Brazil until June 2024, which are represented by 120 species of from Brazil and 91 from South Korea (Table 2). We have found a similarity of 37.19 % (Sørensen 1978) of the Mucorales communities between both countries.

Table 2.

Updated list of Mucorales species from South Korea and Brazil including 15 new species and four new records identified in the current study.

Species South Korea Brazil

Substrate or Host Reference Substrate or Host References
Absidia aguabelensis n.a n.a Soil Leitão et al. (2021)
A. bonitoensis n.a n.a Soil Lima et al. (2021)
A. catingaensis n.a n.a Soil Ariyawansa et al. (2015)
A. cheoangyangensis Leaf in rainwater; soil This study n.a n.a
A. cornuta n.a n.a Soil Lima et al. (2020a)
A. cuneospora n.a n.a Soil Upadhyay (1970)
A. cylindrospora n.a n.a Herbivore dung; leaf litter; soil Souza et al. (2017), Lima et al. (2020c)
A. fluvii Rainwater; soil This study n.a n.a
A. glauca Soil Nguyen et al. (2016) n.a n.a
A. jindoensis Soil Wanasinghe et al. (2018) n.a n.a
A. koreana Soil Ariyawansa et al. (2015) n.a n.a
A. kunryangriensis Rainwater This study n.a n.a
A. multispora n.a n.a Soil Cordeiro et al. (2020)
A. paracylindrospora Soil This study n.a n.a
A. pararepens Rainwater; Scolopendra sp. This study Soil Cordeiro et al. (2022)
A. pernambucoensis n.a n.a Soil Lima et al. (2020a)
A. pseudocylindrospora Soil Nguyen et al. (2016) Soil Lima et al. (2020c)
A. repens n.a n.a Soil Schoenlein-Crusius et al. (2006)
A. saloaensis n.a n.a Soil Cordeiro et al. (2020)
A. spinosa n.a n.a Soil Trufem (1981b)
A. stercoraria Rat dung Li et al. (2016) n.a n.a
A. tarda n.a n.a Soil This study
A. variicolumellata n.a n.a Soil Freitas et al. (2022)
A. varriprojecta n.a n.a Soil This study
A. variispora n.a n.a Soil This study
Actinomucor elegans Freshwater Nguyen et al. (2017) Soil Upadhyay (1967)
Apophysomyces elegans n.a n.a Soil Santiago & Maia (2010)
Backusella azygospora n.a n.a Soil Crous et al. (2019)
B. brasiliensis n.a n.a Soil Lima et al. (2022)
B. chlamydospora Porcellio scaber; Theuronema hilgendorfi hilgendorfi Nguyen et al. (2021) n.a n.a
B. constricta n.a n.a Soil Lima et al. (2016b)
B. circina Soil Nguyen & Lee (2018) n.a n.a
B. gigacellularis n.a n.a Soil De Souza et al. (2014)
B. koreana Scolopendra morsitans Nguyen et al. (2021) n.a n.a
B. lamprospora n.a n.a Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Santiago
et al. (2011)
B. locustae Feces of grasshopper Wanasinghe et al. (2018) n.a n.a
B. obliqua Apis sp. This study Soil Lima et al. (2022)
B. oblongielliptica Timomenus komarovi Nguyen et al. (2021) n.a n.a
B. oblongispora Skin of Bufo gargarizans Nguyen et al. (2021) n.a n.a
B. paraconstricta n.a n.a Soil Santos et al. (2023a)
B. pernambucensis n.a n.a Soil Cordeiro et al. (2023)
B. thermophila Gryllus bimaculatus Nguyen et al. (2021) n.a n.a
B. variabilis n.a n.a Leaf litter; soil Schoenlein-Crusius et al. (2006)
B. varians Decaying leave in soil; Scolopendra sp. This study n.a n.a
Blakeslea trispora Gut of grasshopper and soldier fly larva Nguyen & Lee (2016) n.a n.a
Choanephora cucurbitarum Lactuca sativa; Althaea officinalis; Solanum melongena; Okra (Abelmoschus esculentus) Ryu et al. (2022), Choi et al. (2016), Kwon & Jee (2005), Park et al. (2015) Crotalaria spectabilis; C. striata Schrank.; Hevea brasiliensis; Tephrosia toxicaria; T. vogelii; Capsicum brazilianum Vell.; Vigna sp Mendes & Urben (2017), Alfenas et al. (2018)
Ch. infundibulifera Hibiscus rosa-sinensis Park et al. (2014) Soil Upadhyay (1970)
Circinella angarensis n.a n.a Soil Trufem (1981a)
C. minor n.a n.a Soil Trufem (1981a)
C. muscae Leaves of Toxicodendron sylvestre Nguyen et al. (2018) Herbivore dung; soil Trufem (1981a), Santiago et al. (2011)
C. umbellata Tiphia sp. This study Herbivore dung Santiago et al. (2011)
Cunninghamella bertholletiae Freshwater Nguyen et al. (2017) Soil Lima et al. (2020c)
C. binariae Soil Nguyen et al. (2019) n.a n.a
C. blakesleeana n.a n.a Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Santiago et al. (2011), Lima et al. (2020c)
C. clavata n.a n.a Soil Alves et al. (2017)
C. echinulata Soil Nguyen et al. (2017) Herbivore dung; leaf litter; soil Santiago et al. (2011), Lima et al. (2020c)
C. elegans Soil Nguyen et al. (2017) Soil Lima et al. (2020c)
C. gigacelullaris n.a n.a Soil Hyde et al. (2016)
C. intermedia Pholcus phalangioides This study n.a n.a
C. phaeospora n.a n.a Herbivore dung; soil Santiago et al. (2011), Lima et al. (2020c)
Gilbertella persicaria Freshwater Lee et al. (2018) Herbivore dung; soil Santiago & Cavalcanti (2008), Santiago et al. (2011)
Gongronella brasiliensis n.a n.a Soil Tibpromma et al. (2017)
G. butleri Soil Babu et al. (2015) Soil Lima et al. (2020c)
G. guangdongensis Soil Wajihi et al. (2018) n.a n.a
G. koreana Soil Ariyawansa et al. (2015) n.a n.a
G. lacrispora n.a n.a Soil Upadhyay (1969)
G. namwonensis Freshwater Crous et al. (2020) n.a n.a
G. orasabula Soil Li et al. (2016) n.a n.a
G. pedratalhadensis n.a n.a Soil Freitas et al. (2020)
Hesseltinella vesiculosa n.a n.a Soil Upadhyay (1970)
Isomucor trufemiae n.a n.a Soil De Souza et al. (2012)
Lichtheimia brasiliensis n.a n.a Herbivore dung; soil Santiago et al. (2014), Souza et al. (2017)
L. corymbifera Meju Lee et al. (1993) Soil Santiago & Souza-Motta (2008)
L. hyalospora Meju; Nephila sp. Hong et al. (2012), Nguyen et al. (2023) Soil Lima et al. (2016a)
L. koreana Timomenus komarovi; Theuronema hilgendorfi hilgendorfi; Nephila sp. Nguyen et al. (2023) n.a n.a
L. piauiensis n.a n.a Oryctolagus cuniculus dung Cruz et al. (2024)
L. ornata Meju; Scolopendra morsitans; Theuronema hilgendorfi hilgendorfi Hong et al. (2012), Nguyen et al. (2023) n.a
L. ramosa Meju; nuruk; pregnant cow; Theuronema hilgendorfi hilgendorfi Yang et al. (2011), Hong et al. (2012), Lee et al. (2020), Nguyen et al. (2023) Herbivore dung; soil Souza et al. (2017), Lima et al. (2020c)
Mucor abundans Freshwater Nguyen et al. (2020) n.a n.a
M. albicolonia Nephila sp.; soil This study n.a n.a
M. aligarensis Freshwater Nguyen et al. (2020) n.a n.a
M. amphibiorum n.a n.a Soil Schoenlein-Crusius et al. (2006)
M. ardhlaengiktus Amphibian feces Nguyen et al. (2019) Soil Santos et al. (2023b)
M. aurantiacus Protaetia orientalis This study n.a n.a
M. bacilliformis n.a n.a Soil Upadhyay (1967)
M. bainieri n.a n.a Soil De Souza et al. (2008)
M. caatingaensis n.a n.a Soil Li et al. (2016)
M. cheongyangensis Lycorma delicatula Nguyen & Lee (2020) n.a n.a
M. circinelloides Meju Lee et al. (1993) Herbivore dung; leaf litter; soil; water Schoenlein-Crusius et al. (2006); Santiago et al. (2011)
M. cryophilus Cherry seeds This study n.a n.a
M. durus n.a n.a Herbivore dung; soil Trufem (1981a); Santiago et al. (2011)
M. fluvii Freshwater Wanasinghe et al. (2018) n.a n.a
M. fragilis Gut of soldier fly larvae Nguyen et al. (2016) n.a
M. fuscus n.a n.a Soil Trufem (1981c)
M. genevensis n.a n.a Soil Schoenlein-Crusius et al. (2006)
M. gigasporus Soil Nguyen et al. (2019) n.a n.a
M. glutinatus Soil on the skin surface of toad (Bufo gargarizans) This study n.a n.a
M. griseocyanus Meju Lee et al. (1993) Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Souza et al. (2017)
M. grylli Gryllus sp. Nguyen & Lee (2022) n.a n.a
M. guilliermondii n.a n.a Herbivore dung Santiago et al. (2011)
M. harpali Harpalus sp. Boonmee et al. (2021) n.a n.a
M. heterogamus Freshwater; sediment Nguyen et al. (2020) Soil Schoenlein-Crusius et al. (2006)
M. hiemalis Meju Lee et al. (1993) Herbivore dung; leaf litter; soil; water Souza et al. (2017); Lima et al. (2020c)
M. hyangburmii Gryllus sp. Nguyen & Lee (2022) n.a n.a
M. inaequisporus Persimmon (Diospyros kaki) Lee & Jung (2020) Fruit; soil Santiago et al. (2013b)
M. indicus n.a n.a Herbivore dung; soil Souza et al. (2017), Lima et al. (2020c)
M. irregularis Gut of soldier fly larvae Nguyen et al. (2016) Soil Lima et al. (2018b)
M. jansseni Meju Lee et al. (1993) Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006); Souza et al. (2017)
M. japonicus n.a n.a Leaf litter Schoenlein-Crusius & Milanez (1998)
M. koreanus Tangerine fruit Li et al. (2016) n.a n.a
M. kunryangriensis Gryllus sp. Nguyen & Lee (2022) n.a n.a
M. lusitanicus Meju Hong et al. (2012) Herbivore dung; soil Souza et al. (2017); Lima et al. (2020c)
M. luteus n.a n.a Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Souza et al. (2017)
M. megalocarpus n.a n.a Soil Schoenlein-Crusius & Milanez (1998)
M. merdicola n.a n.a Herbivore dung Li et al. (2016)
M. minutus Soil Nguyen et al. (2017) n.a n.a
M. moelleri Freshwater; sediment Nguyen et al. (2020) Leaf litter; soil Schoenlein-Crusius et al. (2006)
M. mousanensis n.a n.a Herbivore dung; soil Santiago et al. (2011)
M. mucedo Meju Hong et al. (2012) Herbivore dung; soil Santiago et al. (2011)
M. nidicola Air; seed of pumkin Ahn et al. (2016), Nguyen et al. (2016) Soil Lima et al. (2020b)
M. orantomantidis Feces of praying mantis Phookamsak et al. (2019) n.a n.a
M. orodatus n.a n.a Soil Trufem (1981c)
M. paraorantomantidis Isyndus obscurus This study n.a n.a
M. pernambucoensis n.a n.a Soil Lima et al. (2018a)
M. piriformis Sweet persimmon Kwon & Park (2004) Soil Trufem (1981c)
M. plumbeus Nuruk Kim et al. (2009) Soil Trufem (1984)
M. prayagensis n.a n.a Soil Schoenlein-Crusius et al. (2006)
M. racemosus Meju Lee et al. (1993), Hong et al. (2012) Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Santiago et al. (2011)
M. ramosissimus Freshwater Nguyen & Lee (2018) Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006)
M. saturninus Isyndus obscurus This study Leaf litter; soil Schoenlein-Crusius et al. (2006)
M. septatiphorus n.a n.a Soil Souza et al. (2022)
M. silvaticus n.a n.a Leaf litter; soil Schoenlein-Crusius et al. (2006)
M. simplex n.a. n.a Herbivore dung; soil Trufem (1981a), Lima et al. (2017)
M. sinensis n.a n.a Leaf litter; soil Schoenlein-Crusius et al. (2006)
M. souzae n.a n.a Soil Crous et al. (2018)
M. stercorarius Rat dung Tibpromma et al. 2017 n.a n.a
M. subtilissimus n.a n.a Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Santiago et al. (2011)
M. suhagiensis n.a n.a Soil Trufem (1981c)
M. timomeni Timomenus komarovi; soil This study n.a n.a
M. variicolumellatus Storage root Paul et al. (2021) Soil Souza et al. (2020)
M. variosporus n.a n.a Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Souza et al. (2017)
M. zonatus n.a n.a Leaf litter; soil Schoenlein-Crusius et al. (2006)
Mycotypha microspora n.a n.a Soil Santiago & Maia (2010)
Neofennellomyces jeongsukae Euborellia sp.; Porcellio scaber This study n.a n.a
Parasitella parasitica n.a n.a Soil Schoenlein-Crusius et al. (2006)
Phycomyces blakesleeanus Meju Hong et al. (2012) Herbivore dung Trufem & Viriato (1985)
Pilaira anomala n.a n.a Herbivore dung Santiago et al. (2011)
Pilobolus crystallinus Feces of water deer Lee et al. (2018) Herbivore dung Santiago et al. (2011)
P. kleinii n.a n.a Herbivore dung Santiago et al. (2011)
P. lentiger n.a n.a Herbivore dung Santiago et al. (2011)
P. longipes n.a n.a Herbivore dung Santiago et al. (2011)
P. minutus n.a n.a Herbivore dung Santiago et al. (2011)
P. oedipus n.a n.a Herbivore dung Viriato & Trufem (1985)
P. roridus n.a n.a Herbivore dung Santiagoet al. (2011)
P. umbonatus n.a n.a Herbivore dung Santiago et al. (2011)
Poitrasia circinans n.a n.a Leaf litter; soil Upadhyay (1970), Schoenlein-Crusius et al. (2006)
Rhizomucor miehei n.a n.a Soil Antonelli et al. (2021)
Rh. pusillus Meju; air Hong et al. (2012), Song et al. (2022) Humans; soil Schoenlein-Crusius et al. (2006)
Rhizopus arrhizus Meju Lee et al. (1993) Leaf litter; soil; herbivore dung Souza et al. (2017), Lima et al. (2020c)
R. homothallicus n.a n.a n.a n.a
R. koreanus Persimmon fruit Li et al. (2016) n.a n.a
R. microsporus Pulmonary Kim et al. (2014) Leaf litter; soil Lima et al. (2020c)
R. stolonifer Meju Lee et al. (1993), Hong et al. (2012) Herbivore dung; leaf litter; soil Schoenlein-Crusius et al. (2006), Souza et al. (2017), Lima et al. (2020b)
Saksenaea longicolla Soil sediment Nam et al. (2021) n.a n.a
Sa. oblongispora n.a n.a Soil n.a
Syncephalastrum irregulare n.a n.a Soil Santiago et al. (2024)
S. monosporum Seed of Cucurbita pepo Duong et al. (2016) n.a n.a
S. racemosum Nuruk; Meju Yang et al. (2011), Hong et al. (2012) Herbivore dung; soil Souza et al. (2017), Lima et al. (2020c)
Syzygites megalocarpus Tricholoma matsutake Ka et al. (1999) n.a n.a
Thamnostylum piriforme n.a n.a Soil n.a
Utharomyces epallocaulus n.a n.a Herbivore dung Alves et al. (2020)

DISCUSSION

The overarching goal of the current study was to access Mucorales fungi from two countries, South Korea and Brazil. New isolates of the ancient fungal lineage Mucorales were obtained from habitats/ substrates in both Asia and South America, where 12 emerged from the South Korean samples and three from Brazil. The former included the monotypic genus Neofennellomyces. In this study, we also list the species of Mucorales reported in Brazil and in South Korea up to June 2024 (including those isolated in the present study). Considering that about 360 Mucorales species are known (Wijayawardene et al. 2022, Zhao et al. 2023), it is possible that this accessed richness in both countries is underestimated.

Specific inventories of Mucorales fungi in Brazil are scarce, and what is mostly known about these fungi results from specific approaches carried out in only two of the 27 Brazilian states, namely Pernambuco and São Paulo. Of the six biomes found in Brazil, there are records of these fungi in only three: Atlantic Forest (most records), Cerrado (Brazilian savannah), and Caatinga (xeric shrubland and thorn forest), and even so, these studies covered just a few ecosystems belonging to these biomes. For the Amazon Forest, which comprises the greatest biodiversity on the planet, records of these fungi are practically unknown, which is also the case of Pantanal (wetland) and Pampas (grassland) biomes (Flora e Funga do Brasil 2024) that remain unexplored with regards to Mucorales. As for South Korea, Mucorales species were recorded mostly in Cheongyang, Chungnam Province.

Interestingly, after comparing the mucoralean communities between South Korea and Brazil, we found a similarity of only 37.19 % according to Sørensen (1978). Although this is a shallow comparison, given the complexity of tropical and temperate ecosystems, the time frame differences for the species isolation, as well as variations among isolation methodologies, this data draws our attention, as a greater similarity of these communities between the two countries was expected, assuming that some mucoraceous fungi are cosmopolitan (Cannon & Kirk 2007, Voigt 2012). Certainly, the edaphic, and floristic differences between both countries play a key role in structuring fungal communities in each country (Tedersoo et al. 2014). This difference is reinforced by the fact that of 68 species of Mucorales fungi described for the first time since 2012 from Brazil and South Korea (Brazil = 32, South Korea = 36), only Backusella obliqua and M. variicolumellatus were common to both countries, thus with little overlap, indicating that high rates of endemism of mucoralean fungi might be expected in these countries. It is remarkable that South Korea exhibits high richness of species of these fungi, even though it is a small country with a temperate climate. Although high fungal richness and diversity are expected in tropical countries (Hawksworth & Lücking 2017), Tedersoo et al. (2014) showed that the diversity of species in the phylum Mucoromycota increases towards the poles, without declining in the boreal forest- and tundra biomes.

Most Mucorales fungi are common soil saprotrophs, therefore it is not surprising that most newly described species have been isolated from this substrate (Voigt et al. 2021). Species of this order have also been isolated from dung, plant substrates, nuts, water, food products, and humans, which demonstrates that these fungi are ecologically highly diverse (Hoffmann et al. 2013, Nguyen et al. 2023). As for Brazil, most of the Mucorales species recorded to date have been isolated from soil, including the new species A. tarda, A. variiprojecta, and A. variispora described herein, but there are reports of species of this order on dung, plants, leaf litter, stored grains and fruits in this country (Schoenlein-Crusius 2006, Santiago et al. 2011, Melo et al. 2020). Surprisingly, new species of this group have been frequently isolated from animals in South Korea, including toads and different invertebrate species (Nguyen & Lee 2020, 2022, Boonmee et al. 2021, Nguyen et al. 2021, 2023). The new Nefennellomyces described herein was isolated from Euborellia sp. (earwig), and Porcellio scaber (woodlouse), and the current study is the first report of new taxa from Euborellia sp. (earwig). Besides N. jeongsukae, we have also isolated M. albicolonia, M. aurantiacus, M. paraorantomantidis, M. timomeni from invertebrates. However, because these fungi were isolated from the surface and not the intestinal tract, it is likely that these animals are vectors and not hosts of these fungi (Cordeiro et al. 2023). However, the possibility of these fungi being facultative parasites of insects cannot be ruled out without further studies, as there are reports (although rare) in the literature of entomopathogenic mucoraceous fungi as well as arachnid pathogens. For example, Evans & Samson (1977) have isolated Sporodiniella umbellata from insect hosts in cocoa farm in Mexico (this fungus was firstly described from a probable member of the Microlepidoptera by Boedijn in 1959, in Indonesia), thus suggesting the pathogenic potential of this fungi to membracids, and Bibbs et al. (2013) reported a high mortality rate of the brown window spider Latrodectus geometricus caused by a strain of M. fragilis in the USA. The entomopathogenic potential of a strain of M. hiemalis on Bradysia odoriphag larvae in China was reported by Zhu et al. (2022). However, to the best of our knowledge, the infection mechanism of these fungi still remains unknown. An interesting path that could help us understand the ecological behaviour of these fungi would be to profile the nutritional capacity of as many species as possible, as was done by Pawłowska et al. (2019), who showed that species of this group fungi are metabolically much more versatile than previously thought, as they are able to use dozens of complex organic compounds as food, including cellulose, apple pectin and oat xylan. In this work we are also reporting four new species of Absidia from freshwater environments, namely A. fluvii, A. kunryangriensis, A. cheongyangensis, A. pararepens. However, because Mucorales are primarily terrestrial fungi we believe in this case spores of Absidia were occasionally collected while dispersed through water.

The ITS region has been frequently used for diagnostics and molecular phylogenetic identifications in fungi (White et al. 1990) and it was proposed to be a universal DNA barcode marker for fungi (Schoch et al. 2012). The ITS became so popular that it was proposed as a new tool for species differentiation and delineation of systematic relationships in prokaryotes (Man et al. 2010). However, the success rate of species discrimination efficiency is hampered by intraspecies sequence variability which was reported for bacteria, myxomycetes, plants, animals and fungi (Cruz et al. 2006, Winsett & Stephenson 2008, Man et al. 2010, Kiss 2012, Song et al. 2012, Mishra et al. 2021, Bradshaw et al. 2023). Intraspecies sequences variability up to 25.85 % was reported compared to mean interspecies sequence variability of 35.94 % in Campylobacter (Man et al. 2010). Genome mining of 2 414 fungal species revealed that 641 species (~25 %) contained multiple ITS copies, of which 419 (~65 %) contained variation among copies revealing that intragenomic variation is common in fungi (Bradshaw et al. 2023). This heterogeneity among individual ITS copies skews the description of fungal species and evaluations of environmental DNA, especially when making diversity estimates (Bradshaw et al. 2023). Therefore, the use of secondary barcode markers provides high numbers of informative characteristics and the phylogeny based on it appears to be better resolved, especially when single copy genes were used like mcm7 (MS456) and tsr1 (MS277) as proposed by Schmitt et al. (2009) to resolve a wide range of Pezizomycotina (Ascomycota).

In Mucorales, Wagner et al. (2020) separated species from the Mucor circinelloides complex based on the ML analysis of a combination of five multi- and single copy markers (ITS, rpb1, tsr1, mcm7, and cfs). Moreover, a limited number of protein-coding gene sequences are available for Mucorales on GenBank, and therefore phylogenetic analyses were still performed for one or two loci (ITS and LSU rDNA). However, Urquhart et al. (2021) have shown that argA is a reliable marker for Backusella and Cordeiro et al. (2023) have stablished rpb1 as phylogenetic marker for this genus. In our study, single gene sequences, such as ITS rDNA or LSU rDNA did not aid in proper identification of B. obliqua from South Korea. Therefore, we also added these additional markers for a number of previously described species including B. chlamydospora (argA), B. koreana (argA), B. obliqua (argA, rpb1), and B. thermophila (argA) (Nguyen et al. 2021, Lima et al. 2022).

Current fungal diversity estimates range between 2 and 3 million species, with a “best guess” at 2.5 million, of which 155 000 species of fungi have been described, suggesting that between 92.5 % and 95 % of all fungal species remain unknown (Niskanen et al. 2023). New isolations thus remain important and motivate us to describe the missing diversity. At present many new species including Mucorales fungi are only known from a limited number of isolates or are described based on a single strain (Urquhart et al. 2021, Cordeiro at al. 2023), probably because basal fungi are quite rare compared to Ascomycota and Basidiomycota species. Although in this work some new species have been described based on a single strain, our phylogeny clearly separated them from other species, thus justifying their status as new species. However, because some of our analyses involved single isolates of some species, the phenotypic variation within these species could not be accessed.

The newly published records are also important for species documentation to update sequence data, taxonomic nomenclature, examine potential synonyms and revise species concepts (Hyde et al. 2020). They also contribute to our knowledge on the distribution of fungi. The new records will also benefit future studies such as exploration of fungi for potential industrial resources, new fungal and fungus-derived products. In this study, A. pararepens, B. obliqua, C. intermedia, and M. saturninus were recorded for the first time from South Korea as well as the second report of B. obliqua, C. intermedia, and M. saturninus and the third report of A. pararepens worldwide, contributing to our knowledge of the geographic distribution of a mucoralean fungi. Backusella obliqua, and C. intermedia were isolated from soil in Brazil and India in previous studies (Deshpande & Mantri 1966, Lima et al. 2022), and from Apis sp., and Pholcus phalangioides in the present study. Absidia pararepens was isolated from indoor environments, and swabs from the USA (Crous et al. 2020), and soil in Brazil (Cordeiro et al. 2022), and from rainwater, and Scolopendra sp. in the present study.

We believe that studies of Mucorales phylogeny make fundamental contributions to our understanding of their ecology and evolutionary biology (both past and present) as well as global climate change and future climate scenarios (Hug et al. 2016, Cavicchioli et al. 2019, Timmis & Hallsworth 2019, Hallsworth et al. 2023). Studies of ancient microbes preserved in brines, water, or ice, and those preserved in amber or as mineralized fossils can give further insights into ecology, evolutionary trajectories, and the biophysical tolerance of microbes (Krings et al. 2013, Hallsworth 2022, Schreder-Gomes et al. 2022). Mucorales such as those isolated from water, soil, invertebrates, and seeds during the current study can potentially be preserved via such studies. Furthermore, although Mucorales have been isolated from permafrost (Hu et al. 2014), more work is needed on the forensic evidence of the ecologies of ancient Mucorales fungi.

Concluding remarks

In this study, we describe one new genus, 15 new species and four new records of Mucorales for South Korea and Brazil based on phylogenetic, physiological, and morphological comparisons with allied taxa, thus contributing to the knowledge on the taxonomy and distribution of these fungal taxa. The low similarity of Mucorales communities between Brazil and South Korea, as well as the fact that little overlap was observed for new species discovered in the last years in both countries, and even in other countries, reveals that there are high rates of endemism of these fungal taxa in both countries. Because the mycodiversity is vast and the number of taxonomists worldwide is low, more efforts to not only explore but also describe unknown Mucorales species from different countries are needed before they go extinct.

Climate change is a pervasive and growing global threat to biodiversity and ecosystems. It is possible, therefore, that some microbes, including the ancient Mucorales fungi that require specific niches, may also disappear in the coming decades. To document their existence, preserve them in culture collections, keeping them available for potential biotechnological applications is therefore useful.

ACKNOWLEDGEMENTS

This study was supported by the Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (2022R1I1A3068645); the Ministry of Science and ICT (2022M3H9A1082984); the project on the Survey and Discovery of Indigenous Fungal Species of Korea funded by the National Institute of Biological Resources (NIBR) of the Ministry of Environment (MOE); and by the project on the Discovery of Fungi from Freshwater funded by the Nakdonggang National Institute of Biological Resources (NNIBR) of the Ministry of Environment (MOE), Korea. A.L. Santiago thanks the Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq) for the research grant and the Fundação de Amparo à Ciência e Tecnologia do Estado de Pernambuco - project ‘Morphological and molecular approaches to understanding the communities of zygosporic fungi in the Atlantic Forest of Pernambuco and Paraíba (APQ-1346-2.12/22). The authors are grateful to Prof. A. Idnurm (University of Melbourne, Victoria, Australia) and Dr A.S. Urquhart (University of Uppsala, Uppsala, Sweden) for providing comments on a draft version of this paper.

DECLARATION ON CONFLICT OF INTEREST

The authors declare that there is no conflict of interest.

Supplementary Material: https://studiesinmycology.org/

Table S1.

Taxa, strains, locations, and GenBank accession numbers of the strains used in this study.

Table S2.

Maximum growth temperature of strains used in this study.

Fig. S1.

Map showing the sample collection sites (left to right and top to bottom: South Korea and Brazil). Yellow and red cursor represent the main collection areas from South Korea (S1, S2, S3) and Brazil (B1), respectively.

Fig. S2.

Phylogenetic relationship of Neofennellomyces with its related species based the ITS sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages. Bootstrap values ≥ 70 % are shown. Fennellomyces heterothallicus CBS 290.86, F. linderi CBS 158.54, Thamnostylum nigricans CBS 690.76, T. piriforme CBS 316.66, and T. repens CBS 692.76 were used as the outgroups. Ex-type and ex-neotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.

Fig. S3.

Phylogenetic relationship of Neofennellomyces with its related species based the LSU sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages. Bootstrap values ≥ 70 % are shown. Fennellomyces heterothallicus CBS 292.86, F. heterothallicus CBS 290.86, Thamnostylum nigricans CBS 690.76, and T. piriforme CBS 316.66 were used as the outgroups. Ex-type and exneotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.

Fig. S4.

Phylogenetic relationship of Neofennellomyces with its related species based the mcm7 sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages. Bootstrap values ≥ 70 % are shown. Fennellomyces linderi CBS 158.54, Thamnostylum repens CBS 692.76, and Zychaea mexicana CBS 441.76 were used as the outgroups. Ex-type and ex-neotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.

REFERENCES

  1. Abe A, Sone T, Sujaya IN. et al. (2003). rDNA ITS sequence of Rhizopus oryzae: its application to classification and identification of lactic acid producers. Bioscience, Biotechnology, and Biochemistry 67: 1725–1731. [DOI] [PubMed] [Google Scholar]
  2. Ahn GR, Kwon HW, Ko HK. et al. (2016). Unrecorded fungal species isolated from greenhouses used for shiitake cultivation in Korea. The Korean Journal of Mycology 44: 8–15. [Google Scholar]
  3. Alastruey-Izquierdo A, Hoffmann K, de Hoog GS. et al. (2010). Species recognition and clinical relevance of the zygomycetous genus Lichtheimia (syn. Absidia pro parte, Mycocladus). Journal of Clinical Microbiology 48: 2154–2170. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Alfenas RF, Bonaldo SM, Fernandes RAS. et al. (2018). First report of Choanephora cucurbitarum on Crotalaria spectabilis: a highly aggressive pathogen causing a flower and stem blight in Brazil. Plant Disease 102: 1456–1457. [Google Scholar]
  5. Alves ALM, Souza CAF, Oliveira RJV. et al. (2017). Cunninghamella clavata from Brazil: a new record for the western hemisphere. Mycotaxon 132: 381–389. [Google Scholar]
  6. Alves MH, Cruz MA, Santiago ALCMA. (2020). The first record of the coprophilous fungi Utharomyces epallocaulus Boedijn ex P.M. Kirk & Benny (Mucoromycotina, Mucorales, Pilobolaceae) in Brazil. Check List 16: 737–741. [Google Scholar]
  7. Antonelli MA, Gaglioti AL, Silva PR. et al. (2021). Thermophilic fungi in Araucaria Forest, Atlantic Forest Biome, Brazil. Anais da Academia Brasileira de Ciências 93 (Suppl. 4): e20210714. [DOI] [PubMed] [Google Scholar]
  8. Ariyawansa HA, Hyde KD, Jayasiri SC. et al. (2015). Fungal diversity notes 111–252 taxonomic and phylogenetic contributions to fungal taxa. Fungal Diversity 75: 27–274. [Google Scholar]
  9. Babu AG, Kim SW, Adhikari M. et al. (2015). A new record of Gongronella butleri isolated in Korea. Mycobiology 43: 166–169. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Bai F, Yao S, Cai C. et al. (2021). Mucor rongii sp. nov., a new cold-tolerant species from China. Current Microbiology 78: 2464–2469. [DOI] [PubMed] [Google Scholar]
  11. Benjamin RK, Mehrotra BS. (1963). Obligate azygospore formation in two species of Mucor (Mucorales). Aliso 5: 235–245. [Google Scholar]
  12. Benny GL. (2008). The methods used by Dr. R.K. Benjamin, and other mycologists, to isolate Zygomycetes. Aliso 26: 37–61. [Google Scholar]
  13. Benny GL, Humber RA, Voigt K. (2014). Zygomycetous Fungi: Phylum Entomophthoromycota and subphyla Kickxellomycotina, Mortierellomycotina, Mucoromycotina, and Zoopagomycotina. In: The Mycota VII Part A. Systematics and Evolution (McLaughlin DJ, Spatafora JW, eds). Springer, Heidelberg, Germany: 209–250. [Google Scholar]
  14. Benny GL, Smith M, Tretter ED. et al. (2016). Challenges and future perspectives in the systematics of Kickxellomycotina, Mortierellomycotina, Mucoromycotina, and Zoopagomycotina. In: Biology of Microfungi (Li DW, ed). Springer, Cham, Switzerland: 65–126. [Google Scholar]
  15. Bibbs CS, Vitoreli AM, Benny G. et al. (2013). Susceptibility of Latrodectus geometricus (Araneae: Theridiidae) to a Mucor strain discovered in North Central Florida, USA. Florida Entomologist 96: 1052–1061. [Google Scholar]
  16. Boedijn KB. (1959). Notes on the Mucorales of Indonesia. Sydowia 12: 321–362. [Google Scholar]
  17. Boonmee S, Wanasinghe DN, Calabon MS. et al. (2021). Fungal diversity notes 1387–1511: taxonomic and phylogenetic contributions on genera and species of fungal taxa. Fungal Diversity 111: 1–335. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Bradshaw MJ, Aime MC, Rokas A. et al. (2023). Extensive intragenomic variation in the internal transcribed spacer region of fungi. iScience 26: 107317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Bunyard BA, Nicholson MS, Royse DJ. (1994). A systematic assessment of Morchella using RFLP analysis of the 28S rRNA gene. Mycologia 86: 762–772. [Google Scholar]
  20. Cannon PF, Kirk PM. (2007). Fungal families of the world. CAB International, Wallingford. [Google Scholar]
  21. Capella-Gutiérrez S, Silla-Martínez JM, Gabaldón T. (2009). TrimAl: A tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25: 1972–1973. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cavicchioli R, Ripple WJ, Timmis KN. et al. (2019). Scientists’ warning to humanity: microorganisms and climate change. Nature Reviews Microbiology 17: 569–586. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chen GQ, Zheng RY. (1986). A new species of Mucor with giant spores. Acta Mycologica Sinica 1: 56–60. [Google Scholar]
  24. Choi I-Y, Kim JH, Park JH. et al. 2016. First report of Choanephora flower blight caused by Choanephora cucurbitarum on Althaea officinalis in Korea. Plant Disease 100: 1953. [Google Scholar]
  25. Cordeiro TRL, Nguyen TTT, Lima DX. et al. (2020). Two new species of the industrially relevant genus Absidia (Mucorales) from soil of the Brazilian Atlantic Forest. Acta Botanica Brasilica 34: 549–558. [Google Scholar]
  26. Cordeiro TRL, Silva SBG, Cruz MO. et al. (2022). Absidia pararepens Jurjević, M. Kolařík & Hubka (Mucorales, Mucoromycota) was isolated for the first time in South America. Nova Hedwigia 114: 375–387. [Google Scholar]
  27. Cordeiro TRL, Walther G, Lee HB. et al. (2023). A polyphasic approach to the taxonomy of Backusella reveals two new species. Mycological Progress 22: 16. [Google Scholar]
  28. Crous PW, Carnegie AJ, Wingfield MJ. et al. (2019). Fungal Planet description sheets: 868–950. Persoonia 42: 291–473. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Crous PW, Cowan DA, Maggs-Kölling G. et al. (2020). Fungal Planet descriptions sheets: 1112–1181. Persoonia 45: 251–409. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Crous PW, Wingfield MJ, Burgess TI. et al. (2018). Fungal Planet description sheets: 716–784. Persoonia 40: 240–393. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Cruz DO, Jorge DMM, Pereira JOP. et al. (2006). Intraspecific variation in the first internal transcribed spacer (ITS1) of the nuclear ribosomal DNA in Melipona subnitida (Hymenoptera, Apidae), an endemic stingless bee from northeastern Brazil. Apidologie 37: 376–386. [Google Scholar]
  32. Cruz MO, Lee HB, Cordeiro TRL. et al. (2024). Description of a novel coprophilous Lichtheimia (Mucoromycotina, Mucorales) species with notes on Lichtheimia species and an identification key for the genus. Mycological Progress 23: 2. [Google Scholar]
  33. Dade HA. (1937). New Gold Coast Fungi 1. Transactions of the British Mycological Society 21: 16–28. [Google Scholar]
  34. Darriba D, Taboada GL, Doallo R. et al. (2012). jModelTest 2: More models, new heuristics and parallel computing. Nature Methods 9: 772. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. De Hoog GS, Gerrits van den Ende AH. (1998). Molecular diagnostic of clinical strain of filamentous Basidiomycetes. Mycoses 41: 183–189. [DOI] [PubMed] [Google Scholar]
  36. De Souza JI, Marano AV, Pires-Zottarelli CLA. et al. (2014). A new species of Backusella (Mucorales) from a Cerrado reserve in Southeast Brazil. Mycological Progress 13: 975–980. [Google Scholar]
  37. De Souza JI, Pires-Zottarelli CLA, Santos JF. et al. (2012). Isomucor (Mucoromycotina): a new genus from a Cerrado reserve in state of São Paulo, Brazil. Mycologia 104: 232–241. [DOI] [PubMed] [Google Scholar]
  38. De Souza JI, Schoenlein-Crusius IH, Oliveira LHS. (2008). Selected species of Mucorales from soil contaminated with toxic metals in São Paulo State, Brazil. Mycotaxon 106: 273–288. [Google Scholar]
  39. Deshmukh SK, Dufossé L, Chhipa H. et al. (2022). Fungal endophytes: a potential source of antibacterial compounds. Journal of Fungi 8: 164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Deshpande KB, Mantri JM. (1966). A new species of Cunninghamella from India. Mycopathologia et Mycologia Applicata 28: 342–344. [Google Scholar]
  41. Dolatabadi S, Scherlach K, Figge M. et al. (2016). Food preparation with mucoralean fungi: A potential biosafety issue? Fungal Biology 120: 393–401. [DOI] [PubMed] [Google Scholar]
  42. Duong TT, Nguyen TTT, Lee HB. (2016). Phylogenetic status of an undiscovered Zygomycete species, Syncephalastrum monosporum, in Korea. The Korean Journal of Mycology 44: 371–376. [Google Scholar]
  43. Evans HC, Samson RA. (1977). Sporodiniella umbellata, an entomogenous fungus of the Mucorales from cocoa farms in Ecuador. Canadian Journal of Botany 55: 2981–2984. [Google Scholar]
  44. Flora e Funga do Brasil (2024). Fungos Mucorales. Domínios fitogeográficos: Cerrado. in: Flora e Funga do Brasil. Jardim Botânico do Rio de Janeiro. Available: http://floradobrasil.jbrj.gov.br/.
  45. Freitas LWS, Cruz MO, Nguyen TTT. et al. (2022). Absidia variicolumellata, sp. nov., a new mucoralean fungus isolated from Atlantic Forest in Bahia state (Brazil). Sydowia 75: 75–89. [Google Scholar]
  46. Freitas LWS, Oliveira RJV, Cordeiro TRL. et al. (2020). Gongronella pedratalhadensis, a new species of Mucorales (Mucoromycota) isolated from the Brazilian Atlantic Forest, with an identification key for the genus. Sydowia 73: 61–68. [Google Scholar]
  47. Glez-Peña D, Gómez-Blanco D, Reboiro-Jato M. et al. (2010). ALTER: Program–oriented format conversion of DNA and protein alignments. Nucleic Acids Research 38: 14–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Gomes MZR, Lewis RE, Kontoyiannis DP. (2011). Mucormycosis caused by unusual Mucoromycetes, non-Rhizopus, -Mucor, and -Lichtheimia species. Clinical Microbiology Reviews 24: 411–445. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Guindon S, Gascuel O. (2003). A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Systematic Biology 52: 696–704. [DOI] [PubMed] [Google Scholar]
  50. Guo J, Huang H, Liu D. et al. (2015). Isolation of Cunninghamella bigelovii sp. nov. CGMCC 8094 as a new endophytic oleaginous fungus from Salicornia bigelovii. Mycological Progress 14: 11. [Google Scholar]
  51. Hall TA. (1999). BioEdit, a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95–98. [Google Scholar]
  52. Hallur V, Prakash H, Sable M. et al. (2021). Cunninghamella arunalokei a new species of Cunninghamella from India causing disease in an immunocompetent individual. Journal of Fungi 7: 670. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Hallsworth JE. (2022). Water is a preservative of microbes. Microbial Biotechnology 15: 191–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Hallsworth JE, Udaondo Z, Pedrós-Alió C, et al. (2023). Scientific novelty beyond the experiment. Microbial Biotechnology 16: 1131–1173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Hawksworth DL, Lücking R. (2017). Fungal diversity revisited 2.2 to 3.8 million. Microbiology Spectrum 5: FUNK-0052-2016 [DOI] [PubMed] [Google Scholar]
  56. He GQ, Xiao L, Pan Z. et al. (2022). Case report: A rare case of pulmonary mucormycosis caused by Lichtheimia ramosa in pediatric acute lymphoblastic leukaemia and review of Lichtheimia infections in leucemia. Frontiers in Oncology 12: 949910. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Hoffmann K, Voigt K. (2009). Absidia parricida plays a dominant role in biotrophic fusion parasitism among mucoralean fungi (Zygomycetes): Lentamyces, a new genus for A. parricida and A. zychae. Plant Biology 4: 537–554. [DOI] [PubMed] [Google Scholar]
  58. Hoffmann K, Pawłowska J, Walther G. et al. (2013). The family structure of the Mucorales: a synoptic revision based on comprehensive multigene-genealogies. Persoonia 30: 57–76. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Hong S-B, Kim D-H, Lee M. et al. (2012). Zygomycota associated with traditional meju, a fermented soybean starting material for soy sauce and soybean paste. Journal of Microbiology 50: 386–393. [DOI] [PubMed] [Google Scholar]
  60. Houbraken J, Samson RA. (2011). Phylogeny of Penicillium and the segregation of Trichocomaceae into three families. Studies in Mycology 70: 1–51. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Hu W, Zhang Q, Li D. et al. (2014). Diversity and community structure of fungi through a permafrost core profile from the Qinghai-Tibet Plateau of China. Journal of Basic Microbiology 54: 1331–1341. [DOI] [PubMed] [Google Scholar]
  62. Hug LA, Baker BJ, Anantharaman K. et al. (2016). A new view of the tree of life. Nature Microbiology 1: 16048. [DOI] [PubMed] [Google Scholar]
  63. Hurdeal VG, Gareth Jones EB, Santiago ALCMA. et al. (2022). Expanding the diversity of mucoralean fungi from northern Thailand: novel Backusella species from soil. Phytotaxa 559: 275–284. [Google Scholar]
  64. Hurdeal VG, Gentekaki E, Hyde KD. et al. (2021). Novel Mucor species (Mucoromycetes, Mucoraceae) from northern Thailand. MycoKeys 84: 57–78. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Hyde KD, de Silva NI, Jeewon R. et al. (2020) AJOM new records and collections of fungi: 1–100. Asian Journal of Mycology 3: 22–294. [Google Scholar]
  66. Hyde KD, Hongsanan S, Jeewon R. et al. 2016. Fungal diversity notes 367–490: taxonomic and phylogenetic contributions to fungal taxa. Fungal Diversity 80: 1–270. [Google Scholar]
  67. Ka K-H, Park H, Kim H-J, et al. (1999). Syzygites megalocarpus (Mucorales): A necrotrophic mycoparasite of Tricholoma matsutake. The Korean Journal of Mycology 27: 345–8. [Google Scholar]
  68. Katoh K, Rozewicki J, Yamada KD. (2017). MAFFT online service: Multiple sequence alignment, interactive sequence choice and visualization. Briefings in Bioinformatics 20: 1160–1166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Kim MJ, Park PW, Ahn JY. et al. (2014). Fatal pulmonary mucormycosis caused by Rhizopus microsporus in a patient with diabetes. Annals of Laboratory Medicine 34: 76–9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Kiss L. (2012). Limits of nuclear ribosomal DNA internal transcribed spacer (ITS) sequences as species barcodes for Fungi. Proceedings of the National Academy of Sciences of the USA 109: E1811. [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Krings M, Taylor TN, Dotzler N. (2013). Fossil evidence of the zygomycetous fungi. Persoonia 30: 1–10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  72. Kumar S, Stecher G, Tamura K. (2016). MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Molecular Biology and Evolution 33: 1870–1874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Kwon JH, Jee HJ. (2005). Soft rot of eggplant (Solanum melongena) caused by Choanephora cucurbitarum in Korea. Mycobiology 33:163–165. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Kwon JH, Ahn GH, Park CS. (2004). Fruit soft rot of sweet persimmon caused by Mucor piriformis in Korea. Mycobiology 32: 98–101. [Google Scholar]
  75. Lampila LE, Wallen SE, Bullerman LB. (1985). A review of factors affecting biosynthesis of carotenoids by the order Mucorales. Mycopathologia 90: 65–80. [DOI] [PubMed] [Google Scholar]
  76. Lee K, Kim H, Sohn J. et al. (2020). Systemic mucormycosis caused by Lichtheimia ramosa in a pregnant cow. Veterinarni Medicina 65: 506–510. [Google Scholar]
  77. Lee S-S, Park K-H, Choi K-J, et al. (1993). Identification and isolation of zygomycetous fungi found on Maeju, a raw material of Korean traditional soy sources. The Korean Journal of Mycology 21: 172–187. [Google Scholar]
  78. Lee SH, Nguyen TTT, Lee HB. (2018). Isolation and characterization of two rare mucoralean species with specific habitats. Mycobiology 46: 205–214. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Lee SY, Jung HY. (2020). First report of persimmon (Diospyros kaki) fruit rot caused by Mucor inaequisporus in South Korea. Plant Disease 104: 2031. [Google Scholar]
  80. Leitão JDA, Cordeiro TRL, Nguyen TTT. et al. (2021). Absidia aguabelensis sp. nov.: A new mucoralean fungi isolated from a semiarid region in Brazil. Phytotaxa 516: 83–91. [Google Scholar]
  81. Li GJ, Hyde KD, Zhao RL. et al. (2016). Fungal diversity notes 253–366: taxonomic and phylogenetic contributions to fungal taxa. Fungal Diversity 78: 1–237. [Google Scholar]
  82. Lima CLF, Leitão JDA, Nguyen TTT. et al. (2022). Two new species of Backusella (Mucorales, Mucoromycota) from soil in an Upland Forest in Northeastern Brazil with an identification key of Backusella from the Americas. Journal of Fungi 8: 1038. [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Lima CLF, Lima DX, Cordeiro TRL. et al. (2021). Absidia bonitoensis (Mucorales, Mucoromycota), a new species isolated from the soil of an upland Atlantic Forest in Northeastern Brazil. Nova Hedwigia 112: 241–251. [Google Scholar]
  84. Lima CLF, Lima DX, Souza CAF. et al. (2018a). Description of Mucor pernambucoensis (Mucorales, Mucoromycota), a new species isolated from the Brazilian upland rainforest. Phytotaxa 350: 274–282. [Google Scholar]
  85. Lima DX, Cordeiro TR, de Souza CA. et al. (2020a). Morphological and molecular evidence for two new species of Absidia from Neotropic soil. Phytotaxa 446: 61–71. [Google Scholar]
  86. Lima DX, Cordeiro TRL, Lima CLF. et al. (2020b). A new occurrence of Mucor nidicola (Madden, Stchigel, Guarro, Sutton & Starks) (Mucorales, Mucoromycota) in the Upland Rainforest of the Brazilian Northeast and first report as a saprobe in soil. Check List 16: 163–167. [Google Scholar]
  87. Lima DX, Santiago ALCMA, Souza-Motta CM. (2016a). Diversity of Mucorales in natural and degraded semi-arid soils. Brazilian Journal of Botany 39: 1127–1133. [Google Scholar]
  88. Lima DX, Souza CAF, Oliveira RJV. et al. (2018b). Mucor irregularis, a first record for South America. Mycotaxon 33: 429–438. [Google Scholar]
  89. Lima DX, Souza-Motta CM, de Lima CLF. et al. (2020c). Communities of Mucorales (phylum Mucoromycota) in different ecosystems of the Atlantic Forest. Acta Botanica Brasilica 34: 796–806. [Google Scholar]
  90. Lima DX, Souza-Motta CM, Wagner L. et al. (2017). Circinella simplex—a misapplied name of Mucor circinatus sp. nov. Phytotaxa 329: 269–276. [Google Scholar]
  91. Lima DX, Voigt K, Souza CAF. et al. (2016b). Description of Backusella constricta sp. nov. (Mucorales, ex Zygomycota) from the Brazilian Atlantic Rainforest, including a key to species of Backusella. Phytotaxa 289: 59–68. [Google Scholar]
  92. Luo W, Gong Z, Li N, Zhao Y. et al. (2020). A negative regulator of carotenogenesis in Blakeslea trispora. Applied and Environmental Microbiology 86: 1–11. [DOI] [PMC free article] [PubMed] [Google Scholar]
  93. Mamali V, Koutserimpas C, Zarkotou O. et al. (2022). Isolated cerebral mucormycosis caused by Lichtheimia species in a polytrauma patient. Diagnostics 12: 358. [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Man SM, Kaakoush NO, Octavia S. et al. (2010). The internal transcribed spacer region, a new tool for use in species differentiation and delineation of systematic relationships within the Campylobacter genus. Applied and Environmental Microbiology 76: 3071–3081. [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Masclaux F, Guého E, de Hoog GS. et al. (1995). Phylogenetic relationships of human-pathogenic Cladosporium (Xylohypha) species inferred from partial LS rRNA sequences. Journal of Medical and Veterinary Mycology 33: 327–338. [DOI] [PubMed] [Google Scholar]
  96. Melo RFR, Gondim NHB, Santiago ALCMA, et al. (2020). Coprophilous fungi from Brazil: updated identification keys to all recorded species. Phytotaxa 436: 104–124. [Google Scholar]
  97. Mendes MAS, Urben AF. (2017). Fungos relatados em plantas no Brasil,Laboratório de Quarentena Vegetal. Brasília, DF: Embrapa RecursosGenéticos e Biotecnologia. [Google Scholar]
  98. Mirza JH, Khan SM, Begum S, Shagufta S. (1979). Mucorales of Pakistan. Faisalabad: Pakistan University of Agriculture. [Google Scholar]
  99. Mishra S, Sharma G, Das MK. et al. (2021). Intragenomic sequence variations in the second internal transcribed spacer (ITS2) ribosomal DNA of the malaria vector Anopheles stephensi. PLoS ONE 16: e0253173. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Nam B, Lee DJ, Choi YJ. (2021). High-temperature-tolerant fungus and Oomycetes in Korea, including Saksenaea longicolla sp. nov. Mycobiology 49: 476–490. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Nartey L, Pu Q, Zhu W. et al. (2022). Antagonistic and plant growth promotion effects of Mucor moelleri, a potential biocontrol agent. Microbiological Research 255: 126922. [DOI] [PubMed] [Google Scholar]
  102. Nguyen TTT, Lee HB. (2016). Isolation and characterization of Blakeslea trispora isolated from gut of grasshopper and soldier fly larva in Korea. The Korean Journal of Mycology 44: 355–359. [Google Scholar]
  103. Nguyen TTT, Lee HB. (2018). Isolation and characterization of three Zygomycetous fungi in Korea: Backusella circina, Circinella muscae, and Mucor ramosissimus. Mycobiology 46: 317–327. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Nguyen TTT, Lee HB. (2020). Mucor cheongyangensis, a new species isolated from the surface of Lycorma delicatula in Korea. Phytotaxa 446: 33–42. [Google Scholar]
  105. Nguyen TTT, Lee HB. (2022). Discovery of three new Mucor species associated with cricket insects in Korea. Journal of Fungi 8: 601. [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Nguyen TTT, Choi YJ, Lee HB. (2017). Isolation and characterization of three unrecorded Zygomycete fungi in Korea: Cunninghamella bertholletiae, Cunninghamella echinulata, and Cunninghamella elegans. Mycobiology 45: 318–326. [DOI] [PMC free article] [PubMed] [Google Scholar]
  107. Nguyen TTT, Duong TT, Lee HB. (2016). Characterization of two new records of mucoralean species isolated from gut of soldier fly larva in Korea. Mycobiology 44: 310–313. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Nguyen TTT, Hyde KD, Lee HB. (2019). Cunninghamella binariae, Mucor ardhlaengiktus, Mucor gigasporus and Umbelopsis changbaiensis, newly recorded species from amphibian faeces and soil in Korea. Phytotaxa 425: 19–34. [Google Scholar]
  109. Nguyen TTT, Jeon YJ, Mun HY. et al. (2019). Isolation and characterization of four unrecorded Mucor species in Korea. Mycobiology 48: 29–36. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Nguyen TTT, Jung HY, Lee YS. et al. (2017). Phylogenetic status of two undescribed Zygomycete species from Korea: Actinomucor elegans and Mucor minutus. Mycobiology 45: 344–352. [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Nguyen TTT, Lee SH, Bae S. et al. (2016). Characterization of two new records of Zygomycete species belonging to undiscovered taxa in Korea. Mycobiology 44: 29–37. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Nguyen TTT, Santiago ALCMA, Kirk PM. et al. (2023). Discovery of a new Lichtheimia (Lichtheimiaceae, Mucorales) from invertebrate niche and its phylogenetic status and physiological characteristics. Journal of Fungi 9: 317. [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Nguyen TTT, Voigt K, Santiago ALCMA. et al. (2021). Discovery of novel Backusella (Backusellaceae, Mucorales) isolated from invertebrates and toads in Cheongyang, Korea. Journal of Fungi 7: 513. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Niskanen T, Lücking R, Dahlberg A. et al. (2023). Pushing the frontiers of biodiversity research: Unveiling the global diversity, distribution, and conservation of fungi. Annual Review of Environment and Resources 48: 149–176. [Google Scholar]
  115. Nout MJR, Aidoo KE. (2010). Asian fungal fermented food. In: The Mycota, vol. 10: Industrial Applications (Hofrichter M, ed). Springer-Verlag, Berlin, Heidelberg: 29–58. [Google Scholar]
  116. O’Donnell K, Lutzoni FM, Ward TJ. et al. (2001). Evolutionary relationships among mucoralean fungi (Zygomycota): evidence for family polyphyly on a large scale. Mycologia 93: 286–297. [Google Scholar]
  117. Ogawa Y, Tokumasu S, Tubaki K. (2004). An original habitat of tempeh molds. Mycoscience 45: 271–276. [Google Scholar]
  118. Oliveira RJV, Bezerra JL, Lima TEF. et al. (2016). Phaeosphaeria nodulispora, a new endophytic coelomycete isolated from tropical palm (Cocos nucifera) in Brazil. Nova Hedwigia 103: 185–192. [Google Scholar]
  119. Park JH, Cho SE, Han KS. et al. (2014). First report of Choanephora blight caused by Choanephora infundibulifera on Hibiscus rosa-sinensis in Korea. Plant Disease 98: 1275. [DOI] [PubMed] [Google Scholar]
  120. Park JH, Cho SE, Park MJ. et al. (2015). First report of Choanephora cucurbitarum causing Choanephora blight on Phlox paniculata in Korea. Plant Disease 99: 1180. [DOI] [PubMed] [Google Scholar]
  121. Paul NC, Park S, Liu H. et al. (2021). Fungi associated with postharvest diseases of sweet potato storage roots and in vitro antagonistic assay of Trichoderma harzianum against the diseases. Journal of Fungi 7: 927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  122. Pawłowska J, Okrasińska A, Kisło K. et al. (2019). Carbon assimilation profiles of mucoralean fungi show their metabolic versatility. Scientific Reports 9: 11864. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Phookamsak R, Hyde KD, Jeewon R. et al. (2019). Fungal diversity notes 929–1035: taxonomic and phylogenetic contributions on genera and species of fungi. Fungal Diversity 95: 1–273. [Google Scholar]
  124. Rambaut A. (2009). FigTree, Version 1.3. 1. Computer Program Distributed by the Author. Available online: http://www.treebioedacuk/software/fgtree [Google Scholar]
  125. Rashmi M, Kushveer JS, Sarma VV. (2019). A worldwide list of endophytic fungi with notes on ecology and diversity. Mycosphere 10: 798–1079. [Google Scholar]
  126. Ribes JA, Vanover-Sams CL, Baker DJ. (2000). Zygomycetes in human disease. Clinical Microbiology Reviews 13: 236–301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Ronquist F, Teslenko M, van der Mark P. et al. (2012). MrBayes 3.2: Efficient Bayesian phylogenetic inference and model choice across a large model space. Systematic Biology 61: 539–542. [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Ru H, Hong SK, Son KA. et al. (2022). First report of Choanephora rot on Lettuce (Lactuca sativa L.) caused by Choanephora cucurbitarum in Korea. Plant Disease 107: 1217. [DOI] [PubMed] [Google Scholar]
  129. Santiago ALCMA, Cavalcanti MAQ. (2008). Gilbertella persicaria (Mucorales): a new record from Brazil. Mycotaxon 102: 333–337. [Google Scholar]
  130. Santiago ALCMA, Maia LC. (2010). Two new records of Mucorales from the Brazilian semi-arid region. Mycotaxon 114: 171–77. [Google Scholar]
  131. Santiago ALCMA, Souza-Motta CM. (2008). Isolation of Mucorales from processed maize (Zea mays L.) and screening for protease activity. Brazilian Journal of Microbiology 39: 698–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Santiago ALCMA, Hoffmann K, Lima DX. et al. (2013a). A new species of Lichtheimia (Mucoromycotina, Mucorales) isolated from Brazilian soil. Mycological Progress 13: 343–352. [Google Scholar]
  133. Santiago ALCMA, Hoffmann K, Lima DX. et al. (2014). A new species of Lichtheimia (Mucoromycotina, Mucorales) isolated from Brazilian soil. Mycological Progress 13: 343–352. [Google Scholar]
  134. Santiago ALCMA, Lima DX, Freire KTLS. et al. (2023). Syncephalastrum irregulare (Mucorales, Mucoromycota) – a new endophytic species from Brazil. Sydowia 76: 1–9. [Google Scholar]
  135. Santiago ALCMA, Rodrigues A, Canedo EM. et al. (2013b). Taxonomic studies on Mucor inaequisporus, isolated for the first time in South America. Mycotaxon 124: 219–229. [Google Scholar]
  136. Santiago ALCMDA, Santos PJP, Maia LC. (2013c). Mucorales from the semiarid of Pernambuco, Brazil. Brazilian Journal of Microbiology 44: 1678–4405. [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Santiago ALCMA, Trufem SBF, Malosso E. et al. (2011). Zygomycetes from herbivore dung in the Ecological Reserve of Dois Irmãos, Northeast Brazil. Brazilian Journal of Microbiology 42: 89–95. [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Santos FRS, Cordeiro TRL, Lima CLF. et al. (2023a). Discovery of Backusella paraconstricta sp. nov. (Mucorales, Mucoromycota) in an upland forest in northeastern Brazil with an identification key for Backusella from the Americas. Acta Botanica Brasilica 37: e20230061. [Google Scholar]
  139. Santos FRS, Cruz MO, Lima CLF. et al. (2023b). Discovery of the rare Mucor ardhlaengiktus (Mucorales, Mucoromycota) in South America. Nova Hedwigia 116: 77–88. [Google Scholar]
  140. Schipper MAA. (1975). On Mucor mucedo, Mucor flavus and related species. Studies in Mycology 10: 33. [Google Scholar]
  141. Schipper MAA. (1978). On certain species of Mucor with a key to all accepted species. Studies in Mycology 17: 1–52. [Google Scholar]
  142. Schipper MAA. (1984). A revision of the genus Rhizopus I. The Rhizopus stolonifer-group and Rhizopus oryzae. Studies in Mycology 25: 1–19. [Google Scholar]
  143. Schmitt I, Crespo A, Divakar PK. et al. (2009). New primers for promising single-copy genes in fungal phylogenetics and systematics. Persoonia 23: 35–40. [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. Schoch C, Seifert K, Huhndorf S. et al. (2012). Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for Fungi. Proceedings of the National Academy of Sciences of the United States 109: 6241–6246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Schoenlein-Crusius IH, Milanez AI. (1998). Mucorales (Zygomycotina) da Mata Atlântica da Reserva Biológica do Alto da Serra de Paranapiacaba, Santo André, SP. Acta Botanica Brasilica 11: 95–101. [Google Scholar]
  146. Schoenlein-Crusius IH, Milanez AI, Trufem SFB. et al. (2006). Microscopic fungi in the Atlantic Rainforest in Cubatão, São Paulo, Brazil. Brazilian Journal of Microbiology 37: 267–275. [Google Scholar]
  147. Schreder-Gomes SI, Benison KC, Bernau JA. (2022). 830-million-year-old microorganisms in primary fluid inclusions in halite. Geology 50: 918–922. [Google Scholar]
  148. Song CG, Park JH, Lee PM. et al. (2022). Survey of airborne microorganisms in an arcade-type traditional market in Anseong, South Korea. International Journal of Environmental Research and Public Health 19: 6667. [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Song J, Shi L, Li D. et al. (2012). Extensive pyrosequencing reveals frequent intra-genomic variations of internal transcribed spacer regions of nuclear ribosomal DNA. PLoS ONE 7: e43971. [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Sørensen T. (1978). A method of establishing groups of equal amplitude in plant sociology based on similarity of species content and its application to analysis of the vegetation on Danish commons. In: Phytosociology(Mcintosh RP, ed.). Benchmark Papers in Ecology, Stroudsburg: 235–249. [Google Scholar]
  151. Souza CAF, Lima DX, Costa DP. et al. (2020). Mucor variicolumellatus L. Wagner & G. Walther (Mucorales, Mucoromycota): a first record for the Neotropics. Check List 16: 743–747. [Google Scholar]
  152. Souza CAF, Lima DX, Costa DP. et al. (2022). Mucor septatiphorus nom. nov. and other Mucor species recorded from the Brazilian upland forest. Mycotaxon 137: 495–520. [Google Scholar]
  153. Souza CAF, Lima DX, Gurgel LMS. et al. (2017). Coprophilous Mucorales (ex Zygomycota) from three areas in the semi-arid of Pernambuco, Brazil. Brazilian Journal of Microbiology 48: 79–86. [DOI] [PMC free article] [PubMed] [Google Scholar]
  154. Suwannarach N, Kumla J, Supo C. (2021). Cunninghamellasaisamornae (Cunninghamellaceae, Mucorales), a new soil fungus from northern Thailand. Phytotaxa 509: 291–300. [Google Scholar]
  155. Tedersoo L, Bahram M, Põlme S. et al. (2014). Fungal biogeography. Global diversity and geography of soil fungi. Science 346: 1256688. [DOI] [PubMed] [Google Scholar]
  156. Tibpromma S, Hyde KD, Jeewon R. et al. (2017). Fungal diversity notes 491–602: taxonomic and phylogenetic contributions to fungal taxa. Fungal Diversity 83: 1–261. [Google Scholar]
  157. Timmis K, Hallsworth JE. (2022). The darkest microbiome-a post-human biosphere. Microbial Biotechnology 15: 176–185. [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Trufem SFB. (1981a). Mucorales do estado de São Paulo. 3. Gêneros Circinella Van Tieghem & Le Monnier e Cunninghamella Matruchot. Rickia 9: 113–120. [Google Scholar]
  159. Trufem SFB. (1981b). Mucorales do estado de São Paulo. 2. Gêneros Absidia Van Tieghem, Gongronella Ribaldi e Rhizopus Ehrenberg. Rickia: 9: 99–106. [Google Scholar]
  160. Trufem SFB. (1981c). Mucorales do estado de São Paulo. 1. Gênero Mucor Micheli. Rickia 9: 81–91. [Google Scholar]
  161. Trufem SFB. (1984). Mucorales do Estado de São Paulo. 4. Espécies coprófilas. Rickia 11: 53–64. [Google Scholar]
  162. Trufem SFB, Viriato A. (1985). Mucorales do Estado de São Paulo: 6. Mucoraceae coprófilas. Rickia 12: 113–123. [Google Scholar]
  163. Upadhyay HP. (1967). Soil fungi from North-East Brazil III. Phycomycetes. Mycopathologia et Mycologia Applicata 31: 49–62. [Google Scholar]
  164. Upadhyay HP. (1969). Soil fungi from North-East and North Brazil VII. Nova Hedwigia 17: 65–73. [Google Scholar]
  165. Upadhyay HP. (1970). Soil fungi from North-East and North Brazil VIII. Persoonia 6: 111–117. [Google Scholar]
  166. Urquhart AS, Douch JK, Heafield TA. et al. (2021). Diversity of Backusella (Mucoromycotina) in south-eastern Australia revealed through polyphasic taxonomy. Persoonia 46: 1–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  167. van Tuinen D, Zhao B, Gianinazzi-Pearson V. (1998). PCR in studies of AM fungi: from primers to application. In: Mycorrhizal manual. Springer Lab Manual (Varma AK, ed). Springer, Berlin, Heidelberg: 387–400. [Google Scholar]
  168. Vilgalys R, Hester M. (1990). Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several species of Cryptococus. Journal of Bacteriology 172: 4238–4246. [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. Viriato A, Trufem SFB. (1985). Mucorales do estado de São Paulo: 5. Pilobolaceae. Rickia 12: 77–88. [Google Scholar]
  170. Voigt K. (2012). Zygomycota Moreau. In: Syllabus of plant families. A. Engler’s syllabus der Pflanzenfamilien, Blue-green algae, Myxomycetes and Myxomycete-like organisms, Phytoparasitic protists, heterotrophic Heterokontobiota and Fungi (Frey W, ed.). Borntraeger Science Publishers, Berlin: 130–162 [Google Scholar]
  171. Voigt K, Wöstemeyer J. (2001). Phylogeny and origin of 82 Zygomycetes from all 54 genera of the Mucorales and Mortierellales based on combined analysis of actin and translation elongation factor EF-1a genes. Gene 270: 113–120. [DOI] [PubMed] [Google Scholar]
  172. Voigt K, James TY, Kirk PM. et al. (2021). Early-diverging fungal phyla: taxonomy, species concept, ecology, distribution, anthropogenic impact, and novel phylogenetic proposals. Fungal Diversity 109: 59–98. [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Wagner L, Stielow JB, de Hoog GS. et al. (2020). A new species concept for the clinically relevant Mucor circinelloides complex. Persoonia 44: 67–97. [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Wajihi AH, Lee S-Y, Das K. (2018). First report of Gongronella guangdongensis isolated from soil in Korea. The Korean Journal of Mycology 46: 28–33. [Google Scholar]
  175. Walther G, Pawłowska J, Alastruey-Izquierdo A. et al. (2013). DNA barcoding in Mucorales: An inventory of biodiversity. Persoonia 30: 11–47. [DOI] [PMC free article] [PubMed] [Google Scholar]
  176. Walther G, Wagner L, Kurzai O. (2019). Updates on the taxonomy of Mucorales with an emphasis on clinically important taxa. Journal of Fungi 5: 106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  177. Wanasinghe DN, Phukhamsakda C, Hyde K. D, et al. (2018). Fungal diversity notes 709–839: Taxonomic and phylogenetic contributions to fungal taxa with an emphasis on fungi on Rosaceae. Fungal Diversity 89: 1–236. [Google Scholar]
  178. White TJ, Bruns T, Lee S. et al. (1990). Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: PCR protocols: a guide to methods and applications (Innis MA, Gelfand DH, Sninsky JJ, et al., eds). Academic Press, USA: 315–322. [Google Scholar]
  179. Wijayawardene NN, Hyde KD, Al-Ani LKT. et al. (2022). Outline of fungi and fungus-like taxa — 2021. Mycosphere 13: 53–453. [Google Scholar]
  180. Winsett KE, Stephenson SL. (2008). Using ITS sequences to assess intraspecific genetic relationships among geographically separated collections of the myxomycete Didymium squamulosum. Revista Mexicana de Micologia 27: 59–65. [Google Scholar]
  181. Zhao H, Nie Y, Zong TK. et al. (2022). Species diversity and ecological habitat of Absidia (Cunninghamellaceae, Mucorales) with emphasis on five new species from forest and grassland soil in China. Journal of Fungi 8: 471. [DOI] [PMC free article] [PubMed] [Google Scholar]
  182. Zhao H, Nie Y, Zong TK. et al. (2023). Species diversity, updated classification and divergence times of the phylum Mucoromycota. Fungal Diversity 123: 49–157. [Google Scholar]
  183. Zheng RY, Chen GQ. (2001). A monograph of Cunninghamella. Mycotaxon 80: 1–75. [Google Scholar]
  184. Zheng RY, Chen GQ, Huang H. et al. (2007). A monograph of Rhizopus. Sydowia 59: 273–372. [Google Scholar]
  185. Zheng RY, Liu XY, Wang YN. (2017). Circinella (Mucorales, Mucoromycotina) from China. Mycotaxon 132: 43–62. [Google Scholar]
  186. Zhu G, Ding W, Xue M. et al. (2022). Identification and pathogenicity of a new entomopathogenic fungus, Mucor hiemalis (Mucorales: Mucorales), on the root maggot, Bradysia odoriphaga (Diptera: Sciaridae). Journal of Insect Science 22: 2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Ziedan ESHE, Farrag ESH, Sahab AF. (2013). First record and preliminary evaluation of Mucor hiemalis as biocontrol agent on inflorescence brown rot incidence of date palm. Archives of Phytopathology and Plant Protection 46: 617–626. [Google Scholar]
  188. Zong TK, Zhao H, Liu XL. et al. (2021). Taxonomy and phylogeny of four new species in Absidia (Cunninghamellaceae, Mucorales) from China. Frontiers in Microbiology 12: 677836. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Table S1.

Taxa, strains, locations, and GenBank accession numbers of the strains used in this study.

Table S2.

Maximum growth temperature of strains used in this study.

Fig. S1.

Map showing the sample collection sites (left to right and top to bottom: South Korea and Brazil). Yellow and red cursor represent the main collection areas from South Korea (S1, S2, S3) and Brazil (B1), respectively.

Fig. S2.

Phylogenetic relationship of Neofennellomyces with its related species based the ITS sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages. Bootstrap values ≥ 70 % are shown. Fennellomyces heterothallicus CBS 290.86, F. linderi CBS 158.54, Thamnostylum nigricans CBS 690.76, T. piriforme CBS 316.66, and T. repens CBS 692.76 were used as the outgroups. Ex-type and ex-neotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.

Fig. S3.

Phylogenetic relationship of Neofennellomyces with its related species based the LSU sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages. Bootstrap values ≥ 70 % are shown. Fennellomyces heterothallicus CBS 292.86, F. heterothallicus CBS 290.86, Thamnostylum nigricans CBS 690.76, and T. piriforme CBS 316.66 were used as the outgroups. Ex-type and exneotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.

Fig. S4.

Phylogenetic relationship of Neofennellomyces with its related species based the mcm7 sequences. The numbers above or below branches represent maximum likelihood bootstrap percentages. Bootstrap values ≥ 70 % are shown. Fennellomyces linderi CBS 158.54, Thamnostylum repens CBS 692.76, and Zychaea mexicana CBS 441.76 were used as the outgroups. Ex-type and ex-neotype strains are marked with T and NT, respectively. Newly generated sequences are in bold blue.


Articles from Studies in Mycology are provided here courtesy of Westerdijk Fungal Biodiversity Institute

RESOURCES