ArticlePDF Available

Phylogenetic Status of Two Undescribed Zygomycete Species from Korea: Actinomucor elegans and Mucor minutus

Taylor & Francis
Mycobiology
Authors:

Abstract and Figures

During a survey of fungal diversity of the order Mucorales, three zygomycete isolates, CNUFC-YR113-1, CNUFC-KNU16-7, and CNUFC-BS1-1 were isolated from freshwater and soil samples in Korea. The strains were analyzed both morphologically and phylogenetically based on internal transcribed spacer and 28S rDNA gene sequences. Based on their morphology and phylogeny, the CNUFC-YR113-1 and CNUFC-KNU16-7 isolates were identified as Actinomucor elegans, and CNUFC-BS1-1 was identified as Mucor minutus. To the best of our knowledge, the species A. elegans and M. minutus, belonging to an undiscovered taxon, have not been previously described in Korea.
Content may be subject to copyright.
344
Mycobiology
Phylogenetic Status of Two Undescribed
Zygomycete Species from Korea: Actinomucor
elegans and Mucor minutus
Thuong T. T. Nguyen , Hee-Young Jung , Youn Su Lee , Kerstin Voigt and Hyang Burm Lee *
Division of Food Technology, Biotechnology and Agrochemistry, College of Agriculture and Life Sciences, Chonnam National
University, Gwangju 61186, Korea
School of Applied Biosciences, College of Agriculture and Life Sciences, Kyungpook National University, Daegu 41566, Korea
Division of Bioresource Sciences, Kangwon National University, Chuncheon 24341, Korea
JMRC at Leibniz Institute for Natural Product Research and Infection Biology e.V. HKI and Friedrich Schiller University Jena, 07745
Jena, Germany
Abstract During a survey of fungal diversity of the order Mucorales, three zygomycete isolates, CNUFC-YR113-1, CNUFC-
KNU16-7, and CNUFC-BS1-1 were isolated from freshwater and soil samples in Korea. The strains were analyzed both
morphologically and phylogenetically based on internal transcribed spacer and 28S rDNA gene sequences. Based on their
morphology and phylogeny, the CNUFC-YR113-1 and CNUFC-KNU16-7 isolates were identified as Actinomucor elegans, and
CNUFC-BS1-1 was identified as Mucor minutus. To the best of our knowledge, the species A. elegans and M. minutus, belonging
to an undiscovered taxon, have not been previously described in Korea.
Keywords Actinomucor elegans, Mucor minutus, Undiscovered taxa, Zygomycete fungi
Actinomucor and Mucor belong to the subphylum
Mucoromycotina, order Mucorales, family Mucoraceae [1].
The genus Actinomucor was originally described in 1898
by Schostakowitsch [2]. Although the genus is closely
related to Mucor, it differs in having branched stolons that
give rise to rhizoids and sporangiophores. It is also distinct
from the other two genera Rhizopus and Absidia in its
arrangement of the columellae and sporangiophores. The
genus originally contained two species, A. elegans (Eidam)
C. R. Benj. & Hesselt., and A. taiwanensis S. C. Jong &
G. F. Yuan [3, 4]. A. taiwanensis was differentiated from A.
elegans by its larger sporangiospore size and by their differing
maximum growth temperatures: 37
o
C for A. taiwanensis
and 32
o
C for A. elegans. Later, Zheng and Liu [5] renamed
A. taiwanensis to A. elegans var. meitauzae based on
morphological characteristics and molecular analyses. Recently,
Khan et al. [6] proposed the addition of a new variety, A.
elegans var. kuwaitensis. In Index Fungorum (2017; http://
www.indexfungorum.org), the genus Actinomucor contains
only one species named Actinomucor elegans.
Actinomucor species are found in dung, soil, food, and
human sources [5-7]. Some of them are commonly used
for producing popular fermented soybean foods including
Sufu and Chao [8]. In addition, A. elegans is also considered
a good source of glycine aminopeptidase and glucosamine
[9, 10]. A. elegans var. elegans has been reported as a potential
biocontrol agent against the chafer beetle [11].
Mucor Fresen. (Mucoraceae, Mucorales) is characterized
by the formation of non-apophysate sporangia, producing
simple or branched sporangiophores without basal rhizoids.
Zygospores have opposed, non-appendaged suspensors [12].
Mucor species have frequently been detected on substrates
that support the growth of a fungal host, such as in soil,
dung, fruit, and plants [13-15]. Several species are able to
produce enzymes with biotechnological applications [16,
17], while some species are considered the causal agent of
cutaneous zygomycosis in humans [18]. Although there are
more than 300 named species described in the literature,
only approximately 50 are known and described [15].
Research Article
Mycobiology 2017 December, 45(4): 344-352
https://doi.org/10.5941/MYCO.2017.45.4.344
pISSN 1229-8093 • eISSN 2092-9323
© The Korean Society of Mycology
*Corresponding author
E-mail: hblee@jnu.ac.kr
Received October 25, 2017
Revised December 5, 2017
Accepted December 20, 2017
This is an Open Access article distributed under the terms of the
Creative Commons Attribution Non-Commercial License (http://
creativecommons.org/licenses/by-nc/3.0/) which permits unrestricted
non-commercial use, distribution, and reproduction in any medium,
provided the original work is properly cited.
Two New Records of Zygomycete Species in Korea 345
Traditional taxonomy of Mucor species has been determined
based on morphological characteristics such as size and
shape of sporangia as well as the mode of reproduction
(sexual or asexual).
Recently, molecular data have been used to evaluate
mucoralean species [19, 20]. These studies indicated that
Mucor is polyphyletic. Based on the phylogeny of internal
transcribed spacer (ITS) and large subunit (LSU) rDNA
regions of several mucoralean species, Walther et al.
[21] observed that some Mucor species with curved
sporangiophores were grouped with Backusella Hesselt. &
J. J. Ellis. Therefore, these Mucor species were transferred
to Backusella.
In Korea, two new Mucor species have been currently
reported by authors: Mucor koreanus from tangerine fruit
[14] and Mucor stercorarius from rat feces [22]. Only seven
species have been recorded: M. circinelloides, M. hiemalis,
M. mucedo, M. piriformis, M. racemosus, M. fragilis, and M.
irregularis [15, 23]. To our knowledge, there are no specific
published literature records of these species in Korea.
The objective of the present study was to perform
morphological and molecular analyses to characterize two
unrecorded zygomycete species in Korea: Actinomucor
elegans and Mucor minutus.
MATERIALS AND METHODS
Fungal strain isolation from freshwater and soil
samples. Freshwater samples were collected from the
Yeongsan River located in Gwangju, Korea. Soil samples
were collected from the garden of the Chonnam National
University located in Gwangju and a field in Gyeongnam,
Korea. These samples were transported in sterile 50-mL
Falcon tubes, and stored at 4
o
C until examination. Fungi
were isolated using the serial dilution plating method. In
this technique, 1 mL water or 1 g of soil was mixed with
9 mL of sterile distilled water and shaken for 15 min at
25
o
C; serial dilutions ranging from 10
1
to 10
4
were then
made. An aliquot of 0.1 mL from each dilution was
transferred to potato dextrose agar (PDA) and incubated at
25
o
C for 3–7 days. Individual colonies of fungi that showed
varying morphologies were picked up and purely transferred
to another PDA plate. All pure isolates, including A. elegans
and M. minutus, were maintained in PDA slant tubes
and stored in 20% glycerol at 80
o
C at the Environmental
Microbiology Laboratory Fungarium, Chonnam National
University, Gwangju, Korea, as CNUFC-YR113-1, CNUFC-
KNU16-7, and CNUFC-BS1-1. CNUFC-KNU16-7 and
CNUFC-BS1-1 were also deposited at the Collection of
National Institute of Biological Resources (NIBR), Incheon,
Korea; CNUFC-YR113-1 deposited at Culture Collection
of Nakdonggang National Institute of Biological Resources
[NNIBR], Sangju, Gyeongbuk province, Korea.
Morphological studies. For detailed morphological
studies, CNUFC-YR113-1 and CNUFC-BS1-1 strains were
cultured on synthetic mucor agar (SMA; 40 g dextrose, 2 g
asparagine, 0.5 g KH
2
PO
4
, 0.25 g MgSO
4
·7H
2
O, 0.5 g thiamine
chloride, and 15 g agar in 1 L of deionized water). The
plates were incubated at 10, 20, 25, 30, and 35
o
C in the
dark for 7 days. Fragments of mycelia were removed from
cultures, placed on microscope slides with lactophenol
solution (Junsei Chemical Co. Ltd., Tokyo, Japan) and
observed under a light microscope (Olympus, Tokyo, Japan).
DNA extraction, PCR, and sequencing. Genomic DNA
was extracted directly from the mycelia of fungal isolates,
using the Solgent Genomic DNA prep Kit (Solgent Co.
Ltd., Daejeon, Korea). The ITS region and large subunit of
28S rDNA were amplified with the primer pairs ITS4 and
ITS5 [24], and LROR and LR5F [25], respectively. The
PCR amplification mixture (total volume, 20 µL) contained
fungal DNA template, 5 pmol/µL of each primer, and
Accupower PCR Premix (Taq DNA polymerase, dNTPs,
buffer, and a tracking dye; Bioneer Corp., Daejeon,
Korea). PCR products were purified using the Accuprep
PCR Purification Kit (Bioneer Corp.) according to the
manufacturer’s instructions. DNA sequencing was performed
on an ABI 3700 Automated DNA sequencer (Applied
Biosystems Inc., Foster City, CA, USA).
Phylogenetic analysis. The fungal sequences obtained
from the GenBank database (Table 1) were aligned using
Clustal_X v.1.83 [26] and edited with Bioedit v.5.0.9.1 [27].
Table 1 . Taxa, collection numbers, sequences, and GenBank accession numbers used in this study
Taxon name Collection No.
(isolate No.)
GenBank accession No.
ITS 28S
Actinomucor elegans ATCC 46123 AM745430 -
A. elegans CBS 338.72 JN205824 -
A. elegans CBS111562 AB113009 -
A. elegans CBS 100.09 - JN206491
A. elegans CBS154.86 - HM849686
A. elegans CNUFC-YR113-1 MG206066 MG206071
A. elegans CNUFC-YR113-2 MG206067 MG206072
A. elegans CNUFC-KNU16-7 MG206068 MG206073
A. elegans var. elegans ATCC22814
T
AY492092 -
346 Nguyen et al.
Table 1 . Continued
Taxon name Collection No.
(isolate No.)
GenBank accession No.
ITS 28S
A. elegans var. kuwaitiensis CBS117697
T
JN205823 JN206493
A. elegans var. meitauzae ATCC52370
T
AM745432 -
A. elegans var. meitauzae CBS 111558 - JN206492
Backusella circina CBS 128.70
T
- JN206529
B. grandis CBS 186.87
T
- JN206527
B. lamprospora CBS 118.08
T
- JN206531
Benjaminiella multispora CBS 421.70 - KU561719
Blakeslea sinensis CBS 564.91 - JN206515
Choanephora infundibulifera CBS 153.51 - JN206513
Cokeromyces recurvatus CBS 168.59 - JN206408
C. recurvatus CBS 158.50 - KU561716
Mucor aligarensis CBS 993.70
T
- JN206461
M. circinelloides B5-2 KT876701 -
M. circinelloides CBS 108.16 JN205954 -
M. fragilis CBS 236.35 JN205979 -
M. fragilis EML-PUKI06-1 KY047147 -
M. fragilis EML-PUKI06-2 KY047150 -
M. flavus CBS 230.35
T
JN206061 JN206464
M. flavus CBS 681.73 JN206070 -
M. flavus CBS 893.73 - JN206465
M. flavus CBS 182.90 - JN206472
M. fuscus CBS 132.22 JF723619 -
M. fuscus CBS 230.29 JN206204 -
M. genevensis CBS 114.08
T
HM623318 -
M. genevensis CBS 404.71 JN206042 -
M. heterogamus CBS 338.74 JN206169 JN206488
M. heterogamus CBS 252.85 - JN206490
M. heterogamus CBS 405.58 JN206167 -
M. hiemalis CBS 242.35 JN206134 -
M. hiemalis CBS 115.18 JN206127 -
M. irregularis CBS 977.68 JX976259 -
M. irregularis EML-PUKI12-1 KY047151 -
M. irregularis EML-PUKI12-2 KY047146 -
M. koreanus EML-QT1 KT936259 -
M. koreanus EML-QT2 KT936260 -
M. luteus CBS 243.35 JX976254 -
M. minutus CBS 586.67
T
JN206048 JN206463
M. minutus CNUFC-BS1-1 MG206069 MG206074
M. minutus CNUFC-BS1-2 MG206070 MG206075
M. mucedo CBS 542.66 JN206086 JN206480
M. mucedo CBS 987.68 JN206089 JN206480
M. nidicola EML-SBD1 KY047148 -
M. nidicola EML-SBD2 KY047149 -
M. plasmaticus CBS 275.49 - JN206483
M. saturninus CBS 974.68 - JN206458
M. stercoraria CNUFC-UK2-1 KX839689 -
M. stercoraria CNUFC-UK2-2 KX839680 -
M. strictus CBS 100.66 - JN206477
M. racemosus CBS 260.68 JF723556 -
M. velutinosus UTHSC 04-1961 JF299208 -
M. velutinosus UTHSC 04-1981 JF299212 -
U. nan a NRRL 22420 KM017731 KM017708
Bold letters indicate isolates and accession numbers determined in our study.
ITS, internal transcribed spacer; ATCC, American Type Culture Collection, Manassas, VA, USA; CBS, Centraalbureau voor Schim-
melcultures, Utrecht, The Netherlands; CNUFC, Chonnam National University Fungal Collection, Gwangju, South Korea; EML, Environ-
mental Microbiology Laboratory Fungarium, Chonnam National University, Gwangju, South Korea; NRRL (ARS Culture Collection,
Peoria, Illinois); T, ex-type strain.
Two New Records of Zygomycete Species in Korea 347
Phylogenetic analyses were performed using MEGA 6
software [28], and maximum likelihood was constructed
by Kimuras two-parameter correction method. The fungus
Umbelopsis nana was used as an outgroup. The reliability
of internal branches was assessed using the p-distance
substitution model with 1,000 bootstrap replications.
RESULTS
Phylogenetic analysis. Phylogenetic analyses of the
two sequence datasets (ITS and 28S rDNA) showed that
the strains CNUFC-YR113-1, CNUFC-YR113-2, CNUFC-
KNU16-7, CNUFC-BS1-1, and CNUFC-BS1-2 were placed
within the same clade with species of Actinomucor and
Mucor (Figs. 1 and 2).
In the BLASTn analysis of the ITS sequence, CNUFC-
YR113-1 and CNUFC-BS1-1 represented 99.8% (535/536
bp) and 99.4% (613/617 bp) sequence identity values with
A. elegans (GenBank accession No. JN205824) and M.
minutus (GenBank accession No. JN206048), respectively.
In the BLASTn analysis of the 28S sequence, CNUFC-
YR113-1 and CNUFC-BS1-1 strains showed 98.1% (634/
Fig. 1. Phylogenetic tree based on maximum likelihood analysis of internal transcribed rDNA sequences for Actinomucor
elegans CNUFC-YR113-1, A. elegans CNUFC-YR113-2, A. elegans CNUFC-KNU16-7, Mucor minutus CNUFC-BS1-1, and M.
minutus CNUFC-BS1-2. Umbelopsis nana was used as an outgroup. Bootstrap support values of 50% are indicated at the
nodes. The bar indicates the number of substitutions per position.
348 Nguyen et al.
644 bp) and 100% (682/682 bp) identity values with A. elegans
(GenBank accession No. JN205827) and M. minutus
(GenBank accession No. JN206463), respectively.
Taxonomy of CNUFC-YR113-1.
Actinomucor elegans (Eidam) C. R. Benj. & Hesselt.,
Mycologia 49: 241 (1957) (Table 2, Fig. 3).
Table 2 . Morphological characteristics of CNUFC-YR113-1 and the reference Actinomucor elegans grown on synthetic mucor
agar medium at 25
o
C
Character CNUFC-YR113-1 Actinomucor elegans
a
Colony color Rapid-growing, first white then deep olive-buff,
reverse white
Rapid-growing, first white then deep olive-buff,
reverse white to pale olive-buff
Sporangiophores 12.2–20.5 µm in width, variable in length Up to 30 µm in width, variable in length
Primary sporangia Globose to subglobose, multispored, 42.3–83.5 ×
39.9–82.1 µm
Less than 80 µm, multispored
Secondary sporangia Globose to subglobose, multispored, 29.9–46.2 ×
27.5–44.3 µm
Mostly 20–50 µm in diameter, multispored
Columellae inside
primary sporangia
Diverse in shape, oval, pyriform, oblong, 23.3–44.8 ×
22.6–42.9 µm
Elongate-oval to pyriform, 50–60 × 30–40 µm
Columellae inside
secondary sporangia
Globose, 14.5–26.5 × 17.8–30.4 µm Globose, 12–30 µm
Sporangiospores Globose to subglobose, 6.1–8.5 × 5.8–8.1 µm Globose, mostly 6–8 µm in diameter
Chlamydospores Present Present
Zygospores Absent Unknown
a
From the description by Benjamin and Hesseltine [3].
Fig. 2. Phylogenetic tree based on maximum likelihood analysis of 28S rDNA sequences for Actinomucor elegans CNUFC-
YR113-1, A. elegans CNUFC-YR113-2, A. elegans CNUFC-KNU16-7, Mucor minutus CNUFC-BS1-1, and M. minutus CNUFC-
BS1-2. Umbelopsis nana was used as an outgroup. Bootstrap support values of 50% are indicated at the nodes. The bar
indicates the number of substitutions per position.
Two New Records of Zygomycete Species in Korea 349
=Rhizopus elegans Eidam, Jahresber. Schles. Ges. Vaterl.
Kultu. 61: 232 (1884).
=Mucor elegans (Eidam) J. Schröt., Kryptogamen-Flora
von Schlesien 3-1: 207 (1886).
=Mucor corymbosus Harz, Bull. Soc. Imp. Nat. Moscou
44: 143 (1871).
=Actinomucor repens Schostak., Ber. Dtsc. Bot. Ges. 16:
155 (1898).
=Glomerula repens Bainier, Bull. Soc. Mycol. Fr. 19: 154
(1903).
=Mucor botryoides Lendn., Bull. Soc. Bot. Genève 2: 79
(1910).
=Mucor botryoides var. minor C.N. Jensen, Bull.
Cornell Univ. Agric. Exp. Stn. 315: 457 (1912).
=Mucor cunninghamelloides Pispek, Acta Bot. Inst. Bot.
Univ. Zagreb. 4: 91 (1929).
=Actinomucor corymbosus Naumov, Opredelitel
Mukorovykh (Mucorales): 56 (1935).
=Actinomucor corymbosus f. palaestinus Rayss,
Palestine J. Bot. 3: 162 (1945).
Description: Colonies grew rapidly at 25
o
C on SMA,
filling the Petri dish after 5 days of incubation. The colony
color was initially white, later deep olive-buff. The colony
reverse was white. Sporangiophores were 12.2–20.5 µm
wide, erect, branched, irregular, and verticillate. Primary
sporangia were globose to subglobose, and measured 42.3–
83.5 × 39.9–82.1 µm. Secondary sporangia were formed with
same shape as the primary sporangia, and measured 29.9–
46.2 × 27.5–44.3 µm. Columellae inside the primary sporangia
were diverse in shape, oval, pyriform, oblong, and measured
23.3–44.8 × 22.6–42.9 µm. Columellae inside the secondary
sporangia were globose, and measured 14.5–26.5 × 17.8–
30.4 µm. Sporangiospores were globose to subglobose, and
measured 6.1–8.5 × 5.8–8.1 µm. Chlamydospore formations
were well-defined on the medium. Zygospores were not
observed.
Taxonomy of CNUFC-BS1-1.
Mucor minutus (Baijal & B. S. Mehrotra) Schipper, Stud.
Mycol. 10: 24 (1975) (Table 3, Fig. 4).
=Mucor griseoochraceus var. minuta Baijal & B. S.
Mehrotra, Sydowia 19: 206 (1966).
=Mucor saturninus var. minutus (Baijal & B. S. Mehrotra)
Milko, Opredeltiel mukoral’nykh gribov. 129 (1974).
Description: Colonies grew rapidly on SMA, attaining a
diameter of 70–72 mm after 5 days at 25
o
C. The colony
Fig. 3. Morphology of Actinomucor elegans CNUFC-YR113-1. A, Colonies on synthetic mucor agar; B–D, Sporangia on
branched sporangiophores (observed under stereo-microscope); E–H, Branched sporangiophores forming sporangia and
columellae (observed under light microscope); I, Sporangiospores (scale bars: B–D = 200 μm, E–H = 50 μm, I = 20 μm).
Table 3 . Morphological characteristics of CNUFC-BS1-1 and the reference species Mucor minutus grown on synthetic mucor
agar medium at 25
o
C
Character CNUFC-BS1-1 Mucor minutus
a
Colony color First white and later smoke gray Smoke gray, up to 19 mm in height
Sporangiophore 9–24.5 µm wide, variable in length Up to 20 µm, variable in length
Sporangia Globose, 37.1–109.8 µm × 36.4–103.4 µmUp to 175µm
Columella Globose to ellipsoidal, 27.9–95.2 µm × 24.8–84.5 µm Cylindrical to ellipsoid, 110–135 µm in width
Sporangiospores Globose, 4.3–5.6 µm × 4.1–5.0 µm Subspherical, 4–5 µm in diameter
Zygospore Absent Unknown
a
From the description by Schipper [29].
350 Nguyen et al.
color was initially white, later turning to smoke gray.
Sporangiophores were 9–24.5 µm wide, erect, mostly
branched, and irregular. Sporangia were globose, and
measured 37.1–109.8 µm × 36.5–103.4 µm. Columellae were
globose to ellipsoidal, and measured 27.9–95.2 µm×24.8
84.5 µm. Sporangiospores were globose, and measured 4.3–
5.6 µm × 4.1–5.0 µm. Zygospores were not observed on
artificial media.
DISCUSSION
Despite the wide intraspecific variation found among some
taxa, the rDNA ITS and D1/D2 regions have been used as
critical barcode markers for identifying mucoralean fungi
at the species level, including taxa of Actinomucor and
Mucor [21].
In the ITS and LSU phylogenetic trees, our strains
CNUFC-YR113-1, CNUFC-YR113-2, and CNUFC-KNU16-
7 were clustered within the elegans clade including A.
elegans, A. elegans var. meitauzae, and var. kuwaitiensis in a
well-supported clade. However, our strain CNUFC-YR113-
1 differed from A. elegans var. meitauzae and A. elegans
var. kuwaitiensis in sporangiospore size; CNUFC-YR113-1
strain exhibited smaller sporangiospores (6.1–8.5 × 5.8–8.1
µm) than A. elegans var. meitauzae (7–19.5 × 6–15 µm)
and A. elegans var. kuwaitiensis (5–12 µm). The maximum
growth temperature of our strain was 35
o
C, while A.
elegans var. meitauzae and A. elegans var. kuwaitiensis were
able to grow under higher temperatures up to 40
o
C.
Jong and Yuan [4] reported that growth temperature is a
criterion for distinguishing between A. elegans and A.
taiwanensis. These authors showed that A. taiwanensis has
a maximum growth temperature of 37
o
C, while A. elegans
does not grow at this temperature. Contrary to reports by
Jong and Yuan [4], maximum growth temperature is less
useful for distinguishing between the varieties [5, 6].
The morphological features of our isolates were in
line with the description of A. elegans by Benjamin and
Hesseltine [3], as the properties including shape, size of the
sporangiospores (6–8 µm), and maximum temperature for
growth were compared. Under these criteria, our isolate
was identified as A. elegans.
In the tree based on D1/D2 sequence analyses, the strains
CNUFC-BS1-1 and CNUFC-BS1-2 were placed into the
minutus clade within the M. flavus group as presented
by Walther et al. [21] including: M. flavus, M. saturninus,
M. aligarensis, and M. minutus (Fig. 2), and formed a
monophyletic group with M. minutus (type species). The
CNUFC-BS1-1 isolate was morphologically most similar to
M. minutus as described by Schipper [29], although there
were differences in the shape and size of columellae. The
size of columellae described by Schipper [29] was larger
(110–135 µm) than those (27.9–95.2 × 24.8–84.5 µm) observed
in our isolate. According to Schipper [29], the M. minutus
species is similar in morphology and closely related to M.
flavus because they produce columellae with the same size.
However, sporangiospores with different sizes and shapes
have been observed. M. minutus has smaller sporangiospores
(4–5 µm) than M. flavus (7–12 × 4–6.5 µm). Comparing
the colony morphology and culture characteristics of the
isolate with previous descriptions [29], the present isolate
was similar to M. minutus, with some exceptions. Our M.
minutus isolate presented one to three septa below the
columella, which were not described by Schipper [29].
Recently, several studies have focused on the increased
incidence of mucormycosis in both immunocompromised
and immunocompetent patients [30]. Some species belonging
to the order Mucorales (subphylum Mucoromycotina) are
considered opportunistic pathogens. Particularly, four families,
including Cunninghamellaceae, Lichtheimiaceae, Mucoraceae,
Fig. 4. Morphology of Mucor minutus CNUFC-BS1-1. A, Colonies on synthetic mucor agar; B, C, Sporangiophores with long
and short branches and sporangia (observed under stereo-microscope); D, E, Sporangia and sporangiophores (observed under
light microscope); F–J, Columellae with collarette and sporangial septa below the columellae (white arrow); K, Sporangiospores
(scale bars: B, C = 200 μm, D–J = 50 μm, K = 20 μm).
Two New Records of Zygomycete Species in Korea 351
and Syncephalastraceae, have been described to be responsible
for human infections [31].
More recently, A. elegans and A. elegans var. kuwaitiensis
have been reported as the agent of mucormycosis in humans
in several cases [6, 7, 32]. Morphological keys are available
for identifying Actinomucor. However, it is still difficult to
identify taxa to intraspecific rank in Actinomucor. Thus,
taxonomic revision and phylogenetic analysis are needed
in future studies.
Interestingly, A. elegans has been reported as protease
enzyme for generation of small peptides with ACE-inhibitory
activity from razor clam Sinonovacula constricta meat [33].
So this finding suggests that the strain CNUFC-YR113-1
may be a useful source for biotechnological applications.
ACKNOWLEDGEMENTS
This work was in part supported by the Graduate Program
for the Undiscovered Taxa of Korea, and in part by the
Project on Survey and Discovery of Indigenous Fungal
Species of Korea funded by NIBR and Project on Discovery
of Fungi from Freshwater and Collection of Fungarium
funded by NNIBR of the Ministry of Environment (MOE),
and in part carried out with the support of Cooperative
Research Program for Agriculture Science and Technology
Development (PJ012957), Rural Development Administration,
Republic of Korea.
REFERENCES
1. Voigt K. Chytridiomycota. In: Frey W, editor. Syllabus of
plant families–A. Engler’s syllabus der pflanzenfamilien.
Part 1/1: Blue-green algae, Myxomycetes and Myxomycete-
like organisms, phytoparasitic Protists, heterotrophic
Heterokontobionta and Fungi. Stuttgart: Borntraeger Verlag;
2012. p. 106-29.
2. Schostakowitsch W. 1898. Actinomucor repens n. gen. n. sp.
Ber Deut Bot Ges 1898;16:155-8.
3. Benjamin CR, Hesseltine CW. The genus Actinomucor. Mycologia
1957;49:240-9.
4. Jong SC, Yuan GF. Actinomucor taiwanensis sp. nov., for
manufacture of fermented soybean food. Mycotaxon 1985;23:
261-4.
5. Zheng RY, Liu XY. Actinomucor elegans var. meitauzae, the
correct name for A. taiwanensis and Mucor meitauzae
(Mucorales, Zygomycota). Nova Hedwigia 2005;80:419-31.
6. Khan ZU, Ahmad S, Mokaddas E, Chandy R, Cano J, Guarro
J. Actinomucor elegans var. kuwaitiensis isolated from the
wound of a diabetic patient. Antonie Van Leeuwenhoek 2008;
94:343-52.
7. Tully CC, Romanelli AM, Sutton DA, Wickes BL, Hospenthal
DR. Fatal Actinomucor elegans var. kuwaitiensis infection
following combat trauma. J Clin Microbiol 2009;47:3394-9.
8. Hesseltine CW. A millennium of fungi, food and fermentation.
Mycologia 1965;57:149-97.
9. Ma X, Zhou X, Yoshimoto T. Purification and properties of a
novel glycine amino peptidase from Actinomucor elegans and
its potential application. J Appl Microbiol 2004;97:985-91.
10. Wang S, Li P, Su J, Liang R, Wu XK. Enhanced glucosamine
production with Actinomucor elegans based on stimulating
factor of methanol. Indian J Microbiol 2014;54:459-65.
11. Karimi K, Arzanlou M, Ahari AB, Ghazi MM. Phenotypic
and molecular characterization of the causal agent of chafer
beetle mortality in the wheat fields of the Kurdistan province,
Iran. J Plant Prot Res 2015;55:227-34.
12. Benny GL, Humber RA, Voigt K. Zygomycetous fungi: phylum
Entomophthoromycota and subphyla Kickxellomycotina,
Mortierellomycotina, Mucoromycotina, and Zoopagomycotina.
In: McLaughlin DJ, Spatafora JW, editors. The Mycota, Vol.
VII, part A, Systematics and evolution. New York: Springer-
Verlag; 2014. p. 209-50.
13. Benny GL. Methods used by Dr. R. K. Benjamin, and other
mycologists, to isolate Zygomycetes. Aliso 2008;26:37-61.
14. Li GJ, Hyde KD, Zhao RL, Hongsanan S, Abdel-Aziz FA,
Abdel-Wahab MA, Alvarado P, Alves-Silva G, Ammirati JF,
Ariyawansa HA, et al. Fungal diversity notes 253-366: taxonomic
and phylogenetic contributions to fungal taxa. Fungal Divers
2016;78:1-237.
15. Nguyen TT, Duong TT, Lee HB. Characterization of two new
records of mucoralean species isolated from gut of soldier fly
larva in Korea. Mycobiology 2016;44:310-3.
16. Alves MH, Campos-Takaki
GM, Porto AL, Milanez AI.
Screening of Mucor spp. for the production of amylase, lipase,
polygalacturonase and protease. Braz J Microbiol 2002;33:
325-30.
17. Thompson DP, Eribo BE. Extracellular enzyme production by
Rhizopus and Mucor species on solid media. Can J Microbiol
1984;30:126-8.
18. Alvarez E, Stchigel AM, Cano J, Sutton DA, Fothergill AW,
Chander J, Salas V, Rinaldi MG, Guarro J. Molecular
phylogenetic diversity of the emerging mucoralean fungus
Apophysomyces: proposal of three new species. Rev Iberoam
Micol 2010;27:80-9.
19. Jacobs K, Botha A. Mucor renisporus sp. nov., a new
coprophilous species from Southern Africa. Fungal Divers
2008;29:27-35.
20. O’Donnell K, Lutzoni FM, Ward TJ, Benny GL. Evolutionary
relationships among mucoralean fungi (Zygomycota): evidence
for family polyphyly on a large scale. Mycologia 2001;93:286-
97.
21. Walther G, Pawłowska J, Alastruey-Izquierdo A, Wrzosek M,
Rodriguez-Tudela JL, Dolatabadi S, Chakrabarti A, de Hoog
GS. DNA barcoding in Mucorales: an inventory of biodiversity.
Persoonia 2013;30:11-47.
22. Tibpromma S, Hyde KD, Jeewon R, Maharachchikumbura SS,
Liu JK, Bhat DJ, Jones EB, McKenzie EH, Camporesi E,
Bulgakov TS, et al. Fungal diversity notes 491-602: taxonomic
and phylogenetic contributions to fungal taxa. Fungal Divers
2017;83:1-261.
23. Lee YS, Jung HY, Lee HB, Kim SH, Shin KS, Eom AH, Kim
C, Lee SY, Korean Society of Mycology. National list of
species of Korea. Ascomycota, Glomeromycota, Zygomycota,
Myxomycota, Oomycota. Incheon: National Institute of
Biological Resources; 2015.
24. White TJ, Bruns T, Lee S, Taylor J. Amplification and direct
352 Nguyen et al.
sequencing of fungal ribosomal RNA genes for phylogenetics.
In: Innis MA, Gelfand DH, Sninsky JJ, White TJ, editors.
PCR protocols: a guide to methods and applications. San
Diego (CA): Academic Press; 1990. p. 315-22.
25. Lee HB. Molecular phylogenetic status of Korean strain of
Podosphaera xanthii, a causal pathogen of powdery mildew
on Japanese thistle (Cirsium japonicum) in Korea. J Microbiol
2012;50:1075-80.
26. Thompson JD, Gibson TJ, Plewniak F, Jeanmougin F, Higgins
DG. The CLUSTAL_X windows interface: flexible strategies
for multiple sequence alignment aided by quality analysis
tools. Nucleic Acids Res 1997;25:4876-82.
27. Hall TA. BioEdit: a user-friendly biological sequence alignment
editor and analysis program for Windows 95/98/NT. Nucleic
Acids Symp Ser 1999;41:95-8.
28. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S.
MEGA6: Molecular Evolutionary Genetics Analysis version
6.0. Mol Biol Evol 2013;30:2725-9.
29. Schipper MA. On Mucor mucedo, Mucor f lavus and related
species. Stud Mycol 1975;10:1-33.
30. Kwon-Chung KJ. Taxonomy of fungi causing mucormycosis
and entomophthoramycosis (zygomycosis) and nomenclature
of the disease: molecular mycologic perspectives. Clin Infect
Dis 2012;54(Suppl 1):S8-15.
31. Vitale RG, de Hoog GS, Schwarz P, Dannaoui E, Deng S,
Machouart M, Voigt K, van de Sande WW, Dolatabadi S,
Meis JF, et al. Antifungal susceptibility and phylogeny of
opportunistic members of the order Mucorales. J Clin Microbiol
2012;50:66-75.
32. Mahmud A, Lee R, Munfus-McCray D, Kwiatkowski N,
Subramanian A, Neofytos D, Carroll K, Zhang SX. Actinomucor
elegans as an emerging cause of mucormycosis. J Clin Microbiol
2012;50:1092-5.
33.Li Y, Sadiq FA, Fu L, Zhu H, Zhong M, Sohail M.
Identification of angiotensin I-converting enzyme inhibitory
peptides derived from enzymatic hydrolysates of razor clam
Sinonovacula constricta. Mar Drugs 2016;14:E110.
... It was described by Fresenius in 1850 and the type species is Mucor mucedo. Recent molecular studies of mucoralean species have indicated that Mucor is polyphyletic (Nguyen et al. 2017). However, even with definite results showing the polyphyly of Mucor, few clear lineages within Mucor are recognized. ...
... Some of these lineages share innate characteristics, such as sporangium size and branching of tall sporangiophores and the morphology is still widely used in current taxonomy . Analysis of internal transcribed spacer (ITS) and large subunit (LSU) rDNA sequence data of several mucoralean species, showed that some Mucor species with curved sporangiophores grouped with species of Backusella and hence was transferred to Backusella Nguyen et al. 2017). Mucor species are commonly isolated from soil, dung, insect, and fruits (Benny 2008). ...
... Mucor species are commonly isolated from soil, dung, insect, and fruits (Benny 2008). Some species are of biotechnological importance such as biofuel, enzyme, terpernoid production and biotransformation while other species cause mucoromycosis in immunosuppressed humans (Nguyen et al. 2017;Steve et al. 2018;Morin-Sardin et al. 2017). Comparative analyses of five Mucor species based on their lifestyles (M. ...
Article
Full-text available
This is a continuation of a series focused on providing a stable platform for the taxonomy of phytopathogenic fungi and fungus-like organisms. This paper focuses on one family: Erysiphaceae and 24 phytopathogenic genera: Armillaria, Barriopsis, Cercospora, Cladosporium, Clinoconidium, Colletotrichum, Cylindrocladiella, Dothidotthia,, Fomitopsis, Ganoderma, Golovinomyces, Heterobasidium, Meliola, Mucor, Neoerysiphe, Nothophoma, Phellinus, Phytophthora, Pseudoseptoria, Pythium, Rhizopus, Stemphylium, Thyrostroma and Wojnowiciella. Each genus is provided with a taxonomic background, distribution, hosts, disease symptoms, and updated backbone trees. Species confirmed with pathogenicity studies are denoted when data are available. Six of the genera are updated from previous entries as many new species have been described.
... Phylogenetic analysis was performed using MEGA 7.0 [36], and neighbour-joining tree was constructed by Kimura 2-parameter distances, with bootstrapping of 1000 replicates. Umbelopsis nana was used as outgroup to root the Mucor tree, while the Aspergillus fungi tree was rooted with Aspergillus peyronelii [37][38][39][40]. ...
... Phylogenetic analysis clarified the relationships of fungi isolated from crude oil-contaminated soil within Mucor and Aspergillus. Sequences retrieved from the Nigerian sites analyzed strains could be aligned with sequences from fungi isolated from a variety of soils, animal faeces, water and oilcontaminated surface soils [37][38][39]. The neighbour-joining tree of Mucor fungi revealed that the sequence obtained from isolated Mucor strain B-Yorla10 in this study was closely related to KY474527-Mucoromycotina sp. ...
Article
Full-text available
Many free-living saprobic fungi are nature recruited organisms for the degradation of wastes, ranging from lignocellulose biomass to organic/inorganic chemicals, aided by their production of enzymes. In this study, fungal strains were isolated from contaminated crude-oil fields in Nigeria. The dominant fungi were selected from each site and identified as Aspergillus oryzae and Mucor irregularis based on morphological and molecular characterization, with site percentage incidences of 56.67% and 66.70%, respectively. Selected strains response/tolerance to complex hydrocarbon (used engine oil) was studied by growing them on Bushnell Haas (BH) mineral agar supplemented with the hydrocarbon at different concentrations, i.e., 5%, 10%, 15%, and 20%, with a control having dextrose. Hydrocarbon degradation potentials of these fungi were confirmed in BH broth culture filtrates pre-supplemented with 1% engine oil after 15 days of incubation using GC/MS. In addition, the presence of putative enzymes, laccase (Lac), manganese peroxidase (MnP), and lignin peroxidase (LiP) was confirmed in culture filtrates using appropriate substrates. The analyzed fungi grew in hydrocarbon supplemented medium with no other carbon source and exhibited 39.40% and 45.85% dose inhibition response (DIR) respectively at 20% hydrocarbon concentration. An enzyme activity test revealed that these two fungi produced more Lac than MnP and LiP. It was also observed through the GC/MS analyses that while A. oryzae acted on all hydrocarbon components in the used engine oil, M. irregularis only degraded the long-chain hydrocarbons and BTEX. This study confirms that A. oryzae and M. irregularis have the potential to be exploited in the bio-treatment and removal of hydrocarbons from polluted soils.
... identities = 434/447 (97.09%); two gaps), using EF-1α was Wardomyces pulvinatus CBS 112. 65 In the analysis involving species within the Microascaceae, the individual dataset for ITS, LSU, EF-1α, and tub2 showed no conflicts related to the tree topologies for the 70% reciprocal bootstrap trees; thus, a multi-gene analysis was performed. The final concatenated dataset included 43 ingroup strains belonging to the genera Acaulium, Cephalotrichum, Gamsia, Fairmania, Wardomyces and Wardomycopsis, and Microascus longirostris CBS 196. 61 and Scopulariopsis brevicaulis MUCL 40726 as the outgroup. The alignment encompassed a total of 3019 characters, including gaps (661 for ITS, 843 for LSU, 965 for EF-1α, and 550 for tub2), 732 of them parsimony informative (203 for ITS, 73 for LSU, 190 for EF-1α, and 266 for tub2) and 960 of them being variable sites (293 for ITS, 91 for LSU, 264 for EF-1α, and 312 for tub2). ...
Article
Full-text available
The Saladas de Sástago-Bujaraloz is an endorheic and arheic complex of lagoons located in the Ebro Basin and protected by the Ramsar Convention on Wetlands. Due to the semi-arid climate of the region and the high salinity of their waters, these lagoons constitute an extreme environment. We surveyed the biodiversity of salt-tolerant and halophilic fungi residents of the Laguna de Pito, a lagoon belonging to this complex. Therefore, we collected several samples of water, sediments, and soil of the periphery. Throughout the study, we isolated 21 fungal species, including a strain morphologically related to the family Microascaceae. However, this strain did not morphologically match any of genera within this family. After an in-depth morphological characterization and phylogenetic analysis using a concatenated sequence dataset of four phylogenetically informative molecular markers (the internal transcribed spacer region (ITS) of the nuclear ribosomal DNA (nrDNA); the D1-D2 domains of the 28S gene of the nuclear ribosomal RNA (LSU); and a fragment of the translation elongation factor 1-alpha (EF-1α) and the β-tubulin (tub2) genes), we established the new genus Dactyliodendromyces, with Dactyliodendromyces holomorphus as its species. Additionally, as a result of our taxonomic study, we reclassified the paraphyletic genus Wardomyces into three different genera: Wardomyces sensu stricto, Parawardomyces gen. nov., and Pseudowardomyces gen. nov., with Parawardomyces ovalis (formerly Wardomyces ovalis) and Pseudowardomyces humicola (formerly Wardomyces humicola) as the type species of their respective genera. Furthermore, we propose new combinations, including Parawardomyces giganteus (formerly Wardomyces giganteus) and Pseudowardomyces pulvinatus (formerly Wardomyces pulvinatus).
... Identification using molecular methods is sufficient, since the ITS region is a barcode marker with an appropriate discriminative power in Mucorales, being able to clearly distinguish Actinomucor from other morphologically related genera, such as Absidia, Rhizomucor, Mucor, and Rhizopus (Ranjith et al. 2021;Walther et al. 2013). The reconstruction of phylogenetic trees with concatenated molecular markers yielded conclusive results in the same sense, with the grouping of species described by authors such as Nguyen et al. (2017a) and Walther et al. (2013). It is also worth noting that using world-renowned collection strains such as CBS or ATCC ensures that the trees have been constructed with verified sequences. ...
Article
Full-text available
The aim of this work was to isolate fungal strains from phytotoxic agricultural soils, screen them, categorize the most tolerant fungi to three fungicides, and identify them by a molecular approach. In this study, 28 fungal strains were isolated from phytotoxic agricultural soil with intensive use of pesticides. The capacity of fungi to resist and degrade different concentrations of carbendazim, captan, and zineb was determined by an exploratory multivariate analysis. Actinomucor elegans LBM 239 was identified as the most tolerant fungus to these fungicides, degrading a 86.62% of carbendazim after 7 days of treatment. In conclusion, A. elegans LBM 239 demonstrated the highest tolerance and capacity to biodegrade carbendazim, becoming a potential candidate for bioremediation of contaminated soils with carbendazim, captan, or zineb.
... The columellae of short sporangiophores described here are conical and subglobose, unlike those reported for the extype, which are cylindrical to ellipsoid, some conical to flattened. Nguyen et al. (2017) observed the presence of one to three septa below the columellae, unlike our isolate and the ex-type (Schipper 1975), which have no septum below the columellae. ...
Article
Full-text available
Mucor minutus (Baijal & B.S. Mehrotra) Schipper is described for the first time from northeastern Brazil. The specimen was isolated from the soil of Florestas do Jussará, an upland forest fragment in Pernambuco state. In Brazil, M. minutus has previously been isolated only from the state of São Paulo from the soil of the Atlantic Forest. The occurrence of M. minutus in Florestas do Jussará is presented and discussed, and a detailed description and illustration of the specimen are presented. This article contributes to the knowledge of mucoralean distribution.
... Subsequently, single fungal colonies are isolated to obtain pure cultures. Commonly used media include Malt-Extract Agar, and Potato-Dextrose Agar (Madden et al. 2011;Nguyen et al. 2018). Optimal temperature and growth conditions of basal fungi highly depend on the type of fungus. ...
Article
To date, the majority of mycological studies are focused on Ascomycota and Basidiomycota. Research on basal fungi is rather limited with an obvious scarcity of data. Classification of basal fungi has witnessed ongoing changes in the last few decades. There have been numerous taxonomic revisions, establishment of new phyla and debates on the possibility to exclude some basal lineages from the currently circumscribed fungal groups. Zoosporic phyla such as Rozellomycota, Aphelidiomycota and Chytridiomycota are characterized by zoospores, which are motile spores with a flagellum. The position of the first two within the fungal kingdom is currently being debated. Non zoosporic phyla are collectively called zygomycetous fungi and include Mucoromycota and Zoopagomycota. Basal fungi perform important ecological roles. Some groups of basal fungi control the population density of plankton, Oomycetes, aphids, and small invertebrates, while taking part in recycling of energy and matter in food webs. A few basal fungi are known for uptaking toxic heavy metals from the environment. The earliest basal fungi emerged just after the divergence of the nucleariid line of phagotrophs. Various taxonomic studies question the monophyly of some basal phyla, such as Chytridiomycota. Depending on the life mode of the fungi and whether motile or non-motile spores are produced, the isolation methods differ. For the culture of parasitic zoosporic fungi, co-culturing the fungi with their host is recommended while for saprobes ones, baits can be used. For zygomycetous fungi, common methods include dilution plating and direct inoculation. In this review, we pinpoint research gaps and discuss the current classification schemes along with information on evolution, diversity, current methods and challenges to isolate and culture basal fungi.
Article
Full-text available
Mucor representatives are mostly rapidly growing cosmopolitan soil saprotrophs of early diverged Mucoromycotina subphylum. Although this is the most speciose genus within the group, some lineages are still understudied. In this study, new species of Mucor was isolated from the post-mining area in southwestern Poland, where soil chemical composition analysis revealed high concentration of hydrocarbons and heavy metals. Phylogenetic analysis based on multigene phylogeny showed that the new isolate clusters distinctly from other Mucor species as a sister group to Mucor microsporus. New species Mucor thermorhizoides Abramczyk (Mucorales, Mucoromycota) is characterized by the extensive rhizoid production in elevated temperatures and formation of two layers of sporangiophores. It also significantly differs from M. microsporus in the shape of spores and the size of sporangia. M. thermorhizoides was shown to be able to grow in oligotrophic conditions at low temperatures. Together with M. microsporus they represent understudied and highly variable lineage of the Mucor genus.
Article
Full-text available
Research into freshwater fungi has generated a wealth of information over the past decades with various published articles, i.e., reviews, books, and monographs. With the advancement of methodologies used in freshwater fungal research, and numerous mycologists working on this ecological group, our knowledge progress and understanding of freshwater fungi, including novel discoveries and new insights in the ecology of freshwater fungi, has advanced. With this enormous progress, it is timely that an updated account of freshwater fungi be compiled in one volume. Thus, this account is published to give a comprehensive overview of the different facets of freshwater fungal biology. It includes an updated classification scheme based on the latest taxonomic and phylogenetic analysis of freshwater fungal taxa, including their evolutionary history. The biology, diversity, and geographical distribution of higher and basal freshwater fungi are also discussed in the entries. A section on dispersal and adaptation of filamentous freshwater fungi is included in the present work. The ecological importance and role of fungi in the breakdown of wood in freshwater habitats, including their physiology, are discussed in detail. The biotechnological potential of freshwater fungi as producers of bioactive metabolites are reviewed, with methodologies in antimicrobial drug discovery. The present volume also provides an overview of different high throughput sequencing (HTS) platforms for freshwater fungal research highlighting their advantages and challenges, including recent studies of HTS in identification and quantification of fungal communities in freshwater habitats. The present volume also identifies the knowledge gaps and direction of future research in freshwater fungi.
Article
Full-text available
A comprehensive account of fungal classification from freshwater habitats is outlined and discussed in the present review based on literature of biodiversity studies and recent morpho-phylogenetic analyses. A total of 3,870 freshwater fungal species are listed with additional details on the isolation source, habitat, geographical distribution, and molecular data. The Ascomycota (2,968 species, 1,018 genera) dominated the freshwater fungal taxa wherein Sordariomycetes (823 species, 298 genera) had the largest number, followed by Dothideomycetes (677 species, 229 genera), Eurotiomycetes (276 species, 49 genera), and Leotiomycetes (260 species, 83 genera). Other phyla included in the updated classification of freshwater fungi are: Chytridiomycota (333 species, 97 genera), Rozellomycota (221 species, 105 genera), Basidiomycota (218 species, 100 genera), Blastocladiomycota (47 species, 10 genera), Monoblepharomycota (29 species, 6 genera), Mucoromycota (19 species, 10 genera), Aphelidiomycota (15 species, 3 genera), Entomophthoromycota (6 species, 4 genera), Mortierellomycota (5 species, 3 genera), Olpidiomycota (4 species, 1 genus), Zoopagomycota (3 species, 2 genera), and Sanchytriomycota (2 species, 2 genera). The freshwater fungi belong to 1,361 genera, 386 families and 145 orders. The Pleosporales and Laboulbeniaceae are the largest freshwater fungal order and family comprised of 391 and 185 species, respectively. The most speciose genera are Chitonomyces (87, Laboulbeniomycetes), Verrucaria (50, Eurotiomycetes), Rhizophydium (52, Rhizophydiomycetes), Penicillium (47, Eurotiomycetes), and Candida (42, Saccharomycetes).
Article
Full-text available
Fungi are an important and diverse component in various ecosystems. The methods to identify different fungi are an important step in any mycological study. Classical methods of fungal identification, which rely mainly on morphological characteristics and modern use of DNA based molecular techniques, have proven to be very helpful to explore their taxonomic identity. In the present compilation, we provide detailed information on estimates of fungi provided by different mycologistsover time. Along with this, a comprehensive analysis of the importance of classical and molecular methods is also presented. In orderto understand the utility of genus and species specific markers in fungal identification, a polyphasic approach to investigate various fungi is also presented in this paper. An account of the study of various fungi based on culture-based and cultureindependent methods is also provided here to understand the development and significance of both approaches. The available information on classical and modern methods compiled in this study revealed that the DNA based molecular studies are still scant, and more studies are required to achieve the accurate estimation of fungi present on earth.
Article
Full-text available
This is a continuity of a series of taxonomic and phylogenetic papers on the fungi where materials were collected from many countries, examined and described. In addition to extensive morphological descriptions and appropriate asexual and sexual connections, DNA sequence data are also analysed from concatenated datasets to infer phylogenetic relationships and substantiate systematic positions of taxa within appropriate ranks. Wherever new species or combinations are proposed, we apply an integrative approach using morphological and molecular data as well as ecological features wherever applicable. Notes on 112 fungal taxa are compiled in this paper including Biatriosporaceae and Roussoellaceae, Didysimulans gen. nov., 81 new species, 18 new host records and new country records, five reference specimens, two new combinations, and three sexual and asexual morph reports. The new species are Amanita cornelii, A. emodotrygon, Angustimassarina alni, A. arezzoensis, A. italica, A. lonicerae, A. premilcurensis, Ascochyta italica, A. rosae, Austroboletus appendiculatus, Barriopsis thailandica, Berkleasmium ariense, Calophoma petasitis, Camarosporium laburnicola, C. moricola, C. grisea, C. ossea, C. paraincrustata, Colletotrichum sambucicola, Coprinopsis cerkezii, Cytospora gelida, Dacrymyces chiangraiensis, Didysimulans italica, D. mezzanensis, Entodesmium italica, Entoloma magnum, Evlachovaea indica, Exophiala italica, Favolus gracilisporus, Femsjonia monospora, Fomitopsis flabellata, F. roseoalba, Gongronella brasiliensis, Helvella crispoides, Hermatomyces chiangmaiensis, H. chromolaenae, Hysterium centramurum, Inflatispora caryotae, Inocybe brunneosquamulosa, I. luteobrunnea, I. rubrobrunnea, Keissleriella cirsii, Lepiota cylindrocystidia, L. flavocarpa, L. maerimensis, Lophiotrema guttulata, Marasmius luculentus, Morenoina calamicola, Moelleriella thanathonensis, Mucor stercorarius, Myrmecridium fluviae, Myrothecium septentrionale, Neosetophoma garethjonesii, Nigrograna cangshanensis, Nodulosphaeria guttulatum, N. multiseptata, N. sambuci, Panus subfasciatus, Paraleptosphaeria padi, Paraphaeosphaeria viciae, Parathyridaria robiniae, Penicillium punicae, Phaeosphaeria calamicola, Phaeosphaeriopsis yuccae, Pleurophoma italica, Polyporus brevibasidiosus, P. koreanus, P. orientivarius, P. parvovarius, P. subdictyopus, P. ulleungus, Pseudoasteromassaria spadicea, Rosellinia mearnsii, Rubroboletus demonensis, Russula yanheensis, Sigarispora muriformis, Sillia italica, Stagonosporopsis ailanthicola, Strobilomyces longistipitatus, Subplenodomus galicola and Wolfiporia pseudococos. The new combinations are Melanomma populina and Rubroboletus eastwoodiae. The reference specimens are Cookeina tricholoma, Gnomoniopsis sanguisorbae, Helvella costifera, Polythrincium trifolii and Russula virescens. The new host records and country records are Ascochyta medicaginicola, Boletellus emodensis, Cyptotrama asprata, Cytospora ceratosperma, Favolaschia auriscalpium, F. manipularis, Hysterobrevium mori, Lentinus sajor-caju, L. squarrosulus, L. velutinus, Leucocoprinus cretaceus, Lophiotrema vagabundum, Nothophoma quercina, Platystomum rosae, Pseudodidymosphaeria phlei, Tremella fuciformis, Truncatella spartii and Vaginatispora appendiculata and three sexual and asexual morphs are Aposphaeria corallinolutea, Dothiora buxi and Hypocrella calendulina.
Article
Full-text available
While surveying the diversity of fungi of the order Mucorales, two isolates, EML-PUKI12-1 and EML-PUKI06-1, were obtained from the gut of soldier fly larvae inhabiting the bulrush at a pond located in the Chonnam National University Arboretum, Gwangju, Korea. The isolates were confirmed as Mucor irregularis and Mucor fragilis species, respectively, based on the morphological characteristics and phylogenetic analysis of rDNA internal transcribed spacer region. Such mucoralean species belonging to undiscovered taxa has not previously been described in Korea.
Article
Full-text available
Angiotensin I-converting enzyme (ACE) inhibitory activity of razor clam hydrolysates produced using five proteases, namely, pepsin, trypsin, alcalase, flavourzyme and proteases from Actinomucor elegans T3 was investigated. Flavourzyme hydrolysate showed the highest level of degree of hydrolysis (DH) (45.87%) followed by A. elegans T3 proteases hydrolysate (37.84%) and alcalase (30.55%). The A. elegans T3 proteases was observed to be more effective in generating small peptides with ACE-inhibitory activity. The 3 kDa membrane permeate of A. elegans T3 proteases hydrolysate showed the highest ACE-inhibitory activity with an IC50 of 0.79 mg/mL. After chromatographic separation by Sephadex G-15 gel filtration and reverse phase-high performance liquid chromatography, the potent fraction was subjected to MALDI/TOF-TOF MS/MS for identification. A novel ACE-inhibitory peptide (VQY) was identified exhibiting an IC50 of 9.8 μM. The inhibitory kinetics investigation by Lineweaver-Burk plots demonstrated that the peptide acts as a competitive ACE inhibitor. The razor clam hydrolysate obtained by A. elegans T3 proteases could serve as a source of functional peptides with ACE-inhibitory activity for physiological benefits.
Article
Full-text available
Notes on 113 fungal taxa are compiled in this paper, including 11 new genera, 89 new species, one new subspecies, three new combinations and seven reference specimens. A wide geographic and taxonomic range of fungal taxa are detailed. In the Ascomycota the new genera Angustospora (Testudinaceae), Camporesia (Xylariaceae), Clematidis, Crassiparies (Pleosporales genera incertae sedis), Farasanispora, Longiostiolum (Pleosporales genera incertae sedis), Multilocularia (Parabambusicolaceae), Neophaeocryptopus (Dothideaceae), Parameliola (Pleosporales genera incertae sedis), and Towyspora (Lentitheciaceae) are introduced. Newly introduced species are Angustospora nilensis, Aniptodera aquibella, Annulohypoxylon albidiscum, Astrocystis thailandica, Camporesia sambuci, Clematidis italica, Colletotrichum menispermi, C. quinquefoliae, Comoclathris pimpinellae, Crassiparies quadrisporus, Cytospora salicicola, Diatrype thailandica, Dothiorella rhamni, Durotheca macrostroma, Farasanispora avicenniae, Halorosellinia rhizophorae, Humicola koreana, Hypoxylon lilloi, Kirschsteiniothelia tectonae, Lindgomyces okinawaensis, Longiostiolum tectonae, Lophiostoma pseudoarmatisporum, Moelleriella phukhiaoensis, M. pongdueatensis, Mucoharknessia anthoxanthi, Multilocularia bambusae, Multiseptospora thysanolaenae, Neophaeocryptopus cytisi, Ocellularia arachchigei, O. ratnapurensis, Ochronectria thailandica, Ophiocordyceps karstii, Parameliola acaciae, P. dimocarpi, Parastagonospora cumpignensis, Pseudodidymosphaeria phlei, Polyplosphaeria thailandica, Pseudolachnella brevifusiformis, Psiloglonium macrosporum, Rhabdodiscus albodenticulatus, Rosellinia chiangmaiensis, Saccothecium rubi, Seimatosporium pseudocornii, S. pseudorosae, Sigarispora ononidis and Towyspora aestuari. New combinations are provided for Eutiarosporella dactylidis (sexual morph described and illustrated) and Pseudocamarosporium pini. Descriptions, illustrations and / or reference specimens are designated for Aposphaeria corallinolutea, Cryptovalsa ampelina, Dothiorella vidmadera, Ophiocordyceps formosana, Petrakia echinata, Phragmoporthe conformis and Pseudocamarosporium pini. The new species of Basidiomycota are Agaricus coccyginus, A. luteofibrillosus, Amanita atrobrunnea, A. digitosa, A. gleocystidiosa, A. pyriformis, A. strobilipes, Bondarzewia tibetica, Cortinarius albosericeus, C. badioflavidus, C. dentigratus, C. duboisensis, C. fragrantissimus, C. roseobasilis, C. vinaceobrunneus, C. vinaceogrisescens, C. wahkiacus, Cyanoboletus hymenoglutinosus, Fomitiporia atlantica, F. subtilissima, Ganoderma wuzhishanensis, Inonotus shoreicola, Lactifluus armeniacus, L. ramipilosus, Leccinum indoaurantiacum, Musumecia alpina, M. sardoa, Russula amethystina subp. tengii and R. wangii are introduced. Descriptions, illustrations, notes and / or reference specimens are designated for Clarkeinda trachodes, Dentocorticium ussuricum, Galzinia longibasidia, Lentinus stuppeus and Leptocorticium tenellum. The other new genera, species new combinations are Anaeromyces robustus, Neocallimastix californiae and Piromyces finnis from Neocallimastigomycota, Phytophthora estuarina, P. rhizophorae, Salispina, S. intermedia, S. lobata and S. spinosa from Oomycota, and Absidia stercoraria, Gongronella orasabula, Mortierella calciphila, Mucor caatinguensis, M. koreanus, M. merdicola and Rhizopus koreanus in Zygomycota.
Article
Full-text available
Fungi are well known by their ability to excrete enzymes into the environment. Among them, representatives of Mucor Fresen. have important biotechnological potential and some of them produce industrial enzymes. This work studied amylase, lipase, polygalacturonase and protease production by fifty-six isolates of Mucor belonging to 11 different taxa, selected from herbivores dung using solid media. The results showed that the majority of the isolates presented several enzymatic activities with predominance of polygalacturonase (96%), followed by amylase (84%), protease (82%) and lipase (66%).
Article
Procedures for the pure culture preparation of some fermented but nonalcoholic foods are outlined. These include tempeh, ragi, sufu, shoyu, ang-kak, tea fungus, and miso. The fungi used in the manufacture of fermented foods belong to the yeasts, Mucorales, Aspergillus, Penicillium, and Monascus, with the almost complete exclusion of all other fungi. As a supplement, a list of food fermentations and references to their literature in which fungi are involved are given.
Article
Two unique mucoralean strains were isolated from canine dung in the Western Cape. These strains could be recognized by short as well as tall sporangiophores, reaching lengths of up to 80 mm. Morphologically, they are similar to M. piriformis, which is a postharvest pathogen of stone and pome fruit. The strains also resemble M. grandis but can be distinguished from the latter by the presence of short sporangiophores. The coprophilous strains were also compared to other mucorelean fungi on the basis of genetic characteristics with special emphasis on species described in Mucor. For comparisons, we used the ITS region of the ribosomal DNA gene region as well as the Elongation factor 1-α and actin genes. Our results confirmed results from previous studies that species in Mucor do not form a monophyletic group when compared on a genetic level with other Zygomycetes. The ITS region proved to be unreliable in distinguishing between closely related species. The Elongation factor 1-α and actin genes, was found to be more informative as the different species could be distinguished from one another. Based on the morphological and molecular evidence we describe the coprophilous strain as a distinct taxon, namely Mucor renisporus sp. nov.
Article
We report the first case of chafer beetle [Anisoplia austriaca (Herbst 1783)] mortality caused by Actinomucor elegans var. elegans in wheat fields of the Kurdistan province, Iran. For three years, dead larvae of Anisoplia austriaca were collected from wheat fields of the Kurdistan province. Similar isolates of a fast-growing fungus were recovered from all samples. The fungal isolates were identified as A. elegans var. elegans based on morphological and cultural characteristics. The identity of the species was further confirmed using sequence data of the ITS (Internal Transcribed Spacer) region of ribosomal DNA. Koch's postulates were fulfilled by the inoculation of the larvae of A. austriaca and Galleria mellonella (Linnaeus, 1758) (as the model insect) using the spore suspension of A. elegans var. elegans. The viability of sporangiospores was evaluated using a spore dilution technique on germination medium. The results on the pathogenicity (100% mortality in A. austriaca larvae) and viability tests (germination: 95.45%) demonstrated that A. elegans var. elegans can be considered as a potential biocontrol agent against the chafer beetle. Field experiments are still required to evaluate the capacity of A. elegans as a biological control agent.