ArticlePDF Available

Coelomycetous Fungi in the Clinical Setting: Morphological Convergence and Cryptic Diversity

American Society for Microbiology
Journal of Clinical Microbiology
Authors:

Abstract and Figures

Human infections by coelomycetous fungi are becoming more frequent and range from superficial to systemic dissemination. Traumatic implantation of contaminated plant material is the most common cause. The typical morphological feature of these fungi is the production of asexual spores (conidia) within fruiting bodies called conidiomata. This study aimed to determine the distribution of the coelomycetes in clinical samples by a phenotypic and molecular study of a large set of isolates received from a USA reference mycological institution and by obtaining the in vitro antifungal susceptibility pattern of a selected group of strains against nine antifungals. A total of 230 isolates were identified by sequencing the D1 and D2 domains of the LSU nrRNA gene and by morphological characterization. Eleven orders of the phylum Ascomycota were identified: Pleosporales (the largest group; 66.1%), Botryosphaeriales (19.57%), Glomerellales (4.35%), Diaporthales (3.48%), Xylariales (2.17%), Hysteriales and Valsariales (0.87%), and Capnodiales , Helotiales , Hypocreales and Magnaporthales (0.43% each one). The most prevalent species were Neoscytalidium dimidiatum , Paraconiothyrium spp., Phoma herbarum , Didymella heteroderae and Epicoccum sorghinum . The most common anatomical site of isolation was superficial tissue (66.5%), followed by the respiratory tract (17.4%). Most of the isolates tested were susceptible to the majority of antifungals and only flucytosine showed poor antifungal activity.
This content is subject to copyright. Terms and conditions apply.
Coelomycetous Fungi in the Clinical
Setting: Morphological Convergence and
Cryptic Diversity
Nicomedes Valenzuela-Lopez,
a,b
Deanna A. Sutton,
c
José F. Cano-Lira,
a
Katihuska Paredes,
a
Nathan Wiederhold,
c
Josep Guarro,
a
Alberto M. Stchigel
a
Unitat de Micologia, Facultat de Medicina i Ciències de la Salut and IISPV, Universitat Rovira i Virgili, Reus,
Spain
a
; Microbiology Unit, Medical Technology Department, Faculty of Health Science, University of
Antofagasta, Antofagasta, Chile
b
; Fungus Testing Laboratory, University of Texas Health Science Center, San
Antonio, Texas, USA
c
ABSTRACT Human infections by coelomycetous fungi are becoming more frequent
and range from superficial to systemic dissemination. Traumatic implantation of con-
taminated plant material is the most common cause. The typical morphological fea-
ture of these fungi is the production of asexual spores (conidia) within fruiting bod-
ies called conidiomata. This study aimed to determine the distribution of the
coelomycetes in clinical samples by a phenotypic and molecular study of a large set
of isolates received from a U.S. reference mycological institution and by obtaining
the in vitro antifungal susceptibility pattern of nine antifungals against a selected
group of isolates. A total of 230 isolates were identified by sequencing the D1 and
D2 domains of the large subunit (LSU) nuclear ribosomal RNA (nrRNA) gene and by
morphological characterization. Eleven orders of the phylum Ascomycota were iden-
tified: Pleosporales (the largest group; 66.1%), Botryosphaeriales (19.57%), Glomerel-
lales (4.35%), Diaporthales (3.48%), Xylariales (2.17%), Hysteriales and Valsariales
(0.87%), and Capnodiales,Helotiales,Hypocreales and Magnaporthales (0.43% each).
The most prevalent species were Neoscytalidium dimidiatum,Paraconiothyrium spp.,
Phoma herbarum,Didymella heteroderae, and Epicoccum sorghinum. The most com-
mon anatomical site of isolation was superficial tissue (66.5%), followed by the respi-
ratory tract (17.4%). Most of the isolates tested were susceptible to the majority of
antifungals, and only flucytosine showed poor antifungal activity.
KEYWORDS Colletotrichum, coelomycetous fungi, coelomycetes, mycosis,
Neoscytalidium,Phoma,Pyrenochaeta, antifungal susceptibility
The coelomycetous fungi constitute a large number of taxa characterized by the
production of conidia (asexual propagules) within a cavity lined by fungal or host
tissue, called conidiomata (1), and although the majority of the human-opportunistic
infections are caused by fungi producing conidia on conidiophores (modified hyphae,
with one or more conidiogenous cells, which develop free on the substrate), a signif-
icant number of mycoses are produced by coelomycetous fungi (2–4). Coelomycetous
fungi are mostly saprobic and parasites of terrestrial vascular plants, but they can also
infect vertebrates and other fungi. They are ubiquitous in soil, in salty and freshwater
environments, and in sewage (4). Although the term Coelomycetes is still occasionally
used to refer to these fungi, this name is obsolete and is currently considered to refer
to an artificial fungal class. The class Coelomycetes is defined in terms of the morpho-
logical characterization of the asexual reproductive structures and considers the type
and the shape of their conidiomata and the ontogeny of their conidia as the most
useful characteristics (5, 6); the class has traditionally been divided into the orders
Melanconiales and Sphaeropsidales, depending upon the production of either acervular
Received 2 November 2016 Returned for
modification 21 November 2016 Accepted
29 November 2016
Accepted manuscript posted online 7
December 2016
Citation Valenzuela-Lopez N, Sutton DA,
Cano-Lira JF, Paredes K, Wiederhold N,
Guarro J, Stchigel AM. 2017. Coelomycetous
fungi in the clinical setting: morphological
convergence and cryptic diversity. J Clin
Microbiol 55:552–567. https://doi.org/
10.1128/JCM.02221-16.
Editor David W. Warnock, University of
Manchester
Copyright © 2017 American Society for
Microbiology. All Rights Reserved.
Address correspondence to José F. Cano-Lira,
jose.cano@urv.cat.
MYCOLOGY
crossm
February 2017 Volume 55 Issue 2 jcm.asm.org 552Journal of Clinical Microbiology
(cup-shaped) and pycnidial (globose to pyriform) conidiomata, respectively, and the
Pycnothyriales, characterized by the production of pycnothyrial (shield-shaped, flat-
tened, or hemispherical) conidiomata (5, 6). However, molecular studies have demon-
strated that the taxonomy of the Coelomycetes, represented by nearly 1,000 genera and
7,000 species (1), is artificial. Recent studies, have distributed the coelomycetes into at
least three classes of the phylum Ascomycota, i.e., Dothideomycetes,Leotiomycetes, and
Sordariomycetes (7–9).
Infections by coelomycetous fungi are mostly acquired by traumatic implantation of
plant/woody material or soil particles contaminated by their conidia rather than by
inhalation of air-dispersed propagules (2, 4). The coelomycetes are responsible for a
large variety of clinical entities, such as dermatitis, onychomycosis, keratitis, endoph-
thalmitis, subcutaneous phaeohyphomycosis, cysts, mycetoma, sinusitis, osteomyelitis,
bursitis, brain abscesses, and disseminated infections (4). The appropriate treatment of
the infections produced by these fungi is unknown, mainly due to the wide spectrum
of taxa involved and to the difficulties in their identification when the typical repro-
ductive structures are not produced. However, the European Society of Clinical Micro-
biology and Infectious Diseases (ESCMID) and the European Conference of Medical
Mycology (ECMM) have provided joint clinical guidelines for the management of
phaeohyphomycosis, with some recommendations for the treatment of infections due
to the most usual genera of coelomycetes, such as Neoscytalidium,Phoma, and Pyr-
enochaeta, mainly based on the use of amphotericin B and triazoles (10).
For the reasons mentioned above, the spectrum of species of these fungi in the
clinical setting is practically unknown (4, 11). Therefore, the objective of this study has
been to determine the distribution pattern of the coelomycetous fungi isolated from
clinical specimens from the United States using molecular identification of a large set
of isolates based on the sequencing of the D1 and D2 (D1-D2) domains of the large
subunit (LSU) of the nuclear ribosomal RNA (nrRNA) gene. In addition, we have
characterized those isolates morphologically and determined the antifungal suscepti-
bility of a representative number of them to nine antifungal drugs.
RESULTS
A total of 86 (38%) isolates of the 230 studied were able to produce pycnidial
conidiomata; 10 (4%) developed acervuli, and 35 (15%) produced the typical ana-
morphs of Neoscytalidium. The other 99 isolates (43%) remained sterile. The most
common species was Neoscytalidium dimidiatum, representing 15% (35/230) of the
isolates, followed by Paraconiothyrium cyclothyrioides with 7% (16/230), and both were
isolated mostly from superficial tissues. The third most common taxon recovered was
Total=230
19.57% Botryosphaeriales
0.43% Capnodiales
3.48% Diaporthales
4.35% Glomerellales
0.43% Helotiales
0.43% Hypocreales
0.87% Hysteriales
0.87% incertae sed is
0.43% Magnaporthales
66.10% Pleosporales
0.87% Valsariales
2.17% Xylariales
1
2
3
4
5
6
7
8
9
10
11 12
(1)
(2)
(3)
(4)
(5)
(6)
(7)
(8)
(9)
(10)
(11)
(12)
FIG 1 Distribution, by orders, of coelomycetous fungus isolates from clinical samples.
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 553
UTHSC DI16-306 (LN907449)
UTHSC DI16-307 (LN907450)
UTHSC DI16-302 (LN907445)
UTHSC DI16-294 (LN907437)
UTHSC DI16-282 (LN907425)
UTHSC DI16-255 (LN907398)
UTHSC DI16-228 (LN907371)
UTHSC DI16-212 (LN907355)
UTHSC DI16-205 (LN907348)
UTHSC DI16-204 (LN907347)
UTHSC DI16-200 (LN907343)
UTHSC DI16-199 (LN907342)
UTHSC DI16-308 (LN907451)
UTHSC DI16-319 (LN907462)
UTHSC DI16-365 (LN907508)
Phoma herbarum CBS 615.75 (EU754186)
UTHSC DI16-276 (LN880537)
Leptosphaeruli na australi s CBS 317.83 (EU754166)
UTHSC DI16-249 (LN907392)
UTHSC DI16-233 (LN907376)
UTHSC DI16-322 (LN907465)
UTHSC DI16-285 (LN907428)
UTHSC DI16-230 (LN907373)
UTHSC DI16-244 (LN907387)
UTHSC DI16-258 (LN907401)
UTHSC DI16-271 (LN907414)
UTHSC DI16-272 (LN907415)
UTHSC DI16-278 (LN907421)
UTHSC DI16-299 (LN907442)
UTHSC DI16-345 (LN907488)
Epicoc cum s orghinum CBS 179.80 (GU237978)
UTHSC DI16-338 (LN907481)
UTHSC DI16-301 (LN907444)
UTHSC DI16-288 (LN907431)
UTHSC DI16-280 (LN907423)
UTHSC DI16-257 (LN907400)
UTHSC DI16-206 (LN907349)
UTHSC DI16-202 (LN907345)
UTHSC DI16-201 (LN907344)
UTHSC DI16-197 (LN907340)
Epicoc cum ni grum CBS 173. 73T (GU237975)
UTHSC DI16-190 (LN907333)
UTHSC DI16-211 (LN907354)
UTHSC DI16-224 (LN907367)
UTHSC DI16-226 (LN907369)
UTHSC DI16-227 (LN907370)
UTHSC DI16-231 (LN907374)
UTHSC DI16-232 (LN907375)
UTHSC DI16-234 (LN907377)
UTHSC DI16-235 (LN907378)
UTHSC DI16-274 (LN907417)
UTHSC DI16-295 (LN907438)
UTHSC DI16-305 (LN907448)
Didymella heteroderae CBS 109.92T (GU238002)
UTHSC DI16-270 (LN907413)
1/72
0.99/76
0.05
0.99/-
0.98/-
sela
r
op
s
o
el
P
0.96/-
UTHSC DI16-291 (LN907434)
Asc ochyt a hordei var. hordei CBS 544.74 (E U754134)
UTHSC DI16-207 (LN907350)
UTHSC DI16-320 (LN907463)
UTHSC DI16-332 (LN907475)
UTHSC DI16-341 (LN907484)
UTHSC DI16-352 (LN907495)
UTHSC DI16-359 (LN907502)
UTHSC DI16-209 (LN907352)
Paraphoma radic ina CBS 111. 79T (KF251676)
UTHSC DI16-210 (LN907353)
Trematophoma sp. CBS 157.86 ( EU754221)
UTHSC DI16-296 (LN907439)
Paraphoma fimet i CBS 170. 70T (KF251674)
UTHSC DI16-324 (LN907467)
UTHSC DI16-260 (LN907403)
UTHSC DI16-264 (LN907407)
Edenia gomez pompae CBS 124106T (FJ839654)
Pleos pora herbarum CBS 191.86T (GU238160)
-/86
0.99/94
1/95
1/88
0.98/92
-/86
0.97/-
Neoascochyta desmazieri CBS 297.69T (KT389726)
Phoma clade I
Neoascochyta clade
Paraphoma clade
Pleospora clade
Didymella clade
Phoma clade II
FIG 2 Maximum-likelihood tree obtained from the D1-D2 of LSU (555 bp) sequences of the 322 strains, where 92 strains are
type or reference strains. In the tree, the branch lengths are proportional to phylogenetic distance. Bayesian posterior
probability scores of 0.95 and bootstrap support values of 70 are indicated on the nodes. The GenBank accession numbers
are given in parentheses. Saccharomyces castellii and S. cerevisiae were used to root the tree. The type (indicated by a
superscript T) and reference strains are shown in bold type.
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 554
Phoma herbarum (6.5%, 15/230) from superficial and respiratory tract specimens,
followed by Didymella heteroderae (5%, 12/230) and Epicoccum sorghinum (4%, 10/230),
which were isolated from superficial tissues.
In the D1-D2 phylogenetic analysis, the isolates were distributed into 11 orders (Fig.
1), most of which belonged to the Pleosporales (66.1%) and the Botryosphaeriales
UTHSC DI16-229 (LN907372)
Pyrenoc haeta unguis -hominis CBS 378 .92 (GQ387621)
UTHSC DI16-213 (LN907356)
UTHSC DI16-273 (LN907416)
Coniothyri um palmarum CB S 400. 71 (JX681084)
Pyrenoc haeta nobili s CBS 407. 76T (EU754206)
Leptosphaeria rubefac iens CBS 387.80 (JF740311)
UTHSC DI16-192 (LN907335)
UTHSC DI16-290 (LN907433)
Leptosphaeria et heridgei CBS 125980 (JF740291)
UTHSC DI16-238 (LN907381)
UTHSC DI16-203 (LN907346)
UTHSC DI16-189 (LN907332)
UTHSC DI16-236 (LN907379)
Coniothyri um telephii CBS 18 8.71 (GQ387599)
UTHSC DI16-313 (LN907456)
Diederichomy ces c ladoniic ola CBS 128025 (JQ238625)
UTHSC DI16-191 (LN907334)
UTHSC DI16-337 (LN907480)
UTHSC DI16-283 (LN907426)
Neosetophom a samarorum CBS 138.96T (GQ387578)
UTHSC DI16-240 (LN907383)
UTHSC DI16-325 (LN907468)
UTHSC DI16-330 (LN907473)
UTHSC DI16-336 (LN907479)
UTHSC DI16-339 (LN907482)
Parast agonospora nodorum CBS 259.49 (KF251688)
Phaeosphaeria ory zae CBS 110110T (KF251689)
Phaeosphaeria papay ae S528 (KF251690)
Phaeosphaeriops is m usae CBS 120026 (DQ885894)
UTHSC DI16-303 (LN907446)
UTHSC DI16-297 (LN907440)
UTHSC DI16-298 (LN907441)
UTHSC DI10-289 (LN907432)
UTHSC DI16-193 (LN907336)
UTHSC DI16-198 (LN907341)
UTHSC DI16-225 (LN907368)
UTHSC DI16-275 (LN907418)
UTHSC DI16-277 (LN907420)
Pyrenoc haetopsi s leptos pora CBS 101635T (GQ387627)
0.99/84
1/99
0.99/82
1/-
1/-
Medicopsis clade
Acrocalymma clade
Pyrenochaetopsis clade
Coniothyrium clade
al
l
e
o
s
s
u
oR e
dalc
Biatriospora
clade
Trematosphaeria
clade
Keissleriella clade
flavescens clade
Camarographium
clade
Phaeosphaeria clade
Neosetophoma italica MFLU 14 C0809 (KP711361)
Pyrenochaetopsis decipiens CBS 343.85T (GQ387624)
Pyrenochaetopsis indica CBS 124454T (GQ387626)
0.98/84
0.05
selaropsoelP
UTHSC DI16-195 (LN907338)
Acrocalymma walkeri CBS 257. 93T (FJ795454)
UTHSC DI16-315 (LN907458)
Medicops is rom eroi CBS 252.60T (EU754207)
UTHSC DI16-242 (LN907385)
UTHSC DI16-309 (LN907452)
UTHSC DI16-310 (LN907453) UTHSC DI16-220 (LN907363)
UTHSC DI16-356 (LN907499)
Arthopyrenia salicis CBS 368. 94 (AY538339)
UTHSC DI16-362 (LN907505)
UTHSC DI16-334 (LN907477)
Roussoel la nitidula MFLUCC 11 0182 (KJ474843)
Roussoella hysterioides CBS 546.94T (KF443381)
UTHSC DI16-269 (LN907412)
UTHSC DI16-292 (LN907435)
UTHSC DI16-300 (LN907443)
UTHSC DI16-360 (LN907503)
Roussoel la percutanea CB S 868.95 (KF 366449)
UTHSC DI16-342 (LN907485)
UTHSC DI16-241 (LN907384)
Biatri ospora mack innonii CB S 674.75T (GQ387613)
Biat riospora m arina CY 1 228 (GQ925848)
UTHSC DI16-335 (LN907478)
Trematosphaeria pert usa CBS 122368T (FJ201990)
UTHSC DI16-281 (LN907424)
UTHSC DI16-237 (LN907380)
UTHSC DI16-286 (LN907429)
UTHSC DI16-354 (LN907497)
Trematosphaeria gris ea CBS 120271 (K F015613)
UTHSC DI16-326 (LN907469)
Keissleriella cladophila CBS 104.55 ( JX681090)
UTHSC DI16-355 (LN907498)
Paraconiot hyrium flavescens CBS 178. 93 (GU238075)
UTHSC DI16-358 (LN907501)
Pseudoc haetosphaeronem a larense CBS 640.73T (KF015611)
UTHSC DI16-361 (LN907504)
Camarographium koreanum CBS 117159T (JQ044451)
1/99
1/94
1/99
-/73
1/99
1/99
1/99
1/99
1/99
1/91
0.95/96
1/-
0.95/-
1/-
FIG 2 (Continued)
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 555
UTHSC DI16-208 (LN907351)
Montagnula aloes CBS 132531 (NG042676)
UTHSC DI16-251 (LN907394)
Montagnula opul enta CBS 168. 34 (NG027581)
se
la
ro
psoelP
UTHSC DI16-187 (LN907330)
Lophiostoma heterosporum CBS 644. 86 (AY016369)
Sporormiella minima CBS 524.50 (DQ678056)
UTHSC DI16-253 (LN907396)
Preuss ia longis poropsis 16551-g (GQ203742)
UTHSC DI16-287 (LN907430)
Exos porium stylobat um CBS 160.3 0T (JQ044447)
UTHSC DI16-316 (LN907459)
Anteagloni um parvulum SMH5223 (GQ221909)
UTHSC DI16-343 (LN907486)
Phyllosticta flevolandica CBS 998. 72T (DQ377927)
UTHSC DI16-369 (LN907512)
Myrmaec ium rubric osum CB S 139068 (KP 687885)
UTHSC DI16-368 (LN907511)
UTHSC DI16-254 (LN907397)
Chaetophoma s p. CBS 119963 (E U754143)
UTHSC DI16-353 (LN907496)
UTHSC DI16-318 (LN907461)
Neofusicoc cum andi num CBS 117453 (DQ377914)
UTHSC DI16-221 (LN907364)
Lasiodipl odia parva CBS 456.78T (KF766362)
Lasiodipl odia theobromae CB S 287.47 (DQ377858)
UTHSC DI16-364 (LN907507)
UTHSC DI16-214 (LN907357)
UTHSC DI16-217 (LN907360)
Aplosporella sterculiae CBS 342. 78 (JX681073)
UTHSC DI16-250 (LN907393)
UTHSC DI16-248 (LN907391)
Phaeobotry osphaeria visci CBS 100163 (E U754215)
UTHSC DI16-333 (LN907476)
Botry osphaeria dot hidea CBS 115476 (NG027577)
UTHSC DI16-321 (LN907464)
UTHSC DI16-312 (LN907455)
-/72
1/99
1/74
0.99/91
-/92
1/84
1/99
1/99
1/99
1/99
1/99
1/99
1/96
1/91
1/89
-/96
1/99
1X
1X
1.5X
s
el
ai
reahp
so
y
r
t
oB
1/-
0.99/-
0.99/-
Gloniopsis subrugosa CBS 123346 (FJ161210)
Rhytidhysteron rufulum CBS 306.38 (FJ469672)
0.5X
0.5X
1X
UTHSC DI16-284 (LN907427)
Phaeodothis winteri CB S 182. 58 (GU301857)
1/99
Bimuria novae-zelandiae CBS 107.79 (AY016356)
UTHSC DI16-256 (LN907399)
Valsariales
Hysteriales
0.05
1/63
Letendraea eurotioi des CBS 212.31 (AY787935)
1/69
UTHSC DI16-351 (LN907494)
UTHSC DI16-370 (LN907513)
UTHSC DI16-267 (LN907410)
UTHSC DI16-239 (LN907382)
1/86
UTHSC DI16-266 (LN907409)
UTHSC DI16-348 (LN907491)
Paraconi othyrium brasiliens e CBS 254. 88 (JX496171)
UTHSC DI16-311 (LN907454)
UTHSC DI16-357 (LN907500)
Curreya pity ophila CB S 149. 32 (JX681087)
UTHSC DI16-219 (LN907362)
UTHSC DI16-261 (LN907404)
Paraphaeosphaeria neglecta CB S 124078T (JX496152)
UTHSC DI16-263 (LN907406)
UTHSC DI16-363 (LN907506)
Paraconiothyrium fuckeli i CBS 797.95 (JX496226)
Paraconiothyrium cyc lothy rioides CB S 972.95T (JX496232)
UTHSC DI16-265 (LN907408)
UTHSC DI16-215 (LN907358)
UTHSC DI16-216 (LN907359)
UTHSC DI16-218 (LN907361)
UTHSC DI16-222 (LN907365)
UTHSC DI16-243 (LN907386)
UTHSC DI16-246 (LN907389)
UTHSC DI16-252 (LN907395)
UTHSC DI16-268 (LN907411)
UTHSC DI16-279 (LN907422)
UTHSC DI16-314 (LN907457)
UTHSC DI16-327 (LN907470)
UTHSC DI16-328 (LN907471)
UTHSC DI16-346 (LN907489)
UTHSC DI16-347 (LN907490)
UTHSC DI16-349 (LN907492)
UTHSC DI16-367 (LN907510)
Paraconiothyrium cyc lothy rioides CB S 432.75 (JX496201)
Paraconiothyrium estuarinum CBS 109850T (JX496129)
0.98/97
Didymosphaeriaceae
clade
Lophiostoma
clade
Exosporium
clade
Anteaglonium clade
Phyllosticta clade
Neofusicoccum clade
Lasiodiplodia clade
Aplosporella clade
Phaeobotryosphaeria clade
Botryosphaeria clade
FIG 2 (Continued)
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 556
UTHSC DI14-331 (LN907310)
UTHSC DI14-332 (LN907311)
UTHSC DI14-336 (LN907315)
UTHSC DI14-306 (LN907285)
UTHSC DI14-307 (LN907286)
UTHSC DI14-308 (LN907287)
UTHSC DI14-309 (LN907288)
UTHSC DI14-310 (LN907289)
UTHSC DI14-311 (LN907290)
UTHSC DI14-312 (LN907291)
UTHSC DI14-313 (LN907292)
UTHSC DI14-314 (LN907293)
UTHSC DI14-315 (LN907294)
UTHSC DI14-316 (LN907295)
UTHSC DI14-317 (LN907296)
UTHSC DI14-318 (LN907297)
UTHSC DI14-319 (LN907298)
UTHSC DI14-320 (LN907299)
UTHSC DI14-321 (LN907300)
UTHSC DI14-322 (LN907301)
UTHSC DI14-323 (LN907302)
UTHSC DI14-324 (LN907303)
UTHSC DI14-325 (LN907304)
UTHSC DI14-326 (LN907305)
UTHSC DI14-327 (LN907306)
UTHSC DI14-328 (LN907307)
UTHSC DI14-329 (LN907308)
UTHSC DI14-330 (LN907309)
UTHSC DI14-333 (LN907312)
UTHSC DI14-334 (LN907313)
UTHSC DI14-335 (LN907314)
UTHSC DI14-337 (LN907316)
UTHSC DI14-338 (LN907317)
UTHSC DI14-339 (LN907318)
UTHSC DI14-340 (LN907319)
Neoscytalidium dimidiatum CBS 145. 78T (DQ377922)
UTHSC DI16-304 (LN907447)
Myc oleptodis cus terrest ris CB S 231. 53 (JN711859)
UTHSC DI16-196 (LN907339)
Cadophora fastigiat a DAOM 225754 (JN938877)
UTHSC DI16-245 (LN907388)
Pseudoc ercos pora vitis CBS 132012 (KF902011)
UTHSC DI16-350 (LN907493)
Phomatospora bellami nuta AFTOL-ID 766 (FJ176857)
UTHSC DI16-194 (LN907337)
UTHSC DI16-323 (LN907466)
UTHSC DI16-188 (LN907331)
Diatry pe dis ci formis CBS 197.49 (DQ470964)
Cryptos phaeria eunomia CBS 216.87 (KT425296)
UTHSC DI16-366 (LN907509)
UTHSC DI16-371 (LN907514)
1/99
1/86
1/100
1/96
-/73
1/99
1/99
0.99/81
1/100
1/100
1/100
0.96/75
1/99
2X
1X
2X
3X
1X
1X
4X
se
l
a
ir
ea
hpsoyrt
oB
Magnaporthales
Helotiales
Capnodiales
Incertae sedis
Xylariales
Pseudoc ercos pora oenother ae CBS 131885 (JQ324961)
1/96
0.99/90
Mycoleptodisc us indicus CBS 127677 (GU980697)
1/100
Peroneutypa scoparia MFLUCC 11-0615 (KU863140)
1/100
Protoventuria alpina CBS 140.83 (EU035444)
UTHSC DI16-329 (LN907472)
UTHSC DI16-331 (LN907474)
UTHSC DI16-317 (LN907460)
UTHSC DI16-293 (LN907436)
Diaporthe sclerotioides CBS 477 (A F439631)
UTHSC DI16-247 (LN907390)
UTHSC DI16-262 (LN907405)
UTHSC DI16-259 (LN907402)
UTHSC DI16-340 (LN907483)
Valsa ambiens CBS 1 09491 (EU255208)
UTHSC DI16-223 (LN907366)
Phialem oniopsis curvata CBS 490. 82T (KJ573448)
Phialem oniopsis ocularis CBS 110031 (K J573449)
UTHSC DI16-344 (LN907487)
Thyronec tria aus troameric ana A.R. 2808 (GQ505988)
UTHSC DI14-254 (LN907329)
Colletot richum gl oeosporioides CBS 79672 (AY 705727)
UTHSC DI14-250 (LN907325)
UTHSC DI14-248 (LN907323)
UTHSC DI14-245 (LN907320)
UTHSC DI14-249 (LN907324)
UTHSC DI14-246 (LN907321)
UTHSC DI14-252 (LN907327)
Colletot richum t runcatum CBS 112998 (JN940817)
UTHSC DI14-251 (LN907326)
UTHSC DI14-247 (LN907322)
Colletotrichum torulosum CBS 102667 (DQ286173)
UTHSC DI14-253 (LN907328)
Colletotrichum spaethianum CBS 101631 (JN940810)
Saccharomyces castellii NRRL Y 12630 (A Y048167)
Saccharomyces cerevisiae NRRL Y 12632 (AY 048154)
1/99
1/99
-/96
1/99
1/99
1/99
-/94
1/99
0.99/98
1/98
-/75
1/99
1/99
1/99
0.05
3X
1X
1X
2X
2X
se
l
a
h
trop
a
i
D
Incertae sedis
Hypocreales
s
e
la
l
l
er
e
molG
OUT GROUP
Saccharomycetales
1/-
Diatrype
clade
Peroneutypa
clade
Diaporthe
clade
Valsa
clade
Neoscytalidium
clade
Phomatospora
clade
Phialemoniopsis
clade
FIG 2 (Continued)
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 557
(19.57%), followed by the Glomerellales (4.35%), Diaporthales (3.48%), Xylariales (2.17%),
and Hysteriales and Valsariales (0.87% each). The orders Capnodiales,Helotiales,Hypo-
creales, and Magnaporthales were represented by only one isolate each (0.43%), and the
other isolates (0.87%) were incertae sedis (of uncertain taxonomic position).
Figure 2 shows the phylogenetic tree inferred from the analysis of 322 D1-D2
sequences corresponding to our set of isolates and numerous selected type or refer-
ence strains phylogenetically related to them. As mentioned above, the Pleosporales
contained the largest number of isolates (n152), which were distributed into 22
clades and belonged, probably, to 61 species of 44 different genera. These clades have
been named according to the first taxon historically described.
Within the Pleosporales,Phoma clade I (phylogenetically not supported) included 27
isolates, distributed mainly in two genera: Leptosphaerulina, with four isolates charac-
terized by a phoma-like asexual morph, which clustered with a reference strain of
Leptosphaerulina australis, and Phoma, with 15 isolates placed close to a reference strain
of Phoma herbarum and morphologically characterized by producing pycnidia and
hyaline aseptate conidia. The taxonomic position of the other eight isolates of this clade
was unresolved; in fact, they formed a separate, unsupported sister clade and displayed
a phoma-like asexual morph. The University of Texas Health Science Center (UTHSC)
isolate DI16-270 also showed the typical morphology of Phoma (Phoma clade II) but has
been placed phylogenetically distant from the mentioned genera and probably be-
longs to a new genus.
The Didymella clade included 22 isolates, 10 of which clustered with a reference
strain of Epicoccum sorghinum; unfortunately, the morphological features of these
isolates could not be studied because they produced only sterile mycelia in all the
culture media tested. Twelve isolates grouped with the type strain of Didymella
heteroderae, producing a phoma-like asexual morph, but were particularly character-
ized by the production of chlamydospores in long chains.
The Neoascochyta clade included seven isolates, six clustering with the type strain of
Neascochyta desmazieri and another one placed together with a reference strain of
Ascochyta hordei var. hordei. Morphologically, the species of this clade are mainly
characterized by the production of one-septate conidia that vary in size.
The Paraphoma clade contained only one isolate, which showed an identical
sequence to the type strain of Paraphoma radicina and morphologically was charac-
terized by setose (covered with bristle-like structures) pycnidia and hyaline aseptate
conidia.
The Pleospora clade was made up of five isolates and, with the exception of one of
them, was distributed into three well-supported sister clades corresponding to the
genera Edenia,Paraphoma, and Trematophoma. The isolate that clustered with the type
strain of Paraphoma fimeti was separate from the type species of Paraphoma (Para-
phoma radicina) and showed glabrous pycnidia instead the setose pycnidia produced
by the rest of the species. Interestingly, instead of the ellipsoidal, subhyaline conidia
typical of Edenia spp., the isolate UTHSC DI16-324 produced fusiform, hyaline, two- to
three-septate conidia that are probably indicative of a new genus. The other isolates of
this clade remained sterile.
The Coniothyrium clade included nine isolates, and its topology shows that the
genera Coniothyrium,Leptosphaeria, and Pyrenochaeta are clearly polyphyletic using
this conserved marker. Three of these isolates formed a well-supported sister clade
together with a reference strain of Coniothyrium telephii, which is characterized by
setose pycnidia. The other six isolates were distributed into the genera Leptosphaeria
and Pyrenochaeta. These had a pyrenochaeta-like anamorph, producing conidiophores
within pycnidia and hyaline aseptate conidia.
The Phaeosphaeria clade grouped nine isolates, with four of them clustering with
Neosetophoma and producing confluent pycnidia and small hyaline conidia. The other
five isolates were associated with the genera Diederichomyces,Parastagonospora,Pha-
eosphaeria, and Phaeosphaeriopsis. Only one isolate (UTHSC DI16-325), morphologically
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 558
resembling Phaeosphaeriopsis spp., was able to sporulate, displaying small conidio-
phores within pycnidia and one-septate, pigmented, variable-in-shape conidia.
The Pyrenochaetopsis clade included nine isolates, with four of them matching the
type strain of Pyrenochaetopsis leptospora, another four isolates forming a supported
sister clade separate from P. leptospora, and one not clustering to any of the type strains
included in the analysis. All of the isolates displayed the typical phoma-like morphol-
ogy, i.e., glabrous pycnidia and hyaline aseptate conidia, instead of setose pycnidia of
the genus Pyrenochaetopsis.
The five isolates assigned to the Acrocalymma and Medicopsis clades were grouped
with the type strains of Acrocalymma walkeri and Medicopsis romeroi, respectively, but
differed in 4.5% of the nucleotide sequences of the respective strains of reference.
These isolates remained sterile throughout.
The Roussoella clade was made up of eight isolates, two of which were associated
with a supposed reference strain of Arthopyrenia salicis (CBS 368.94), whose correct
identification was questioned by Liu et al. (12), and the remaining ones were associated
with Roussoella spp.; only three isolates were able to sporulate and had a morphology
similar to that of this genus, i.e., production of glabrous pycnidia and pigmented
aseptate conidia.
Two isolates nested in the Biatriospora clade but remained sterile. The Trematospha-
eria clade comprised five sterile isolates, two of which were phylogenetically related
with the type strain of Trematosphaeria pertusa and the rest of which were associated
with a reference strain of Trematosphaeria grisea.
The Keissleriella clade had only one isolate, which showed a phoma-like morphology
and clustered with a reference strain of Keissleriella cladophila. Another isolate was
associated with a reference strain of Paraconiothyrium flavescens and displayed a
morphology similar to that of Paraconiothyrium (pycnidia, phialidic conidiogenous cells,
and pigmented aseptate conidia); however, the taxonomic placement of that isolate
remains doubtful because it grouped phylogenetically distant from the type species of
the genus (Paraconiothyrium estuarinum). In the Camarographium clade, two sterile
isolates were located that were related to the genera Camarographium and
Pseudochaetosphaeronema.
The Didymosphaeriaceae clade comprised 33 isolates, of which 22 were phyloge-
netically related to Paraconiothyrium spp., 2 were related to Montagnula spp., and 2
were related to the type strain of Paraphaeosphaeria neglecta. Three isolates were
distributed into each of the genera Bimuria,Curreya, and Phaeodothis, and four isolates
formed a well-supported monophyletic sister clade separated from any known taxa of
the family. Only three isolates (UTHSC DI16-261, UTHSC DI16-266, and UTHSC DI16-363)
were able to sporulate, showing glabrous pycnidia and pale brown conidia displaying
morphological features similar to those of Paraconiothyrium spp. The Exosporium clade
comprised only two sterile isolates, one of which was related to the genus Preussia
while the other was related to Exosporium. The Anteaglonium,Lophiostoma, and
Phyllosticta clades comprised only one sterile isolate each one.
In the Valsariales clade, two isolates matched a reference strain of Myrmaecium
rubricosum. These isolates were characterized by producing free, well-differentiated
conidiophores instead of simply conidiogenous cells (phialides) inside the pycnidia.
The Hysteriales clade contained two sterile isolates, one related to an unidentified
strain of Chaetophoma and one of uncertain taxonomical placement but phylogeneti-
cally related to Chaetophoma,Gloniopsis, and Rhytidhysteron.
The second largest clade, corresponding to the order Botryosphaeriales, included 45
isolates distributed in six clades but mostly concentrated into the Neoscytalidium clade.
The fungi included in these clades were characterized by the production of stromatic
conidiomata (a hard, compact mass of cells or of vegetative hyphae), holoblastic
instead of phialidic conidiogenous cells, and aseptate, hyaline to brown, thick-walled
conidia. The Botryosphaeriales included the genera Botryosphaeria (three isolates),
Lasiodiplodia (two isolates), Neofusicoccum (one isolate), Aplosporella (two isolates), and
Phaeobotryosphaeria (two isolates). Additionally, 35 isolates of Neoscytalidium dimidia-
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 559
tum were also placed in this order. This fungus is characterized typically by the
production of holoarthric conidia (formed by disarticulation of the preexisting hyphae)
in chains.
The Capnodiales,Helotiales, and Magnaporthales clades each included only one
sterile isolate. Only the isolate of the Helotiales was not phylogenetically related to
any previously known described species. The isolate of the Capnodiales was closely
related to a reference strain of Pseudocercospora oenotherae. This genus is charac-
terized by producing stromata in the (plant) host, subhyaline to brown conidio-
phores, and small or large, subhyaline to brown conidia; unfortunately, our isolate
failed to sporulate. In the Magnaporthales, the isolate matched a reference strain of
Mycoleptodiscus indicus. This genus is characterized by producing sporodochia (a
cushion-like, densely aggregated group of conidiophores) and curved conidia; in
this case, the morphological study was not possible due to the absence of sporu-
lation of the isolate.
The Xylariales clade included five sterile isolates, three of which were related to the
genus Diatrype but phylogenetically distant from a reference strain of Diatrype disci-
formis. The remaining two isolates were associated with the Peroneutypa clade, with
one of them matching a reference strain of Peroneutypa scoparia and the other
uncertainly placed taxonomically.
The Diaporthales clade grouped eight isolates, six of which belonged to the Dia-
porthe clade and were characterized by the production of hyaline conidiophores within
pycnidial conidiomata, phialidic conidiogenous cells, and small conidia. The other two
sterile isolates were located in the Valsa clade.
The Hypocreales clade included only a single sterile isolate that matched a reference
strain of Thyronectria austroamericana.
The Glomerellales clade comprised 10 isolates, all of which belonged to the genus
Colletotrichum and were characterized by the production of acervular conidiomata,
phialidic conidiogenous cells, conidia variable in shape, and the presence of appres-
soria. Six of the isolates were identified as Colletotrichum gloeosporioides, two were
identified as Colletotrichum truncatum, and one was identified as Colletotrichum
spaethianum. One isolate (UTHSC DI14-247) was molecularly closely related to a refer-
ence strain of Colletotrichum torulosum.
Two isolates (UTHSC DI16-350 and UTHSC DI16-223) were not located in any of the
previously known orders and consequently were treated as incertae sedis. The first one
was assigned to the Phomatospora clade and the other, characterized by the produc-
tion of sporodochia and hyaline conidia, was identified as Phialemoniopsis curvata.
From a total of 224 clinical isolates, 153 were recovered mainly from superficial
tissues (epidermis and dermis) (66.5%), followed by 40 from the respiratory tract (17.4),
22 from miscellaneous deep tissues or fluids (9.6%), and 9 isolates from subcutaneous
tissues (3.9%) (Table 1).
Approximately half of all the fungi tested (44%; 101/230) were able to grow at 37°C
(Table 1); they were distributed within the orders at the following percentages: 100%
(10/10) of the Glomerellales, 100% (2/2) of Hysteriales, 100% (2/2) of the Valsariales, 98%
(44/45) of the Botryosphaeriales, 50% (1/2) of the isolates incertae sedis, and 28%
(42/152) of the Pleosporales.
Table 2 summarizes the results of the antifungal susceptibility testing. In general, all
the drugs tested, but especially terbinafine and amphotericin B, showed good activity
against the coelomycetous fungi, with terbinafine being the most active (geometric
mean [GM] of 0.04
g/ml; MIC
90
of 0.03
g/ml). Among the triazoles, itraconazole was
the least active, with an overall GM of 1
g/ml and a MIC
90
of 16
g/ml. Colletotrichum
gloeosporioides,Neoscytalidium dimidiatum, and Didymella heteroderae showed high
MICs for all the antifungals tested. Posaconazole and voriconazole demonstrated similar
in vitro potencies, with the only exceptions being activity against Colletotrichum
gloeosporioides and Neoascochyta desmazieri, for which the voriconazole GMs were 2.64
and 2
g/ml, respectively, and against Neoscytalidium dimidiatum, for which the
posaconazole GM was 2.26
g/ml. All the echinocandins showed good in vitro activity
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 560
against these fungi, with a GM of 0.06
g/ml. Flucytosine was the least active antifungal
tested, with elevated MICs against all isolates.
DISCUSSION
This is, to our knowledge, the largest taxonomic study on coelomycetous fungi of
clinical origin. It has demonstrated, based on DNA sequencing, a wider diversity of taxa
than previously reported. Although two recent reviews have reported approximately 35
species of coelomycetes involved in human infections (3, 4), the present study identifies
88 species; unfortunately, the role of many of them as pathogens for human still
remains uncertain because the clinical data of the patients are not allowed to be
TABLE 1 Anatomical sites of coelomycetous fungus isolates from clinical specimens
Order Clade
No. of isolates obtained from:
37°C
growth
Total no. of
isolates
Superficial
tissue
Subcutaneous
tissue
Deep
tissue/fluids
Respiratory
tract
Environment
and animal
Botryosphaeriales Aplosporella 22
Botryosphaeria 33
Lasiodiplodia 22
Neofusicoccum 11
Neoscytalidium 27 3 5 35
Phaeobotryosphaeria 22
Capnodiales 11
Diaporthales Diaporthe 3216
Valsa 112
Glomerellales 71 1 1 10
Helotiales 11
Hypocreales 11
Hysteriales 112
incertae sedis Phialemoniopsis 11
Phomatospora 11
Magnaporthales 11
Pleosporales Acrocalymma 11
Anteaglonium 11
Biatriospora 112
Camarographium 22
Coniothyrium 71 1 9
Didymella 17 1 4 22
Didymosphaeriaceae 26 1 2 3 1 33
Exosporium 112
flavescens 11
Keissleriella 11
Lophiostoma 11
Medicopsis 31 4
Neoascochyta 617
Paraphoma 11
Phaeosphaeria 21 1 4 1 9
Phoma I 13 3 10 1 27
Phoma II 1 1
Phyllosticta 11
Pleospora 1135
Pyrenochaetopsis 6129
Roussoella 5218
Trematosphaeria 41 5
Valsariales 22
Xylariales Diatrype 111 3
Peroneutypa 112
Total no. of isolates (%) 153 (66.5) 9 (3.9) 22 (9.6) 40 (17.4) 6 (2.6) 230 (100)
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 561
published. In general, the coelomycetous fungi are involved in many kinds of mycoses,
with superficial to deep infections, onychomycosis, cutaneous infections, keratitis, and
endophthalmitis being relatively frequent. In general, the most commonly reported
species clinically are Colletotrichum spp. (13–20), Neoscytalidium dimidiatum (21–25),
and Phoma spp. (11, 26–35). Our study partly confirms the data from previous studies
in which Neoscytalidium dimidiatum (approximately 15%), Paraconiothyrium cyclothyri-
oides (approximately 7%), and Phoma herbarum (approximately 6.5%) were the most
common species, having been recovered mainly from superficial tissue and respiratory
tract specimens. However, of these fungi, the only species that is relatively easy to
identify by phenotypic criteria is N.dimidiatum, which is the best known coelomycetous
fungus found clinically (22, 23, 36). The identification of the other fungi mentioned
above generally requires the use of molecular tools due to the difficulty of achieving in
vitro sporulation. Although Paraconiothyrium cyclothyrioides was relatively common in
our studied samples, there are only two clinical reports that refer to this species. Both
TABLE 2 Results of in vitro antifungal susceptibility testing of coelomycetous fungi
Taxon (no. of isolates) Parameter
a
Value for the drug (
g/ml)
b
AMB VRC ITC PSC AFG CFG MFG TRB 5FC
Neoascochyta desmazieri (5) GM 0.44 2 0.57 0.21 0.03 0.03 0.03 0.03 1.15
Range 0.25–1 1–4 0.25–1 0.06–0.5 0.03–0.06 0.03 0.03 0.03 0.5–2
MIC
90
0.5 2 1 0.5 0.03 0.03 0.03 0.03 2
Colletotrichum gloeosporioides (5) GM 0.57 2.64 8 0.87 0.03 0.03 0.03 0.03 16
Range 0.03–2 0.5–4 1–16 0.5–1 0.03 0.03 0.03 0.03 16
MIC
90
2 4 16 1 0.03 0.03 0.03 0.03 16
Epicoccum sorghinum (8) GM 0.25 0.92 0.59 0.30 0.03 0.04 0.03 0.03 2.97
Range 0.12–1 0.5–2 0.5–1 0.12–0.5 0.03–0.06 0.03–0.5 0.03 0.03 1–8
MIC
90
0.5 1 1 0.5 0.03 0.03 0.03 0.03 4
Neoscytalidium dimidiatum (16) GM 0.22 0.59 2.56 2.26 0.13 0.2 0.47 0.08 2.83
Range 0.06–1 0.03–16 0.06–16 0.03–16 0.03–0.5 0.03–1 0.06–8 0.03–2 0.25–16
MIC
90
0.5 4 16 16 0.25 0.5 4 0.03 8
Paraconiothyrium cyclothyrioides (15) GM 0.25 0.25 0.3 0.15 0.03 0.03 0.03 0.03 2.61
Range 0.03–8 0.06–0.5 0.06–0.5 0.03–0.5 0.03 0.03 0.03 0.03 1–16
MIC
90
0.5 0.5 0.5 0.25 0.03 0.03 0.03 0.03 4
Didymella heteroderae (11) GM 1.76 1.87 3.31 1.07 0.34 0.13 0.14 0.03 4
Range 0.5–8 0.06–16 0.5–16 0.5–2 0.03–8 0.03–4 0.03–2 0.03 1–16
MIC
90
4 16 16 2 8 4 2 0.03 16
Phoma herbarum (10) GM 0.43 0.57 0.81 0.40 0.04 0.04 0.03 0.03 2
Range 0.12–2 0.06–4 0.25–4 0.12–1 0.03–0.12 0.03–0.12 0.03–0.06 0.03 0.5–16
MIC
90
1 1 1 1 0.06 0.12 0.06 0.03 16
Phoma sp. (7) GM 0.1 0.17 0.17 0.14 0.03 0.03 0.03 0.03 1.78
Range 0.03–4 0.03–2 0.03–2 0.03–1 0.03 0.03 0.03 0.03 0.5–16
MIC
90
0.25 1 0.5 0.5 0.03 0.03 0.03 0.03 4
Diaporthe sclerotioides (4) GM 0.06 0.21 2 0.5 0.04 0.03 0.03 0.03 4
Range 0.03–0.12 0.12–0.25 1–4 0.5 0.03–0.06 0.03 0.03 0.03 0.5–16
MIC
90
0.12 0.25 2 0.5 0.03 0.03 0.03 0.03 8
Pyrenochaetopsis leptospora (4) GM 0.7 0.59 0.7 0.21 0.03 0.03 0.03 0.03 4
Range 0.03–4 0.25–2 0.06–16 0.03–1 0.03 0.03 0.03 0.03 0.5–16
MIC
90
2 1 1 0.5 0.03 0.03 0.03 0.03 16
Overall (85) GM 0.33 0.61 1 0.46 0.06 0.06 0.06 0.04 2.9
Range 0.03–8 0.03–16 0.03–16 0.03–16 0.03–8 0.03–4 0.03–8 0.03–2 0.25–32
MIC
90
2 4 16 16 0.25 0.5 2 0.03 16
a
GM, geometric mean; MIC
90
, drug concentration that inhibited 90% of isolates.
b
AMB, amphotericin B; VRC, voriconazole; ITC, itraconazole; PSC, posaconazole; AFG, anidulafungin; CFG, caspofungin; MFG, micafungin; TRB, terbinafine; 5FC,
flucytosine.
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 562
cases are from immunocompromised patients; in one case P.cyclothyrioides caused skin
lesions of the lower extremities, and in the second case it produced a systemic
coinfection together with Phaeoacremonium parasiticum (37, 38). Even though Phoma
sporulates easily, it is commonly misidentified as other related genera, such as Asco-
chyta, because the genera have similar morphologies, physiologies, and nucleotide
sequences (39, 40). Boerema et al. carried out one of the most comprehensive revisions
of the taxonomy of the genus Phoma. Using systematic criteria that predominated then,
approximately 220 species were accepted, distributed into nine sections (41). In a
recent multilocus study based on the sequence data of the 18S nrRNA (SSU) and LSU
genes, other authors demonstrated that such classification was totally artificial (42).
Currently, Phoma sensu stricto is included in the family Didymellaceae, and the other
Phoma-like fungi belong to other phylogenetic families, i.e., Cucurbitariaceae,Lepto-
sphaeriaceae,Phaeosphaeriaceae, etc. (39, 40, 42, 43).
It is of note that one of the frequently isolated species in our study, Didymella
heteroderae (5.2% of isolation frequency), has never been mentioned as an etiologic
agent of human infections even though our results reveal its ability to grow and to
sporulate at 37°C, which is uncommon in that genus and suggests its potential
pathogenicity.
An important clinical presentation of the coelomycetous fungi is eumycetoma,
which is restricted to a specific group of pleosporalean species of fungi, namely,
Medicopsis romeroi (formerly, Pyrenochaeta romeroi)(
44–46), Biatriospora mackinnonii
(formerly Pyrenochaeta mackinnonii)(
46), and Trematosphaeria grisea (formerly, Ma-
durella grisea)(
47–49), among others. However, in the present study only nine of the
isolates that were isolated from superficial and, less frequently, from deep tissues
belonged to these genera. This might be explained by the fact that the habitat of these
fungi is usually restricted to arid zones of East Africa and India and, occasionally, South
America (46, 50, 51).
Despite several studies in recent years devoted to infections by coelomycetous
fungi, little clinical data exist. The first well-documented review of human infections
caused by these fungi was carried out by Punithalingham (11), who referenced a total
of 12 species belonging primarily to the genera Botryodiplodia,Dothiorella,Hender-
sonula,Phoma,Phyllosticta,Pseudochaetosphaeronema, and Pyrenochaeta. In that work,
a morphological description of these taxa and their clinical origin was provided,
together with a dichotomous key for their identification. However, in our study, just
under 12% of the total isolates identified belonged to such genera. In a recent study,
Stchigel and Sutton (4) provided detailed information about the species of these fungi
isolated from clinical samples, described useful tools for their isolation and identifica-
tion, and gave general guidelines for infection management and treatment. These
authors concluded that these organisms are easy to isolate but that it was difficult to
induce in vitro fructification and sporulation. Our results are in agreement with theirs
because 43% of our isolates failed to sporulate, and it was only possible to identify
them and to determine their phylogenetic relationships by DNA sequencing.
The prevalence of coelomycetous fungi found in these clinical specimens—more
than 200 isolates recovered in a 9-year period— goes against the fact that so few
studies have described infections by them. This highlights the difficulty in conducting
a comprehensive study of these fungi and in establishing their real occurrence in
clinical settings. The taxonomy of these fungi is very complex because numerous
isolates are usually unable to sporulate in vitro or to produce different synanamorphs,
which sometimes predominate over the traditional coelomycete structures, making
their phenotypic recognition difficult; reliable identification can be done, therefore,
only by gene sequencing (9, 46, 52). However, even in this case, there are a very high
number of genera and species of coelomycetous fungi, and the phylogenetic bound-
aries of numerous taxa are still unresolved. Therefore, we carried out a phylogenetic
analysis of a large set of coelomycetous fungi using LSU sequences. This marker proved
useful for solving the phylogeny of most of the isolates included in the study, identi-
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 563
fying them, at least at genus level, and showing, in front of the internal transcribed
spacer (ITS), the advantage of an easy alignment of sequences.
The increasing use of molecular tools in fungal taxonomy has allowed the recog-
nition of numerous new taxa that are impossible to detect by traditional methods.
Recently, several new species of coelomycetous fungi, namely, Roussoella percutanea,
Truncatella angustata,Hongkongmyces pedis,Rhytidhysteron spp., Pseudochaetospha-
eronema martinelli, and Emarellia spp., have been involved in cases of subcutaneous
infections and eumycetoma (53–58), and some of our Pleosporales isolates, having
failed to sporulate, could represent new taxa.
Although clinical breakpoints for coelomycetous fungi have not been defined and
although in vitro antifungal susceptibility studies on these fungi are scarce, most of the
species seem to be inhibited by amphotericin B (4). Our results show that posaconazole
is the most active of the triazoles tested, and results for amphotericin B are similar in
vitro to those reported by Chowdhary et al. (10). Currently, only disseminated infections
due to N.dimidiatum have been conducted in animal models, and amphotericin B,
voriconazole, and posaconazole have been shown to be effective in the treatment of
this experimental mycosis (36). Guidelines for the management of infections due to
coelomycetous fungi include only a small group of taxa (Neoscytalidium,Phoma, and
Pyrenochaeta spp.) (10) although our study supports those protocols. A recent study by
Guégan et al. (59) analyzed several coelomycetous fungi that were implicated in human
mycosis and concluded that the surgical resection of infected tissues is advisable for
treating well-delimited lesions and that surgery together with new triazoles could be
used if lesions are extensive.
In conclusion, this study demonstrates that a wide variety of fungal taxa, identified
through their morphology as coelomycetous fungi, are involved in human infections in
the United States. However, more studies are necessary to understand the real preva-
lence of coelomycete infections throughout the world. The most active antifungal
drugs to treat them seem to be terbinafine, echinocandins, and amphotericin B, while
results for the azoles varied. Although the LSU gene sequence is useful for preliminary
identification and for establishing phylogenetic relationships between the majority of
coelomycetous fungi, future molecular studies testing a higher number genes are
essential to properly identify doubtful isolates at the species level.
MATERIALS AND METHODS
Fungal isolates and sequences. A total of 230 isolates of coelomycetous fungi were included in this
study, consisting of 224 from human clinical specimens, 3 from animal sources, and 3 from environ-
mental samples. All of the isolates were provided by the Fungus Testing Laboratory of the University of
Texas Health Science Center at San Antonio (UTHSC; San Antonio, Texas, USA). In addition, 92 D1-D2
sequences corresponding to type or reference strains were retrieved from GenBank and CBS databases
and included in the phylogenetic analysis.
Morphological and physiological characterization. For cultural characterization, the isolates were
grown on oatmeal agar (OA; 30 g of filtered oat flakes, 15 g of agar-agar, 1 liter of tap water) and malt
extract agar (MEA; 40 g of malt extract, 15 g of agar-agar, 1 liter of distilled water) at 20 1°C for 14 days
in darkness. The ability of the isolates to grow at 37°C was determined on potato dextrose agar (PDA;
Pronadisa, Madrid, Spain) after 7 days of incubation in darkness. The morphological features of the
vegetative and reproductive structures were studied using an Olympus CH2 light-field microscope
(Olympus Corporation, Tokyo, Japan) in wet mounts (on water and lactic acid) and slide cultures (isolates
grown on OA and MEA). The isolates were characterized phenotypically according to traditional criteria
(4, 5, 41, 60). Color standards are from Kornerup and Wanscher (61). Photomicrographs were taken with
an Axio-Imager M1 light-field microscope (Zeiss, Oberkochen, Germany).
DNA extraction, amplification, and sequencing. The total genomic DNA was extracted from
colonies grown on PDA after 7 days of incubation at 20 1°C, using a FastDNA kit protocol (Bio101; Vista,
CA) with a FastPrep FP120 instrument (Thermo Savant, Holbrook, NY) according to the manufacturer’s
protocol. DNA was quantified using a NanoDrop 2000 instrument (Thermo Scientific, Madrid, Spain). The
D1-D2 domains were amplified with the primer pair LR0R and LR5 (62). The amplicons were sequenced
in both directions with the same primer pair used for amplification at Macrogen Europe (Macrogen, Inc.,
Amsterdam, The Netherlands). The consensus sequences were obtained using SeqMan software, version
7.0.0 (DNAStar Lasergene, Madison, WI, USA).
Molecular identification and phylogenetic analysis. Preliminary molecular identification of the
isolates was made using the D1-D2 nucleotide sequences in blastn searches (https://blast.ncbi.nlm.nih.
gov/Blast.cgi) and the CBS database (www.cbs.knaw.nl). Only the sequences of type or reference strains
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 564
deposited in CBS/GenBank databases were considered for identification purposes. A level of identity of
98% was considered for species-level identification.
For the phylogenetic study, the sequences were aligned using the ClustalW application (63)ofthe
MEGA, version 6.06 (64), computer program, refined with MUSCLE (65), and manually adjusted using the
same software platform. Phylogenetic reconstructions were made by maximum-likelihood (ML) and
Bayesian inference (BI) with MEGA, version 6.06, and MrBayes, version 3.2.4 (66), respectively. The best
substitution model for the gene matrix (general time-reversal model incorporating invariable sites and
a discrete gamma distribution [GTRIG]) was estimated using MrModelTest, version 2.3 (67). For ML
analyses, a nearest-neighbor interchange was used as the heuristic method for tree inference. Support
for internal branches was assessed by 1,000 ML bootstrapped pseudoreplicates. Bootstrap support (BS)
of 70 was considered significant. For BI analyses, Markov chain Monte Carlo (MCMC) sampling was
carried out with 23 million generations, with samples taken every 1,000 generations. The 50% majority
rule consensus trees and posterior probability values (PP) were calculated after the first 25% of the
resulting trees was removed for burn-in. A PP value of 0.95 was considered significant. Saccharomyces
castellii (NRRL Y-12630; GenBank accession number AY048167) and Saccharomyces cerevisiae (NRRL
Y-12632; GenBank accession number AY048154) were used as outgroups.
Antifungal susceptibility testing. Using a broth microdilution reference method (68), the in vitro
antifungal susceptibilities of 85 isolates were determined of selected species of the genera Colletotri-
chum,Diaporthe,Didymella,Epicoccum,Neoascochyta,Neoscytalidium,Paraconiothyrium,Phoma sp., and
Pyrenochaetopsis. The following antifungals were tested: amphotericin B, voriconazole, posaconazole,
itraconazole, caspofungin, anidulafungin, micafungin, terbinafine, and flucytosine. The minimal effective
concentration (MEC) was determined after 48 h for the echinocandins, and the MIC was determined after
48 h and 72 h for the other drugs. Candida parapsilosis ATCC 22019 and Paecilomyces variotii ATCC
MYA-3630 were used as controls. The inocula for the coelomycetous fungi that did not sporulate were
prepared according to the method of Chowdhary et al. (69).
Accession number(s). The DNA sequences determined in this study have been deposited in
GenBank under accession numbers LN907285 to LN907514.
ACKNOWLEDGMENTS
This work was supported by the Spanish Ministerio de Economía y Competitividad,
grant CGL2013-43789-P.
We have no conflicts of interest to declare.
REFERENCES
1. Kirk PM, Cannon PF, Stalpers JA, Minter DW. 2008. Ainsworth & Bisby’s
dictionary of the fungi, 10th ed. CAB International, Wallingford, United
Kingdom.
2. Sutton DA. 1999. Coelomycetous fungi in human disease. A review:
clinical entities, pathogenesis, identification and therapy. Rev Iberoam
Micol 16:171–179.
3. Revankar SG, Sutton DA. 2010. Melanized fungi in human disease. Clin
Microbiol Rev 23:884 –928. https://doi.org/10.1128/CMR.00019-10.
4. Stchigel AM, Sutton DA. 2013. Coelomycete fungi in the clinical lab.
Curr Fungal Infect Rep 7:171–191. https://doi.org/10.1007/s12281
-013-0139-9.
5. Sutton BC. 1980. The Coelomycetes. Fungi Imperfecti with pycnidia,
acervuli and stromata. Commonwealth Mycological, Kew, England.
6. Nag Raj TR. 1993. Coelomycetous anamorphs with appendage-bearing
conidia. Mycologue Publications, Waterloo, Ontario, Canada.
7. Schoch CL, Crous PW, Groenewald JZ, Boehm EWA, Burgess TI, de
Gruyter J, de Hoog GS, Dixon LJ, Grube M, Gueidan C, Harada Y,
Hatakeyama S, Hirayama K, Hosoya T, Huhndorf SM, Hyde KD, Jones EBG,
Kohlmeyer J, Kruys A, Li YM, Lücking R, Lumbsch HT, Marvanová L,
Mbatchou JS, McVay AH, Miller AN, Mugambi GK, Muggia L, Nelsen MP,
Nelson P, Owensby CA, Phillips AJL, Phongpaichit S, Pointing SB, Pujade-
Renaud V, Raja HA, Plata ER, Robbertse B, Ruibal C, Sakayaroj J, Sano T,
Selbmann L, Shearer CA, Shirouzu T, Slippers B, Suetrong S, Tanaka K,
Volkmann-Kohlmeyer B, Wingfield MJ, Wood AR, Woudenberg JHC,
Yonezawa H, Zhang Y, Spatafora JW. 2009. A class-wide phylogenetic
assessment of Dothideomycetes. Stud Mycol 64:1–15. https://doi.org/
10.3114/sim.2009.64.01.
8. Maharachchikumbura SSN, Hyde KD, Groenewald JZ, Xu J, Crous PW.
2014. Pestalotiopsis revisited. Stud Mycol 79:121–186. https://doi.org/
10.1016/j.simyco.2014.09.005.
9. Wijayawardene NN, Hyde KD, Wanasinghe DN, Papizadeh M, Goo-
nasekara ID, Camporesi E, Bhat DJ, McKenzie EHC, Phillips AJL, Diederich
P, Tanaka K, Li WJ, Tangthirasunun N, Phookamsak R, Dai D-Q, Dissanay-
ake AJ, Weerakoon G, Maharachchikumbura SSN, Hashimoto A, Mat-
sumura M, Bahkali AH, Wang Y. 2016. Taxonomy and phylogeny of
dematiaceous coelomycetes. Fungal Divers 77:1–316. https://doi.org/
10.1007/s13225-016-0360-2.
10. Chowdhary A, Meis JF, Guarro J, de Hoog GS, Kathuria S, Arendrup MC,
Arikan-Akdagli S, Akova M, Boekhout T, Caira M, Guinea J, Chakrabarti A,
Dannaoui E, van Diepeningen A, Freiberger T, Groll AH, Hope WW,
Johnson E, Lackner M, Lagrou K, Lanternier F, Lass-Flörl C, Lortholary O,
Meletiadis J, Muñoz P, Pagano L, Petrikkos G, Richardson MD, Roilides E,
Skiada A, Tortorano AM, Ullmann AJ, Verweij PE, Cornely OA, Cuenca-
Estrella M. 2014. ESCMID and ECMM joint clinical guidelines for the
diagnosis and management of systemic phaeohyphomycosis: diseases
caused by black fungi. Clin Microbiol Infect 20:47–75. https://doi.org/
10.1111/1469-0691.12515.
11. Punithalingam E. 1979. Sphaeropsidales in culture from humans. Nova
Hedwigia 31:119–158.
12. Liu JK, Phookamsak R, Dai DQ, Tanaka K, Jones EG, Xu JC, Chukeatirote
E, Hyde K. 2014. Roussoellaceae, a new pleosporalean family to accom-
modate the genera Neoroussoella gen. nov., Roussoella and Roussoellop-
sis. Phytotaxa 181:1–33. https://doi.org/10.11646/phytotaxa.181.1.1.
13. Guarro J, Svidzinski TE, Zaror L, Forjaz MH, Gené J, Fischman O. 1998.
Subcutaneous hyalohyphomycosis caused by Colletotrichum gloeospori-
oides. J Clin Microbiol 36:3060 –3065.
14. Cano J, Guarro J, Gené J. 2004. Molecular and morphological identifica-
tion of Colletotrichum species of clinical interest. J Clin Microbiol 42:
2450 –2454. https://doi.org/10.1128/JCM.42.6.2450-2454.2004.
15. Castro LG, Da Silva Lacaz C, Guarro J, Gené J, Heins-Vaccari EM, de Freitas
Leite RS, Arriagada GL, Reguera MM, Ito EM, Valente NY, Nunes RS. 2001.
Phaeohyphomycotic cyst caused by Colletotrichum crassipes. J Clin Mi-
crobiol 39:2321–2324. https://doi.org/10.1128/JCM.39.6.2321-2324.2001.
16. Fernandez V, Dursun D, Miller D, Alfonso EC. 2002. Colletotrichum kera-
titis. Am J Ophthalmol 134:435– 438. https://doi.org/10.1016/S0002
-9394(02)01576-3.
17. Chakrabarti A, Shivaprakash MR, Singh R, Tarai B, George VK, Fomda BA,
Gupta A. 2008. Fungal endophthalmitis: fourteen years’ experience from
a center in India. Retina 28:1400 –1407. https://doi.org/10.1097/
IAE.0b013e318185e943.
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 565
18. Shivaprakash MR, Appannanavar SB, Dhaliwal M, Gupta A, Gupta S,
Gupta A, Chakrabarti A. 2011. Colletotrichum truncatum: An unusual
pathogen causing mycotic keratitis and endophthalmitis. J Clin Micro-
biol 49:2894–2898. https://doi.org/10.1128/JCM.00151-11.
19. Shiraishi A, Araki-Sasaki K, Mitani A, Miyamoto H, Sunada A, Ueda A,
Asari S, Zheng X, Yamamoto Y, Hara Y, Ohashi Y. 2011. Clinical charac-
teristics of keratitis due to Colletotrichum gloeosporioides. J Ocul Phar-
macol Ther 27:487– 491. https://doi.org/10.1089/jop.2011.0011.
20. Figtree M, Weeks K, Chan L, Leyton A, Bowes A, Giuffre B, Sullivan M,
Hudson BJ. 2013. Colletotrichum gloeosporioides sensu lato causing deep
soft tissue mycosis following a penetrating injury. Med Mycol Case Rep
2:40 43. https://doi.org/10.1016/j.mmcr.2013.01.003.
21. al-Rajhi AA, Awad AH, Al-Hedaithy SS, Forster RK, Caldwell KC. 1993.
Scytalidium dimidiatum fungal endophthalmitis. Br J Ophthalmol 77:
388 –390. https://doi.org/10.1136/bjo.77.6.388.
22. Elewski BE. 1996. Onychomycosis caused by Scytalidium dimidiatum.J
Am Acad Dermatol 35:336 –338. https://doi.org/10.1016/S0190
-9622(96)90664-7.
23. Madrid H, Ruíz-Cendoya M, Cano J, Stchigel A, Orofino R, Guarro J. 2009.
Genotyping and in vitro antifungal susceptibility of Neoscytalidium
dimidiatum isolates from different origins. Int J Antimicrob Agents 34:
351–354. https://doi.org/10.1016/j.ijantimicag.2009.05.006.
24. Machouart M, Menir P, Helenon R, Quist D, Desbois N. 2013. Scytalidium
and scytalidiosis: what’s new in 2012? J Mycol Med 23:40 46. https://
doi.org/10.1016/j.mycmed.2013.01.002.
25. Bakhshizadeh M, Hashemian HR, Najafzadeh MJ, Dolatabadi S, Zarrinfar
H. 2014. First report of rhinosinusitis caused by Neoscytalidium dimidia-
tum in Iran. J Med Microbiol 63:1017–1019. https://doi.org/10.1099/
jmm.0.065292-0.
26. Bakerspigel A. 1970. The isolation of Phoma hibernica from lesions on a
leg. Sabouraudia 7:261–264.
27. Punithalingam E. 1976. Phoma oculo hominis sp. nov. from corneal ulcer.
Trans Br Mycol Soc 67:142–143. https://doi.org/10.1016/S0007-1536
(76)80022-8.
28. Bakerspigel A, Lowe D, Rostas A. 1981. The isolation of Phoma eupyrena
from a human lesion. Arch Dermatol 117:362–363. https://doi.org/
10.1001/archderm.1981.01650060052024.
29. Shukla NP, Rajak RK, Agarwasl GP, Gupta D. 1984. Phoma minutispora as
a human pathogen. Mykosen 27:255–258.
30. Baker JG, Salkin IF, Forgacs P, Haines JH, Kemna ME. 1987. First report of
subcutaneous phaeohyphomycosis of the foot caused by Phoma minu-
tella. J Clin Microbiol 25:2395–2397.
31. Rai MK. 1989. Phoma sorghina infection in human being. Mycopatholo-
gia 105:167–170. https://doi.org/10.1007/BF00437250.
32. Rosen T, Rinaldi MJ, Tschen JA, Stern JK, Cernoch P. 1996. Cutaneous
lesions due to Pleurophoma (Phoma) complex. South Med J 89:431– 434.
https://doi.org/10.1097/00007611-199604000-00018.
33. Hirsh AH, Schiff TA. 1996. Subcutaneous phaeohyphomycosis caused by
an unusual pathogen: Phoma species. J Am Acad Dermatol 34:679 680.
https://doi.org/10.1016/S0190-9622(96)80083-1.
34. Tullio V, Banche G, Allizond V, Roana J, Mandras N, Scalas D, Panzone M,
Cervetti O, Valle S, Carlone N, Cuffini AM. 2010. Non-dermatophyte
moulds as skin and nail foot mycosis agents: Phoma herbarum,Chaeto-
mium globosum and Microascus cinereus. Fungal Biol 114:345–349.
https://doi.org/10.1016/j.funbio.2010.02.003.
35. Roehm CE, Salazar JC, Hagstrom N, Valdez TA. 2012. Phoma and Acre-
monium invasive fungal rhinosinusitis in congenital acute lymphocytic
leukemia and literature review. Int J Pediatr Otorhinolaryngol 76:
1387–1391. https://doi.org/10.1016/j.ijporl.2012.06.026.
36. Ruíz-Cendoya M, Madrid H, Pastor J, Guarro J. 2010. Evaluation of
antifungal therapy in a neutropenic murine model of Neoscytalidium
dimidiatum infection. Int J Antimicrob Agents 35:152–155. https://
doi.org/10.1016/j.ijantimicag.2009.09.028.
37. Gordon RA, Sutton DA, Thompson EH, Shrikanth V, Verkley GJ, Stielow
JB, Mays R, Oleske D, Morrison LK, Lapolla WJ, Galfione S, Tyring S,
Samathanam CA, Fu J, Wickes BL, Mulanovich V, Wanger A, Arias CA.
2012. Cutaneous phaeohyphomycosis caused by Paraconiothyrium cy-
clothyrioides. J Clin Microbiol 50:3795–3798. https://doi.org/10.1128/
JCM.01943-12.
38. Colombier MA, Alanio A, Denis B, Melica G, Garcia-Hermoso D, Levy B,
Peraldi MN, Glotz D, Bretagne S, Gallien S. 2015. Dual invasive infection
with Phaeoacremonium parasiticum and Paraconiothyrium cyclothyrioides
in a renal transplant recipient: Case report and comprehensive review of
the literature of Phaeoacremonium phaeohyphomycosis. J Clin Microbiol
53:2084 –2094. https://doi.org/10.1128/JCM.00295-15.
39. Aveskamp MM, de Gruyter J, Woudenberg JH, Verkley GJ, Crous PW.
2010. Highlights of the Didymellaceae: a polyphasic approach to char-
acterise Phoma and related pleosporalean genera. Stud Mycol 65:1– 60.
https://doi.org/10.3114/sim.2010.65.01.
40. Chen Q, Jiang JR, Zhang GZ, Cai L, Crous PW. 2015. Resolving the Phoma
enigma. Stud Mycol 82:137–217. https://doi.org/10.1016/j.simyco
.2015.10.003.
41. Boerema GH, de Gruyter J, Noordeloos ME, Hamers M. 2004. Phoma
identification manual. Differentiation of specific and infraspecific taxa in
culture. CABI Publishing, Cambridge, United Kingdom.
42. de Gruyter J, Aveskamp MM, Woudenberg JH, Verkley GJ, Groenewald
JZ, Crous PW. 2009. Molecular phylogeny of Phoma and allied anamorph
genera: towards a reclassification of the Phoma complex. Mycol Res
113:508 –519. https://doi.org/10.1016/j.mycres.2009.01.002.
43. de Gruyter J, Woudenberg JH, Aveskamp MM, Verkley GJ, Groenewald
JZ, Crous PW. 2013. Redisposition of Phoma-like anamorphs in Pleospo-
rales. Stud Mycol 75:1–36. https://doi.org/10.3114/sim0004.
44. Borelli D. 1979. Opportunistic fungi as producers of gray colonies and
mycetomata. Dermatologica 159:168 –174. https://doi.org/10.1159/
000250685.
45. Mathuram Thiyagarajan U, Bagul A, Nicholson ML. 2011. A nodulo-cystic
eumycetoma caused by Pyrenochaeta romeroi in a renal transplant
recipient: a case report. J Med Case Rep 5:460. https://doi.org/10.1186/
1752-1947-5-460.
46. Ahmed SA, Van De Sande WW, Stevens DA, Fahal A, Van Diepeningen
AD, Menken SB, de Hoog GS. 2014. Revision of agents of black-grain
eumycetoma in the order Pleosporales. Persoonia 33:141–154. https://
doi.org/10.3767/003158514X684744.
47. Butz WC, Ajello L. 1971. Black grain mycetoma: a case due to Madurella
grisea. Arch Dermatol 104:197–201. https://doi.org/10.1001/archderm
.1971.04000200085015.
48. Gulati V, Bakare S, Tibrewal S, Ismail N, Sayani J, Baghla DPS. 2012. A rare
presentation of concurrent Scedosporium apiospermum and Madurella
grisea eumycetoma in an immunocompetent host. Case Rep Pathol
2012:154201. https://doi.org/10.1155/2012/154201.
49. de Hoog GS, Van Diepeningen AD, Mahgoub ES, Van de Sande WWJ.
2012. New species of Madurella, causative agents of black-grain
mycetoma. J Clin Microbiol 50:988 –994. https://doi.org/10.1128/JCM
.05477-11.
50. Mackinnon JE, Ferrada-Urzúa LV, Montemayor L. 1943. Madurella grisea
n. sp. Mycopathologia 4:384 –393. https://doi.org/10.1007/BF01237166.
51. McGinnis MR. 1996. Mycetoma. Dermatol Clin 41:97–104.
52. Borman AM, Desnos-Ollivier M, Campbell CK, Bridge PD, Dannaoui E,
Johnson EM. 2016. Novel Taxa Associated with Human Fungal Black-
Grain Mycetomas: Emarellia grisea gen. nov., sp. nov., and Emarellia
paragrisea sp. nov. J Clin Microbiol 54:1738 –1745. https://doi.org/
10.1128/JCM.00477-16.
53. Ahmed SA, Stevens DA, Van de Sande WWJ, Meis JF, de Hoog GS. 2014.
Roussoella percutanea, a novel opportunistic pathogen causing subcu-
taneous mycoses. Med Mycol 52:689 698. https://doi.org/10.1093/
mmy/myu035.
54. Jagielski T, Zak I, Tyrak J, Bryk A. 2015. First probable case of subcuta-
neous infection due to Truncatella angustata: a new fungal pathogen of
humans? J Clin Microbiol 53:1961–1964. https://doi.org/10.1128/JCM
.00400-15.
55. Tsang CC, Chan JF, Trendell-Smith NJ, Ngan AH, Ling IW, Lau SK, Woo PC.
2014. Subcutaneous phaeohyphomycosis in a patient with IgG4-related
sclerosing disease caused by a novel ascomycete, Hongkongmyces pedis
gen. et sp. nov.: first report of human infection associated with the
family Lindgomycetaceae. Med Mycol 52:736 –747. https://doi.org/
10.1093/mmy/myu043.
56. Mahajan VK, Sharma V, Prabha N, Thakur K, Sharma NL, Rudramurthy SM,
Chauhan PS, Mehta KS, Abhinav C. 2014. A rare case of subcutaneous
phaeohyphomycosis caused by a Rhytidhysteron species: a clinico-
therapeutic experience. Int J Dermatol 53:1485–1489. https://doi.org/
10.1111/ijd.12529.
57. Mishra K, Das S, Goyal S, Gupta C, Rai G, Ansari MA, Saha R, Singal A.
2014. Subcutaneous mycoses caused by Rhytidhysteron species in an
immunocompetent patient. Med Mycol Case Rep 5:32–34. https://
doi.org/10.1016/j.mmcr.2014.07.002.
58. Ahmed SA, Desbois N, Quist D, Miossec C, Atoche C, Bonifaz A, de Hoog
GS. 2015. Phaeohyphomycosis caused by a novel species, Pseudochaeto-
Valenzuela-Lopez et al. Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 566
sphaeronema martinelli. J Clin Microbiol 53:2927–2934. https://doi.org/
10.1128/JCM.01456-15.
59. Guégan S, Garcia-Hermoso D, Sitbon K, Ahmed S, Moguelet P, Dromer F,
Lortholary O, French Mycosis Study Group. 2016. Ten-year experience of
cutaneous and/or subcutaneous infections due to coelomycetes in
France. Open Forum Infect Dis 3:ofw106. https://doi.org/10.1093/ofid/
ofw106.
60. de Hoog GS, Guarro J, Gené J, Figueras MJ. 2000. Atlas of clinical fungi,
2nd ed. Centraalbureau voor Schimmelcultures, Utrecht, The Nether-
lands.
61. Kornerup A, Wanscher JH. 1978. Methuen handbook of colour, 3rd ed.
Methuen, London, England.
62. Vilgalys R, Hester M. 1990. Rapid genetic identification and mapping of
enzymatically amplified ribosomal DNA from several Cryptococcus spe-
cies. J Bacteriol 172:4238 4246. https://doi.org/10.1128/jb.172.8.4238
-4246.1990.
63. Thompson JD, Higgins DG, Gibson TJ. 1994. CLUSTAL W: improving the
sensitivity of progressive multiple sequence alignment through se-
quence weighting, position-specific gap penalties and weight matrix
choice. Nucleic Acids Res 22:4673– 4680. https://doi.org/10.1093/nar/
22.22.4673.
64. Tamura K, Stecher G, Peterson D, Filipski A, Kumar S. 2013. MEGA6:
Molecular Evolutionary Genetics Analysis version 6.0. Mol Biol Evol
30:2725–2729. https://doi.org/10.1093/molbev/mst197.
65. Edgar RC. 2004. MUSCLE: multiple sequence alignment with high accu-
racy and high throughput. Nucleic Acids Res 32:1792–1797. https://
doi.org/10.1093/nar/gkh340.
66. Huelsenbeck JP, Ronquist F. 2001. MRBAYES: Bayesian inference of
phylogenetic trees. Bioinformatics 17:754 –755. https://doi.org/10.1093/
bioinformatics/17.8.754.
67. Nylander JA. 2004. MrModeltest v2. Uppsala University, Uppsala,
Sweden.
68. Clinical and Laboratory Standards Institute. 2008. Reference method for
broth dilution antifungal susceptibility testing of filamentous fungi;
approved standard, 2nd ed. Document M38-A2. Clinical and Laboratory
Standards Institute, Wayne, PA.
69. Chowdhary A, Kathuria S, Singh PK, Agarwal K, Gaur SN, Roy P, Rand-
hawa HS, Meis JF. 2013. Molecular characterization and in vitro antifun-
gal susceptibility profile of Schizophyllum commune, an emerging ba-
sidiomycete in bronchopulmonary mycoses. Antimicrob Agents
Chemother 57:2845–2848. https://doi.org/10.1128/AAC.02619-12.
Coelomycetous Fungi of Clinical Origin Journal of Clinical Microbiology
February 2017 Volume 55 Issue 2 jcm.asm.org 567
... The molecular identification of F. solani followed the DNA extraction methodology of Valenzuela-Lopez et al. [18], employing glass beads and adaptations for chitin cell wall lysis. After DNA extraction, electrophoresis was performed using a 1% verification gel to assess DNA quality. ...
... DNA quantification and purity determination were conducted using a nanodrop® spectrophotometer. For DNA amplification, Internal Transcribed Spacer (ITS), the gene encoding the IT-rDNA region, was targeted with the following nucleotide sequences: ITS 1 ( T C C G T A G G T G A A C C T G C G G) and ITS 4 ( T C C T C C G C T T A T T G A T A T G C), selected according to Khan et al. [19] and Valenzuela-Lopez et al. [18], confirming the primers' utility in fungal identification. ...
... p-NPP (p-nitrophenyl palmitate) (C 16 ) 482.30 ± 0.08 100 p-NPS (p-nitrophenyl stearate) (C 18 ) 443.70 ± 0. 01 91.99 * Carbonic acid chain of the fatty acid and its quantity in the substrate composition. Relative activity was calculated based on the fact that the activity with C16 as the substrate was defined as 100% from SmF in comparison with lipase F2 from FSS, and it was found that the latter exhibited higher activity. ...
Article
Full-text available
Omega-3 fatty acids, such as eicosapentaenoic acid (EPA) and docosahexaenoic acid (DHA), offer numerous health benefits. Enriching these fatty acids in fish oil using cost-effective methods, like lipase application, has been studied extensively. This research aimed to investigate F. solani as a potential lipase producer and compare its efficacy in enhancing polyunsaturated omega-3 fatty acids with commercial lipases. Submerged fermentation with coconut oil yielded Lipase F2, showing remarkable activity (215.68 U/mL). Lipase F2 remained stable at pH 8.0 (activity: 93.84 U/mL) and active between 35 and 70 °C, with optimal stability at 35 °C. It exhibited resistance to various surfactants and ions, showing no cytotoxic activity in vitro, crucial for its application in the food and pharmaceutical industries. Lipase F2 efficiently enriched EPA and DHA in fish oil, reaching 22.1 mol% DHA and 23.8 mol% EPA. These results underscore the economic viability and efficacy of Lipase F2, a partially purified enzyme obtained using low-cost techniques, demonstrating remarkable stability and resistance to diverse conditions. Its performance was comparable to highly pure commercially available enzymes in omega-3 production. These findings highlight the potential of F. solani as a promising lipase source, offering opportunities for economically producing omega-3 and advancing biotechnological applications in the food and supplements industry.
... The coelomycetous fungi involved in mycoses are poorly known due to complexity of their identification and relatively low frequency of infections. However, they are responsible for a large variety of clinical entities from superficial to deep and systemic mycoses [11]. Previously reported fish infections were mostly attributed to coelomycete genus Phoma and reviewed by Ř ehulka et al. [7]. ...
... Although our attempts to find characteristic fruiting bodies in isolated fungus were not successful, phylogenetic analysis clearly identified it as a recently described species of Neopyrenochaeta, N. submersa, described from plant debris in freshwater in Spain [12]. Problems with induction of sporulation in colelomomycetous fungi are relatively common and make their identification based on phenotype challenging or impossible [11]. In these cases, the use of molecular techniques is the only possibility to achieve reliable identification. ...
Article
Full-text available
A mycotic infection manifesting as abdominal distension with free serous fluid accumulation in the coelomic cavity is documented in farmed rainbow trout. Histological examination using PAS and silver staining revealed the presence of numerous fungal hyphae in the spleen and gastrointestinal wall. The isolated fungus was sterile and identified by using phylogenetic analysis based on four loci as Neopyrenochaeta submersa. This is the first time this fungus has been reported as pathogen.
... The biochemical DNA extraction protocol was used following the methodology of Valenzuela-Lopez et al. [20] with the addition of glass spherules and adaptations to perform cell lysis of the chitin wall. Subsequently, the quality of the DNA was checked, followed by DNA quantification and purity using a Nanodrop One C, Waltham, MA, USA). ...
Article
Full-text available
Cervical cancer is caused by a persistent and high-grade infection. It is caused by the Human Papillomavirus (HPV), which, when entering cervical cells, alters their physiology and generates serious lesions. HPV 18 is among those most involved in carcinogenesis in this region, but there are still no drug treatments that cause cure or total remission of lesions caused by HPV. It is known that L-asparaginase is an amidohydrolase, which plays a significant role in the pharmaceutical industry, particularly in the treatment of specific cancers. Due to its antitumor properties, some studies have demonstrated its cytotoxic effect against cervical cancer cells. However, the commercial version of this enzyme has side effects, such as hypersensitivity, allergic reactions, and silent inactivation due to the formation of antibodies. To mitigate these adverse effects, several alternatives have been explored, including the use of L-asparaginase from other microbiological sources, which is the case with the use of the fungus Aspergillus niger, a high producer of L-asparaginase. The study investigated the influence of the type of fermentation, precipitant, purification, characterization, and in vitro cytotoxicity of L-asparaginase. The results revealed that semisolid fermentation produced higher enzymatic activity and protein concentration of A. niger. The characterized enzyme showed excellent stability at pH 9.0, temperature of 50 • C, resistance to surfactants and metallic ions, and an increase in enzymatic activity with the organic solvent ethanol. Furthermore, it exhibited low cytotoxicity in GM and RAW cells and significant cytotoxicity in HeLa cells. These findings indicate that L-asparaginase derived from A. niger may be a promising alternative for pharmaceutical production. Its attributes, including stability, activity, and low toxicity in healthy cells, suggest that this modified enzyme could overcome challenges associated with antitumor therapy.
... The biochemical DNA extraction protocol was used following the methodology of Valenzuela-Lopez et al. [18] with the addition of glass spherules and adaptations to perform cell lysis of the chitin wall. Subsequently, the quality of the DNA was checked, followed by DNA quantification and purity using a Nanodrop One C spectrophotometer (Thermo Scientific). ...
Preprint
Full-text available
Cervical cancer is caused by a persistent and high-grade infection, caused by the Human Papilloma Virus (HPV), which, when entering cervical cells, alters their physiology and generates serious lesions. HPV 18 - is among those most involved in carcinogenesis in this region, but there are still no drug treatments that cause cure or total remission of lesions caused by HPV. Knowing that L-Asparaginase is an amidohydrolase, which plays a significant role in the pharmaceutical industry, particularly in the treatment of specific cancers, due to its antitumor properties, some studies have demonstrated its cytotoxic effect against cervical cancer cells. However, the commercial version of this enzyme has side effects, such as hypersensitivity, allergic reactions and silent inactivation due to the formation of antibodies. To mitigate these adverse effects, several alternatives have been explored, including the use of L-asparaginase from other microbiological sources, which is the case with the use of the fungus Aspergillus niger, a high producer of L-asparaginase. The study investigated the influence of the type of fermentation, precipitant, purification, characterization and in vitro cytotoxicity of L-asparaginase. The results revealed that semisolid fermentation produced higher enzymatic activity and protein concentration of A. niger. The characterized enzyme showed excellent stability at pH 9.0, temperature of 50ºC, resistance to surfactants and metallic ions, and an increase in enzymatic activity was also noted with the organic solvent ethanol. Furthermore, it exhibited low cytotoxicity in GM and RAW cells, and significant cytotoxicity in HeLa cells. These findings indicate that L-asparaginase derived from Aspergillus niger may be a promising alternative for pharmaceutical production. Its attributes, including stability, activity and low toxicity in healthy cells, suggest that this modified enzyme could overcome challenges associated with antitumor therapy.
... The species of the genus Rhizopus used in this work was isolated from the air, provided by the Mycology Laboratory of the Basic and Applied Immunology Center of the Federal University of Maranhão. The biochemical DNA extraction protocol was used according to Valenzuela -Lopez et al . ( 2017)with the addition of glass spherules and adaptations to carry out cell lysis of the chitin wall. The quality of the DNA was verified, followed by DNA quantification and purity in a Nanodrop spectrophotometer One C Thermo Scientific . Samples between 50-100 ƞg /µl were amplified via polymerase chain reaction. ...
Article
Full-text available
Fungi have been recognized as a great source of bioactive metabolites with great medical and pharmaceutical applicability. The objective of this work was to produce a fungal extract from Rhizopus spp and identify the secondary metabolites obtained, followed by preparation of a microemulsion formulated with Rhizopus sp extract; characterizing its antioxidant and antimicrobial activity. Among the metabolites obtained, sorbitol, linoleic acid, oleic acid, palmitic acid, and stearic acid stand out. The formulated microemulsion proved to be an effective pharmacological presentation in preserving the metabolites obtained from the fungal extract, and also potentiating in relation to the extract, when comparing the antioxidant activity and antimicrobial action.
... Our study found that relative abundances of the phylum Ascomycota, class Sordariomycetes, order Glomerellales were enriched in PE group. Glomerellales is a common opportunistic pathogenic fungus, which could activate cellular immunity [30]. ...
Article
Full-text available
Objective Studies recently acknowledged that the fungal community in the gut plays an important role in many inflammatory diseases of noninfectious origin. but the role of gut fungi in the pathogenesis of preeclampsia (PE) remains unknown. Methods We performed a case-case–control study to compare the gut mycobiota of PE, pregnancy with chronic hypertension (PCH), and the normal group (Normal pregnancy group without any underlying disease) by internal transcribed spacer sequencing. In addition, LC/MS was used to explore the relationship between the fecal metabolites and gut mycobiota. Results Compared with the PCH and the normal group, α diversity (represented the species abundance in a single sample) of mycobiota were lower in the PE group, but there was no statistically significant difference among these three groups. However, Linear discriminant analysis Effect Size (LEfSe) analysis found 3 differentially abundant fungal taxa in PE group when compared with the normal group. The gut metabolites of PE patients were significantly different from PCH and the normal group. Choline metabolism molecule glycerophosphocholine was the most discriminant metabolite between PE and the normal group. Correlation analysis found that Candida spp.was positively correlated with glycerophosphocholine which increased in PE. Conclusion We found that gut mycobiota changed in the third trimester of pregnant women with preeclampsia.
Article
Full-text available
The family Cucurbitariaceae is rich in species diversity and has a wide host range and geographic distribution. In this study, we identified 12 Cucurbitariaceae isolates which were obtained from disease symptoms in two forest trees in Khuzestan province, Iran. In addition, this family is reassessed using phylogenetic analyses based on DNA sequences from five nuclear regions (ITS, LSU, TUB2, TEF1α, and RPB2). The phylogenetic analyses showed that the present isolates represent one new genus, Nothocucurbitaria, and three new species, Allocucurbitaria galinsogisoli, Nothocucurbitaria izehica, and Parafenestella quercicola, which are described and illustrated. Furthermore, the genus Allocucurbitaria is emended to accommodate Seltsamia ulmi that grouped with the type species of Allocucurbitaria. Parafenestella pittospori and A. prunicola are recombined into the genera Neocucurbitaria and Nothocucurbitaria, respectively. Comparative analysis of single-locus trees revealed that the TUB2 and TEF1α can distinguish most genera and species in Cucurbitariaceae, while the ITS and LSU phylogenies show low resolution at both generic and species level. The best single-locus marker, RPB2, was able to distinguish all generic and most species lineages in Cucurbitariaceae.
Article
Full-text available
The study was conducted in the Faculty of Agricultural Engineering Sciences in the laboratories of the Plant Protection Department 2019, with the aim of diagnosing the pathogen of Branch Wilt and blackening of stems, outwardly and molecularly. The results of isolation from infected Malus domestica, Morus alba, Punica granatum, Citrus sinensis, C. aurantium, Ficus elastic, Ficus benjamina, Ricinus communis and Populus euphratica trees showed that twelve isolates of Neoscytalidium spp. were obtained. As a percentage of the frequency of the fungus in the visited orchards of apple trees (Al-Tarmiya and Saffronia) and berries (Al-Tarmiya) 88.9%, and in pomegranates (Al-Tarmiya) and sour oranges(Al-Jadriya) 66.6%, while the lowest frequency was recorded in orange trees (Al-Jadriya), which amounted to 33.3%. The results of the nucleotide sequencing of the isolates of the fungus Neoscytalidium spp. The presence of three types of fungi: N. novaehollandiae isolated from apple, Morus and pomegranate trees, Ficus, Castor and N. dimidiatum isolated from apple and Morus trees, Populus and N. hyalinum isolated from apple, orange, sour orange and ficus trees. The apple isolates 1, 2, and 3, Morus isolate 1, the pomegranate isolate, and the Ficus isolate showed a 100% congruence rate, while the Morus 2, orange, ficus, and Populus isolates showed a congruence rate of 99%, while the orange isolate congruence 98%, and the Castor isolate recorded a congruence rate of 97% with the isolates. Global registered in the International Gen Bank. The nucleotide sequences of the three species N. hyalinum, N. novaehollandiae , and N. dimidiatum in the International Genebank Organization, and this is the first record of these species in Iraq on the hosts isolated from them in this study.
Article
Full-text available
Pleosporales comprise a diverse group of fungi with a global distribution and significant ecological importance. A survey on Pleosporales (in Didymosphaeriaceae, Roussoellaceae and Nigrogranaceae) in Guizhou Province, China, was conducted. Specimens were identified, based on morphological characteristics and phylogenetic analyses using a dataset composed of ITS, LSU, SSU, tef 1 and rpb 2 loci. Maximum Likelihood (ML) and Bayesian analyses were performed. As a result, three new species ( Neokalmusia karka , Nigrograna schinifolium and N. trachycarpus ) have been discovered, along with two new records for China ( Roussoella neopustulans and R. doimaesalongensis ) and a known species ( Roussoella pseudohysterioides ). Morphologically similar species and phylogenetically close taxa are compared and discussed. This study provides detailed information and descriptions of all newly-identified taxa.
Article
Full-text available
This study documents the morphology and phylogeny of ascomycetes collected from karst landscapes of Guizhou Province, China. Based on morphological characteristics in conjunction with DNA sequence data, 70 species are identified and distributed in two classes (Dothideomycetes and Sordariomycetes), 16 orders, 41 families and 60 genera. One order Planisphaeriales, four families Leptosphaerioidaceae, Neoleptosporellaceae, Planisphaeriaceae and Profundisphaeriaceae, ten genera Conicosphaeria, Karstiomyces, Leptosphaerioides, Neoceratosphaeria, Neodiaporthe, Neodictyospora, Planisphaeria, Profundisphaeria, Stellatus and Truncatascus, and 34 species (Amphisphaeria karsti, Anteaglonium hydei, Atractospora terrestris, Conicosphaeria vaginatispora, Corylicola hydei, Diaporthe cylindriformispora, Dictyosporium karsti, Hysterobrevium karsti, Karstiomyces guizhouensis, Leptosphaerioides guizhouensis, Lophiotrema karsti, Murispora hydei, Muyocopron karsti, Neoaquastroma guizhouense, Neoceratosphaeria karsti, Neodiaporthe reniformispora, Neodictyospora karsti, Neoheleiosa guizhouensis, Neoleptosporella fusiformispora, Neoophiobolus filiformisporum, Ophioceras guizhouensis, Ophiosphaerella karsti, Paraeutypella longiasca, Paraeutypella karsti, Patellaria guizhouensis, Planisphaeria karsti, Planisphaeria reniformispora, Poaceascoma herbaceum, Profundisphaeria fusiformispora, Pseudocoleophoma guizhouensis, Pseudopolyplosphaeria guizhouensis, Stellatus guizhouensis, Sulcatispora karsti and Truncatascus microsporus) are introduced as new to science. Moreover, 13 new geographical records for China are also reported, which are Acrocalymma medicaginis, Annulohypoxylon thailandicum, Astrosphaeriella bambusae, Diaporthe novem, Hypoxylon rubiginosum, Ophiosphaerella agrostidis, Ophiosphaerella chiangraiensis, Patellaria atrata, Polyplosphaeria fusca, Psiloglonium macrosporum, Sarimanas shirakamiense, Thyridaria broussonetiae and Tremateia chromolaenae. Additionally, the family Eriomycetaceae was resurrected as a non-lichenized family and accommodated within Monoblastiales. Detailed descriptions and illustrations of all these taxa are provided.
Article
Full-text available
Fresh collections, type studies and molecular phylogenetic analyses of a multigene matrix of partial nuSSU-ITS-LSU rDNA, rpb2, tef1 and tub2 sequences were used to evaluate the boundaries of Cucurbitaria in a strict sense and of several related genera of the Cucurbitariaceae. Two species are recognised in Cucurbitaria and 19 in Neocucurbitaria. The monotypic genera Astragalicola, Cucitella, Parafenestella, Protofenestella, and Seltsamia are described as new. Fenestella is here included as its generic type F. fenestrata (= F. princeps), which is lecto- and epitypified. Fenestella mackenzei and F. ostryae are combined in Parafenestella. Asexual morphs of Cucurbitariaceae, where known, are all pyrenochaeta- or phoma-like. Comparison of the phylogenetic analyses of the ITS-LSU and combined matrices demonstrate that at least rpb2 sequences should be added whenever possible to improve phylogenetic resolution of the tree backbone; in addition, the tef1 introns should be added as well to improve delimitation of closely related species.
Article
Full-text available
A concatenated dataset of LSU, SSU, ITS and tef1 DNA sequence data was analysed to investigate the taxonomic position and phylogenetic relationships of the genus Camarosporium in Pleosporineae (Dothideomycetes). Newly generated sequences from camarosporium-like taxa collected from Europe (Italy) and Russia form a well-supported monophyletic clade within Pleosporineae. A new genus Camarosporidiella and a new family Camarosporidiellaceae are established to accommodate these taxa. Four new species, Neocamarosporium korfii, N. lamiacearum, N. salicorniicola and N. salsolae, constitute a strongly supported clade with several known taxa for which the new family, Neocamarosporiaceae, is introduced. The genus Staurosphaeria based on S. lycii is resurrected and epitypified, and shown to accommodate the recently introduced genus Hazslinszkyomyces in Coniothyriaceae with significant statistical support. Camarosporium quaternatum, the type species of Camarosporium and Camarosporomyces flavigena cluster together in a monophyletic clade with significant statistical support and sister to the Leptosphaeriaceae. To better resolve interfamilial/intergeneric level relationships and improve taxonomic understanding within Pleosporineae, we validate Camarosporiaceae to accommodate Camarosporium and Camarosporomyces. The latter taxa along with other species are described in this study.
Article
Full-text available
The familial placement of four genera, Mycodidymella, Petrakia, Pseudodidymella, and Xenostigmina, was taxonomically revised based on morphological observations and phylogenetic analyses of nuclear rDNA SSU, LSU, tef1, and rpb2 sequences. ITS sequences were also provided as barcode markers. A total of 130 sequences were newly obtained from 28 isolates which are phylogenetically related to Melanommataceae (Pleosporales, Dothideomycetes) and its relatives. Phylogenetic analyses and morphological observation of sexual and asexual morphs led to the conclusion that Melanommataceae should be restricted to its type genus Melanomma, which is characterised by ascomata composed of a well-developed, carbonaceous peridium, and an aposphaeria-like coelomycetous asexual morph. Although Mycodidymella, Petrakia, Pseudodidymella, and Xenostigmina are phylogenetically related to Melanommataceae, these genera are characterised by epiphyllous, lenticular ascomata with well-developed basal stroma in their sexual morphs, and mycopappus-like propagules in their asexual morphs, which are clearly different from those of Melanomma. Pseudodidymellaceae is proposed to accommodate these four genera. Although Mycodidymella and Xenostigmina have been considered synonyms of Petrakia based on sexual morphology, we show that they are distinct genera. Based on morphological observations, these genera in Pseudodidymellaceae are easily distinguished by their synasexual morphs: sigmoid, multi-septate, thin-walled, hyaline conidia (Mycodidymella); globose to ovoid, dictyosporus, thick-walled, brown conidia with cellular appendages (Petrakia); and clavate with a short rostrum, dictyosporus, thick-walled, brown conidia (Xenostigmina). A synasexual morph of Pseudodidymella was not observed. Although Alpinaria was treated as member of Melanommataceae in a previous study, it has hyaline cells at the base of ascomata and pseudopycnidial, confluent conidiomata which is atypical features in Melanommataceae, and is treated as incertae sedis.
Article
Full-text available
The Didymellaceae is one of the most species-rich families in the fungal kingdom, and includes species that inhabit a wide range of ecosystems. The taxonomy of Didymellaceae has recently been revised on the basis of multi-locus DNA sequence data. In the present study, we investigated 108 Didymellaceae isolates newly obtained from 40 host plant species in 27 plant families, and various substrates from caves, including air, water and carbonatite, originating from Argentina, Australia, Canada, China, Hungary, Israel, Italy, Japan, South Africa, the Netherlands, the USA and former Yugoslavia. Among these, 68 isolates representing 32 new taxa are recognised based on the multi-locus phylogeny using sequences of LSU, ITS, rpb2 and tub2, and morphological differences. Within the Didymellaceae, five genera appeared to be limited to specific host families, with other genera having broader host ranges. In total 19 genera are recognised in the family, with Heracleicola being reduced to synonymy under Ascochyta. This study has significantly improved our understanding on the distribution and biodiversity of Didymellaceae, although the placement of several genera still need to be clarified.
Article
Full-text available
The current paper represents the fourth contribution in the Genera of Fungi series, linking type species of fungal genera to their morphology and DNA sequence data. The present paper focuses on two genera of microfungi, Camarosporium and Dothiora, which are respectively epi- and neotypified. The genus Camarosporium is typified by C. quaternatum, which has a karstenula-like sexual morph, and phoma-like synasexual morph. Furthermore, Camarosporomyces, Foliophoma and Hazslinszkyomyces are introduced as new camarosporiumlike genera, while Querciphoma is introduced as a new phoma-like genus. Libertasomycetaceae is introduced as a new family to accommodate Libertasomyces and Neoplatysporoides. Dothiora, which is typified by D. pyrenophora, is shown to produce dothichiza- and hormonema-like synasexual morphs in culture, and D. cactacearum is introduced as a new species. In addition to their typification, ex-type cultures have been deposited in the Westerdijk Fungal Biodiversity Institute (CBS Culture Collection), and species-specific DNA barcodes in GenBank. Authors interested in contributing accounts of individual genera to larger multi-authored papers in this series should contact the associate editors listed on the List of Protected Generic Names for Fungi.
Article
Full-text available
During a survey of saprophytic microfungi on decomposing woody, herbaceous debris and soil from different regions in Southern Europe, a wide range of interesting species of asexual ascomycetes were found. Phylogenetic analyses based on partial gene sequences of SSU, LSU and ITS proved that most of these fungi were related to Sordariomycetes and Dothideomycetes and to lesser extent to Leotiomycetes and Eurotiomycetes. Four new monotypic orders with their respectively families are proposed here, i.e. Lauriomycetales, Lauriomycetaceae; Parasympodiellales, Parasympodiellaceae; Vermiculariopsiellales, Vermiculariopsiellaceae, and Xenospadicoidales, Xenospadicoidaceae. One new order and three families are introduced here to accommodate orphan taxa, viz. Kirschsteiniotheliales, Castanediellaceae, Leptodontidiaceae, and Pleomonodictydaceae. Furthermore, Bloxamiaceae is validated. Based on morphology and phylogenetic affinities Diplococcium singulare, Trichocladium opacum and Spadicoides atra are moved to the new genera Paradiplococcium, Pleotrichocladium and Xenospadicoides, respectively. Helicoon fuscosporum is accommodated in the genus Magnohelicospora. Other novel genera include Neoascotaiwania with the type species N. terrestris sp. nov., and N. limnetica comb. nov. previously accommodated in Ascotaiwania; Pleomonodictys with P. descalsii sp. nov. as type species, and P. capensis comb. nov. previously accommodated in Monodictys; Anapleurothecium typified by A. botulisporum sp. nov., a fungus morphologically similar to Pleurothecium but phylogenetically distant; Fuscosclera typified by F. lignicola sp. nov., a meristematic fungus related to Leotiomycetes; Pseudodiplococcium typified by P. ibericum sp. nov. to accommodate an isolate previously identified as Diplococcium pulneyense; Xyladictyochaeta typified with X. lusitanica sp. nov., a foliicolous fungus related to Xylariales and similar to Dictyochaeta, but distinguished by polyphialidic conidiogenous cells produced in setiform conidiophores. Other novel species proposed are Brachysporiella navarrica, Catenulostroma lignicola, Cirrenalia iberica, Conioscypha pleiomorpha, Leptodontidium aureum, Pirozynskiella laurisilvatica, Parasympodiella lauri and Zanclospora iberica. To fix the application of some fungal names, lectotypes and/or epitypes are designated for Magnohelicospora iberica, Sporidesmium trigonellum, Sporidesmium opacum, Sporidesmium asperum, Camposporium aquaticum and Psilonia atra.
Article
A description is provided for Phoma sorghina . Information is included on the disease caused by the organism, its transmission, geographical distribution, and hosts. HOSTS: Gramineae and all kinds of plants. Also isolated from soil, air and various animal sources. DISEASE: A minor leaf spot of cereals and grasses. The visible symptoms vary considerably; on sorghum leaves spots are usually irregular or rounded, yellowish-brown or grey with definite reddish-purple margins or indefinite in outline, reaching 1 cm or more in width. Pycnidia develop within spots on leaves, glumes and seeds. Also the fungus has been implicated with pre- and post-emergence death of seedlings of Macroptilium and Sylosanthes species (54, 1779) crown rot of bananas (61, 3556), leaf spot of Agave americana and stem rot of Euphorbia tirucalli (63, 3383), brown stem canker of Leucosperum cordifolium (56, 253). GEOGRAPHICAL DISTRIBUTION: A ubiquitous fungus occurring in tropical and subtropical regions. Africa (Botswana, Cameroon, Egypt, Ethiopia, Gambia, Kenya, Malawi, Mauritius, Mozambique, Nigeria, Senegal, Sierra Leone, South Africa, Sudan, Tanzania, Togo, Uganda, Zaire, Zambia, Zimbabwe); Asia (Bangladesh, Brunei, Burma, China, Hong Kong, India, Indonesia (Irian Jaya), Laos, Malaysia, Nepal, Pakistan, Saudi Arabia, Sri Lanka, Taiwan, USSR); Australasia and Oceania (Australia, Hawaii, New Zealand, Papua New Guinea, Solomon Islands); Europe (Germany, Portugal, Italy, UK); North America (Canada, USA); Central America and West Indies (Antigua, Guatemala, Honduras, Jamaica, Nicaragua, Puerto Rico, Trinidad); South America (Argentina, Bolivia, Brazil, Colombia). TRANSMISSION: Probably by contaminated seed; the fungus has been found on or isolated from several seed samples (1, 289; 33, 599; 47, 2153; 54, 1779; 60, 367; 61, 4102). In Taiwan P. sorghina has been found to be transmitted from seed to seedlings (62, 4281). The fungus has also been claimed to persist on trash and weed hosts and remain viable up to 1 yr but lose its viability after 2 yr storage on dry infected leaves (Koch & Rumbold, 1921).
Article
Halojulellaceae fam. nov. and Halojullela gen. nov. are introduced to accommodate Julella avicenniae, a marine species in the suborder Pleosporineae, order Pleosporales, Dothideomycetes. Justification for the new family is based on combined gene analysis of the large and small subunits of the nuclear ribosomal RNA genes (LSU, SSU) and two protein coding genes RPB2 and TEF1, as well as morphological characters. Halojulellaceae and Halojulella are characterized by immersed to semi-immersed, clypeate ascomata, with short, papillate ostioles, cellular, hyphae-like, pseudoparaphyses, 8-spored, fissitunicate, clavate to cylindrical asci with a well-developed apical apparatus, a moderately long pedicel with a club-like base and hyaline or golden brown, ellipsoidal, muriform ascospores and is typified by Halojulella avicenniae. Halojullela differs from Julella, (type J. buxi) in its marine habitat and distinctly differing ascus with the apical apparatus being well-developed and moderately long club-like pedicel. Morphological characters and molecular data show that H. avicenniae belongs in the Pleosporales, outside any of the known families, and thus a new family is introduced to accommodate it. Julella is maintained as a distinct genus which is presently most likely polyphyletic with saprobic and lichenized elements and needs further study as no molecular data is presently available for any species.