ArticlePDF Available

Diketopiperazines from Batnamyces globulariicola, gen. & sp. nov. (Chaetomiaceae), a fungus associated with roots of the medicinal plant Globularia alypum in Algeria

Authors:

Abstract and Figures

Eight diketopiperazines including five previously unreported derivatives were isolated from an endophytic fungus cultured from the medicinal plant Globularia alypum collected in Algeria. The strain was characterised by means of morphological studies and molecular phylogenetic methods and was found to represent a species of a new genus in the Chaetomiaceae, for which we propose the name Batnamyces globulariicola. The taxonomic position of the new genus, which appears phylogenetically related to Stolonocarpus and Madurella, was evaluated by a multi-locus genealogy and by morphological studies in comparison to DNA sequence data reported in the recent monographs of the family. The culture remained sterile on several culture media despite repeated attempts to induce sporulation, and only some chlamydospores were formed. After fermentation in submerged culture and extraction of the cultures with organic solvents, the major secondary metabolites of B. globulariicola were isolated and their chemical structures were elucidated by extensive spectral analysis including nuclear magnetic resonance (NMR) spectroscopy, high-resolution electrospray ionisation mass spectrometry (HRESIMS), and electronic circular dichroism (ECD) measurements. The isolated compounds were tested for their biological activities against various bacteria, fungi, and two mammalian cell lines, but only three of them exhibited weak cytotoxicity against KB3.1 cells, but no antimicrobial effects were observed.
Content may be subject to copyright.
ORIGINAL ARTICLE
Diketopiperazines from Batnamyces globulariicola,gen.&sp.nov.
(Chaetomiaceae), a fungus associated with roots of the medicinal
plant Globularia alypum in Algeria
Sara R. Noumeur
1,2
&Rémy B. Teponno
1,3
&Soleiman E. Helaly
1,4
&Xue-Wei Wang
5
&Daoud Harzallah
6
&
Jos Houbraken
7
&Pedro W. Crous
7
&Marc Stadler
1
Received: 13 October 2019 / Revised: 26 March 2020 /Accepted: 27 March 2020
Abstract
Eight diketopiperazines including five previously unreported derivatives were isolated from an endophytic fungus
cultured from the medicinal plant Globularia alypum collected in Algeria. The strain was characterised by means of
morphological studies and molecular phylogenetic methods and was found to represent a species of a new genus in the
Chaetomiaceae, for which we propose the name Batnamyces globulariicola. The taxonomic position of the new genus,
which appears phylogenetically related to Stolonocarpus and Madurella, was evaluated by a multi-locus genealogy and
by morphological studies in comparison to DNA sequence data reported in the recent monographs of the family. The
culture remained sterile on several culture media despite repeated attempts to induce sporulation, and only some
chlamydospores were formed. After fermentation in submerged culture and extraction of the cultures with organic
solvents, the major secondary metabolites of B. globulariicola were isolated and their chemical structures were eluci-
dated by extensive spectral analysis including nuclear magnetic resonance (NMR) spectroscopy, high-resolution
electrospray ionisation mass spectrometry (HRESIMS), and electronic circular dichroism (ECD) measurements. The
isolated compounds were tested for their biological activities against various bacteria, fungi, and two mammalian cell
lines, but only three of them exhibited weak cytotoxicity against KB3.1 cells, but no antimicrobial effects were
observed.
Keywords Diketopiperazines .New genus .New species .Phylogenetic methods .Sordariomycetes
Sara R. Noumeur and Rémy B. Teponno contributed equally tothis work.
Section Editor: Hans-Josef Schroers
Electronic supplementary material The online version of this article
(https://doi.org/10.1007/s11557-020-01581-9) contains supplementary
material, which is available to authorized users.
*Marc Stadler
marc.stadler@helmholtz-hzi.de
1
Department of Microbial Drugs, Helmholtz Centre for Infection
Research and German Centre for Infection Research (DZIF), partner
site Hannover/Braunschweig, Inhoffenstrasse 7,
38124 Braunschweig, Germany
2
Department of Microbiology and Biochemistry, Faculty of Natural
and Life Sciences, University of Batna 2, 05000 Batna, Algeria
3
Department of Chemistry, Faculty of Science, University of
Dschang, P.O. Box 67, Dschang, Cameroon
4
Department of Chemistry, Faculty of Science, Aswan University,
Aswan 81528, Egypt
5
State Key Laboratory of Mycology, Institute of Microbiology,
Chinese Academy of Sciences, No. 3, 1st Beichen West Road,
Chaoyang District, Beijing 100101, China
6
Laboratory of Applied Microbiology, Department of Microbiology,
Faculty of Natural and Life Sciences, University Sétif 1 Ferhat
Abbas, 19000 Sétif, Algeria
7
Westerdijk Fungal Biodiversity Institute, P.O. Box 85167, 3508
AD Utrecht, The Netherlands
Mycological Progress (2020) 19:589603
https://doi.org/10.1007/s11557-020-01581-9
#The Author(s) 2020
Introduction
In recent years, it has been demonstrated that fungal endo-
phytes are playing an important role in most ecosystems of
the world. These fungi, which colonise their host plants with-
out causing any disease symptoms, have been shown to rep-
resent various known and new phylogenetic lineages
(Blackwell and Vega 2018).
Moreover, several studies have led to the isolation and
structure elucidation of important secondary metabolites from
endophytic fungi, thus raising the prospect of using such or-
ganisms as alternative sources of lead compounds for devel-
opment of new drugs or agrochemical pesticides
(Suryanarayanan et al. 2009; Bills and Gloer 2016).
Moreover, it is now also being attempted to use these organ-
isms as biocontrol agents and biofertilisers (Hyde et al. 2019;
White et al. 2019).
In our ongoing search for bioactive fungal metabolites
from fungi associated with Algerian medicinal plants
(Noumeur et al. 2017; Teponno et al. 2017), 12 secondary
metabolites including five previously unreported
diketopiperazines (15) were isolated from the culture of a
fungus that was obtained from Globularia alypum
(Plantaginaceae) using a well-established protocol that has
been used for the isolation of endophytic fungi for several
decades. The producer organism turned out to be new to
Science and is formally described in the present paper.
Moreover, the isolation of its secondary metabolites, and their
structure elucidation and their preliminary biological charac-
terisation are reported.
Material and methods
Origin and isolation of the strain
Fresh healthy roots of the medicinal plant Globularia alypum
(Plantaginaceae) were collected in June 2015 from Ain Touta
(Batna, Algeria). Along with other cultures, strain CBS
144474 was isolated according to established protocols in-
volving surface disinfection (Noumeur et al. 2017; Teponno
et al. 2017) and was maintained in liquid nitrogen at the HZI
culture collection, Braunschweig, Germany. Morphology and
macroscopic features of the culture were determined on sev-
eral different culture media including yeast-malt-glucose
(YMG; Fig. 1), see Richter et al. (2016), and all other the
standard media that are currently in use for induction of spor-
ulation of Sordariomycetes, i.e. potato dextrose agar (PDA),
oatmeal agar (OA), synthetic nutrient-poor agar (SNA), and
malt extract agar (MEA) (cf. Crous et al. 2009;Wangetal.
2019b). Cardinal temperature requirements for growth were
only checked on YMG at temperatures ranging from 24 to
43 °C.
Molecular analysis and sequencing
Genomic DNA extraction from the fungal colonies growing
on YMG was performed using an EZ-10 Spin Column
Genomic DNA Miniprep kit (Bio Basic Canada Inc.,
Markham, Ontario, Canada) as specified by the manufacturer.
The internaltranscribed spacer 1 and 2 including the interven-
ing 5.8S nrDNA (ITS), the 28S nrDNA (LSU) including the
D1/D2 domains, a part of the DNA-directed RNA polymerase
II second largest subunit gene (RPB2), and the β-tubulin gene
(TUB2) were selected for phylogenetic inference. PCR and
generation of DNA sequences followed the procedure
outlined by Wendt et al. (2018).
After a BLAST result based on GenBank data, which allo-
cated the new fungus to the Chaetomiaceae, the phylogenetic
affinities were studied based on an analysis of a combined
ITS, LSU, RPB2,andTUB2 dataset in comparison with
DNA sequence data reported in the recent monographs of
the family (Wang et al. 2016a,b,2019a,b). Alignments were
madeusingthewebinterfaceMAFFTv.7(Katohand
Standley 2013), followed by manual adjustments with
MEGA v. 6 (Tamura et al. 2013). Phylogenetic analysis was
performed using maximum likelihood (ML) and Bayesian in-
ference (BI) approaches under RAxML-HPC2 on XSEDE
8.2.10 (Stamatakis 2014) using the Cipres Science gateway
portal (Miller et al. 2010) and MrBayes v. 3.2.6 (Ronquist
et al. 2012), respectively. For BI, the best evolutionary model
for each locus was determined using MrModeltest v. 2.0
(Nylander 2004). The maximum likelihood analysis used the
GTRGAMMA model. Obtained trees were viewed in FigTree
v. 1.1.2 (Rambaut 2009) and subsequently visually prepared
and edited in Adobe
®
Illustrator
®
CS6. Confident branch sup-
port is defined as Bayesian posterior probabilities (PP) 0.95
and maximum likelihood bootstrap values (ML-BS) 70%.
Instrumentation
Optical rotations were determined with a Perkin Elmer
(Überlingen, Germany) 241 MC polarimeter (using the sodi-
um D line and a quartz cuvette with a 10-cm path length and
0.5-mL volume). Circular dichroism (CD) spectra were re-
corded on a JASCO spectropolarimeter, model J-815
(JASCO, Pfungstadt, Germany). Nuclear magnetic resonance
(NMR) spectra were recorded on a Bruker (Bremen,
Germany) 500 MHz Avance III spectrometer with a BBFO
(plus) SmartProbe (
1
H 500 MHz,
13
C 125 MHz) and a Bruker
700 MHz Avance III spectrometer with a 5-mm TCI cryo-
probe (
1
H 700 MHz,
13
C 175 MHz), locked to the deuterium
signal of the solvent. Chemical shifts are given in parts per
million (ppm) and coupling constants in Hertz (Hz). Spectra
were measured at 24.8 °C in CD
3
OD, acetone-d
6
, and deuter-
ated chloroform; chemical shifts were referenced to residual
solvent signals with resonances at δ
H
/
C
3.31/49.15 for
590 Mycol Progress (2020) 19:589603
CD
3
OD, δ
H
/
C
2.05/29.92 for acetone-d
6
,andδ
H
/
C
7.24/77.23
for deuterated chloroform. High-performance liquid chroma-
tography coupled with diode array detector and mass spectro-
metric detection (HPLC-DAD/MS) analysis was performed
usinganamaZonspeedETDiontrapmassspectrometer
(Bruker Daltonics) in positive and negative ionisation modes.
The mass spectrometer was coupled to an Agilent 1260 series
HPLC-UV system (Agilent Technologies) (Santa Clara, CA,
USA) (column 2.1 × 50 mm, 1.7 μm, C18 Acquity uPLC
BEH (Waters); solvent A: H
2
O + 0.1% formic acid; solvent
B: acetonitrile (ACN) + 0.1% formic acid; gradient: 5% B for
0.5 min, increasing to 100% B in 20 min, maintaining
isocratic conditions at 100% B for 10 min, flow = 0.6 mL/
min, UVvis detection 200600 nm). High-resolution
electrospray (HR-ESI) MS spectra were recorded on a
maXis ESI TOF mass spectrometer (Bruker Daltonics) (scan
range m/z1002500, rate 2 Hz, capillary voltage 4500 V, dry
temperature 200 °C), coupled to an Agilent 1200 series
HPLC-UV system (column 2.1 × 50 mm, 1.7 μm, C18
Acquity uPLC BEH (Waters); solvent A: H
2
O + 0.1% formic
acid; solvent B: ACN + 0.1% formic acid; gradient: 5% B for
0.5 min, increasing to 100% B in 19.5 min, maintaining 100%
Bfor5min,FR=0.6mL/min,UVvis detection 200
600 nm). The molecular formulas were calculated including
the isotopic pattern (Smart Formula algorithm). Preparative
HPLC purification was performed at room temperature on
an Agilent 1100 series preparative HPLC system
(ChemStation software (Rev. B.04.03 SP1); binary pump sys-
tem; column: Kinetex 5u RP C18 100 Å, dimensions 250×
21.20 mm; mobile phase: ACN + 0.05% trifluoroacetic acid
(TFA) (solvent B) and water + 0.05% TFA (solvent A); flow
rate 20 mL/min; diode-array UV detector; 226 fraction
collector).
Fermentation, extraction, and isolation
Pieces of well-colonised agar of strain CBS 144474 from
YMG plates were inoculated in Q61/2 medium (Chepkirui
et al. 2019) in a 500-mL Erlenmeyer flask containing
200 mL of media and incubated at 23 °C for 7 days. This seed
culture was used to inoculate 25 other flasks (5 L) of the same
medium composition after homogenisation with a Heidolph
Silent Crusher. The flasks were incubated at 23 °C under con-
stant shaking at 140 rpm on a rotary shaker for 12 days. After
separationof the mycelia byvacuum filtration, the supernatant
was extracted with ethyl acetate (EtOAc).The EtOAc fraction
was dried over anhydrous Na
2
SO
4
, filtered, and concentrated
under vacuum to yield 590 mg of extract. The wet mycelia
were extracted twice with acetone, then with methanol in an
ultrasonic bath at 40 °C for 30 min. The resulting solution was
evaporated to yield an aqueous phase, which was further ex-
tracted with EtOAc (3 × 500 mL). After drying over anhy-
drous Na
2
SO
4
, the EtOAc fraction was concentrated under
vacuum to yield 416.7 mg of crude extract. In the meantime,
pieces of a well-grown YMG agar plate of strain CBS 144474
were inoculated in a YMG medium (Richter et al. 2016), in a
500-mL Erlenmeyer flask containing 200 mL of media, and
incubated at 23 °C for 8 days. The homogenised seed culture
was used to inoculate 25 other flasks (5 L) of the same medi-
um composition. The flasks were incubated at 23 °C under
constant shaking at 140 rpm on a rotary shaker for 11 days.
After separation, the supernatant and the mycelia were extract-
ed as described above to give 408.5 and 512.8 mg of extracts,
respectively.
The crude supernatant extract from the fermentation in
Q61/2 medium (ca. 400 mg) was purified by preparative
HPLC using a gradient of 2550% solvent B for 40 min,
50100% B for 10 min, and 100% B for 10 min. The fractions
were combined according to UVabsorption at 220, 280, and
325 nm and concurrent HPLC-MS analyses. Compounds 5
(0.9 mg; Rt = 11.12 min), 1(1 mg; Rt = 13.80 min), 2
(0.8 mg; Rt = 15.40 min), 8(0.8 mg; Rt = 21.34 min), 3
(1.2 mg; Rt = 25.42 min), 6(1.4 mg; Rt = 29.84 min), 7
(0.7 mg; Rt = 32.16 min), 4(0.7 mg; Rt = 36.12 min), and 9
(0.5 mg; Rt = 38.69 min) were eluted.
The supernatant extract from the fermentation in YMG
medium(ca.400mg)wasalsopurifiedbypreparative
HPLC. The gradient used was 535% solvent B for 40 min,
35100% solvent B for 20 min, and 100% B for 10 min. The
fractions were combined according to UV absorption at 220,
280, and 325 nm to yield compounds 10 (8.7 mg; Rt =
18.17 min), 11 (9.3 mg; Rt = 19.69 min), and 12 (3.7 mg;
Rt = 23.11 min).
Mycol Progress (2020) 19:589603 591
Summary of spectral data for the new compounds (15)
(3R,6Z)-3-Thiomethyl-6-[4-O-[(2E)-4-hydroxy-3-methylbut-
2-enyl]benzylidene]piperazine-2,5-dione (1): yellowish gum;
α½
25
D+22.0(c0.05, MeOH); UV (MeOH) λ
max
(log ε): 204
(4.37), 229 (4.26), 323 nm (4.34); CD (c0.5 mg/mL, EtOH)
λ
max
309 (+), 235 nm (+);
1
H NMR (CD
3
OD, 700 MHz) and
13
C NMR (CD
3
OD, 175 MHz) data (see Table 1); HRESIMS:
m/z 349.1217 [M + H]
+
(calcd for C
17
H
21
N
2
O
4
S
+
, 349.1217).
(3R,6Z)-3-Thiomethyl-6-[4-O-[(2Z)-4-hydroxy-3-methylbut-2-
enyl]benzylidene]piperazine-2,5-dione (2): yellowish gum; α½
25
D+37.5(c
0.04, MeOH); UV (MeOH) λ
max
(log ε): 203 (4.13), 229
(3.97), 323 nm (3.96); CD (c0.5 mg/mL, EtOH) λ
max
316
(+), 237 nm (+);
1
H NMR (CD
3
OD, 700 MHz) and
13
CNMR
(CD
3
OD, 175 MHz) data (see Table 1); HRESIMS: m/z
349.1217 [M + H]
+
(calcd for C
17
H
21
N
2
O
4
S
+
, 349.1217).
(3R,6Z)-3-Hydroxy-6-[4-O-(3-methylbut-2-
enyl)benzylidene]piperazine-2,5-dione (3): yellowish gum;
α½
25
D+26.7(c0.06, MeOH); UV (MeOH) λ
max
(log ε):
206 (4.22), 224 (4.26), 321 nm (4.15); CD (c0.5 mg/mL,
EtOH) λ
max
309 (+), 228 nm (+);
1
HNMR(CD
3
OD,
Table 2
13
Cand
1
H NMR spectroscopic data of compounds 4(MeOH-
d
4
)and5(Acetone d
6
)
Position 4 5
δ
13
C, type δ
1
Hδ
13
C, type δ
1
H
2 166.1, C / 165.2, C /
3 59.6, CH 4.97 s 59.0, CH 5.01 s
5 163.5, C / 165.1, C /
6 125.2, C / 69.0, C /
7 119.4, CH 6.87 s 43.0, CH
2
2.94 d (13.5)
3.60 d (13.5)
8 126.7, C / 128.1, C /
9/9132.1, CH 7.44 d (8.6) 133.1, CH 7.26 d (9.0)
10/10116.5, CH 6.99 d (8.6) 115.4, CH 6.83 d (9.0)
11 161.0, C / 159.3, C /
12 66.1, CH
2
4.58 br d (6.6) 64.8, CH
2
4.61 br d (6.3)
13 121.1, CH 5.46 m 122.7, CH 5.47 m
14 139.1, C / 141.3, C /
15 26.0, CH
3
1.79 s 61.5, CH
2
4.15 s
16 18.3, CH
3
1.76 s 21.4, CH
3
1.81 s
17 13.1, CH
3
2.26 s 10.6, CH
3
1.36 s
18 13.4, CH
3
2.23 s
Table 1
13
Cand
1
HNMR
spectroscopic data of compounds
13in methanol-d
4
Position 1 2 3
δ
C
, type δ
H
δ
C
,type δ
H
δ
C
,type δ
H
1 10.17 s
a
10.17 s
a
9.92 s
a
2 166.0, C / 166.0, C / 167.0, C /
3 59.6, CH 4.97 s 59.6, CH 4.97 s 76.1, CH 5.09 s
49.03brs
a
9.02 br s
a
8.92 d 3.9
a
5 163.5, C / 163.5, C / 163.8, C /
6 125.3, C / 125.3, C / 125.4, C /
7 119.3, CH 6.87 s 119.3, CH 6.86 s 119.5, CH 6.80 s
8 126.8, C / 126.8, C / 126.8, C /
9/9132.1, CH 7.45 d (8.6) 132.1, CH 7.45 d (8.6) 132.1, CH 7.46 d (8.7)
10/10116.4, CH 7.00 d (8.6) 116.5, CH 7.00 d (8.6) 116.4, CH 6.98 d (8.7)
11 160.9, C / 160.0, C / 160.9, C /
12 65.8, CH
2
4.68 br d (6.4) 65.3, CH
2
4.67 br d (6.4) 66.1, CH
2
4.58 br d (6.6)
13 120.8, CH 5.73 m 123.5, CH 5.57 m 121.1, CH 5.47 m
14 141.5, C / 141.8, C / 139.1, C /
15 67.9, CH
2
3.99 s 21.5, CH
3
1.85 br d (1.1) 26.0, CH
3
1.79 s
16 14.2, CH
3
1.76 s 61.8, CH
2
4,16 s 18.3, CH
3
1.76 s
17 13.1, CH
3
2.26 s 13.1, CH
3
2.26 s
a
Measured in DMSO-d
6
592 Mycol Progress (2020) 19:589603
700 MHz) and
13
CNMR(CD
3
OD, 175 MHz) data (see
Tab le 1); HRESIMS: m/z 303.1334 [M + H]
+
(calcd for
C
16
H
19
N
2
O
4
+
, 303.1339).
(3R,6Z)-3-Thiomethyl-6-[4-O-(3-methylbut-2-
enyl)benzylidene]piperazine-2,5-dione (4): yellowish gum;
α½
25
D+ 48.6 (c0.035, MeOH); UV (MeOH) λ
max
(log ε):
202 (4.19), 229 (4.02), 324 nm (4.00); CD (c0.5 mg/mL,
EtOH) λ
max
310 (+), 230 nm (+);
1
HNMR(CD
3
OD,
700 MHz) and
13
CNMR(CD
3
OD, 175 MHz) data (see
Tab le 2); HRESIMS: m/z 355.1088 [M + Na]
+
(calcd for
C
17
H
20
N
2
NaO
3
S
+
, 355.1092), 687.2282 [2 M + Na]
+
.
(3S,6R)-3,6-Bisthiomethyl-6-[4-O-[(2Z)-4-hydroxy-3-
methylbut-2-enyl]phenylmethyl]piperazine-2,5-dione (5):
yellowish gum; α½
25
D+ 33.3 (c0.045, MeOH); UV
(MeOH) λ
max
(log ε): 208 (4.16), 226 (4.06), 275 nm
(3.38); CD (c0.5 mg/mL, EtOH) λ
max
315 (+), 248 (),
220 nm ();
1
H NMR (CD
3
OD, 700 MHz) and
13
C NMR
(CD
3
OD, 175 MHz) data (see Table 2); HRESIMS: m/z
397.1249 [M + H]
+
(calcd for C
18
H
25
N
2
O
4
S
2
+
, 397.1250),
419.1068 [M + Na]
+
(calcd for C
18
H
24
N
2
O
4
S
2
Na
+
,
419.1075).
Screening for biological activities
The antimicrobial activity and the in vitro cytotoxicity (IC
50
)
were evaluated according to our previously reported proce-
dures (Sandargo et al. 2018; Teponno et al. 2017). Briefly,
minimum inhibitory concentrations (MICs) in μg/mL of the
isolated compounds were determined by serial dilution assays
against Schizosaccharomyces pombe DSM 70572, Pichia
anomala DSM 6766, Mucor hiemalis DSM 2656, Candida
albicans DSM 1665, Rhodotorula glutinis DSM 10134,
Mycol Progress (2020) 19:589603 593
Micrococcus luteus DSM 1790, Bacillus subtilis DSM 10,
Escherichia coli DSM 1116, Staphylococcus aureus DSM
346, Mycobacterium smegmatis ATCC 700084,
Chromobacterium violaceum DSM 30191, and
Pseudomonas aeruginosa DSM PA14. The assays were car-
ried out in 96-well microtiter plates in YMG media for fila-
mentous fungi and yeast and EBS for bacteria. Gentamycin,
kanamycin, nystatin, and oxytetracycline were used as posi-
tive control, and the negative control was methanol. The cy-
totoxicity against HeLa cells KB3.1 and mouse fibroblasts
L929 cells was determined by using the MTT (2-(4,5-dimeth-
ylthiazol-2-yl)-2,5-diphenyltetrazolium bromide) method in
96-well microplates. The cell lines were cultured in DMEM
(Gibco). Briefly, 60-μL aliquots of serial dilutions from an
initial stock of 1 mg/mL in MeOH of the test compounds were
added to 120-μL aliquots of a cell suspension (5 × 10
4
cells/
mL) in 96-well microplates. After 5 days of incubation, an
MTT assay was performed, and the absorbance was measured
at 590 nm using an ELISA plate reader (Victor). The concen-
tration at which the growth ofcells was inhibited to 50% of the
control (IC
50
) was obtained from the dose-response curves.
Epothilone B was used as a positive control, while methanol
was used as a negative control.
Results and discussion
Taxonomy
The fungus was isolated as sterile mycelia without any repro-
ductive structures, and its identification by phenotypic char-
acters was impossible by conventional methods. The mor-
phology and macroscopic features of the culture on the plate
were determined on YMG, PDA, and OA.
Molecular phylogeny
The concatenated alignment consisted of 65 taxa including
representatives of 35 genera in the Chaetomiaceae (cf.
Tab le 3). Five isolates representing four species of the family
Podosporaceae were selected as outgroups. The alignment
contained 3055 characters (including gaps) and is composed
of four partitions: 858 characters for RPB2, 961 characters for
TUB2, 664 characters for ITS, and 572 characters for LSU. Of
the total characters, 1573 were constant, 1220 were parsimo-
ny-informative, and 262 were parsimony-uninformative. For
the Bayesian inference, the GTR+I+G model was selected as
optimal for RPB2,TUB2, ITS, and LSU based on the result of
the MrModeltest. Isolate CBS 144474 was located in a sepa-
rate clade along with representatives of the genera
Stolonocarpus,Madurella,andCanariomyces (ML-BS =
100%, PP = 0.99), but could not be accommodated in either
of these genera (Fig. 2). Therefore, a novel genus Batnamyces
is proposed.
Batnamyces Noumeur, gen. nov. MB 832844
Etymologyin reference to the town in Algeria where the
type was collected.
DiagnosisDiffers from the genera Canariomyces,
Stolonocarpus,andMadurella, to which it appears phyloge-
netically most closely related, in the absence of sexual features
and conidiogenous structures, except for producing terminal
chains of hyphal chlamydospores.
Type species: Batnamyces globulariicola Noumeur, sp.
nov. MB 832845 (Fig. 1)
Typus:Algeria,Batna,AinToutafromrootsofGlobularia
alypum Plantaginaceae, June 2015, S. R. Noumeur (holotype
CBS H-23624), ex-type culture in CBS 144474; GenBank
Acc nos of DNA sequences: MT075917 (ITS/LSU),
MT075918 (RPB2), MT075919 (TUB2).
Colonies on YMG and OMA at 23 °C spread over the
whole 9-cm Petri dish after 14 days while they attain a diam-
eter of 53 mm on PDA, initially appearing dark brown, then
becoming covered with white patches with age (Fig. 1b).
Mycelium on SNA, with brown, thick-walled, smooth,
branched, septate, 56μm diam hyphae, giving rise to hya-
line, thin-walled, smooth, branched, septate, 1.52μmdiam
hyphae. Colonies remained sterile on SNA, PDA, OA, and
MEA (see images in Supplementary information). After sev-
eral transfers onto new OA agar plates, oval, chlamydospores
(512 × 8 μm) were formed in short chains, arising from the
hyphal tips (Fig. 1e). Even after several months of
subcultivation, no other conidiogenous structures or sexual
morph were observed either in the old or the newly inoculated
plates. The growth maximum was determined to be 2830 °C,
and no growth was observed at 37 °C.
Notes
The new genus Batnamyces is primarily defined
based on its molecular phylogeny, since we neither observed
the characteristic structures of the asexual nor sexual morph of
the species in this genus. Its classification in the family was
inferred from the molecular phylogeny that was established on
the basis of a multi-locus genealogy comprising representa-
tives of all important genera of Chaetomiaceae. The genera
Batnamyces,Canariomyces,Madurella, and Stolonocarpus
formed a single lineage (Fig. 2). Canariomyces species typi-
cally produce non-ostiolate ascomata together with single-
celled conidia arising from reduced conidiophores that are
reduced to a hyphal cell (cf. figs. 1922 in Wang et al.
2019b). Madurella species usually produce only sterile (non-
sporulating) hyphae and sparse aerial mycelium, growing re-
strictedly in culture and often producing buff, cinnamon, si-
enna, or orange exudates diffusing into the agar (cf. fig. 18 in
Wang et al. 2019b). On the other hand, Stolonocarpus is
characterised by non-ostiolate ascomata arising from a
Table 3 Details of strains used in this study
Current name Culture accession number
1
GenBank accession numbers
2
References
ITS LSU RPB2 TUB2
Chaetomiaceae
Achaetomium globosum CBS 332.67 TKX976570 KX976695 KX976793 KX976911 Wang et al. 2016a
Achaetomium strumarium CBS 333.67 TAY681204 AY681170 KC503254 AY681238 Cai et al. 2006,
Wang et al. 2016a
Acrophialophora nainiana CBS 100.60 TMK926793 MK926793 MK876755 MK926893 Wang et al. 2019b
CBS 417.67 MK926794 MK926794 MK876756 MK926894 Wang et al. 2019b
Amesia atrobrunnea CBS 379.66 TJX280771 JX280666 KX976798 KX976916 de Hoog et al. 2013,
Wang et al. 2016a
Amesia nigricolor CBS 600.66 TKX976578 KX976703 KX976802 KX976920 Wang et al. 2016a
Arcopilus aureus CBS 153.52 KX976582 KX976707 KX976806 KX976924 Wang et al. 2016a
Arcopilus flavigenus CBS 337.67 TKX976587 KX976712 KX976811 KX976929 Wang et al. 2016a
Batnamyces globulariicola CBS 144474 TPresent study
Botryotrichum murorum CBS 163.52 KX976591 KX976716 KX976815 KX976933 Wang et al. 2016a
Botryotrichum piluliferum CBS 654.79 KX976597 KX976722 KX976821 KX976939 Wang et al. 2016a
Brachychaeta variospora CBS 414.73 TMK926797 MK926797 MK876759 MK926897 Wang et al. 2019b
Canariomyces microsporus CBS 276.74 TMK926799 MK926799 MK876760 MK926899 Wang et al. 2019b
CBS 161.80 MK926800 MK926800 MK876761 MK926900 Wang et al. 2019b
Canariomyces notabilis CBS 548.83 TMK926802 MK926802 MK876763 MK926902 Wang et al. 2019b
Carteria arctostaphyli CBS 229.82 TMK926807 MK926807 MK876767 MK926907 Wang et al. 2019b
Chaetomium elatum CBS 142034 TKX976612 KX976733 KX976832 KX976954 Wang et al. 2016a
Chaetomium globosum CBS 160.62 TKT214565 KT214596 KT214666 KT214742 Wang et al. 2016b
Collariella bostrychodes CBS 163.73 KX976641 KX976738 KX976837 KX976983 Wang et al. 2016a
Collariella robusta CBS 551.83 TKX976652 KX976747 KX976846 KX976994 Wang et al. 2016a
Collariella virescens CBS 148.68 TKX976654 KX976749 KX976848 KX976996 Wang et al. 2016a
Condenascus tortuosus CBS 610.97 MK926817 MK926817 MK876777 MK926917 Wang et al. 2019b
Corynascus sepedonium CBS 111.69 THQ871751 KX976777 HQ871827 KX977027 van den Brink et al. 2012,
Wang et al. 2016a
Chrysanthotrichum allolentum CBS 644.83 TMK926808 MK926808 MK876768 MK926908 Wang et al. 2019b
Chrysanthotrichum lentum CBS 339.67 TMK926809 MK926809 MK876769 MK926909 Wang et al. 2019b
Chrysocorona lucknowensis CBS 727.71 eT MK926813 MK926813 MK876773 MK926913 Wang et al. 2019b
CBS 385.66 MK926816 MK926816 MK876776 MK926916 Wang et al. 2019b
Corynascella humicola CBS 337.72 TKX976656 KX976751 KX976850 KX976998 Wang et al. 2016a
CBS 379.74 KX976657 KX976752 KX976851 KX976999 Wang et al. 2016a
Dichotomopilus funicola CBS 159.52 TGU563369 GU563354 KX976856 JF772461 Wang et al. 2016a
Dichotomopilus indicus CGMCC 3.14184 TGU563367 GU563360 KX976861 JF772453 Wang et al. 2016a
Floropilus chiversii CBS 558.80 TMK926818 MK926818 MK876778 MK926918 Wang et al. 2019b
Humicola fuscoatra CBS 118.14 TLT993579 LT993579 LT993498 LT993660 Wang et al. 2019a
Hyalosphaerella fragilis CBS 456.73 TKX976693 KX976791 MK876779 KX977042 Wang et al. 2019b
Madurella fahalii CBS 129176 TMK926819 MK926819 MK876780 MK926919 Wang et al. 2019b
Madurella mycetomatis CBS 109801 TMK926820 MK926820 MK876781 MK926920 Wang et al. 2019b
Madurella pseudomycetomatis CBS 129177 TMK926821 MK926821 MK876782 MK926921 Wang et al. 2019b
Madurella tropicana CBS 201.38 TMK926824 MK926824 MK876785 MK926924 Wang et al. 2019b
CBS 206.47 MK926825 MK926825 MK876786 MK926925 Wang et al. 2019b
Melanocarpus albomyces CBS 638.94 TKX976679 KX976773 KX976886 KX977021 Wang et al. 2016a
CBS 747.70 KX976680 KX976774 KX976887 KX977022
594 Mycol Progress (2020) 19:589603
stolon-like mycelium and covered by pigmented hypha-like
hairs (fig. 41 in Wang et al. 2019b). The genus Batnamyces is
more similar to Canariomyces and Stolonocarpus than to
Madurella with respect to the morphology of the colonies
and mycelia but can be easily distinguished from them by
the lack of reproductive structures. Since the ex-type strain
of Batnamyces was obtained by using an isolation procedure,
which is well established for endophytes, from an endemic
plant in an area that has never been studied intensively for
the biodiversity of its mycobiota, it did not come as a surprise
that no reproductive structures are produced. After all, it is
pretty well known that endophytic fungi often do not produce
any propagules. However, we isolated the fungus only one
time and can therefore not be sure about its actual lifestyle.
Poor statistic support (PP < 0.95; ML-BS = 85%) also implied
that B. globulariicola was not a member of either Madurella
or Stolonocarpus.
Isolation and structure elucidation of compounds 15
Fractionation of crude ethyl acetate extracts from the culture
of Batnamyces globulariicola in Q61/2 and YMG media led
to the isolation and structure elucidation of 5 new 2,5-
diketopiperazines (15) together with seven known metabo-
lites identified by spectroscopic analysis and comparison with
literature data, such as Sch 54796 6(Chu et al. 1993;Usami
Tabl e 3 (continued)
Current name Culture accession number
1
GenBank accession numbers
2
References
ITS LSU RPB2 TUB2
Wang et al. 2016a
Microthielavia ovispora CBS 165.75 TMK926826 MK926826 MK876787 MK926926 Wang et al. 2019b
Myceliophthora lutea CBS 145.77 THQ871775 KM655351 HQ871816 KX977026 van den Brink et al. 2012,
Wang et al. 2016a
Mycothermus thermophilus CBS 625.91 TLT993604 LT993604 LT993523 LT993685 Wang et al. 2019a
Ovatospora medusarum CBS 148.67 TKX976684 KX976782 KX976897 KX977032 Wang et al. 2016a
Ovatospora mollicella CBS 583.83 TKX976685 KX976783 KX976898 KX977033 Wang et al. 2016a
Parathielavia hyrcaniae CBS 353.62 TKM655329 KM655368 KM655401 KX977043 van den Brink et al. 2015,
Wang et al. 2016a
Parathielavia kuwaitensis CBS 945.72 TKM655332 KM655371 KM655404 KX977044 van den Brink et al. 2015,
Wang et al. 2016a
Pseudothielavia terricola CBS 165.88 TKX976694 KX976792 MK876795 KX977045 van den Brink et al. 2015,
Wang et al. 2016a
CBS 487.74 MK926834 MK926834 MK876796 MK926934 Wang et al. 2019b
Remersonia thermophila CBS 645.91 LT993611 LT993611 LT993530 LT993692 Wang et al. 2019a
Staphylotrichum coccosporum CBS 364.58 TLT993620 LT993620 LT993539 LT993701 Wang et al. 2019a
Stellatospora terricola CBS 811.95 TMK926835 MK926835 MK876797 MK926935 Wang et al. 2019b
Stolonocarpus gigasporus CBS 112062 TMK926836 MK926836 MK876798 MK926936 Wang et al. 2019b
Subramaniula thielavioides CBS 122.78 TKP862597 KP970654 KP900670 KP900708 Wang et al. 2016a
Thermothielavioides terrestris CBS 117535 TMK926837 MK926837 MK876799 MK926937 Wang et al. 2019b
CBS 492.74 MK926838 MK926838 MK876800 MK926938 Wang et al. 2019b
Thermothelomyces heterothallica CBS 202.75 THQ871771 KM655354 HQ871798 KX977025 van den Brink et al. 2012,
Wang et al. 2016a
Trichocladium asperum CBS 903.85 TLT993632 LT993632 LT993551 LT993713 Wang et al. 2019a
Trichocladium griseum CBS 119.14 TLT993639 LT993639 LT993558 LT993720 Wang et al. 2019a
Podosporaceae
Cladorrhinum foecundissimum CBS 180.66 TMK926856 MK926856 MK876818 MK926956 Wang et al. 2019b
Podospora fimicola CBS 482.64 TMK926862 MK926862 MK876824 MK926962 Wang et al. 2019b
CBS 990.96 MK926863 MK926863 MK876825 MK926963 Wang et al. 2019b
Triangularia anserina CBS 433.50 MK926864 MK926864 MK876826 MK926964 Wang et al. 2019b
Triangularia bambusae CBS 352.33 TMK926868 MK926868 MK876830 MK926968 Wang et al. 2019b
Mycol Progress (2020) 19:589603 595
et al. 2002), Sch 54794 7(Chu et al. 1993; Usami et al. 2002),
cyclo-(glycyl-
L
-tyrosyl)-3,3-dimethylallyl ether 8(Koolen
et al. 2012), 4-O-(3-methylbut-2-enyl)benzoic acid 9
(Nozawa et al. 1989),
L
-Pro-
L
-Ile 10 (Ren et al. 2010),
L
-
Pro-
L
-Leu 11 (Ren et al. 2010; Sansinenea et al. 2016), and
L
-Pro-
L
-Phe) 12 (Sansinenea et al. 2016)(Fig.3).
Compound 1was isolated as a yellowish gum. Its molecu-
lar formula C
17
H
20
N
2
O
4
S was deduced from the HRESIMS
which exhibited the pseudomolecular ion peak at m/z
349.1217 [M + H]
+
(calcd for C
17
H
21
N
2
O
4
S
+
, 349.1217).
This was confirmed by the ESIMS ion cluster at m/z371.11
[M + Na]
+
and a prominent ion fragment and 301 [M + H
48]
+
revealing the loss of a methanethiol (CH
3
SH) unit (Chu
et al. 1993). Its
1
H NMR spectrum displayed resonances for an
AABBspin system at δ
H
7.45 (d, J= 8.6 Hz, H-9, and H-9)
and 7.00 (d, J= 8.6 Hz, H-10, and H-10) suggesting the pres-
ence of a 1,4-disubstituted benzene ring in the molecule
(Table 1). It also showed a signal assigned to a vinyl proton
δ
H
6.87 (H-7) and a set of resonances depicted at δ
H
4.68 (brd,
J= 6.4 Hz, H-12), 5.73 (m, H-12), 3.99 (s), 3.99 (s, H-15), and
1.76 (s, H-16) attributed to an O-isoprenol group. Other sig-
nals were those of a thiomethyl singlet at δ
H
2.26 (s, H-17) and
a methine singlet at δ
H
4.97 (s, H-3). The downfield shift of
the latter revealed its bis-heteroatom connectivity (Chu et al.
1997)(Table1). The
13
C NMR spectrum showed two
amidocarbonyl carbon signals characteristic of a
diketopiperazine core at δ
C
166.0 (C-2) and 163.5 (C-5)
(Chu et al. 1993; Guimarães et al. 2010;Fuetal.2011;Fan
et al. 2017). It also displayed resonances at δ
C
125.3 (C-6),
119.3 (C-7), 126.8 (C-8), 132.1 (C-9, C-9), 116.4 (C-10,
C-10), and 160.9 (C-11) evidencing the presence of an
oxybenzylidene moiety. In addition, the signals of two meth-
ylenes at δ
C
66.8 (C-12) and 67.9 (C-15), a methine at δ
C
120.8 (C-13), a methyl at δ
C
14.2 (C-13), and a quaternary
carbon at δ
C
141.5 (C-14) were assigned to the O-isoprenol
group. The remaining signals were those of a methine at δ
C
Fig. 1 Morphological characteristics of the new genus Batnamyces.a
Colonies obverse and reverse on PDA at 23 °C after 21 days. b
Colonies obverse and reverse on OM at 23 °C after 21 days. cColonies
obverse and reverse on YMG at 23 °C after 21 days. dand e
chlamydospore-like structures formed at hyphal tips. Scale bars d, e =
10 μm
Fig. 2 Phylogenetic tree resulting from a maximum likelihood analysis of
the concatenated RPB2,TUB2, ITS, and LSU sequence data sets. The
confidence values are indicated at the notes: bootstrap proportions from
the ML analysis above branches, and the posterior probabilities from the
Bayesian analysis below branches. -means lacking statistical support
(< 70% for bootstrap proportions from ML or MP analyses; < 0.95 for
posterior probabilities from Bayesian analyses). The branches with full
statistical support (MP-BS = 100%; ML-BS = 100%; PP = 1.0) are
indicated by thickened branches. Genus clades are discriminated with
boxes of different colours. The scale bar shows the expected number of
changes per site. The tree is rooted with members of the Podosporaceae.
Tafter the respective DNA sequences are derived from ex-type (includ-
ing ex-epitype and ex-neotype) specimens
596 Mycol Progress (2020) 19:589603
0.2
Trichocladium griseum CBS 119.14 T
Humicola fuscoatra CBS 118.14 T
Stolonocarpus gigasporus CBS 112062 T
Subramaniula thielavioides CBS 122.78 T
Chrysanthotrichum lentum CBS 339.67 T
Ovatospora medusarum CBS 148.67 T
Parathielavia hyrcaniae CBS 353.62 T
Thermothielavioides terrestris CBS 492.74
Pseudothielavia terricola CBS 165.88 T
Mycothermus thermophilus CBS 625.91 T
Collariella robusta CBS 551.83 T
Pseudothielavia terricola CBS 487.74
Myceliophthora lutea CBS 145.77 T
Collariella virescens CBS 148.68 T
Podospora fimicola CBS 482.64 T
Carteria arctostaphyli CBS 229.82 T
Canariomyces microsporus CBS 161.80
Triangularia anserina CBS 433.50
Madurella tropicana CBS 201.38 T
Podospora fimicola CBS 990.96
Microthielavia ovispora CBS 165.75 T
Cladorrhinum foecundissimum CBS 180.66 T
Melanocarpus albomyces CBS 638.94 T
Chrysocorona lucknowensis CBS 385.66
Trichocladium asperum CBS 903.85 T
Melanocarpus albomyces CBS 747.70
Arcopilus aureus CBS 153.52
Triangularia bambusae CBS 352.33 T
Brachychaeta variospora CBS 414.73 T
Floropilus chiversii CBS 558.80 T
Chrysanthotrichum allolentum CBS 644.83 T
Acrophialophora nainiana CBS 100.60 T
Chaetomium globosum CBS 160.62 T
Condenascus tortuosus CBS 610.97
Amesia atrobrunnea CBS 379.66 T
Achaetomium globosum CBS 332.67 T
Dichotomopilus indicus CGMCC 3.14184 T
Botryotrichum murorum CBS 163.52
Madurella pseudomycetomatis CBS 129177 T
Dichotomopilus funicola CBS 159.52 T
Chrysocorona lucknowensis CBS 727.71 T
Corynascella humicola CBS 337.72 T
Botryotrichum piluliferum CBS 654.79
Canariomyces microsporus CBS 276.74 T
Thermothelomyces heterothallica CBS 202.75 T
Stellatospora terricola CBS 811.95 T
Amesia nigricolor CBS 600.66 T
Corynascus sepedonium CBS 111.69 T
Achaetomium strumarium CBS 333.67 T
Remersonia thermophila CBS 645.91
Ovatospora mollicella CBS 583.83 T
Madurella fahalii CBS 129176 T
Canariomyces notabilis CBS 548.83 T
Madurella mycetomatis CBS 109801 T
Chaetomium elatum CBS 142034 T
Corynascella humicola CBS 379.74
Thermothielavioides terrestris CBS 117535 T
Staphylotrichum coccosporum CBS 364.58 T
Parathielavia kuwaitensis CBS 945.72 T
Arcopilus flavigenus CBS 337.67 T
Hyalosphaerella fragilis CBS 456.73 T
Batnamyces globularicola CBS 144474 T
Madurella tropicana CBS 206.47
Collariella bostrychodes CBS 163.73
Acrophialophora nainiana CBS 417.67
0.99/100
0.99/85
0.99/100
0.98/100
-/93
0.98/100
0.95/-
0.99/100
0.99/100
0.96/96
0.98/91
0.99/100
0.99/100
0.99/100
0.99/100
0.96/95
0.99/100
0.99/100
0.95/92
0.98/94
0.99/97
0.99/100
0.96/92
0.97/82
0.99/100
0.99/100
0.97/100
0.95/71
0.99/90
0.99/100
0.99/100
0.99/100
-/85
0.99/100
0.99/100
0.99/100
0.99/100
0.96/81 0.99/100
0.99/100
0.99/100
0.99/100
0.97/100
-/86
0.98/100
0.99/100
Mycol Progress (2020) 19:589603 597
59.6 (C-3) and the thiomethyl group at δ
C
13.1 (C-17)
(Table 1). The location of the thiomethyl group at C-3 was
further evidenced by the HMBC correlation observed between
thiomethyl protons signal at δ
H
2.26 (s, H-17) and the carbon
at δ
C
59.6 (C-3). Furthermore, the HMBC correlation from H-
12 (δ
H
4.68) to carbon C-11 (δ
C
160.9) confirmed that the O-
isoprenol group was linked at C-11 (Fig. 4). Careful examina-
tion of the
1
H-
1
H COSY, HSQC, and HMBC spectra proved
that 1was related to Sch 56396, a metabolite produced by the
fungus Tolypocladium sp. (Chu et al. 1997), the main differ-
ence was the hydroxylation of one methyl of the isopentenyl
group to form 1.TheEgeometry for the Δ
13,14
double bond
was determined from the NOESY correlation depicted be-
tween the olefinic proton H-13 (δ
H
5.73) and the methylene
protons H-15 (δ
H
3.99). To solve the stereochemistry of Δ
6,7
double bond, a NOESY spectrum was measured in DMSO-d
6
.
The lack of NOESY correlation between NH-1 depicted at δ
H
10.17 (s) and the vinyl proton H-7 (δ
H
6.87) was in favour of
the Zconfiguration. This was further confirmed by the
NOESY correlation depicted between NH-1 (δ
H
10.17 s)
and H-9/H-9(δ
H
7.45, d, 8.6). The absolute configuration
at C-3 was determined to be Rby comparison of the experi-
mental ECD of 1(Fig. S9, Supporting information) with the
calculated ECD spectra for the four stereoisomers (3R,6E;
3S,6E;3R,6Z;3S,6Z) of a related compound (Guimarães
et al. 2010). Although the absorption on the experimental
spectrum was not intense, the Cotton effects of 1were in
accordance with the experimental Cotton effects of 3R,6Zste-
reoisomer especially with positive absorption in the regions of
200240 nm and 275350 nm, respectively. The structure of
compound 1was unambiguously elucidated as (3R,6Z)-3-
thiomethyl-6-[4-O-[(2E)-4-hydroxy-3-methylbut-2-
enyl]benzylidene]piperazine-2,5-dione.
Metabolite 2was obtained as a yellowish gum. It possessed
the same molecular formula as 1as evidenced by the
HRESIMS which showed a protonated molecular ion peak
at m/z349.1217 [M + H]
+
(calcd for C
17
H
21
N
2
O
4
S
+
,
349.1217) despite the fact that both compounds had different
retention times. The
1
Hand
13
C NMR spectra of 2(Table 1)
were closely related to those of 1, especially for signals of the
diketopiperazine core and the oxybenzylidene moiety. The
only difference was the downfield or upfield shifts of some
1
Hand
13
C signals probably due to the change of configura-
tion at the Δ
13,14
double bond. This configuration was de-
duced to be Zfrom careful analysis of the NOESY spectrum
which exhibited a cross-peak correlation between the methy-
lene protons at δ
H
4.67 (brd, J= 6.4 Hz, H-12) and 4,16 (s,
H-16). The NOESY correlation between the olefinic proton
H-13 (δ
H
, 5.57, m) and the methyl protons H-15 (δ
H
1.85) was
also depicted. In view of determining the configuration of the
Δ
6,7
double bond, the NOESY spectrum of 2was also mea-
suredinDMSO-d
6
. The NOESY correlation observed
N
H
H
NO
O
O
R
3
R
2
R
1
1
2
345
6
7
8
9
10
11
12
13
14
9'
10'
R
1
R
2
R
3
SMe CH
2
OH CH
3
SMe CH
3
CH
2
OH
OH CH
3
CH
3
SMe CH
3
CH
3
1:
2:
3:
4:HN
N
O
O
HN
N
O
O
HN
N
O
O
10
11 12
O
CH
2
OH
N
H
H
NO
O
MeS
SMe
O
N
H
H
NO
O
MeS
SMe
O
N
H
H
NO
O
MeS
SMe
O
O
HO
O
N
H
H
NO
O
5
6
7
8
9
1
2
345
678
9
10
11
12
13
14
9'
10' 15
16
17
18
Fig. 3 Structures of compounds 112 isolated from the culture of Batnamyces globulariicola
598 Mycol Progress (2020) 19:589603
between the NH-1 proton signal (δ
H
10.17) and H-9/H-9(δ
H
7.45, d, J=8.6)showedthat Δ
6,7
has the same geometry in
metabolites 1and 2. The experimental ECD spectrum of 2
(Fig. S18, Supporting information) was nearly identical to that
of 1, leading to the deduction of the 3Rabsolute configuration
for 2. The structure of metabolite 2was thus concluded as
(3R,6Z)-3-thiomethyl-6-[4-O-[(2Z)-4-hydroxy-3-methylbut-
2-enyl]benzylidene]piperazine-2,5-dione.
Compound 3was obtained as a yellowish gum. Its molecular
formula was determined as C
16
H
18
N
2
O
4
from the HRESIMS
analysis which showed the pseudomolecular ion at m/z
303.1334 [M + H]
+
(calcd for C
16
H
19
N
2
O
4+
, 303.1339). Its
1
H
NMR spectrum exhibited resonances for an AABBspin system
at δ
H
7.46 (d, J=8.7Hz,H-9,andH-9) and 6.98 (d, J=8.7Hz,
H-10, and H-10) suggesting a para-disubstituted aromatic ring
in the structure. It also showed signals for a vinyl proton at δ
H
5.47 (m, H-13), an oxymethylene at δ
H
4.58 (brd, J=6.6 Hz,
H-12), and two vinyl connected methyl singlets at δ
H
1.79 (H-15)
and1.76(H-16)characteristicofanO-isoprenyl moiety. Signals
of the vinyl proton H-7 and the methine proton H-3 were ob-
served at δ
H
6.80 (s) and 5.09 (s), respectively. The
13
CNMR
spectrum exhibited signals of two amidocarbonyl carbons at δ
C
167.0 (C-2) and 163.8 (C-5). Those characteristic of the
oxybenzylidene moiety were observed at δ
C
125.4 (C-6), 119.5
(C-7), 126.8 (C-8), 132.1 (C-9, C-9), 116.4 (C-10, C-10), and
160.9 (C-11). The remaining resonances depicted at δ
C
66.1
(C-12), 121.1 (C-13), 139.1 (C-14), δ
C
26.0 (C-15), and 18.3
(C-16) confirmed the presence of the O-prenyl group in the mol-
ecule (Table 1). The downfield shift of C-3 in metabolite 3(δ
C
76.1) with respect to compounds 1and 2(δ
C
59.6) suggested the
presence of an OH group at C-3 in 3instead of the thiomethyl
moiety. This was further supported by mass analysis and the
absence of the thiomethyl signal on the
1
Hand
13
CNMRspectra.
The structure was confirmed by a comprehensive analysis of the
2D NMR data, particularly
1
H-
1
H COSY, HSQC, and HMBC
spectra (Fig. 4). The configuration of the Δ
6,7
double bond was
determined to be Zby the NOESY spectrum measured in
DMSO-d
6
on which correlations were observed between NH-1
and H-9 (or H-9) but not between NH-1 (δ
H
9.92, s) and H-7 (δ
H
6.80, s). The ECD spectrum of 3(Fig. S27, Supporting
information) indicated the same absolute configuration at C-3
(R) as for metabolites 1and 2. The structure was finally conclud-
ed as (3R,6Z)-3-hydroxy-6-[4-O-(3-methylbut-2-
enyl)benzylidene]piperazine-2,5-dione.
The molecular formula of 4also obtained as a yellowish
gum was deduced to be C
17
H
20
N
2
O
3
S from the HRESIMS
which showed ion clusters [M + Na]
+
at m/z 355.1088 (Calcd
355.1087) and [2M + Na]
+
at m/z 687.2283 (Calcd 687.2281).
The NMR data (Table 2) showed similarities with the previ-
ously described metabolites 13.Its
1
H-NMR spectrum ex-
hibited in addition to signals of the diketopiperazine and the
oxybenzylidene moieties two olefinic methyl resonances at δ
H
1.76 and 1.79, an oxygenated methylene doublet at δ
H
4.58
(J= 6.6 Hz), and a vinyl proton signal at δ
H
5.46 (m) charac-
teristic of a γ,γ-dimethylallyloxy moiety (Sritularak and
Likhitwitayawuid 2006). Careful examination of the
1
H-
1
H
COSY, HSQC, and HMBC spectra proved that 4was similar
to Sch 56396 previously isolated from the fermentation broth
of the fungus Tolypocladium sp. (Chu et al. 1997), but the only
difference was on the sign of their optical rotations.
Compound 4showed a positive optical rotation, while the
optical rotation of Sch 56396 was negative, confirming that
they are stereoisomers. The Zconfiguration of the Δ
6,7
was
deduced from the chemical shift of H-7 (δ
H
6.87) in compar-
ison with those of the same proton in compounds 13, since it
was reported that the (Z)-vinyl proton of the 6-benzylidene-
substituted piperazine-2,5-diones is farther downfield than the
(E)-vinyl proton because of the deshielding effect of the 5-
ketone (Fu et al. 2011). Since the configuration of the chiral
centre C-3 of Sch 56396 was not determined, we measured the
ECD spectrum of 4(Fig. S36, Supporting information) and its
comparison with those of compounds 13allowed us to as-
sign the 3Rconfiguration. Compound 4was then elucidated as
(3R,6Z)-3-thiomethyl-6-[4-O-(3-methylbut-2-
enyl)benzylidene]piperazine-2,5-dione, the stereoisomer of
Sch 56396.
Compound 5was obtained as a yellow gum from metha-
nol. Its HRESIMS showed ion clusters at m/z 397.1249 [M +
N
H
H
NO
O
O
CH
3
CH
2
OH
H
3
CS
O
CH
2
OH
CH
3
N
H
H
NO
O
H
3
CS
SCH
3
N
H
H
NO
O
O
CH
2
OH
CH
3
H
3
CS
N
H
H
NO
O
O
CH
3
CH
3
HO
1
H-
1
HCOSY
HMBC
123
5
Fig. 4 Selected
1
H-
1
H COSY and HMBC correlations for compounds 13and 5
Mycol Progress (2020) 19:589603 599
H]
+
and 419.1068 [M + Na]
+
consistent with the molecular
formula of C
18
H
24
N
2
O
4
S
2
(calcd for C
18
H
25
N
2
O
4
S
2+
,
397.1250; calcd for C
18
H
24
N
2
O
4
S
2
Na
+
, 419.1075). The pres-
ence of two thiomethyl groups was confirmed by the ion frag-
ments depicted at m/z 349.1215 [M + H48]
+
and 301.1177
[M + H2×48]
+
revealing the loss of two methanethiol
(CH
3
SH) units (Chu et al. 1993). Its
1
HNMRspectrum
showed in addition to the signals of the O-isoprenoltyrosine
moiety at δ
H
7.26 (d, J= 9.0 Hz, H-9, and H-9), 6.83 (d, J=
9.0 Hz, H-10, and H-10), 4.61 (brd, J= 6.3 Hz, H-12), 5.47
(m, H-12), 4.15 (s, H-15), and 1.81 (s, H-16) those of two
thiomethyl groups at δ
H
1.36 (s, H-18) and 2.23 (s, H-17) as
well as a methylene group observed as an AX spin system at
δ
H
2.94 (d, J= 13.5, H-7A) and 3.60 (d, J= 13.5, H-7X)
(Table 2). Careful examination of the
13
C,
1
H-
1
HCOSY,
HSQC, and HMBC spectra proved metabolite 5to have the
same planar structure as meromutides A and B recently iso-
lated after pleiotropic activation of natural products in
Metarhizium robertsii by deletion of a histone acetyltransfer-
ase (Fan et al. 2017). The Zgeometry for the Δ
13,14
double
bond was determined from the NOESY correlation depicted
between the methylene protons at δ
H
4.61 (brd, J= 6.4 Hz,
H-12) and 4.15 (s, H-16). Compound 5with the Zgeometry of
the olefinic double bond can possess 4 stereoisomers (3S,6S),
(3R,6S), (3S,6R), and (3R,6R). Up to now, only two of them,
namely meromutide A (3S,6S) and meromutide B (3R,6S)
were isolated and characterised. Its absolute configuration
was determined by careful comparison of the chemical shifts
for both protons and carbons of the thiomethyl groups linked
at C-3 and C-6 of its known stereoisomers. For meromutide A,
these chemical shifts were as follows: C-3 (δ
H
2.24; δ
C
13.9)
and C-6 (δ
H
2.30; δ
C
13.4), while for meromutide B, they
were C-3 (δ
H
2.21; δ
C
10.4) and C-6 (δ
H
1.20; δ
C
12.9) (Fan
et al. 2017). The chemical shifts of the corresponding protons
and carbons of the thiomethyl groups at C-3 and C-6 in com-
pound 5(also measured in methanol-d
6
) were different from
those of the known stereoisomers ((δ
H
1.36; δ
C
10.6) and (δ
H
2.23; δ
C
13.4), respectively) indicating a change of configu-
ration in one of the chiral centres. On the other hand, since
metabolite 5is the C-15 hydroxyl derivative of Sch54796 (6)
and Sch54794 (7) possessing the known (3R,6S)and(3S,6S)
configurations, its absolute configuration at C-6 must be Ras
supported by the positive optical rotation (+ 33.3, MeOH) in
comparison with those of the known congeners (Sch54796 (
25, DMSO) and Sch54794 (70, DMSO); see Chu et al.
(1993). Furthermore, the NOESY correlation depicted be-
tween H-3 (δ
H
5.01, s) and CH
3
-18 (2.23, s) revealed the
trans-orientation of the two thiomethyl groups on the
diketopiperazine ring. To further confirm the 3S,6Rconfigu-
ration, we compared the ECD spectrum of metabolite 5with
that of fusaperazine, a related thiodiketopiperazine possessing
a3R,6Rconfiguration obtained from a marine algae-derived
fungus Penicillium sp. KMM 4672 (Yurchenko et al. 2019).
Both compounds exhibited a negative Cotton effect between
220 and 240 nm, probably due to the common 6Rconfigura-
tion. However, a negative Cotton effect was observed on the
ECD spectrum of compound 5between 240 and 290 nm while
a positive Cotton effect was depicted in the same zone of the
ECD spectrum of fusaperazine. The structure of 5was finally
elucidated as (3S,6R)-3,6-bisthiomethyl-6-[4-O-[(2Z)-4-hy-
droxy-3-methylbut-2-enyl]phenylmethyl]piperazine-2,5-
dione.
Diketopiperazines are cyclic peptides produced by bacteria
and fungi arising from the cyclisation of two or more amino
acids catalysed either by two-modular non-ribosomal peptide
synthetases or by cyclodipeptide synthases (Huang et al.
2014; Brockmeyer and Li 2017). Since the isolated com-
pounds are biogenetically related, their biosynthetic relation-
ships were proposed. Compounds 10,11, and 12 could be
obtained from the cyclisation of L-proline and
L
-isoleucine,
L-proline and L-leucine, and L-proline and L-phenylalanine,
respectively. While working on the biosynthesis of the
epidithiodiketopiperazine gliotoxin, Scharf et al. 2011 discov-
ered that a specialised glutathione S-transferase (GliG) plays a
key role in C-S bond formation (sulfurisation) and that
bishydroxylation of the diketopiperazine by oxygenase
(GliC) is a prerequisite for glutathione adduct formation.
Cyclisation of glycine and L-tyrosine followed by O-
prenylation affords cyclo-(glycyl-L-tyrosyl)-3,3-dimethylallyl
ether 8which could undergo bishydroxylation by oxygenase
GliC to yield the intermediate 13 (not isolated). Thiolation of
13 in the presence of GliG could afford Sch 54796 (6) and Sch
54794 (7). The C-6 epimerisation of Sch 54794 (7)followed
by hydroxylation at C-15 could lead compound 5. The inter-
mediate 13 could also undergo dehydration to afford metabo-
lite 3which could give compound 9after an oxidative cleav-
age of the C-6C-7 double bond. Thiolation of metabolite 3
could also lead compound 4, which could undergo hydroxyl-
ation of one of the prenylmethyl groups to give 1and 2(Fig.
5).
Biological activities
Since some derivatives of diketopiperazines were reported to
display significant antibiotic, antitumor, and immunosuppres-
sant properties (Ameur et al. 2004), while others show a wide
range of biological effects in cell cycle progression (Cui et al.
1996), the isolated compounds were tested for their antimicro-
bial and cytotoxic activities against various bacteria, fungi,
and two mammalian cell lines, but only weak cytotoxic activ-
ity was observed for metabolites 1,4,and9against KB3.1
cells. The evaluation of these compounds in additional bioas-
says is presently underway.
In general, Batnamyces globulariicola belongs to a group
of Chaetomiaceae that has been poorly studied for secondary
metabolites, suggesting that it will be worthwhile to examine
600 Mycol Progress (2020) 19:589603
further strains that appear phylogenetically related for the pro-
duction of diketopiperazines and other secondary metabolites.
The lack of suitable morphological features for the classifica-
tion of these fungi using a polythetic approach could thus be
compensated by chemotaxonomic methodology, as recently
accomplished for some genera of the Xylariales (cf.
Samarakoon et al. 2020; Wittstein et al. 2020).
Conclusion
Batnamyces globulariicola was only isolated once among
several hundreds of cultures that inhabited the rhizosphere of
the host plant during the course of a PhD thesis (Noumeur
2018). We selected this strain out of many others that were
isolated concurrently because it turned out to represent a hith-
erto unknown phylogenetic lineage among the Sordariales.
We were at first unsure whether to report the fungus as an
unknown memberof the Chaetomiaceae, along with the
secondary metabolites, but finally decided to name the taxon,
and place it in its phylogenetic context, as part of an ongoing
taxonomic study (Wang et al. 2019b). It was rather surprising
to see that it still did not fit in any of the known genera of the
Chaetomiaceae even by comparison with the latest mono-
graphic work.
We are aware of the potential pitfalls regarding the classi-
fication of such a single strain as a member of a monotypic
genus, and in particular the requests by some members of the
mycological community that more than one culture needs to
be deposited to justify the formal description of a new taxon.
However, such arguments mostly come from scientists who
are working with ubiquitous fungal genera that make up the
bulk of new species in the large classes of Eurotiomycetes,
Sordariomycetes, and Dothideomycetes, where ex-type
strains as well as large numbers of congeneric isolates are
readily available.
We cannot be absolutely certain whether our new fungus
really represents an endophyte, because it was only isolated
once. However, a sterile mycelium could hardly have survived
the applied, rather harsh root disinfection procedure unless it
forms persistent propagules which may have survived the
chemical treatment. Nevertheless, we cannot exclude that it
is heterothallic and that a sexual state may exist in nature.
Even though we did not observe sporulation of the culture,
O
N
H
H
NO
O
MeS
SMe
O
N
H
H
NO
O
MeS
SMe
O
O
HO
6
7
8
9
O
N
H
H
NO
O
1
2
345
678
9
10
11
12
13
14
9'
10' 15
16
HO
O
OH
NH
2
O
OH
H
2
N
Glycine
L-Tyrosine Cyclisation
Prenylation
Bishydroxylation
O
N
H
H
NO
O
HO
OH
Thiolation
13
C-6 Epimerization
C-15 Hydroxylation
O
CH
2
OH
N
H
H
NO
O
MeS
SMe
N
H
H
NO
O
O
HO
Dehydration
Oxidative
cleavage
N
H
H
NO
O
O
MeS
N
H
H
NO
O
O
R
2
R
1
MeS
R
1
R
2
CH
2
OH CH
3
CH
3
CH
2
OH
1:
2:
3
Thiolation
5
Hydroxylation
4
Fig. 5 Proposed biogenetic pathway for the formation of compounds 19from glycine and
L
-tyrosine
Mycol Progress (2020) 19:589603 601
aside from the production of chlamydospores, we cannot ex-
clude that it forms persistent propagules in its natural habitat
that we have not been able to induce in the laboratory. There
remains a chance that a yet unknown stage of the fungus has
persisted in a soil particle that was attached to the roots, and it
should be attempted to re-isolate Batnamyces from other parts
of the plant. The taxonomy of other fungi that were already
frequently isolated from similar biotopes, such as the so-called
dark septate root endophytes, which mostly belong to the
Dothideomycetes (Knapp et al. 2015; Bonfim et al. 2016),
poses a similar challenge. Their taxonomy cannot be resolved
by using a morphocentric approach based on conidiogenous
structures but must rely on molecular methods. Even here, a
chemotaxonomic approach would be interesting to pursue to
attain complementary phenotype-derived data.
Acknowledgements The technical assistance of Nadine Wurzler and
Vanessa Stiller is appreciated. We thank Wera Collisi for conducting the
bioassays, Christel Kakoschke for recording NMR spectra and Sabrina
Karwehl for HRESIMS measurements.
Funding information Open Access funding provided by Projekt DEAL.
RBT and SEH are grateful for financial support from the Alexander von
Humboldt Foundation. SRN acknowledges the Ministry of Higher
Education and Scientific Research (MESRS) of Algeria for the financial
support.
Compliance with ethical standards
Conflict of interest The authors declare that they have no conflict of
interest.
Open Access This article is licensed under a Creative Commons
Attribution 4.0 International License, which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as
long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons licence, and indicate if
changes were made. The images or other third party material in this article
are included in the article's Creative Commons licence, unless indicated
otherwise in a credit line to the material. If material is not included in the
article's Creative Commons licence and your intended use is not
permitted by statutory regulation or exceeds the permitted use, you will
need to obtain permission directly from the copyright holder. To view a
copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.
References
Ameur RMB, Mellouli L, Chabchoub F, Fotso S, Bejar S (2004)
Purification and structure elucidation of two biologically active mol-
ecules from a new isolated Streptomyces sp. US 24 strain. Chem Nat
Compd 40:510513
Bills GF, Gloer JB (2016) Biologically active secondary metabolites from
the fungi. Microbiol Spectrum 4:6
Blackwell M, Vega FE (2018) Lives within lives: hidden fungal biodi-
versity and the importance of conservation. Fungal Ecol 35:127134
Bonfim JA, Vasconcellos RL, Baldesin LF, Sieber TN, Cardoso EJ
(2016) Dark septate endophytic fungi of native plants along an
altitudinal gradient in the Brazilian Atlantic forest. Fungal Ecol 20:
202210
Brockmeyer K, Li SM (2017) Mutations of residues in pocket P1 of a
cyclodipeptide synthase strongly increase product formation. J Nat
Prod 80:29172922
Cai L, Jeewon R, Hyde KD (2006) Molecular systematics of
Sordariaceae based on multiple gene sequences and morphology.
Mycol Res 110:137150
Chepkirui C, Cheng T, Sum WC, Matasyoh JC, Decock C, Praditya DF,
Wittstein K, Steinmann E, Stadler M (2019) Skeletocutins A-L:
antibacterial agents from the Kenyan wood-inhabiting basidiomy-
cete, Skeletocutis sp. J Agric Food Chem 31:84688475
Chu M, Mierzwa R, Truumees I, Gentile F, Patel M, Gullo V, Chan TM,
Puar MS (1993) Two novel diketopiperazines isolated from the fun-
gus Tolypocladium sp. Tetrahedron Lett 34:75377540
Chu M, Truumees I, Mierzwa R, Patel M, Puar MS (1997) Sch 56396: a
new c-fos proto-oncogene inhibitor produced by the fungus
Tolypocladium sp.J Antibiot 50:10611063
Crous PW, Verkley GJ, Groenewald JZ, Samson RA (2009) Fungal bio-
diversity. CBS laboratory manual series 1. Centraalbureau voor
Schimmelcultures, Utrecht
Cui CB, Kakeya H, Okada G, Onose R, Osada H (1996) Novel mamma-
lian cell cycle inhibitors, tryprostatins A, B and other
diketopiperazines produced by Aspergillus fumigatus. J Antibiot
49:527533
De Hoog GS, Ahmed SA, Najafzadeh MJ, Sutton DA, Saradeghi Keisari
M, Fahal AH, Eberhart U, Verkley GJ, Xin L, Stielow B, van de
Sande WWJ (2013) Phylogenetic findings suggest possible new
habitat and routes of infection of human eumyctoma. PLoS Negl
Trop Dis 7:e2229
Fan A, Mi W, Liu Z, Zeng G, Zhang P, Hu Y, Fang W, Yin WB (2017)
Deletion of a histone acetyltransferase leads to the pleiotropic acti-
vation of natural products in Metarhizium robertsii. Org Lett 19:
16861689
Fu P, Liu P, Qu H, Wang Y, Chen D, Wang H, Li J, Zhu W (2011) α-
Pyrones and diketopiperazine derivatives from the marine-derived
actinomycete Nocardiopsis dassonvillei HR10-5. J Nat Prod 74:
22192223
Guimarães DO, Borges WS, Vieira NJ, De Oliveira LF, Da Silva CHTP,
Norberto P, Lopes NP, Dias LG, Durán-Patrón R, Collado IG, Pupo
MT (2010) Diketopiperazines produced by endophytic fungi found
in association with two Asteraceae species. Phytochemistry 71:
14231429
Huang RM, Yi XX, Zhou Y, Su X, Peng Y, Gao CH (2014) An update on
2,5-diketopiperazines from marine organisms. Mar Drugs 12:6213
6235
Hyde KD, Xu JC, Rapior S, Jeewon R, Lumyong S et al (2019) The
amazing potential of fungi, 50 ways we can exploit fungi industri-
ally. Fungal Divers 97:1136
Katoh K, Standley DM (2013) MAFFT multiple sequence alignment
software version 7: improvements in performance and usability.
MolBiolEvol30:772780
Knapp DG, Kovács GM, Zajta E, Groenewald JZ, Crous PW (2015) Dark
septate endophytic pleosporalean genera from semiarid areas.
Persoonia 35:87
Koolen HHF, Soares ER, Da Silva FMA, De Souza AQL, De Medeiros
LS, Filho ER, De Almeida RA, Ribeiro IA, Pessoa C, De Morais
MO, Da Costa PM, De Souza ADL (2012) An antimicrobial
diketopiperazine alkaloid and co-metabolites from an endophytic
strain of Gliocladium isolated from Strychnos cf. toxifera.Nat
Prod Res 26:20132019
Miller MA, Pfeiffer W, Schwartz T (2010) Creating the CIPRES Science
Gateway for inference of large phylogenetic trees. In: Proceedings
of the Gateway Computing Environments Workshop (GCE),
November 14, 2010, New Orleans, Louisiana: pp. 18
602 Mycol Progress (2020) 19:589603
Noumeur SR (2018) Identification of bioactive natural products from
endophytic fungi isolated from Globularia alypum. PhD disserta-
tion, University of Setif, Algeria
Noumeur SR, Helaly SE, Jansen R, GerekeM, Stradal TEB, Harzallah D,
Stadler M (2017) Preussilides AF, bicyclic polyketides from the
endophytic fungus Preussia similis with antiproliferative activity. J
Nat Prod 80:15311540
Nozawa K, Takada M, Udagawa SL, Nakajima S,Kawai KI (1989) Three
p-hydroxybenzoic acid derivatives from Talaromyces derxii.
Phytochemistry 28:655656
Nylander JAA (2004) MrModeltest v. 2. Programme distributed by the
author. Evolutionary Biology Centre. Uppsala University
Rambaut A (2009) FigTree v. 1.3.1. Computer program and documenta-
tion distributed by the author at. http://tree.bio.ed.ac.uk/software/
Ren S, Ma W, Xu T, Lin X, Yin H, Yang B, Zhou XF, Yang XW, Long L,
Lee KJ, Gao Q, Liu Y (2010) Two novel alkaloids from the South
China Sea marine sponge Dysidea sp. J Antibiot 63:699701
Richter C, Helaly SE, Thongbai B, Hyde KD, Stadler M (2016)
Pyristriatins A and B: pyridino-cyathane antibiotics from the basid-
iomycete Cyathus cf. striatus. J Nat Prod 79:16841688
Ronquist F, Teslenko M, van der Mark P, Ayres DL, Darling A, Höhna S,
Larget B, Liu L, Suchard MA, Huelsenbeck JP (2012) MrBayes 3.2:
efficient Bayesianphylogenetic inference and model choice across a
large model space. Syst Biol 61:539542
Samarakoon MC, Thongbai B, Hyde KD, Brönstrup M, Beutling U,
Lambert C, Miller AN, Liu JK, Promputtha I, Stadler M (2020)
Elucidation of the life cycle of the endophytic genus Muscodor
and its transfer into the genus Induratia in Induratiaceae fam. nov.,
based on a polyphasic taxonomic approach. Fungal Divers. https://
doi.org/10.1007/s13225-020-00443-9
Sandargo B, Thongbai B, Praditya D, Steinmann E, Stadler M, Surup F
(2018) Antiviral 4-hydroxypleurogrisein and antimicrobial pleurotin
derivatives from cultures of the nematophagous basidiomycete
Hohenbuehelia grisea. Molecules 23:2697
Sansinenea E, Salazar F, Jiménez J, Mendoza A, Ortiz A (2016)
Diketopiperazines derivatives isolated from Bacillus thuringiensis
and Bacillus endophyticus, establishment of their configuration by
X-ray and their synthesis. Tetrahedron Lett 57:26042007
Scharf DH, Remme N, Habel A, Chankhamjon P, Scherlach K,
Heinekamp T, Hortschansky P, Brakhage AA, Hertweck C (2011)
A dedicated glutathione S-transferase mediates carbon-sulfur bond
formation in gliotoxin biosynthesis. J Am Chem Soc 133:12322
12325
Sritularak B, Likhitwitayawuid K (2006) Flavonoids from the pods of
Millettia erythrocalyx. Phytochemistry 67:812817
Stamatakis A (2014) RAxML version 8: a tool for phylogenetic analysis
and post-analysis of large phylogenies. Bioinformatics 30:1312
1313
Suryanarayanan TS, Thirunavukkarasu N, Govindarajulu MB, Sasse F,
Jansen R, Murali TS (2009) Fungal endophytes and bioprospecting.
Fungal Biol Rev 23(12):919
Tamura K, Stecher G, Peterson D, Filipski A, Kumar S (2013) MEGA6:
molecular evolutionary genetics analysis version 6.0. Mol Biol Evol
30:27252729
Teponno RB, Noumeur SR, Helaly SE, Hüttel S, Harzallah D, Stadler M
(2017) Furanones and anthranilic acid derivatives from the endo-
phytic fungus Dendrothyrium variisporum. Molecules 22:1674
1685
Usami Y, Aoki S, Hara T, Nu A (2002) New dioxopiperazine metabolites
from a Fusarium species separated from a marine alga. J Antibiot
55:655659
Van den Brink J, Samson RA, Hagen F, Boekhout T, de Vries RP (2012)
Phylogeny of the industrial relevant, thermophilic genera
Myceliophthora and Corynascus. Fungal Divers 52:197207
Van den Brink J, Facun K, de Vries M et al (2015) Thermophilic growth
and enzymatic thermostability are polyphyletic traits within
Chaetomiaceae. Fungal Biol 119:12551266
Wang XW, Houbraken J, Groenewald JZ, Meijier M, Andersen B,
Nielsen KF, Crous PW, Samson RA (2016a) Diversity and taxono-
my of Chaetomium and chaetomium-like fungi from indoor envi-
ronments. Stud Mycol 84:145224
Wang XW, Lombard L,Groenewald JZ, Li J, Videira S, Samson RA, Liu
XZ, Crous PW (2016b) Phylogenetic reassessment of the
Chaetomium globosum species complex. Persoonia 36:83133
Wang XW, Yang FY, Meijer M, Kraak B, Sun BD, Jiang YL, Wu YM,
Bai FY, Seifert KA, Crous PW, Samson RA (2019a) Redefining
Humicola sensu stricto and related genera in the Chaetomiaceae.
Stud Mycol 93:65153
Wang XW, Bai FY, Bensch K, Meijer M, Sun BD, Han YF, Crous PW,
Samson RA, Yang FY, Houbraken J (2019b) Phylogenetic re-
evaluation of Thielavia with the introduction of a new family
Podosporaceae. Stud Mycol 93:155252
Wendt L, Sir EB, Kuhnert E, Heitkämper S, Lambert C, Hladki AI,
Romero AI, Luangsa-ard JJ, Srikitikulchai P, Peršoh D, Stadler M
(2018) Resurrection and emendation of the Hypoxylaceae,
recognised from a multi-gene genealogy of the Xylariales. Mycol
Prog 17:115154
White JF, Kingsley KL, Zhang Q, Verma R, Obi N, Dvinskikh S, Elmore
MT, Verma SK, Gond SK, Kowalski KP (2019) Endophytic mi-
crobes and their potential applications in crop management. Pest
Manag Sci 75(10):25582565
Wittstein K, Cordsmeier A, Lambert C, Wendt L, Sir EB, Weber J,
Wurzler N, Petrini LE, Stadler M (2020) Identification of
Rosellinia species as producers of cyclodepsipeptide PF1022 A
and resurrection of the genus Dematophora as inferred from
polythetic taxonomy. Stud Mycol 96:116
Yurchenko AN, Berdyshev DV, Smetanina OF, Ivanets EV, Zhuravleva
OI, Rasin AB, Khudyakova YV, Popov RS, Sergey A, Dyshlovoy
SA, Amsberg GV, Afiyatullov SS (2019) Citriperazines A-D pro-
duced by a marine algae-derived fungus Penicillium sp. KMM
4672. Nat Prod Res, in press. https://doi.org/10.1080/14786419.
2018.1552696
Publishersnote Springer Nature remains neutral with regard to jurisdic-
tional claims in published maps and institutional affiliations.
Mycol Progress (2020) 19:589603 603
... The order also includes well-known model-organisms such as Neurospora crassa and Podospora anserina, both of which have been key-players in important scientific discoveries (Davis & Perkins, 2002;Gladieux et al., 2020;Roche et al., 2014;Silar, 2020). It furthermore contains species producing a diversity of biologically active secondary metabolites with interesting drug-like properties (Charria-Girón et al., 2022;Noumeur et al., 2020), and the highest known number of thermophilic species, which have large industrial relevance (Hutchinson et al., 2019;Patel & Rawat, 2021;van den Brink et al., 2015). ...
... Most species are saprobes and can occur in a broad range of habitats, such as soil, air, compost, animal dung, plants tissues, seeds, and indoor environments (Rodríguez et al. 2002;Wang et al. 2016a, b;Zhang et al. 2017aZhang et al. , b, 2021Sousa et al. 2020). They were also identified as endophytes (Noumeur et al. 2020;Mehrabi et al. 2020) or opportunistic pathogens of humans (Ryan et al. 2021). Some members of the Chaetomiaceae are also known as remarkable secondary metabolite producers with antimicrobial, antioxidant, anticancer, or cytotoxic properties (Wang et al. 2017;Jiang et al. 2021;Promgool et al. 2022;Tavares et al. 2022). ...
Article
Interest in cave fungal diversity is flourishing because it may represent a reservoir of new species and metabolites. However, the mycobiota remains poorly studied in the underground environment, especially in neotropical regions. During surveys that aimed to investigate the fungal diversity in quartzite and limestone caves in the Southern Espinhaço Mountain in Brazil, six Chaetomiaceae isolates were obtained from different cave substrates. Five taxonomical novelties of Chaetomiaceae in Brazilian caves were discovered based on phylogenetic analyses using DNA sequences from the ITS, LSU, TUB, and RPB2 genes. Chaetomium meridionalense, Pseudohumicola alba, and Pseudohumicola lutea are new species found in Gruta da Extração and Gruta Velha Nova caves. Parahumicola is introduced as a new genus representing a novel phylogenetic lineage with unique morphological characteristics in the family Chaetomiaceae. This new monotypic genus is typified by P. guana, which was found in a bat guano sample in the Gruta Monte Cristo cave. Furthermore, this is the first report of Collariella bostrychodes in a neotropical cave. Overall, these findings emphasise that Brazilian caves constitute an untapped source of fungal resources.
... The mycelial culture was deposited at Deutsche Sammlung von Mikroorganismen und Zellkulturen (DSMZ) in Braunschweig under the designation number DSM 105465. The fungus was identified by morphological studies and sequencing of the ITS rDNA (5.8S gene region, the internal transcribed spacers ITS1 and ITS2) and LSU (large subunit) ribosome RNA genes, according to the well established procedure of Noumeur et al. [23]. The genomic DNA sequence was deposited to GenBank under the accession number MK463979. ...
Article
Full-text available
In our continued search for biologically active metabolites from cultures of rare Basidiomycota species, we found eight previously undescribed cyathane-xylosides from submerged cultures of Dentipellis fragilis, which were named dentifragilins A–H. In addition, the known cyathane deriv-atives striatal D and laxitextine A were isolated. All compounds were characterized by high-resolution electrospray ionization mass spectrometry (HR-ESIMS) as well as by 1D and 2D nuclear magnetic resonance (NMR) spectroscopy. Several of the compounds exhibited significant activities in standardized cell-based assays for the determination of antimicrobial and cytotoxic effects. The discovery of cyathanes in the genus Dentipellis has chemotaxonomic implications, as this class of diterpenoids has already been shown to be characteristic for mycelial cultures of the related genera Hericium and Laxitextum, which are classified as Dentipellis in the family Heri-ciaceae.
... Within the phylum Ascomycota, the order Sordariales is one of the largest and most diverse taxonomic groups within the class Sordariomycetes, and contains soil-borne, lignicolous, herbicolous, and coprophilous taxa currently arranged among seven families (Lundquist 1972;Hawksworth 1986;Huhndorf et al. 2004;Marin-Felix et al. 2020). It also represents a source of prolific producers of a great diversity of biologically active secondary metabolites with interesting drug-like properties (Guo et al. 2019;Noumeur et al. 2020). As an example, the antifungal diterpene glycosides, the sordarins, have attracted considerable interest due to their specific action against elongation factor 2 in protein biosynthesis in fungi (Dominguez and Martin 1998;Vicente et al. 2009). ...
Article
Full-text available
Ascomycetes belonging to the order Sordariales are a well-known reservoir of secondary metabolites with potential beneficial applications. Species of the Sordariales are ubiquitous, and they are commonly found in soils and in lignicolous, herbicolous, and coprophilous habitats. Some of their species have been used as model organisms in modern fungal biology or were found to be prolific producers of potentially useful secondary metabolites. However, the majority of sordarialean species are poorly studied. Traditionally, the classification of the Sordariales has been mainly based on morphology of the ascomata, ascospores, and asexual states, characters that have been demonstrated to be homoplastic by modern taxonomic studies based on multi-locus phylogeny. Herein, we summarize for the first time relevant information about the available knowledge on the secondary metabolites and the biological activities exerted by representatives of this fungal order, as well as a current outlook of the potential opportunities that the recent advances in omic tools could bring for the discovery of secondary metabolites in this order.
Article
Full-text available
The field of mycology has grown from an underappreciated subset of botany, to a valuable, modern scientific discipline. As this field of study has grown, there have been significant contributions to science, technology, and industry, highlighting the value of fungi in the modern era. This paper looks at the current research, along with the existing limitations, and suggests future areas where scientists can focus their efforts, in the field mycology. We show how fungi have become important emerging diseases in medical mycology. We discuss current trends and the potential of fungi in drug and novel compound discovery. We explore the current trends in phylogenomics, its potential, and outcomes and address the question of how phylogenomics can be applied in fungal ecology. In addition, the trends in functional genomics studies of fungi are discussed with their importance in unravelling the intricate mechanisms underlying fungal behaviour, interactions, and adaptations, paving the way for a comprehensive understanding of fungal biology. We look at the current research in building materials, how they can be used as carbon sinks, and how fungi can be used in biocircular economies. The numbers of fungi have always been of great interest and have often been written about and estimates have varied greatly. Thus, we discuss current trends and future research needs in order to obtain more reliable estimates. We address the aspects of machine learning (AI) and how it can be used in mycological research. Plant pathogens are affecting food production systems on a global scale, and as such, we look at the current trends and future research needed in this area, particularly in disease detection. We look at the latest data from High Throughput Sequencing studies and question if we are still gaining new knowledge at the same rate as before. A review of current trends in nanotechnology is provided and its future potential is addressed. The importance of Arbuscular Mycorrhizal Fungi is addressed and future trends are acknowledged. Fungal databases are becoming more and more important, and we therefore provide a review of the current major databases. Edible and medicinal fungi have a huge potential as food and medicines, especially in Asia and their prospects are discussed. Lifestyle changes in fungi (e.g., from endophytes, to pathogens, and/or saprobes) are also extremely important and a current research trend and are therefore addressed in this special issue of Fungal Diversity.
Article
Full-text available
In this study, 15 Lulworthiales strains isolated from the marine tunicate Halocynthia papillosa collected in the central Tyrrhenian Sea were characterized using a polyphasic approach (morpho-physiological, molecular, and phylogenetic analyses). Based on multi-locus phylogenetic inference and morphological characters, a new genus, Rambellisea, and two new species, R. halocynthiae and R. gigliensis (Lulworthiales), were proposed. Multi-locus phylogenetic analyses using the nuclear ribosomal regions of DNA (nrITS1-nr5.8S-nrITS2, nrLSU, and nrSSU) sequence data strongly supported the new taxa. Phylogenetic inference, estimated using Maximum Likelihood and Bayesian Inference, clearly indicates that Rambellisea gen. nov. forms a distinct clade within the order Lulworthiales. Moreover, the two new species were separated into distinct subclades, solidly supported by the analyses. This is the first report of Lulworthiales species isolated from animals.
Article
Full-text available
Abundisporin A (1), together with seven previously undescribed drimane sesquiterpenes named abundisporins B−H (2−8), were isolated from a polypore, Abundisporus violaceus MUCL 56355 (Polyporaceae), collected in Kenya. Chemical structures of the isolated compounds were elucidated based on exhaustive 1D and 2D NMR spectroscopic measurements and supported by HRESIMS data. The absolute configurations of the isolated compounds were determined by using Mosher's method for 1−4 and TDDFT-ECD calculations for 4 and 5−8. None of the isolated compounds exhibited significant activities in either antimicrobial or cytotoxicity assays. Notably, all of the tested compounds demonstrated neurotrophic effects, with 1 and 6 significantly increasing outgrowth of neurites when treated with 5 ng/mL NGF.
Article
Full-text available
This paper provides an updated classification of the Kingdom Fungi (including fossil fungi) and fungus-like taxa. Five-hundred and twenty-three (535) notes are provided for newly introduced taxa and for changes that have been made since the previous outline. In the discussion, the latest taxonomic changes in Basidiomycota are provided and the classification of Mycosphaerellales are broadly discussed. Genera listed in Mycosphaerellaceae have been confirmed by DNA sequence analyses, while doubtful genera (DNA sequences being unavailable but traditionally accommodated in Mycosphaerellaceae) are listed in the discussion. Problematic genera in Glomeromycota are also discussed based on phylogenetic results.
Article
Full-text available
Rosellinia (Xylariaceae) is a large, cosmopolitan genus comprising over 130 species that have been defined based mainly on the morphology of their sexual morphs. The genus comprises both lignicolous and saprotrophic species that are frequently isolated as endophytes from healthy host plants, and important plant pathogens. In order to evaluate the utility of molecular phylogeny and secondary metabolite profiling to achieve a better basis for their classification, a set of strains was selected for a multi-locus phylogeny inferred from a combination of the sequences of the internal transcribed spacer region (ITS), the large subunit (LSU) of the nuclear rDNA, beta-tubulin (TUB2) and the second largest subunit of the RNA polymerase II (RPB2). Concurrently, various strains were surveyed for production of secondary metabolites. Metabolite profiling relied on methods with high performance liquid chromatography with diode array and mass spectrometric detection (HPLC-DAD/MS) as well as preparative isolation of the major components after re-fermentation followed by structure elucidation using nuclear magnetic resonance (NMR) spectroscopy and high resolution mass spectrometry (HR-MS). Two new and nine known isopimarane diterpenoids were identified during our mycochemical studies of two selected Dematophora strains and the metabolites were tested for biological activity. In addition, the nematicidal cyclodepsipeptide PF1022 A was purified and identified from a culture of Rosellinia corticium, which is the first time that this endophyte-derived drug precursor has been identified unambiguously from an ascospore-derived isolate of a Rosellinia species. While the results of this first HPLC profiling were largely inconclusive regarding the utility of secondary metabolites as genus-specific chemotaxonomic markers, the phylogeny clearly showed that species featuring a dematophora-like asexual morph were included in a well-defined clade, for which the genus Dematophora is resurrected. Dematophora now comprises all previously known important plant pathogens in the genus such as D. arcuata, D. bunodes, D. necatrix and D. pepo, while Rosellinia s. str. comprises those species that are known to have a geniculosporium-like or nodulisporium-like asexual morph, or where the asexual morph remains unknown. The extensive morphological studies of L.E. Petrini served as a basis to transfer several further species from Rosellinia to Dematophora, based on the morphology of their asexual morphs. However, most species of Rosellinia and allies still need to be recollected in fresh state, cultured, and studied for their morphology and their phylogenetic affinities before the infrageneric relationships can be clarified.
Article
Full-text available
The genus Thielavia is morphologically defined by having non-ostiolate ascomata with a thin peridium composed of textura epidermoidea, and smooth, singlecelled, pigmented ascospores with one germ pore. Thielavia is typified with Th. basicola that grows in close association with a hyphomycete which was traditionally identified as Thielaviopsis basicola. Besides Th. basicola exhibiting the mycoparasitic nature, the majority of the described Thielavia species are from soil, and some have economic and ecological importance. Unfortunately, no living type material of Th. basicola exists, hindering a proper understanding of the classification of Thielavia. Therefore, Thielavia basicola was neotypified by material of a mycoparasite presenting the same ecology and morphology as described in the original description. We subsequently performed a multi-gene phylogenetic analyses (rpb2, tub2, ITS and LSU) to resolve the phylogenetic relationships of the species currently recognised in Thielavia. Our results demonstrate that Thielavia is highly polyphyletic, being related to three family-level lineages in two orders. The redefined genus Thielavia is restricted to its type species, Th. basicola, which belongs to the Ceratostomataceae (Melanosporales) and its host is demonstrated to be Berkeleyomyces rouxiae, one of the two species in the “Thielaviopsis basicola” species complex. The new family Podosporaceae is sister to the Chaetomiaceae in the Sordariales and accommodates the re-defined genera Podospora, Trangularia and Cladorrhinum, with the last genus including two former Thielavia species (Th. hyalocarpa and Th. intermedia). This family also includes the genetic model species Podospora anserina, which was combined in Triangularia (as Triangularia anserina). The remaining Thielavia species fall in ten unrelated clades in the Chaetomiaceae, leading to the proposal of nine new genera (Carteria, Chrysanthotrichum, Condenascus, Hyalosphaerella, Microthielavia, Parathielavia, Pseudothielavia, Stolonocarpus and Thermothielavioides). The genus Canariomyces is transferred from Microascaceae (Microascales) to Chaetomiaceae based on its type species Can. notabilis. Canariomyces is closely related to the human-pathogenic genus Madurella, and includes three thielavia-like species and one novel species. Three monotypic genera with a chaetomium-like morph (Brachychaeta, Chrysocorona and Floropilus) are introduced to better resolve the Chaetomiaceae and the thielavia-like species in the family. Chrysocorona lucknowensis and Brachychaeta variospora are closely related to Acrophialophora and three newly introduced genera containing thielavia-like species; Floropilus chiversii is closely related to the industrially important and thermophilic species Thermothielavioides terrestris (syn. Th. terrestris). This study shows that the thielavia-like morph is a homoplastic form that originates from several separate evolutionary events. Furthermore, our results provide new insights into the taxonomy of Sordariales and the polyphyletic Lasiosphaeriaceae.
Article
Full-text available
Fungi are an understudied, biotechnologically valuable group of organisms. Due to the immense range of habitats that fungi inhabit, and the consequent need to compete against a diverse array of other fungi, bacteria, and animals, fungi have developed numerous survival mechanisms. The unique attributes of fungi thus herald great promise for their application in biotechnology and industry. Moreover, fungi can be grown with relative ease, making production at scale viable. The search for fungal biodiversity, and the construction of a living fungi collection, both have incredible economic potential in locating organisms with novel industrial uses that will lead to novel products. This manuscript reviews fifty ways in which fungi can potentially be utilized as biotechnology. We provide notes and examples for each potential exploitation and give examples from our own work and the work of other notable researchers. We also provide a flow chart that can be used to convince funding bodies of the importance of fungi for biotechnological research and as potential products. Fungi have provided the world with penicillin, lovastatin, and other globally significant medicines, and they remain an untapped resource with enormous industrial potential. Keywords Biocontrol · Biodiversity · Biotechnology · Food · Fungi · Mushrooms
Article
Full-text available
Endophytes are microbes (mostly bacteria and fungi) present asymptomatically in plants. Endophytic microbes are often functional in that they may carry nutrients from the soil into plants, modulate plant development, increase stress tolerance of plants, suppress virulence in pathogens, increase disease resistance in plants, and suppress development of competitor plant species. Endophytic microbes have been shown to: (i) obtain nutrients in soils and transfer nutrients to plants in the rhizophagy cycle and other nutrient‐transfer symbioses; (ii) increase plant growth and development; (iii) reduce oxidative stress of hosts; (iv) protect plants from disease; (v) deter feeding by herbivores; and (vi) suppress growth of competitor plant species. Because of the effective functions of endophytic microbes, we suggest that endophytic microbes may significantly reduce use of agrochemicals (fertilizers, fungicides, insecticides, and herbicides) in the cultivation of crop plants. The loss of endophytic microbes from crop plants during domestication and long‐term cultivation could be remedied by transfer of endophytes from wild relatives of crops to crop species. Increasing atmospheric carbon dioxide levels could reduce the efficiency of the rhizophagy cycle due to repression of reactive oxygen used to extract nutrients from microbes in roots. © 2019 The Authors. Pest Management Science published by John Wiley & Sons Ltd on behalf of Society of Chemical Industry.
Article
Full-text available
4-Hydroxypleurogrisein, a congener of the anticancer-lead compound pleurotin, as well as six further derivatives were isolated from the basidiomycete Hohenbuehelia grisea, strain MFLUCC 12-0451. The structures were elucidated utilizing high resolution electron spray ionization mass spectrometry (HRESIMS) and 1D and 2D nuclear magnetic resonance (NMR) spectral data and evaluated for their biological activities; for leucopleurotin, we provide Xray data. While most congeners showed moderate antimicrobial and cytotoxic activity, 4-hydroxypleurogrisein emerged as an inhibitor of hepatitis C virus infectivity in mammalian liver cells.
Article
Full-text available
The traditional concept of the genus Humicola includes species that produce pigmented, thick-walled and single-celled spores laterally or terminally on hyphae or minimally differentiated conidiophores. More than 50 species have been described in the genus. Species commonly occur in soil, indoor environments, and compost habitats. The taxonomy of Humicola and morphologically similar genera is poorly understood in modern terms. Based on a four-locus phylogeny, the morphological concept of Humicola proved to be polyphyletic. The type of Humicola, H. fuscoatra, belongs to the Chaetomiaceae. In the Chaetomiaceae, species producing humicola-like thick-walled spores are distributed among four lineages: Humicola sensu stricto, Mycothermus, Staphylotrichum, and Trichocladium. In our revised concept of Humicola, asexual and sexually reproducing species both occur. The re-defined Humicola contains 24 species (seven new and thirteen new combinations), which are described and illustrated in this study. The species in this genus produce conidia that are lateral, intercalary or terminal on/in hyphae, and conidiophores are not formed or are minimally developed (micronematous). The ascospores of sexual Humicola species are limoniform to quadrangular in face view and bilaterally flattened with one apical germ pore. Seven species are accepted in Staphylotrichum (four new species, one new combination). Thick-walled conidia of Staphylotrichum species usually arise either from hyphae (micronematous) or from apically branched, seta-like conidiophores (macronematous). The sexual morph represented by Staphylotrichum longicolleum (= Chaetomium longicolleum) produces ascomata with long necks composed of a fused basal part of the terminal hairs, and ascospores that are broad limoniform to nearly globose, bilaterally flattened, with an apical germ pore. The Trichocladium lineage has a high morphological diversity in both asexual and sexual structures. Phylogenetic analysis revealed four subclades in this lineage. However, these subclades are genetically closely related, and no distinctive phenotypic characters are linked to any of them. Fourteen species are accepted in Trichocladium, including one new species, twelve new combinations. The type species of Gilmaniella, G. humicola, belongs to the polyphyletic family Lasiosphaeriaceae (Sordariales), but G. macrospora phylogenetically belongs to Trichocladium. The thermophilic genus Mycothermus and the type species My. thermophilum are validated, and one new Mycothermus species is described. Phylogenetic analyses show that Remersonia, another thermophilic genus, is sister to Mycothermus and two species are known, including one new species. Thermomyces verrucosus produces humicola-like conidia and is transferred to Botryotrichum based on phylogenetic affinities. This study is a first attempt to establish an inclusive modern classification of Humicola and humicola-like genera of the Chaetomiaceae. More research is needed to determine the phylogenetic relationships of “humicola”-like species outside the Chaetomiaceae.
Article
Full-text available
Nothing is sterile. Insects, plants, and fungi, highly speciose groups of organisms, conceal a vast fungal biodiversity. An approximation of the total number of fungal species on Earth remains an elusive goal, but estimates should include fungal species hidden in associations with other organisms. Some specific roles have been discovered for the fungi hidden within other life forms, including contributions to nutrition, detoxification of foodstuffs, and production of volatile organic compounds. Fungi rely on associates for dispersal to fresh habitats and, under some conditions, provide them with competitive advantages. New methods are available to discover microscopic fungi that previously have been overlooked. In fungal conservation efforts, it is essential not only to discover hidden fungi but also to determine if they are rare or actually endangered.
Article
Molecular phylogenetic studies of cultures derived from some specimens of plant-inhabiting Sordariomycetes using ITS, LSU, rpb2 and tub2 DNA sequence data revealed close affinities to strains of Muscodor. The taxonomy of this biotechnologically important genus, which exclusively consists of endophytes with sterile mycelia that produce antibiotic volatile secondary metabolites, was based on a rather tentative taxonomic concept. Even though it was accommodated in Xylariaceae, its phylogenetic position had so far remained obscure. Our phylogeny shows that Muscodor species have affinities to the xylarialean genera Emarcea and Induratia, which is corroborated by the fact that their sexual states produce characteristic apiospores. These data allow for the integration of Muscodor in Induratia, i.e. the genus that was historically described first. The multi-locus phylogenetic tree clearly revealed that a clade comprising Emarcea and Induratia forms a monophylum separate from representatives of Xylariaceae, for which we propose the new family Induratiaceae. Divergence time estimations revealed that Induratiaceae has been diverged from the Xylariaceae + Clypeosphaeriaceae clade at 93 (69–119) million years ago (Mya) with the crown age of 61 (39–85) Mya during the Cretaceous period. The ascospore-derived cultures were studied for the production of volatile metabolites, using both, dual cultures for assessment of antimicrobial effects and extensive analyses using gas chromatography coupled with mass spectrometry (GC–MS). The antimicrobial effects observed were significant, but not as strong as in the case of the previous reports on Muscodor species. The GC–MS results give rise to some doubt on the validity of the previous identification of certain volatiles. Many peaks in the GC–MS chromatograms could not be safely identified by database searches and may represent new natural products. The isolation of these compounds by preparative chromatography and their subsequent characterisation by nuclear magnetic resonance (NMR) spectroscopy or total synthesis will allow for a more concise identification of these volatiles, and they should also be checked for their individual contribution to the observed antibiotic effects. This will be an important prerequisite for the development of biocontrol strains.
Article
Fermentation of the fungal strain Skeletocutis sp. originating from Mount Elgon Natural Reserve in Kenya, followed by bioassay guided fractionation led to the isolation of twelve previously undescribed metabolites named skeletocutins A-L (1-5, 7-13) together with the known tyromycin A (6). Their structures were assigned by NMR spectroscopy complemented by HR-ESIMS. Compounds 1-6 and 11-13 exhibited selective activities against Gram-positive bacteria, while compound 10 weakly inhibited the formation of biofilm of Staphylococcus aureus. The isolated metabolites were also evaluated for inhibition of L-leucine aminopeptidase, since tyromycin A had previously been reported to possess such activities but only showed weak effects. Furthermore, all compounds were tested for antiviral activity against Hepatitis C virus (HCV), and compound 6 moderately inhibited HCV infectivity with an IC50 of 6.6µM.
Article
Four new diketopiperazine alkaloids, citriperazines A-D were isolated from algae-derived Penicillium sp. KMM 4672. The structures of compounds 1–4 were determined using spectroscopic methods. The absolute configurations of compounds 1 and 4 were established by comparison of calculated and experimental ECD spectra. The cytotoxicity of compounds 1–4 against several human prostate cell lines was evaluated. © 2018