ArticlePDF Available

Metabolomic-guided discovery of cyclic nonribosomal peptides from Xylaria ellisii sp. nov., a leaf and stem endophyte of Vaccinium angustifolium

Authors:

Abstract and Figures

Fungal endophytes are sources of novel bioactive compounds but relatively few agriculturally important fruiting plants harboring endophytes have been carefully studied. Previously, we identified a griseofulvin-producing Xylaria species isolated from Vaccinium angustifolium, V. corymbosum, and Pinus strobus. Morphological and genomic analysis determined that it was a new species, described here as Xylaria ellisii. Untargeted high-resolution LC-MS metabolomic analysis of the extracted filtrates and mycelium from 15 blueberry isolates of this endophyte revealed differences in their metabolite profiles. Toxicity screening of the extracts showed that bioactivity was not linked to production of griseofulvin, indicating this species was making additional bioactive compounds. Multivariate statistical analysis of LC-MS data was used to identify key outlier features in the spectra. This allowed potentially new compounds to be targeted for isolation and characterization. This approach resulted in the discovery of eight new proline-containing cyclic nonribosomal peptides, which we have given the trivial names ellisiiamides A-H. Three of these peptides were purified and their structures elucidated by one and two-dimensional nuclear magnetic resonance spectroscopy (1D and 2D NMR) and high-resolution tandem mass spectrometry (HRMS/MS) analysis. The remaining five new compounds were identified and annotated by high-resolution mass spectrometry. Ellisiiamide A demonstrated Gram-negative activity against Escherichia coli BW25113, which is the first reported for this scaffold. Additionally, several known natural products including griseofulvin, dechlorogriseofulvin, epoxy/cytochalasin D, zygosporin E, hirsutatin A, cyclic pentapeptides #1–2 and xylariotide A were also characterized from this species.
Content may be subject to copyright.
1
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
Metabolomic-guided discovery
of cyclic nonribosomal peptides
from Xylaria ellisii sp. nov., a leaf
and stem endophyte of Vaccinium
angustifolium
Ashraf Ibrahim1,7, Joey B. Tanney2,3,4, Fan Fei1, Keith A. Seifert4, G. Christopher Cutler5,
Alfredo Capretta1, J. David Miller2 & Mark W. Sumarah2,6*
Fungal endophytes are sources of novel bioactive compounds but relatively few agriculturally
important fruiting plants harboring endophytes have been carefully studied. Previously, we identied a
griseofulvin-producing Xylaria species isolated from Vaccinium angustifolium, V. corymbosum, and Pinus
strobus. Morphological and genomic analysis determined that it was a new species, described here as
Xylaria ellisii. Untargeted high-resolution LC-MS metabolomic analysis of the extracted ltrates and
mycelium from 15 blueberry isolates of this endophyte revealed dierences in their metabolite proles.
Toxicity screening of the extracts showed that bioactivity was not linked to production of griseofulvin,
indicating this species was making additional bioactive compounds. Multivariate statistical analysis
of LC-MS data was used to identify key outlier features in the spectra. This allowed potentially new
compounds to be targeted for isolation and characterization. This approach resulted in the discovery
of eight new proline-containing cyclic nonribosomal peptides, which we have given the trivial names
ellisiiamides A-H. Three of these peptides were puried and their structures elucidated by one and two-
dimensional nuclear magnetic resonance spectroscopy (1D and 2D NMR) and high-resolution tandem
mass spectrometry (HRMS/MS) analysis. The remaining ve new compounds were identied and
annotated by high-resolution mass spectrometry. Ellisiiamide A demonstrated Gram-negative activity
against Escherichia coli BW25113, which is the rst reported for this scaold. Additionally, several
known natural products including griseofulvin, dechlorogriseofulvin, epoxy/cytochalasin D, zygosporin
E, hirsutatin A, cyclic pentapeptides #1–2 and xylariotide A were also characterized from this species.
Vaccinium angustifolium (wild lowbush blueberries or commonly wild blueberries) were consumed fresh and
preserved for the winter by the Indigenous peoples of northeastern North America and rapidly incorporated into
the diets of European settlers in Canada from the early 17th century1,2. Today, blueberries comprise more than
half of all fruit production in Canada. Wild blueberries oen grow in forests where Pinus strobus (eastern white
pine) is the dominant tree species. Eastern white pine is an economically, ecologically, and culturally important
keystone tree species in eastern N. American forests, especially for bird species3,4.
Endophytes are an ecological category of phylogenetically diverse fungi that can asymptomatically colonize
healthy plant tissues. Ascomycetous endophytes of various species of Vaccinium have been reported over the
past three decades. is includes from surface-sterilized tissues of Vaccinium vitis-idaea (lingonberry, European
1Department of Chemistry and Chemical Biology, McMaster University, Hamilton, Ontario, L8S 4M1, Canada.
2Department of Chemistry, Carleton University, Ottawa, Ontario, K1S 5B6, Canada. 3Pacic Forestry Centre, Canadian
Forest Service, Natural Resources Canada, Victoria, British Columbia, V8Z 1M5, Canada. 4Ottawa Research and
Development Centre, Agriculture and Agri-Food Canada, Ottawa, Ontario, K1A 0C6, Canada. 5Department of Plant,
Food, and Environmental Sciences, Faculty of Agriculture, Dalhousie University, Truro, NS, B2N 5E3, Canada. 6London
Research and Development Centre, Agriculture and Agri-Food Canada, London, Ontario, N5V 4T3, Canada. 7Present
address: LifeMine Therapeutics, Cambridge, Massachusetts, 02140, USA. *email: mark.sumarah@canada.ca
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
2
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
blueberry) and V. myrtillus (bilberry, whortleberry) in Europe5, V. dunalianum var. urophyllum (South China
blueberry) in China6, as well as from stems of V. macrocarpon (cranberries) and V. corymbosum (northern high-
bush blueberry) in New Jersey7,8. ere is some evidence of the same endophyte species occurring in both conifer
and Vaccinium species, e.g. Nemania diusa (Xylariaceae)6,9,10 and Phacidiaceae species such as Allantophomopsis
lycopodina, Phacidium lacerum, and Strasseria geniculata1115. Indirect evidence of a conifer-Vaccinium shared
endophyte includes the aquatic hyphomycete Dwayaangam colodena, a common needle endophyte of Picea spp.,
which was reported from rainwater collected from foliage of Picea abies, Pinus sylvestris, and Vaccinium myrtillus
in Europe1619. Discovery of an aquatic hyphomycete conifer endophyte and reports of hardwood saprotrophs as
conifer endophytes (e.g. Phialocephala piceae20) are evocative of more complex interactions between endophytes
and their environment.
Endophytes belonging to the family Xylariaceae (Xylariales, Sordariomycetes) are ubiquitous and detected in
varying abundance in most studies involving woody plants, regardless of geographic location or host, whether
by isolation of cultures or by studies of DNA, oen exhibiting little host preference and including known sapro-
trophs2124. Xylariaceae endophytes are common but dicult to identify to species because of a lack of reference
sequences and the limited taxonomic resolution of the asexual states (usually the only morphological characters
produced in vitro). However, careful eld observations can provide connections between the oen conspicu-
ous Xylariaceae stromata found in nature and the corresponding endophytes isolated in culture or detected by
DNA sequences from the same forests2325. Taxonomically, Xylariaceae comprises at least 37 genera with likely
more than 1,000 species26. Many endophyte studies based on morphological identication of cultures report
geniculosporium-like morphs attributable to Anthostomella, Rosellinia and Xylaria species2731. e classical
nature of most taxonomic studies of Xylariaceae is reected by the need for the sexual state to conrm identica-
tion, with a relative paucity of species-specic DNA barcodes and phylogenetic markers compared to many other
ascomycete groups. us, xylariaceous endophytes may include species and genera known to classical taxonomy
but not included in sequence databases (i.e.: named-but-unsequenced species).
Species of Xylariaceae are a rich source of secondary metabolites, and chemotaxonomy is oen part of taxo-
nomic studies. Species in this family can produce diverse metabolites from multiple biosynthetic families includ-
ing dihydroisocoumarins, punctaporonins, cytochalasins, butyrolactones, and succinic acid derivatives32,33.
Exploration of Xylaria metabolites using newer chemical methods led to discovery of a broad array of metabolites
from both tissues of stromata and culture extracts34.
Although there have been many studies of metabolites from fungal endophytes35,36, there are few reports from
endophytes of Vaccinium37,38. We previously described production of the antifungal compounds griseofulvin and
piliformic acid from an unknown Xylaria species isolated as a foliar endophyte from wild blueberry in natural and
commercial sites, and from white pine. Aer the Richardson et al. (2014) study, we continued to isolate the same
unidentied species of Xylaria as an endophyte of white pine needles and as an endophyte of leaves and stems
of both wild and highbush blueberry at three dierent locations in Nova Scotia, New Brunswick, and Ontario,
Canada38. We also conducted eld sampling to discover the putative sexual state of this unknown Xylaria species.
is would provide information on morphological characters of sexual structures, permitting its identication.
is previously unknown endophyte is described here as Xylaria ellisii based on morphological and genomic
evidence. Representative sequences in NCBI GenBank from other studies indicate that X. ellisii has been isolated
many times as an unidentied endophyte from a wide variety of plant hosts, allowing us to infer additional infor-
mation about its distribution, biology, and chemistry.
In our eort to discover novel natural products, we applied a LC-MS metabolomic-guided discovery approach
to these Xylaria strains from wild and highbush blueberry plants (Fig.1). is approach allows for a global survey
of small molecule metabolites from an extract and visual representation of metabolite variances between group-
ings or extracts. us, discriminating between like and dierent features allows extracts to be prioritized for fur-
ther investigation39,40. Fieen strains were grown on two media and the resulting ethyl acetate extracted ltrates
and mycelium were screened using standardized LC-UV/MS conditions. Multivariate statistical analysis was used
to organize resulting analytical data to reveal extracts that appeared to have dierences in their major secondary
metabolites. is approach led to the discovery of a family of eight new proline-containing cyclic nonribosomal
pentapeptides named ellisiiamides A–H. Ellisiiamide A is an alanine (Ala) substituted variant, a rst report for
this scaold, and demonstrated modest activity against Escherichia coli.
Results
Identication, biology and ecology of Xylaria sp. Approximately 30 strains of Xylaria sp. were isolated
from surface-sterilized blueberry tissues collected from highbush and wild blueberry elds within a ~300 ×
100 km triangular area. All elds were surrounded by forested lands. Preliminary phylogenetic analysis using the
internal transcribed spacer (ITS) barcode combined with morphological features conrmed conspecicity of iso-
lated endophytic Xylaria sp. strains. However, identication of the strains to species was not possible using molec-
ular or in vitro morphological data. Based on a BLAST query of the Xylaria sp. ITS and RPB2 sequences with
available GenBank sequences, the endophyte strains were closest related to sequences identied as Xylaria berteri,
X. castorea, X. cubensis, X. laevis, and X. longipes, species that form conspicuous sexual reproductive structures
(stromata) from decaying hardwood. Given the close phylogenetic relationship of the unknown Xylaria endo-
phyte to these species and evidence of prevalent endophytic-saprotrophic life histories within Xylariaceae2325,41,42,
we inferred that the unknown Xylaria endophyte likely produces stromata from decaying hardwood in mixed-
wood stands in the Acadian forest. us, Xylaria stromata were selectively sampled during ongoing eld surveys
to collect the putative sexual state of the endophyte. is would provide material for identication and insight
into its life history20,43.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
3
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
A Xylaria sp. producing stroma reminiscent of X. corniformis and X. curta was collected from decaying, oen
partially buried, Acer saccharum branches or logs in late summer and autumn. Sequences (ITS, SSU, LSU, BenA,
EF1-α, RPB2) obtained from stromatal tissue and ascospore cultures were identical to those obtained from
the Xylaria sp. endophyte cultures, indicating they are conspecic and evincing a saprotrophic-endophytic life
history. Based on morphological study of the stromata, this species is equivalent to X. corniformis var. obovata
Sacc., Xylaria corniformis sensu Laessøe44, and Xylaria curta sensu Rogers45. From the RPB2 phylogeny, X. corni-
formis var. obovata is weakly supported (posterior probability value (PP) = 0.56) sister to X. laevis and other
species within the strongly-supported (PP = 1.0) X. cubensis aggregate clade. Xylaria is polyphyletic, including
Amphirosellinia nigrospora, Stilbohypoxylon quisquiliarum, and Nemania serpens, and the type species (X. hypox-
ylon) occurs in a basal clade sister to X. bambusicola. Additional RPB2 sequences for related Xylaria species are
needed to generate a more comprehensive phylogeny (Fig.2).
Several DAOMC herbarium specimens identied as X. corniformis from Acer spp. wood in Ontario and
Quebec were morphologically similar to X. laevis. e resulting ITS sequences from these specimens showed that
they formed a clade sister to X. longipes and X. primorskensis and were distinct from the griseofulvin-producing
X. corniformis var. obovata (Fig.3). We support the distinction of X. corniformis var. obovata from X. corni-
formis, and thus describe a new species, Xylaria ellisii, to accommodate its novelty and fulll the need to delineate
boundaries in species complexes with robust species concepts connected to authenticated reference sequences
and specimens.
LC-MS analysis of culture extracts and multivariate data analysis. Fieen strains of X. ellisii were
subject to further study: four from cultivated highbush blueberry plants and 11 from wild blueberry plants. Ethyl
acetate extracts of the culture ltrate and associated mycelium were screened using standardized LC-UV/MS
conditions.
In order to identify unique secondary metabolite dierences between extracts of Xylaria isolates of highbush
and wild blueberry plants we compared the extracted ltrates and mycelium with three dierent pair-wise com-
parisons. ese comparisons included: ethyl acetate extracts of Xylaria strains grown on 2% malt extract broth
(ML) versus those grown in potato dextrose broth (PDB) cultures; ML media cultures of highbush versus wild
varieties; and, PDB medium cultures of highbush versus wild isolates (Fig.1S). A supervised multivariant analysis
Figure 1. Discovery of new griseofulvin-producing fungal endophyte species Xylaria ellisii isolated from
highbush and wild blueberry leaves and stems. (A) Isolation and culturing of fungal endophyte, (B) LC-UV
comparative prole analysis of crude ltrate extracts at λ 210 nm, revealing dierences in metabolite
production, (C) Most likely tree from a RAxML analysis of ITS dataset containing representative endophytes.
Culture numbers precede the species name and RAxML bootstrap support percentages 50 from a summary
of 1000 replicates are presented at the branch nodes. is tree was rooted with Mucor ellipsoideus (ATCC MYA-
4767; NR_111683) and the scale bar represents the number of substitutions per site.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
4
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
method, Orthogonal Partial Least Squares Discriminant Analysis (OPLS-DA), was used to identify outlier metab-
olites biosynthesized under the dierent culture conditions tested. OPLS-DA correlates dierences in secondary
metabolite feature abundances (X variables) to various treatment groups (Y variables) by identifying principle
components that describe dierences. R2X, R2Y, and Q2 parameters are important validation parameters used for
OPLS-DA, where R2X and R2Y describes the percentage of X and Y variables described by the model (Fig.4 and
Supplementary Fig.2S). A valid model is dened as having a prediction statistic of Q2 > 0.4, with values above 0.7
being highly signicant46. Metabolite features with a high Variable Importance in Projection (VIP) scores (>0.7)
are responsible for driving the dierences between treatment groups, and these values are considered signi-
cant47. eir metabolic features can be viewed at both ends of the OPLS-DA S-plot.
Fractions with VIP scores above 0.7 were selected for further study and compounds were identied where pos-
sible. OPLS-DA validation parameters for each of the extracted ltratesand myceliummetabolite models tested
are summarized in Table1 and Supplementary Table1S. In total, 3856 metabolite features were identiedfrom
Figure 2. Bayesian 50% majority rule RPB2 consensus tree containing Xylaria ellisii and related species. All
unlabeled branches have Bayesian posterior probability values of 1.0; values lower than 1.0 are presented at
nodes. e tree was rooted to Barrmaelia rhamnicola (CBS:142772) and the scalebar indicates the expected
number ofchanges per site. Strain numbers follow species names and type specimens are indicated in bold.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
5
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
Figure 3. Bayesian 50% majority rule ITS consensus tree containing Xylaria ellisii and related species. All unlabeled
branches have Bayesian posterior probability values of 1.0; values lower than 1.0 are presented at nodes. e tree was
rooted to Nemania serpens and the scalebar indicates the expected number ofchanges per site. GenBank accession
numbers and host information follow species names (when applicable). Type specimens are indicated in bold.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
6
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
the extracted ltrates, with a Q2 value of 0.615 for ML versus PDB, and Q2 values of 0.778 for ML and PDB, as well
as highbush versus wild varieties.
Metabolomic-guided discovery and metabolite identication of knowns 1–11. We rst evaluated
metabolites with the top VIP (30, 50, 100) scores for ethyl acetate extracts of the ltrates and methanol/acetone (1:)
extractedmycelia from Xylaria. e initial focus was on metabolites that displayed UV absorption maxima at ~210,
254, 275 or 350 nm (Table2 and Supplementary TablesS1–8). Compounds (100–2000 μg) were puried by reverse
phase semi-preparative HPLC and characterized by NMR (Bruker Advance III 700 MHz NMR with cryoprobe)
(Supplementary Fig.1S). Metabolites were dereplicated against natural product databases including Antibase
(https://www.wiley.com/en-us/AntiBase%3A+The+Natural+Compound+Identifier-p-9783527343591),
Dictionary of Natural Products (http://dnp.chemnetbase.com/faces/chemical/ChemicalSearch.xhtml) and
NORINE (https://bioinfo.li.fr/norine/) using molecular formulas dictated by HRMS data. In addition, a com-
parative analysis was conducted against known fungal metabolites4851. Using this dereplication approach,
the previously reported compounds 111 were identified; (1) griseofulvin, (2) dechlorogriseofulvin, (3)
Figure 4. Supervised multivariate analyses of extracted ltrates from blueberry isolates of X. ellisii endophytes.
e OPLS-DA score plot (a) and S-plot (b) for comparison between X. ellisii endophytes cultured in ML or PDB
media. e OPLS-DA score plots and S-plots compared the X. ellisii endophytes isolates from highbush or wild
blueberries cultured in ML (c,d) or PDB (e,f) medium, respectively.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
7
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
cytochalasin D, (4) zygosporin E, (5) epoxycytochalasin D, (6) hirsutain A, (7) pilformic acid, (8) 2, 3-dihydro-2
,4-dimethylbenzofuran-7-carboxylic acid, (9) cyclic pentapeptide 1, (10) xylarotide A, and (11) cyclic pentapep-
tide 2 (Table2). LC-HRMS, NMR, and spectroscopic data for compounds 111 conrming their structures can
be found in the Supplementary Methods, Figs.2 and 3S, 32–49S and Tables1–9S).
Structure elucidation of ellisiiamides A-C (12–14). Ellisiiamides A–C (1214) were identi-
ed by metabolomic analysis of the extracted ltrates and mycelium with high VIP scores (2.6–11.59; Fig.4,
Supplementary Tables1–8S). ese new cyclic pentapeptides are structurally similar to cyclic pentapeptide 1 (9),
with amino acid dierences at positions 2 (Ala/IsoLeu vs. Val) and 3 (Val vs. IsoLeu) within the peptide scaold
(Fig.5 and Supplementary Fig5S, Tables9–12S).
Ellisiiamide A (12) was isolated as a white powder and afforded a protonated molecular ion at m/z 556
(C30H45N5O5 with 11 double bond equivalents). Examination of the 1H and 13C NMR data revealed the
Model Variables*R2X(cum) R2Y(cum) Q2(cum) Conditions
1a 3856 0.15 0.939 0.615 ML, PDB
1b 100 0.313 0.935 0.737 ML, PDB (including top 100 VIP)
1c 3756 0.299 0.982 0.434 ML, PDB (excluding top 100 VIP)
1d 3556 0.207 0.954 0.394 ML, PDB (excluding top 300 VIP)
2a 3856 0.477 0.998 0.778 ML-H, ML-L
2b 30 0.861 0.984 0.864 ML-H, ML-L (including top 30 VIP)
2c 3826 0.544 1 0.718 ML-H, ML-L (excluding top 30 VIP)
2d 3156 0.121 0.73 -0.187 ML-H, ML-L (excluding top 700 VIP)
3a 3856 0.648 1 0.778 PDB-H, PDB-L
3b 50 0.589 0.995 0.885 PDB-H, PDB-L, (including top 50 VIP)
3c 3806 0.651 1 0.668 PDB-H, PDB-L (excluding top 50 VIP)
3d 2956 0.0851 0.894 -0.175 PDB-H, PDB-L (excluding top 900 VIP)
Table 1. A summary of validation parameters (R2X, R2Y, Q 2) of all calculated OPLS-DA models for extracted
ltrates of X. ellisii endophytes isolates from wild and highbush blueberries cultured in ML and PDB media.
ML-H, endophyte isolates from highbush blueberries cultured in ML medium; ML-L, endophyte isolated from
wild blueberries cultured in ML medium; PDB-H, endophyte isolates from highbush blueberries cultured
in PDB medium; PDB-L, endophyte isolates from wild blueberries cultured in PDB medium. *Number of
metabolomic features included in the OPLS-DA analysis.
#Compound Class Rt Molecular
Formula Measure and
Calculated [M+H]+ppm
error
Known
1Griseofulvin PKS 11.42 C17H18ClO6353.0793 353.0786 1.98
2Dechlorogriseofulvin*PKS 10.01 C17H19O6319.1173 319.1176 0.94
3Cytochalasin D*PKS-NRPS 11.81 C30H38NO6508.2687 508.2694 1.38
4Zygosporin E*PKS-NRPS 13.94 C30H38NO5492.2742 492.2744 0.41
5Epoxycytochalasin D PKS-NRPS 10.87 C30H38N07524.2651 524.2661 1.91
6Hirsutatin A*NRPS 15.85 C34H53N4O10 677.3741 677.3756 2.21
7Piliformic acid PKS 10.84 C11H18O4Na 237.1094 237.1097 1.27
82,3-dihydro,2,4- dimethylbenzofuran
-7- carboxylic acid PKS 11.15 C11H13O3193.0857 193.0859 1.04
9Cyclic pentapeptide 1*NRPS 16.20 C32H50N5O5584.3816 584.3806 1.71
10 Xylarotide A NRPS 15.97 C29H52N5O5550.3973 550.3963 1.82
11 Cyclic pentapeptide 2 NRPS 14.28 C28H50N5O5536.3819 536.3806 2.42
New
12 Ellisiiamide A*NRPS 14.76 C30H46N5O5556.3501 556.3493 1.44
13 Ellisiiamide B*NRPS 15.19 C31H48N5O5570.3656 570.3650 1.05
14 Ellisiiamide C*NRPS 17.04 C33H52N5O5598.3968 598.3963 0.84
15 Ellisiiamide D NRPS 14.47 C27H48N5O5522.3662 522.3650 2.30
16 Ellisiiamide E NRPS 16.89 C30H54N5O5564.4132 564.4119 2.30
17 Ellisiiamide F NRPS 14.11 C31H48N5O6586.3616 586.3599 1.72
18 Ellisiiamide G NRPS 14.00 C32H50N5O6600.3768 600.3756 2.00
19 Ellisiiamide H NRPS 14.89 C33H52N5O6614.3936 614.3912 2.41
Table 2. Identication of known and new secondary metabolites from X. ellisii via LC- UV/HRMS and LC-
HRMS/MS analysis. Select metabolites have been further isolated and characterized by 1D and 2D NMR.
*Structures elucidated by 1D and 2D NMR, HRMS and MS/MS analysis.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
8
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
presence of ve α protons (δ 5.08/55.8, 4.49/46.0, 4.17/55.7, 4.74/46.5, 5.10/58.9 ppm), and three key amide
N-H protons for Ala (δ 8.52), ΙsoLeu (δ 6.94) and Leu (δ 8.49) and the N-Methyl group at (δ 3.04/30.2 ppm).
Examination of the multiplicity edited 1H-13C HSQC, 1H-13C HMBC, and 1H-1H COSY NMR data revealed the
individual amino acid spin systems within the peptide scaold based on α proton correlations to the individual
carbonyl carbon, including amide protons to neighboring amino acid carbonyls, and α, β and γ proton corre-
lations (Fig.5, Supplementary Figs.4–13S and Tables9 and 10S). ese correlations supported the amino acid
sequence of cyclo-(NMePhe-Ala-IsoLeu-Leu-Pro). NOESY through-space correlations of αΗ(Ν−MePhe)/NH
(Ala), NH(Ala)/β Η (Ala), NH(IsoLeu)/αΗ (Ala), NH(Leu)/αΗ (IsoLeu) and H3-NMe (N-MePhe)/β Η (Pro)
further supported the amino acid sequence and relative stereochemistry. Analysis of the LC-MS/MS spectra of
ellisiiamide A revealed key diagnostic b-ion fragments of m/z 459.3 (-Pro), 346.2 (-Leu), 233.1 (-IsoLeu), 162.1
(-Ala) and the presence of two fragmentation pathways as seen in cyclic pentapeptide 1 with ring-opening cleav-
age events at the N-MePhe-Pro and Pro-Leu sites52.
Ellisiiamide B (13) was isolated as a white powder with a protonated molecular ion at m/z 570 aording a
molecular formula of C31 H47N5O5 with 11 double bond equivalents. Examination of 1H and 13C NMR data
revealed presence of ve α protons (δ 5.10/56.0, 3.95/57.6, 4.10/56.8, 4.72/46.6, 5.08/58.7 ppm), key amide N-H
protons for Val (δ 8.18), Val2 (δ 6.98), and Leu (δ 8.43), and the N-Methyl group at (δ 3.03/30.2 ppm). Ellisiiamide
B (13) diers from (9) with Val substituted for IsoLeuc at position # 3 (Fig.4, Supplementary Fig.4S and 14–21S,
Table9 and 11S). Examination of the MS/MS spectra revealed a similar fragmentation pattern as in (9) and
(12), with key diagnostic b-ion fragment ions at m/z 471.3 (-Pro), 360.2 (-Leu), 261.2 (-Val), and 162.1 (-Val).
e cyclo-(NMePhe-Val1-Val2-Leu-Pro) amino acid sequence was conrmed with key HMBC correlations of
αΗ(Ν−MePhe)/CO (N-MePhe), H3-NMe (N- MePhe)/CO (Pro), αΗ(Val1)/CO(Val1), αΗ(Val2)/CO(Val2), αΗ
(Leu)/CO (Leu), and αΗ (Pro)/CO (Pro). Key NOESY correlations of αΗ(Ν−MePhe)/NH(Val1), NH(Val1)/αΗ
(Val1), NH(Val2)/αΗ (Val1), NH(Val2)/αΗ (Val2), NH(Val2)/β Η (Val2), NH(Leu)/ΝΗ (Val2), and H3-NMe
(N-MePhe)/β Η (Pro) further supported the assignments.
Ellisiiamide C (14) was isolated as a white powder with a protonated molecular ion at m/z 598 aording
a molecular formula of C33H51N5O5 with 11 double bond equivalents. Examination of 1H and 13C NMR data
revealed the presence of ve α protons (δ 5.10/56.0, 4.05/55.8, 4.25/55.1, 4.72/46.5, 5.09/58.7 ppm), key amide
Figure 5. Ellisiiamides A–H (12–19), new cyclic nonribosomal peptides from Xylaria ellisii. (a) ellisiiamides
A–C (12–14) isolated and characterized by 1D and 2D NMR, LC-HRHMS and LC-HRMS/MS analysis with
new amino acid substituent highlighted. Corresponding COSY/TOCSY (1H -1H), HMBC (1H-13C) and long-
range through-space NOESY/ROESY correlations are shown. (b) Structures of ellisiiamides D–H (15–19)
based on LC-MS/MS, comparative LC-MS/MS analysis of cyclic pentapeptides 9, 11 and 12–14 (c) Amino
acid scaold of the cyclic pentapeptide family of compounds. Cyclic pentapeptide 1 (9) shown with established
amino acid substituents.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
9
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
N-H protons for IsoLeu1 (δ 8.12), IsoLeu2 (δ 6.94) and Leu (δ 8.45), and the N-Methyl group at (δ 3.04/30.2
ppm). Ellisiiamide C (14) diers from (9) with IsoLeu substituted for Val at position # 2 (Fig.4, Supplementary
Fig.4S and 22–31S, and Tables9 and 11S). Examination of the 1H-1H COSY, multiplicity edited 1H-13C HSQC
and HMBC NMR data revealed the individual spin system for the new IsoLeuc group with correlations of
β3Η(IsoLeu)/αΗ (IsoLeu) and δΗ(IsoLeu)/βΗ (IsoLeu). Correlations of the remaining α protons to the indi-
vidual carbonyl carbons, amide protons to neighboring amino acid carbonyl, and HMBC α, β and γ proton
correlations for Leu, Pro and N-MePhe is consistent with the cyclic peptide scaold (Supplementary Table9S.).
NOESY through-space correlations of αΗ(Ν−MePhe)/NH (IsoLeu1), NH (IsoLeu1)/NΗ (IsoLeu2), NH (Leu)/
αΗ (IsoLeu2)) further supported the amino acid sequence. Analysis of the MS/MS spectra of (14) revealed key
diagnostic b-ion fragments of m/z 501.3 (-Pro), 388.3 (-Leu), 275.2 (-IsoLeu2), and 162.1 (-IsoLeu1) further con-
rming the amino acid sequence of cyclo-(N- MePhe-IsoLeu1-IsoLeu2-Leu-Pro).
e optical rotation for ellisiiamides A–C were measured at [α]21 -86.1 (0.06, MeOH), [α]21 -43.1 (0.04,
MeOH), and [α]20 -47.8 (0.06, MeOH) respectively, and were consistent with 9 at [α]21 -63.4 (0.18, MeOH)
(Supplementary Table9S).
LC-MS/MS analysis and putative identication of new cyclic pentapeptides. Ellisiiamide D–H
(15–19) was identied by metabolomic analysis of the extracted ltrates and mycelium models as unique outliers
with high VIP scores (1.92–6.46). Evaluation of the HRESIMS derived molecular formulas and MS/MS fragmen-
tation patterns of (1519) indicated that the fragmentation sequence and ring-opening events were consistent
with ellisiiamide A–C and cyclic pentapeptide 1. We have therefor assigned putative identication and annotated
structures for ellisiiamides D–H. LC-HRMS/MS characterization data can be found in the Supporting Methods
and Figs.4S and Table 1S and 9S.
Bioactivity activity screening. Compounds 9 and 1214 were screened for biological activity against
three species of microorganisms in accordance with the Clinical Laboratory Standards Institute (CLSI) proto-
cols (National Committee for Clinical Laboratory Standards, 2000, 1997). e microorganism included E. coli
BW25113 ΔbamBΔtolC, Saccharomyces cerevisiae B4741, and Candida albicans ATCC# 90028.
Ellisiiamide A (12) showed modest activity against E. coli with a minimum inhibitory concentration (MIC) of
100 μg/mL. Such activity against E. coli is a rst report for the cyclic pentapeptide scaold. Compound 9 showed
no antifungal activity against S. cerevisiae or C. albicans at 100 μg/mL, which is consistent with reported data52.
Similarly, compounds 1314 showed no activity against any test microorganisms at concentrations between
50–200 μg/mL.
Taxonomy of Xylaria ellisii. Xylaria ellisii. J.B. Tanney, Seifert & Y.M. Ju, sp. nov. MycoBank MB832257
(Fig.6)
= Xylaria corniformis (Fr.: Fr.) Fr. var. obovata M.C. Cooke & J.B. Ellis, Grevillea 6: 92. 1878.
Etymology. Named for the prolic mycologist Job Bicknell Ellis who, with Mordecai Cubitt Cooke, described
Xylaria corniformis var. obovata Sacc., a synonym of X. ellisii.
Typus. Canada: New Brunswick, Alma, Fundy National Park, East Branch Trail, 45.6433-65.1156, stromata on
partially buried, mostly decorticated Acer saccharum branch, 28 Sep 2014, J.B. Tanney NB-623 (holotype DAOM
628556). Ex-type culture DAOMC 252031.
Colonies 32–38 mm diam aer 14 d in the dark at 20 °C on MEA; white, velvety, appressed, sometimes sec-
tored; margin diuse, hyaline; surface and reverse white. Exudates and soluble pigments absent. Mycelium con-
sisting of hyaline, smooth, septate, branched, hyphae 1.5–3 µm diam.
Conidiophores on MEA macronematous, arising vertically from mycelium, hyaline to pale brown, smooth,
cylindrical, thin-walled, dichotomously branched several times, septate, 30–60 × 3–4 µm, or occurring in
synnemata, grey to olive brown (4D2–4E3). Synnemata cylindrical to clavate, occurring singly, gregariously,
or in clusters joined at base, up to 10 mm high by 1–3 mm diam, surface appearing powdery due to conidia.
Conidiogenous cells intercalary and terminal, cylindrical, straight or undulating to geniculate, 7–16(20) ×
3–4 µm, hyaline to pale brown, producing one or more conidia holoblastically from lateral or apical regions,
crater-shaped protruding secession scars 1–1.5 × 1–1.5 µm. Conidia pyriform to obovoid, subhyaline to pale
brown, (5)5.5–7(7.5) × (2.5)3(3.5) µm, attened basal scar indicating former site of attachment to conid-
iogenous cell.
Stromata upright, solitary, unbranched or occasionally branched once, cylindrical to spathulate or clavate, api-
ces broadly rounded, divided into fertile head and sterile stipe, (2)2.54(5) × 0.8–1.2 cm including stipes (0.4–
1.5 cm high); surface even to irregularly attened or wrinkled, frequently cracked into a network of light brown
to brownish orange (6D4–6D5) angular plates above black basal layer; stromatal interior white; stipes brownish
orange to light brown (6D4–6D5) frequently with black longitudinal cracks extending from fertile head; arising
from brown (7D7) to black pannose bases, basal mycelia oen appearing iridescent. Perithecia immersed, subglo-
bose to globose, 0.3–1 mm diam, lining the perimeter of the stromata. Ostioles conspicuous, papillate, 100–300 µm
diam. Asci 95–130 × 6–7 µm, partis sporiferae 50–80 µm, eight-spored, cylindrical, with ascospores arranged
uniseriately; apical apparatus inverted hat-shaped, amyloid, 1.5–2 µm long. Ascospores (8)9–9.5(10) ×
(4.5)5–5.5(6) µm, dark brown, smooth, unicellular, ellipsoid-inequilateral, narrowly or broadly rounded ends,
1–2 guttules frequently observed, inconspicuous long, straight germ slits which are more or less the spore length,
occurring on convex side; small ephemeral cellular appendage 1.5–2 × 1.5 µm, visible on less pigmented imma-
ture ascospores and disappearing as spores reach maturity.
Cardinal temperatures: Range 5–30 °C, optimum 20 °C, minimum slightly <5 °C, maximum slightly >30 °C.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
10
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
Host range: Stromata on decaying hardwood including Acer, Betula, Fagus, and other hardwood trees. Foliar
endophyte of Abies balsamea, Picea glauca, P. mariana, P. rubens, and Pinus strobus. Foliar and stem endophyte of
Vaccinium angustifolium and V. corymbosum. Closely related ITS sequences in GenBank suggest a broad endo-
phytic and endolichenic host range.
Figure 6. Xylaria ellisii morphology. (A,B) Stromata on partially buried, decaying Acer saccharum branches,
arrow pointing to longitudinal section of stroma with perithecia lining outer surface. (C) Base of stroma
showing ostioles and reticulations. (D) Ostioles on stroma surface. (E) Eight-week-old colony on oatmeal agar.
(F) Longitudinal section of perithecium. (G) Asci and paraphyses. (H) Conidiogenous cells. (I) Conidia. (J,K)
Asci with amyloid, inverted hat-shaped apical apparatuses. (L) Ascospores, arrow denoting germ slit. Scale bars:
(F,G) = 100 µm, (H,JL) = 10 µm, I = 5 µm.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
11
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
Distribution: Eastern Canada and U.S.A.
Additional specimens and cultures examined: DAOM 696463, DAOM 696464, DAOM 696466, DAOM 696480,
DAOM 696488, DAOM 696489, DAOM 696492, DAOM 696493, DAOM 696503, NB-699, NB-701, NB-702,
NB-703, NB-708, NB-721, NB-722, NB-723, NB-727, NB-746, CH-12, CH-15, CH-16, CH-37, CH-38, CH-4,
CH-5, DT-181, DT-6, NB-236-1F, NB-236-2F, NB-236-2I, NB-285-10A, NB-285-10D, NB-285-1A, NB-285-3A,
NB-285-6B, NB-285-7A, NB-285-7B, NB-285-7C, NB-285-7D, NB-365-4E, NB-365-71G, NB-365-
8A, NB-366-1F, NB-366-2E, NB-366-3L, NB-366-4C, NB-382-1C, NB-382-3A, NB-382-3C, NB-382-3D, NB-382-
4B, NB-391-1E, NB-391-2C, NB-391-4C, NB-406-2A, NB-406-2B, NB-406-5A, NB-421-1B, NB-437-5E,
NB-464-10A, NB-487-5B, NB-487-5C, NB-487-6A, NB-487-6H, NB-488-6L, NB-505-4D, NB-746, RS9-10E,
RS9-12C, T1-3B-2, T1-4B-1, T2-4A-2, T3-2A-2, T3-2B-1, T3-3A-3, T4-3A-1, T5-1A-1, T5-3B-1-1, T6-4B-1,
T6-5A-1-2.
Notes: Xylaria ellisii is equivalent to X. corniformis var. obovata, e.g.: Ju et al. (2016) recorded blackish-brown
ascospores, 8–10.5 × 4.5–5.5(6) µm from the X. corniformis var. obovata holotype53. e X. corniformis species
concept is unresolved and consequently the name has been misapplied to various species within the X. corni-
formis and X. polymorpha aggregates45. e Xylaria corniformis aggregate is a polyphyletic morphotaxonomic
concept comprising species characterized by stromata with a wrinkled surface and a thin outer layer that gradu-
ally cracks into ne scales with maturation, including X. bipindensis, X. cuneata, X. curta, X. divisa, X. feejeensis,
X. humosa, X. luteostromata, X. maumeei, X. montagnei, X. plebeja, and X. rhytidophloea45,5355. Rogers (1983)
noted the taxonomic confusion surrounding X. corniformis and its misapplication to X. bulbosa, X. castorea, X.
curta, and other morphologically similar species, and recommended that Xylaria taxonomy would be best served
if the name X. corniformis were no longer used45. Xylaria corniformis s.s. is possibly a rare species known only
from Swedish and Polish collections and is characterized by delicate, horn-like stromata with attenuated or sterile
apices versus the robust stromata of X. ellisii, which also have darker coloured ascospores44,55,56. Ju et al. (2009)
concluded that X. corniformis var. obovata was an equivalent of X. corniformis sensu Læssøe (1987)55. Læssøe
(1987) noted that X. corniformis var. obovata was probably the most frequently encountered member of the X.
corniformis complex in northern temperate regions44. Ju et al. (2009) considered X. corniformis and X. corniformis
var. obovata as distinct species but refrained from making a formal taxonomic decision pending additional evi-
dence55. Xylaria ellisii is common on decaying fallen Acer saccharum branches in New Brunswick during late
summer and autumn and is a frequently isolated endophyte of Picea, Pinus strobus, and Vaccinium angustifolium
in Eastern Canada38. Conspecic ITS sequences in GenBank suggest that X. ellisii is capable of endophytically
infecting a wide range of hosts.
Discussion
Xylaria ellisii was the most commonly isolated Xylariaceae endophyte from Picea and Pinus in Eastern Canada57.
Stromata of X. ellisii were commonly found on decaying Acer saccharum branches or stems in the same forest
stands where it was isolated as a Picea endophyte. Endophyte ITS sequences in GenBank corresponding to X.
ellisii originate from an exceptional diversity of hosts, including Tsuga canadensis, bryophytes (e.g.: Hypnum sp.),
liverworts (e.g.: Metzgeria furcata, Trichocolea tomentella), and lichens (e.g.: Flavoparmelia caperata, Sticta beau-
voisii, Xanthoparmelia conspersa) (Fig.3). In New Brunswick, corresponding X. ellisii stromata were commonly
found in late summer and early fall only on decaying Acer saccharum wood; however, the stromatal host range
is likely broad. For example, Læssøe (1987) examined European specimens of Xylaria corniformis (probably X.
ellisii) from Carpinus and Fagus44 and Rogers (1983) examined North American collections from Betula, Fagus,
Malus, and Tsuga45.
Xylaria ellisii is a common Picea and Pinus endophyte even in conifer-dominated stands lacking Acer saccha-
rum or any other hardwood hosts possibly suitable for the production of stromata. is indicates that the fungus
is capable of persisting in the environment in the prolonged absence of a suitable primary host. e method of
transmission between foliage is currently unknown. It is conceivable that the dry, powdery masses of conidia
produced from conidiomata in vitro are also produced on dead foliage and capable of infecting new foliage by
means of air currents or insect vectors58,59. Abscised foliage infected with X. ellisii is probably capable of sapro-
trophically colonizing hosts by means of direct contact (viaphytism), as demonstrated in other Xylaria species25.
e known range of hosts that X. ellisii can endophytically infect includes lichens and various understory and
overstory plant species with dierent successional statuses, allowing for its persistence across forest succession
pathways and disturbances (e.g.: as an endophyte of the re-adapted seral species Vaccinium angustifolium). A
proposed endophytic-saprotrophic life history is described and illustrated for Xylaria ellisii (as Xylaria sp.) by
Tanney et al.60.
e production of the potently antifungal compound griseofulvin by X. ellisii, an apparently ubiquitous endo-
phyte with a broad host range, is signicant. Griseofulvin is toxic to a wide variety of plant pathogens6164 and
is systemically translocated within plants65, suggesting that X. ellisii endophyte infections could increase host
resistance to plant pathogens. For example, Park et al. (2005) described griseofulvinproduction in an unidentied
Xylaria endophyte of Abies holophylla and showed its ability to control the development of plant diseases such
as barley powdery mildew (Blumeria graminis f. sp. hordei), rice sheath blight (Corticium sasaki), wheat leaf rust
(Puccinia recondita), and rice blast (Magnaporthe grisea)64. Griseofulvin and related compounds are reported
from Xylaria endophytes of Asimina triloba, Chrysobalanus icaco, and Garcinia hombroniana6668. Richardson
et al. (2014) reported the production of the antifungal compound griseofulvin by Xylaria ellisii (as Xylaria sp.)
isolated as a foliar endophyte of Pinus strobus and Vaccinium angustifolium38. ese isolates produced griseof-
ulvin and its de-halogenated analogue (Fig.1), along with piliformic acid38. Subsequent investigations of white
pine seedlings infected with this Xylaria species found griseofulvin at biologically eective concentrations in the
needles69.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
12
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
Nonribosomal peptides (NRPS) are of great interest as they represent a unique class of natural products with
diverse therapeutic applications such as antimicrobial agents (caspofungin, penicillin, vancomycin), antican-
cer compounds (bleomycin, daptomycin), immunosuppressants (cyclosporine, rapamycin) and as insect toxins
(beauvercin, enniatin)7073. is complex structural diversity of linear, cyclic, and cyclic branched architectures is
synthesized through a modular enzymatic assembly line process70,73. In principle, this enzyme complex is capable
of incorporating >500 proteinogenic and nonproteinogenic building blocks, including polyketide and terpene
hybrid moieties.
In this study, we have applied a LC-MS metabolomic guided discovery approach to prole the chemical space
of a novel endophytic species described here as Xylaria ellisii. Our collections of isolates have identical ITS DNA
sequences yet dier in their LC-MS metabolite proles and bioactivity. OPLS-DA and S-plot analysis identi-
ed features separated by a statistical toll, Variable Importance in Projection (VIP) scores. VIP scores from the
extracted ltrates and mycelium extracts were calculated and extracts dierentiated by this method were targeted
for compound isolation and structural characterization. is approach resulted in the discovery of three new
cyclic pentapeptides given the trivial names ellisiiamides A–C (12–14) and the putative identication and anno-
tation of ellisiiamides D–H by LC-HRMS and LC-HRMS/MS analysis. Additionally, 11 known compounds are
reported to be produced by these strains. Ellisiiamide A (12) was active against Gram-negative bacteria and is
a rst report for this scaold. ese ndings are of interest as the isolates were also reported from eastern white
pine needles in a pine-blueberry forest ecotype. Endophytes from wild Vaccinium species may be an interesting
source of novel bioactive compounds. is information provides a better understanding of the chemical ecology
of plant-fungi microbiomes. In the long term, opportunities may present to employ this information for inte-
grated pest management crop protection strategies.
Methods
Sampling, isolation, and culturing. Plant material, including leaves and stems from highbush and wild
blueberries, were collected from three dierent locations within the Acadian forest region of Nova Scotia, Canada.
Highbush blueberry endophyte isolates were obtained from a commercial eld in Rawdon, Nova Scotia and wild
blueberry endophytes isolates were collected from commercial elds in Mount om, Debert, and Portapique,
Nova Scotia. Specimens were collected in labelled bags and stored at -20 °C prior to fungal isolation. Plant tissues
were rst washed with sterile deionized water to remove any loose debris and surface contaminants, followed by
a chemical surface-sterilization process using sodium hypochlorite bleach (6%) and ethanol (70%). Small seg-
ments were then cut and/or incised and placed in Petri plates containing 2% malt extract agar (MEA; 20 g Bacto
malt extract, Difco Laboratories, Sparks, USA; 15 g agar, EMD Chemicals Inc., Gibbstown, USA; 1 L deionized
H2O). Inoculated plates were incubated at 25 °C for 4–8 weeks, depending on the presence of lamentous hyphae.
Endophytic fungi that grew from cut ends were then transferred to potato dextrose agar (PDA, Sigma-Aldrich,
Canada) plates and incubated at 25 °C.
Field specimens of stromata were collected and stored in paper bags. Single-ascospore isolates were made
by axing with petroleum jelly a small (ca. 5 mm2) piece of stroma containing mature perithecia to the lid of
a Petri dish containing water agar (WA; 15 g agar, EMD Chemicals Inc., Gibbstown, USA; 1 L deionized H2O).
Germination of ejected ascospores on the agar surface was conrmed by stereo microscope (Olympus SZX12,
Olympus, Tokyo, Japan) and germinating ascospores were transferred to individual Petri plates containing
2% MEA and incubated at 20 °C. Dried specimens were accessioned in the Canadian National Mycological
Herbarium (Ottawa, Ont.; DAOM). Living cultures were deposited in the Canadian Collection of Fungal Cultures
(Ottawa, Ont.; DAOMC). Additional specimens used for morphological comparison and phylogenetic analyses
were also obtained from DAOM, DAOMC, and the personal culture collection of J.B. Tanney.
Xylaria strains from highbush blueberry and wild blueberry were cultured in PDB (24 g/L potato dextrose
broth) and ML (30 g/L malt) fermentation media. Each strain was grown in 1 L Roux bottles containing 200 mL
of media and grown statically for 4–6 weeks at 25 °C. e culture broth was then separated from the mycelium by
vacuum ltration using a Whatman #4 lter paper. e ltrate was extracted with equal volumes of ethyl acetate,
while the mycelium was rst lyophilized for 24 h and then extracted with equivalent volumes of methanol and
acetone (1:1). Organic fractions were then dried under reduced pressure by rotary vacuum. Extracts were then
re-suspended in 600 μL of HPLC grade acetonitrile with minimal amounts of DMSO added for solubility. e
ltrates were then centrifuged at 13,000 rpm for 15 min and Acro-disk (13 mm, 0.45 μm GHP) ltered prior to
LC-MS analysis.
Morphological study. Sections of stromata were cut by hand using a safety razor blade or with a freezing
microtome (ca. 15–30 µm thick) and mounted in either water, 5% KOH, 85% lactic acid, or Lugols solution with
or without 5% KOH pretreatment to test amyloid reactions74. Stromata and colony colours were described using
alphanumeric codes75. Observations of the asexual morph were made from living cultures grown on oatmeal agar
(OA)76. Microscopic measurements were taken from living material mounted in deionized water and are pre-
sented as ranges calculated from the mean ± standard deviation of each measured value with outliers in brackets.
Observations were made using an Olympus BX50F4 light microscope and an Olympus SZX12 stereo microscope
(Olympus, Tokyo, Japan). Images were captured with an InnityX-32 camera (Lumenera Corp., Ottawa, Canada)
using Innity Analyze v. 6.5.2 (Lumenera Corp.) soware. Photographic plates were assembled using Adobe
Photoshop CC 2017.1.1 (Adobe Systems, San Jose, California, USA). Cardinal temperatures were assessed for the
type strain (DAOMC 252031) by incubating single-point inoculated Petri dishes containing MEA at 5 °C inter-
vals from 5–40 °C. Each treatment was conducted in triplicate and colony diameters were measured two weeks
aer inoculation.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
13
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
DNA extraction, sequencing, and phylogenetic analyses. DNA was extracted from cultures and
stromata using the Ultraclean Microbial DNA Isolation Kit (Mo Bio, Carlsbad, CA) or NucleoSpin Plant II Kit
(Macherey-Nagel, Düren, Germany). Stromatal tissue from fresh collections and herbarium specimens under-
went an initial grinding stage in liquid nitrogen using an Axygen polypropylene pestle (PES-15-B-SI, Union City,
CA, USA).
Loci chosen for sequencing included the internal transcribed spacer rDNA region (ITS), β-tubulin (BenA),
translation elongation factor 1-alpha (EF1-α), the second largest subunit of RNA polymerase II (RPB2), 18 s nuc
rDNA (SSU), and 28 S nuc rDNA (LSU). Primer pairs used for PCR amplication and sequencing included: ITS1
and ITS477 or ITS4A and ITS578 for ITS; Bt2a and Bt2b for BenA79; RPB2-5f2 and RPB2-7CR80 for RPB2; and
EF1-728F and EF1-986R81 for EF1-α. LSU was amplied using LR0R and LR5 and sequenced using the primers
LR0R, LR3, LR3R, and LR582. SSU was amplied using the primers NS1 and NS4, and sequenced using the prim-
ers NS1, NS2, NS3, and NS477. PCR and sequencing were performed as described by Tanney and Seifert (2017)57.
To improve ITS amplication in herbarium specimens, 0.5 μm of 20 mg/ml bovine serum albumin (BSA) was
added per reaction.
For all analyses, sequences were aligned using MAFFT v783 and visually inspected and manually aligned
when necessary in Geneious R8 v8.1.5 (Biomatters, Auckland, New Zealand). e most suitable sequence evolu-
tion model was determined based on the optimal Akaike information criterion scores in MrModeltest v2.2.684.
Consensus trees were visualized in FigTree 1.4.2 (available at http://tree.bio.ed.ac.uk/soware/gtree/) and
exported as SVG vector graphics for assembly in Adobe Illustrator v10 (Adobe Systems, San Jose, CA, USA).
Three separate phylogenetic analyses were performed. The first phylogeny included ITS sequences of
diverse representative endophytes isolated from highbush and wild blueberry leaves and stem. e ex-type
of Mucor ellipsoideus (ATCC MYA-4767; NR_111683) was selected as outgroup because of its basal position
(Mucoromycotina). Maximum likelihood (ML) analysis was performed using RAxML v8.2.4 in PAUP v4.0b10
starting from a random starting tree with 1000 bootstrap replicates85,86.
e second phylogenetic analysis included RPB2 sequences from related Xylaria species. e resulting align-
ment was 1058 bp long and consisted of 47 taxa, including the outgroup Barrmaelia rhamnicola (CBS 142772).
Bayesian analysis was performed using MrBayes v3.2.687. Three independent Markov Chain Monte Carlo
(MCMC) samplings were performed with 12 chains (11 heated and one cold) with sampling every 500 genera-
tions until the standard deviation of split frequencies was <0.01. e rst 25% of trees were discarded as burn-in
and the remaining trees were kept and combined into one consensus tree with 50% majority rule consensus.
Convergence was assessed from the three independent runs using Tracer v1.688. e third phylogenetic analysis
included ITS sequences from related endophytic Xylaria isolates. e alignment was 593 bp long and included
sequences from 107 isolates or samples. e resulting phylogenetic analysis was performed in the same manner
as described above, with Nemania serpens (GU292820) as the outgroup.
All novel sequences used in this study were accessioned in GenBank (Supplementary Table13S) and taxo-
nomic novelties and associated metadata were deposited in MycoBank (www.MycoBank.org).
LC-UV/HRMS and LC-UV/HRMS/MS screening. Extracts of endophytic cultures were screened using a
Dionex Ultimate 3000 HPLC-UV system coupled to a Bruker maXis 4 G ultra-high-resolution-qTOF mass spec-
trometer operated in positive electrospray ionization (ESI) with calibrations done using sodium formate ion clus-
ters. LC-MS data were collected using a scan range of 150–1100 m/z, with the nebulizer gas (nitrogen) at 3 bar, dry
gas ow at 8 L/min, dry gas temperature at 240 °C, and capillary voltage at 4500 V. Chromatographic separations
were performed using a standardized HPLC-UV method with a Supelco Ascentis Express C18 reverse-phase
core-shell column (150 × 4.6 mm, 2.7 μm, Sigma Aldrich, USA) operating at 750 μL/min and at 40 °C. UV/vis
data were acquired from 190–600 nm and monitored at four wavelengths (210, 254, 275 and 350 nm). Mobile
phase composition was linear with a gradient of 5% organic from 0 to 1 min, 5–95% from 1 to 24 min, 95–100%
from 24 to 25 min, and 100% from 25 to 31 min. Solvent A was H2O + 0.1% formic acid and solvent B was ace-
tonitrile with 0.1% formic acid (v/v). HR-MS/MS analysis was performed on a ermo Q-Exactive Orbitrap mass
spectrometer operated in positive electrospray ionization (ESI+) and coupled to an Agilent 1290 HPLC system.
Data processing and multivariate statistical analysis. Data processing and analyses were modied
from a previously published protocol (Fei et al., 2014). Post-acquisition internal calibration using sodium for-
mate clusters in both ESI+ and ESI- were performed with Bruker’s Data Analysis 4.0 SP4. LC-MS data les were
converted to.mzXML format using Bruker CompassXport. Metabolic features were extracted and aligned using
open source XCMS with centWave algorithm89; adducts, isotopic ions, and in-source fragments were identied
using CAMERA90,91. To acquire the nal metabolite feature list, isotopic ions and features with integrated peak
area under 10,000 were removed. For myceliummetabolome, metabolite features that eluted aer 25 min were
eliminated.
Both extracted ltrates and mycelium were analyzed using supervised multivariate OPLS-DA aer pareto
scaling by SIMCA-P+ 12.0.1 (Umetrics, Kinnelon, NJ). e statistical parameters R2X(cum), R2Y(cum), and
Q2(cum) of OPLS-DA were used to assess the tness of the model. R2X (R2Y) indicated the fraction in which
metabolite features (X) and group (Y) matrix was were explained by the model. A prediction statistic (Q2) above
0.4 was indicative of a statistically robust model, i.e. true dierences between the comparing groups, and Q2
between 0.7–1.0 indicated the model was statistically signicant46. Both R2 and Q2 followed an upward trend from
0 to 1. For an over-t model, R2 approached 1, and Q2 fell toward 092. Signicant features between classes were
identied based on OPLS-DA S-plot and their Variable Importance in Projection (VIP) score. To ensure the iden-
tied metabolites are the sole important markers, the two OPLS-DA analyses were conducted in parallel by only
including the signicant features or by removing the signicant features from the raw data92. A useful metabolite
subset was produced if the rst model was successful and the later model failed.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
14
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
Metabolite Isolation and characterization. NMR experiments for 1D and 2D measurements were per-
formed on a Bruker Advance III 700 MHz NMR spectrometer equipped with a 5 mm QNP cryoprobe, operating
at 700.17 MHz for 1H NMR and 176.08 MHz for 13C NMR or a Bruker Advance III HD 850 MhZ NMR spectrom-
eter equipped with a 5 mm TXI probe operating at 850.21 MHz for 1 H NMR and 213.81 MHz for 13 C NMR,
with chemical shis referenced to the residual solvent signal93. Nitrogen dried compounds were re-suspended
in 200 μL of deuterated solvent (C6D6, CD3OD, or DMSO-d6) and transferred to 3 mm NMR tubes (Wilmad
335-pp-7) for NMR measurements. NMR data processing was done using MNOVA NMR soware ver. 10.0.1
by Mestrelab Research. Optical rotation measurements were done using an Autopol IV Polarimeter (Rudolph
Research Analytical).
Purication of metabolomic targeted metabolites was performed on a semi-preparative HPLC system consist-
ing of an Agilent 1100 series HPLC with a G1311A Quaternary Pump, a G1379A Degasser, a G1367A Wellplate
Autosampler, a G1316A Column ermostat, a G1315B Diode Array Detector (DAD), and a G1364C Automatic
Fraction Collector controlled by Agilent ChemStation soware (Rev. B.03.02-SR2). Metabolites were isolated
using a Phenomenex Synergi-Max reverse-phase C-12 column (250 × 10 mm, 4 μm) (Torrence, CA, USA) oper-
ating at 5 mL/min and 40 °C. Mobile phase composition was a linear gradient of 5% organic from 0 to 3 min,
5–30% from 3 to 16 min, 30% from 16 to 20 min, and 30–85% from 20–37 min with fractions collected every 20 s.
Known isolated compounds (mg/L): dechlorogriseofulvin (2) eluted at 27.1 min (4 mg); griseofulvin (1) eluted at
29.1 min (2.8 mg); cytochalasin D (3) eluted at 30.2 min (2.5 mg); zygosporin E (4) eluted at 32.5 min (2 mg); hir-
sutatin A (6) eluted at 33.9 min (2 mg); and cyclic pentapeptide #1 (9) eluted at 34.9 min (4 mg) (Supplementary
Figs.3S and 32–49S)
Newly-isolated compounds (mg/L): ellisiiamide G (18) eluted at 31.6 min (0.3 mg); ellisiiamide A (12) eluted
at 32.8 min (2.0 mg); ellisiiamide B (13) eluted at 33.4 min (1.3 mg); and ellisiiamide C (14) eluted at 35.6 min
(2.3 mg). Compound fractions, from multiple HPLC runs, were pooled together and dried under N2 gas in
pre-weighed vials prior to NMR and optical rotation measurements (Supplementary Figs.5–31S, Tables10–12S).
Ellisiiamide A (12) C30H45N5O5; white powder; [α]21 86.1 (0.18, MeOH); For 1H and 13C NMR (DMSOd6)
spectroscopic data see Supporting Table9S: HRESIMS (m/z) 556.3501 [M+H]+ (calcd for C30H46N5O5, 556.3493).
Ellisiiamide B (13) C31H47N5O5; white powder; [α]21 43.1 (0.04, MeOH); For 1H and 13C NMR (DMSOd6)
spectroscopic data see Supporting Table10S: HRESIMS (m/z) 570.3656 [M+H]+ (calcd for C31H48N5O5,
570.3650).
Ellisiiamide C (14) C33H51N5O5; white powder; [α]21 47.8 (0.06, MeOH); For 1H and 13C NMR (DMSOd6)
spectroscopic data see Supporting Table11S: HRESIMS (m/z) 598.3968 [M+H]+ (calcd for C33H52N5O5,
598.3963).
Biological activity screening. Compounds were tested for their minimum inhibitory concentration
(MIC) according to the Clinical Laboratory Standards Institute (CLSI) protocols M7-A5 and M27-A (National
Committee for Clinical Laboratory Standards, 2000, 1997). Stock working solutions were made to 5, 10, and
20 mg/mL and tested at a maximum concentration of 200 μg/mL in 96-well liquid culture (National Committee
for Clinical Laboratory Standards, 1997, 2003) as previously described37. Preliminary evaluation of biologi-
cal activity was against E. coli BW25113 ΔbamBΔtolC, a membrane and eux pump compromised strain,
Staphylococcus aureus ATCC# 29213, Bacillus subtilis 1A1, Micrococcus luteus, Saccharomyces cerevisiae B4741,
and Candida albicans ATCC# 90028. A cut-o of <25% growth was used for inhibition, with the trend across
dilutions also considered37.
Received: 2 October 2019; Accepted: 28 January 2020;
Published: xx xx xxxx
References
1. Cutler, M. An Account of Some of the Vegetable Productions, Naturally Growing in this Part of America: Botanically Arranged. (Lloyd
Library, 1903).
2. Turner, N. J. & Aderas, P. v. J. A. S. B. P. Sustained by First Nations: European newcomers use of Indigenous plant foods in
tempterate North America. 81 (2012).
3. Miller, J. D. In Endophytes of Forest Trees: Biology and Applications (eds. Anna Maria Pirttilä & A. Carolin Fran) 237–249 (Springer
Netherlands, 2011).
4. ir, D. et al. Avian assemblages dier between old-growth and mature white pine forests of Ontario, Canada: a role for supercanopy
trees? 7 (2012).
5. Petrini, O., Sieber, T. N., Toti, L. & Viret, O. Ecology, metabolite production, and substrate utilization in endophytic fungi. 1,
185–196, https://doi.org/10.1002/nt.2620010306 (1993).
6. Li, Z.-J., Shen, X.-Y. & Hou, C.-L. Fungal endophytes of South China blueberry (Vaccinium dunalianum var. urophyllum). 63,
482–487, https://doi.org/10.1111/lam.12673 (2016).
7. Carris, L. M. Chalara Vaccinii Sp. Nov., A Vaccinium Endophyte. Mycologia 80, 875–879, https://doi.org/10.1080/00275514.1988.1
2025742 (1988).
8. Carris, L. M. A New Species of Dwayalomella from Vaccinium Corymbosum. Mycologia 81, 638–642, https://doi.org/10.1080/0027
5514.1989.12025797 (1989).
9. Martínez-Álvarez, P., Fernández-González, . A., Sanz-os, A. V., Pando, V. & Diez, J. J. Two fungal endophytes reduce the severity
of pitch caner disease in Pinus radiata seedlings. Biological Control 94, 1–10, https://doi.org/10.1016/j.biocontrol.2015.11.011
(2016).
10. Prihatini, I., Glen, M., Wardlaw, T. J. & Mohammed, C. L. Diversity and identication of fungi associated with needles of Pinus
radiata in Tasmania. Southern Forests: a Journal of Forest Science 78, 19–34, https://doi.org/10.2989/20702620.2015.1092345 (2016).
11. aj, T. N. J. C. Jo. B. Genera coelomycetum. XXI. Strasseria and two new anamorph-genera. Apostrasseria and Nothostrasseria. 61,
1–30 (1983).
12. Carris, L. M. Cranberry blac rot fungi: Allantophomopsis cytisporea and Allantophomopsis lycopodina. Canadian Journal of
Botany 68, 2283–2291, https://doi.org/10.1139/b90-291 (1990).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
15
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
13. ouol, O., olaří, M., olářová, Z. & Baldrian, P. J. F. D. V. Diversity of foliar endophytes in wind-fallen Picea abies trees. 69–77 (2011).
14. Prihatini, I., Glen, M., Wardlaw, T. J., atowsy, D. A. & Mohammed, C. L. J. N. Z. J. o. F. S. Needle fungi in young Tasmanian Pinus
radiata plantations in relation to elevation and rainfall. 45, 25, https://doi.org/10.1186/s40490-015-0055-6 (2015).
15. Tanney, J. B. & Seifert, . A. Phacidiaceae endophytes of Picea rubens in Eastern Canada. Botany 96, 555–588, https://doi.
org/10.1139/cjb-2018-0061 (2018).
16. Soolsi, S., Piché, Y., Chauvet, É. & Bérubé, J. A. A fungal endophyte of blac spruce (Picea mariana) needles is also an aquatic
hyphomycete. 15, 1955–1962, https://doi.org/10.1111/j.1365-294X.2006.02909.x (2006).
17. Soolsi, S., Piché, Y., Laitung, B. & Bérubé, J. A. Streams in Quebec boreal and mixed-wood forests reveal a new aquatic
hyphomycete species, Dwayaangam colodena sp. nov. Mycologia 98, 628–636, https://doi.org/10.1080/15572536.2006.11832666
(2006).
18. évay, Á. & Gönczöl, J. ainborne hyphomycete conidia from evergreen trees. Nova Hedwigia 91, 151–163, https://doi.
org/10.1127/0029-5035/2010/0091-0151 (2010).
19. Sumarah, M. W., Puniani, E., Sørensen, D., Blacwell, B. A. & Miller, J. D. Secondary metabolites from anti-insect extracts of
endophytic fungi isolated from Picea rubens. Phytochemistry 71, 760–765, https://doi.org/10.1016/j.phytochem.2010.01.015 (2010).
20. Tanney, J. B., Douglas, B. & Seifert, . A. Sexual and asexual states of some endophytic Phialocephala species of Picea. Mycologia
108, 255–280, https://doi.org/10.3852/15-136 (2016).
21. Dreyfuss, M. J. B. H. Further investigations on the occurrence and distribution of endophytic fungi in tropical plants. 94, 33–40
(1984).
22. Petrini, L. J. S. Xylariaceous fungi as endophytes. 38, 216–234 (1985).
23. U’en, J. M. et al. Contributions of North American endophytes to the phylogeny, ecology, and taxonomy of Xylariaceae
(Sordariomycetes, Ascomycota). Molecular Phylogenetics and Evolution 98, 210–232, https://doi.org/10.1016/j.ympev.2016.02.010
(2016).
24. Oane, I. et al. Study of endophytic Xylariaceae in ailand: diversity and taxonomy inferred from rDNA sequence analyses with
saprobes forming fruit bodies in the eld. Mycoscience 49, 359–372, https://doi.org/10.1007/S10267-008-0440-6 (2008).
25. omas, D. C., Vandegri, ., Ludden, A., Carroll, G. C. & oy, B. A. Spatial Ecology of the Fungal Genus Xylaria in a Tropical
Cloud Forest. 48, 381–393, https://doi.org/10.1111/btp.12273 (2016).
26. Daranagama, D. A. et al. Towards a natural classification and bacbone tree for Graphostromataceae, Hypoxylaceae,
Lopadostomataceae and Xylariaceae. 88, 1–165 (2018).
27. Petrini, O. & Carroll, G. J. C. J. o. B. Endophytic fungi in foliage of some Cupressaceae in Oregon. 59, 629–636 (1981).
28. Sieber, T. N. J. M. . Endophytic fungi in twigs of healthy and diseased Norway spruce and white r. 92, 322–326 (1989).
29. Laessøe, T. & Spooner, B. J. . B. osellinia & Astrocystis (Xylariaceae): new species and generic concepts. 1–70 (1993).
30. Barlund, P. & owalsi, T. J. C. J. o. B. Endophytic fungi in branches of Norway spruce with particular reference to Tryblidiopsis
pinastri. 74, 673–678 (1996).
31. Hata, ., Futai, . & Tsuda, M. J. C. J. o. B. Seasonal and needle age-dependent changes of the endophytic mycobiota in Pinus
thunbergii and Pinus densiora needles. 76, 245–250 (1998).
32. Helaly, S. E., ongbai, B. & Stadler, M. Diversity of biologically active secondary metabolites from endophytic and saprotrophic
fungi of the ascomycete order Xylariales. Natural Product eports 35, 992–1014, https://doi.org/10.1039/C8NP00010G (2018).
33. Whalley, A. J. S. & Edwards, . L. Secondary metabolites and systematic arrangement within the Xylariaceae. Canadian Journal of
Botany 73, 802–810, https://doi.org/10.1139/b95-325 (1995).
34. Stadler, M. J. C. . E. A. M. Importance of secondary metabolites in the Xylariaceae as parameters for assessment of their taxonomy,
phylogeny, and functional biodiversity. 1, 75–133 (2011).
35. Schulz, B., Boyle, C., Draeger, S., ömmert, A.-. & rohn, . J. M. . Endophytic fungi: a source of novel biologically active
secondary metabolites. 106, 996–1004 (2002).
36. usari, S., Hertwec, C. & Spiteller, M. J. C. & biology. Chemical ecology of endophytic fungi: origins of secondary metabolites. 19,
792–798 (2012).
37. Ibrahim, A. et al. Epoxynemanione A, nemanifuranones A–F, and nemanilactones A–C, from Nemania serpens, an endophytic
fungus isolated from iesling grapevines. Phytochemistry 140, 16–26, https://doi.org/10.1016/j.phytochem.2017.04.009 (2017).
38. ichardson, S. N. et al. Griseofulvin-producing Xylaria endophytes of Pinus strobus and Vaccinium angustifolium: evidence for a
conifer-understory species endophyte ecology. Fungal Ecology 11, 107–113, https://doi.org/10.1016/j.funeco.2014.05.004 (2014).
39. Hou, Y. et al. Microbial Strain Prioritization Using Metabolomics Tools for the Discovery of Natural Products. Analytical Chemistry
84, 4277–4283, https://doi.org/10.1021/ac202623g (2012).
40. Worley, B. & Powers, . Multivariate Analysis in Metabolomics. Curr Metabolomics 1, 92–107, https://doi.org/10.2174/221323
5X11301010092 (2013).
41. Bills, G. F. et al. Hypoxylon pulicicidum sp. nov. (Ascomycota, Xylariales), a Pantropical Insecticide-Producing Endophyte. PLOS
ONE 7, e46687, https://doi.org/10.1371/journal.pone.0046687 (2012).
42. Carroll, G. C. e foraging ascomycete. 16th International Botanical Congress, St. Louis, MO, USA, 1999 (1999).
43. Tanney, J. B., McMullin, D. ., Green, B. D., Miller, J. D. & Seifert, . A. Production of antifungal and antiinsectan metabolites by the
Picea endophyte Diaporthe maritima sp. nov. Fungal. Biology 120, 1448–1457, https://doi.org/10.1016/j.funbio.2016.05.007 (2016).
44. Laessøe, T. Xylaria Corniformis econsidered (1987).
45. ogers, J. D. Xylaria bulbosa, Xylaria curta, and Xylaria longipes in Continental United States. Mycologia 75, 457–467, https://doi.
org/10.2307/3792687 (1983).
46. Jones, O. A. H., Spurgeon, D. J., Svendsen, C. & Grin, J. L. A metabolomics based approach to assessing the toxicity of the
polyaromatic hydrocarbon pyrene to the earthworm Lumbricus rubellus. Chemosphere 71, 601–609, https://doi.org/10.1016/j.
chemosphere.2007.08.056 (2008).
47. Erisson, L., Byrne, T., Johansson, E., Trygg , J. & Viström, C. Multi-and megavariate data analysis basic principles and applications.
Vol. 1 (Umetrics Academy, 2013).
48. Caboche, S. et al. NOINE: a database of nonribosomal peptides. Nucleic Acids es 36, D326–D331, https://doi.org/10.1093/nar/
gm792 (2008).
49. Nielsen, . F., Månsson, M., an, C., Frisvad, J. C. & Larsen, T. O. Dereplication of Microbial Natural Products by LC-DAD-
TOFMS. Journal of Natural Products 74, 2338–2348, https://doi.org/10.1021/np200254t (2011).
50. Nielsen, . F. & Smedsgaard, J. Fungal metabolite screening: database of 474 mycotoxins and fungal metabolites for dereplication by
standardised liquid chromatography–UV–mass spectrometry methodology. Journal of Chromatography A 1002, 111–136, https://
doi.org/10.1016/S0021-9673(03)00490-4 (2003).
51. Li, Y.-Y., Hu, Z.-Y. & Shen, Y.-M. J. N. p. c. Two new cyclopeptides and one new nonenolide from Xylaria sp. 101. 6,
1934578X1100601214 (2011).
52. Wu, W. et al. Isolation and Structural Elucidation of Proline-Containing Cyclopentapeptides from an Endolichenic Xylaria sp.
Journal of Natural Products 74, 1303–1308, https://doi.org/10.1021/np100909y (2011).
53. Ju, Y.-M., Hsieh, H.-M. & Dominic, S. J. N. A. F. e Xylaria names proposed by CG Lloyd. 11, 1-31 (2016).
54. Hsieh, H.-M. et al. Phylogenetic status of Xylaria subgenus Pseudoxylaria among taxa of the subfamily Xylarioideae (Xylariaceae)
and phylogeny of the taxa involved in the subfamily. Molecular Phylogenetics and Evolution 54, 957–969, https://doi.org/10.1016/j.
ympev.2009.12.015 (2010).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
16
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
55. Ju, Y.-M., Hsieh, H.-M., Vasilyeva, L. & Aulov, A. ree new Xylaria species from ussian Far East. Mycologia 101, 548–553 (2009).
56. Fries, E. Elenchus fungorum, sistens commentarium in Systema mycologicum. Vol. 2 (Symptibus Ernesti Mautitii, 1828).
57. Tanney, J. B. & Seifert, . A. Lophodermium resinosum sp. nov. from red pine (Pinus resinosa) in Eastern Canada. Botany 95,
773–784, https://doi.org/10.1139/cjb-2017-0012 (2017).
58. Pažoutová, S. et al. A new endophytic insect-associated Daldinia species, recognised from a comparison of secondary metabolite
proles and molecular phylogeny. 60, 107–123, https://doi.org/10.1007/s13225-013-0238-5 (2013).
59. Pažoutová, S., Šrůta, P., Holuša, J., Chudíčová, M. & olaří, M. Diversity of xylariaceous symbionts in Xiphydria woodwasps: role
of vector and a host tree. Fungal. Ecology 3, 392–401 (2010).
60. Tanney, J. B., McMullin, D. . & Miller, J. D. In Endophytes of Forest Trees: Biology and Applications (eds. Anna Maria Pirttilä & A.
Carolin Fran) 343–381 (Springer International Publishing, 2018).
61. Brian, P. Griseofulvin. Transactions of the British Mycological Society 43, 1–13 (1960).
62. Deer, J. & Tullenaars, I. J. M. L. G. e antibiotic griseofulvin, some aspects of its mode of action. 28, 574–579 (1963).
63. Napier, E. J., Turner, D. I. & hodes, A. e In Vitro Action of Griseofulvin against Pathogenic Fungi of Plants. Annals of Botany 20,
461–466, https://doi.org/10.1093/oxfordjournals.aob.a083536%J (1956).
64. Par, J.-H. et al. Griseofulvin from Xylaria sp. strain F0010, an endophytic fungus of Abies holophylla and its antifungal activity
against plant pathogenic fungi. 15, 112–117 (2005).
65. Crowdy, S. H., Grove, J. F., Hemming, H. G. & obinson, . C. e Translocation of Antibiotics in Higher Plants: II. e movement
of griseofulvin in broad bean and tomato. Journal of Experimental Botany 7, 42–64 (1956).
66. Cas ella, T. M. et al. Antimicrobial and cytotoxic secondary metabolites from tropical leaf endophytes: Isolation of antibacterial agent
pyrrocidine C from Lewia infectoria SNB-GTC2402. Phytochemistry 96, 370–377, https://doi.org/10.1016/j.phytochem.2013.10.004
(2013).
67. uachaisiriul, V. et al. Indanone and mellein derivatives from the Garcinia-derived fungus Xylaria sp. PSU-G12. Phytochemistry
Letters 6, 135–138 (2013).
68. Sica, V. P. et al. Spatial and Temporal Proling of Griseofulvin Production in Xylaria cubensis Using Mass Spectrometry Mapping.
Front Microbiol 7, 544–544, https://doi.org/10.3389/fmicb.2016.00544 (2016).
69. McMullin, D. ., Nguyen, H. D. T., Daly, G. J., Menard, B. S. & Miller, J. D. Detection of foliar endophytes and their metabolites in
Picea and Pinus seedling needles. Fungal. Ecology 31, 1–8, https://doi.org/10.1016/j.funeco.2017.09.003 (2018).
70. Fischbach, M. A. & Walsh, C. T. Assembly-Line Enzymology for Polyetide and Nonribosomal Peptide Antibiotics: Logic,
Machinery, and Mechanisms. Chemical eviews 106, 3468–3496, https://doi.org/10.1021/cr0503097 (2006).
71. Gallo, A., Ferrara, M. & Perrone, G. Phylogenetic Study of Polyetide Synthases and Nonribosomal Peptide Synthetases Involved in
the Biosynthesis of Mycotoxins. 5, 717–742 (2013).
72. Schwarzer, D., Fining, . & Marahiel, M. A. Nonribosomal peptides: from genes to products. Natural Product eports 20, 275–287,
https://doi.org/10.1039/B111145 (2003).
73. Walsh, C. T. A chemocentric view of the natural product inventory. Nature Chemical Biology 11, 620, https://doi.org/10.1038/
nchembio.1894 (2015).
74. Baral, H. O. Lugols solution/II versus Melzer’s reagent: hemiamyloidity, a universal feature of the ascus wall. Mycotaxon 29,
399–450 (1987).
75. ornerup, A. & Wanscher, J. H. J. M. H. o. C. Methuen handboo of colour (1963).
76. Visagie, C. M. et al. Identication and nomenclature of the genus Penicillium. Stud Mycol 78, 343–371, https://doi.org/10.1016/j.
simyco.2014.09.001 (2014).
77. White, T. J., Bruns, T., Lee, S. & Taylor, J. J. Ppagtm & applications. Amplication and direct sequencing of fungal ribosomal NA genes
for phylogenetics. 18, 315–322 (1990).
78. Larena, I., Salazar, O., González, V., Julián, Ma. C. & ubio, V. Design of a primer for ribosomal DNA internal transcribed spacer
with enhanced specicity for ascomycetes. Journal of Biotechnology 75, 187–194, https://doi.org/10.1016/S0168-1656(99)00154-6
(1999).
79. Glass, N. L. & Donaldson, G. C. Development of primer sets designed for use with the PC to amplify conserved genes from
lamentous ascomycetes. Applied and Environmental Microbiology 61, 1323 (1995).
80. Liu, Y. J., Whelen, S. & Hall, B. D. Phylogenetic relationships among ascomycetes: evidence from an NA polymerse II subunit.
Molecular Biology and Evolution 16, 1799–1808, https://doi.org/10.1093/oxfordjournals.molbev.a026092%J (1999).
81. Carbone, I. & ohn, L. M. A method for designing primer sets for speciation studies in lamentous ascomycetes. Mycologia 91,
553–556 (1999).
82. Vilgalys, . & Hester, M. apid genetic identification and mapping of enzymatically amplified ribosomal DNA from several
Cryptococcus species. J Bacter iol 172, 4238–4246, https://doi.org/10.1128/jb.172.8.4238-4246.1990 (1990).
83. atoh, . & Standley, D. M. MAFFT Multiple Sequence Alignment Soware Version 7: Improvements in Performance and Usability.
Molecular Biology and Evolution 30, 772–780, https://doi.org/10.1093/molbev/mst010%J (2013).
84. Nylander, J. A. A. MrModeltest v2. Program distributed by the author (2004).
85. Bruns, T. M., Gaunt, . A. & Weber, D. J. Estimating bladder pressure from sacral dorsal root ganglia recordings. Conference
proceedings: Annual International Conference of the IEEE Engineering in Medicine and Biology Society. IEEE Engineering in
Medicine and Biology Society. Conference 2011, 4239–4242 (2011).
86. Stamatais, A. AxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30,
1312–1313 (2014).
87. onquist, F. et al. MrBayes 3.2: Ecient Bayesian Phylogenetic Inference and Model Choice Across a Large Model Space. Systematic
Biology 61, 539–542, https://doi.org/10.1093/sysbio/sys029%J (2012).
88. ambaut, A., Suchard, M., Xie, D. & Drummond, A. (etrieved from htp://beast. bio. ed. ac. u/Tracer, 2014).
89. Smith, C. A., Want, E. J., O’Maille, G., Abagyan, . & Siuzda, G. XCMS: Processing Mass Spectrometry Data for Metabolite
Proling Using Nonlinear Pea Alignment, Matching, and Identication. Analytical Chemistry 78, 779–787, https://doi.org/10.1021/
ac051437y (2006).
90. uhl, C., Tautenhahn, . & Neumann, S. J. A. C. LC-MS pea annotation and identication with CAMEA. 84, 1–14 (2010).
91. uhlisch, C. & Pohnert, G. J. N. p. r. Metabolomics in chemical ecology. 32, 937–955 (2015).
92. Broadhurst, D. I. & ell, D. B. J. M. Statistical strategies for avoiding false discoveries in metabolomics and related experiments. 2,
171–196, https://doi.org/10.1007/s11306-006-0037-z (2006).
93. Gottlieb, H. E., otlyar, V. & Nudelman, A. J. T. J. o. o. c. NM chemical shis of common laboratory solvents as trace impurities. 62,
7512–7515 (1997).
Acknowledgements
The authors would like to thank M. Kelman (AAFC), T. McDowell (AAFC) and D. Sorensen (McMaster
University) for technical assistance. We would like to thank Linda Ejim (McMaster University) for antimicrobial
testing of purified compounds. We thank The Center for Microbial Chemical Biology (CMCB), and the
Biointerfaces Institutes (BI) at McMaster University for access to state-of-the-art instrumentation. J.B. Tanney
thanks Jacques Fournier, Ju Yu-Ming, and Marc Stadler for insightful discussion on Xylaria taxonomy. A.I was
Content courtesy of Springer Nature, terms of use apply. Rights reserved
17
SCIENTIFIC REPORTS | (2020) 10:4599 | https://doi.org/10.1038/s41598-020-61088-x
www.nature.com/scientificreports
www.nature.com/scientificreports/
funded through an Ontario Graduate Scholarship (OGS) Doctoral Research Award. is project was funded by
an AAFC grant to MWS and KAS. Additional support was provided by the Natural Sciences and Engineering
Research Council of Canada (NSERC SYN 479724-15) to J.D. Miller and by J.D. Irving Ltd.
Author contributions
Ashraf Ibrahim, Joey Tanney, and Mark Sumarah conducted the primary research. Fan Fei performed the
statistical analysis. Ashraf Ibrahim, Mark Sumarah, J. David Miller and Keith Seifert conceived the experiments.
Alfredo Capretta discussed research and structural characterization. Chris Cutler provided blueberry samples for
endophyte isolation. All authors contributed to the manuscript writing and review.
Competing interests
e authors declare no competing interests.
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41598-020-61088-x.
Correspondence and requests for materials should be addressed to M.W.S.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
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 Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted 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 license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2020
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com
... To date, our preliminary investigations within this collection resulted in the discovery of several novel and bioactive secondary metabolites including the antifungal polyketides trienylfuranones from the raspberry endophyte Hypomontagnella submonticulosa, 5 the antibacterial non-ribosomal peptides ellisiiamides from the blueberry-Pinus endophyte Xylaria ellisii, 6,7 and the antimicrobial polyketides nemanilactones and nemanifuranones from the grape endophyte tentatively identified as Nemania serpens. 8 However, only a fraction of the species within the collection have been investigated for their ability to produce novel compounds. ...
... A metabolomics-guided approach may also be taken, where culture extracts are first screened by LC−HRMS and are grouped statistically using multivariate analysis, such as orthogonal projections to latent structures discriminant analysis (OPLS-DA) or principal component analysis (PCA), based on their secondary metabolite profiles. 6 This approach allows strains with common metabolites to be grouped together, while unique metabolite profiles can be readily identified as divergent outliers. Furthermore, the datasets can be rapidly dereplicated or, in other words, have known compounds identified, saving time and costly efforts in purification and structural characterization. ...
... Seed spectra of known compounds were included in the network and belonged to compounds previously isolated from our fungal endophyte collection, including Coniochaeta tetraspora, Nigrospora sphaerica, Sphaerulina vaccinii, Xylaria castorea, and X. ellisii. 6,19 All but X. ellisii were previously identified tentatively based on ITS data. Processing the LC−MS data by XCMS yielded ∼8000 molecular features in the PCA plot, therefore in the resulting molecular network, the cosine score cutoff was set at 0.75�slightly higher than the default setting of 0.6�to reduce the number of connections that warrant investigation. ...
Article
Full-text available
Many diverse species of fungi naturally occur as endophytes in plants. The majority of these fungi produce secondary metabolites of diverse structures and biological activities. Culture extracts from 288 fungi isolated from surface-sterilized blueberries, cranberries, raspberries, and grapes were analyzed by LC-HRMS/MS. Global Natural Products Social (GNPS) Molecular Networking modeling was used to investigate the secondary metabolites in the extracts. This technique increased the speed and simplicity of dereplicating the extracts, targeting new compounds that are structurally related. In total, 60 known compounds were dereplicated from this collection and seven new compounds were identified. These previously unknown compounds are targets for purification, characterization, and bioactivity testing in future studies. The fungal endophytes characterized in this study are potential candidates for providing bio-protection to the host plant with a reduced reliance on chemical pesticides.
... Three new proline-containing cyclic nonribosomal peptides, including ellisiiamides A-C (240-242), were isolated from the blueberry Vaccinium angustifolium-derived fungus Xylaria ellisii. Compound 240 showed modest inhibitory activity against E. coli, with an MIC value of 100 µg/mL [82]. ...
... (BCC 21097) Mangrove plant Cytotoxicity [24] 7α,10α-Dihydroxy-1β-methoxyeremophil-11(13)-en-12,8β-olide) (55) Xylaria sp. (BCC 21097) Mangrove plant Cytotoxicity, antimalarial activity [24] 1α,10α-Epoxy-7α-hydroxyeremophil-11 (13) Ellisiiamides B-C (241-242) Xylaria ellisii Blueberry Vaccinium angustifolium - [82] Cyclo ...
Article
Full-text available
The fungus genus Xylaria is an important source of drug discoveries in scientific fields and in the pharmaceutical industry due to its potential to produce a variety of structured novel and bioactive secondary metabolites. This review prioritizes the structures of the secondary metabolites of Xylaria spp. from 1994 to January 2024 and their relevant biological activities. A total of 445 new compounds, including terpenoids, nitrogen-containing compounds, polyketides, lactones, and other classes, are presented in this review. Remarkably, among these compounds, 177 compounds show various biological activities, including cytotoxic, antimicrobial, anti-inflammatory, antifungal, immunosuppressive, and enzyme-inhibitory activities. This paper will guide further investigations into the structures of novel and potent active natural products derived from Xylaria and their potential contributions to the future development of new natural drug products in the agricultural and medicinal fields.
... The most representatives of genus Xylaria are considered saprophytes, sometimes from slight to strongly parasites most often found in association with the stem and leaves and rarely on fruits [4]. Despite being saprophytes, Xylaria species are among the predominant fungal endophytes that can colonize asymptomatically wide range of hosts such as herbs and woody plants, seeds, fallen leaves and fruits or are the mutualistic symbionts in lichens (endolichens) at least in one phase of their life cycle [4,[9][10][11][12][13][14][15][16][17][18]. Also, some species could be found on dung or are associated with insects' nests [11,12,19]. ...
... The cultural characteristics and morphology are a reliable way for identification of xylariaceous fungi, but the limited taxonomic resolution of the asexual state (only morphological characters) often could be a reason for inaccurate identification [13,17,64]. ...
Article
Full-text available
The present study is the first to report Xylaria karsticola isolated from the basidiocarp of Macrolepiota procera (Basidiomycota), from Stara Planina Mountain, Bulgaria and second report for such species found in Europe. The fungal isolate was in vitro cultivated and the morphology was observed. It was primarily determined as a xylariaceous morphotype at the intragenus level, based on the evaluation of colony growth rate, color, and stromatic structure formation and was confirmed by unique conidiophores and conidia. The molecular identification of the isolate was performed by amplification of ITS1-5.8S-ITS2 region and the strain was identified as Xylaria karsticola with 97.57% of confidence. The obtained sequence was deposited in the GenBank database under the accession number MW996752 and in the National Bank of Industrial Microorganisms and Cell Cultures of Bulgaria under accession number NBIMCC 9097. The phylogenetic analysis of the isolate was also conducted by including 26 sequences obtained from different Xylaria isolates. Considering the phylogenetic data, X. karsticola NBIMCC 9097 was grouped along with other X. karsticola isolates, although the DNA sequence of the novel X. karsticola was rather distantly related to the other X. karsticola sequence data. The results were supported by the bootstrap analysis (100%) and indicated the different origin of the examined X. karsticola NBIMCC 9097.
... The family Xylariaceae contains a large number of commonly reported endophytes in a very broad range of host plants, including bryophytes, liverworts, angiosperms, and conifers, and even occurs in the form of endolichenic fungi [126]. These fungi are not only ubiquitous endophytes but also remarkably prolific producers of bioactive natural products [63,127]. ...
Article
Full-text available
Plant diseases and pests reduce crop yields, accounting for global crop losses of 30% to 50%. In conventional agricultural production systems, these losses are typically controlled by applying chemical pesticides. However, public pressure is mounting to curtail agrochemical use. In this context, employing beneficial endophytic microorganisms is an increasingly attractive alternative to the use of conventional chemical pesticides in agriculture. A multitude of fungal endophytes are naturally present in plants, producing enzymes, small peptides, and secondary metabolites due to their bioactivity, which can protect hosts from pathogens, pests, and abiotic stresses. The use of beneficial endophytic microorganisms in agriculture is an increasingly attractive alternative to conventional pesticides. The aim of this study was to characterize fungal endophytes isolated from apparently healthy, feral wine grapes in eastern Canada that have grown without agrochemical inputs for decades. Host plants ranged from unknown seedlings to long-lost cultivars not widely propagated since the 1800s. HPLC-MS was used to identify unique endophyte-derived chemical compounds in the host plants, while dual-culture competition assays showed a range in endophytes’ ability to suppress the mycelial growth of Botrytis, which is typically controlled in viticulture with pesticides. Twelve of the most promising fungal endophytes isolated were identified using multilocus sequencing and morphology, while DNA barcoding was employed to identify some of their host vines. These fungal endophyte isolates, which consisted of both known and putative novel strains, belonged to seven genera in six families and five orders of Ascomycota. Exploring the fungal endophytes in these specimens may yield clues to the vines’ survival and lead to the discovery of novel biocontrol agents.
... Analytical reproducibility is regarded as one of the most crucial factors in metabolomic studies. Among the most common studies used for this function are principal component analysis (PCA), partial least square discriminant analysis (PLS-DA), and orthogonal partial least square discriminant analysis (OPLS-DA) which have been employed in order to identify novel proline-containing cyclic non-ribosomal peptides from Xylaria ellisii which is a leaf and stem endophyte of Vaccinium angustifolium (Ibrahim et al. 2020) and investigate the fungal endophytes connected to the well-known antimalarial herb Artemisia annua for their capacity to produce additional antimalarial drugs (Alhadrami et al. 2021). Another study used untargeted metabolomic analysis to investigate the effect of histone deacetylase (HDAC) inhibition on the production of a bioactive metabolite by an endophytic Aspergillus nidulans. ...
Chapter
Full-text available
The microbial communities that reside in the plant endosphere without exhibiting any apparent disease symptoms are known as endophytes, and they have the most distinct relationships in close proximity with their host plants in comparison to other plant-associated microbiota. The sturdy foundation for natural metabolic scaffolds with a wide range of applications is provided by bioactive secondary metabolites from endophytic fungus. The highly significant metabolites of these filamentous fungi are encoded by gene clusters, but many of these gene clusters are hidden in lab-like environments. Due to this reason, various approaches involving genetic modification(s) and/or metabolomic interventions are essential for stimulating such dormant gene clusters and enhancing metabolite production. In this chapter, chemical profiling of fungal endophyte-derived bioactive compounds has been discussed using omics-based approaches. The study about the association between plants and fungal endophytes using a variety of techniques, including genome sequencing, comparative genomics, microarray, next-generation sequencing, metagenomics, and metatranscriptomics, has also been covered in this topic.
... The family Xylariaceae contains a large number of commonly reported endophytes in a very broad range of host plants, including bryophytes, liverworts, angiosperms and conifers, even occurring in the form of endolichenic fungi [85]. These fungi are not only ubiquitous endophytes, but also remarkably prolific producers of bioactive natural products [86,87]. ...
Preprint
Full-text available
Plant diseases and pests reduce crop yields, accounting for global crop losses of 30% to 50%. In conventional agricultural production systems, these losses are typically controlled by applying chemical pesticides. However, public pressure is mounting to curtail agrochemical use. In this context, employing beneficial endophytic microorganisms is an increasingly attractive alternative to the use of conventional chemical pesticides in agriculture. A multitude of fungal endophytes are naturally present in plants, producing enzymes, small peptides and secondary metabolites due to their bioactivity, can protect hosts from pathogens, pests and abiotic stresses. The use of beneficial endophytic microorganisms in agriculture is an increasingly attractive alternative to conventional pesticides. The aim of this study was to characterize fungal endophytes isolated from apparently healthy, feral wine grapes in eastern Canada that have grown without agrochemical inputs for decades. Host plants ranged from unknown seedlings to long-lost cultivars not widely propagated since the 1800s. HPLC-MS was used to identify unique endophyte-derived chemical compounds in the host plants, while dual-culture competition assays showed a range in endophytes’ ability to suppress the mycelial growth of Botrytis, which is typically controlled in viticulture with pesticides. Twelve of the most promising fungal endophytes isolated were identified using multilocus sequencing and morphology, while DNA barcoding was employed to identify some of their host vines. These fungal endophyte isolates, which consisted of both known and putative novel strains, belonged to seven genera in six families and five orders of Ascomycota. Exploring the fungal endophytes in these specimens may yield clues to the vines’ survival and lead to the discovery of novel biocontrol agents.
... These findings suggest this tropical tree has a fungal endophyte rich. Remarkably, it was found cases of genera such as Trametes and Xylaria inside leaves or stems, similar to several endophyte species from these genera are already present in different parts of plants (Bertini et al., 2022;Ibrahim et al., 2020;Macías-Rubalcava and S anchez-Fern andez, 2017;Malik et al., 2020;Nelson et al., 2020;Rai et al., 2022;Rungjindamai et al., 2008;Santamaría and Bayman, 2005). In the case of T. elegans, was not only isolated as endophyte by this research group , but also was reported by many others (Guerrero et al., 2018;Liu et al., 2016;Pinruan et al., 2010;Rungjindamai et al., 2008), suggesting the idea that this wellknown macrofungus it can grow as an endophyte in nature. ...
Article
Full-text available
Fungal laccases are promising for biotechnological applications, including bioremediation and dye biotransformation, due to their high redox potential and broad substrate specificity. However, current bioprospecting methods for identifying laccase-producing fungi can be challenging and time-consuming. For early detection, it was developed a three-step, multi-criteria weighting system that evaluates fungal strains based on: First, the biotransformation capacity of three dyes (i.e., Congo red, brilliant blue G-250, and malachite green), at three different pH values, and with a relative weighting supported for the redox potential of each colorant. The relative decolorization coefficient (RDC), used as th2e first classification criterion, expressed their potential performance. Second, under the same conditions, laccase activity was estimated by observing the different degrees of oxidation of a given substrate. The selection criterion was the relative oxidation coefficient (ROC). Finally, laccase activity was quantified in submerged fermentations using three inducers (i.e., loofah sponge, Tween 80, and veratyl alcohol). This multicriteria screening strategy evaluated sixteen isolated endophytic fungal strains from Otoba gracilipes. The system identified Beltraniopsis sp. ET-17 (at pH values of 5.00 and 5.50) as a promising strain for dye biotransformation, and Phlebia floridensis as the best laccase producer, achieving a high activity of 116 μmol min−1 L−1 with loofah sponge as an inducer. In-vitro testing confirmed the efficacy of P. floridensis, with 53.61 % decolorization of a dye mixture (brilliant blue-Congo red. ratio 1:1) after 15 days of incubation. Thus, with the proposed screening strategy it was possible to highlight two species of interest at an early bioprospecting stage on a Colombian native tree poorly explored.
Article
Full-text available
The genus Xylaria comprises a diverse group of fungi with a global distribution and significant ecological importance, known for being a source of bioactive secondary metabolites with antibacterial, antioxidative, anticarcinogenic, and additional properties. In this study, we present a comprehensive taxonomic revision of the species of Xylaria found in some parts of southern China, characterized by an extensive multilocus phylogeny analysis based on internal transcribed spacer (ITS), TUB2 (β‐tubulin), and DNA‐directed RNA polymerase II subunit 2 ( rpb2 ) gene regions. Morphological examination and detailed comparative analyses of the collected specimens were conducted to determine the distinctiveness of each species. The multilocus phylogeny approach allowed us to infer evolutionary relationships and assess species boundaries accurately, leading to the identification of 40 novel Xylaria species hitherto unknown to science. The newly described species are: X. baoshanensis , X. bawanglingensis , X. botryoidalis , X. dadugangensis , X. doupengshanensis , X. fanglanii , X. glaucae , X. guizhouensis , X. japonica , X. jinghongensis , X. jinshanensis , X. kuankuoshuiensis , X. liboensis , X. negundinis , X. orbiculati , X. ovata , X. pseudoanisopleura , X. pseudocubensis , X. pseudobambusicola , X. pseudoglobosa , X. pseudohemisphaerica , X. pseudohypoxylon , X. puerensis , X. qianensis , X. qiongzhouensis , X. rhombostroma , X. serratifoliae , X. shishangensis , X. shuqunii , X. shuangjiangensis , X. sinensis , X. tongrenensis , X. umbellata , X. xishuiensis , X. yaorenshanensis , X. yinggelingensis , X. yumingii , X. yunnanensis , X. zangmui , and X. zonghuangii . The study's findings shed light on the distinctiveness of the newly described species, supported by both morphological distinctions and phylogenetic relationships with their close relatives. This taxonomic revision significantly contributes to our understanding the diversity of Xylaria in China and enriches the knowledge of fungal biodiversity worldwide.
Article
Full-text available
A review of selected studies on fungal endophytes confirms the paucity of Basidiomycota and basal fungi, with almost 90% attributed to Ascomycota. Reasons for the low number of Basidiomycota and basal fungi, including the Chytridiomycota, Mucoromycota, and Mortierellomycota, are advanced, including isolation procedure and media, incubation period and the slow growth of basidiomycetes, the identification of non-sporulating isolates, endophyte competition, and fungus–host interactions. We compare the detection of endophytes through culture-dependent methods and culture-independent methods, the role of fungi on senescence of the host plant, and next-generation studies.
Article
Full-text available
Microbial endophytes are microorganisms that reside within plant tissues without causing any harm to their hosts. These microorganisms have been found to confer a range of benefits to plants, including increased growth and stress tolerance. In this review, we summarize the recent advances in our understanding of the mechanisms by which microbial endophytes confer abiotic and biotic stress tolerance to their host plants. Specifically, we focus on the roles of endophytes in enhancing nutrient uptake, modulating plant hormones, producing secondary metabolites, and activating plant defence responses. We also discuss the challenges associated with developing microbial endophyte-based products for commercial use, including product refinement, toxicology analysis, and prototype formulation. Despite these challenges, there is growing interest in the potential applications of microbial endophytes in agriculture and environmental remediation. With further research and development, microbial endophyte-based products have the potential to play a significant role in sustainable agriculture and environmental management.
Article
Full-text available
More than 100 fungal endophyte strains belonging to the family Phacidiaceae were isolated from surface-sterilized Picea rubens Sarg. needles collected from the Acadian Forest Region in New Brunswick, Canada. The strains were characterized morphologically by their asexual states, and phylogenetic analyses were conducted using the nuclear internal transcribed spacer rDNA (ITS) marker and the second largest subunit of ribosomal polymerase II (RPB2). Morphological and phylogenetic data revealed seven species: Darkera cf. parca; Strasseria geniculata; two novel Phacidium species: Phacidium dicosmoanum and Phacidium faciforme; and three novel monotypic genera described to accommodate distinct species: Calvophomopsis rubenticola, Cornibusella ungulata, and Gloeopycnis protuberans. Further analyses of Darkera spp. were performed with ITS and partial translation elongation factor 1-α (TEF1α), and the results suggest that D. parca is a species complex. Phacidiaceae includes hundreds of known species that are unrepresented by sequence data; therefore, ITS sequences were generated from herbarium material including type specimens of Darkera parca, Phacidium lunatum, and specimens of Allantophomopsiella, Allantophomopsis, Bulgaria, Phacidium, and Pseudophacidium species. The description of novel species combined with morphological observations and reference sequences will facilitate the identification of conifer endophytes, both from specimens or cultures and in environmental sequence data, and improve our understanding of this large and mostly neglected family.
Article
Full-text available
A morphologically distinct Lophodermium species was collected from fallen secondary needles of Pinus resinosa Aiton over two consecutive years in Eastern Ontario; subsequent herbarium studies confirmed its presence in Nova Scotia, Quebec, and Maine. Symptomatic needles frequently exhibited red bands and completely subepidermal ascomata and conidiomata. Ascospore isolates from specimens were used to reconstruct phylogenies inferred from internal transcribed spacer rDNA and partial actin gene sequences. Both phylogenetic analyses delineated the specimens from other sequenced Lophodermium species. Phylogenetic evidence combined with morphological characters of ascomata and conidiomata supported the distinctiveness of this species, described here as Lophodermium resinosum sp. nov.
Article
A simple method is described for designing primer sets that can amplify specific protein-encoding sequences in a wide variety of filamentous ascomycetes. Using this technique, we successfully designed primers that amplified the intergenic spacer region of the nuclear ribosomal DNA repeat, portions of the translation elongation factor 1 alpha, calmodulin, and chitin synthase 1 genes, and two other genes encoding actin and ras protein. All amplicons were sequenced and determined to amplify the target gene. Regions were successfully amplified in Sclerotinia sclerotiorum and other sclerotiniaceous species, Neurospora crassa, Trichophyton rubrum, Aspergillus nidulans, Podospora anserina, Fusarium solani, and Ophiostoma novo-ulmi. These regions are a potentially rich source of characters for population and speciation studies in filamentous ascomycetes. Each primer set amplified a DNA product of predicted size from N. crassa.
Chapter
This chapter describes the ecology of foliar endophytes of the Acadian Forest that dominates Canada’s Maritime Provinces extending into Eastern Quebec and Maine. Recent evidence has illuminated the ‘foraging ascomycete’ life habit of fungi that can be endophytic in conifer needles. These fungi can occupy several eco-niches other than the needles including as saprophytes in aquatic or terrestrial environments or as endophytes of understory species. Structurally diverse secondary antifungal and antiinsectan metabolites appear to mediate the exchange between plant and fungus. The plant provides nutrients and shelter, the fungus increases plant fitness by contributing to tolerance to herbivorous insects or needle pathogens. This work is enabled by the advent of affordable sequencing capability, a dedication to fieldwork and alpha taxonomy, and directed investigations of the metabolites produced by these interesting fungi.
Article
The diversity of secondary metabolites in the fungal order Xylariales is reviewed with special emphasis on correlations between chemical diversity and biodiversity as inferred from recent taxonomic and phylogenetic studies. The Xylariales are arguably among the predominant fungal endophytes, which are the producer organisms of pharmaceutical lead compounds including the antimycotic sordarins and the antiparasitic nodulisporic acids, as well as the marketed drug, emodepside. Many Xylariales are “macromycetes”, which form conspicuous fruiting bodies (stromata), and the metabolite profiles that are predominant in the stromata are often complementary to those encountered in corresponding mycelial cultures of a given species. Secondary metabolite profiles have recently been proven highly informative as additional parameters to support classical morphology and molecular phylogenetic approaches in order to reconstruct evolutionary relationships among these fungi. Even the recent taxonomic rearrangement of the Xylariales has been relying on such approaches, since certain groups of metabolites seem to have significance at the species, genus or family level, respectively, while others are only produced in certain taxa and their production is highly dependent on the culture conditions. The vast metabolic diversity that may be encountered in a single species or strain is illustrated based on examples like Daldinia eschscholtzii, Hypoxylon rickii, and Pestalotiopsis fici. In the future, it appears feasible to increase our knowledge of secondary metabolite diversity by embarking on certain genera that have so far been neglected, as well as by studying the volatile secondary metabolites more intensively. Methods of bioinformatics, phylogenomics and transcriptomics, which have been developed to study other fungi, are readily available for use in such scenarios.
Article
The needles of Picea glauca (white spruce) and Pinus strobus (white pine) trees infected with toxigenic fungal endophytes contain varying concentrations of their secondary metabolites that are toxic to either insect pests or needle pathogens. In the present study, liquid chromatography-mass spectrometric methods to determine needle concentrations of metabolites of four endophyte species were developed. The endophytes considered were a Phialocephala sp. (vermiculine) and Phialocephala scopiformis (rugulosin) from white spruce, as well as a Xylaria sp. (griseofulvin) and Lophodermium nitens (pyrenophorol) from white pine needles. To ensure that needles were infected with the associated fungal endophyte, suitable qPCR-based methods were also developed. There was a high degree of concordance between the qPCR analysis of the fungal mycelium and the LC-MS/MS quantification of the associated metabolites. Concentrations of the antifungal compounds griseofulvin and pyrenophorol were present in amounts that affect conifer needle diseases including white pine blister rust caused by Cronartium ribicola. Similarly, concentrations of the antiinsectan compounds vermiculine and rugulosin were in the range known to reduce the growth of Choristoneura fumiferana and mitigate foliage damage.
Article
Species and generic recognition in the order Xylariales has been uncertain due to lack of molecular data from authentic cultures, as well as overlapping morphological characteristics. In this study, we revise the families Graphostromataceae, Hypoxylaceae, Lopadostomataceae and Xylariaceae in Xylariales. Our study is based on DNA sequence data derived from living cultures of fresh isolates, data from GenBank and morphological observation of type and worldwide herbarium specimens. We also collected new specimens from Germany, Italy and Thailand. Combined analyses of ITS, LSU, RPB2 and β-tubulin sequence data were used to reconstruct the molecular phylogeny of the above families. Generic and familiar boundaries between these families are revised and presented in an updated combined phylogenetic tree. We accept six genera in Graphostromataceae, 19 genera in Hypoxylaceae, four in Lopadostomataceae and 37 genera in Xylariaceae. Five genera previously treated in Xylariaceae are placed in Amphisphaeriales genera incertae sedis and seven genera are placed in Xylariales genera incertae sedis. Two genera are placed in Sordariomycetes genera incertae sedis, while four genera are placed as Xylariomycetidae genera incertae sedis. Three genera are considered as doubtful. Barrmaelia and Cannonia, presently included in Xylariaceae are transferred to Diatrypaceae and Coniochaetales respectively, based on their morphology and phylogeny. Areolospora and Myconeesia are excluded from Xylariaceae and synonymized with Phaeosporis and Anthostomella respectively. Updated descriptions and illustrations are provided for all taxa with notes provided on each genus. Excluded and doubtful genera are listed with notes on their taxonomy and phylogeny. Taxonomic keys are provided for all revised families with morphological details for genera within the families.
Article
Ten polyketide specialized metabolites, epoxynemanione A, nemanifuranones A-F, and nemanilactones A-C, were isolated from the culture filtrate of Nemania serpens (Pers.) Grey (1821), an endophytic fungus from a Riesling grapevine (Vitis vinifera) found in Canada's Niagara region. Additionally, four known metabolites 2-(hydroxymethyl)-3-methoxy-benzoic acid, phyllostine, 5-methylmellein and a nordammarane triterpenoid were isolated. A related known metabolite 2,3-dihydro-2-hydroxy-2,4-dimethyl-5- trans-propenylfuran-3-one has also been included for structural and biological comparison to the nemanifuranones. The latter was isolated from the culture filtrates of Mollisia nigrescens, an endophytic fungus from the leaves and stems of lowbush blueberry (Vaccinium angustifolium) found in the Acadian forest of Nova Scotia, Canada. Their structures were elucidated based on 1D and 2D NMR, HRESIMS measurements, X-ray crystallographic analysis of nemanifuranone A, the nordammarane triterpenoid and 2,3-dihydro-2-hydroxy-2,4-dimethyl-5-trans-propenylfuran-3-one compounds, and comparison of NOE and vicinal 1H-1H coupling constants to literature data for relative stereochemical assignments. Nemanifuranone A possesses a rare C2 hemiacetal and was active against both Gram-negative and Grampositive bacteria.