ArticlePDF Available

Phylogenetic analysis of the distribution of deadly amatoxins among the little brown mushrooms of the genus Galerina

PLOS
PLOS ONE
Authors:

Abstract and Figures

Some but not all of the species of ’little brown mushrooms’ in the genus Galerina contain deadly amatoxins at concentrations equaling those in the death cap, Amanita phalloides . However, Galerina ’s ~300 species are notoriously difficult to identify by morphology, and the identity of toxin-containing specimens has not been verified with DNA barcode sequencing. This left open the question of which Galerina species contain toxins and which do not. We selected specimens for toxin analysis using a preliminary phylogeny of the fungal DNA barcode region, the ribosomal internal transcribed spacer (ITS) region. Using liquid chromatography/mass spectrometry, we analyzed amatoxins from 70 samples of Galerina and close relatives, collected in western British Columbia, Canada. To put the presence of toxins into a phylogenetic context, we included the 70 samples in maximum likelihood analyses of 438 taxa, using ITS, RNA polymerase II second largest subunit gene ( RPB2 ), and nuclear large subunit ribosomal RNA (LSU) gene sequences. We sequenced barcode DNA from types where possible to aid with applications of names. We detected amatoxins only in the 24 samples of the G . marginata s.l. complex in the Naucoriopsis clade. We delimited 56 putative Galerina species using Automatic Barcode Gap Detection software. Phylogenetic analysis showed moderate to strong support for Galerina infrageneric clades Naucoriopsis , Galerina , Tubariopsis , and Sideroides . Mycenopsis appeared paraphyletic and included Gymnopilus . Amatoxins were not detected in 46 samples from Galerina clades outside of Naucoriopsis or from outgroups. Our data show significant quantities of toxin in all mushrooms tested from the G . marginata s.l. complex. DNA barcoding revealed consistent accuracy in morphology-based identification of specimens to G . marginata s.l. complex. Prompt and careful morphological identification of ingested G . marginata s.l. has the potential to improve patient outcomes by leading to fast and appropriate treatment.
Content may be subject to copyright.
RESEARCH ARTICLE
Phylogenetic analysis of the distribution of
deadly amatoxins among the little brown
mushrooms of the genus Galerina
Brandon LandryID
1
, Jeannette Whitton
1
, Anna L. BazzicalupoID
1
, Oldriska CeskaID
2
, Mary
L. BerbeeID
1
*
1Department of Botany, University of British Columbia, Vancouver, British Columbia, Canada, 21809
Penshurst Road, Victoria, British Columbia, Canada
*mary.berbee@gmail.com
Abstract
Some but not all of the species of ’little brown mushrooms’ in the genus Galerina contain
deadly amatoxins at concentrations equaling those in the death cap, Amanita phalloides.
However, Galerina’s ~300 species are notoriously difficult to identify by morphology, and
the identity of toxin-containing specimens has not been verified with DNA barcode sequenc-
ing. This left open the question of which Galerina species contain toxins and which do not.
We selected specimens for toxin analysis using a preliminary phylogeny of the fungal DNA
barcode region, the ribosomal internal transcribed spacer (ITS) region. Using liquid chroma-
tography/mass spectrometry, we analyzed amatoxins from 70 samples of Galerina and
close relatives, collected in western British Columbia, Canada. To put the presence of toxins
into a phylogenetic context, we included the 70 samples in maximum likelihood analyses of
438 taxa, using ITS, RNA polymerase II second largest subunit gene (RPB2), and nuclear
large subunit ribosomal RNA (LSU) gene sequences. We sequenced barcode DNA from
types where possible to aid with applications of names. We detected amatoxins only in the
24 samples of the G.marginata s.l. complex in the Naucoriopsis clade. We delimited 56
putative Galerina species using Automatic Barcode Gap Detection software. Phylogenetic
analysis showed moderate to strong support for Galerina infrageneric clades Naucoriopsis,
Galerina,Tubariopsis, and Sideroides.Mycenopsis appeared paraphyletic and included
Gymnopilus. Amatoxins were not detected in 46 samples from Galerina clades outside of
Naucoriopsis or from outgroups. Our data show significant quantities of toxin in all mush-
rooms tested from the G.marginata s.l. complex. DNA barcoding revealed consistent accu-
racy in morphology-based identification of specimens to G.marginata s.l. complex. Prompt
and careful morphological identification of ingested G.marginata s.l. has the potential to
improve patient outcomes by leading to fast and appropriate treatment.
Introduction
Galerina, a genus of small, yellow-orange or yellow-brown mushrooms, includes species that
have been implicated in dozens of poisoning cases worldwide [1]. However, information
PLOS ONE
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 1 / 19
a1111111111
a1111111111
a1111111111
a1111111111
a1111111111
OPEN ACCESS
Citation: Landry B, Whitton J, Bazzicalupo AL,
Ceska O, Berbee ML (2021) Phylogenetic analysis
of the distribution of deadly amatoxins among the
little brown mushrooms of the genus Galerina.
PLoS ONE 16(2): e0246575. https://doi.org/
10.1371/journal.pone.0246575
Editor: Stefan Lo¨tters, Universitat Trier, GERMANY
Received: May 26, 2020
Accepted: January 22, 2021
Published: February 10, 2021
Peer Review History: PLOS recognizes the
benefits of transparency in the peer review
process; therefore, we enable the publication of
all of the content of peer review and author
responses alongside final, published articles. The
editorial history of this article is available here:
https://doi.org/10.1371/journal.pone.0246575
Copyright: ©2021 Landry et al. This is an open
access article distributed under the terms of the
Creative Commons Attribution License, which
permits unrestricted use, distribution, and
reproduction in any medium, provided the original
author and source are credited.
Data Availability Statement: All sequences are
available through GenBank and accession numbers
are listed in Supplementary S1 Fig. The alignment
is available through DRYAD https://doi.org/10.
5061/dryad.r7sqv9s9z.
about exactly which of the >300 species in the genus [2] pose a poisoning risk is incomplete
and confusing. This is partly because Galerina species are difficult to identify using just mor-
phological characters. In part, toxin analysis has usually involved destructive sampling, leaving
no voucher material to confirm identification. DNA barcoding has not previously been applied
to link identifications of specimens with toxin analysis, and toxins have not been assayed from
diverse Galerina species. Here, we connect vouchered Galerina specimens to DNA barcode
sequences and to amatoxin presence and absence in the context of the most complete molecu-
lar phylogeny of the genus to date.
Although individual Galerina mushrooms are small, the amatoxins can have dramatic con-
sequences if ingested. Given the amatoxin LD
50
(amount of substance required to kill 50% of
the test population) of 0.1 mg/kg body weight, 10 fruiting bodies of one of the toxic species
would be sufficient to poison a child weighing 20 kg [1]. Serious illness has resulted in people
of various ages when Galerina mushrooms have been confused with edible or hallucinogenic
mushrooms and eaten in quantity. By the time serious symptoms appear, 2–4 days after eating
mushrooms, the toxins have inflicted serious damage on the liver and other internal organs. A
family in Japan including a six-year-old boy ate soup containing what were probably Galerina
fasciculata, possibly mistaken for wild enoki mushrooms [3]. The older family members expe-
rienced nausea and diarrhea and then recovered, but the boy’s condition became progressively
worse. Some 36 hours after eating the soup, the boy was admitted to the hospital; 72 hours
after the meal, his liver failed. Following treatment, he slowly recovered, to be discharged after
15 days [3]. A 32-year-old Swedish woman saute
´ed and ate Galerina marginata, mistaking
them for honey mushrooms (Armillaria species). She was admitted to the hospital 17 h later
with vomiting and diarrhea, and with blood enzyme levels indicating liver damage [4]. She
recovered after nine days in the hospital. Two days after their cafeteria erred by serving a
locally sourced ’mushroom dish’ that likely contained Galerina sulcipes, a group of 13 cowork-
ers in China, aged 19–56 required 10 days of hospitalization to recover from liver and kidney
damage [5]. Although details are unavailable, in 2011, three Galerina poisoning cases including
one fatality were reported in North America [6]. There is no known antidote for amatoxin
ingestion, but case studies show that supportive therapy, such as replacing electrolytes and
keeping the patient hydrated saves lives [7,8]. Better knowledge of the taxonomic distribution
of amatoxin production may allow for better documentation of the geographic range and
abundance of toxic species. If ingested mushrooms can be identified as amatoxin-containing
species earlier, appropriate treatment can be initiated earlier, likely improving outcomes.
Deadly amatoxins in Galerina mushrooms have been documented since the mid-20
th
cen-
tury. In 1954, two patients consumed what was later identified as Galerina venenata and pre-
sented with symptoms mirroring poisoning by the death cap, Amanita phalloides [9].
Prompted by these poisoning cases, Tyler and Smith [10] used paper chromatography to show
that G.venenata contains α- and β-amanitin–two of the amatoxins, the toxic peptides identi-
fied from the genus Amanita.
To discuss the relationships of the toxin producers among the large number of Galerina
species, infrageneric clades become relevant. A series of authors have subdivided the genus
into subgenera and sections; e.g. Gulden and Hallgrı
´msson [11] and Smith and Singer [12].
The infrageneric taxa applied by different authors are only partially congruent with one
another or with molecular phylogenies [13]. For clarity of communication, Gulden et al. [13]
designated four infrageneric clades in their molecular phylogenies as informal groups "Naucor-
iopsis," "Galerina," "Tubariopsis," and "Mycenopsis," pointing out that the names "largely reflect
already recognized morphology-based subgenera or sections within Galerina." Our results are
largely congruent with these earlier studies and so we recognize Gulden et al. [13]’s four
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 2 / 19
Funding: This research was funded by grants from
the Natural Sciences and Engineering Research
Council of Canada (https://www.nserc-crsng.gc.ca/
index_eng.asp) including a Canada Graduate
Scholarship-Master’s Program and a Michael
Smith Foreign Travel Supplement to BL, and a
Discovery Grant RGPIN-2016–03746 to MLB.
Competing interests: The authors have declared
that no competing interests exist.
provisional clades as subgenera. We also apply "Sideroides" as a subgenus, based on an infrage-
neric taxon first used in Smith and Singer’s monograph [12].
Previous phylogenetic and toxin studies placed known Galerina toxin-producers in subge-
nus Naucoriopsis [1,13]. Within Naucoriopsis, amatoxins have been reported in the G.margin-
ata s.l. species complex [1]. Five other species that are also reported to contain amatoxins
are likely to be members of Naucoriopsis, although without verification by DNA barcoding.
Muraoka et al. [14] and Muraoka and Shinozawa [15] purified amatoxins from cultures of G.
fasciculata and G.helvoliceps. Besl [16] extracted amatoxins from cultures of G.beinrothii;
from dried mushrooms of G.badipes; and from both cultures and dried mushrooms of the G.
marginata species complex. Besl et al. [16] also reported negative results; toxins were not
detected from four specimens selected from among the ~200 Galerina species from outside
Naucoriopsis.
Of the toxin producers associated with specimen vouchers, the culture Galerina marginata
CBS 339.88 is the best studied. The Joint Genome Institute sequenced its complete genome.
Luo et al. [17] characterized its genes responsible for α-amanitin synthesis and used hybridiza-
tion to indicate that the same genes are present in G.venenata CBS 924.72, and G.badipes
(CBS 268.50). Surprisingly, G.badipes reportedly produced γ-amanitin but not the more com-
mon α- or β-amanitin [16].
The number of Galerina species that produce toxins is unclear. Until recently, most Galer-
ina species have been described and delimited based on micro- and macromorphological dif-
ferences. Smith and Singer’s [12] monograph on the genus distinguished 199 species of
Galerina. However, Gulden et al. [18] showed that nuclear ribosomal internal transcribed
spacer (ITS) sequence variation did not support the monophyly of species from vouchers
labeled G.marginata,G.autumnalis,G.unicolor,G.oregonensis, and G.venenata. Gulden et al.
synonymized all of these under G.marginata. The study left unclear whether other species
should be included in G.marginata. The possibility remained that cryptic species may be con-
tained in a group that we refer to as ’G.marginata s.l.’.
Galerina appears polyphyletic in molecular phylogenies that draw on ITS and large ribo-
somal subunit (LSU) data [13,18]. Gymnopilus appears nested within Galerina’s subgenus
Mycenopsis with a Bayesian posterior probability of 1.00. Other Galerina species were inter-
mingled with Phaeocollybia,Hebeloma and other genera, mostly without strong Bayesian sup-
port [13]. Suggesting that some of the apparent Galerina polyphyly reflected lack of data, when
Matheny et al. [19] used more data, 4508 aligned sites from a combination of ribosomal and
RPB2 (encoding the RNA polymerase II second largest subunit B150) gene sequences, phylog-
enies no longer showed Phaeocollybia and Hebeloma intermingled with Galerina. Matheny
et al. transferred Galerina clavus, which was clearly not a Galerina, to a new genus, Romagne-
siella [19]. These results suggested encouragingly that including RPB2 with ribosomal gene
data might clarify the infrageneric structure of Galerina, putting the toxic species in a larger
phylogenetic context.
Our goal was to resolve relationships and clarify the phylogenetic distribution of amatoxins
among Galerina species. To more closely characterize poisonous species, we aimed to analyze
DNA and toxins of vouchered Galerina collections from the UBC Herbarium in the Beaty Bio-
diversity Museum (https://herbweb.botany.ubc.ca/herbarium/search.php?Database=fungi).
Many of these are recently accessioned collections made by regional mycologists, especially
Oldriska Ceska and Paul Kroeger. Discovering which Galerina species contain amatoxins is
technically straightforward because a small amount of fungal tissue suffices for both toxin anal-
ysis and DNA barcoding. Two studies [20,21] have demonstrated that amatoxins are readily
detected and quantified via liquid chromatography-mass spectrophotometry from as little as 8
mg dried Amanita, even in herbarium specimens that were 17 years old. Using preliminary
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 3 / 19
ITS phylogenies to represent the diversity of clades in Galerina, we selected specimens for
toxin analysis and for sequencing of partial LSU and RPB2 regions. To help guide applications
of names, we borrowed specimens including types determined by A.H. Smith and sequenced
their ITS1 regions. We quantified α-amanitin concentrations from a diverse sample of 62
DNA-barcoded UBC Galerina specimens and eight species from closely related genera. Inte-
grating toxin data in a broad phylogenetic framework gives us new power to predict toxicity
from morphology and to speed identifications of specimens involved in possible poisoning
cases.
Materials and methods
Taxon sampling, DNA amplification and phylogenetic analysis
For this study, we re-analyzed ITS sequences of Galerina specimens from UBC determined
previously by Bazzicalupo et al. [22]. For each collection, DNA extraction, PCR amplification,
and ITS sequencing had been replicated [22]. We analyzed the ITS sequences of 147 Galerina
collections from which we recovered the same sequence in each of two independent extrac-
tions (S1 Table). For RPB2 and LSU amplifications, we extracted additional DNA from speci-
mens selected to represent the diversity of lineages as estimated from preliminary analyses of
the ITS data. We extracted DNA from 5–20 mg of gill tissue following instructions in the Qia-
gen DNEasy Plant Mini Kit for PCR amplification with Illustra PuReTaq Ready-To-Go PCR
beads (GE Healthcare: Mississauga, ON, Canada). We used primers LR0R and LR5 [23] for
LSU gene amplifications. For RPB2, we initially used primers bRPB2-6F and bRPB2-7.1R [24].
The PCR cycles began with an initial denaturation at 95˚C for 5 min, followed by 30 cycles of
95˚C denaturation for 30 sec, 55˚C annealing for 30 sec, 72˚C elongation for 30 sec, increasing
the elongation time by 4 sec per cycle and concluding with a final elongation at 72˚C for 7
minutes. For RPB2 samples that gave only weak bands or no bands at all, we re-amplified the
product in nested PCR reactions using bRPB2-7R [24] and a re-designed internal forward
primer berniF 5’ ATG GTG TGC CCT GCG GAA AC. For forward and reverse Sanger
sequencing, we used BigDye Terminator v3.1 (Thermo Fisher Scientific: MA, USA) following
the manufacturer’s instructions. The UBC Sequencing and Bioinformatics Consortium per-
formed the electrophoresis.
The 368 ITS sequences analyzed included 161 sequences from UBC specimens of Galerina,
Hebeloma and Gymnopilus, genera representing the family Hymenogastraceae. To help associ-
ate names with clades, we sequenced the ITS1 regions from 14 Galerina specimens from
MICH and examined by A.H. Smith, including types where possible. Also to help associate
names with clades, we used sequences from Gulden et al. [13,18]. We used a series of BLAST
searches to select additional GenBank sequences to represent the known diversity in the genus
and we included ITS sequences of Psilocybe in addition to Galerina,Hebeloma and Gymnopilus
in the analysis. We selected 154 sequences from the 5’ end of the LSU, 28 of them determined
for this study, and 78 RPB2 sequences, 24 from this study to represent Galerina and closely-
related families Hymenogastraceae, Strophariaceae, Crepidotaceae, Inocybaceae, Tubariaceae,
Bolbitiaceae and Cortinariaceae. For voucher information and GenBank accession numbers,
see S1 Table.
We used the MAFFT online server with the L-INS setting [25] to obtain initial alignments
for each locus, then refined the alignments manually using Mesquite 3.5 [26]. For the RPB2
dataset, we excluded introns from the final alignment. Using jModelTest 2 [27] implemented
on the CIPRES portal [28], we selected, as best nucleotide substitution models, (AICc) GTR+I
+G for the ITS and LSU datasets; TIM1+I+G for RPB2 codon position 1; and TVM+I+G for
RPB2 codon positions 2 and 3. For analyses, we approximated the best models using GTR+I
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 4 / 19
+G throughout. For each individual alignment and the concatenated alignment, we used
RAxML v.8.2.10 [29] on the CIPRES portal to infer maximum likelihood trees from 200 ’thor-
ough’ searches. We used 500 bootstrap replicates to assess branch support. Conflicts in the
topologies from individual loci generally involved weakly to moderately supported nodes
(<70% bootstrap) (S2S4 Figs), so we concatenated the alignments in Mesquite.
For subsequent analyses of concatenated LSU, RPB2 and ITS data, the ITS regions of the
more distant outgroups were too variable to align, and so we included only species of Galerina,
Gymnopilus,Psilocybe and Hebeloma. We included sequence data from each specimen ana-
lyzed for toxins and from each specimen represented by data from LSU or RPB2 sequence
regions. We included a representative of each unique ITS haplotype. To increase geographical
sampling, we included a representative of each country of origin from among sequences with
the same haplotype. The resulting dataset included 337 taxa and 4401 aligned positions and is
available through DRYAD: https://doi.org/10.5061/dryad.r7sqv9s9z. We partitioned the input
alignments by locus, and for RPB2, by codon position. We again used RAxML for 200 likeli-
hood searches and 500 bootstrap replicates.
Amatoxin detection
We analyzed amatoxin concentrations from 70 specimens, from 62 Galerina, four Gymnopilus,
three Hebeloma, and one specimen of Flammula alnicola. For Galerina specimens, we analyzed
two ~5 mg replicate tissue samples for 36 of these specimens. We analyzed only one ~5 mg
sample each from 26 specimens that were too small to allow replicated sampling. We tested
four tissue disruption methods to compare and maximize amatoxin extraction efficiency:
(1) no tissue grinding, (2) grinding with a plastic pestle, (3) grinding with a wooden stir stick
and (4) vortexing the tissue with a glass bead. Tissue grinding with a wooden stir stick was
most efficient and we used it for all subsequent samples. After grinding, we added 50% metha-
nol to each tube at a ratio of 40 μL/mg starting tissue.
After 24 hours, we centrifuged samples at 13,300 rpm for 10 minutes in an accuSpin Micro
17 centrifuge (Thermo Fisher Scientific: MA, USA) and transferred the supernatant to a new
1.5 mL tube. To remove 50% of the 50% methanol solution, we spun samples for 30–60 min-
utes in a Savant SPD111V SpeedVac (Thermo Fisher Scientific: MA, USA) and then added
sterile water to reconstitute the solution to a final volume of 200 μL. We centrifuged samples
again at 13,300 rpm, for 10 minutes. Finally, we loaded 110 μL of the supernatant into individ-
ual 1.5 mL glass autosampler vials with 0.15 mL glass inserts. As a positive control, we included
one vial containing 110 μL of 0.2 μg/μLα-amanitin standard (SIGMA A2263) dissolved in
water. Injection volume for high-performance liquid chromatography/mass spectrometry
(HPLC/MS) analysis was 100 μL.
We performed chromatographic separation using a Proto 300 C18 column (RS-2546-
W185, Higgins Analytical: CA, USA) attached to an Agilent 1200 series HPLC, multi-wave-
length detector, and Agilent 6120 Quadrupole MS (Agilent Technologies: CA, USA), with
detection at 220, 280, 295 and 310 nm [30]. Elution solution A was 20 mM ammonium acetate
pH 5 and solution B was 100% acetonitrile. The flow rate was 1 mL/min, with a gradient of
100% solution A to 100% solution B over 20 minutes. A column re-equilibration period of 10
minutes at 100% solution A was included at the end of each run.
We first determined presence or absence of α-amanitin via HPLC and UV absorbance and
confirmed the results by MS. The α-amanitin standard showed an absorption peak at 310 nm
at 8.5-minute retention time, coupled with strong MS signals for an ion with a mass/charge
(m/z) ratio of 919. We first checked the chromatograms for each Galerina sample for 310 nm
peaks at 8.5 minutes and we scanned extracted ion chromatogram MS data for compounds
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 5 / 19
with a mass/charge ratio of 919 at 8.5 minutes. Where UV absorbance, retention time, and MS
showed evidence of α-amanitin, samples were recorded as positive. Samples were recorded as
positive for β-amanitin based on a peak with the retention time of 8.0 minutes that is expected
under the chromatography conditions used [30]. Samples that did not produce a distinct peak
at 310 nm at 8.0 or 8.5 minutes and that lacked compounds with the expected mass/charge
ratio were considered toxin-negative.
Species delimitation
To delimit putative Galerina species, we used the online version of Automatic Barcode Gap
Discovery (ABGD) [31] under the assumption that within species sequence variation is usually
lower than the variation between species. We included the 314 ITS sequences from Galerina,
Psilocybe, and Gymnopilus samples that were at least 500 bp long, repeating the analysis with
and without a Kimura 2-parameter correction for multiple substitutions.
The ABGD software gives a range of broader or narrower estimates of species boundaries.
To choose among alternative estimates, we assumed that characters of sister species evolve to
show reciprocal monophyly [32], that conspecific isolates would in many cases form well-sup-
ported clades, but would lack well supported subclades [33], and that closely related species
might differ in ecology [18,34]. We did not apply a correction for multiple hits in the final
analysis because preliminary results showed that a Kimura correction increased both the num-
ber of single-sequence species and the number of paraphyletic species (with no evidence of
reciprocal monophyly). Our final ABGD analysis produced seven alternative estimates of pos-
sible species boundaries, based on a set of priors for the maximum percent within-species
divergence that ranged from 0.001 to 0.0215. These priors bracket the range of reasonable lev-
els of within-species divergence. The prior of 0.001 gave 71 putative species, many represented
by only a single sequence and nested within another species. The prior of 0.0215 put all collec-
tions in one species in spite of many supported subclades. A prior of 0.0028 with recursive par-
titions resulted in 63 putative Galerina species, six of them nested among G.marginata s.l. No
arbitrary prior is likely to be perfect and in some cases, the 63-group partition lumped well-
supported sister taxa with consistent identifications or created paraphyletic putative species.
Of the alternatives, the partition giving 63 Galerina species had the advantages of producing a
high proportion of putative species that formed clades with moderate to high bootstrap sup-
port of 70% or more, and relatively few paraphyletic species, while dividing the G.marginata s.
l. clade into species consistent with patterns of sequence variation in the ITS regions.
For additional support for species delimitation, we examined alignments for patterns of
polymorphisms among ITS sequences from closely related putative species [34] in the G.mar-
ginata s.l. complex. Where collection localities of delimited species were near one another, as
for many of the B.C. collections, interbreeding between close relatives with different ITS
sequence variants would be expected to lead to double peaks in ITS sequences that represent
heterozygosity. We examined chromatograms, correcting sequences to note double peaks in
areas of otherwise clean sequence, with special attention to sites that were polymorphic across
species. We considered that fixed sequence differences between sympatric populations of 10 or
more specimens pointed to reproductive isolation.
Results
Amatoxins in the Galerina marginata species complex
We examined the distribution of amatoxins across the Galerina phylogeny (Figs 1and 2). Of
the 62 Galerina samples assayed, all 24 amatoxin-positive samples belonged to G.marginata s.
l. in Naucoriopsis (Figs 1and 2). We detected amatoxins in dried herbarium samples collected
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 6 / 19
from 2004–2013 (S2 Table). Quantification was more difficult in samples from some of the
herbarium specimens than others due to high background noise in the chromatograms. When
amatoxin was detected, its concentration showed no obvious correlation with sample age (S2
Table).
The 24 samples that were positive for α-amanitin fell into two delimited species within G.
marginata s.l.: G.venenata and G.castaneipes. Average amatoxin concentrations in G.vene-
nata were significantly higher than the toxin concentration in G.castaneipes at P <0.05 (t-
value = 2.56; p-value = 0.018; Cohen’s d = 1.1). The average toxin concentration from the nine
G.venenata samples was 1.58 mg/g dry weight or (assuming that 88% of fresh samples was
water, p. 75 in Walton [35]), ~189 μg/g estimated wet weight (S2 Table). Based on expected
HPLC retention times, all nine G.venenata samples also contained β-amanitin. The average
toxin concentration from 14 G.castaneipes samples was 0.99 mg/g dry weight or (assuming
88% of fresh weight is water) ~117 μg/g estimated wet weight. A peak with the expected reten-
tion time for amatoxin appeared to be present but could not be quantified in one of the 15
samples of G.castaneipes, and for two additional G.castaneipes samples, toxin concentrations
were too low to quantify in at least one of the replicated measurements. Nine of the 14 G.
Fig 1. All 24 toxin-positive mushroom specimens are in subgenus Naucoriopsis of Galerina.We assayed for toxins
in 70 collections representing 17 species of Galerina and 8 species in related genera. Each fraction is the number of
samples positive for α-amanitin over the total number of specimens tested. Clade colors correspond to Galerina
subgenera or to species of Gymnopilus and Psilocybe that appear nested in Galerina (S3 Table).
https://doi.org/10.1371/journal.pone.0246575.g001
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 7 / 19
Fig 2. Distribution of toxins across 56 species of Galerina and allies. In this maximum likelihood tree, thickened branches represent 70% or more
bootstrap support from concatenated ITS, LSU and RPB2 data. Light grey boxes show monophyletic, delimited Galerina species. Dark grey boxes show
paraphyletic species. Names outside of boxes correspond to sequences that were <500 bp long and not included in delimitations. +TOX in magenta, α-
amanitin is present; -TOX in green, no amatoxins were detected; the number of collections tested is in parentheses. Verticallines designate infrageneric
groups as follows: black, G.marginata s.l.; solid purple, Naucoriopsis; dashed purple, possible Naucoriopsis; green, Galerina; blue Tubariopsis; gold
Mycenopsis; red Sideroides.
https://doi.org/10.1371/journal.pone.0246575.g002
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 8 / 19
castaneipes samples contained β-amanitin. In two samples, the presence of a β-amanitin peak
was ambiguous. Four samples of G.castaneipes showed no trace of β-amanitin.
Amatoxins were not found in any of the genera closely related to Galerina; amatoxins were
not detected from the four Gymnopilus spp., the three samples of Hebeloma or the sample of
Flammula alnicola. (Figs 1and 2). We did not detect α- or β-amanitin in Galerina badipes
F27620, which represents the sister clade to G.marginata complex within Naucoriopsis (Fig 2
and S1 Fig). No amatoxins were detected among 37 Galerina samples representing the diver-
sity of sections outside of Naucoriopsis (Fig 3).
Molecular and morphological identification of toxic Galerina
Herbarium specimens were accurately identified to Galerina and its infrageneric groups (S3
Table), based on morphological identifications later confirmed by DNA barcoding. Impor-
tantly, collections of the toxin-containing G.marginata s.l. were usually correctly identified to
this clade, and all those tested had been recognized as members of Naucoriopsis. This is
encouraging evidence that toxic galerinas can be distinguished from other mushrooms in
cases of accidental ingestion and possible poisoning, albeit with some level of expertise and
with the use of microscopic characters.
Phylogenies show that many putative Galerina species recognized by ABGD are monophy-
letic, many with >70% bootstrap support (Fig 2,S1S3 Figs; S1 and S3 Tables). However,
within each infrageneric group, the application of names to species-level clades is inconsistent
(S1 Fig). The inconsistency of species-level identifications even by specialists in the genus
points to the lack of congruence between morphological characters and genetically defined
species.
Fig 3. Toxin containing specimens in Galerina subgenus Naucoriopsis are shown in the top row; in the lower row
are examples of species in the non-toxin producing subgenera. Each species name is followed by the specimen’s
UBC voucher accession number; the Mushroom Observer photograph accession number; and in italics, the name of
the subgenus that includes the species. (a, b) Specimens producing positive tests for amatoxins. (a) G.castaneipes
F28078 MO119849, Naucoriopsis. White arrow points to inrolled cap margin in a young mushroom. (b) G.venenata
F26281 MO153552 Naucoriopsis. Black arrows point to membranous rings around the stems. (c) G.nana F25541
MO102538 Naucoriopsis (affiliation is uncertain). (d) G.atkinsoniana F28226 MO137762 Galerina. (e) G.
dimorphocystis F25868 MO129940 Tubariopsis. (f) G.subcerina F25303 MO84732 Mycenopsis. (d, e) Specimens not
tested, but ITS sequences match specimens without detectable toxins. (f) Specimens tested, no toxins detected. Scale
bar (f) is 1 cm. Scales are not available for the other images, but estimating from the mosses and cone, caps on
mushrooms (a, b) are up to ~3 cm wide. Caps on mushrooms (c-f) are ~1 cm or less wide.
https://doi.org/10.1371/journal.pone.0246575.g003
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 9 / 19
If defined phylogenetically as the sister clade to G.badipes (Fig 2,S1 Fig), Galerina margin-
ata s.l. receives 92% bootstrap support and encompasses six putative species represented by
sequences of 500 bp or longer (Fig 2,S1 Fig). Internal bootstrap support values >70% indicates
that G.marginata s.l. has more genetic structure than expected from a single species but the
putative species do not show the reciprocal monophyly expected of well-established species
(S1A Fig). Collections with identical or nearly identical ITS (S2 Fig) or RPB2 (S3 Fig)
sequences were identified under various names, frequently as G.marginata but also as G.
autumnalis,G.castaneipes,G.oregonensis,G.pseudomycenopsis,G.unicolor and G.venenata
(S1 Table).
Of the toxin-containing species, Galerina castaneipes (Figs 3a,4a and 4b), as delimited by
ABGD, appears monophyletic in all analyses (S1S4 Figs). It includes the type specimen G.cas-
taneipes AH Smith 55523, collected on rotting oak wood in Grant’s Pass, Oregon. Although
conifer wood is more common in the region, all of the other 20 collections of G.castaneipes
identified from sequencing come from collections (where wood type was recorded) from rot-
ting hardwood, from Quercus garryana or Arbutus menziesii, geographically from the south-
eastern tip of Vancouver Island, British Columbia.
Galerina venenata contains A.H. Smith’s 1953 type specimen of that species and is common
among North American and European collections (Figs 2,3b,4c and 4d,S1 Fig). A.H. Smith’s
1958 type of G.cinnamomea var. cinnamomea falls within the same clade. The G.venenata
clade appears monophyletic in the RPB2 tree (S3 Fig) but not in the ITS or concatenated trees
with better taxon sampling (Fig 2 and S1 Fig). Collection localities of the UBC specimens of G.
venenata and G.castaneipes overlapped, suggesting that parental mycelia of the two species
would have had opportunities to interbreed. However, the alignment of the ITS regions shows
three sites with fixed differences between the two species and little evidence of continuing
genetic exchange in the form of shared ITS polymorphisms (S5 Fig). Three sequences from
collections identified as species from outside G.marginata s.l. appeared in the G.venenata
clade. Of these, UBC F27894 and UBC F22840 were initially identified as G.badipes, and UBC
F24580 was identified as G.jaapii. On reexamination, all three specimens had predominantly
4-spored basidia, characteristic of G.venenata, rather than the 2-spored basidia characteristic
of G.badipes and G.jaapii. The specimen UBC F24580 had a few pleurocystidia; this character
and the shape of its cystidia led to its re-identification as G.venenata.
We label one clade "G.marginata" in the absence of another name that would apply to the
group. No type specimen of G.marginata is available to clarify the application of the name.
The clade receives 87% bootstrap support but to be monophyletic, it would have to include
specimen G.marginata UWODD6MO221929, designated by ABGD as a different species
(S1A Fig). Specimens identified as G.marginata appear in four of the putative species of G.
marginata s.l.
Pattern of confused application of names to species in non-toxic clades. Application of
species names is similarly problematical in subgenera Galerina and Sideroides, two clades
receiving >90% bootstrap support in analyses of concatenated data (Fig 2,S1 Fig). In both of
these clades, the number of monophyletic putative species is greater than the number of spe-
cies names applied to collections, and application of species identifications appears almost to
be random within delimited species (S1 Fig). In subgenus Galerina, four names are applied to
collections, but eight putative species are delimited by ABGD (S3 Table). Other than the G.
alpestris clade, each delimited species includes specimens with two or more different herbar-
ium identifications. The clade we label as ’G.vittiformis’ includes a paratype of Smith’s G.vitti-
formis var. bispora and specimens from N. America, Norway and Greenland. It is unclear
whether this clade would also include the European type of G.vittiformis. Four clades labeled
here as ’Galerina aff. vittiformis sp. 2–5’, received over 90% bootstrap support each. Some
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 10 / 19
Fig 4. Microscopic characters of toxic Galerina marginata complex include brown, minutely roughened spores
with a plage and bottle shaped cystidia. Although not specific to toxic Galerina species, these characters in any
ingested mushrooms justify medical action to mitigate possible poisoning by amatoxins. (a-d) Basidiospores. (a, b) G.
castaneipes F26244. (c-e) G.venenata, (c) F26281, (d) F18374, (e) cystidium of F26281. The alphanumeric codes are
each specimen’s UBC voucher accession number. Arrows designate the plage, the smooth area on the adaxial side of
the spore just above the apiculus (arrowheads). Scale bars, 10 μm. Spores are all to the same scale.
https://doi.org/10.1371/journal.pone.0246575.g004
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 11 / 19
clades appear to show geographical structuring. Sister to the G.vittiformis clade are five collec-
tions in two subclades, two of G.alpestris from Italy and in another subclade, three ’G.minima
collections from Norway and Greenland. All 13 collections of G. aff. vittiformis sp. 3 (Fig 3d)
are from British Columbia; both collections of its sister taxon G. aff. vittiformis sp. 4 are from
Greenland (S1 Fig). Similar problems plague naming in other clades (S3 Table).
Galerina infrageneric clades. Several Galerina infrageneric clades, variously considered
as subgenera, sections, or stirpes in previous publications (see S3 Table and Gulden et al. [13])
receive strong support from concatenated data. The divergence order of taxa at the base of
Naucoriopsis is unsupported but a core clade in Naucoriopsis that includes G.jaapii and G.cas-
taneipes receives 79% bootstrap support (S1 Fig). Galerina sect. Galerina appears as the sister
group to Naucoriopsis, with 95% bootstrap support from RPB2 (S3 Fig) and 76% support from
the concatenated dataset (S1 Fig). Section Tubariopsis appears as sister to the clade comprising
Naucoriopsis and Galerina, although with <50% bootstrap support (S1 Fig).
Gymnopilus species are consistently nested within Galerina subgenus Mycenopsis in each
individual gene tree (S2S4 Figs) and the concatenated tree (Fig 2 and S1 Fig). A subset of spe-
cies of Mycenopsis share a most recent common ancestor with Gymnopilus with 88% bootstrap
support and the clade including all Mycenopsis and Gymnopilus species receives 66% bootstrap
support (S1 Fig). The clade of five Galerina species from Sideroides receives 98% support from
concatenated data but it is distantly related to the other Galerina species and instead appears,
without strong support, as sister to Psilocybe (Fig 2,S1 Fig).
The phylogeny of RPB2 sequences (S3 Fig) shows greater resolution and overall higher sup-
port levels for relationships among Galerina species compared with the phylogenies from the
LSU (S4 Fig). With very low support values, the LSU phylogeny shows Galerina as highly para-
phyletic with other genera including Agrocybe,Hebeloma,Psilocybe and Cortinarius.
Discussion
Toxin-producing Galerina species are in sect. Naucoriopsis
All known producers of amatoxins in Galerina fall into subgenus Naucoriopsis and most are in
Galerina marginata s.l. This includes the 24 specimens that we identified by sequence data as
G.castaneipes and G.venenata, all containing detectable amatoxin quantities. Accurate quanti-
fication from the dried specimens was difficult in some cases due to unidentifiable background
peaks in chromatograms, possibly attributable to products of tissue breakdown before drying
was complete. The estimated concentrations of amatoxin in fresh samples, 189 μg/g in G.vene-
nata and 116 μg/g in G.castaneipes are comparable to 78–244 μg/g fresh weight, levels Enjal-
bert et al. [1] reported from 27 samples from specimens in the G.marginata complex. It is also
comparable to amatoxin concentrations ranging from 172–367 μg/g fresh weight in Amanita
phalloides [1].
Also in G.marginata s.l., in Naucoriopsis, and reported as toxin-positive [36], Galerina sul-
ciceps is a tropical species found in greenhouses. Toxin tests and DNA sequence barcodes are
not yet available for the same collection of G.sulciceps. The ABGD delimitation shows that the
sequence from a single collection of the species is distinctive enough to be delimited along
with G.physospora in a species separate from G.marginata,G.castaneipes and G.venenata.
Because G.physospora is close to, if not synonymous with G.sulciceps, it seems likely to also
contain amatoxins, as does G.patagonica, also in the G.marginata s.l. species complex, based
on similar reasoning.
Three other species reported in the literature as toxin-positive, Galerina beinrothii [16], G.
helvoliceps and G.fasciculata [14,15], could not be included in our molecular analyses due to
lack of DNA sequence data. Galerina beinrothii [37] and G.fasciculata [38] were originally
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 12 / 19
described as close to G.marginata. Smith and Singer [12] similarly placed G.helvoliceps near
G.marginata. These results further support our conclusion that the amatoxin-producing
Galerina species are found within the G. marginata s.l. species complex in subgenus
Naucoriopsis.
While we detected α-amanitin in all samples tested from G.marginata s. l., β-amanitin was
consistently present in the nine G.venenata samples but was undetectable from four of 15 G.
castaneipes (S2 Table). Tyler and Smith [10] detected β-amanitin in North American samples
in the initial discovery of amatoxins in Galerina. Besl et al. [16] detected β-amanitin in all sam-
ples assayed that contained α-amanitin. However, Luo et al. [17] did not detect β-amanitin or
a gene encoding it in the published genome of G.marginata CBS 339.88 [39], which based on
its ITS sequence (GenBank MH862132.1) falls in G.venenata. The β-amanitin toxin appears to
be genetically encoded in Amanita [40,41]. Sgambelluri et al. [30] speculated that some toxin
producing fungi contain an enzyme such as a deaminase that could convert the asparagine in
α-amanitin to the aspartic acid in β-amanitin. Walton (p. 75) [35] suggested that the low levels
of β-amanitin peaks may also be an artifactual deamination product of α-amanitin breakdown
but that the levels of β-amanitin reported by Enjalbert et al. [1] are much too high to be
explained by this phenomenon.
Toxin status in G.badipes (sect. Naucoriopsis) is uncertain. Galerina badipes is the only
Galerina species outside of G.marginata s.l. that is reported to contain amatoxins but we did
not detect α- or β-amanitin in our sample of G.badipes. Besl et al. [16] detected γ-amanitin, a
post-translational variant of α-amanitin [35]. Post-translational conversion of α-amanitin to
γ-amanitin could explain why neither α- nor β-amanitin have been detected in G.badipes
mushrooms, even though Luo et al. [17] detected the genes necessary for α-amanitin synthesis
in a mycelial culture of the species. Further, RNA blotting showed a much weaker α-amanitin
signal from G.badipes compared with G.marginata [17]. We note, however, that Luo et al. did
not test for amatoxin presence using HPLC/-MS. A possible explanation that is consistent with
our results and those of Besl et al. [16] is that in G.badipes,α-amanitin may be present but
below detection limits. We believe that the UBC F27620 collection of G.badipes is correctly
identified because its sequence matches others from G.badipes from Gulden et al.’s [13] study.
Toxins in vouchers of G.badipes from across its geographical range should be analyzed. Given
the confusing results, G.badipes has to be presumed to be toxic when implicated in accidental
ingestions.
We did not test other members of sect. Naucoriopsis such as G.jaapii, which may be
restricted to Europe, or other species such as Galerina triscopa that appear to be related to sect.
Naucoriopsis, although with bootstrap support <50%. While this study adds to the evidence
that amatoxins evolved once in the common ancestor of the G.marginata species complex,
further analysis of additional early diverging Naucoriopsis species could point to earlier origin
or to a more complex pattern of toxin gain and loss.
Potential pharmaceuticals from amatoxins and associated genes from Galerina spe-
cies. Although best known as toxins, amatoxins and other cycloamanides may also have uses
in medical therapies. Amatoxins conjugated with anti-tumor antibodies show potential for
treating cancer [35,4244]. Cyclic peptides with other biological activities may find other uses
as pharmaceutical products. Some have desirable pharmaceutical properties such as stability
and rapid absorption into the bloodstream [45].
Amatoxins are expensive because they are purified from the mycorrhizal Amanita phal-
loides mushrooms [45]. Unlike the as-yet-uncultured A.phalloides, the saprotrophic Galerina
species like members of Naucoriopsis grow at least slowly in culture, yielding from 0.5–1 mg
amatoxin/g dry weight [17]. Isolating a wider range of Galerina species in pure culture may
lead to the discovery of strains that grow faster and produce more amatoxin. Genetic
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 13 / 19
engineering may expand the range of useful cycloamanides produced from Galerina species’
genes. Sgambelluri et al. [45] expressed POPB, encoding the enzyme prolyl oligopeptidase B,
important in post-translational processing of amatoxins [17] from G.marginata in Saccharo-
myces cerevisiae to catalyze the cyclization of 100 different straight-chain peptide substrates
ranging from 7–16 residues to cycloamanide configurations. The POPB genes from other
Galerina species may further expand the range of potentially therapeutic cycloamanides.
Morphological and ecological characteristics to recognize toxin producing
Galerina in poisoning cases
Mushroom poisoning by amatoxins is difficult to diagnose because it takes two to four days
after ingestion before serious symptoms appear. Basidiospores and cystidia can survive cook-
ing or ingestion and should be sought in stomach contents or the remains of a meal containing
the mushroom if poisoning is a possibility. Individual mushrooms may be atypical of their
genus or species; different species often grow in close proximity and a patient may have eaten
a mixture of different mushroom species. Despite these caveats, a combination of habitat,
mushroom size and habit, and microscopic characters allow for recognition of Galerina and of
the toxic species in sect. Naucoriopsis [46] (Table 1,Fig 4).
Evolutionary relationships of clades within Galerina
The RPB2 data contributed here improves the resolution of infrageneric relationships among
Galerina. In contrast to phylogenies in Gulden et al. [13], our gene trees from concatenated
data support Galerina sections Naucoriopsis and Galerina as sister clades, consistent with their
shared microscopic features [11]. Also consistent with morphology, trees that include new
RPB2 sequences remove various other genera (Phaeocollybia,Agrocybe,Alnicola,Hebeloma,
Flammula) from the nested positions within Galerina that they take in LSU gene trees in S4
Fig and in Gulden et al. [13].
On the other hand, this study, like Gulden et al. [13] shows Gymnopilus spp. evolving from
within Galerina subgenus Mycenopsis. Gulden et al.’s analysis supported this relationship with
a posterior probability of 1.0 from LSU data. With a smaller sample of Gymnopilus and Galer-
ina but with RPB2 as well as rDNA data, Matheny et al. [19] showed the same nested relation-
ship. Gymnopilus and Galerina share spore characters including shape, ornamentation,
presence of a plage and a dextrinoid reaction, and their cystidia may be similar in form, pro-
viding support for a recent shared ancestry [13,47]. Still problematical and in need of analysis
Table 1. Comparison of characters for recognizing toxin-containing species [12,46].
Toxic Toxins not detected
Species Galerina marginata s.l., G.venenata,G.castaneipes,G.badipes 14 species representing Galerina subgenera listed below.
Subgenera Naucoriopsis Galerina,Tubariopsis,Mycenopsis,Sideroides
Cap 5–40 mm, robust compared with other Galerina spp.; hemispherical to
convex, margin inrolled when young.
Most are <20 mm diam.; larger in a few of the species. Delicate, conical to
bell-shaped, becoming convex with age. Margin is not usually inrolled when
young.
Stem width 1–4 mm, commonly with membranous ring or ring zone. A ring is
usually lacking in the less common species G.castaneipes.
Varies, but 1–2 mm in many species, mostly without a ring but white veil often
present
Cystidia On sides and edges of gills, ~30–70 μm long, often lageniform,
rounded at base, tapering to tip or occasionally subcapitate, slightly
expanded at tip.
Various; can be similar to G.marginata s.l., in others with more or less inflated
tip; or ’tibiiform’, bone-shaped with a thin, well delimited neck between an
expanded base and tip. In some species only at gill edges, not on gill faces.
Basidiospores Almond shaped, roughened, with a distinct plage. Spores brown,
dextrinoid, turning reddish in Melzer’s iodine solution.
Various, some as in G.marginata s.l.; others differ in shape, ornamentation, or
by being completely smooth or lacking a dextrinoid reaction.
Habitat On rotting wood, turf, grass or moss. Often in moss, some on rotten wood and herbs.
https://doi.org/10.1371/journal.pone.0246575.t001
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 14 / 19
from more loci is the unsupported sister relationship between a clade of Psilocybe species and
five Galerina species that form subgenus Sideroides.
Conclusion
This study combines a multi-locus sequence phylogeny with HPLC/MS toxin analysis data.
The identifications of herbarium specimens to species correlated poorly with genetic species in
this study as in previous analyses [13,18], possibly because keys based on morphology fail to
capture the amount of within- and among-species morphological variation. In spite of this, at
a higher taxonomic level specimens are reliably identified as members of Naucoriopsis, the
clade of species that produce toxins. Prompt morphological identification should enable recog-
nition of likely amatoxin-containing mushrooms, speeding diagnosis and treatment for
patients who have ingested these deadly toxic mushrooms.
Supporting information
S1 Fig. Phylogeny showing Galerina collections tested for amatoxins with species delimita-
tions and country of provenance. In this maximum likelihood tree numbers at nodes repre-
sent bootstrap support >50% from concatenated ITS, LSU and RPB2 data. Support values are
omitted from some deeply nested clades due to graphic constraints. Light grey boxes show
monophyletic, delimited Galerina species. Darker grey boxes show delimited but paraphyletic
species. A species/clade name is given in each box. Sequence names from original identifica-
tions are followed by a voucher identifier. Where applicable, the number of collections from
the same country with the same sequence is given in parentheses. +TOX in magenta, α-amani-
tin is present; -TOX in green, no amanitins were detected. Vertical lines designate subgenera
as follows: Black, G.marginata s. l.; solid purple, Naucoriopsis; dashed purple, possible Naucor-
iopsis; green, Galerina; blue Tubariopsis; gold Mycenopsis; red Sideroides. Orange designates
Gymnopilus spp. nested within Galerina.
(DOCX)
S2 Fig. Phylogeny of ITS sequences. In this maximum likelihood tree of 368 sequences, thick-
ened branches represent bootstrap support >50% from ITS data. Branch thickening is omitted
from some deeply nested clades due to graphic constraints. Light grey boxes show monophy-
letic delimited Galerina species. Darker grey boxes show delimited but paraphyletic species.
Sequences that are not boxed were less than 500 bp in length and not included in ABGD spe-
cies delimitation. A species/clade name is given in each box. Sequence names from original
identifications are followed by a voucher identifier and preceded by a number to help locate
the same voucher in RPB2 and LSU gene trees. Vertical lines designate subgenera as follows:
Black, G.marginata s. l.; solid purple, Naucoriopsis; dashed purple, possible Naucoriopsis;
green, Galerina; blue Tubariopsis; gold Mycenopsis; red Sideroides. Orange designates Gymno-
pilus spp. nested within Galerina.
(DOCX)
S3 Fig. Phylogeny of RPB2 sequences. In this maximum likelihood tree with 78 taxa, numbers
at nodes represent bootstrap support >70% from RPB2 data. Support values are omitted from
some deeply nested clades due to graphic constraints. Light grey boxes show monophyletic,
delimited Galerina species. A species/clade name is given in each box. Sequence names from
original identifications are followed by a voucher identifier and preceded by a number to help
locate the same voucher in ITS and LSU gene trees. Vertical lines designate subgenera as fol-
lows: Solid purple, Naucoriopsis; dashed purple, possible Naucoriopsis; green, Galerina; blue
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 15 / 19
Tubariopsis; gold Mycenopsis; red Sideroides. Orange designates Gymnopilus sapineus nested
within Galerina.
(DOCX)
S4 Fig. Phylogeny of LSU sequences. In this maximum likelihood tree with 154 taxa, numbers
at nodes represent bootstrap support >70% from LSU data. Support values are omitted from
some deeply nested clades due to graphic constraints. Light grey boxes show monophyletic,
delimited Galerina species. Darker grey boxes show delimited but paraphyletic species. A spe-
cies/clade name is given in each box. Sequence names from original identifications are fol-
lowed by a voucher identifier and preceded by a number to help locate the same voucher in
ITS and LSU gene trees. Vertical lines designate subgenera as follows: Solid purple, Naucoriop-
sis; dashed purple, possible Naucoriopsis; green, Galerina; blue Tubariopsis; gold Mycenopsis;
brown Sideroides. Orange designates Gymnopilus spp. nested within Mycenopsis.
(DOCX)
S5 Fig. Alignment of variable sites from the ITS region supports genetic separation of
three species in Galerina marginata s.l. Color of names at the left designates delimited spe-
cies: G.castaneipes, brown; G.venenata, blue; and G.marginata, red. Under a scenario of ran-
dom interbreeding, frequent heterozygosity rather than private alleles would be expected.
Instead, each of the three delimited species has unique nucleotide substitutions.
(DOCX)
S1 Table. Species delimitations, specimens, and GenBank accessions. ’Species Name ABGD
delimitation’ is a unique name applied to a clade within Galerina that is delimited by ABGD
software; specimens with the same ABGD group number are in the same putative species.
ABGD delimitations are only available for clades within Galerina and for ITS sequences >500
bp. ’Specimen ID’ designates the collection; under ’Toxin?’ +TOX indicates that amatoxins
were detected, -TOX indicates no amatoxins detected; ’Country’ is the two-digit code for
country of collection; ’Section’ refers to infrageneric taxon of Galerina; ’Taxon #’ is an arbitrary
specimen tracking number also used in supplementary figures; ’Original ID’ is the species
name originally applied to the herbarium collection or to the sequence.
(XLSX)
S2 Table. Results of assays to detect toxins α- and β-amanitin in Galerina species and
closely related outgroups. Taxon # is an arbitrary specimen tracking number used in supple-
mentary figures; Accession number specifies a specimen in UBC; Section is the suprageneric
classification of each species; Species is the identification based on sequence barcode analysis.
(XLSX)
S3 Table. Galerina classification, justification and notes on the application of names to
subgenera and to species.
(DOCX)
Acknowledgments
Jonathan Walton made this work possible by welcoming Brandon Landry into his laboratory
to analyze the Galerina amatoxins. More broadly, Jonathan’s research advanced the under-
standing of the toxins and their biosynthesis in a range of fungus species. His death on October
18, 2018 left a hole in the hearts of his many collaborators who benefited from his help and his
ongoing generosity. Adolf Ceska provided photographs of specimens, aided in collection, and
provided useful comments on the text. We thank Collection Managers Patricia Rogers
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 16 / 19
(MICH) and Olivia Lee (UBC) for loans and processing of specimens. Berni van der Meer con-
tributed to primer design and sequencing. We thank Gro Gulden, Natural History Museum,
University of Oslo, and reviewers Heather Hallen-Adams, University of Nebraska, Lincoln,
and Todd Osmundson, University of Wisconsin, La Crosse for helpful suggestions.
Author Contributions
Conceptualization: Anna L. Bazzicalupo, Mary L. Berbee.
Data curation: Brandon Landry, Anna L. Bazzicalupo, Oldriska Ceska, Mary L. Berbee.
Formal analysis: Brandon Landry, Anna L. Bazzicalupo, Mary L. Berbee.
Funding acquisition: Mary L. Berbee.
Investigation: Brandon Landry, Mary L. Berbee.
Methodology: Brandon Landry, Anna L. Bazzicalupo, Mary L. Berbee.
Project administration: Mary L. Berbee.
Resources: Oldriska Ceska, Mary L. Berbee.
Software: Brandon Landry, Anna L. Bazzicalupo, Mary L. Berbee.
Supervision: Jeannette Whitton, Anna L. Bazzicalupo, Mary L. Berbee.
Validation: Brandon Landry, Oldriska Ceska, Mary L. Berbee.
Visualization: Brandon Landry, Mary L. Berbee.
Writing – original draft: Brandon Landry.
Writing – review & editing: Brandon Landry, Jeannette Whitton, Anna L. Bazzicalupo, Old-
riska Ceska, Mary L. Berbee.
References
1. Enjalbert F, Cassanas G, Rapior S, Renault C, Chaumont JP. Amatoxins in wood-rotting Galerina mar-
ginata. Mycologia. 2004; 96(4):720–9. https://doi.org/10.1080/15572536.2005.11832920 PMID:
21148893
2. Horak E. Addenda ad Galerinam. 1. Galerina robertii sp. n., eine neue Art aus den Franzo
¨sischen
Alpen. Zeitschrift fu¨r Mykologie. 1994; 60:85–90.
3. Kaneko H, Tomomasa T, Inoue Y, Kunimoto F, Fukusato T, Muraoka S, et al. Amatoxin poisoning from
ingestion of Japanese Galerina mushrooms. Journal of Toxicology-Clinical Toxicology. 2001; 39
(4):413–6 https://doi.org/10.1081/clt-100105164 PMID: 11527238
4. Hulting J, Green K, Persson H. Amanitin poisoning from Galerina marginata—clinical and laboratory
findings in the 1st Swedish case. Journal of Toxicology-Clinical Toxicology. 1985; 23(4–6):428–9
5. Hyde KD, Norphanphoun C, Abreu VP, Bazzicalupo A, Chethana KWT, Clericuzio M, et al. Fungal
diversity notes 603–708: taxonomic and phylogenetic notes on genera and species. Fungal Diversity.
2017; 87(1):1–135.
6. Beug MW. NAMA Toxicology Committee Report for 2011. McIlvainea. 2011;21(May 6, 2017).
7. Berger KJ, Guss DA. Mycotoxins revisited: Part I. Journal of Emergency Medicine. 2005; 28(1):53–62.
https://doi.org/10.1016/j.jemermed.2004.08.013 PMID: 15657006
8. Ward J, Kapadia K, Brush E, Salhanick SD. Amatoxin poisoning: case reports and review of current
therapies. Journal of Emergency Medicine. 2013; 44(1):116–21. https://doi.org/10.1016/j.jemermed.
2012.02.020 PMID: 22555054
9. Grossman CM, Malbin B. Mushroom poisoning: a review of the literature and report of two cases
caused by a previously undescribed species. Ann Intern Med. 1954; 40(2):249–59. Epub 1954/02/01.
https://doi.org/10.7326/0003-4819-40-2-249 PMID: 13125209.
10. Tyler VE, Smith AH. Chromatographic detection of Amanita toxins in Galerina venenata. Mycologia.
1963; 55(3):358–9. https://doi.org/10.2307/3756526
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 17 / 19
11. Gulden G, Hallgrı
´msson H. The genera Galerina Earle and Phaeogalera Ku¨hner (Basidiomycetes,
Agaricales) in Iceland. Acta Botanica Islandica. 2000; 13:3–54.
12. Smith AH, Singer R. A Monograph on the Genus Galerina Earle. New York: Hafner; 1964. 384 p.
13. Gulden G, Stensrud O, Shalchian-Tabrizi K, Kauserud H. Galerina Earle: A polyphyletic genus in the
consortium of dark-spored agarics. Mycologia. 2005; 97(4):823–37. https://doi.org/10.3852/mycologia.
97.4.823 PMID: 16457352
14. Muraoka S, Fukamachi N, Mizumoto K, Shinozawa T. Detection and identification of amanitins in the
wood-rotting fungi Galerina fasciculata and Galerina helvoliceps. Applied and Environmental Microbiol-
ogy. 1999; 65(9):4207–10 https://doi.org/10.1128/AEM.65.9.4207-4210.1999 PMID: 10473435
15. Muraoka S, Shinozawa T. Effective production of amanitins by two-step cultivation of the basidiomy-
cete, Galerina fasciculata GF-060. Journal of Bioscience and Bioengineering. 2000; 89(1):73–6. https://
doi.org/10.1016/s1389-1723(00)88053-6 PMID: 16232701
16. Besl H, Mack P, Schmid-Heckel H. Giftpilze in den Gattungen Galerina und Lepiota. Zeitschrift fu¨r
Mykologie. 1984; 50(2):183–92.
17. Luo H, Hallen-Adams HE, Scott-Craig JS, Walton JD. Ribosomal biosynthesis of alpha-amanitin in
Galerina marginata. Fungal Genetics and Biology. 2012; 49(2):123–9. https://doi.org/10.1016/j.fgb.
2011.12.005 PMID: 22202811
18. Gulden G, Dunham S, Stockman J. DNA studies in the Galerina marginata complex. Mycological
Research. 2001; 105:432–40. https://doi.org/10.1017/s0953756201003707
19. Matheny PB, Moreau PA, Vizzini A, Harrower E, De Haan A, Contu M, et al. Crassisporium and Romag-
nesiella: two new genera of dark-spored Agaricales. Systematics and Biodiversity. 2015; 13(1):28–41.
https://doi.org/10.1080/14772000.2014.967823
20. Cai Q, Tulloss RE, Tang LP, Tolgor B, Zhang P, Chen ZH, et al. Multi-locus phylogeny of lethal amani-
tas: Implications for species diversity and historical biogeography. BMC Evolutionary Biology. 2014;
14(1):143. https://doi.org/10.1186/1471-2148-14-143 PMID: 24950598
21. Hallen HE, Adams GC, Eicker A. Amatoxins and phallotoxins in indigenous and introduced South Afri-
can Amanita species. South African Journal of Botany. 2002; 68(3):322–6. https://doi.org/10.1016/
s0254-6299(15)30393-8
22. Bazzicalupo AL, Whitton J, Berbee ML. Over the hills, but how far away? Estimates of mushroom geo-
graphic range extents. Journal of Biogeography. 2019; 46(7):1547–57. https://doi.org/10.1111/jbi.
13617
23. Vilgalys R, Hester M. Rapid genetic identification and mapping of enzymatically amplified ribosomal
DNA from several Cryptococcus species. Journal of Bacteriology. 1990; 172(8):4238–46 https://doi.
org/10.1128/jb.172.8.4238-4246.1990 PMID: 2376561
24. Matheny PB. Improving phylogenetic inference of mushrooms with RPB1 and RPB2 nucleotide
sequences (Inocybe; Agaricales). Molecular Phylogenetics and Evolution. 2005; 35(1):1–20 https://doi.
org/10.1016/j.ympev.2004.11.014 PMID: 15737578
25. Katoh K, Misawa K, Kuma K, Miyata T. MAFFT: a novel method for rapid multiple sequence alignment
based on fast Fourier transform. Nucleic Acids Research. 2002; 30(14):3059–66. https://doi.org/10.
1093/nar/gkf436 PMID: 12136088
26. Maddison W, Maddison D. Mesquite: a modular system for evolutionary analysis. 3.51 ed2018.
27. Darriba D, Taboada GL, Doallo R, Posada D. jModelTest 2: more models, new heuristics and parallel
computing. Nature Methods. 2012; 9(8):772-. https://doi.org/10.1038/nmeth.2109 PMID: 22847109
28. Miller MA, Pfeiffer W, Schwartz T, editors. Creating the CIPRES Science Gateway for inference of large
phylogenetic trees. Proceedings of the Gateway Computing Environments Workshop (GCE); 2010;
New Orleans, LA.
29. Stamatakis A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phyloge-
nies. Bioinformatics. 2014; 30(9):1312–3. https://doi.org/10.1093/bioinformatics/btu033 PMID:
24451623
30. Sgambelluri RM, Epis S, Sassera D, Luo H, Angelos ER, Walton JD. Profiling of amatoxins and phallo-
toxins in the genus Lepiota by liquid chromatography combined with UV absorbance and mass spec-
trometry. Toxins. 2014; 6(8):2336–47. https://doi.org/10.3390/toxins6082336 PMID: 25098279
31. Puillandre N, Lambert A, Brouillet S, Achaz G. ABGD, Automatic Barcode Gap Discovery for primary
species delimitation. Molecular Ecology. 2012; 21(8):1864–77. https://doi.org/10.1111/j.1365-294X.
2011.05239.x PMID: 21883587
32. De Queiroz K. Species concepts and species delimitation. Systematic Biology. 2007; 56(6):879–86.
https://doi.org/10.1080/10635150701701083 PMID: 18027281
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 18 / 19
33. Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, et al. Phylogenetic species rec-
ognition and species concepts in fungi. Fungal Genetics and Biology. 2000; 31(1):21–32. https://doi.
org/10.1006/fgbi.2000.1228 PMID: 11118132
34. Bazzicalupo AL, Buyck B, Saar I, Vauras J, Carmean D, Berbee ML. Troubles with mycorrhizal mush-
room identification where morphological differentiation lags behind barcode sequence divergence.
Taxon. 2017; 66(4):791–810. https://doi.org/10.12705/664.1
35. Walton JD. The Cyclic Peptide Toxins of Amanita and Other Poisonous Mushrooms. Cham, Switzer-
land: Springer International Publishing AG; 2018. 240 p.
36. Besl H. Amatoxine im Gewa
¨chshaus: Galerina sulciceps, ein tropischer Giftpilz. Zeitschrift fu¨r Mykolo-
gie. 1981; 47(2):253–6.
37. Bresinsky A. "Galerina beinrothii" nov. spec., "Panaeolus uliginosus" J. Schaeff. und andere Agaricales
aus Flachmooren Oberbayerns. Zeitschrift fu¨r Pilzkunde: ZfP. 1966; 32(1/2):8–17.
38. Hongo T. Notes on Japanese larger fungi (21). Journal of Japanese Botany. 1974; 49:294–305.
39. Riley R, Salamov AA, Brown DW, Nagy LG, Floudas D, Held BW, et al. Extensive sampling of basidio-
mycete genomes demonstrates inadequacy of the white-rot/brown-rot paradigm for wood decay fungi.
Proceedings of the National Academy of Sciences. 2014; 111(27):9923. https://doi.org/10.1073/pnas.
1400592111 PMID: 24958869
40. Hallen HE, Luo H, Scott-Craig JS, Walton JD. Gene family encoding the major toxins of lethal Amanita
mushrooms. Proceedings of the National Academy of Sciences of the United States of America. 2007;
104(48):19097–101. https://doi.org/10.1073/pnas.0707340104 PMID: 18025465
41. Pulman JA, Childs KL, Sgambelluri RM, Walton JD. Expansion and diversification of the MSDIN family
of cyclic peptide genes in the poisonous agarics Amanita phalloides and A.bisporigera. Bmc Genomics.
2016; 17. https://doi.org/10.1186/s12864-016-3378-7 PMID: 27978833
42. Kume K, Ikeda M, Miura S, Ito K, Sato KA, Ohmori Y, et al. Alpha-amanitin restrains cancer relapse
from drug-tolerant cell subpopulations via TAF15. Scientific Reports. 2016; 6. https://doi.org/10.1038/
srep25895 PMID: 27181033
43. Moldenhauer G, Salnikov AV, Luttgau S, Herr I, Anderl J, Faulstich H. Therapeutic potentialof amani-
tin-conjugated anti-epithelial cell adhesion molecule monoclonal antibody against pancreatic carci-
noma. Journal of the National Cancer Institute. 2012; 104(8):622–34. https://doi.org/10.1093/jnci/
djs140 PMID: 22457476
44. Moshnikova A, Moshnikova V, Andreev OA, Reshetnyak YK. Antiproliferative effect of pHLIP-amanitin.
Biochemistry. 2013; 52(7):1171–8. https://doi.org/10.1021/bi301647y PMID: 23360641
45. Sgambelluri RM, Smith MO, Walton JD. Versatility of Prolyl Oligopeptidase B in Peptide Macrocycliza-
tion. Acs Synthetic Biology. 2018; 7(1):145–52. https://doi.org/10.1021/acssynbio.7b00264 PMID:
28866879
46. Gulden G. Galerina Earle. In: Knudsen H, Vesterholt J, editors. Funga Nordica Agaricoid, boletoid, cla-
varioid, cyphelloid and gastroid genera. 2. Copenhagen, Denmark: Nordsvamp; 2012. p. 886–906.
47. Rees BJ, Orlovich DA, Marks PBD. Treading the fine line between small-statured Gymnopilus and
excentrically stipitate Galerina species in Australia. Mycological Research. 1999; 103:427–42. https://
doi.org/10.1017/s095375629800745x
PLOS ONE
Phylogeny of deadly amatoxins among the little brown mushrooms
PLOS ONE | https://doi.org/10.1371/journal.pone.0246575 February 10, 2021 19 / 19
... The taxonomic classification of Galerina Earle (Agaricales) can be traced back to Fries (1821), who initially designated it as the tribe Galera, comprising "Mycena-like ocher-brown spored fungi". Galerina was traditionally classified in the Hymenogastraceae Vittad., and species of the genus are characterized by small to medium-sized basidiomata with yellow-orange or yellowbrown pileus colour, slender and delicate stipe, the distinctive presence of verrucose basidiospores with rusty brown to brown and typically the presence of cheilocystidia (Ammirati et al., 1985;Gulden, 2012;Landry et al., 2021). Originally, it was assigned to the Cortinariaceae Singer due to the ocher brown and tuberculate nature of its basidiospores (Kirk et al., 2001), but subsequent reevaluation placed it within the Strophariaceae Singer & A.H. Sm. (Kirk et al., 2008). ...
... A recent study reported that Galerina contains lethal amatoxins at levels comparable to those in Amanita phalloides (Vaill. ex Fr.) Link (Landry et al., 2021 (Landry et al., 2021). Amatoxins, αand β-amanitin, which are toxic peptides found in Amanita, were also detected in G. venenata (Tyler and Smith, 1963). ...
... A recent study reported that Galerina contains lethal amatoxins at levels comparable to those in Amanita phalloides (Vaill. ex Fr.) Link (Landry et al., 2021 (Landry et al., 2021). Amatoxins, αand β-amanitin, which are toxic peptides found in Amanita, were also detected in G. venenata (Tyler and Smith, 1963). ...
Article
Full-text available
A deadly poisonous fungus, Galerina venenata collected from rhizospheric region of Olea europaea, which is distributed in the area with dominance of Mediterranean climate, is reported for the first time for Türkiye. Molecular phylogenetic analysis based on ITS rDNA sequences from Turkish collections confirmed the position of G. venenata in the genus, and being the closest relative of G. marginata. A detailed morphological description of the present species, photographs, and comparison with taxonomically and phylogenetically close species are provided.
... This genus of basidiomycetous fungi encompasses ∼300 species worldwide (Horak 1994, Gulden et al. 2005, which form relatively small, yellowish to reddish-brown fruiting bodies with campanulate, convex to flat pilei and slender stipes. Several Galerina species are well known for posing a poisoning risk due to the production of deadly amatoxins (Landry et al. 2021). The genus shows a broad distribution in Mediterranean, temperate and boreal regions in the Northern Hemisphere (GBIF 2022), where saprotrophic species generally grow on dead parts of bryophytes in peat bogs or are associated with woody remnants or other plant debris in forests, on which this genus degrades wood cell wall components (Gulden et al. 2005, Grzesiak & Wolski 2015, Kohler et al. 2015. ...
... The present study validated by means of molecular phylogenetics the existence in Antarctica of populations of G. marginata, G. badipes and G. fallax. The former species (and probably G. badipes too) is well known for producing amatoxins, which can have dramatic consequences for human ingestion (Landry et al. 2021). The samples of Galerina collected from Amsler Island, Antarctica, were also found to contain alpha-amanitin (unpublished data 2013, analyses completed by Jonathan Walton, University of Michigan). ...
Article
Full-text available
Fungi are probably the most diverse group of eukaryotic organisms in the Antarctic continent and nearby archipelagos, and they dominate communities in either mild or harsh habitats. However, our knowledge of their global distribution ranges and the temporal origins of their Antarctic populations is rather limited or almost absent, especially for species that do not lichenize. We focused for the first time on elucidating the taxonomic identity and phylogenetic relationships of several Antarctic collections of the deadly fungal Basidiomycota genus Galerina . By using molecular sequence data from the universal fungal barcode and a dataset encompassing 178 specimens, the inferred phylogeny showed that the Antarctic specimens corresponded with the sub-cosmopolitan species Galerina marginata , Galerina badipes and Galerina fallax , and their most closely related intraspecific genetic lineages were from northern Europe and North America. We found that these species probably host Antarctic-endemic intraspecific lineages. Furthermore, our dating analyses indicated that their Antarctic populations originated in the Pleistocene, a temporal frame that agrees with that proposed for the Antarctic colonization of plants such as the grass Deschampsia antarctica , mosses and some amphitropical lichens. Altogether, these findings converge on the same temporal scenario for the assembly of the most conspicuous terrestrial Antarctic plant and fungal communities.
... Within the genus, species grouped in the section Phalloideae are known to produce amatoxins and are the leading cause of fatal intoxication worldwide[28,85,86]. Nonetheless, some species of at least two other genera, Galerina and Lepiota, produce amatoxins and can cause hepatotoxicity comparable to the best-known species of the genus Amanita[87,88]. The toxins produced by these fungi are numerous and peptidic. They are essentially grouped into three chemical families: amatoxins, phallotoxins and virotoxins, all cyclopeptides with a tryptophan sulfur bond and some hydroxylated amino acids. ...
Article
Fungi are often considered a delicacy and are primarily cultivated and harvested, although numerous species are responsible for intoxication due to toxin content. Foodborne diseases are a significant public health concern, causing approximately 420 000 deaths and 600 million morbidities yearly, of which mushroom poisoning is one of the leading causes. Epidemiological data on non-cultivated mushroom poisoning in individual countries are often unrepresentative, as intoxication rarely requires emergency intervention. On the other hand, the lack of specialist knowledge among medical personnel about the toxicological manifestations of mushroom consumption may result in ineffective therapeutic interventions. This work aims to provide an easy-to-consult and wide-ranging tool useful for better understanding the variability of mushroom intoxications, the associated symptoms, and the main treatments for the most severe cases, given the absence of a complete species mapping tool toxic. Moreover, we establish an effective collection network that describes the incidence of mushroom poisonings by reporting the species and associated toxicological manifestations for each case. In conclusion, we highlight the need to establish appropriate primary prevention interventions, such as training the affected population and increasing consultancy relationships between mycological experts and specialised healthcare personnel. https://academic.oup.com/mmy/article-abstract/doi/10.1093/mmy/myae033/7640032?utm_source=etoc&utm_campaign=mmy&utm_medium=email
... Within the genus, species grouped in the section Phalloideae are known to produce amatoxins and are the leading cause of fatal intoxication worldwide[28,85,86]. Nonetheless, some species of at least two other genera, Galerina and Lepiota, produce amatoxins and can cause hepatotoxicity comparable to the best-known species of the genus Amanita[87,88]. The toxins produced by these fungi are numerous and peptidic. They are essentially grouped into three chemical families: amatoxins, phallotoxins and virotoxins, all cyclopeptides with a tryptophan sulfur bond and some hydroxylated amino acids. ...
Article
https://academic.oup.com/mmy/advance-article-abstract/doi/10.1093/mmy/myae033/7640032?redirectedFrom=fulltext Fungi are often considered a delicacy and are primarily cultivated and harvested, although numerous species are responsible for intoxication due to toxin content. Foodborne diseases are a significant public health concern, causing approximately 420,000 deaths and 600 million morbidities yearly , of which mushroom poisoning is one of the leading causes. Epidemiological data on non-cultivated mushroom poisoning in individual countries are often unrepresentative, as intoxication rarely requires emergency intervention. On the other hand, the lack of specialist knowledge among medical personnel about the toxicological manifestations of mushroom consumption may result in ineffective therapeutic interventions. This work aims to provide an easy-to-consult and wide-ranging tool useful for better understanding the variability of mushroom intoxications, the associated symptoms, and the main treatments for the most severe cases, given the absence of a complete species mapping tool toxic. Moreover, we establish an effective collection network that describes the incidence of mushroom poisonings by reporting the species and associated toxicological manifestation for each case. In conclusion, we highlight the need to establish appropriate primary prevention interventions, such as training the affected population and increasing consultancy relationships between mycological experts and specialised healthcare personnel. Lay Summary We propose a review of the literature that describes the main syndromes resulting from the consumption of toxic fungal species, reporting symptoms and clinical manifestations, latency times and, where possible, diagnostic tools for recognising the species involved and interventions to be carried out.
... Mart ın (Enjalbert et al., 2002;Sgambelluri et al., 2014;Diaz, 2018;Landry et al., 2021). After eating an amanitin-producing mushroom, symptoms develop in four periods. ...
Article
Full-text available
Mushrooms have long been considered as delicacies as well as used as important dietary supplements and food. However, there are major concerns with poisonous mushrooms as these pose threats to public health and safety. In this paper, we provide a review focusing on poisonous mushrooms, their toxins, symptoms and utilizations. In addition, this paper establishes a poisonous mushroom list which includes 643 species from two phyla, 16 orders, 51 families and 148 genera. The toxicity of all these species was verified and 332 species were ranked as P1 signifying that these species have toxic studies and or clinical poisoning case records and 311 species were P2 meaning they had previously been recorded as poisonous in other studies. Furthermore, we discuss advances in technology including how genomic studies could be used as a breakthrough tool in the field of toxic mushrooms. With this comprehensive review, we aim to promote public awareness of poisonous mushrooms, including how to avoid mushroom poisoning, and how to better utilize poisonous mushroom resources.
... The mentioned species, unlike Macrolepiota procera, are poisonous fungi due to the content of amatoxins. There is a large differentiation of amatoxins within one genus of mushrooms, which is demonstrated by phylogenetic analysis carried out in close connection with chemical tests [24]. Amatoxins (amanitotoxins) are cyclic octapeptides containing sulfoxide and indole groups. ...
Article
Full-text available
Parasol mushroom (Macrolepiota procera) is a fungus that is often included in the menu of people looking for replacements for meat products and at the same time appreciating mushrooms. Its fruiting bodies are known for their delicate flavor and aroma. The aim of the publication was to analyze the latest information (mainly from 2015 to 2021) on the chemical composition of the M. procera fruiting bodies and their antioxidant properties. The data on other health-promoting properties and the possibilities of using these mushrooms in medicine were also compiled and summarized, taking into account their antibacterial, antioxidant, anti-inflammatory, regulatory, antidepressant, and anticancer effects. Moreover, the influence of various forms of processing and conservation of raw mushroom on its health-promoting properties was discussed. The possibilities of controlling the quality of both the raw material and the prepared dishes were also discussed. Such an opportunity is offered by the possibility of modifying the growing conditions, in particular, the appropriate selection of the substrate for mushroom cultivation and the deliberate enrichment of its composition with the selected substances, which will then be incorporated into the fungus organism.
Article
Amanitin (AMA) is responsible for human fatalities after ingestion of poisonous mushrooms, thus, a rapid and accurate detection method is urgently needed. Here, gold nanoparticle-based immunosensor with monoclonal antibody against AMA was constructed for rapid detection of alpha- and beta-amanitin (α- and β-AMA) in mushroom, serum and urine samples. Under optimized conditions, the visual limits of detection (vLOD) and calculated LOD for α-AMA and β-AMA in mushroom were 10 ng/g, 20 ng/g, and 0.1 ng/g, 0.2 ng/g, respectively. Analysis of wild mushroom samples was also performed using a strip scan reader in the 10 min range. Furthermore, in mushrooms containing amatoxins results were confirmed and compared with those determined by liquid chromatography tandem mass spectrometry. Thus, this immunosensor provided a useful monitoring tool for rapid detection and screening of mushroom samples and in serum and urine from subjects who accidentally consumed AMA-containing mushrooms.
Article
Full-text available
Second ProofHeading (revised on May 26th, 2022). In Europe, amatoxin-containing mushrooms are responsible for most of the deadly poisonings caused by macrofungi. The present work presents a multidisciplinary revision of the European species of Amanita sect. Phalloideae based on morphology, phylogeny, epidemiology, and biochemistry of amatoxins and phallotoxins. Five distinct species of this section have been identified in Europe to date: A. phalloides, A. virosa, A. verna, the recently introduced North American species A. amerivirosa, and A. vidua sp. nov., which is a new name proposed for the KOH-negative Mediterranean species previously described as A. verna or A. decipiens by various authors. Epitypes or neotypes are selected for species lacking suitable reference collections, namely A. verna and A. virosa. Three additional taxa, Amanita decipiens, A. porrinensis, and A. virosa var. levipes are here considered later heterotypic synonyms of A. verna, A. phalloides, and A. amerivirosa, respectively. in Special Issue Molecular Characterization of Fungi. Keywords: Agaricales; amatoxins; phallotoxins; epidemiology; toxicology
Article
Full-text available
This is the sixth in a series of papers where we bring collaborating mycologists together to produce a set of notes of several taxa of fungi. In this study we introduce a new family Fuscostagonosporaceae in Dothideomycetes. We also introduce the new ascomycete genera Acericola, Castellaniomyces, Dictyosporina and Longitudinalis and new species Acericola italica, Alternariaster trigonosporus, Amarenomyces dactylidis, Angustimassarina coryli, Astrocystis bambusicola, Castellaniomyces rosae, Chaetothyrina artocarpi, Chlamydotubeufia krabiensis, Colletotrichum lauri, Collodiscula chiangraiensis, Curvularia palmicola, Cytospora mali-sylvestris, Dictyocheirospora cheirospora, Dictyosporina ferruginea, Dothiora coronillae, Dothiora spartii, Dyfrolomyces phetchaburiensis, Epicoccum cedri, Epicoccum pruni, Fasciatispora calami, Fuscostagonospora cytisi, Grandibotrys hyalinus, Hermatomyces nabanheensis, Hongkongmyces thailandica, Hysterium rhizophorae, Jahnula guttulaspora, Kirschsteiniothelia rostrata, Koorchalomella salmonispora, Longitudinalis nabanheensis, Lophium zalerioides, Magnibotryascoma mali, Meliola clerodendri-infortunati, Microthyrium chinense, Neodidymelliopsis moricola, Neophaeocryptopus spartii, Nigrograna thymi, Ophiocordyceps cossidarum, Ophiocordyceps issidarum, Ophiosimulans plantaginis, Otidea pruinosa, Otidea stipitata, Paucispora kunmingense, Phaeoisaria microspora, Pleurothecium floriforme, Poaceascoma halophila, Periconia aquatica, Periconia submersa, Phaeosphaeria acaciae, Phaeopoacea muriformis, Pseudopithomyces kunmingnensis, Ramgea ozimecii, Sardiniella celtidis, Seimatosporium italicum, Setoseptoria scirpi, Torula gaodangensis and Vamsapriya breviconidiophora. We also provide an amended account of Rhytidhysteron to include apothecial ascomata and a J+ hymenium. The type species of Ascotrichella hawksworthii (Xylariales genera incertae sedis), Biciliopsis leptogiicola (Sordariomycetes genera incertae sedis), Brooksia tropicalis (Micropeltidaceae), Bryochiton monascus (Teratosphaeriaceae), Bryomyces scapaniae (Pseudoperisporiaceae), Buelliella minimula (Dothideomycetes genera incertae sedis), Carinispora nypae (Pseudoastrosphaeriellaceae), Cocciscia hammeri (Verrucariaceae), Endoxylina astroidea (Diatrypaceae), Exserohilum turcicum (Pleosporaceae), Immotthia hypoxylon (Roussoellaceae), Licopolia franciscana (Vizellaceae), Murispora rubicunda (Amniculicolaceae) and Doratospora guianensis (synonymized under Rizalia guianensis, Trichosphaeriaceae) were re-examined and descriptions, illustrations and discussion on their familial placement are given based on phylogeny and morphological data. New host records or new country reports are provided for Chlamydotubeufia huaikangplaensis, Colletotrichum fioriniae, Diaporthe subclavata, Diatrypella vulgaris, Immersidiscosia eucalypti, Leptoxyphium glochidion, Stemphylium vesicarium, Tetraploa yakushimensis and Xepicula leucotricha. Diaporthe baccae is synonymized under Diaporthe rhusicola. A reference specimen is provided for Periconia minutissima. Updated phylogenetic trees are provided for most families and genera. We introduce the new basidiomycete species Agaricus purpurlesquameus, Agaricus rufusfibrillosus, Lactifluus holophyllus, Lactifluus luteolamellatus, Lactifluus pseudohygrophoroides, Russula benwooii, Russula hypofragilis, Russula obscurozelleri, Russula parapallens, Russula phoenicea, Russula pseudopelargonia, Russula pseudotsugarum, Russula rhodocephala, Russula salishensis, Steccherinum amapaense, Tephrocybella constrictospora, Tyromyces amazonicus and Tyromyces angulatus and provide updated trees to the genera. We also introduce Mortierella formicae in Mortierellales, Mucoromycota and provide an updated phylogenetic tree.
Article
Full-text available
Species of Russula (Russulaceae), a large, cosmopolitan, ectomycorrhizal fungal genus are notoriously difficult to identify. To delimit species and to evaluate their morphology, we sequenced the ~400 bp ITS2 ribosomal DNA region from 713 Pacific Northwest Russula specimens from Benjamin Woo’s exceptional collection. As a topological constraint for analysis of the ITS2, we sequenced and inferred a phylogeny from the ITS, LSU, RPB2 and EF1-α regions from 50 European and North American specimens of major clades in Russula. We delimited 72 candidate species from Woo’s collection’s ITS2 sequences using ABGD, GMYC, PTP, and mothur software. To guide application of names, we sequenced a ~200 bp portion of the ITS from 18 American type specimens. Of the 72 delimited species, 28 matched a type or a currently barcoded European species. Among the remaining 44 are poorly known or undescribed species. We tested the congruence of morphology with delimitations for 23 species represented by 10 or more specimens each. No morphological character alone was consistently diagnostic across all specimens of any of the 23 candidate species. Ordination of combined field characters followed by pairwise multivariate analysis of variance showed that centroids were significantly different in 221 of 253 species pair comparisons. Ordination also showed that specimens from the same species were widely dispersed, overlapping with specimens from other species. This explains why only 48.5% of specimens were correctly assigned to their species in a canonical variates analysis of combined field and spore characters. Based on sequence comparisons, we contribute to correcting the broad and confusing misapplications of European names that have long obscured patterns of Russula’s geographical distribution and diversification. Our evidence suggests that morphology in Russula diverges slowly, and that phenotypic plasticity, convergence, or retention of ancestral polymorphisms blur the distinctions among recently derived species.
Article
Full-text available
The basidiomycete genus Galerina Earle accommodates more than 300 small brown-spored agarics worldwide, predominantly described from the Northern hemisphere. The delimitation of species and infrageneric units hitherto has been based on morphological and, to some extent, ecological characters. In this study we have analyzed nuclear ribosomal LSU and ITS sequences to reveal infrageneric phylogeny and the phylogenetic placement of Galerina among the dark-spored agarics. Sequences from 36 northern hemisphere Galerina species and 19 other dark-spored taxa were analyzed, some of them obtained from EMBL/GenBank. Our results, received from Bayesian and distance methods, strongly suggest that Galerina is a polyphyletic genus. The LSU analysis shows that Galerina is composed of three or four separate monophyletic main groups. In addition, a few species cluster together with other dark-spored agarics. The same groups are recognized in the ITS tree and they correspond roughly to previously recognized subgenera or sections in Galerina. With high support our LSU analysis suggests that Gymnopilus is a monophyletic genus and that Gymnopilus and one of the Galerina lineages (“mycenopsis”) are sister groups. The analyses further indicate that the Galerina lineages, as well as the genus Gymnopilus, could be referred to a strongly emendated family Strophariaceae, which corresponds largely to the family as circumscribed by Kühner (1980) ———. 1980. Les Hyménomycètes agaricoïdes (Agaricales, Tricholomatales, Pluteales, Russulales). Etude générale et classification. Bull Soc Linn Lyon 49 Numéro special 1027 p. . Our results affirm that morphological characters often are highly homoplastic in the agarics. At the present stage formal taxonomic consequences or nomenclatural changes are not proposed.
Article
Full-text available
Amatoxins, bicyclic octapeptide derivatives responsible for severe hepatic failure, are present in several Basidiomycota species belonging to four genera, i.e. Amanita, Conocybe, Galerina and Lepiota. DNA studies for G. autumnalis, G. marginata, G. oregonensis, G. unicolor and G. venenata (section Naucoriopsis) determined that these species are the same, supporting the concept of Galerina marginata complex. These mostly lignicolous species are designated as white-rot fungi having a broad host range and capable of degrading both hardwoods and softwoods. Twenty-seven G. marginata basidiomes taken from different sites and hosts (three sets) as well as 17 A. phalloides specimens (three sets) were collected in French locations. The 44 basidiomes were examined for amatoxins and phallotoxins using high-performance liquid chromatography. Toxinological data for the wood-rotting G. marginata and the ectomycorrhizal A. phalloides species were compared and statistically analyzed. The acidic and neutral phallotoxins were not detected in any G. marginata specimen, whereas the acidic (β-Ama) and neutral (α-Ama and γ-Ama) amanitins were found in all basidiomes from either Angiosperms or Gymnosperms hosts. The G. marginata amatoxin content varied from 78.17 to 243.61 μg.mg⁻¹ of fresh weight and was elevated significantly in one set out of three. The amanitin amounts from certain Galerina specimens were higher than those from some A. phalloides basidiomes. Relationship between the amanitin distribution and the chemical composition of substrate was underlined and statistically validated for the white-rot G. marginata. Changes in nutritional components from decayed host due to enzymatic systems and genetic factors as well as environmental conditions seem to play a determinant role in the amanitin profile. Variability noticed in the amanitin distribution for the white-rot G. marginata basidiomes was not observed for the ectomycorrhizal A. phalloides specimens.
Article
Full-text available
Background The cyclic peptide toxins of Amanita mushrooms, such as α-amanitin and phalloidin, are encoded by the “MSDIN” gene family and ribosomally biosynthesized. Based on partial genome sequence and PCR analysis, some members of the MSDIN family were previously identified in Amanita bisporigera, and several other members are known from other species of Amanita. However, the complete complement in any one species, and hence the genetic capacity for these fungi to make cyclic peptides, remains unknown. Results Draft genome sequences of two cyclic peptide-producing mushrooms, the “Death Cap” A. phalloides and the “Destroying Angel” A. bisporigera, were obtained. Each species has ~30 MSDIN genes, most of which are predicted to encode unknown cyclic peptides. Some MSDIN genes were duplicated in one or the other species, but only three were common to both species. A gene encoding cycloamanide B, a previously described nontoxic cyclic heptapeptide, was also present in A. phalloides, but genes for antamanide and cycloamanides A, C, and D were not. In A. bisporigera, RNA expression was observed for 20 of the MSDIN family members. Based on their predicted sequences, novel cyclic peptides were searched for by LC/MS/MS in extracts of A. phalloides. The presence of two cyclic peptides, named cycloamanides E and F with structures cyclo(SFFFPVP) and cyclo(IVGILGLP), was thereby demonstrated. Of the MSDIN genes reported earlier from another specimen of A. bisporigera, 9 of 14 were not found in the current genome assembly. Differences between previous and current results for the complement of MSDIN genes and cyclic peptides in the two fungi probably represents natural variation among geographically dispersed isolates of A. phalloides and among the members of the poorly defined A. bisporigera species complex. Both A. phalloides and A. bisporigera contain two prolyl oligopeptidase genes, one of which (POPB) is probably dedicated to cyclic peptide biosynthesis as it is in Galerina marginata. Conclusion The MSDIN gene family has expanded and diverged rapidly in Amanita section Phalloideae. Together, A. bisporigera and A. phalloides are predicted to have the capacity to make more than 50 cyclic hexa-, hepta-, octa-, nona- and decapeptides. Electronic supplementary material The online version of this article (doi:10.1186/s12864-016-3378-7) contains supplementary material, which is available to authorized users.
Article
Aim Geographic distributions of mushroom species remain poorly understood despite their importance for advancing our understanding of the habitat requirements, species interactions and ecosystem functions of this key group of organisms. Here, we estimate geographic range extents (maximum within‐species geographic distance) of genetically defined operational taxonomic units (OTUs). Location World‐wide, with emphasis on the American Pacific Northwest. Taxa Amanita, Agaricus, Cortinarius, Galerina, Hebeloma, Hydnum, Hygrocybe, Hygrophorus, Inocybe, Lepiota, Pholiota, and Russula + Lactarius; other genera in Agaricomycotina. Method We used publicly available OTUs from ribosomal internal transcribed spacer (ITS) sequences (n = 15,373) from 12 mushroom genera with worldwide distributions. For each of 2,324 ~ species‐level OTUs, we estimated the maximum within‐species range extent based on sample locality records. In parallel, we estimated range extents for species in four tree genera. Contrasting estimates from well‐studied trees allowed us to test for potential biases in our range estimates of less well inventoried mushrooms. Results The median range extents across the 2,324 mushroom OTUs varied from ~ 1,200 to 4,039 km, depending on assumptions. These extents were significantly lower than estimates from permuted or randomized data. Mushroom ranges were comparable to the median natural range extent of tree species (1,613 km). In contrast, the tree median species range increased to 16,581 km when anthropogenic range extensions were included. At least 10 mushroom species were similarly broadly distributed, eight of which have been associated with human activity. Main conclusions Overall, like tree species, mycorrhizal and saprotrophic fungi show evidence of biogeographic structure rather than global distributions. This reconstruction of geographic range extents drew upon investments into ITS barcoding of extensive herbarium collections. Large scale analyses such as ours can yield estimates of fungal geographic range extents that are a prerequisite to a deeper understanding of the diverse roles of fungi in ecosystems.
Book
Poisonous mushrooms have fascinated scientists and laypersons alike for thousands of years. Almost all mushroom fatalities are due to the genus Amanita, whose poetic common names (death cap, destroying angel) attest to their lethality. In his classic 1986 book, Theodor Wieland covered the state of our knowledge about the chemistry and biochemistry of the toxins of Amanita mushrooms up until that time, with a particular focus on the decades of chemical research by him and the Wieland dynasty (including his father, brother, brother-in-law, and cousin). Wieland’s book is now mainly of historical interest, with its exhaustive overview of the early chemical studies done without benefit of methods taken for granted by modern chemists. This book is a complete top-to-bottom revision of Wieland’s 1986 book. The material covers history, chemistry, and biology with equal thoroughness. It should be of interest to natural products chemists and biologists, professional and amateur mycologists, and toxicologists. The three scientific fields that are most relevant to the book are natural products chemistry, mycology, and fungal molecular genetics. Dr. Walton is an expert in all three. To maximize the broad utility and appeal of the book, care has been taken to define all technical terms specific to a particular discipline, so that, for example, mycologists will be able to understand the relevant chemistry, and chemists will be able to understand the relevant fungal biology.
Article
Cyclic peptides are promising compounds for new chemical biological tools and therapeutics due to their structural diversity, resistance to proteases, and membrane permeability. Amatoxins, the toxic principles of poisonous mushrooms, are biosynthesized on ribosomes as 35-mer precursor peptides which are ultimately converted to hydroxylated bicyclic octapeptides. The initial cyclization steps, catalyzed by a dedicated prolyl oligopeptidase (POPB), involves removal of the 10-amino acid leader sequence from the precursor peptide and transpeptidation to produce a monocyclic octapeptide intermediate. The utility of POPB as a general catalyst for peptide cyclization was systematically characterized using a range of precursor peptide substrates produced either in E. coli or chemically. Substrates produced in E. coli were expressed either individually or in mixtures produced by codon mutagenesis. A total of 127 novel peptide substrates were tested, of which POPB could cyclize 100. Peptides of 7 to 16 residues were cyclized at least partially. Synthetic 25mer precursor peptide substrates containing modified amino acids including D-Ala, β-Ala, N-methyl-Ala, and 4-hydroxy-Pro were also successfully cyclized. Although a phalloidin heptapeptide with all L amino acids was not cyclized, partial cyclization was seen when L-Thr at position #5 was replaced with the naturally occurring D amino acid. POPB should have broad applicability as a general catalyst for macrocyclization of peptides containing 7 to at least 16 amino acids, with an optimum of 8-9 residues.