ArticlePDF Available

Deadly mushrooms of the genus Galerina found in Antarctica colonized the continent as early as the Pleistocene

Authors:
  • university of Minnesota
  • The National Museum of Natural Sciences-CSIC

Abstract and Figures

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.
Content may be subject to copyright.
Deadly mushrooms of the genus Galerina found in Antarctica
colonized the continent as early as the Pleistocene
ISAAC GARRIDO-BENAVENT 1, ROBERT A. BLANCHETTE 2and ASUNCIÓN DE LOS RÍOS 3
1
Departament de Botànica i Geologia, Facultat de Ciències Biològiques, Universitat de València, C/Doctor Moliner 50, E-46100 Burjassot,
Valencia, Spain
2
Department of Plant Pathology, 1991 Upper Buford Circle, 495 Borlaug Hall, University of Minnesota, St. Paul, MN 55108-6030, USA
3
Department of Biogeochemistry and Microbial Ecology, National Museum of Natural Sciences (MNCN), CSIC, E-28006 Madrid, Spain
Isaac.Garrido@uv.es
Abstract: 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 rst 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 intraspecic genetic lineages were
from northern Europe and North America. We found that these species probably host Antarctic-
endemic intraspecic 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 ndings converge on the same temporal scenario for the
assembly of the most conspicuous terrestrial Antarctic plant and fungal communities.
Received 8 November 2022, accepted 29 April 2023
Key words: biogeography, dating analysis, fungal endemism, Galerina marginata, long-distance dispersal,
non-lichenized fungi
Introduction
Fungi are probably the most widespread and diverse group
of eukaryotic organisms inhabiting Antarctica, with a
known fossil record dating back to the Permian period
(White Jr & Taylor 1991, Harper et al. 2016). They are
involved in key processes in terrestrial ecosystems, such
as decomposition and symbiotic mutualism (Treseder &
Lennon 2015, Asplund & Wardle 2017), and therefore
they contribute greatly to biogeochemical cycles in
otherwise low-nutrient habitats. The number of known
species in territories south of 60°S and archipelagos at
lower latitudes, such as South Georgia, is 1500
(Øvstedal & Lewis Smith 2001,2011, Onofri et al. 2005,
Bridge et al. 2008, Bridge & Spooner 2012). Almost a
third of them associate symbiotically with eukaryotic
algae (chlorophytes) or cyanobacteria, forming
macroscopic lichen thalli. Lichens are in fact one of the
most conspicuous elements of Antarctic terrestrial
habitats, and their communities develop profusely in
maritime areas (Søchting et al. 2004, Peat et al. 2007),
and even in rocky outcrops at harsher locations in the
continent (Kappen et al. 1981, Broady & Weinstein
1998, Pérez-Ortega et al. 2012). On the other hand,
non-lichenized fungi generally remain unnoticeable
because either they are unicellular (e.g. yeasts) or they
form unseen mycelia and small reproductive structures.
Microfungi are the most abundant in Antarctic soils
(Vishniac & Hemping 1979, Ruisi et al. 2007, Arenz &
Blanchette 2011, Arenz et al. 2014), and many
Ascomycota and even some Basidiomycota have been
reported on wood brought to Antarctica (Arenz &
Blanchette 2009, Blanchette et al. 2010, Arenz et al.
2011, Held & Blanchette 2017). Additional research
using next-generation sequencing techniques has also
revealed a high diversity of fungi even in the most
unexpected habitats (Coleine et al. 2018,
Garrido-Benavent et al. 2020, Rosa et al. 2021), with
similarities in species diversity and community
composition to the Arctic (Cox et al. 2016).
Antarctic Science page 1 of 14 (2023) © The Author(s), 2023. Published by Cambridge
University Press on behalf of Antarctic Science Ltd. This is an Open Access article,
distributed under the terms of the Creative Commons Attribution licence (http://
creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution
and reproduction, provided the original article is properly cited. doi:10.1017/S0954102023000196
1
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
Macrofungi (i.e. non-lichenized species that form
relatively large fruiting bodies or 'mushrooms') are
infrequently reported in Antarctica, with a few dozen
species growing in the climatically milder sub-Antarctic,
Maritime Antarctica and occasionally in the western
Antarctic Peninsula, often occurring on large carpets of
mosses and vascular plants (Pegler et al. 1980,
Gumin
ska et al. 1994, Bridge et al. 2008, Putzke et al.
2012, Held & Blanchette 2017, Newsham et al. 2021).
Bridge & Spooner (2012) and Newsham et al. (2021)
suggested that the general absence of large land animals
and higher plants with woody components in Antarctica
is a limiting factor for the development of these fungi.
From a taxonomic and biogeographical viewpoint, the
scarcity of available collections of mushroom-forming
fungi from the Antarctic has so far impeded detailed
comparisons with species that are known from elsewhere.
More specically, the lack of genetic data has prevented
the deciphering of the most probable temporal and
spatial origins of the Antarctic populations of certain
species. In fact, despite being one of the most diverse
groups in the whole continent, there is still a general lack
of knowledge as to whether non-lichenized fungi, and
particularly the Antarctic macrofungi, form specic
populations of cosmopolitan species or constitute true
endemic species (Bridge & Spooner 2012, Arenz et al.
2014).
To provide an answer to this question, the present work
uses recent collections of fruiting bodies of Galerina
species to assess their taxonomic identity and estimate a
date for their Antarctic origin within a phylogenetic
framework. 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 at 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). In the Antarctic continent and nearby
archipelagos, the number of Galerina species reported is
11, with Galerina antarctica Singer, Galerina glebarum
(Berk.) Singer and Galerina perrara Singer originally
being described, and these are known only from these
regions (Fig. 1; Berkeley 1847, Singer & Corte 1962,
Pegler et al. 1980, Bridge et al. 2008). Based on the
inferred phylogeny, we aim to ascertain whether the
sequenced Antarctic specimens belong to geographically
restricted, species-level lineages (i.e. putative endemic
species) or conform to particular intraspecic lineages of
cosmopolitan Galerina (non-endemic species). In
lichenized fungi, Antarctic endemic species have been
shown to have a relictual, pre-Pleistocene origin, whereas
Antarctic populations of amphitropical lichens are
generally much younger, dating back from the
Pleistocene onwards (Fernández-Mendoza & Printzen
2013, Garrido-Benavent et al. 2016,2018,2021). The
temporal frame estimated with the time-calibrated
Galerina phylogeny will further help us to discern
whether their evolution in Antarctica conforms to either
of these two scenarios.
Material and methods
Fieldwork and morpho-anatomical study of fruiting bodies
Several fruiting bodies of Galerina growing in a localized
area, and therefore probably corresponding to a single
mycelium, were collected in March 2018 from
Livingston Island (South Shetland Islands) and more
specically in Punta Hannah (62°39'15.37" S,
60°36'27.44" W, 377 m above sea level), which is the
second largest island in the South Shetland Islands, a
mountainous archipelago located in Maritime
Antarctica. These fruiting bodies grew abundantly on
soil, with a profuse development of cryptogams (mosses
and the chlorophyte macroalgae Prasiola) and the
Antarctic hair grass Deschampsia antarctica Desv.
(Fig. 2). Sampling permit no. CPE-2017-3 was obtained
through the Spanish Polar Committee. Specimens were
frozen until further processing at the laboratory, where
they were observed under a Leica S8APO dissecting
microscope equipped with a Leica EC3 image capture
system. Handmade sections of lamellae were rehydrated
in distilled H
2
O to describe anatomical characteristics.
Microscopic observations were made using a Zeiss
Axioplan 2 microscope tted with 'Nomarski'
differential interference contrast, and photographs were
taken with a Zeiss AxioCam digital camera.
Microscopic measurements were made by means of the
Zeiss Axiovision 4.8 imaging system. Reported data are
averages followed by standard deviations, and the
maximum and minimum values are given in parentheses.
DNA extraction and polymerase chain reaction
amplication
The isolation of genomic DNA from a single Galerina
basidioma (pl. basidiomata; i.e. basidiomycete fruiting
bodies) was done from a piece of lamellae and using the
Speed Tools DNA Extraction Kit (Biotools, Madrid,
Spain), following the manufacturer's recommendations.
The extracted DNA was eluted in a nal volume of 60 μl
2ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
with sterile puried water (SIGMA). Sequence data of the
internal transcribed spacer of the nuclear ribosomal DNA
(the so-called fungal barcode marker; Schoch et al. 2012)
was amplied using the primer pair ITS1F-KYO2 and
ITS4-KYO2 (Toju et al. 2012). Polymerase chain
reaction (PCR) experiments were performed in a total
volume of 10 μl, containing 1 μl of reaction buffer
(Biotools
®
), 2 μl of dNTPs (1 mM), 0.5 μlofeach
primer (10 μM), 0.2 U of DNA polymerase (Biotools
®
)
and 1.5 μl of the genomic DNA elution; the nal
volume was reached by adding distilled water (SIGMA).
The following PCR temperature prole was employed:
5 min at 95°C, then 30 cycles of 30 s at 95°C, 1 min at
52°C and 1.5 min at 72°C, with a nal extension of
10 min at 72°C. The PCR experiments were visualized
on 1% agarose gel stained with PRONASAFE nucleic
acid stain solution (CONDA Laboratories). The PCR
products were puried and cleaned using the UltraClean
PCR Clean-Up Kit (MOBIO Laboratories, Inc.). Both
complementary DNA strands were sequenced at
Macrogen Europe (Spain) using the same primer set as
for the initial amplication. Electropherograms were
checked and assembled using SeqManII v.5.07
©
(DNASTAR, Inc.).
Fig. 1. Diversity and distribution of Galerina species in Antarctica based on collection data provided by Bridge et al. (2008), Arenz et al.
(2014), Krishnan et al. (2016) and Canini et al. (2020). Species for which samples have been included in the present work are in bold.
3PLEISTOCENE ANTARCTIC COLONIZATION OF THE MUSHROOM GENUS GALERINA
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
Compilation of the specimen-based nrITS dataset and
sequence alignment
The newly produced sequence was submitted to the
BLAST online tool (Altschul et al. 1990) to check for
possible PCR product contamination and to identify
and retrieve available, highly similar nrITS sequences. To
this purpose, the GenBank (http://www.ncbi.nlm.nih.gov/),
UNITE (Nilsson et al. 2019) and BOLD (Ratnasingham
& Hebert 2007) nucleotide databases were used as
references. A total of 118 sequences (97 GenBank,
13 UNITE and 8 BOLD) spanning a 97100%
similarity range were downloaded. Most were accessions
labelled with the species name Galerina marginata
(Batsch) Kühner. A closely related nrITS sequence of a
Galerina collection from Amsler Island (Antarctic
Peninsula) was included as well. This was also collected
from an area where mosses were growing under permit
ACA-2012-013. DNA extraction and sequencing were
done using methods previously described (Blanchette
et al. 2016). An additional search in public databases
was conducted to select and retrieve any available
nrITS data for other Antarctic Galerina collections. Two
sequences obtained from soil isolates were found:
MK537266, which was generated in a study by Canini
et al. (2020) from Victoria Land; and MF692967, from
King George Island (Krishnan et al. 2018). BLAST
searches against the GenBank database revealed a close
Fig. 2. Galerina marginata:a. fruiting bodies growing on a carpet of Deschampsia antarctica and Prasiola sp. in Punta Hannah
(Livingston Island), b. a detail of the fruiting body pileus, c. a basidium (i.e. basidiomycete sporangium) with developing spores and
d. spores. Scale bars: 10 μm.
4ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
match of these accessions to nrITS sequences labelled
with the species names Galerina badipes (Pers.) Kühner
and Galerina fallax A.H. Sm. & Singer. Twenty-ve
sequences hosted in GenBank belonging to these two
species were downloaded and incorporated into the
dataset as well. Finally, the dataset was completed by
including additional Galerina species based on works
by Gulden et al. (2005)andLathaet al. (2015), which
provided two of the most comprehensive Galerina
phylogenies published to date. The nal nrITS
consisted of 178 sequences. We followed Gulden et al.
(2005) in selecting adequate outgroup taxa for our
phylogenetic analyses. Although the genus Galerina
was revealed to be polyphyletic by these authors,
they considered a group of taxa referred to 'tubariopsis'
to be a suitable outgroup. In our dataset, this is
represented by the following species: Galerina arctica
(Singer) Nezdojm., Galerina clavata (Velen.) Kühner,
Galerina discreta E. Horak, Senn-Irlet, M. Curti
& Musumeci, Galerina laevis Singer, Galerina
pseudocerina A.H. Sm. & Singer and Galerina stordalii
A.H. Sm.
The program MAFFT v.7.308 (Katoh & Standley
2013) was used to generate a multiple-sequence
alignment with the following parameters: the FFT-NS-I
x1000 algorithm, the 200PAM/k= 2 scoring matrix, a
gap open penalty of 1.5 and an offset value of 0.123. The
resulting alignment was manually optimized in Geneious
v.9.0.2 to 1) trim alignment ends of longer sequences that
included part of the 18S28S ribosomal subunits, 2) replace
gaps at the ends of shorter sequences with an International
Union of Pure and Applied Chemistry (IUPAC) base
representing any base ('N') and 3) replace doubtful base
calls at the extremes with 'N'. The software GBlocks 0.91b
(Castresana 2000) was subsequently used to automatically
deal with ambiguously aligned regions, implementing the
least stringent parameters but allowing gaps in 50% of the
sequences. Alignments were deposited in FigShare
(DOI: 10.6084/m9.gshare.22219546).
Maximum-likelihood phylogenetic analyses
The online version of RAxML-HPC2 hosted at the
CIPRES Science Gateway (Stamatakis 2006, Stamatakis
et al. 2008, Miller et al. 2010) was used to estimate two
maximum-likelihood (ML) phylogenies based on the
GBlocks-trimmed (GB) and untrimmed (ORG) alignments.
This approach would allow us to evaluate the effect of
alignment uncertainty on the inferred nodal support. The
analyses used the GTRGAMMA nucleotide substitution
model for the two delimited partitions within the nrITS
(ITS1+2, 5.8S), and nodal support was evaluated by
conducting 1000 rapid bootstrap pseudoreplicates.
The resulting phylogenetic trees were visualized in FigTree
v.1. 4 ( http://tree.bio.ed.ac.uk/software/tracer/), and
Adobe Illustrator CS5 was used for artwork. Tree nodes
with bootstrap support (BS) values 70% were regarded
as signicantly supported.
Haplotype networks, DNA polymorphism and neutrality
tests
The genealogical relationships among specimens included
in the G. marginata clade were calculated under a
statistical parsimony framework in PopA RT v.1.7 (Leigh
&Bryant2015) using the method of Templeton et al.
(1992). To this purpose, a sub-alignment of 120 sequences
was extracted from the GBlocks-untrimmed, original
alignment. Because the inference of haplotype networks is
sensitive to ambiguous base calls and missing data (Joly
et al. 2007), the sub-alignment was edited to remove 16
sequences with a high proportion of missing data at their
extremes and 34 sequences with ambiguous base calls
occurring at polymorphic positions. Haplotypes were
subsequently inferred with DnaSP v.5.10 (Librado &
Roza s 2009) considering sites with alignment gaps and
removing invariable sites. The network was artistically
edited in Adobe Illustrator CS5 and haplotypes were
labelled according to their geographical origin. DNA
polymorphism in the 70 remaining sequences was
evaluated with the software DnaSP v.5.10 (Librado &
Roza s 2009). The computed indices were the number of
segregating sites (s), the number of haplotypes (h),
haplotype diversity (Hd) calculated without considering
gap positions and the nucleotide diversity (π) using the
Jukes & Cantor (1969) correction. Deviations from
neutrality with Tajima's D and Fu's Fs statistics were also
assessed to infer past population size changes. The tests
were carried out in DnaSP v.5.10 using the nu mb e r o f
segregating sites, and their signicance was assessed based
on 10
4
coalescent simulations. We did not infer haplotype
networks, nor do we evaluate DNA polymorphism for G.
badipes and G. fallax because of the few sequences these
species encompassed and due to the substantial amount
of missing data in sequences masking the existing
polymorphism.
Dating analyses
The inference of a time frame for the global evolutionary
history of Galerina was conducted under a Bayesian
framework with BEAST 1.8.1 (Drummond et al. 2012).
Because this Basidiomycota genus lacks a suitable fossil
record, the dating analysis used a secondary calibration
imposed on the nrITS substitution rate. Hence, the
BEAST analysis implemented the average rate of
4.61 × 10
-3
substitutions per site per million years (s/s/Ma)
inferred for the genus Phaeocollybia R. Heim in Ryberg
&Matheny(2012), because this genus and Galerina
belong into the family Hymenogastraceae (Matheny et al.
5
PLEISTOCENE ANTARCTIC COLONIZATION OF THE MUSHROOM GENUS GALERINA
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
2015). This analysis was referred to as Dating A. To take
into account the uncertainty associated with that rate, we
re-ran analyses using the estimates representing the
minimum (2.92 × 10
-3
s/s/Ma, Dating B) and maximum
(6.45 × 10
-3
s/s/Ma, Dating C) values of the rate's
95% credibility interval provided by Ryberg & Matheny
(2012). The analyses were conducted with the two
alignment versions (GB and ORG) to learn about the
impact on age estimates of keeping ambiguously aligned
positions in the alignment. Redundant sequences were
removed from the alignments using the FaBox v.1.41
online toolbox (Villesen 2007). PartitionFinder 1.1.1
(Lanfear et al. 2012) was used to infer optimal
substitution models for the two nrITS partitions
considering a model with linked branch lengths and the
Bayesian information criterion. This analysis favoured the
GTR+Γmodel for the ITS1+ITS2 partition and the
K80 model for the 5.8S partition. We conducted
preliminary Bayes factor comparisons (Kass & Raftery
1995) of ML estimates (MLEs) calculated with path
sampling and stepping-stone approaches (Lartillot &
Philippe 2006,Xieet al. 2011) to choose among different
BEAST tree priors and molecular clocks. The use of an
uncorrelated lognormal relaxed molecular clock over the
strict clock was strongly supported for the GB and ORG
datasets (Tables SI & SII). As for the tree priors, models
incorporating the coalescent-constant size produced
substantially higher MLE values than models using the
birth-death and Yule process priors. Runs using chain
lengths of 1.5 × 10
8
steps were implemented, and
parameters were logged every 1.5 × 10
4
steps. Resulting log
les were checked in Tracer 1.7 to ensure that all
parameters had effective sample sizes > 200 after removing
the rst 20% of saved trees as burn-in. Then, the median
heights of the 1 × 10
4
post-burn-in tree samples were
annotated with TreeAnnotator 1.8.1, and the chronograms
were drawn with FigTree 1.4. Tracer 1.7 and
TreeAnnotator 1.8.1 are available at http://tree.bio.ed.ac.uk/.
We set the value of Bayesian posterior probabilities (PPs)
at a minimum of 0.97 for considering tree nodes to be well
supported.
Results
Specimen study
The G. marginata specimens studied morphologically and
phylogenetically showed relatively small pilei of up to 4 cm
in diameter, rst convex and then turning at (Fig. 2).
Basidia were tetrasporic and produced ellipsoid to
broadly amygdaloid, brownish and verrucose spores,
with a rounded apex and a visible hilar appendix and
containing one to two guttules. The size (length × width)
of 25 spores measured in water was (11.2) 13.0 ± 1.0
(15.8) × (6.2) 7.4 ± 0.8 (9.9) μm, and the length/width
ratio was (1.4) 1.8 ± 0.1 (2.0). Care must be taken when
comparing the size of pilei and microscopic
characteristics with literature data because specimens
were brought back from Antarctica frozen and were
examined after melting, a process that might have
affected the structure of these characteristics.
Alignments and phylogenies under ML
The original alignment (ORG) done with MAFFT
consisted of 178 Galerina nrITS sequences and
665 positions, of which 279 were variable and
82 corresponded to singleton sites. After processing the
alignment with GBlocks (GB), 603 positions (90% of the
original alignment) were retained in 29 selected blocks;
255 positions were variable and 73 were singleton sites.
The ML analyses in RAxML estimated phylogenies
with lnL = -5168.2 (GB) and lnL = -5493.96 (ORG).
Although the topologies inferred based on the two
alignments were not identical, they showed no supported
conicts (Supplemental Figs 1 & 2). In general, sister
relationships among the different Galerina species
included in the ingroup lacked support. Bootstrap values
> 70% were obtained for the crown nodes encompassing
all G. marginata,G. badipes and G. fallax sequences,
where the data obtained from Antarctic material are
placed in both topologies. It must be highlighted that the
G. marginata clade, hereinafter referred to as G. marginata
s.l., included sequences from specimens originally
labelled as Galerina autumnalis,G. hygroph ila,
G. pseudomycenopsis,G. unicolor and G. venenata
(including its type sequence, MH827070). Furthermore,
BS values > 78% were found for the sister relationship
of G. margi nat a s.l. and G. bad ipes,G. mi nima and
G. atkin son iana,G. pseudobadipes and G. stylifera,and
G. ce phal otri cha and G. mniophila. The sister relationship
of G. jaapii with the clade containing G. marginata s.l.
and G. badip es received a BS value of 82% (GB), whereas
the clade containing the latter three species along with
G. in dica had a BS value of 81%. Within G. margin ata s.l.,
the newly generated sequence from Livingston Island
(GenBank accession OQ569484) was located at the
bottom and close to three other sequences from
Antarctica: OP795715, which was obtained from a
basidioma collected at Norsel Point on Amsler Island
and differed by one nucleotide; KU559684, an
environmental sequence co-occurring in Antarctica and
the Arctic (Cox et al. 2016), which is shorter than the
other sequences and therefore was composed of a
number of missing nucleotides; and KT990212, labelled
as 'Arrhenia antarctica' and collected by Halina Galera
from an uncertain location within Antarctica that
showed missing and ambiguous positions together with
one diverging nucleotide. Two sequences labelled as
G. pseudomycenopsis (AJ585503 and GU234057) collected
6ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
in the USA and Svalbard were closely related as well
and differed, in general, by fewer than ve alignment
positions.
Genetic diversity in G. marginata s.l.
Forty-two haplotypes, producing a haplotype diversity
(Hd) of 0.930, were recovered from the 70 analysed
sequences of G. marginata s.l. (Fig. 3). The haplotype
network revealed a close relationship between the two
Antarctic haplotypes and others obtained from Northern
Hemisphere specimens, including either North America
(USA and Canada) or northern Europe (Scandinavia,
Baltic countries and the UK). Single mutations
segregated these haplotypes. Furthermore, the haplotype
network showed two star-like sub-networks separated by
just one mutation. The most evident sub-network was
composed of a central haplotype with a wide distribution
in North America (especially in Canada) that also
occurred in Switzerland (central Europe). Connected to
this central haplotype by just one or two mutations were
several minor haplotypes from North America and
northern Europe. In contrast, the second star-like
sub-network had a central haplotype distributed in
Europe overall and the Caucasus mountainous region,
and this was linked to minor haplotypes distributed in
Europe as well. At the bottom of the network in Fig. 3,a
number of haplotypes from Asia (Altay Republic, China,
South Korea and Japan) were connected by one or up to
four mutations to haplotypes occurring in North America
and Mexico. A couple of these haplotypes were shared by
regions on both sides of the PacicOcean.Finally,
Tajima's and Fu's neutrality tests estimated negative
values of D (-1.46358, P> 0.10; not signicant) and Fs
(-24.755, P< 0.001; signicant). The negative values of D
and Fs indicate an excess of low-frequency
polymorphisms relative to expectation and an excess of
the number of alleles, respectively. Collectively, the results
from these two tests suggest a population or demographic
expansion in G. marginata s. l.
BEAST phylogenies and age estimates
The average effective sample sizes were > 200 for all
parameters in the Bayesian dating analyses conducted
with BEAST based on the two alignments (GB and
ORG). No supported topological conicts were observed
between the two BEAST topologies (Fig. 4) or between
Fig. 3. Statistical parsimony networks connecting Galerina marginata haplotypes and summary of DNA polymorphism indices.
Haplotypes were coloured according to the geographical origin of samples. The sizes of the circles in the networks are proportional to
the numbers of individuals bearing the haplotype; black-lled smaller circles indicate missing haplotypes. Mutations are shown as
hatch marks. S= segregating sites; h= number of haplotypes; Hd = haplotype diversity; π= nucleotide diversity.
7PLEISTOCENE ANTARCTIC COLONIZATION OF THE MUSHROOM GENUS GALERINA
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
Fig. 4. Chronogram obtained with BEAST based on nrITS data depicting the evolutionary history of Galerina species. Dashed red
rectangles highlight the clades where the Antarctic collections are included. The mean age estimate for the divergence of selected nodes
is provided in million years ago (Ma). For each terminal in the tree, the GenBank, UNITE or BOLD nrITS accession number, the
taxonomic identity as originally deposited in these databases and the geographical origin are given. Green-lled rectangles indicate
nodal support (posterior probability (PP) 0.97) in analyses using the two versions of the nrITS alignment (GB and ORG). The
newly produced nrITS sequences with the corresponding GenBank codes are highlighted in bold. Numbers 14 in white circles
indicate phylogenetic clades where sequenced specimens of Antarctic Galerina are placed: 12=Galerina marginata,
3=Galerina badipes) and 4 = Galerina fallax.
8ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
them and those obtained in RAxML under ML
(Supplemental Figs 1 & 2). However, the BEAST
analysis using the GB alignment found support (PP >
0.97) for a sister relationship between a Galerina sp.
collected from the Balearic Islands (MH817980) and the
bulk of G. marginata s.l. sequences. Within G. marginata
s.l., sequences from Antarctica were phylogenetically
close to the same sequences reported in the RAxML
analyses. Moreover, two highly supported clades were
revealed: one encompassing several sequences from
North America and Asia, which correspond with the
haplotypes shown at the bottom of the haplotype
network (Fig. 3); and a second clade containing several
sequences obtained from North American as well as one
from Switzerland. This clade is represented by the most
evident star-like sub-network in Fig. 3. High support
Fig. 4. Continued.
9PLEISTOCENE ANTARCTIC COLONIZATION OF THE MUSHROOM GENUS GALERINA
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
was also revealed for a clade containing G. marginata s.l.
and G. badipes together with G. jaapii,G. indica,
G. triscopa f. telamonioides,G. pseudocamerina,
G. pruinatipes,G. chionophila and G. nana. Supported
sister relationships were also found for G. calyptrata,
G. sphagnicola and G. luteolosperma and for the pair
G. mniophila-cephalotricha.
The dating analysis based on the GB alignment
generated slightly lower age estimates than the analysis
that used the ORG alignment (Table I). However, the
inferred 95% highest posterior density (HPD) intervals
obtained with the two alternative analyses overlapped to
a considerable extent (Tab l e I ). For example, when the
average nrITS substitution rate of 4.61 × 10
-3
s/s/Ma was
employed (Dating A), the crown node of G. marginata
s.l. was dated back to 2.35 Ma (3.221.42 Ma,
95% HPD; GB alignment) and 3.09 Ma (4.441.74 Ma,
95% HPD; ORG alignment), a time interval at the
transition from the Pliocene to the Pleistocene. For
simplicity, the discussion below is based on the
chronogram estimated with the GB alignment (Fig. 4)
because it did not include potentially misaligned
(ambiguous) regions. This chronogram reveals that major
diversication events in Galerina took place since the
Miocene epoch (ca. 23.03 to 5.30 Ma) and extended into
the Pliocene (ca. 5.30 to 2.58 Ma). Speciation occurred
also in the Pleistocene (ca. 2.58 Ma to 11 700 years ago),
as is observed between the species G. mniophila and
G. cephalotricha and between G. minima and
G. atkin soni ana. Intraspecicdiversication in
G. marginat a s.l., G. bad ipes and G. fal lax, which include
sequences obtained from Antarctic material, occurred
mainly in the Pleistocene, with the Antarctic haplotypes
originating during the last 500 000 years on average
(Fig. 4). Finally, it should be highlighted that Dating B
and C, which used the minimum and maximum values of
the rate's 95% credibility interval provided by Ryberg &
Matheny (2012), produced older and younger age
estimates, respectively, compared with results using the
ave r ag e r ate val u e. Table I summarizes the age estimates
and corresponding 95% HPD intervals for selected nodes
(see Fig. 4) based on the three dating analyses (Dating A,
B and C).
Discussion
The present study validated by means of molecular
phylogenetics the existence in Antarctica of populations
of G. ma rg inata,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). These three Galerina
species, G. marginata,G. badipes and G. fallax, represent
relatively common macrofungi in Mediterranean and
Temperate-Arctic ecosystems in the Northern
Hemisphere (GBIF 2022). The closest genetic lineages to
sequenced Antarctic Galerina were in fact collected in
northern Europe (G. marginata), Greenland (G. badipes)
and North America (G. fallax), according to the inferred
phylogenies. Therefore, their distribution is here shown
to be potentially sub-cosmopolitan or amphitropical,
given the few occurrences in tropical regions (GBIF
2022). This biogeographical interpretation of our results
supports the opinion of Pegler et al. (1980), who used
morphological evidence to suggest a close similarity
between macrofungal species from the sub-Antarctic and
the Temperate-Arctic regions of the Northern
Hemisphere. The distribution patterns of the studied
Galerina match to a great extent with the global
geographical distribution of some non-lichenized
Antarctic microfungi (Bridge & Newsham 2009, Bridge
& Spooner 2012,Coxet al. 2016), but, most
interestingly, they also match with the amphitropical
Table I. Estimated divergence ages for the selected crown nodes in Fig. 4 representing Galerina species with Antarctic populations. The dating analyses in
BEAST used alternative nrITS alignment versions (GB vs ORG). For each of these versions, the mean age value and the corresponding 95% highest
posterior density intervals are provided in million years ago (Ma) considering the mean, minimum or maximum values of the nrITS substitution rate
inferred for the genus Phaeocollybia in Ryberg & Matheny (2012), which was used here for calibration purposes. The time frame proposed in the
'Epoch' column considered the six mean ages estimated in each row.
GBlocks-trimmed alignment (GB) Original alignment (ORG)
Mean Minimum Maximum Mean Minimum Maximum Epoch
(1) Crown Galerina marginata s.l. 2.35
(3.221.42)
3.91
(5.032.22)
1.61
(2.301.04)
3.09
(4.441.74)
5.47
(7.002.87)
1.76
(2.381.01)
Pliocene-Pleistocene
(2) Crown of clade with Antarctic Galerina
marginata s.l.
0.51
(0.880.07)
1.1
(1.630.14)
0.33
(0.720.08)
0.57
(0.900.13)
2.04
(2.160.49)
0.55
(0.830.23)
Pleistocene
(3) Crown Galerina badipes 0.69
(1.310.17)
1.33
(2.070.24)
0.43
(0.920.12)
0.79
(1.380.16)
1.78
(2.140.26)
0.55
(1.000.10)
Pleistocene
(4) Crown Galerina fallax 0.84
(1.560.23)
1.55
(2.420.35)
0.54
(1.100.19)
0.95
(1.630.24)
2.04
(2.550.39)
0.67
(1.160.17)
Pleistocene
10 ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
distribution pattern displayed by a signicant proportion
of lichenized fungi, which in Antarctica account for
almost 40% of lichens (Øvstedal & Lewis Smith 2001).
The existence of nearly identical global distribution
patterns in various lichenized and non-lichenized
Antarctic fungi makes us hypothesize that, at the
geological time scale, these species overcame similar
ecological and geographical lters to acquire their
current distribution. To accumulate evidence for
supporting or rejecting this hypothesis, other Antarctic
species of Galerina and members of additional
non-lichenized macrofungi genera should be surveyed
and studied phylogenetically.
The phylogenetic and haplotype network analyses
indicated that the studied intraspecic genetic lineages of
the Antarctic Galerina might be geographically restricted
and therefore endemic to this polar region. The fact
that the considered species produced fruiting bodies in
the surveyed Antarctic localities indicates that these
macrofungi have established permanent populations, and
therefore they are not transient visitors. Basidiomata
formation represents the last step in the life cycle of
Basidiomycota fungi. Briey, it starts with spore
germination and mycelium growth once abiotic and
biotic requirements are met, followed usually by mating
of two compatible, distinct mycelia, and, as a result,
basidiomata develop and spores are produced and
released after meiosis. We suggest that the three Galerina
species have been established in this region long enough
for mutations to accumulate in the studied genetic locus,
the nrITS, which is known to evolve at a higher rate
compared to other commonly used fungal molecular
markers (Schoch et al. 2012). Moreover, the existence of
Antarctic-endemic intraspecic lineages of these fungi is
of the utmost importance for designing conservation
policies that consider a broad spectrum of eukaryotic
organisms and not only plants and animals.
However, assessing endemicity in fungi poses some
risks, even at the intraspecic genetic level. In fact,
the existence or not of true Antarctic-endemic non-
lichenized fungal lineages has been hotly debated (e.g.
Bridge & Spooner 2012, Arenz et al. 2014) because of
the obvious difculties in observing and/or isolating
macro- and microfungi in Antarctica, or elsewhere, due
to their complex life cycles, ecologies and/or sizes. In this
sense, Bridge & Spooner (2012) mentioned that some of
the allegedly Antarctic-endemic species reported by
Onofri et al. (2005) were found later elsewhere. Assessing
endemicity in lichenized fungi is comparatively more
straightforward because they usually form macroscopic
and enduring lichen thalli (but see Hale et al. 2019).
Hence, the proportion of Antarctic-endemic lichens has
been estimated at 30% (Øvstedal & Lewis Smith 2001).
In our opinion, the analysis of biogeographical patterns
in non-lichenized Antarctic fungi may be more accurate
if approached phylogenetically as long as extensive
specimen and molecular datasets are compiled. For
example, genotypes restricted to Antarctica were
revealed for widespread fungi, such as the ascomycete
Thelebolus microsporus (Berk. & Broome) Kimbr. and
other microfungi (de Hoog et al. 2005, Bridge &
Newsham 2009, Bridge & Spooner 2012, Gonçalves
et al. 2017). Even so, the lack of availability of sequence
data and collections from as yet unexplored areas in the
Southern Hemisphere makes it difcult to assess
endemicity in Antarctic fungi. Although we compiled a
large specimen dataset in the present Galerina study, it
lacked sequence data associated with reports of
G. marginata and G. badipes from Australia and New
Zealand (GBIF 2022). Because of the geographical
proximity of these austral regions to Antarctica, it would
be worth checking whether Antarctic, Australian and
New Zealand populations of Galerina share the same
genotype or are at least closely related. In this way,
endemicity or colonization routes would be judged
more correctly. Furthermore, DNA sequence length
and quality are crucial for precise interpretations
of biogeographical patterns. This is of particular
importance today due to the abundance of short
sequences from DNA metabarcoding studies. If short
and long sequences are combined in datasets, the
resulting biogeographical interpretations could be
misleading to some extent, as unrealistic phylogenetic
afnities could be inferred. For example, we used a short
nrITS sequence of G. marginata that co-occurred in
Antarctica and the Arctic (KU559684; Cox et al. 2016).
Despite the existence of identical DNA sequences in
individuals from both poles having been observed in, for
example, the lichenized fungus Pseudephebe minuscula
(Arnold) Brodo & D. Hawksw. (Garrido-Benavent et al.
2021), it cannot be ruled out that comparison of the
Galerina sequences along their entire length would reveal
some genetic differences.
To the best of our knowledge, the present work is the
rst examining the temporal origins of Antarctic
macrofungal populations. Thus, the Antarctic lineages of
Galerina probably diverged from their Northern
Hemisphere relatives during the Pleistocene, based on a
consensus estimate of divergence times inferred with the
various dating strategies implemented in the present
study. In G. ma rgi nata, the divergence was probably
linked to a demographic expansion, as revealed by the
calculated neutrality tests. Moreover, the calculated
divergence time intervals for the three Galerina agree with
the inferred Pleistocene origin of Antarctic populations of
amphitropical lichens (Fernández-Mendoza & Printzen
2013,Garrido-Benaventet al. 2021). In the lichenized
fungi Cetraria aculeata (Schreb.) Fr. and P. minuscula,a
close genetic afnity of Maritime Antarctica specimens
to South American (Chilean) collections suggested a
11
PLEISTOCENE ANTARCTIC COLONIZATION OF THE MUSHROOM GENUS GALERINA
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
colonization route through the Sea of Hoces (Drake
Passage). In addition, continental specimens of the
second species were genetically close to Svalbard
(Northern Hemisphere) specimens, which indicated an
independent colonization route. The closest relatives of
G. marginata,G. badi pes and G. fallax, based on the
inferred phylogenies, grew in North America and northern
Europe, so that a direct, long-distance dispersal across the
tropics and ending in the establishment of Antarctic
Galerina populations might be assumed. However, the
possibility that these species colonized the Antarctic in a
series of stepping-stone movements from other territories
in the Southern Hemisphere, for which neither specimen
nor sequence data are yet available, must not be ruled
out. South America was in fact the region from which the
vascular plant D. antarctica is believed to have colonized
the Antarctic region, also during the Pleistocene
(Fasanella et al. 2017). It is worth recalling that the
studied Galerina species grew in tight association with
carpets of this plant as well as mosses, where these fungi
behave as saprophytes. Even some common populations
of Antarctic mosses had a Pleistocene origin (Pisa et al.
2014,Biersmaet al. 2017,2018). The overlapping
temporal frameworks for the origins of these plants and
fungi that coexist in the same Antarctic terrestrial
communities further support a relatively recent Antarctic
colonization of Galerina. The meiotic spores produced by
their basidiomata, which in general are ellipsoidal or
amygdaliform and < 15 μm in length, constitute the
expected mode of dispersal, and either wind currents or
migratory birds could be involved in such transoceanic
dispersals (Muñoz et al. 2004, Viana et al. 2016). For
example, Biersma et al. (2018) inferred aerial models that
indicated local wind patterns as the most probable
transfer mechanisms from southern South America to the
northern Maritime Antarctic. A greater research effort is
needed to corroborate these means of dispersal.
The biogeographical history of the studied Antarctic
Galerina has been interpreted on the basis of time
trees inferred using a secondary calibration (i.e. nrITS
substitution rate). Although this approach would be
expected to lead to more inaccurate dating results than
phylograms calibrated using fossil data (Schenk 2016),
the divergence times calculated for the whole Galerina
phylogeny in this study are largely coherent with
those reported for the divergence of species in other
Basidiomycota genera (e.g. Amanita,Heterobasidion,
Russula) that used different calibration strategies and
more extensive molecular sequence datasets (Chen et al.
2015, Sánchez-Ramírez et al. 2015, Looney et al. 2020).
Furthermore, the diversication events within the three
Antarctic Galerina species also agree with the inferred
colonization events of D. antarctica and mosses that,
together with these fungi, form typical terrestrial
habitats in Antarctica.
Acknowledgements
The UTM-CSIC technicians are thanked for their
assistance during Antarctic eldwork. The authors also
acknowledge Esther Rodríguez (MNCN) for technical
assistance in the laboratory; Alexandra Isern, Director
Antarctic Earth Sciences, US National Science
Foundation, Carolyn Lipke and other personnel at
Palmer Station, Polar Programs; the US National
Science Foundation for their help collecting Galerina
samples from the Antarctic Peninsula; and Benjamin
Held, University of Minnesota, for laboratory assistance.
This research was part of POLARCSIC activities. Two
anonymous peer reviewers are thanked for their feedback.
Funding
This work was supported by grants CTM2017-84441-R
(MINECO/FEDER, UE) and PID2019-105469RB-C22
(MICINN, AEI).
Author contributions
IG-B: conceived the project, conducted the analyses and
wrote the rst draft of the manuscript. RAB: contributed
to the molecular dataset and editing of the nal
manuscript. AdlR: obtained funding and eld resources,
conducted eldwork (including specimen collection) and
contributed to the editing of the nal manuscript.
Supplemental material
A supplemental table will be found at https://doi.org/10.
1017/S0954102023000196.
References
ALTSCHUL, S.F., GISH, W., MILLER, W., MYERS, E.W. & LIPMAN, D.J. 1990.
Basic Local Alignment Search Tool. Journal of Molecular Biology,
215, 403410.
ARENZ, B.E. & BLANCHETTE R.A. 2009. Investigations of fungal diversity
in wooden structures and soils at historic sites on the Antarctic
Peninsula. Canadian Journal of Microbiology,55,4656.
ARENZ, B.E. & BLANCHETTE R.A. 2011. Distribution and abundance of
soil fungi in Antarctica at sites on the Peninsula, Ross Sea region
and McMurdo Dry Valleys. Soil Biology and Biochemistry,43,
308315.
ARENZ, B.E., BLANCHETTE, R.A. & FARRELL, R.L. 2014. Fungal diversity
in Antarctic soils. In COWAN, D., ed. Antarctic terrestrial microbiology.
Berlin, Springer, 3553.
ARENZ, B.E., HELD, B.W., JURGENS, J.A. & BLANCHETTE R.A. 2011.
Fungal colonization of exotic substrates in Antarctica. Fungal
Diversity,49,1322.
ASPLUND,J.&WARDLE, D.A. 2017. How lichens impact on terrestrial
community and ecosystem properties. Biological Reviews,92,
17201738.
BERKELEY, M.J. 1847. Fungi. In HOOKER J.D., ed. The botany of the
Antarctic voyage of H.M. Discovery ships Erebus and Terror in the
years 18391843. Part 2. Flora Antarctica. Botany of Fuegia, The
Falklands, Kerguelen's Land, etc. London, Reeve Brothers, 447454.
12 ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
BIERSMA, E.M., JACKSON, J.A., BRACEGIRDLE, T.J., GRIFFITHS, H., LINSE,
K. & CONVEY, P. 2018. Low genetic variation between South American
and Antarctic populations of the bank-forming moss Chorisodontium
aciphyllum (Dicranaceae). Polar Biology,41, 599610.
BIERSMA, E.M., JACKSON, J.A., HYVÖNEN, J., KOSKINEN, S., LINSE, K.,
GRIFFITHS,H.&CONVEY, P. 2017. Global biogeographic patterns in
bipolar moss species. Royal Society Open Science,4, 170147.
BLANCHETTE, R.A., HELD, B.W., HELLMANN, L., MILLMAN,L.&
BÜNTGEN, U. 2016. Arctic driftwood reveals unexpectedly rich fungal
diversity. Fungal Ecology,23,5865.
BLANCHETTE, R.A., HELD, B.W., ARENZ, B.E. JURGENS, J.A., BALTES, N.J.,
DUNCAN, S.M. & FARRELL, R.L. 2010. An Antarctic hot spot for fungi
at Shackleton's historic hut on Cape Royds. Microbial Ecology,60,
2938.
BRIDGE, P.D. & NEWSHAM K.K. 2009. Soil fungal community of
composition at Mars Oasis, a southern Maritime Antarctic site,
assessed by PCR amplication and cloning. Fungal Ecology,2,6674.
BRIDGE,P.D.&SPOONER, B.M. 2012. Non-lichenized Antarctic fungi:
transient visitors or members of a cryptic ecosystem? Fungal
Ecology,5, 381394.
BRIDGE, P.D., SPOONER, B.M. & ROBERTS, P.J. 2008. List of
Non-Lichenized Fungi from the Antarctic Region. Version 2.3.3.
Retrieved from https://legacy.bas.ac.uk/bas_research/data/access/
fungi/ (accessed 13 September 2022).
BROA DY, P.A. & WEINSTEIN, R.N. 1998. Algae, lichens and fungi in La
Gorce Mountains, Antarctica. Antarctic Science,10, 376385.
CANINI, F., GEML,J.,D'ACQUI, L.P., SELBMANN, L.,ONOFRI, S., VENTURA,S.
&Z
UCCONI, L. 2020. Exchangeable cations and pH drive diversity and
functionality of fungal communities in biological soil crusts from
coastal sites of Victoria Land, Antarctica. Fungal Ecology,45, 100923.
CASTRESANA, J. 2000. Selection of conserved blocks from multiple
alignments for their use in phylogenetic analysis. Molecular Biology
and Evolution,17, 540552.
CHEN, J.J., CUI, B.K., ZHOU, L.W., KORHONEN,K.&DAI, Y.C. 2015.
Phylogeny, divergence time estimation, and biogeography of the
genus Heterobasidion (Basidiomycota, Russulales). Fungal Diversity,
71, 185200.
COLEINE, C., STAJICH, J.E., ZUCCONI, L., ONOFRI, S., POMBUBPA, N., EGIDI,
E., et al. 2018. Antarctic cryptoendolithic fungal communities are
highly adapted and dominated by Lecanoromycetes and
Dothideomycetes. Frontiers in Microbiology,9, 1392.
COX, F., NEWSHAM, K.K., BOL, R., DUNGAIT, J.A. & ROBINSON, C.H.
2016. Not poles apart: Antarctic soil fungal communities show
similarities to those of the distant Arctic. Ecology Letters,19, 528536.
DE HOOG, G.S., GOTTLICH, E., PLATAS, G., GENILLOUD, O., LEOTTA,G.&
VA N BRUMMELEN, J. 2005. Evolution, taxonomy and ecology of the
genus Thelebolus in Antarctica. Studies in Mycology,51,3376.
DRUMMOND, A.J., SUCHARD, M.A., XIE,D.&RAMBAUT, A. 2012.
Bayesian phylogenetics with BEAUti and the BEAST 1.7. Molecular
Biology and Evolution,29, 19691973.
FASANELLA, M., PREMOLI, A.C., URDAMPILLETA, J.D., GONZÁLEZ, M.L. &
CHIAPELLA, J.O. 2017. How did a grass reach Antarctica? The
Patagonian connection of Deschampsia antarctica (Poaceae).
Botanical Journal of the Linnean Society,185, 511524.
FERNÁNDEZ-MENDOZA,F.&PRINTZEN, C. 2013. Pleistocene expansion of
the bipolar lichen Cetraria aculeata into the Southern hemisphere.
Molecular Ecology,22, 19611983.
GARRIDO-BENAVENT, I., DELOS RÍOS, A., FERNÁNDEZ-MENDOZA,F.&
PÉREZ-ORTEGA, S. 2018. No need for stepping stones: direct, joint
dispersal of the lichen-forming fungus Mastodia tessellata
(Ascomycota) and its photobiont explains their bipolar distribution.
Journal of Biogeography,45, 213224.
GARRIDO-BENAVENT, I., PÉREZ-ORTEGA, S., DELOS RÍOS, A., MAYRHOFER,
H. & FERNÁNDEZ-MENDOZA, F. 2021. Neogene speciation and
Pleistocene expansion of the genus Pseudephebe (Parmeliaceae,
lichenized fungi) involving multiple colonizations of Antarctica.
Molecular Phylogenetics and Evolution,155, 107020.
GARRIDO-BENAVENT,I.,SØCHTING, U., DELOS RÍOS MURILLO,A.&PÉREZ-
ORTEGA, S. 2016. Shackletonia cryodesertorum (Teloschistaceae,
Ascomycota), a new species from the McMurdo Dry Valleys
(Antarctica) with notes on the biogeography of the genus
Shackletonia. Mycological Progress,15, 743754.
GARRIDO-BENAVENT, I., PÉREZ-ORTEGA, S., DURÁN, J., ASCASO,C.,
POINTING, S.B., RODRÍGUEZ-CIELOS, R., et al. 2020. Differential
colonization and succession of microbial communities in rock and
soil substrates on a Maritime Antarctic glacier foreeld. Frontiers in
Microbiology,11, 126.
GBIF. 2022. Galerina marginata (Batsch) Kühner. In GBIF Secretariat
(2021), GBIF Backbone Taxonomy. Checklist dataset. Retrieved
from https://doi.org/10.15468/39omei (accessed 15 September 2022).
GONÇALVES, V.N., VITORELI, G.A., DE MENEZES, G.C., MENDES, C.R.,
SECCHI, E.R., ROSA, C.A. & ROSA, L.H. 2017. Taxonomy, phylogeny
and ecology of cultivable fungi present in seawater gradients across
the Northern Antarctica Peninsula. Extremophiles,21, 10051015.
GRZESIAK,B.&WOLSKI, G.J. 2015. Bryophilous species of the genus
Galerina in peat bogs of Central Poland. Herzogia,28, 607623.
GULDEN, G., STENSRUD, Ø., SHALCHIAN-TABRIZI,K.&KAUSERUD,H.
2005. Galerina Earle: a polyphyletic genus in the consortium of
dark-spored agarics. Mycologia,97, 823837.
GUMIN
SKA, B., HEINRICH,Z.&OLECH, M. 1994. Macromycetes of the
South Shetland Islands (Antarctica). Polish Polar Research,15,103109.
HALE, E., FISHER, M.L., KEULER, R., SMITH,B.&LEAVITT, S.D. 2019. A
biogeographic connection between Antarctica and montane regions of
western North America highlights the need for further study of
lecideoid lichens. The Bryologist,122, 315324.
HARPER, C.J., TAY LO R , T.N., KRINGS,M.&TAY L O R , E.L. 2016.
Structurally preserved fungi from Antarctica: diversity and
interactions in late Palaeozoic and Mesozoic polar forest ecosystems.
Antarctic Science,28, 153173.
HELD,B.W.&BLANCHETTE R.A. 2017. Deception Island Antarctica
harbors a diverse assemblage of wood decay fungi. Fungal Biology,
121, 145157.
HORAK, E. 1994. Addenda ad Galerinam. 1. Galerina robertii sp. n., eine
neue Art aus den französischen Alpen. Zeitschrift für Mykologie,60,
8590.
JOLY, S., STEVENS, M.I. & VA N VUUREN, B.J. 2007. Haplotype networks can
be misleading in the presence of missing data. Systematic Biology,56,
857862.
JUKES, T.H. & CANTOR, C.R. 1969. Evolution of protein molecules.
Mammalian Protein Metabolism,3,21132.
KAPPEN, L., FRIEDMANN, E.I. & GARTY, J. 1981. Ecophysiologyof lichens in
the dry valleys of Southern Victoria Land, Antarctica I. Microclimate
of the cryptoendolithic lichen habitat. Flora,171,216235.
KASS, R.E. & RAFTERY, A.E. 1995. Bayes factors. Journal of the American
Statistical Association,90, 773795.
KATOH,K.&STANDLEY, D.M. 2013. MAFFT multiple sequence
alignment software version 7: improvements in performance and
usability. Molecular Biology and Evolution,30, 772780.
KOHLER, A., KUO, A., NAGY, L.G., MORIN, E., BARRY, K.W., BUSCOT,F.,
et al. 2015. Convergent losses of decay mechanisms and rapid
turnover of symbiosis genes in mycorrhizal mutualists. Nature
Genetics,47, 410415.
KRISHNAN, A., CONVEY, P., GONZÁLEZ-ROCHA,G.&ALIAS, S.A. 2016.
Production of extracellular hydrolase enzymes by fungi from King
George Island. Polar Biology,39,6576.
KRISHNAN, A., CONVEY, P., GONZÁLEZ, M., SMYKLA,J.&ALIAS, S.A. 2018.
Effects of temperature on extracellular hydrolase enzymes from soil
microfungi. Polar Biology,41, 537551.
LANDRY, B., WHITTON, J., BAZZICALUPO, A.L., CESKA,O.&BERBEE, M.L.
2021. Phylogenetic analysis of the distribution of deadly amatoxins
13PLEISTOCENE ANTARCTIC COLONIZATION OF THE MUSHROOM GENUS GALERINA
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
among the little brown mushroomsof the genus Galerina.PLoS ONE,
16, e0246575.
LANFEAR,R.,CALCOTT,B.,HO,S.Y.&GUINDON, S. 2012. PartitionFinder:
combined selection of partitioning schemes and substitution models for
phylogenetic analyses. Molecular Biology and Evolution,29, 16951701.
LARTILLOT,N.&PHILIPPE, H. 2006. Computing Bayes factors using
thermodynamic integration. Systematic Biology,55, 195207.
LATHA, K.D., RAJ, K.A., PARAMBAN,R.&MANIMOHAN, P. 2015. Two new
bryophilous agarics from India. Mycoscience,56,7580.
LEIGH,J.W.&BRYANT, D. 2015. PopART: full-feature software for
haplotype network construction. Methods in Ecology and Evolution,
6, 11101116.
LIBRADO,P.&ROZAS, J. 2009. DnaSP v5: a software for comprehensive
analysis of DNA polymorphism data. Bioinformatics,25, 14511452.
LOONEY, B.P., ADAMC
ÍK,S.&MATHENY, P.B. 2020. Coalescent-based
delimitation and species-tree estimations reveal Appalachian origin
and Neogene diversication in Russula subsection Roseinae.
Molecular Phylogenetics and Evolution,147, 106787.
MATHENY, P.B., MOREAU, P.A., VIZZINI, A., HARROWER, E., DEHAAN, A.,
CONTU,M.&CURTI, M. 2015. Crassisporium and Romagnesiella:two
new genera of dark-spored Agaricales. Systematics and Biodiversity,
13,2841.
MILLER, M.A., PFEIFFER,W.&SCHWARTZ, T. 2010. Creating the CIPRES
Science Gateway for inference of large phylogenetic trees. In
Proceedings of the Gateway Computing Environments Workshop
(GCE ), New Orleans, LA, 14 November 2010. Red Hook, NY:
Curran Associates, 18.
MUÑOZ, J., FELICÍSIMO, A.M., CABEZAS, F., BURGAZ, A.R. & MARTÍNEZ,I.
2004. Wind as a long-distance dispersal vehicle in the Southern
Hemisphere. Science,304, 11441147.
NEWSHAM, K.K., DAVEY, M.L., HOPKINS,D.W.&DENNIS, P.G. 2021.
Regional diversity of Maritime Antarctic soil fungi and predicted
responses of guilds and growth forms to climate change. Frontiers in
Microbiology,11, 615659.
NILSSON, R.H., LARSSON, K.H., TAYL O R , A.F.S., BENGTSSON-PALME,J.,
JEPPESEN, T.S., SCHIGEL,D.,et al. 2019. The UNITE database for
molecular identication of fungi: handling dark taxa and parallel
taxonomic classications. Nucleic Acids Research,47, D259D264.
ONOFRI, S., SELBMANN, L., ZUCCONI, L., TOSI,S.&DE HOOG, G.S. 2005.
The mycota of continental Antarctica. Terra Antarctic Reports,11,
3742.
ØVSTEDAL, D.O. & LEWIS SMITH, R.I. 2001. Lichens of Antarctica and
South Georgia: a guide to their identication and ecology. Cambridge
University Press, Cambridge, UK, 424 pp.
ØVSTEDAL,D.O.&LEWIS SMITH, R.I. 2011. Four additional lichens from
the Antarctic and South Georgia, including a new Leciophysma
species. Folia Cryptogamica Estonica,48,6568.
PEAT, H.J., CLARKE,A.&CONVEY, P. 2007. Diversity and biogeography of
the Antarctic ora. Journal of Biogeography,34, 132146.
PEGLER, D.N., SPOONER, B.M. & LEWIS SMITH, R.I. 1980. Higher fungi of
Antarctica, the subantarctic zone and Falkland Islands. Kew Bulletin,
35, 499562.
PÉREZ-ORTEGA, S., ORTI ZLVAREZ, R., GREEN, T.G.A. & DE LOS RÍOS,A.
2012. Lichen myco- and photobiont diversity and their relationships at
the edge of life (McMurdo Dry Valleys, Antarctica). FEMS
Microbiology Ecology,82, 429448.
PISA, S., BIERSMA, E.M., CONVEY, P., PATIÑO, J., VANDERPOORTEN, A.,
WERNER,O.&ROS, R.M. 2014. The cosmopolitan moss Bryum
argenteum in Antarctica: recent colonisation or in situ survival?
Polar Biology,37, 14691477.
PUTZKE,J.,PUTZKE, M.T.L., PEREIRA,A.B.&ALBUQUERQUE, M.P. 2012.
Agaricales (Basidiomycota) fungi in the South Shetland Islands,
Antarctica. INCT-APA Annual Activity Report. Retrieved from https://
doi.editoracubo.com.br/10.4322/apa.2014.065 (accessed 15 June 2023).
RATNASINGHAM,S.&HEBERT, P.D. 2007. BOLD: The Barcode of Life
Data System (http://www.barcodinglife.org). Molecular Ecology
Notes,7, 355364.
ROSA, L.H., PINTO, O.H.B., CONVEY, P., CARVALHO-SILVA, M., ROSA, C.A.
&C
ÂMARA, P.E.A.S. 2021. DNA metabarcoding to assess the diversity
of airborne fungi present over Keller Peninsula, King George Island,
Antarctica. Microbial Ecology,82, 165172.
RUISI, S., BARRECA, D., SELBMANN, L., ZUCCONI,L.&ONOFRI, S. 2007.
Fungi in Antarctica. Reviews in Environmental Science and Bio/
Technology,6, 127141.
RYBERG,M.&MATHENY, P.B. 2012. Asynchronous origins of
ectomycorrhizal clades of Agaricales. Proceedings of the Royal
Society B: Biological Sciences,279, 20032011.
SÁNCHEZ-RAMÍREZ, S., TULLOSS, R.E., AMALFI,M.&MONCALVO, J.M.
2015. Palaeotropical origins, boreotropical distribution and
increased rates of diversication in a clade of edible ectomycorrhizal
mushrooms (Amanita section Caesareae). Journal of Biogeography,
42, 351363.
SCHENK, J.J. 2016. Consequences of secondary calibrations on divergence
time estimates. PLoS ONE,11, e0148228.
SCHOCH, C.L., SEIFERT, K.A., HUHNDORF, S., ROBERT, V., SPOUGE, J.L.,
LEVESQUE, C.A., et al. 2012. Nuclear ribosomal internal transcribed
spacer (ITS) region as a universal DNA barcode marker for fungi.
Proceeding of the National Academy of Sciences of the United States
of America,109, 62416246.
SINGER,R.&CORTE, A. 1962. Estudio sobre los Basidiomycetes
antárcticos. Contribuciones cientícas del Instituto Antártico
Argentino,71,143.
SØCHTING, U., ØVSTEDAL,D.O.&SANCHO, L.G. 2004. The lichens of
Hurd Peninsula, Livingston Island, South Shetlands, Antarctica.
Bibliotheca Lichenologica,88, 607658.
STAMATAKIS, A. 2006. RAxML-VI-HPC: maximum likelihood-based
phylogenetic analyses with thousands of taxa and mixed models.
Bioinformatics,22, 26882690.
STAMATAKIS, A., HOOVER,P.&ROUGEMONT, J. 2008. A rapid bootstrap
algorithm for the RAxML web-servers. Systematic Biology,57,
758771.
TEMPLETON, A.R., CRANDALL, K.A. & SING, C.F. 1992. A cladistic
analysis of phenotypic associations with haplotypes inferred from
restriction endonuclease mapping and DNA sequence data. III.
Cladogram estimation. Genetics,132, 619633.
TOJU, H., TANABE, A.S., YAMAMOTO,S.&SATO, H. 2012. High-coverage
ITS primers for the DNA-based identication of ascomycetes and
basidiomycetes in environmental samples. PLoS ONE,7, e40863.
TRESEDER, K.K. & LENNON, J.T. 2015. Fungal traits that drive ecosystem
dynamics on land. Microbiology and Molecular Biology Reviews,79,
243262.
VIANA, D.S., SANTAMARÍA,L.&FIGUEROLA, J. 2016. Migratory birds as
global dispersal vectors. Trends in Ecology and Evolution,31, 763775.
VILLESEN, P. 2007. FaBox: an online toolbox for fasta sequences.
Molecular Ecology Notes,7, 965968.
VISHNIAC, H.S. & HEMPFLING, W.P. 1979. Evidence of an indigenous
microbiota (yeast) in the dry valleys of Antarctica. Microbiology,
112, 301314.
WHITE JR,J.F.&TAYL O R , T.N. 1991. Fungal sporocarps from Triassic
peat deposits in Antarctica. Review of Palaeobotany and Palynology,
67, 229236.
XIE, W., LEWIS, P.O., FAN, Y., KUO,L.&CHEN, M.H. 2011. Improving
marginal likelihood estimation for Bayesian phylogenetic model
selection. Systematic Biology,60, 150160.
14 ISAAC GARRIDOBENAVENT et al.
https://doi.org/10.1017/S0954102023000196 Published online by Cambridge University Press
Article
Full-text available
The investigation of Agaricales diversity in the Antarctica is limited, with only seven genera reported for the region. Galerina stands out as the genus with the highest species diversity, including 12 species in Antarctica. This research reports the presence of G. marginata in the region, providing the first complete morphological description for the specimen developing in Antarctica. Sampling was conducted during the Austral summer of 2022/2023 as part of the XLI Brazilian Antarctic Operation in Point Smellie, Byers Peninsula, Livingston Island, South Shetland Archipelago, Antarctica. Phylogenetic relationships reconstructed by Maximum Likelihood demonstrate that G. marginata forms a monophyletic clade with over 60% bootstrap support in most branches. The isolate in this study was found to be internal to the main cluster. Evolutionary reconstructions using the Maximum Likelihood method indicate that the branches correspond to the Antarctic isolate being an internal clade within the marginata group. Recording fungal populations in polar regions offers information about their adaptation and survival in inhospitable environments. Understanding the species' distribution in Antarctica encourages future investigations into its ecology and interactions with other organisms. Here, data are presented to establish an initial foundation for monitoring the G. marginata population in Antarctica and assessing the potential impacts of climate change on its development and survival in the forthcoming years. We report the third occurrence of Galerina marginata (Batsch) Kühner in Antarctica and provide, for the first time, a comprehensive morphological description of an individual of the species for the Antarctic continent, accompanied by phylogenetic analyses and comprehensive discussions regarding its diversity and global distribution.
Article
Full-text available
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.
Article
Full-text available
We report a metabarcoding study documenting the fungal taxa in 29 barren fellfield soils sampled from along a 1,650 km transect encompassing almost the entire maritime Antarctic (60–72°S) and the environmental factors structuring the richness, relative abundance, and taxonomic composition of three guilds and growth forms. The richness of the lichenised fungal guild, which accounted for 19% of the total fungal community, was positively associated with mean annual surface air temperature (MASAT), with an increase of 1.7 operational taxonomic units (OTUs) of lichenised fungi per degree Celsius rise in air temperature. Soil Mn concentration, MASAT, C:N ratio, and pH value determined the taxonomic composition of the lichenised guild, and the relative abundance of the guild was best predicted by soil Mn concentration. There was a 3% decrease in the relative abundance of the saprotrophic fungal guild in the total community for each degree Celsius rise in air temperature, and the OTU richness of the guild, which accounted for 39% of the community, was negatively associated with Mn concentration. The taxonomic composition of the saprotrophic guild varied with MASAT, pH value, and Mn, NH 4 ⁺ -N, and SO 4 ²⁻ concentrations. The richness of the yeast community, which comprised 3% of the total fungal community, was positively associated with soil K concentration, with its composition being determined by C:N ratio. In contrast with a similar study in the Arctic, the relative abundance and richness of lichenised fungi declined between 60°S and 69°S, with those of saprotrophic Agaricales also declining sharply in soils beyond 63°S. Basidiomycota, which accounted for 4% of reads, were much less frequent than in vegetated soils at lower latitudes, with the Ascomycota (70% of reads) being the dominant phylum. We conclude that the richness, relative abundance, and taxonomic composition of guilds and growth forms of maritime Antarctic soil fungi are influenced by air temperature and edaphic factors, with implications for the soils of the region as its climate changes during the 21st century.
Article
Full-text available
We assessed fungal diversity present in air samples obtained from King George Island, Antarctica, using DNA metabarcoding through high-throughput sequencing. We detected 186 fungal amplicon sequence variants (ASVs) dominated by the phyla Ascomycota, Basidiomycota, Mortierellomycota, Mucoromycota, and Chytridiomycota. Fungi sp. 1, Agaricomycetes sp. 1, Mortierella parvispora, Mortierella sp. 2, Penicillium sp., Pseudogymnoascus roseus, Microdochium lycopodinum, Mortierella gamsii, Arrhenia sp., Cladosporium sp., Mortierella fimbricystis, Moniliella pollinis, Omphalina sp., Mortierella antarctica, and Pseudogymnoascus appendiculatus were the most dominant ASVs. In addition, several ASVs could only be identified at higher taxonomic levels and may represent previously unknown fungi and/or new records for Antarctica. The fungi detected in the air displayed high indices of diversity, richness, and dominance. The airborne fungal diversity included saprophytic, mutualistic, and plant and animal opportunistic pathogenic taxa. The diversity of taxa detected reinforces the hypothesis that the Antarctic airspora includes fungal propagules of both intra- and inter-continental origin. If regional Antarctic environmental conditions ameliorate further in concert with climate warming, these fungi might be able to reactivate and colonize different Antarctic ecosystems, with as yet unknown consequences for ecosystem function in Antarctica. Further aeromycological studies are necessary to understand how and from where these fungi arrive and move within Antarctica and if environmental changes will encourage the development of non-native fungal species in Antarctica.
Article
Full-text available
Glacier forefields provide a unique chronosequence to assess microbial or plant colonization and ecological succession on previously uncolonized substrates. Patterns of microbial succession in soils of alpine and subpolar glacier forefields are well documented but those affecting high polar systems, including moraine rocks, remain largely unexplored. In this study, we examine succession patterns in pioneering bacterial, fungal and algal communities developing on moraine rocks and soil at the Hurd Glacier forefield (Livingston Island, Antarctica). Over time, changes were produced in the microbial community structure of rocks and soils (ice-free for different lengths of time), which differed between both substrates across the entire chronosequence, especially for bacteria and fungi. In addition, fungal and bacterial communities showed more compositional consistency in soils than rocks, suggesting community assembly in each niche could be controlled by processes operating at different temporal and spatial scales. Microscopy revealed a patchy distribution of epilithic and endolithic lithobionts, and increasing endolithic colonization and microbial community complexity along the chronosequence. We conclude that, within relatively short time intervals, primary succession processes at polar latitudes involve significant and distinct changes in edaphic and lithic microbial communities associated with soil development and cryptogamic colonization.
Article
Full-text available
A previously unreported population of Salicornia europaea has been found on the edges of a small tidal inlet in the emirate of Umm al-Qaiwain. It appears to be different from two ecotypes of the species found in other parts of the United Arab Emirates (UAE). The location has large numbers of an introduced halophytic plant species, Sesuvium portulacastrum, which serve to inhibit the growing of local species. The number of plants of the local S. europaea ecotype is low and the species may disappear. The eradication of the alien species of S. portulacastrum from the area is necessary to protect the S. europaea ecotype and other local flora in the area.
Article
Full-text available
UNITE (https://unite.ut.ee/) is a web-based database and sequence management environment for the molecular identification of fungi. It targets the formal fungal barcode-the nuclear ribosomal internal transcribed spacer (ITS) region-and offers all ∼1 000 000 public fungal ITS sequences for reference. These are clustered into ∼459 000 species hypotheses and assigned digital object identifiers (DOIs) to promote unambiguous reference across studies. In-house and web-based third-party sequence curation and annotation have resulted in more than 275 000 improvements to the data over the past 15 years. UNITE serves as a data provider for a range of metabarcoding software pipelines and regularly exchanges data with all major fungal sequence databases and other community resources. Recent improvements include redesigned handling of unclassifiable species hypotheses, integration with the taxonomic backbone of the Global Biodiversity Information Facility, and support for an unlimited number of parallel taxonomic classification systems.
Article
Widespread geographic distributions in lichens have been usually explained by the high dispersal capacity of their tiny diaspores. However, recent phylogenetic surveys have challenged this assumption and provided compelling evidence for cryptic speciation and more restricted distribution ranges in diverse lineages of lichen-forming fungi. To evaluate these scenarios, we focus on the fungal genus Pseudephebe (Parmeliaceae) which includes amphitropical species, a distribution pattern whose origin has been a matter of debate since first recognized in the nineteenth century. In our study, a six-locus dataset and a broad specimen sampling covering almost all Earth’s continents is used to investigate species delimitation in Pseudephebe. Population structure, gene flow and dating analyses, as well as genealogical reconstruction methods, are employed to disentangle the most plausible transcontinental migration routes, and estimate the timing of the origin of the amphitropical distribution and the Antarctic populations. Our results demonstrate the existence of three partly admixed phylogenetic species that diverged between the Miocene and Pliocene, and whose Quaternary distribution has been strongly driven by glacial cycles. Pseudephebe minuscula is the only species showing an amphitropical distribution, with populations in Antarctica, whereas the restricted distribution of P. pubescens and an undescribed Alaskan species might reflect the survival of these species in European and North American refugia. Our microevolutionary analyses suggest a Northern Hemisphere origin for P. minuscula, which could have dispersed into the Southern Hemisphere directly and/or through “mountain-hopping” during the Pleistocene. The Antarctic populations of this species are sorted into two genetic clusters: populations of the Antarctic Peninsula were grouped together with South American ones, and the Antarctic Continental populations formed a second cluster with Bolivian and Svalbard populations. Therefore, our data strongly suggest that the current distribution of P. minuscula in Antarctica is the outcome of multiple, recent colonizations. In conclusion, our results stress the need for integrating species delimitation and population analyses to properly approach historical biogeography in lichen-forming fungi.
Article
Ice-free regions in coastal areas of Victoria Land, Antarctica, are patchily distributed, limited in extent and characterized by a simple vegetation of lichens and mosses, growing only for a short period during the austral summer. These organisms are associated with soil particles and microorganisms (e.g., algae, microfungi and bacteria) to make up biological soil crusts (BSCs), found worldwide in cold and/or arid and semi-arid regions, where plant growth is impaired. Despite BSCs being among the most widespread ecosystems throughout coastal ice-free areas of continental Antarctica, fungal components of these communities have received little focus. Through ITS1 DNA metabarcoding of samples from 17 sites of six different localities from 73 to 77°S, in a distance scale from 29 to 411 km among different sites, we provide insights into the diversity, community composition, and functionality of fungal communities of these peculiar ecosystems, deepening our knowledge on how they are related to different edaphic variables (i.e. chemical properties and texture). Although fungal richness was low (59 ± 27 OTUs per sample), we found numerous previously unsequenced, putatively unknown fungal species representing a great part of the sampled communities. Community composition was spatially auto-correlated and appeared to be driven by site-specific differences in environmental conditions, particularly edaphic factors, such as exchangeable cations and pH. These results are of particular interest, as they give a wide characterization of the parameters determining soil colonization in a such limiting environment, especially in the light of global changes that are expected to deeply modify the conditions of this environment.
Article
Numerous lineages of mushroom-forming fungi have been subject to bursts of diversification throughout their evolutionary history, events that can impact our ability to infer well-resolved phylogenies. However, groups that have undergone quick genetic change may have the highest adaptive potential. As the second largest genus of mushroom-forming fungi, Russula provides an excellent model for studying hyper-diversification and processes in evolution that drives it. This study focuses on the morphologically defined group – Russula subsection Roseinae. Species hypotheses based on morphological differentiation and multi-locus phylogenetic analyses are tested in the Roseinae using different applications of the multi-species coalescent model. Based on this combined approach, we recognize fourteen species in Roseinae including the Albida and wholly novel Magnarosea clades. Reconstruction of biogeographic and host association history suggest that parapatric speciation in refugia during glacial cycles of the Pleistocene drove diversification within the Roseinae, which is found to have a Laurasian distribution with an evolutionary origin in the Appalachian Mountains of eastern North America. Finally, we detect jump dispersal at a continental scale that has driven diversification since the most recent glacial cycles.
Article
Processes that shape biogeographic patterns in lichens have been of long-standing interest, especially in extreme environments. Lecideoid lichens in Antarctica have been relatively well-studied and are thought to have a high degree of endemism. However, recent collection efforts in montane regions of western North America have uncovered lecideoid lichens morphologically similar to some from Antarctic and South American Sub-Antarctic regions, including Lecidea andersonii Filson and L. polypycnidophora U.Rupr. & Türk. To explore the similarity between these putative conspecifics, we used ITS and MCM7 sequence data to infer relationships within a phylogenetic framework. The resulting phylogenetic reconstructions provided further evidence of the close relationship between specimens from western North America with geographically distant lecideoid populations in the Southern Hemisphere. We recovered two diverse, well-supported clades, provisionally called the L. andersonii complex and the ‘Lecidea NA-ARG clade,’ both including species-level subclades. A number of lecideoid species and species-level lineages thought to be endemic to Antarctica, the South American Sub-Antarctic region, or with bipolar distributions also occur in montane regions of western North America. In some cases, these species-level clades corresponded to distinct geographic regions, e.g., L. polypycnidophora, while other species failed to show phylogenetic structure corresponding to distinct geographic regions, e.g., L. andersonii. These findings bring into question the origin and prevalence of truly endemic Antarctic lichen-forming fungi. Additional investigations into lecideoid lichens, including the biogeographic connection between western North America with Antarctica, will likely provide novel perspectives into the connection between Antarctica and North America, potential colonization routes and the timing of dispersal events.