ArticlePDF Available

Turnover of Lecanoroid Mycobionts and Their Trebouxia Photobionts Along an Elevation Gradient in Bolivia Highlights the Role of Environment in Structuring the Lichen Symbiosis

Frontiers
Frontiers in Microbiology
Authors:

Abstract and Figures

Shifts in climate along elevation gradients structure mycobiont–photobiont associations in lichens. We obtained mycobiont (lecanoroid Lecanoraceae) and photobiont ( Trebouxia alga) DNA sequences from 89 lichen thalli collected in Bolivia from a ca. 4,700 m elevation gradient encompassing diverse natural communities and environmental conditions. The molecular dataset included six mycobiont loci (ITS, nrLSU, mtSSU, RPB1 , RPB2 , and MCM7 ) and two photobiont loci (ITS, rbc L); we designed new primers to amplify Lecanoraceae RPB1 and RPB2 with a nested PCR approach. Mycobionts belonged to Lecanora s.lat., Bryonora , Myriolecis , Protoparmeliopsis , the “ Lecanora ” polytropa group, and the “ L .” saligna group. All of these clades except for Lecanora s.lat. occurred only at high elevation. No single species of Lecanoraceae was present along the entire elevation gradient, and individual clades were restricted to a subset of the gradient. Most Lecanoraceae samples represent species which have not previously been sequenced. Trebouxia clade C, which has not previously been recorded in association with species of Lecanoraceae, predominates at low- to mid-elevation sites. Photobionts from Trebouxia clade I occur at the upper extent of mid-elevation forest and at some open, high-elevation sites, while Trebouxia clades A and S dominate open habitats at high elevation. We did not find Trebouxia clade D. Several putative new species were found in Trebouxia clades A, C, and I. These included one putative species in clade A associated with Myriolecis species growing on limestone at high elevation and a novel lineage sister to the rest of clade C associated with Lecanora on bark in low-elevation grassland. Three different kinds of photobiont switching were observed, with certain mycobiont species associating with Trebouxia from different major clades, species within a major clade, or haplotypes within a species. Lecanoraceae mycobionts and Trebouxia photobionts exhibit species turnover along the elevation gradient, but with each partner having a different elevation threshold at which the community shifts completely. A phylogenetically defined sampling of a single diverse family of lichen-forming fungi may be sufficient to document regional patterns of Trebouxia diversity and distribution.
Content may be subject to copyright.
fmicb-12-774839 December 14, 2021 Time: 13:52 # 1
ORIGINAL RESEARCH
published: 20 December 2021
doi: 10.3389/fmicb.2021.774839
Edited by:
Lucia Muggia,
University of Trieste, Italy
Reviewed by:
Gábor M. Kovács,
Eötvös Loránd University, Hungary
Julieta Orlando,
University of Chile, Chile
*Correspondence:
Jolanta Miadlikowska
jolantam@duke.edu
These authors have contributed
equally to this work and share first
authorship
These authors share senior
authorship
Specialty section:
This article was submitted to
Microbe and Virus Interactions with
Plants,
a section of the journal
Frontiers in Microbiology
Received: 13 September 2021
Accepted: 19 November 2021
Published: 20 December 2021
Citation:
Medeiros ID, Mazur E,
Miadlikowska J, Flakus A,
Rodriguez-Flakus P,
Pardo-De la Hoz CJ, Cie ´
slak E,
´
Sliwa L and Lutzoni F (2021) Turnover
of Lecanoroid Mycobionts and Their
Trebouxia Photobionts Along an
Elevation Gradient in Bolivia Highlights
the Role of Environment in Structuring
the Lichen Symbiosis.
Front. Microbiol. 12:774839.
doi: 10.3389/fmicb.2021.774839
Turnover of Lecanoroid Mycobionts
and Their Trebouxia Photobionts
Along an Elevation Gradient in Bolivia
Highlights the Role of Environment in
Structuring the Lichen Symbiosis
Ian D. Medeiros1, Edyta Mazur2, Jolanta Miadlikowska1*, Adam Flakus2,
Pamela Rodriguez-Flakus2, Carlos J. Pardo-De la Hoz1, El ˙
zbieta Cie´
slak2,
Lucyna ´
Sliwa2and François Lutzoni1
1Department of Biology, Duke University, Durham, NC, United States, 2W. Szafer Institute of Botany, Polish Academy
of Sciences (PAS), Kraków, Poland
Shifts in climate along elevation gradients structure mycobiont–photobiont associations
in lichens. We obtained mycobiont (lecanoroid Lecanoraceae) and photobiont (Trebouxia
alga) DNA sequences from 89 lichen thalli collected in Bolivia from a ca. 4,700
m elevation gradient encompassing diverse natural communities and environmental
conditions. The molecular dataset included six mycobiont loci (ITS, nrLSU, mtSSU,
RPB1,RPB2, and MCM7) and two photobiont loci (ITS, rbcL); we designed new
primers to amplify Lecanoraceae RPB1 and RPB2 with a nested PCR approach.
Mycobionts belonged to Lecanora s.lat., Bryonora,Myriolecis,Protoparmeliopsis, the
Lecanorapolytropa group, and the “L.” saligna group. All of these clades except for
Lecanora s.lat. occurred only at high elevation. No single species of Lecanoraceae was
present along the entire elevation gradient, and individual clades were restricted to a
subset of the gradient. Most Lecanoraceae samples represent species which have not
previously been sequenced. Trebouxia clade C, which has not previously been recorded
in association with species of Lecanoraceae, predominates at low- to mid-elevation
sites. Photobionts from Trebouxia clade I occur at the upper extent of mid-elevation
forest and at some open, high-elevation sites, while Trebouxia clades A and S dominate
open habitats at high elevation. We did not find Trebouxia clade D. Several putative new
species were found in Trebouxia clades A, C, and I. These included one putative species
in clade A associated with Myriolecis species growing on limestone at high elevation
and a novel lineage sister to the rest of clade C associated with Lecanora on bark in
low-elevation grassland. Three different kinds of photobiont switching were observed,
with certain mycobiont species associating with Trebouxia from different major clades,
species within a major clade, or haplotypes within a species. Lecanoraceae mycobionts
and Trebouxia photobionts exhibit species turnover along the elevation gradient, but
Frontiers in Microbiology | www.frontiersin.org 1December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 2
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
with each partner having a different elevation threshold at which the community shifts
completely. A phylogenetically defined sampling of a single diverse family of lichen-
forming fungi may be sufficient to document regional patterns of Trebouxia diversity
and distribution.
Keywords: elevation gradients, systematics, symbiosis, Andes mountains, new PCR primer, Lecanoromycetes,
Trebouxiophyceae, lichen biogeography
INTRODUCTION
Alexander von Humboldt’s Essai sur La Géographie des Plantes
was revolutionary for biogeography (Schrodt et al., 2019),
but, with respect to lichens, Humboldt was limited by the
nascent state of lichenology in the early 19th century. Drawing
on observations from Europe and South America, Humboldt
described lichens as “independent of the influence of the
climates” (von Humboldt and Bonpland, 2009)—a statement
that, with a modern understanding of lichen systematics and
symbiosis, we now know is far from the truth. Climate does
matter for lichens. There is turnover in lichenized fungi along
elevation gradients, including those in the northern Andes (Wolf,
1993;Soto-Medina et al., 2019) that inspired Humboldt’s work.
Furthermore, the mycobiont–photobiont interaction that defines
the lichen symbiosis is affected by environmental conditions at
micro and macro scales (e.g., James and Henssen, 1976;Peksa
and Škaloud, 2011;Singh et al., 2017). What appears externally
as the same lichen may actually represent a mycobiont species
occurring with different photobiont species at different points in
its range (Werth and Sork, 2014;Dal Grande et al., 2018).
Lichen photobiont biodiversity and patterns in the
mycobiont–photobiont association have been most commonly
studied at a regional, intra-biome scale at which climate is largely
consistent [Hestmark et al., 2016;Jüriado et al., 2019;Yahr et al.,
2004; but see Lu et al. (2018) for an intra-biome study along
a climatic gradient] or at a global scale, at which climate may
vary widely but regional patterns may not be apparent (Muggia
et al., 2014;Nyati et al., 2014;Lutsak et al., 2016;Magain et al.,
2017, 2018;Vanˇ
curová et al., 2018;Chagnon et al., 2019). Less
commonly, regional photobiont turnover across adjacent biomes
has been studied along gradients of latitude (Werth and Sork,
2014) or altitude (Dal Grande et al., 2018). In most studies,
investigations of lichen photobionts have been structured around
a specific mycobiont taxon, whether that taxon is a species (e.g.,
Blaha et al., 2006;Werth and Sork, 2014), genus (e.g., Vargas
Castillo and Beck, 2012;Magain et al., 2017, 2018), or family
(e.g., Helms, 2003;Nyati et al., 2014).
Lecanoraceae is one of the three largest families of lichenized
fungi that associate with the green alga Trebouxia, the most
common genus of lichen photobionts (Miadlikowska et al.,
2006;Lücking et al., 2017;Muggia et al., 2018). The family is
cosmopolitan and occurs on a range of substrates, typically with
a crustose growth form (Brodo et al., 2001). Despite the family’s
global distribution and diversity of corticolous species, the
literature on the Lecanoraceae–Trebouxia association has mostly
dealt with mycobiont taxa that are saxicolous, phylogenetically
outside Lecanora s.str., and from Europe, North America, or
Antarctica (Blaha et al., 2006;Guzow-Krzemi´
nska, 2006;Pérez-
Ortega et al., 2012;Ruprecht et al., 2012, 2020;Leavitt et al.,
2015, 2016;Wagner et al., 2020). Photobiont studies that have
included corticolous species of Lecanoraceae have generally done
so in the context of sampling photobionts from across the
lichen community of a small geographic area in the northern
hemisphere (e.g., Singh et al., 2019).
Trebouxia is divided into five major clades: A, C, D, I, and
S (Muggia et al., 2020;Xu et al., 2020). Clade D was recently
identified and is known from Iceland, Svalbard, and Tierra del
Fuego (Xu et al., 2020). Clade C is largely tropical, with a few
temperate representatives. The remaining clades, A, I, and S, are
cosmopolitan, including Antarctica. At the level of these major
clades, lineages are distinguished by differences in pyrenoid
structure (Muggia et al., 2020). The species-level classification of
Trebouxia is still incompletely understood. Leavitt et al. (2015,
2016) and Muggia et al. (2020) implemented an alphanumeric
classification system for putative Trebouxia species based on a
so-called “barcode gap” species delimitation method (Puillandre
et al., 2012). Although some of the species recognized by Muggia
et al. (2020) correspond to named taxa, the majority have not
been formally described.
In this paper, we investigate mycobiont–photobiont
interactions in Lecanoraceae lichens from across an elevation
gradient in Bolivia. Recent studies on lichen photobionts in
Bolivia have uncovered putative new species of Trentepohlia and
Asterochloris (Kosecka et al., 2020, 2021a), but the diversity of
Trebouxia photobionts in Bolivia remains largely unexplored
outside of one global study on photobionts of Cetraria aculeata
(Lutsak et al., 2016). Our dataset includes corticolous and
saxicolous species from Lecanora and other genera in
Lecanoraceae and encompasses nearly 5,000 m of elevation
change in the Andes mountains, south of where Humboldt
developed his ideas on biogeography. We use this dataset to ask
three main questions: (1) How is the biodiversity of Bolivian
Lecanoraceae distributed, both in the phylogeny of this fungal
family and along the elevation gradient? (2) What major clades
and species of Trebouxia are associated with Lecanoraceae in
Bolivia? (3) How do climate, substrate, and mycobiont host
structure photobiont communities?
MATERIALS AND METHODS
Field Sampling
Specimens of Bolivian lichens suspected of belonging to
Lecanoraceae based on their morphology were collected from
2004 to 2019. Collection sites spanned over 4,700 m of
Frontiers in Microbiology | www.frontiersin.org 2December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 3
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
elevation and included diverse natural communities across
multiple biomes (Navarro and Maldonado, 2002;Navarro and
Ferreira, 2007). Habitats ranged from patches of trees in
Los Llanos de Moxos savanna below 200 m, to Yungas and
Tucumano-Boliviano montane forests at mid-elevation sites, to
dry, rocky grasslands of the high Andean Puna at about 4,000
m and subnival vegetation above 4,600 m (Figure 1). Voucher
specimens are stored in KRAM and LPB.
Morphological Studies
Morphology and anatomy were studied by standard techniques.
Cross sections of apothecia were mounted in water for
observation of apothecium anatomy or in ca. 25% KOH
for observation of ascospores. Crystals were observed under
polarized light in a compound fluorescence microscope (Nikon
Eclipse 80i) and their solubility was tested with KOH and
65% nitric acid. Lichen secondary metabolites were analyzed
by thin-layer chromatography (TLC) in solvents A, B’, and C
following Culberson and Kristinsson (1970) and Orange et al.
(2001). For saxicolous samples, rocks were tested for carbonate
minerals with 10% HCl.
DNA Extraction, PCR, and Sequencing
From a collection of approximately 550 specimens, 235
specimens representing diverse morphotypes and habitats were
selected for molecular study. Note that the genus Lecidella,
although part of Lecanoraceae, was not targeted in this
sampling focused on lecanoroid species. Material collected in
the years 2015–2018 was kept frozen at 20C prior to DNA
extraction, while collections from earlier years were stored under
standard herbarium conditions. Specimens collected in 2019
were extracted when fresh. Apothecia excised from a single
thallus—or soralia, if apothecia were absent—were cleaned in
sterile, distilled water on a microscope slide. Visible contaminants
(e.g., lichenicolous fungi) were removed with ultrathin tweezers
and a razor blade. DNA was isolated using either the QIAamp
DNA Investigator Kit or DNeasyTM Plant Mini Kit (Qiagen,
Germany), depending on the amount of lichen tissue available,
following the manufacturer’s instructions.
To conduct a multi-locus phylogenetic analysis of
Lecanoraceae, we amplified the mycobiont nuclear ribosomal
large subunit (nrLSU), internal transcribed spacer (ITS),
mitochondrial small subunit (mtSSU), RNA polymerase
II largest (RPB1) and second-largest (RPB2) subunits, and
minichromosome maintenance factor 7 (MCM7). The ribosomal
RNA (rRNA) loci ITS, mtSSU, and to some extent nrLSU
have been frequently sequenced for phylogenetic studies of
Lecanoraceae and represent most of the reference data for this
family on GenBank. The ITS region, including the ITS1 spacer,
5.8S rRNA, and ITS2 spacer, was amplified with primers ITS1F
(Gardes and Bruns, 1993) and either LR3 or ITS4 (Vilgalys and
Hester, 1990;White et al., 1990); nrLSU was amplified with
primers AL2R and LR6 (Vilgalys and Hester, 1990;Döring et al.,
2000); and mtSSU was amplified with primers MSU1 and MSU7
(Zhou and Stanosz, 2001). Thermal cycler conditions for the
three rRNA loci were 95C for 3 min, followed by 35 cycles of
95C for 40 s, 52C for 40 s, and 72C for 150 s, with a final
extension at 72C for 10 min.
Due to the very small amount of DNA present in our
extractions, the protein-coding loci RPB1,RPB2, and MCM7
were amplified with a nested PCR approach. For RPB1, the initial
amplification was performed with primers RPB1-Af and either
RPB1-Cr or RPB1-Dr (Stiller and Hall, 1997;Matheny et al.,
2002). In the first round of PCR, the thermal cycler conditions
were 94C for 10 min, followed by 25 cycles of 94C for 45 s,
50C for 50 s, and 72C for 100 s, with a final extension at 72C
for 10 min. The second round of PCR used forward and reverse
primers lecRPB1-F and lecRPB1-R (Table 1), which we designed
based on the Lecanoraceae RPB1 alignment from Zhao et al.
(2016). Suitable priming sites were identified by visual inspection
of alignments in Mesquite (Maddison and Maddison, 2018) and
potential primer sequences were evaluated using online tools1,2.
The thermal cycler conditions for the second round were 94C
for 5 min, followed by 35 cycles of 94C for 45 s, 55C for
60 s, and 72C for 90 s, with a final extension at 72C for
10 min. For RPB2, the first round of PCR used primers fRPB2-5F
and fRPB2-7cR (Liu et al., 1999), and thermal cycler conditions
were identical to the first-round program for RPB1. The second
round used primers lecRPB2-6F and fRPB2-7cR. The newly
designed forward primer lecRPB2-6F (Table 1) was modified
from bRPB2-6F (Matheny, 2005) using Lecanoraceae sequences
from Miadlikowska et al. (2014). Thermal cycler conditions for
this round were the same as for the first round, except that the
number of cycles was increased to 35 and the extension time was
reduced to 85 sec per cycle. For MCM7, the first round of PCR
used primers MCM7-709f and MCM7-1443r, while the second
round used MCM7-709f and MCM7-1348r (Schmitt et al., 2009).
The thermal cycler programs for both rounds followed Schmitt
et al. (2009), except that the number of cycles in the first round
was reduced to 25.
Sanger sequencing of PCR amplicons has been shown
to reliably recover the dominant photobiont in Trebouxia-
associated lichens even when other photobiont genotypes are
present at low levels (Paul et al., 2018). Photobiont ITS was
amplified with primers ITS1T and ITS4T (Kroken and Taylor,
2000) using the thermal cycler program described above for
the mycobiont rRNA loci. For a subset of specimens, we
also amplified a portion of the Rubisco large subunit (rbcL)
with primers a-ch-rbcL-203-5’-MPN and a-ch-rbcL-991-3’-MPN
(Nelsen et al., 2011). Thermal cycler conditions followed Nelsen
et al. (2011). Selection of specimens for rbcL sequencing was
based on the photobiont ITS results and focused on specimens
that were: (1) outliers in the elevation distribution or host breadth
of a photobiont clade, (2) photobionts that appeared to be
new species-level lineages, (3) species-level photobiont lineages
only found in a single specimen, or (4) previously recognized
Trebouxia lineages for which rbcL data were lacking.
1https://eurofinsgenomics.eu/en/ecom/tools/pcr-primer-design/ (accessed June
24, 2021).
2https://www.thermofisher.com/us/en/home/brands/thermo-scientific/
molecular-biology/molecular-biology- learning-center/molecular-
biologyresource-library/thermo-scientific-web-tools/multiple-primer-analyzer.
html (accessed June 24, 2021).
Frontiers in Microbiology | www.frontiersin.org 3December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 4
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
FIGURE 1 | Habitats where Lecanoraceae lichens were collected. (A) La Paz, Bautista Saavedra, 1448’9”S, 6910’51”W, 4850 m, open high Andean vegetation
with siliceous rocks. (B) La Paz, Bautista Saavedra, 1515’0”S, 692’50”W, 4549 m, open high Andean area with limestone rocks. (C) La Paz, Franz Tamayo,
1446’39”S, 690’35”W, 2550 m, Yungas montane cloud forest. (D) Tarija, Aniceto Arce, 2141’36”S, 6429’33”W, 2195 m, Tucumano-Boliviano montane forest.
(E) Santa Cruz, Cordillera, 1827’29”S, 6123’1”W, 292 m, Chaqueño forest. (F) Beni, Yacuma, 1451’7”S, 6620’23”W, 175 m, Los Llanos de Moxos savanna. All
photographs by AF.
PCR amplicons were checked on 1% agarose gels to confirm
the presence of a single fragment size and cleaned following
an enzymatic cleanup protocol with exonuclease I and shrimp
alkaline phosphatase (ThermoFisher Scientific, Waltham, MA,
United States). Sanger sequencing was performed by Eurofins
Genomics (Louisville, KY, United States). Sequencing reactions
used PCR primers except in two cases. Primers mrSSU1
and mrSSU3R (Zoller et al., 1999) were used in place of
MSU1 and MSU7 to sequence the mtSSU amplicons. A newly
designed primer, lecRPB2-seq7R (Table 1) was used in place of
fRPB2-7cR for sequencing the RPB2 amplicons. Forward and
reverse sequence reads were assembled, trimmed, and checked
for base-calling errors in SeqMan Pro (DNASTAR, Madison,
WI, United States) or Geneious Prime 2021.0.33. When ITS
3https://www.geneious.com
Frontiers in Microbiology | www.frontiersin.org 4December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 5
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
TABLE 1 | Lecanoraceae-specific RPB1 and RPB2 primers designed in
the present study.
Locus Primer
name
Use Nucleotide sequence (5’–3’)
RPB1 lecRPB1-F PCR GAR ACN GTY TGY CAY AAY TGY GGC AAG
lecRPB1-R PCR C RAA YTC RTT NAC NAC RTG NGC NGG
RPB2 lecRPB2-
6F
PCR TGG GGN YTR GTM TGY CCD GC
lecRPB2-
seq7R
sequencing G RTT RTG RTC NGG RAA NGG
and partial nrLSU were sequenced as one amplicon, they
were split using the program ITSx (Bengtsson-Palme et al.,
2013). Sequences were checked for contamination or specimen
misidentification using two complementary methods: (1) BLAST
with the NCBI nucleotide database and (2) placement in a
phylogeny of Lecanoromycetes with T-BAS (Miadlikowska et al.,
2014;Carbone et al., 2017, 2019). GenBank numbers for all newly
obtained sequences are given in Supplementary Data Sheet 1.
Alignments and Phylogenetic Analyses
We compiled a reference dataset (Supplementary Data Sheet 2)
of Lecanoraceae and outgroups based on taxa included in
Miadlikowska et al. (2014) and Zhao et al. (2016). In
general, reference taxa were only included if multiple loci,
including protein-coding loci, were available. The families
Parmeliaceae (represented by one species each of Protoparmelia
and Letharia) and Gypsoplacaceae (represented by two species
of Gypsoplaca) were used as outgroups. Our Lecanoraceae
dataset included several species groups within Lecanora,
Palicella,Lecidella,Protoparmeliopsis,Rhizoplaca,Myriolecis, the
Lecanorapolytropa group, and the “Lecanorasaligna group,
as well as Haematomma,Ramboldia, and Miriquidica. Reference
sequences were checked for specimen misidentifications or other
metadata errors using BLAST with the NCBI nucleotide database.
The mycobiont rRNA loci (mycobiont ITS, nrLSU, and
mtSSU) were initially aligned using the MAFFT online server4
with the G-INS-1 option (Katoh et al., 2019). The alignments
were corrected by eye in Mesquite (Maddison and Maddison,
2018) and introns and ambiguously aligned regions were
delimited manually and excluded from downstream analyses
(Lutzoni et al., 2000). Protein-coding loci were aligned by
translated amino acid in Mesquite. Introns were delimited
manually and excluded from downstream analyses.
Each single-locus alignment was used as input for a maximum
likelihood phylogenetic analysis in IQ-TREE version 2.1.2
(Nguyen et al., 2015;Chernomor et al., 2016) run on the
CIPRES server (Miller et al., 2010). We performed 5,000 ultrafast
bootstrap pseudoreplicates to calculate support for each single-
locus tree (Hoang et al., 2018). We inspected the resulting
trees for well-supported conflicts (95% ultrafast bootstrap;
Hoang et al., 2018) among the six mycobiont loci, which were
then concatenated to form a single dataset. The concatenated
mycobiont alignment was used as input for a partitioned
4https://mafft.cbrc.jp/alignment/server/
maximum likelihood analysis in IQ-TREE. We ran IQ-TREE with
the -p and -m MFP + MERGE options so that ModelFinder
(Kalyaanamoorthy et al., 2017) would both optimize the
partitioning scheme and find the best-fitting substitution model
for each partition. The input partitioning scheme divided the
concatenated alignment by locus and by codon position for
the protein-coding loci. The final partitioning schemes and
substitution models used in this analysis are provided in Table 2.
We performed 5,000 ultrafast bootstrap pseudoreplicates to
calculate bipartition support for the tree topology.
We performed two phylogenetic analyses with the photobiont
data. First, all sequences were included in a global ITS-rbcL
alignment for Trebouxia to assign each sample to one of the
major Trebouxia clades. We used the ITS and rbcL sequences
curated by Muggia et al. (2020) as a reference dataset, with
slight modifications. These sequences are listed in the file
“Supplementary Data 2” from Muggia et al. (2020). We removed
sequence L138 from the Trebouxia clade S ITS alignment
and sequences L1107, L1360, L1401, L1417, L1425, L1504,
AJ007387, AJ969549, AJ249482, and 9493 from the Trebouxia
clade A ITS alignment because they were missing a significant
portion of ITS1 or ITS2. None of these ITS sequences had a
corresponding rbcL sequence. L1107 and 9493 were the only
representatives of the putative species A37 and A49, respectively,
but preliminary analyses indicated that our new sequences
were not closely related to these taxa. We also added all
ITS and rbcL reference sequences for Trebouxia clade D from
“Supplementary Table 5” of Xu et al. (2020). Representatives
from three other genera of Trebouxiophyceae were used as
outgroup taxa: Asterochloris, GenBank accession JN573844;
Myrmecia, KM462861; and Vulcanochloris, KR952313. Only rbcL
was included for the outgroup taxa because ITS was not alignable
across the four genera.
ITS sequences were initially aligned with MAFFT using
the G-INS-1 option and regions that were not conserved
across Trebouxia were excluded by eye in Mesquite; rbcL
sequences were aligned by translated amino acid in Mesquite.
Phylogenies for each locus were inferred as described above
for the mycobiont loci. The two alignments were concatenated
after inspecting single-locus trees for well-supported conflicts.
We used the concatenated dataset as input for a maximum
likelihood phylogenetic analysis in IQ-TREE. Model selection,
tree inference, and bootstrapping parameters were the same as
described above for the concatenated mycobiont dataset, with the
initial partitioning scheme dividing the alignment by locus and
further splitting rbcL by codon position.
For the second set of analyses, we performed a separate
phylogenetic analysis for each major clade of Trebouxia
represented in our sampling. The genus-wide alignment allowed
unknown sequences to be assigned to one of the major clades, but
because variable regions of ITS1 and ITS2 had to be excluded,
relationships within these clades were often poorly resolved. For
each major clade, reference taxa and sample sequences were
extracted from the global ITS and rbcL alignments. The ITS
alignments were corrected by eye in Mesquite and ambiguous
regions were re-delimited. One rbcL sequence from each of the
other major clades of Trebouxia was retained for the outgroup.
Frontiers in Microbiology | www.frontiersin.org 5December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 6
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
TABLE 2 | Statistics on the alignments for Lecanoraceae (mycobiont) after the removal of introns and ambiguously aligned regions.
Locus Length Invariable (%) PI (%) Missing taxa (%) Model/Partition
ITS 252 144(57)69(27)14(9)1
nrLSU 1024 701(68)214(21)35(23)1
mtSSU 789 531(67)172(22)29(19)2
RPB1 630 298(47)297(47)42(27)3:3:4
RPB2 744 329(44)367(49)56(36)3:2:4
MCM7 537 269(50)249(46)102(66)3:2:4
The full concatenated alignment included 154 taxa and 3,976 sites. PI, parsimony-informative sites. Substitution models were as follows: 1, TN + F + R4; 2, HKY + F + R3;
3, GTR + F + R3; 4, TIM2e + I + G4. Models for protein-coding loci are given as first:second:third codon positions. Loci and codon positions with the same substitution
model were fused into a single partition.
After checking for well-supported conflicts between the two loci
as described above, we concatenated the ITS and rbcL alignments
and performed a maximum-likelihood analysis in IQ-TREE.
Model selection, tree inference, and bootstrapping parameters
were the same as described above for the concatenated mycobiont
dataset. The initial partitioning scheme divided the alignment
by locus and further split the ITS by region (ITS1, 5.8S, and
ITS2) and rbcL by codon position. The final partitioning schemes
and substitution models used for these analyses are provided in
Table 3.
To explore intraspecific diversity in a subset of putative
Trebouxia species, we inferred haplotype networks in R version
3.6.1 (R Core Team, 2019) using the packages ape (Paradis and
Schliep, 2019) and pegas (Paradis, 2010). Sites with ambiguous
bases were excluded from the haplotype analysis.
Classification of Mycobionts and
Photobionts
Our preliminary delimitation of mycobiont species was based on
morphology, chemistry, and the inferred phylogeny. We used
the program bPTP (Zhang et al., 2013) to inform our species
delimitation, using the concatenated maximum likelihood
phylogeny as input and running the Markov chain Monte Carlo
(MCMC) analysis for 500,000 generations, sampling every 500
generations5. The first 25% of samples were discarded as burn-in.
Based on the ITS-rbcL phylogeny, photobiont sequences were
assigned to a clade and putative species based on the framework
from Muggia et al. (2020). Specimens that were phylogenetically
distinct from any previously delimited species were considered to
represent novel putative species.
Ecological Analyses
To better understand the ecological diversity across the elevation
gradient, we extracted 19 bioclimatic variables from the
WorldClim database (Fick and Hijmans, 2017) for each sampling
locality. Bioclimatic variables were sampled at a resolution of
10 arc-minutes using the R packages sp (Pebesma and Bivand,
2005) and raster (Hijmans, 2021). We used the hclust function to
perform a hierarchical clustering of the 19 bioclimatic variables,
which was visualized as a dendrogram. Vegetation types for each
site were classified with field locality descriptions and vegetation
maps from Navarro and Ferreira (2007).
5https://species.h-its.org/
We analyzed patterns of species richness (αdiversity) and
turnover (βdiversity) along the elevation gradient. From the
original data for each lichen specimen (Supplementary Data
Sheet 1), we prepared inferred presence/absence matrices for the
mycobiont and photobiont species. The elevation gradient was
split into 500 m bins starting at sea level and extending to 4,999 m.
For each Lecanoraceae or Trebouxia species, we noted the lowest
and highest elevations at which the species was found, and coded
the species as “present” for all bins included in that range. For
example, a Trebouxia species found at 2,400, 3,050, and 3,600 m
would be coded as present from 2,000 to 3,999 m even though it
was absent from the 2,500 to 2,999 m bin. The Sørensen index of
dissimilarity was calculated for all pairwise comparisons between
elevation bins using the vegdist function in the R package vegan
(Oksanen et al., 2020).
RESULTS
Sequenced Specimens
We obtained DNA sequence data from approximately half of
the 235 specimens for which extractions were performed. Most
of the specimens that failed to yield any sequence data were
those collected before 2015 and stored under standard herbarium
conditions. Of the specimens from which we obtained sequence
data, approximately ten were not Lecanoraceae and were
excluded from further analysis. A further ca. 20 specimens yielded
only mycobiont or only photobiont sequences; the sequences
from specimens without paired mycobiont–photobiont data
were not used for the present paper. Finally, five specimens
were excluded from the analysis because of well-supported
conflicts between loci or discordance between molecular and
morphological data. A total of 89 specimens with both mycobiont
and photobiont data were used for the remaining analyses. We
generated over 500 new DNA sequences from this group of
specimens: 83 mycobiont ITS, 68 nrLSU, 79 mtSSU, 70 RPB1, 64
RPB2,22MCM7, 89 photobiont ITS, and 27 rbcL.
Alignments
Summary statistics for the Lecanoraceae and Trebouxia
alignments are presented in Tables 2,3, respectively. For the
mycobiont, almost all of ITS1 and a large fraction of ITS2 could
not be aligned across Lecanoraceae and were excluded from the
phylogenetic analysis. The three protein-coding loci contained
Frontiers in Microbiology | www.frontiersin.org 6December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 7
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
TABLE 3 | Statistics on the alignments for Trebouxia (photobiont) after the removal of introns and ambiguously aligned regions.
Clade ITS rbcL
Length PI (%) Model/Partition PI (%) Model/Partition
A 516 150(29)1:2:1 83(11)2:2:3
C 514 129(25)4:5:4 121(15)5:5:6
I 505 101(20)7:5:7 86(11)5:5:8
S 552 123(22)9:5:9 39(5)5:5:3
PI, parsimony-informative sites. The rbcL alignment was 789 bases long for all clades and had no excluded positions. Substitution models were as follows: 1 = TVMe + R3;
2 = K2P + R2; 3 = TIM3 + F + R3; 4 = SYM + R3; 5 = K2P + I; 6 = TIM2 + F + R3; 7 = TPM2 + F + G4; 8 = TIM2 + F + I; 9 = TIM3e + G4. Models are given as ITS1:5.8S:ITS2
for the ITS and first:second:third codon positions for rbcL. ITS regions and codon positions with the same substitution model were fused into a single partition.
the greatest number of parsimony-informative sites in absolute
terms and relative to their length (Table 2). For the Trebouxia
sequences, most of ITS1 and ITS2 could be aligned within each
major clade. ITS provided most of the parsimony-informative
sites for all four Trebouxia single-clade alignments, while rbcL
varied from being nearly as informative as ITS in Trebouxia clade
C to being minimally informative in Trebouxia clade S (Table 3).
Complete alignments with excluded regions delimited are
available in the supplementary materials (Supplementary Data
Sheets 3–8).
Lecanoraceae and Trebouxia Phylogeny
and Biodiversity
Photobionts from members of the Lecanoraceae collected in
Bolivia represented approximately 21 putative species in four
of the five major clades within Trebouxia (Figure 2 and
Supplementary Figure 1). The majority of photobionts were
phylogenetically nested within previously delimited species,
but six lineages appear to represent new species-level taxa in
Trebouxia clades A, C, and I (Supplementary Figures 2–4).
All samples from Trebouxia clade S corresponded to previously
delimited species (Supplementary Figure 5). No samples were
from Trebouxia group D (Supplementary Figure 1).
The single-locus trees for the mycobiont ITS, nrLSU, and
mtSSU had some well-supported nodes near the tips, but
deeper relationships were for the most part poorly supported
(Supplementary Figures 68). RPB1,RPB2, and MCM7 were
better able to resolve deeper relationships in Lecanoraceae,
with RPB1 yielding the best-supported tree of any single
locus (Supplementary Figures 9–11). The maximum likelihood
tree from the concatenated six-locus dataset (Figure 3 and
Supplementary Figure 12) included strong support for many
putative genus- and subgenus-level clades and some relationships
among those clades, but the backbone was poorly supported
in several regions of the tree. For example, the relationships
among Haematomma,Miriquidica,Ramboldia, and the rest of
the family were not well supported, nor was the position of
Lecidella. However, none of our mycobiont specimens belonged
to these genera.
Sampled mycobionts belong to two well-supported clades
within Lecanoraceae (Figure 3). The first clade, which we call
Lecanora s.lat., includes the Lecanora subfusca group (Lecanora
clades I and II), L. subcarnea group (Lecanora clade III), a group
we call Lecanora clade IV, L. formosa and Palicella,L. cavicola,
and L. rupicola. None of the relationships among these clades
were well supported. Species of Lecanora s.lat. were found
predominantly on bark at low- to mid-elevation sites, but some
collections were from rock at high elevation (Figure 4). The high-
elevation, saxicolous specimens mostly belonged to L. rupicola,
L. cavicola, and L. formosa, but also included taxa from Lecanora
clades II and III (Figures 4,5A).
The second major clade, which we call the MPRPS clade for
lack of a formal name, included high-elevation collections from
Myriolecis,Protoparmeliopsis, the “Lecanorapolytropa group,
Bryonora, and the “Lecanorasaligna group. MPRPS also includes
Rhizoplaca, although none of our collections belong to that
genus (Figure 3). Myriolecis and Protoparmeliopsis form a well-
supported clade, and the “L.saligna group is a well-supported
sister to the rest of the MPRPS clade, but the phylogenetic
placements of the “L.polytropa group, Rhizoplaca, and Bryonora
are uncertain. Except for the single specimen of Bryonora and one
Myriolecis specimen that were found on moss and two specimens
of the “L.saligna group that were lichenicolous lichens, species
in the MPRPS clade occurred on rock (Figure 4).
Based on morphology, chemistry, and the molecular
phylogeny, our sampling included approximately 27 species in
Lecanora s.lat. and approximately 10 species in the MPRPS clade
(Figure 4). The tree topology and species delimitation analysis
suggest that several of the morphospecies contain multiple
cryptic species. For example, there are three well-supported
lineages on long branches in the morphospecies Lecanora
ecoronata (Lecanora clade I), and the specimens identified as L.
caesiorubella were distributed across three lineages in Lecanora
clade III that did not form a monophyletic group (Figure 4).
Photobiont Ecology and Mycobiont
Associations
The distribution of Trebouxia species was structured by elevation,
substrate, and mycobiont identity (Figures 4,5). Species of
the MPRPS clade were always found with photobionts from
Trebouxia clades A and S. Myriolecis,Protoparmeliopsis, and the
single specimen of Bryonora were exclusively associated with
Trebouxia clade A, while the “Lecanorapolytropa and “L.
saligna groups occurred only with Trebouxia clade S. All of these
mycobiont genera were restricted to high elevations (Figure 5A).
This pattern is also related to substrate chemistry: Saxicolous
Frontiers in Microbiology | www.frontiersin.org 7December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 8
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia






   
 
0.04
FIGURE 2 | Summary phylogeny showing relationships among major clades of Trebouxia based on maximum likelihood analysis of the ITS-rbcL dataset. Bold clade
names indicate clades represented in our Bolivian sampling. Bold branches indicate UFboot2 support 95. Scale represents substitutions per site. The color
scheme introduced here will be used for these Trebouxia clades throughout the rest of the paper. For the full tree with all tip names and bootstrap values, see
Supplementary Figure 1.



















 




0.04

FIGURE 3 | Summary phylogeny showing relationships within Lecanoraceae based on maximum likelihood analysis of the six-locus dataset. The clade names
Lecanora I–IV, Lecanora s.lat., and MPRPS (Myriolecis,Protoparmeliopsis,Rhizoplaca, “L.” polytropa, “L.” saligna) will be used throughout the paper in the sense
shown here. Bold clade names and filled triangles indicate clades represented in our Bolivian sampling. Bold branches indicate UFboot2 support 95. Scale
represents substitutions per site. For the full tree with all tip names and bootstrap values, see Supplementary Figure 12.
Frontiers in Microbiology | www.frontiersin.org 8December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 9
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia





































1000
2000
3000
4000
elevation (m) vegetation
Beni savanna
Chaceño−Amazon forest
Tucumano−Boliviano forest
Yungas forest
Humid Puna high−Andean
Humid Puna subnival
Other
substrate
bark
rock (siliceous)
moss
rock (calcareous)






 





lichens
Trebouxia
substrate
elevation
vegetation

mycobiont

























































































FIGURE 4 | Photobiont association, substrate, elevation, and vegetation type in the context of the mycobiont phylogeny. The tree is from the same six-locus analysis
as Figure 3 and Supplementary Figure 12; reference taxa were removed in R with the drop.tip function in ape and data were added to the tree with the package
ggtree (Yu et al., 2017). Bold branches indicate clades with UFboot2 support 95 in Supplementary Figure 12. Lecanoraceae species highlighted in green were
found with photobionts from more than one major clade of Trebouxia. Lecanoraceae species highlighted in yellow were found with multiple Trebouxia species from
the same clade. Colors used for Trebouxia clades are the same as in Figure 2. Species-level relationships within Myriolecis are not well-resolved in this analysis.
species of Myriolecis and Protoparmeliopsis grew on calcareous
rock, while species of the “L.” polytropa group and the lichen
hosts of members of the “L.” saligna group occurred on siliceous
rock (Figure 4). All of these mycobiont genera were restricted to
high-elevation sites (Figure 5A).
Species of the Lecanora s.lat. clade were found in association
with Trebouxia clades A, C, I, and S. There was a pronounced
turnover in photobiont clade along the elevation gradient, with
Trebouxia clade C at low to mid-elevations (up to ca. 3,000 m),
clade I mostly at upper mid elevations (ca. 2,000–3,500 m), and
clade S at high elevations (above 3,500 m) (Figure 5A). Trebouxia
clade A was only rarely associated with Lecanora s.lat., always
above 2,000 m (Figure 5A).
Clades within Lecanora s.lat. occupied varying portions of
the total elevation gradient (Figure 5A). Lecanora clade I was
only absent from the highest elevations, while Lecanora clade
III was missing at low elevations; Lecanora clade IV occupied
intermediate elevations. Lecanora clade II was primarily at low
elevations, except for a single high elevation species. Lecanora
clades I–IV were each associated with photobionts from at least
two, and often three, Trebouxia clades. The same elevational
turnover of photobiont clades that is apparent for Lecanoraceae
as a whole also occurs within these smaller clades; see, for
example, Lecanora clades I and III (Figure 5A).
At the level of putative Trebouxia species, there is considerable
variation in the level of specificity exhibited toward habitat and
mycobiont (Figures 4,5B). Most Trebouxia species occupied a
500–1,000 m subset of the elevation gradient (Figure 5B). Some
taxa occurred only with a single mycobiont clade in that narrow
elevation range. For example, Trebouxia C13 was only found
Frontiers in Microbiology | www.frontiersin.org 9December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 10
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
Mycobiont clade
Elevation (m)
Putative Trebouxia species
Elevation (m)
B
A
0
1000
2000
3000
4000
5000
Lecanora−II
Lecanora−IV
Lecanora−I
Lecanora−III
L. calvicola
Protoparmeliopsis
Bryonora
Myriolecis
“L.” polytropa group
L. rupicola
“L.” saligna group
L. formosa
0
1000
2000
3000
4000
5000
C36
C13
C35
C10
C34
C02
A19
I19
C09
I13 s.lat.
I20
I15
A05
I18
S10
A53
A aff. 07
A01
A04
S02
A13
I
II
III
IV
V
VI
VII
VIII
C09 haplotype network
Trebouxia
A
C
I
S
Substrate
bark
lichen
moss
rock (siliceous)
rock (calcareous)
FIGURE 5 | Elevation distribution of Lecanoraceae mycobionts and their Trebouxia photobionts. (A) Elevation distribution of the major mycobiont clades, showing
Trebouxia clades and substrates. Note that photobiont turnover across the elevation gradient is apparent for Lecanoraceae as a whole and within individual
mycobiont clades. There are relatively few samples from the 3,000 to 4,000 m range because specimens for molecular study were originally selected for a taxonomic
study of Lecanoraceae. Although we have many collections from this altitudinal range, they were not particularly diverse morphologically and therefore were not
sampled heavily in our sequencing. (B) Elevation distribution of putative Trebouxia species. A single photobiont species typically occupies 500–1,000 m of the
elevation gradient. Inset shows haplotype network for Trebouxia C09, which occupies 1,500 m of the gradient. The haplotypes clustered on the left occur across the
elevation range of this species, while those clustered on the right only occur at the bottom of the range.
Frontiers in Microbiology | www.frontiersin.org 10 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 11
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
below 500 m with species of Lecanora clade I, while Trebouxia
A53 only occurred with species of Myriolecis at about 4,500 m
(Figure 4). Other photobionts were generalists with respect to
mycobiont but were only found in a specific habitat. Trebouxia
A13, S02, and S10 occurred with mycobiont species from both
the Lecanora s.lat. and MPRPS clades, but nearly always above
4,000 m (Figure 4). On the other hand, some photobionts
exhibited substantial niche breadth for both mycobiont host
and elevation. Trebouxia C09, the most broadly distributed
photobiont species in our sampling, was associated with species
from Lecanora clades I–IV and occurred on both bark and rock
from 1,500 to 3,000 m (Figures 4,5B).
Numbers of mycobiont and photobiont species were tightly
linked across the elevation gradient, with the number of
mycobiont species nearly always greater; both Lecanoraceae
and Trebouxia had the greatest number of species at high
elevation (Figure 6A). Species turnover was consistently high
across the elevation gradient for both partners (Figure 6B).
Each partner had a mid- to mid-high-elevation threshold
across which no species were shared, but this threshold
differed between Lecanoraceae (2,500 m) and Trebouxia
(3,500 m) (Figure 6B).
Photobiont Switching
Three mycobiont species were associated with photobionts from
two or more Trebouxia clades (Figure 4). Lecanora sp. 1
(n= 4) occurred on sandstone in open areas between 2,700
and 3,300 m. Below 3,000 m, this species was found with
Trebouxia C09, while above 3,000 m it was found with Trebouxia
I15. Lecanora rupicola (n= 2) was found in association with
Trebouxia A13 and S02 at a single high-elevation site, where
it grew on siliceous rocks. Lecanora aff. farinacea (n= 4)
occurred with Trebouxia C09 at ca. 2,900 m and with Trebouxia
I18 and S02 at ca. 4,250 m. All four collections were from
siliceous rocks, either sandstone at the lower site or schist at
the upper sites.
A larger number of mycobiont species associated with
multiple photobionts within one of the major Trebouxia clades
(Figure 4). “Lecanoraintricata (n= 4) and L. cavicola (n= 3)
each associated with Trebouxia S02 and S10. At one site at
ca. 4,260 m, L. cavicola was found with both photobionts. All
specimens of these two species were collected from siliceous rocks
in open vegetation above 4,000 m. Protoparmeliopsis garovaglii
(n= 2), which was only collected at a single site at ca. 4,550 m, on
limestone, occurred with Trebouxia A01 and A aff. 07. Lecanora
caesiorubella 1 (n= 2) occurred with Trebouxia C34 at ca.
1,900 m and Trebouxia C02 at 2,250 m; both sites were in
Yungas forest. Lecanora ecoronata 1 (n= 9) was predominantly
associated with Trebouxia C13, but at the same savanna sites also
occurred with Trebouxia C35 and C36. Lecanora sp. 2 (n= 3)
occurred with Trebouxia C13 at the bottom of its range and
C10 at the top of its range. Lecanora flavidomarginata (n= 2)
occurred with Trebouxia I18 and I20, at ca. 4,000 m and ca. 3,400
m, respectively.
Haplotypes of Trebouxia C09 fell into two clusters
(Figure 5B). One cluster spanned the entire elevation range of
this putative species, while the other occurred only in the lower
portion of its range. Lecanora aff. stramineoalbida was associated
with both haplotype clusters.
DISCUSSION
Mycobiont Phylogeny
Phylogenetic studies of Lecanoraceae have long been bedeviled
by a poorly supported backbone (Grube et al., 2004;Pérez-
Ortega et al., 2010;Zhao et al., 2016;Yakovchenko et al.,
2019;Davydov et al., 2021). Most of these studies, including
recent publications, have used ITS and mtSSU to infer the
phylogeny of Lecanoraceae, but those loci are most useful for
resolving relationships within genera or species groups and
cannot resolve deeper relationships across the entire family.
Zhao et al. (2016) obtained better support with a six-locus
dataset, but their analysis excluded many morphologically
distinct species groups for which the protein-coding loci
were lacking. In the present study, we used the same six
loci as Zhao et al. (2016): ITS, nrLSU, mtSSU, RPB1,RPB2,
and MCM7. We expanded their taxon sampling, bringing
in additional clades that were missing from their six-locus
analysis: Bryonora,Lecanora IV, Lecanora cavicola,Lecanora
rupicola, and the “Lecanora” saligna group. Despite this expanded
sampling, we found that several regions of the phylogeny remain
poorly supported.
One particularly difficult problem in Lecanoraceae systematics
is the delimitation of Lecanora s.str. In the present paper, we
have sidestepped this issue by referring to a well-supported
Lecanora s.lat. and well-supported clades nested within it,
without making a taxonomic judgment about what should be
included in Lecanora s.str. If morphologically distinctive groups
such as Palicella,Pulvinora, and the Lecanora rupicola group
are to be excluded from Lecanora s.str., as various authors have
proposed (Grube et al., 2004;Rodriguez Flakus and Printzen,
2014;Davydov et al., 2021), Lecanora s.str. should ideally be
tied to a well-supported node within our Lecanora s.lat. Our
analysis recovered no such node deeper than the species groups
delimited in Figure 3. For example, the Lecanora subfusca group
as traditionally circumscribed (Lecanora clades I and II; Zhao
et al., 2016) was not well supported in our analysis (Figure 3
and Supplementary Figure 12). Further research, potentially
with genomic data, will be required to stabilize the genus-level
nomenclature of Lecanoraceae.
Based on our results and the existing literature, we
recommend sequencing at a minimum ITS, mtSSU, and RPB1 in
order to place a specimen within the phylogeny of Lecanoraceae.
The new primers and nested PCR protocols we describe in
this paper can be used to sequence RPB1 and RPB2 even from
extractions with a very small quantity of DNA. In this study,
we have substantially increased the number of RPB1,RPB2,
and MCM7 sequences available for Lecanoraceae, which should
facilitate the use of these loci in future phylogenetic studies of
this family. The inclusion of Bryonora in our multilocus dataset
illustrates how the protein-coding loci can clarify relationships in
Lecanoraceae. Bryonora was previously represented on GenBank
by a single ITS sequence from Grube et al. (2004), and its
Frontiers in Microbiology | www.frontiersin.org 11 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 12
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
0
5
10
15
1000 2000 3000 4000 5000
elevation
alpha diversity (number of species)
0.00
0.25
0.50
0.75
1.00
1000 2000 3000 4000
elevation
beta diversity (pairwise Sørensen)
AB
mycobiont
photobiont
FIGURE 6 | Species richness and β-diversity of Lecanoraceae and Trebouxia along the elevation gradient. (A) Number of Lecanoraceae (solid line) and Trebouxia
(dashed line) species observed and inferred for 500 m elevation bins. (B) Pairwise comparisons of 500 m elevation bins by the Sørensen index of dissimilarity, where
0 means all species are shared and 1 means no species are shared. Note that each partner has a mid- to mid-high-elevation threshold not spanned by any species,
but this threshold is not the same for mycobiont and photobiont. Comparisons below 1,500 m (reduced line weight) may be unreliable because of the absence of
samples from the 500–999 m elevation band.
phylogenetic position was unknown. The phylogenetic analysis
of Grube et al. (2004) placed it outside Lecanoraceae, albeit
without support. Our results, which included ITS, nrLSU, mtSSU,
RPB2, and MCM7 sequences for a specimen of Bryonora,
strongly support Bryonora as a distinct genus belonging to
the MPRPS clade.
Although there have been morphological studies on
Lecanoraceae in Bolivia (´
Sliwa et al., 2013, 2014; to a minor
extent, Guderley, 1999), ours is the first study to generate
molecular data for Bolivian Lecanoraceae and one of the only
molecular phylogenetic studies of tropical Lecanoraceae (Kirika
et al., 2012;Papong et al., 2013). These data will be used in
forthcoming publications on the systematics of Lecanoraceae. As
a resource for the community, we have made our Lecanoraceae
and Trebouxia trees and alignments available on the T-BAS
online portal6to facilitate the phylogenetic placement of
specimens from these clades (Carbone et al., 2017, 2019).
Trebouxia Diversity and Ecology
We found Lecanoraceae mycobionts from Bolivia in association
with four of the five major clades of Trebouxia, with only
Trebouxia clade D absent from our sampling. Clades A, I, and
S have been recorded with Lecanoraceae in numerous previous
studies (Beck, 1999;Blaha et al., 2006;Guzow-Krzemi´
nska, 2006;
Hauck et al., 2007;Pérez-Ortega et al., 2012;Ruprecht et al.,
2012, 2020;Leavitt et al., 2015, 2016;Voytsekhovich and Beck,
2016;Singh et al., 2019;Muggia et al., 2020;Wagner et al.,
2020). Lecanoraceae species have not previously been reported
to associate with Trebouxia clade C. Our results show that
species of Lecanora s.lat.—specifically Lecanora clades I–IV
frequently associate with Trebouxia clade C photobionts in low
to mid-elevation habitats (Figure 5A). That this association has
not been seen before is a result of the dearth of molecular
6https://decifr.hpc.ncsu.edu/index.php
phylogenetic studies on tropical Lecanora. Our results reinforce
the importance of Trebouxia clade C photobionts in tropical
lichens. The prevalence of Trebouxia clade C photobionts in
the tropics has been recognized for nearly two decades (Helms,
2003;Cordeiro et al., 2005), but it was not until recently
that the species richness of this clade was well understood
(Muggia et al., 2020).
Several Trebouxia ecological patterns seen in other regions
and with other mycobiont taxa are evident in our results. These
include the association of Trebouxia clade A with calcareous
rock (Helms, 2003) and open, intermittently humid Andean
vegetation (Vargas Castillo and Beck, 2012), as well as the
association of Trebouxia clade S with high elevation (Blaha
et al., 2006;Muggia et al., 2008). In an analysis of over 6,000
publicly available sequences of lichenized Trebouxia, originally
sampled from around the world and associated with diverse
mycobiont hosts, Nelsen et al. (2021) found that Trebouxia
clade I photobionts occupy a bioclimatic space intermediate
between clade C and clade S, while clade A—the most speciose
lineage in Trebouxia—overlaps with all three. Our results
confirm this pattern (Figures 5A,B) and show that it can be
observed on a regional scale within the symbionts of a single
mycobiont family.
Along an elevation gradient from ca. 700–2,100 m in Europe,
Dal Grande et al. (2018) found that Trebouxia photobionts of
the mycobiont genus Lasallia had larger altitudinal ranges at
higher elevations. We saw no evidence for this pattern across the
entire elevation range represented in our sampling (Figure 5B).
It is possible that greater inter-biome variation in climate and
vegetation along our nearly 5,000 m elevation gradient leads to
species turnover that overrides any such pattern.
Our photobiont data included putative new species of
Trebouxia in clades A, C, and I (Supplementary Figures 24). In
clade S, which is globally less diverse than the other three clades
(Muggia et al., 2020), all of our specimens belonged to previously
Frontiers in Microbiology | www.frontiersin.org 12 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 13
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
delimited species (Supplementary Figure 5). Most of the putative
novel species were recovered from a single specimen (C34, C35,
C36, I19, and I20) (Figure 5B). The exception is A53, which
we found in 11 specimens of Myriolecis across four localities
in the humid puna subnival vegetation (Figures 4,5). A53 was
always found in lichen thalli growing on exposed, calcareous rock
(Figure 4).
Several Trebouxia species we found in our sampling were
previously known primarily, or exclusively, from east Africa (see
below). Shared Trebouxia genotypes between Bolivia and Kenya
have been previously reported (Lutsak et al., 2016), highlighting
that long-distance dispersal probably plays a major role in the
distribution of lichenized algae (Rolshausen et al., 2020).
Notes on Selected Trebouxia Species
A01 — We found this species associated with a single
thallus of Protoparmeliopsis garovaglli.Guzow-Krzemi´
nska
(2006) recorded this photobiont in association with P. muralis in
Europe, and Leavitt et al. (2015) found it associated with species
of Protoparmeliopsis,Lecanora,Rhizoplaca, and Xanthoparmelia.
A04 — We recovered Trebouxia A04 from a single specimen
of Bryonora. This is the first photobiont DNA sequence obtained
from a lichen thallus of this mycobiont genus.
A05 — This species was found in association with a species
of Lecanora clade IV from ca. 3,500 m. It has otherwise been
reported from various species of Parmeliaceae: Hypotrachyna and
Punctelia in Kenya (Muggia et al., 2020) and Oropogon in Central
and South America (Leavitt et al., 2015).
A13 — This species is globally distributed (De los Ríos et al.,
2002;Blaha et al., 2006;Guzow-Krzemi´
nska, 2006;Muggia et al.,
2008, 2014;Nyati et al., 2014;Leavitt et al., 2015;Voytsekhovich
and Beck, 2016).
A19 — We provide the first rbcL sequence from this previously
delimited species (Muggia et al., 2020).
A53 — This is the most common novel putative species we
recovered in our sampling (Figure 4). It was always found in
association with species of Myriolecis growing on calcareous rock.
C02 — We provide the first rbcL sequence from this previously
delimited species (Muggia et al., 2020).
C09 — This species was previously reported from multiple
genera of Parmeliaceae in Kenya and from Tephromela in Kenya,
Russia, and Peru (Muggia et al., 2014, 2020). In our sampling,
Trebouxia C09 was found in association with species of Lecanora
s.lat. It is clearly a generalist with respect to mycobiont host and
substrate, occurring with species of Lecanora clade IV on bark
at the lower end of its altitudinal range, then with species from
clades I, II, and III, on bark and on rock, at the upper end of its
range (Figures 4,5B).
C10 — This species was previously reported from multiple
genera of Parmeliaceae in Kenya and Parmotrema tinctorum in
Japan (Ohmura et al., 2006;Muggia et al., 2020). In our sampling,
Trebouxia C10 was only found in association with species of
Lecanora clades I and II (i.e., the Lecanora subfusca group)
(Figure 4).
C13 — This species is known from three specimens of
Parmotrema and Canoparmelia from Kenya (Muggia et al., 2020).
In our sampling, Trebouxia C13 was only found in association
with species of Lecanora clades I and II (i.e., the Lecanora
subfusca group) (Figure 4).
C34 — This novel putative species was also found by Kosecka
et al. (2021b). See discussion below.
C35 and C36 — These OTUs form a novel lineage sister to
the rest of Trebouxia clade C (Supplementary Figure 3). They
were both found in association with Lecanora ecoronata 1 in the
low-elevation savanna. The phylogenetic position of these species
is supported by both ITS and rbcL, but additional investigations
of the ultrastructure of these species should be conducted to
evaluate whether they would be better treated as a new major
clade (i.e., if they do not have a corticola-type pyrenoid).
I13 s.lat. — We found several specimens associated with
Trebouxia I13 but, despite having both ITS and rbcL data, the
clade is not well supported (Supplementary Figure 4). This
region of the Trebouxia clade I tree, which also includes I17, I18,
I19, and I20, appears to be primarily tropical (Muggia et al., 2020)
and requires further phylogenetic study.
I15 — We provide the first rbcL sequence from this previously
delimited species. Trebouxia I15 was reported from Punctelia and
Hypotrachyna in Kenya (Leavitt et al., 2015;Muggia et al., 2020).
I18 — This species was reported from Hypotrachyna in Peru
and Kenya and Tephromela in Peru (Muggia et al., 2014, 2020).
I19 and I20 — These novel putative species were also found by
Kosecka et al. (2021b). See discussion below.
S02 — This species is a generalist with respect to mycobiont
host, associating with species of multiple genera within both the
Lecanora s.lat. and MPRPS clades.
S10 — This OTU corresponds to the formally described
species Trebouxia simplex (Tschermak-Woess, 1978), the type
of which was originally isolated from Chaenotheca in Europe,
and has also been reported from Bryoria in North America and
Europe (Lindgren et al., 2014) and Lasallia in Europe (Sadowska-
De´
s et al., 2014). Our results confirm that S10/T. simplex is
a generalist with respect to mycobiont host, associating with
species from both the Lecanora s.lat and MPRPS clades.
Multiple Photobionts in a Single Thallus
Although intra-thalline genetic diversity has been documented
in lichenized Trebouxia (e.g., Mansournia et al., 2012;Dal
Grande et al., 2014), Sanger sequencing of PCR amplicons
has been shown to consistently identify the most abundant
photobiont genotype (Paul et al., 2018). Furthermore, when
multiple photobiont genotypes are present, they are generally
closely related, belonging to the same species (Mansournia et al.,
2012;Dal Grande et al., 2014), sister taxa (Casano et al., 2011), or
at least the same group within Trebouxia (Paul et al., 2018).
For one specimen (Flakus 29597), Sanger sequencing
recovered two different ITS sequences with the forward and
reverse primers, corresponding to the photobionts I18 and S02.
Repeated PCR and sequencing of this specimen generated the
same result: two clean chromatograms with different sequences.
The rbcL sequence obtained from this specimen was consistent
with I18, so we used that photobiont in the figures and
analyses, although S02 is also consistent with the ecology of this
specimen. A technique such as fluorescence in situ hybridization
Frontiers in Microbiology | www.frontiersin.org 13 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 14
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
(FISH) would be required to conclusively demonstrate that both
photobionts occur in the same lichen thallus.
Specimens for which the photobiont ITS PCR was successful
(as determined by a single clear band on an agarose gel) but
sequencing failed because the chromatograms were unusable
due to double peaks might plausibly represent cases where
multiple algal species were present in the thallus. Likewise,
cases where isolated double peaks occurred in otherwise high-
quality chromatograms may indicate the presence of multiple
algal haplotypes in a single thallus. Further investigation
with next-generation sequencing would be required to verify
these conjectures.
Photobiont Switching Within a
Mycobiont Species
Photobiont switching within a mycobiont species can occur at
multiple phylogenetic levels: between photobiont genera (Ertz
et al., 2018;Vanˇ
curová et al., 2018), between major clades within
a photobiont genus (Blaha et al., 2006), between photobiont
species within a subgeneric clade (Magain et al., 2017, 2018;
Dal Grande et al., 2018), or between haplotypes within a
photobiont species (Gasulla et al., 2020). We found no evidence
of switching at the genus level, reaffirming that Lecanoraceae
mycobionts are strict specialists on Trebouxia (Miadlikowska
et al., 2006). Conversely, we found evidence for all three of the
intrageneric types of switching, suggesting that the specificity
of these mycobionts for particular algal species may be lower
than commonly thought. It should be emphasized, however,
that our sampling included few thalli from any one mycobiont
species and is therefore not ideal for studying photobiont
interactions within a mycobiont species. The conclusions drawn
from the instances of switching we did observe should be
considered preliminary and should be tested with a more
focused sampling.
One of the mycobiont species we found with photobionts
from two different Trebouxia clades was Lecanora rupicola
(Figure 4). Blaha et al. (2006) sequenced the photobiont for
specimens of L. rupicola from across Europe. They found that
this species associated with diverse photobionts from three
of the main clades of Trebouxia, including in some cases
multiple photobiont species at a single site. The mycobiont was
not sequenced in that study, so we do not know what role
population structure or cryptic diversity may have played in the
observed patterns, although the authors noted that there was
no correlation between variation in secondary chemistry and
photobiont identity.
All of the mycobiont species associated with photobionts from
multiple Trebouxia clades occurred on siliceous rocks in open
areas at mid-high to high elevation (Figure 4). The same was true
for many of the mycobiont species that partnered with multiple
photobionts from a single Trebouxia clade (Figure 4). This is
consistent with previous work that has shown that mycobiont–
photobiont associations in lichens are more specialized in
warmer climates (Singh et al., 2017). One hypothesis to explain
this pattern is that there are simply more photobiont species,
and therefore greater opportunity for partner switching, at
high elevation (Figure 6A). However, many of the instances
of switching we observed were related to elevation change
(Figure 5), lending support to the alternate hypothesis that
mycobionts switch between locally-adapted photobionts along
their range (Muggia et al., 2014;Dal Grande et al., 2018).
Turnover of Mycobionts and Photobiont
Communities
Our results add to a growing body of literature showing that
elevation gradients have a substantial effect on the composition of
lichen photobiont communities (Dal Grande et al., 2018;Devkota
et al., 2019;Gasulla et al., 2020;Rolshausen et al., 2020;Wagner
et al., 2020). This literature is part of a broader realization that
microbial diversity is subject to some of the same biogeographic
influences as macroscopic organisms (e.g., Nottingham et al.,
2018; but see Bryant et al., 2008). For example, various taxa of
free-living cyanobacteria and green algae in the soil may vary
in their abundance along elevation gradients (ˇ
Reháková et al.,
2011;Stewart et al., 2021). The elevation difference between the
lowest and highest sites in our dataset—nearly 5,000 m, greater
than any other study of this type—presents a unique opportunity
to study mycobiont and photobiont turnover accompanying
the transition between geographically close but ecologically
distant biomes (Figure 7). In contrast to previous studies where
turnover of mycobiont and photobiont communities occurred
at different spatial scales (Dal Grande et al., 2018;Lu et al.,
2018;Gasulla et al., 2020;Rolshausen et al., 2020), we observed
turnover in Lecanoraceae communities and their associated
Trebouxia photobionts at roughly equivalent rates (Figures 5,6).
However, because there are no mycobiont species that occur
along the entire length of the gradient, we cannot tease apart the
impact of substrate, climate, and mycobiont host on photobiont
distribution to the degree possible in other studies (e.g., Dal
Grande et al., 2018). For example, species of Myriolecis are
almost always found on calcareous rock at high elevation with
Trebouxia clade A.
Our findings on species richness and β-diversity along the
elevation gradient are consistent with previously observed
patterns and also suggest phenomena which should be
investigated in a larger sampling replicated across multiple
transects. The finding that, within each elevation bin, Trebouxia
species were less diverse than Lecanoraceae (Figure 6A) recalls
the conventional wisdom that natural communities contain
more mycobiont species than photobiont species (e.g., Singh
et al., 2019;Wagner et al., 2020). The one exception at 3,500
m, where the number of mycobiont and photobiont species are
equal, corresponds to a poorly-sampled elevation band and may
be an artifact (Figure 5).
Our results for β-diversity suggest an interesting
phenomenon: Along an elevation gradient, species turnover
occurs for both partners in a mutualistic interaction, but the
elevation thresholds at which community composition shifts
most dramatically are different for the two partners (Figure 6B).
This pattern has not been seen in similar studies of smaller
elevation gradients and fewer mycobiont taxa (e.g., Dal Grande
et al., 2018). These thresholds may be equivalent to the “symbiont
turnover zones” described by Rolshausen et al. (2020), but while
Frontiers in Microbiology | www.frontiersin.org 14 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 15
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
1000
2000
3000
4000
elevation
elevation
vegetation
MAT (BIO1)
MAP (BIO12)
500
1000
1500
MAP (mm)
0.5
10
15
20
25
MAT (°C)
N
600 km
22qS
20qS
18qS
16qS
14qS
12qS
10qS
70qW 68qW66qW64qW62qW60qW 58qW
vegetation
Beni savanna
Chaceño−Amazon forest
Tucumano−Boliviano forest
Yungas forest
Humid Puna high-Andean
Humid Puna subnival
Other
FIGURE 7 | Hierarchical clustering dendrogram of 19 bioclimatic variables for the sampling localities. Sampling locations and vegetation types are displayed on the
inset map. MAT, mean annual temperature (WorldClim variable BIO1). MAP, mean annual precipitation (WorldClim variable BIO12).
those authors found a turnover zone for one partner, we show
turnover zones for both partners and, crucially, these are at
widely spaced elevations. No Lecanoraceae species are found
both above and below 2,500 m, while no Trebouxia species
are found on both sides of 3,500 m. The former elevation is
the threshold at which rock begins to appear as a substrate
for Lecanoraceae, and is also the transition point between
aseasonal Yungas forest and seasonal Tucumano-Boliviano
forest (Figures 4,7). The latter elevation corresponds to the
transition from Trebouxia clades C and I to Trebouxia clades A
and S (Figure 5B); this is also the threshold at which open, rocky
habitats fully replace forest vegetation (Figure 7). Perhaps not
coincidentally, 3,500 m is also very close to a similar threshold
for woody plant ranges in the Bolivian Andes (Tello et al., 2015).
These findings suggest two alternative hypotheses. One
possibility is that mycobiont and photobiont ranges are
responding to different biotic or abiotic factors (e.g., substrate
versus temperature; rainfall versus the presence of pathogens).
Another possibility is that both thresholds are a response to the
same factor, but the two partners are responding differently. Lu
et al. (2018) found that ranges of Peltigera species and Nostoc
phylogroups did not have the same response to climate shifts
along a latitudinal gradient, but their study was at an intra-biome
scale and lacked the sharp community discontinuities we see here.
While our sampling is not adequate to distinguish between these
hypotheses, it provides guidance as to an ideal region and study
design that would have the power to resolve this question.
Phylogenetically Versus Ecologically
Designed Sampling to Study Photobiont
Communities
Kosecka et al. (2021b) examined the community of Trebouxia
photobionts along a largely identical elevation gradient in Bolivia.
That study included a four- to five-fold larger sample size
from multiple families of Trebouxia-associated Lecanoromycetes,
including some Lecanoraceae, but only dealt with the photobiont
sequences from these specimens. The results of Kosecka et al.
(2021b) in terms of Trebouxia diversity and distribution are
largely consistent with our own findings, from the broad pattern
of turnover among major clades of Trebouxia along the elevation
gradient to many of the individual Trebouxia species found. Our
study recovered several novel lineages not found by Kosecka et al.
(2021b)—A53, C35, and C36; conversely, and commensurate
with their larger sample size, Kosecka et al. (2021b) found
several putative Trebouxia species which were missing from our
sampling. Nevertheless, the close correspondence between the
two studies suggests that dense sampling of a single mycobiont
family may be sufficient to reveal regional patterns of biodiversity
in lichenized Trebouxia. This stands in contrast to the situation
with cyanolichens, where the Nostoc community will look very
different if, for example, one samples from only Collemataceae,
Peltigeraceae, or Lobariaceae in the same geographic area
(Magain et al., 2018). Studies of Trebouxia photobionts from
a single mycobiont clade are the norm in the literature, but
it has not been clear whether their results can be generalized
to represent the entire Trebouxia community of a region. Our
results suggest that the answer may be yes, as long as the focal
mycobiont clade occurs in a sufficiently diverse range of micro-
and macrohabitats.
DATA AVAILABILITY STATEMENT
The datasets presented in this study can be found in online
repositories. New sequence data generated for this study are
available on GenBank with accession numbers OK665489–
OK665671, OL603980–OL604141, OL625025–OL625113, and
OL663852–OL663919. Other data are included in the online
Supplementary Material for this article.
Frontiers in Microbiology | www.frontiersin.org 15 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 16
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
AUTHOR CONTRIBUTIONS
FL, JM, and L´
S contributed to the conception and design of
the study and supervised Ph.D. students. L´
S and JM secured
funding. AF and PR-F collected the specimens and ecological
data in the field. EM and L´
S studied specimen morphology and
delimited morphospecies. EM and EC prepared DNA extractions.
IM designed primers for PCR and sequencing and generated
the figures. EM, IM, and CP-DH collected the molecular data.
IM and EM curated, validated, and analyzed the data and wrote
the original draft of the manuscript. All authors participated in
review and editing of the final manuscript and approved the
submitted version.
FUNDING
Funding for this research was provided by the National
Science Centre, Poland, under a grant to L´
S (project
no. 2016/21/B/NZ8/02463). IM was supported by a
US National Science Foundation Graduate Research
Fellowship (DGE 1644868).
ACKNOWLEDGMENTS
AF and PR-F are greatly indebted to the staff of the
Herbario Nacional de Bolivia, Instituto de Ecología, Universidad
Mayor de San Andrés, La Paz, for their cooperation and
to the Servicio Nacional de Áreas Protegidas (SERNAP) and
all protected area staff for providing permits for scientific
studies, as well as assistance and logistical support during field
work. We thank the two reviewers for their helpful feedback
on the manuscript.
SUPPLEMENTARY MATERIAL
The Supplementary Material for this article can be found
online at: https://www.frontiersin.org/articles/10.3389/fmicb.
2021.774839/full#supplementary-material
REFERENCES
Beck, A. (1999). Photobiont inventory of a lichen community growing on
heavy-metal-rich rock. Lichenologist 31, 501–510. doi: 10.1006/lich.1999.
0232
Bengtsson-Palme, J., Ryberg, M., Hartmann, M., Branco, S., Wang, Z., Godhe, A.
et al. (2013). Improved software detection and extraction of ITS1 and ITS2
from ribosomal ITS sequences of fungi and other eukaryotes for analysis of
environmental sequencing data. Methods Ecol. Evol. 4, 914–919. doi: 10.1111/
2041-210X.12073
Blaha, J., Baloch, E., and Grube, M. (2006). High photobiont diversity
associated with the euryoecious lichen-forming ascomycete Lecanora rupicola
(Lecanoraceae, Ascomycota). Biol. J. Linn. Soc. 88, 283–293. doi: 10.1111/j.
1095-8312.2006.00640.x
Brodo, I. M., Sharnoff, S. D., and Sharnoff, S. (2001). Lichens of North America.
New Haven, CT: Yale University Press.
Bryant, J. A., Lamanna, C., Morlon, H., Kerkhoff, A. J., Enquist, B. J.,
and Green, J. L. (2008). Microbes on mountainsides: contrasting
elevational patterns of bacterial and plant diversity. Proc. Natl. Acad.
Sci. U.S.A. 105(Suppl. 1), 11505–11511. doi: 10.1073/pnas.080192
0105
Carbone, I., White, J. B., Miadlikowska, J., Arnold, A. E., Miller, M. A., Kauff, F.,
et al. (2017). T-BAS: Tree-Based Alignment Selector toolkit for phylogenetic-
based placement, alignment downloads, and metadata visualization: an example
with the Pezizomycotina tree of life. Bioinformatics 33, 1160–1168. doi: 10.1093/
bioinformatics/btw808
Carbone, I., White, J. B., Miadlikowska, J., Arnold, A. E., Miller, M. A., Magain,
N., et al. (2019). T-BAS version 2.1: Tree-Based Alignment Selector toolkit for
evolutionary placement and viewing of alignments and metadata on curated
and custom trees. Microbiol. Resour. Announce. 8:e00328-19. doi: 10.1128/
MRA.00328-19
Casano, L. M., del Campo, E. M., García-Breijo, F. J., Reig-Armiñana, J.,
Gasulla, F., del Hoyo, A., et al. (2011). Two Trebouxia algae with different
physiological performances are ever-present in lichen thalli of Ramalina
farinacea. Coexistence versus competition? Environ. Microbiol. 13, 806–818.
doi: 10.1111/j.1462-2920.2010.02386.x
Chagnon, P. L., Magain, N., Miadlikowska, J., and Lutzoni, F. (2019). Species
diversification and phylogenetically constrained symbiont switching generated
high modularity in the lichen genus Peltigera.J. Ecol. 107, 1645–1661. doi:
10.1111/1365-2745.13207
Chernomor, O., Von Haeseler, A., and Minh, B. Q. (2016). Terrace aware data
structure for phylogenomic inference from supermatrices. Syst. Biol. 65, 997–
1008. doi: 10.1093/sysbio/syw037
Cordeiro, L. M. C., Reis, R. A., Cruz, L. M., Stocker-Wörgötter, E., Grube, M.,
and Iacomini, M. (2005). Molecular studies of photobionts of selected lichens
from the coastal vegetation of Brazil. FEMS Microbiol. Ecol. 54, 381–390. doi:
10.1016/j.femsec.2005.05.003
Culberson, C. F., and Kristinsson, H. (1970). A standardized method for the
identification of lichen products. J. Chromatogr. 46, 85–93. doi: 10.1016/S0021-
9673(00)83967-9
Dal Grande, F., Alors, D., Divakar, P. K., Bálint, M., Crespo, A., and Schmitt, I.
(2014). Insights into intrathalline genetic diversity of the cosmopolitan lichen
symbiotic green alga Trebouxia decolorans Ahmadjian using microsatellite
markers. Mol. Phylogenet. Evol. 72, 54–60. doi: 10.1016/j.ympev.2013.12.010
Dal Grande, F., Rolshausen, G., Divakar, P. K., Crespo, A., Otte, J., Schleuning, M.,
et al. (2018). Environment and host identity structure communities of green
algal symbionts in lichens. New Phytol. 217, 277–289. doi: 10.1111/nph.14770
Davydov, E. A., Yakovchenko, L. S., Hollinger, J., Bungartz, F., Parrinello,
C., and Printzen, C. (2021). The new genus Pulvinora (Lecanoraceae)
for species of the ‘Lecanora pringlei’ group, including the new species
Pulvinora stereothallina.Bryologist 124, 242–256. doi: 10.1639/0007-2745-
124.2.242
De los Ríos, A., Ascaso, C., and Grube, M. (2002). An ultrastructural, anatomical
and molecular study of the lichenicolous lichen Rimularia insularis.Mycol. Res.
106, 946–953. doi: 10.1017/S0953756202006238
Devkota, S., Chaudhary, R. P., Werth, S., and Scheidegger, C. (2019). Genetic
diversity and structure of the epiphytic foliose lichen Lobaria pindarensis in
the Himalayas depends on elevation. Fungal Ecol. 41, 245–255. doi: 10.1016/
j.funeco.2019.07.002
Döring, H., Clerc, P., Grube, M., and Wedin, M. (2000). Mycobiont-specific
PCR primers for the amplification of nuclear ITS and LSU rDNA from
lichenized ascomycetes. Lichenologist 32, 200–204. doi: 10.1006/lich.1999.
0250
Ertz, D., Guzow-Krzemi´
nska, B., Thor, G., Łubek, A., and Kukwa, M. (2018).
Photobiont switching causes changes in the reproduction strategy and
phenotypic dimorphism in the Arthoniomycetes. Sci. Rep. 8:4952. doi: 10.1038/
s41598-018-23219-3
Fick, S. E., and Hijmans, R. J. (2017). WorldClim 2: new 1-km spatial resolution
climate surfaces for global land areas. Int. J. Climatol. 37, 4302–4315. doi:
10.1002/joc.5086
Frontiers in Microbiology | www.frontiersin.org 16 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 17
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
Gardes, M., and Bruns, T. D. (1993). ITS primers with enhanced specificity for
basidiomycetes—application to the identification of mycorrhizae and rusts.
Mol. Ecol. 2, 113–118. doi: 10.1111/j.1365-294X.1993.tb00005.x
Gasulla, F., Barrasa, J. M., Casano, L. M., and del Campo, E. M. (2020).
Symbiont composition of the basidiolichen Lichenomphalia meridionalis varies
with altitude in the Iberian Peninsula. Lichenologist 52, 17–26. doi: 10.1017/
S002428291900046X
Grube, M., Baloch, E., and Arup, U. (2004). A phylogenetic study of the Lecanora
rupicola group (Lecanoraceae, Ascomycota). Mycol. Res. 108, 506–514. doi:
10.1017/S0953756204009888
Guderley, R. (1999). Die Lecanora subfusca-gruppe in Süd-und Mittelamerika.
J. Hattori Bot. Lab. 87, 131–257.
Guzow-Krzemi´
nska, B. (2006). Photobiont flexibility in the lichen
Protoparmeliopsis muralis as revealed by ITS rDNA analyses. Lichenologist 38,
469–476. doi: 10.1017/S0024282906005068
Hauck, M., Helms, G., and Friedl, T. (2007). Photobiont selectivity in the epiphytic
lichens Hypogymnia physodes and Lecanora conizaeoides.Lichenologist 39,
195–204. doi: 10.1017/S0024282907006639
Helms, G. (2003). Taxonomy and Symbiosis in Associations of Physciaceae and
Trebouxia. Ph.D. dissertation. Göttingen: Georg-August Universität Göttingen.
Hestmark, G., Lutzoni, F., and Miadlikowska, J. (2016). Photobiont associations
in co-occurring umbilicate lichens with contrasting modes of reproduction
in coastal Norway. Lichenologist 48, 545–557. doi: 10.1017/S002428291600
0232
Hijmans, R. J. (2021). raster: Geographic Data Analysis and Modeling. R Package
Version 3.4-13. Available online at: https://CRAN.R-project.org/package=raster
Hoang, D. T., Chernomor, O., Von Haeseler, A., Minh, B. Q., and Vinh, L. S. (2018).
UFBoot2: improving the ultrafast Bootstrap approximation. Mol. Biol. Evol. 35,
518–522. doi: 10.1093/molbev/msx281
James, P. W., and Henssen, A. (1976). “The morphological and taxonomical
significance of cephalodia,” in Lichenology: Progress and Problems, eds D. H.
Brown, D. L. Hawksworth, and R. H. Bailey (London: Academic Press), 27–77.
Jüriado, I., Kaasalainen, U., Jylhä, M., and Rikkinen, J. (2019). Relationships
between mycobiont identity, photobiont specificity and ecological preferences
in the lichen genus Peltigera (Ascomycota) in Estonia (northeastern Europe).
Fungal Ecol. 39, 45–54. doi: 10.1016/j.funeco.2018.11.005
Kalyaanamoorthy, S., Minh, B. Q., Wong, T. K., Von Haeseler,A., and Jermiin, L. S.
(2017). ModelFinder: fast model selection for accurate phylogenetic estimates.
Nat. Methods 14, 587–589. doi: 10.1038/nmeth.4285
Katoh, K., Rozewicki, J., and Yamada, K. D. (2019). MAFFT online service:
multiple sequence alignment, interactive sequence choice and visualization.
Brief. Bioinform. 20, 1160–1166. doi: 10.1093/bib/bbx108
Kirika, P., Parnmen, S., and Lumbsch, T. (2012). Two new species of Lecanora
sensu stricto (Lecanoraceae, Ascomycota) from east Africa. MycoKeys 3, 37–47.
doi: 10.3897/mycokeys.3.3201
Kosecka, M., Guzow-Krzemi´
nska, B., ˇ
Cernajová, I., Škaloud, P., Jabło´
nska, A., and
Kukwa, M. (2021a). New lineages of photobionts in Bolivian lichens expand
our knowledge on habitat preferences and distribution of Asterochloris algae.
Sci. Rep. 11:8701. doi: 10.1038/s41598-021- 88110-0
Kosecka, M., Guzow-Krzemi´
nska, B., Jabło´
nska, A., Flakus, A., Rodriguez-
Flakus, P., and Kukwa, M. (2021b). “Adaptation strategies of Bolivian lichens:
biodiversity and ecology of Trebouxia photobionts,” in Proceedings of the IAL9
Program and Abstract Book, 82.
Kosecka, M., Jabło´
nska, A., Flakus, A., Rodriguez-Flakus, P., Kukwa, M.,
and Guzow-Krzemi´
nska, B. (2020). Trentepohlialean algae (Trentepohliales,
Ulvophyceae) show preference to selected mycobiont lineages in lichen
symbioses. J. Phycol. 56, 979–993. doi: 10.1111/jpy.12994
Kroken, S., and Taylor, J. W. (2000). Phylogenetic species, reproductive mode, and
specificity of the green alga Trebouxia forming lichens with the fungal genus
Letharia.Bryologist 103, 645–660.
Leavitt, S. D., Kraichak, E., Nelsen, M. P., Altermann, S., Divakar, P. K., Alors,
D., et al. (2015). Fungal specificity and selectivity for algae play a major role
in determining lichen partnerships across diverse ecogeographic regions in the
lichen-forming family Parmeliaceae (Ascomycota). Mol. Ecol. 24, 3779–3797.
doi: 10.1111/mec.13271
Leavitt, S. D., Kraichak, E., Vondrak, J., Nelsen, M. P., Sohrabi, M., Perez-Ortega, S.,
et al. (2016). Cryptic diversity and symbiont interactions in rock-posy lichens.
Mol. Phylogenet. Evol. 99, 261–274. doi: 10.1016/j.ympev.2016.03.030
Lindgren, H., Velmala, S., Hognabba, F., Goward, T., Holien, H., and Myllys,
L. (2014). High fungal selectivity for algal symbionts in the genus Bryoria.
Lichenologist 46, 681–695. doi: 10.1017/S0024282914000279
Liu, Y. J., Whelen, S., and Hall, B. D. (1999). Phylogenetic relationships among
ascomycetes: evidence from an RNA polymerase II subunit. Mol. Biol. Evol. 16,
1799–1808. doi: 10.1093/oxfordjournals.molbev.a026092
Lu, J., Magain, N., Miadlikowska, J., Coyle, J. R., Truong, C., and Lutzoni, F. (2018).
Bioclimatic factors at an intrabiome scale are more limiting than cyanobiont
availability for the lichen-forming genus Peltigera.Am. J. Bot. 105, 1198–1211.
doi: 10.1002/ajb2.1119
Lücking, R., Hodkinson, B. P., and Leavitt, S. D. (2017). The 2016 classification
of lichenized fungi in the Ascomycota and Basidiomycota–approaching one
thousand genera. Bryologist 119, 361–416. doi: 10.1639/0007-2745-119.4.361
Lutsak, T., Fernández-Mendoza, F., Kirika, P., Wondafrash, M., and Printzen, C.
(2016). Mycobiont-photobiont interactions of the lichen Cetraria aculeata in
high alpine regions of East Africa and South America. Symbiosis 68, 25–37.
doi: 10.1007/s13199-015- 0351-1
Lutzoni, F., Wagner, P., Reeb, V., and Zoller, S. (2000). Integrating ambiguously
aligned regions of DNA sequences in phylogenetic analyses without
violating positional homology. Syst. Biol. 49, 628–651. doi: 10.1080/
106351500750049743
Maddison, W. P., and Maddison, D. R. (2018). Mesquite: A Modular System
for Evolutionary Analysis. Version 3.51. Available online at: http://www.
mesquiteproject.org
Magain, N., Miadlikowska, J., Goffinet, B., Sérusiaux, E., and Lutzoni, F. (2017).
Macroevolution of specificity in cyanolichens of the genus Peltigera section
Polydactylon (Lecanoromycetes, Ascomycota). Syst. Biol. 66, 74–99. doi: 10.
1093/sysbio/syw065
Magain, N., Truong, C., Goward, T., Niu, D., Goffinet, B., Sérusiaux, E., et al.
(2018). Species delimitation at a global scale reveals high species richness with
complex biogeography and patterns of symbiont association in Peltigera section
Peltigera (lichenized Ascomycota: Lecanoromycetes). Taxon 67, 836–870. doi:
10.12705/675.3
Mansournia, M. R., Bingyun, W. U., Matsushita, N., and Hogetsu, T.
(2012). Genotypic analysis of the foliose lichen Parmotrema tinctorum
using microsatellite markers: association of mycobiont and photobiont,
and their reproductive modes. Lichenologist 44, 419–440. doi: 10.1017/
S0024282911000909
Matheny, P. B. (2005). Improving phylogenetic inference of mushrooms with RPB1
and RPB2 nucleotide sequences (Inocybe; Agaricales). Mol. Phylogenet. Evol. 35,
1–20. doi: 10.1016/j.ympev.2004.11.014
Matheny, P. B., Liu, Y. J., Ammirati, J. F., and Hall, B. D. (2002). Using RPB1
sequences to improve phylogenetic inference among mushrooms (Inocybe,
Agaricales). Am. J. Bot. 89, 688–698. doi: 10.3732/ajb.89.4.688
Miadlikowska, J., Kauff, F., Hofstetter, V., Fraker, E., Grube, M., Hafellner, J., et al.
(2006). New insights into classification and evolution of the Lecanoromycetes
(Pezizomycotina, Ascomycota) from phylogenetic analyses of three ribosomal
RNA-and two protein-coding genes. Mycologia 98, 1088–1103. doi: 10.1080/
15572536.2006.11832636
Miadlikowska, J., Kauff, F., Högnabba, F., Oliver, J. C., Molnár, K., Fraker, E.,
et al. (2014). Multigene phylogenetic synthesis for the class Lecanoromycetes
(Ascomycota): 1307 fungi representing 1139 infrageneric taxa, 312 genera and
66 families. Mol. Phylogenet. Evol. 79, 132–168. doi: 10.1016/j.ympev.2014.
04.003
Miller, M. A., Pfeiffer, W., and Schwartz, T. (2010). “Creating the CIPRES science
gateway for inference of large phylogenetic trees,” in Proceedings of the Gateway
Computing Environments Workshop (GCE), 14 Nov. 2010, New Orleans, LA,
1–8.
Muggia, L., Grube, M., and Tretiach, M. (2008). Genetic diversity and photobiont
associations in selected taxa of the Tephromela atra group (Lecanorales,
lichenised Ascomycota). Mycol. Prog. 7, 147–160. doi: 10.1007/s11557-008-
0560-6
Muggia, L., Leavitt, S. D., and Barreno, E. (2018). The hidden diversity of lichenised
Trebouxiophyceae (Chlorophyta). Phycologia 57, 503–524. doi: 10.2216/17-
134.1
Muggia, L., Nelsen, M. P., Kirika, P. M., Barreno, E., Beck, A., Lindgren,
H., et al. (2020). Formally described species woefully underrepresent
phylogenetic diversity in the common lichen photobiont genus Trebouxia
Frontiers in Microbiology | www.frontiersin.org 17 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 18
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
(Trebouxiophyceae, Chlorophyta): an impetus for developing an integrated
taxonomy. Mol. Phylogenet. Evol. 149:106821. doi: 10.1016/j.ympev.2020.
106821
Muggia, L., Pérez-Ortega, S., Kopun, T., Zellnig, G., and Grube, M. (2014).
Photobiont selectivity leads to ecological tolerance and evolutionary divergence
in a polymorphic complex of lichenized fungi. Ann. Bot. 114, 463–475. doi:
10.1093/aob/mcu146
Navarro, G., and Ferreira, W. (2007). Mapa de Vegetación de Bolivia, Escala
1: 250000. Edición CD-ROM. RUMBOL SRL. Depósito Legal 2-7-116-11.
Edición auspiciada por The Nature Conservancy (TNC), CONDESAN, The
Natureserve.
Navarro, G., and Maldonado, M. (2002). Geografía Ecológica de Bolivia: Vegetación
y Ambientes Acuáticos. Cochabamba: Centro de Ecología Simón I. Patiño,
Departamento de Difusión.
Nelsen, M. P., Leavitt, S. D., Heller, K., Muggia, L., and Lumbsch, H. T.
(2021). Macroecological diversification and convergence in a clade of
keystone symbionts. FEMS Microbiol. Ecol. 97:fiab072. doi: 10.1093/femsec/
fiab072
Nelsen, M. P., Rivas Plata, E., Andrew, C. J., Lücking, R., and Lumbsch, H. T.
(2011). Phylogenetic diversity of Trentepohlialean algae associated with lichen-
forming fungi. J. Phycol. 47, 282–290.
Nguyen, L. T., Schmidt, H. A., Von Haeseler, A., and Minh, B. Q. (2015). IQ-TREE:
a fast and effective stochastic algorithm for estimating maximum-likelihood
phylogenies. Mol. Biol. Evol. 32, 268–274. doi: 10.1093/molbev/msu300
Nottingham, A. T., Fierer, N., Turner, B. L., Whitaker, J., Ostle, N. J., McNamara,
N. P., et al. (2018). Microbes follow Humboldt: temperature drives plant and
soil microbial diversity patterns from the Amazon to the Andes. Ecology 99,
2455–2466. doi: 10.1002/ecy.2482
Nyati, S., Scherrer, S., Werth, S., and Honegger, R. (2014). Green-algal
photobiont diversity (Trebouxia spp.) in representatives of Teloschistaceae
(Lecanoromycetes, lichen-forming ascomycetes). Lichenologist 46, 189–212.
doi: 10.1017/S0024282913000819
Ohmura, Y., Kawachi, M., Kasai, F., Watanabe, M. M., and Takeshita, S.
(2006). Genetic combinations of symbionts in a vegetatively reproducing
lichen, Parmotrema tinctorum, based on ITS rDNA sequences. Bryologist 109,
43–59.
Oksanen, J., Blanchet, F. G., Friendly, M., Kindt, R., Legendre, P., McGlinn, D., et al.
(2020). vegan: Community Ecology Package. R Package Version 2.5-7. Available
online at: https://CRAN.R-project.org/package=vegan
Orange, A., James, P. W., and White, F. J. (2001). Microchemical Methods for the
Identification of Lichens. London: British Lichen Society.
Papong, K., Boonpragob, K., Parnmen, S., and Lumbsch, H. T. (2013).
Molecular phylogenetic studies on tropical species of Lecanora sensu stricto
(Lecanoraceae, Ascomycota). Nova Hedwigia 96, 1–13. doi: 10.1127/0029-5035/
2012/0072
Paradis, E. (2010). pegas: an R package for population genetics with an integrated–
modular approach. Bioinformatics 26, 419–420. doi: 10.1093/bioinformatics/
btp696
Paradis, E., and Schliep, K. (2019). ape 5.0: an environment for modern
phylogenetics and evolutionary analyses in R. Bioinformatics 35, 526–528. doi:
10.1093/bioinformatics/bty633
Paul, F., Otte, J., Schmitt, I., and Dal Grande, F. (2018). Comparing Sanger
sequencing and high-throughput metabarcoding for inferring photobiont
diversity in lichens. Sci. Rep. 8:8624. doi: 10.1038/s41598-018-26947-8
Pebesma, E. J., and Bivand, R. S. (2005). Classes and methods for spatial data in R.
R News 5, 9–13.
Peksa, O., and Škaloud, P. (2011). Do photobionts influence the ecology of lichens?
A case study of environmental preferences in symbiotic green alga Asterochloris
(Trebouxiophyceae). Mol. Ecol. 20, 3936–3948. doi: 10.1111/j.1365-294X.2011.
05168.x
Pérez-Ortega, S., Ortiz-Álvarez, R., Allan Green, T. G., and de Los Ríos, A. (2012).
Lichen myco-and photobiont diversity and their relationships at the edge of
life (McMurdo Dry Valleys, Antarctica). FEMS Microbiol. Ecol. 82, 429–448.
doi: 10.1111/j.1574-6941.2012.01422.x
Pérez-Ortega, S., Spribille, T., Palice, Z., Elix, J. A., and Printzen, C. (2010). A
molecular phylogeny of the Lecanora varia group, including a new species from
western North America. Mycol. Prog. 9, 523–535. doi: 10.1007/s11557-010-
0660-y
Puillandre, N., Lambert, A., Brouillet, S., and Achaz, G. (2012). ABGD, Automatic
Barcode Gap Discovery for primary species delimitation. Mol. Ecol. 21, 1864–
1877. doi: 10.1111/j.1365-294X.2011.05239.x
R Core Team (2019). R: A Language and Environment for Statistical Computing.
Vienna: R Foundation for Statistical Computing.
ˇ
Reháková, K., Chlumská, Z., and Doležal, J. (2011). Soil cyanobacterial and
microalgal diversity in dry mountains of Ladakh, NW Himalaya, as related to
site, altitude, and vegetation. Microb. Ecol. 62, 337–346. doi: 10.1007/s00248-
011-9878-8
Rodriguez Flakus, P., and Printzen, C. (2014). Palicella, a new genus of lichenized
fungi and its phylogenetic position within Lecanoraceae. Lichenologist 46,
535–552. doi: 10.1017/S0024282914000127
Rolshausen, G., Hallman, U., Dal Grande, F., Otte, J., Knudsen, K., and Schmitt,
I. (2020). Expanding the mutualistic niche: parallel symbiont turnover along
climatic gradients. Proc. R. Soc. B 287:20192311. doi: 10.1098/rspb.2019.2311
Ruprecht, U., Brunauer, G., and Printzen, C. (2012). Genetic diversity of
photobionts in Antarctic lecideoid lichens from an ecological view point.
Lichenologist 44, 661–678. doi: 10.1017/S0024282912000291
Ruprecht, U., Fernández-Mendoza, F., Türk, R., and Fryday, A. M. (2020). High
levels of endemism and local differentiation in the fungal and algal symbionts
of saxicolous lecideoid lichens along a latitudinal gradient in southern South
America. Lichenologist 52, 287–303. doi: 10.1017/S0024282920000225
Sadowska-De´
s, A. D., Dal Grande, F., Lumbsch, H. T., Beck, A., Otte, J., Hur,
J. S., et al. (2014). Integrating coalescent and phylogenetic approaches to delimit
species in the lichen photobiont Trebouxia.Mol. Phylogenet. Evol. 76, 202–210.
doi: 10.1016/j.ympev.2014.03.020
Schmitt, I., Crespo, A., Divakar, P. K., Fankhauser, J. D., Herman-Sackett, E.,
Kalb, K., et al. (2009). New primers for promising single-copy genes in
fungal phylogenetics and systematics. Persoonia 23, 35–40. doi: 10.3767/
003158509X470602
Schrodt, F., Santos, M. J., Bailey, J. J., and Field, R. (2019). Challenges and
opportunities for biogeography—what can we still learn from von Humboldt?
J. Biogeogr. 46, 1631–1642. doi: 10.1111/jbi.13616
Singh, G., Dal Grande, F., Divakar, P. K., Otte, J., Crespo, A., and Schmitt, I. (2017).
Fungal–algal association patterns in lichen symbiosis linked to macroclimate.
New Phytol. 214, 317–329. doi: 10.1111/nph.14366
Singh, G., Kukwa, M., Dal Grande, F., Łubek, A., Otte, J., and Schmitt, I. (2019).
A glimpse into genetic diversity and symbiont interaction patterns in lichen
communities from areas with different disturbance histories in Białowie˙
za
forest, Poland. Microorganisms 7:335. doi: 10.3390/microorganisms7090335
´
Sliwa, L., Rodriguez Flakus, P., Wilk, K., and Flakus, A. (2014). New records of
Lecanora for Bolivia. II. Pol. Bot. J. 59, 97–103. doi: 10.2478/pbj-2014-0021
´
Sliwa, L., Wilk, K., Rodriguez Flakus, P., and Flakus, A. (2013). New records of
Lecanora for Bolivia. Mycotaxon 121, 385–392. doi: 10.5248/121.385
Soto-Medina, E., Lücking, R., Silverstone-Sopkin, P. A., and Torres, A. M. (2019).
Changes in functional and taxonomic diversity and composition of corticolous
lichens in an altitudinal gradient in Colombia. Cryptogam. Mycol. 40, 97–115.
doi: 10.5252/cryptogamie-mycologie2019v40a6
Stewart, A., Rioux, D., Boyer, F., Gielly, L., Pompanon, F., Saillard, A., et al. (2021).
Altitudinal zonation of green algae biodiversity in the French Alps. Front. Plant
Sci. 12:679428. doi: 10.3389/fpls.2021.679428
Stiller, J. W., and Hall, B. D. (1997). The origin of red algae: implications for plastid
evolution. Proc. Natl. Acad. Sci. U.S.A. 94, 4520–4525.
Tello, J. S., Myers, J. A., Macía, M. J., Fuentes, A. F., Cayola, L., Arellano, G.,
et al. (2015). Elevational gradients in β-diversity reflect variation in the strength
of local community assembly mechanisms across spatial scales. PLoS One
10:e0121458. doi: 10.1371/journal.pone.0121458
Tschermak-Woess, E. (1978). Über die phycobionten der sektion Cystophora von
Chaenotheca, insbesondere Dictyochloropsis splendida und Trebouxia simplex,
spec. nova. Plant Syst. Evol. 129, 185–208. doi: 10.1007/BF00990760
Vanˇ
curová, L., Muggia, L., Peksa, O., ˇ
Rídká, T., and Škaloud, P. (2018). The
complexity of symbiotic interactions influences the ecological amplitude of
the host: a case study in Stereocaulon (lichenized Ascomycota). Mol. Ecol. 27,
3016–3033. doi: 10.1111/mec.14764
Vargas Castillo, R., and Beck, A. (2012). Photobiont selectivity and specificity
in Caloplaca species in a fog-induced community in the Atacama Desert,
northern Chile. Fungal Biol. 116, 665–676. doi: 10.1016/j.funbio.2012.
04.001
Frontiers in Microbiology | www.frontiersin.org 18 December 2021 | Volume 12 | Article 774839
fmicb-12-774839 December 14, 2021 Time: 13:52 # 19
Medeiros et al. Lecanoraceae and Trebouxia in Bolivia
Vilgalys, R., and Hester, M. (1990). Rapid genetic identification and mapping
of enzymatically amplified ribosomal DNA from several Cryptococcus species.
J. Bacteriol. 172, 4238–4246. doi: 10.1128/jb.172.8.4238-4246.1990
von Humboldt, A., and Bonpland, A. (2009). Essay on the Geography of Plants, ed.
S. T. Jackson, trans. S. Romanowski (Chicago, IL: University of Chicago Press).
Voytsekhovich, A., and Beck, A. (2016). Lichen photobionts of the rocky outcrops
of Karadag massif (Crimean Peninsula). Symbiosis 68, 9–24. doi: 10.1007/
s13199-015-0346-y
Wagner, M., Bathke, A. C., Cary, S. C., Green, T. A., Junker, R. R., Trutschnig,
W., et al. (2020). Myco-and photobiont associations in crustose lichens in
the McMurdo Dry Valleys (Antarctica) reveal high differentiation along an
elevational gradient. Polar Biol. 43, 1967–1983. doi: 10.1007/s00300-020-
02754-8
Werth, S., and Sork, V. L. (2014). Ecological specialization in Trebouxia
(Trebouxiophyceae) photobionts of Ramalina menziesii (Ramalinaceae) across
six range-covering ecoregions of western North America. Am. J. Bot. 101,
1127–1140. doi: 10.3732/ajb.1400025
White, T. J., Bruns, T., Lee, S., and Taylor, J. (1990). “Amplification and
direct sequencing of fungal ribosomal RNA genes for phylogenetics,” in
PCR Protocols: A Guide to Methods and Applications, eds M. A. Innis,
D. H. Gelfand, J. J. Sninsky, and T. J. White (London: Academic Press),
315–322.
Wolf, J. H. (1993). Diversity patterns and biomass of epiphytic bryophytes and
lichens along an altitudinal gradient in the northern Andes. Ann. Mo. Bot. Gard.
80, 928–960.
Xu, M., De Boer, H., Olafsdottir, E. S., Omarsdottir, S., and Heidmarsson,
S. (2020). Phylogenetic diversity of the lichenized algal genus Trebouxia
(Trebouxiophyceae, Chlorophyta): a new lineage and novel insights from
fungal-algal association patterns of Icelandic cetrarioid lichens (Parmeliaceae,
Ascomycota). Bot. J. Linn. Soc. 194, 460–468. doi: 10.1093/botlinnean/boaa050
Yahr, R., Vilgalys, R., and DePriest, P. T. (2004). Strong fungal specificity and
selectivity for algal symbionts in Florida scrub Cladonia lichens. Mol. Ecol. 13,
3367–3378. doi: 10.1111/j.1365-294X.2004.02350.x
Yakovchenko, L. S., Davydov, E. A., Ohmura, Y., and Printzen, C. (2019). The
phylogenetic position of species of Lecanora sl containing calycin and usnic
acid, with the description of Lecanora solaris Yakovchenko & Davydov sp. nov.
Lichenologist 51, 147–156. doi: 10.1017/S0024282919000045
Yu, G., Smith, D. K., Zhu, H., Guan, Y., and Lam, T. T. Y. (2017). ggtree: an
R package for visualization and annotation of phylogenetic trees with their
covariates and other associated data. Methods Ecol. Evol. 8, 28–36. doi: 10.1111/
2041-210X.12628
Zhang, J., Kapli, P., Pavlidis, P., and Stamatakis, A. (2013). A general
species delimitation method with applications to phylogenetic placements.
Bioinformatics 29, 2869–2876. doi: 10.1093/bioinformatics/btt499
Zhao, X., Leavitt, S. D., Zhao, Z. T., Zhang, L. L., Arup, U., Grube, M., et al. (2016).
Towards a revised generic classification of lecanoroid lichens (Lecanoraceae,
Ascomycota) based on molecular, morphological and chemical evidence.
Fungal Divers. 78, 293–304. doi: 10.1007/s13225-015-0354-5
Zhou, S., and Stanosz, G. R. (2001). Primers for amplification of mt
SSU rDNA, and a phylogenetic study of Botryosphaeria and associated
anamorphic fungi. Mycol. Res. 105, 1033–1044. doi: 10.1016/S0953-7562(08)61
965-6
Zoller, S., Scheidegger, C., and Sperisen, C. (1999). PCR primers for the
amplification of mitochondrial small subunit ribosomal DNA of lichen-forming
ascomycetes. Lichenologist 31, 511–516. doi: 10.1006/lich.1999.0220
Conflict of Interest: The authors declare that the research was conducted in the
absence of any commercial or financial relationships that could be construed as a
potential conflict of interest.
Publisher’s Note: All claims expressed in this article are solely those of the authors
and do not necessarily represent those of their affiliated organizations, or those of
the publisher, the editors and the reviewers. Any product that may be evaluated in
this article, or claim that may be made by its manufacturer, is not guaranteed or
endorsed by the publisher.
Copyright © 2021 Medeiros, Mazur, Miadlikowska, Flakus, Rodriguez-Flakus,
Pardo-De la Hoz, Cie´
slak, ´
Sliwa and Lutzoni. This is an open-access article
distributed under the terms of the Creative Commons Attribution License (CC BY).
The use, distribution or reproduction in other forums is permitted, provided the
original author(s) and the copyright owner(s) are credited and that the original
publication in this journal is cited, in accordance with accepted academic practice.
No use, distribution or reproduction is permitted which does not comply with
these terms.
Frontiers in Microbiology | www.frontiersin.org 19 December 2021 | Volume 12 | Article 774839
... The few tropical Lecanora specimens sequenced so far are mainly from the Paleotrop ics [7,49] and a recent study with molecular data focusing on Bolivia [50]. With additiona sequences from Brazil, we found that some of the species' identifications of publishe sequences are incongruent; for example, the GenBank (GB) sequences identified as L. vain ioi from Thailand were clustered in a distinct branch separate from the Brazilian materia the type of L. vainioi being is from Brazil [51]. ...
... The few tropical Lecanora specimens sequenced so far are mainly from the Paleotropics [7,49] and a recent study with molecular data focusing on Bolivia [50]. With additional sequences from Brazil, we found that some of the species' identifications of published sequences are incongruent; for example, the GenBank (GB) sequences identified as L. vainioi from Thailand were clustered in a distinct branch separate from the Brazilian material, the type of L. vainioi being is from Brazil [51]. ...
... Curiously, all our sequences of pruinose specimens are positioned in a distinct branch of GB sequences from L. caesiorubella, viz. with L. albella, L. farinacea, and L. subcarnea; similar to Medeiros et al. [50]. Reports of L. caesiorubella Ach. ...
Article
Full-text available
We sequenced over 200 recent specimens of Lecanora s.lat. from Brazil, delimiting 28 species in our material. Many seem to represent undescribed species, some of which being morphologically and chemically similar to each other or to already described species. Here, we present a phylogenetic analysis based on ITS, including our specimens and GenBank data. We describe nine new species. The purpose of the paper is to illustrate the diversity of the genus in Brazil, not to focus on segregate genera. However, we found that all Vainionora species cluster together and these will be treated separately. Other Lecanora species with dark hypothecium clustered in several different clades. Species with the morphology of Lecanora caesiorubella, in which currently several subspecies with different chemistry and distribution are recognized, fall apart in different, distantly related clades, so they cannot be regarded as subspecies but should be recognized at species level. A key is given for the Lecanora species from Brazil.
... SH-aLRT = 100%, PP = 0.99) combines species belonging to Carbonea, the L. polytropa-, L. saligna-and L. varia groups, Polyozosia, Protoparmeliopsis and Rhizoplaca. These genera and species groups conform largely to the 'MPRPS' clade (Medeiros et al. 2021) which, however, also comprised Bryonora, Carbonea and the L. varia group, but were not included in the analysis of Medeiros et al. (2021). The multispored Polyozosia sambuci forms a clade with P. altunica and P. contractula (SH-aLRT = 100%, UFBoot = 100%, PP = 1.00). ...
... SH-aLRT = 100%, PP = 0.99) combines species belonging to Carbonea, the L. polytropa-, L. saligna-and L. varia groups, Polyozosia, Protoparmeliopsis and Rhizoplaca. These genera and species groups conform largely to the 'MPRPS' clade (Medeiros et al. 2021) which, however, also comprised Bryonora, Carbonea and the L. varia group, but were not included in the analysis of Medeiros et al. (2021). The multispored Polyozosia sambuci forms a clade with P. altunica and P. contractula (SH-aLRT = 100%, UFBoot = 100%, PP = 1.00). ...
Article
Full-text available
Two new multispored species from China, Lecanora anhuiensis Li J. Li & Printzen, sp. nov. and Lecanora pseudojaponica Li J. Li & Printzen, sp. nov. are described and illustrated here, based on morphological, chemical and molecular evidence. Lecanora anhuiensis is characterised by an epruinose, yellowish-brown to deep brown apothecial disc, an epihymenium with fine crystals, an amphithecium with small crystals, 16-spored asci and the presence of zeorin, in addition to atranorin. Lecanora pseudojaponica is characterised by an epruinose, red-brown apothecial disc, an epihymenium without crystals , an amphithecium with small crystals, 8 or 16-spored asci and the presence of zeo-rin and the stictic acid complex, in addition to atranorin. Phylogenetic reconstructions, based on mtSSU, nrITS and nrLSU suggest that these two species are members of the Lecanora subfusca group. They are compared with morphologically similar and phylo-genetically related species, based on a nrITS dataset. Phylogenetic results show that the multispored taxa of Lecanora are polyphyletic. The number of ascospores per ascus appears to be a taxonomic character of minor importance. Detailed descriptions, discussions and figures for the two new species from China and a key for the multispored species of Lecanora worldwide are provided.
... However, it is widely acknowledged that specificity of the mycobiont-photobiont association as well as the spatial distribution of their component species largely drives the formation of new lichen organisms 6,[15][16][17][18] . Indeed, many symbionts are quite stringent in the partnerships they form, which are often determined by biophysical gradients [19][20][21][22][23][24] or substrate preferences 18 . Selective mycobionts do not always use the entire niche of their photobionts and are therefore sometimes restricted to certain climatic conditions. ...
... Much of our current understanding on photobiont-mycobiont partnerships comes from studies involving a few lichen species or geographical areas 17,19,23,24,[33][34][35][36][37][38] . However, mycobiont-photobiont partnerships do not happen in isolation, they are part of a larger web of interactions supporting the assembly and maintenance of communities, and understanding them requires a systemic approach. ...
Article
Full-text available
Symbiosis is a major engine of evolutionary innovation underlying many extant complex organisms. Lichens are a paradigmatic example that offers a unique perspective on the role of symbiosis in ecological success and evolutionary diversification. Lichen studies have produced a wealth of information regarding the importance of symbiosis, but they frequently focus on a few species, limiting our understanding of large-scale phenomena such as guilds. Guilds are groupings of lichens that assist each other’s proliferation and are intimately linked by a shared set of photobionts, constituting an extensive network of relationships. To characterize the network of lichen symbionts, we used a large data set (n=206\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$n=206$$\end{document} publications) of natural photobiont-mycobiont associations. The entire lichen network was found to be modular, but this organization does not directly match taxonomic information in the data set, prompting a reconsideration of lichen guild structure and composition. The multiscale nature of this network reveals that the major lichen guilds are better represented as clusters with several substructures rather than as monolithic communities. Heterogeneous guild structure fosters robustness, with keystone species functioning as bridges between guilds and whose extinction would endanger global stability.
... According to previous studies, we selected two species of the genus Protoparmeliopsis and two species of Polyozosia A. Massal. as the outgroup for the genus Rhizoplaca (Medeiros et al. 2021;Zhao et al. 2016;Zhang et al. 2020). Geneious R8 was used to assemble the raw sequences and generate one matrix per locus. ...
Article
Full-text available
In this study, two new species, Rhizoplaca adpressa, are described from Southwest China, based on their morphology, phylogeny and chemistry. In phylogeny, the two new species are monophyletic, and sister to each other within Rhizoplaca chrysoleuca-complex. Rhizo-placa adpressa is characterized by its placodioid and closely adnate thallus, pale green and heavily pruinose upper surface, narrow (ca. 1 mm) and white free margin on the lower surface of marginal squamules, the absence of a lower cortex, and its basally non-constricted apothecia with orange discs that turn reddish-brown at maturity. Rhi-zoplaca auriculata is characterized by its squamulose to placodioid thallus, yellowish green and marginally pruinose squamules, wide (1−3 mm) and bluish-black free margin on the lower surface of marginal squamules, the absence of a lower cortex, and its basally constricted apothecia with persistently orange discs. Rhizoplaca adpressa and R. auriculata share the same secondary metabolites of usnic and placodiolic acids.
... Polymerase chain reaction (PCR) was performed with 1μl of DNA extract in a total volume of 25 μl. Thermal cycler conditions for the locus involved a denaturation step at 95˚C for 3 min, followed by 35 cycles of denaturation at 95˚C for 40 s, annealing at 52˚C for 40 s, and elongation at 72˚C for 150 s, with a final extension at 72˚C for 10 min (Medeiros et al. 2021). Before sequencing PCR, the amplicon was checked on a 1 % agarose gel and cleaned with exonuclease I and shrimp alkaline phosphatase (ExoSap). ...
Article
Śliwa, L., Mazur, E. & Wirth, V. 2023. A new and intriguing species of Myriolecis in a revised phylogenetic framework for the genus. – Herzogia 36: 371–386. On the basis of morphological and chemical data, and a phylogenetic reconstruction using six DNA regions, a new crustose, lecanoroid lichen species from the region of Baden-Württemberg (SW Germany) is described. Myriolecis suevica is characterized by the presence of an unexpected secondary metabolic compound, namely isousnic acid. Phylogenetic analysis revealed that this new species is closely related to M. sambuci, the nomenclatural type for the genus Myriolecis. ‘Lecanora’ casimceana and ‘L.’ elenkinii, which resemble representatives of the genus Myriolecis also contain isousnic acid and are therefore described in detail. The issue of replacing the generic name Myriolecis with the former name Polyozosia for the group traditionally circumscribed as Lecanora dispersa gr. is briefly discussed. Śliwa, L., Mazur, E. & Wirth, V. 2023. Eine im phylogenetischen Kontext der Gattung bemerkenswerte neue Myriolecis-Art. – Herzogia 36: 371–386. Auf Grundlage morphologischer und chemischer Daten und der molekularphylogenetischen Untersuchung von sechs DNA-Regionen wird eine neue lecanoroide Krustenflechte aus Baden-Württemberg (SW-Deutschland) beschrieben. Myriolecis suevica ist durch den in der Gattung ungewÖhnlichen Gehalt an Isousninsäure charakterisiert. Die phylogenetische Analyse zeigt eine nahe Verwandtschaft zu M. sambuci, dem nomenklatorischen Typus der Gattung Myriolecis. ‘Lecanora’ casimceana und ‘Lecanora’ elenkinii, die beide Vertretern der Gattung Myriolecis ähneln, enthalten ebenfalls Isousninsäure im Thallus. Sie werden ausführlich beschrieben. Die vorliegende Arbeit diskutiert kurz die Synonymisierung des Gattungsnamen Myriolecis mit Polyozosia, dem älteren Namen für die Lecanora dispersaGruppe in traditioneller Umschreibung.
... Alongside three fungi and one bacterium, we cultured a green alga ( Chlorophyta , genome 4) of the Asterochloris genus (Table 1 ). This is a common lichen photobiont with members found in cold environments of the Northern Hemisphere, such as Canada, Scandinavia, and Russia (Brown and Jumpponen 2019 , Pino-Bodas and Stenroos 2021 ), and in high altitude ecosystems (Medeiros et al. 2021 ). Asterochloris can also be the photobiont component of a lichen in association with a mycobiont member of the Lecidea genus (Meeßen et al. 2013 ), which would support the taxonomic assignment of isolate 2 to its basionym Lecidea genus. ...
Article
Full-text available
Snow is the largest component of the cryosphere, with its cover and distribution rapidly decreasing over the last decade due to climate warming. It is imperative to characterize the snow (nival) microbial communities to better understand the role of microorganisms inhabiting these rapidly changing environments. Here, we investigated the core nival microbiome, the cultivable microbial members, and the microbial functional diversity of the remote Uapishka mountain range, a massif of alpine sub-arctic tundra and boreal forest. Snow samples were taken over a two-month interval along an altitude gradient with varying degree of anthropogenic traffic and vegetation cover. The core snow alpine tundra/boreal microbiome, which was present across all samples, constituted of Acetobacterales, Rhizobiales and Acidobacteriales bacterial orders, and of Mycosphaerellales and Lecanorales fungal orders, with the dominant fungal taxa being associated with lichens. The snow samples had low active functional diversity, with Richness values ranging from 0 to 19.5. The culture-based viable microbial enumeration ranged from 0 to 8.05 × 103 CFUs/mL. We isolated and whole-genome sequenced five microorganisms which included three fungi, one alga, and one potentially novel bacterium of the Lichenihabitans genus; all of which appear to be part of lichen-associated taxonomic clades.
... Similar findings have been observed in many other symbiosis analyses, where symbiotic partners primarily drive mutualism, while ecological mechanisms and geographical distributions such as extreme environments or cooler climate zones explain part of the variation [15,21,30]. However, within a specific taxonomic scale, ecological factors like climate exhibit a more significant impact on structuring LFF-LFA associations [43,45]. In the island radiation of LFF, macroclimate plays a more critical role in the LFF-LFA association than LFA [40]. ...
Article
Full-text available
Biotic and abiotic factors influence the formation of fungal-algal pairings in lichen sym-biosis. However, the specific determinants of these associations, particularly when distantly related fungi are involved, remain poorly understood. In this study, we investigated the impact of different drivers on the association patterns between taxonomically diverse lichenized fungi and their tre-bouxioid symbiotic partners. We collected 200 samples from four biomes and identified 41 species of lichenized fungi, associating them with 16 species of trebouxioid green algae, of which 62% were previously unreported. The species identity of both the fungal and algal partners had the most significant effect on the outcome of the symbiosis, compared to abiotic factors like climatic variables and geographic distance. Some obviously specific associations were observed in the temperate zone; however, the nestedness value was lower in arid regions than in cold, polar, and temperate regions according to interaction network analysis. Cophylogenetic analyses revealed congruent phyloge-nies between trebouxioid algae and associated fungi, indicating a tendency to reject random associations. The main evolutionary mechanisms contributing to the observed phylogenetic patterns were "loss" and "failure to diverge" of the algal partners. This study broadens our knowledge of fungal-algal symbiotic patterns in view of Trebouxia-associated fungi.
... The number of species-level lineages recognized in Trebouxia increases as soon as new ecological niches or different lichen symbioses are investigated (e.g., [10][11][12]). However, only 30 Trebouxia species-level lineages have so far been formally described based on their cell morphology/ultrastructure and genetic diversity [6,[13][14][15][16]. ...
Article
Full-text available
Two microalgal species, Trebouxia jamesii and Trebouxia sp. TR9, were detected as the main photobionts coexisting in the thalli of the lichen Ramalina farinacea. Trebouxia sp. TR9 emerged as a new taxon in lichen symbioses and was successfully isolated and propagated in in vitro culture and thoroughly investigated. Several years of research have confirmed the taxon Trebouxia sp. TR9 to be a model/reference organism for studying mycobiont–photobiont association patterns in lichen symbioses. Trebouxia sp. TR9 is the first symbiotic, lichen-forming microalga for which an exhaustive characterization of cellular ultrastructure, physiological traits, genetic and genomic diversity is available. The cellular ultrastructure was studied by light, electron and confocal microscopy; physiological traits were studied as responses to different abiotic stresses. The genetic diversity was previously analyzed at both the nuclear and organelle levels by using chloroplast, mitochondrial, and nuclear genome data, and a multiplicity of phylogenetic analyses were carried out to study its intraspecific diversity at a biogeographical level and its specificity association patterns with the mycobiont. Here, Trebouxia sp. TR9 is formally described by applying an integrative taxonomic approach and is presented to science as Trebouxia lynnae, in honor of Lynn Margulis, who was the primary modern proponent for the significance of symbiosis in evolution. The complete set of analyses that were carried out for its characterization is provided.
Article
Full-text available
Recent work has suggested exceptional species-level diversity in the lichen-forming Lecanora polytropa complex (Lecanoraceae, Ascomycota). However, biogeographic patterns and the spatial structuring of this diversity remains poorly known. To investigate diversity across multiple spatial scales, we sampled members of this species complex from two distinct regions—the Pacific Coast Ranges in southern Alaska, USA, and montane habitats in Spain. We also included sequence data from several species within this complex that were recently described from populations in France. Using the standard DNA barcoding marker and a sequence-based species delimitation approach (ASAP), we inferred a total of 123 candidate species (SHs) within the Lecanora polytropa complex, 32 of which were sampled for the first time here. Of 123 SHs, 21 had documented intercontinental distributions, while the vast majority were found at much smaller spatial scales. From our samples collected from Alaska, USA, and Spain, representing 36 SHs, we found high genetic diversity occurring within each sampled site, but limited overlap among all sites. Mountain ranges in both regions had high proportions of endemic lineages, with the highest diversity and endemism occurring in mountain ranges in Spain. Our sequence data generally support the recent taxonomic proposals, and an integrative taxonomy may help partly resolve the taxonomic conundrums within this hyper-diverse lineage.
Article
Arthonia frostiicola and A. galligena are described as new to science based on collections from mountainous regions of southeastern North America. Arthonia frostiicola infects the saxicolous lichen Dirinaria frostii, producing emarginate black apothecia which erupt from within the host thallus. It is characterized by a dark hypothecium and 1-septate, obovoid ascospores which turn brownish and verruculose in age. It is known from five collections made in the southern Appalachian Mountains and Ozark Mountains in southeastern North America. Arthonia galligena produces galls in the thallus and apothecia of the corticolous lichens Lecanora masana and L. rugosella, and is apparently endemic to the high elevations of the southern Appalachian Mountains. It is characterized by a variably pigmented, pale to red-brown hypothecium and 2-septate, macrocephalic ascospores which turn brownish and verruculose in age. Keys to the species of Arthonia on Caliciales and Lecanoraceae are provided.
Article
Full-text available
Mountain environments are marked by an altitudinal zonation of habitat types. They are home to a multitude of terrestrial green algae, who have to cope with abiotic conditions specific to high elevation, e.g., high UV irradiance, alternating desiccation, rain and snow precipitations, extreme diurnal variations in temperature and chronic scarceness of nutrients. Even though photosynthetic green algae are primary producers colonizing open areas and potential markers of climate change, their overall biodiversity in the Alps has been poorly studied so far, in particular in soil, where algae have been shown to be key components of microbial communities. Here, we investigated whether the spatial distribution of green algae followed the altitudinal zonation of the Alps, based on the assumption that algae settle in their preferred habitats under the pressure of parameters correlated with elevation. We did so by focusing on selected representative elevational gradients at distant locations in the French Alps, where soil samples were collected at different depths. Soil was considered as either a potential natural habitat or temporary reservoir of algae. We showed that algal DNA represented a relatively low proportion of the overall eukaryotic diversity as measured by a universal Eukaryote marker. We designed two novel green algae metabarcoding markers to amplify the Chlorophyta phylum and its Chlorophyceae class, respectively. Using our newly developed markers, we showed that elevation was a strong correlate of species and genus level distribution. Altitudinal zonation was thus determined for about fifty species, with proposed accessions in reference databases. In particular, Planophila laetevirens and Bracteococcus ruber related species as well as the snow alga Sanguina genus were only found in soil starting at 2,000 m above sea level. Analysis of environmental and bioclimatic factors highlighted the importance of pH and nitrogen/carbon ratios in the vertical distribution in soil. Capacity to grow heterotrophically may determine the Trebouxiophyceae over Chlorophyceae ratio. The intensity of freezing events (freezing degree days), proved also determinant in Chlorophyceae distribution. Guidelines are discussed for future, more robust and precise analyses of environmental algal DNA in mountain ecosystems and address green algae species distribution and dynamics in response to environmental changes.
Article
Full-text available
Lichens are classic models of symbiosis, and one of the most frequent nutritional modes among fungi. The ecologically and geographically widespread lichen-forming algal (LFA) genus Trebouxia is one of the best-studied groups of LFA and associates with over 7000 fungal species. Despite its importance, little is known about its diversification. We synthesized twenty years of publicly-available data by characterizing the ecological preferences of this group and testing for time-variant shifts in climatic regimes over a distribution of trees. We found evidence for limited shifts among regimes, but that disparate lineages convergently evolved similar ecological tolerances. Early Trebouxia lineages were largely forest specialists or habitat generalists that occupied a regime whose extant members occur in moderate climates. Trebouxia then convergently diversified in non-forested habitats and expanded into regimes whose modern representatives occupy wet-warm and cool-dry climates. We rejected models in which climatic diversification slowed through time, suggesting climatic diversification is inconsistent with that expected under an adaptive radiation. In addition, we found that climatic and vegetative regime shifts broadly coincided with the evolution of biomes and associated or similar taxa. Together, our work illustrates how this keystone symbiont from an iconic symbiosis evolved to occupy diverse habitats across the globe.
Article
Full-text available
Phylogenetic reconstructions based on ITS/5.8S, mtSSU and nuLSU DNA sequence data suggest that Lecanora pringlei from North America and a closely related new species from the Altai Mountains, Russia, should be transferred to a new genus Pulvinora, phylogenetically related to Frutidella. It is distinguished by Lecanora-type asci, mycolecanorine apothecia soon becoming convex with an algal layer pushed below the hypothecium, and a pulvinate thallus with squamules at the tip of pseudopodetia-like, branched, pale brownish structures. Lecanora subcavicola and L. pringlei subsp. brandegeei do not belong to this new genus Pulvinora; consequently, we propose the new combination L. brandegeei to accommodate the latter taxon. Pulvinora stereothallina is distinguished from P. pringlei in the shape and size of its squamules (plane to concave, up to 3.0 mm long vs. remaining convex, up to 1.5 mm long), by the lack of maculae, the presence of a whitish pruina on the margins and elevated parts of its squamules, by apothecia coalescing into clusters (vs. single), an ochre-yellow to brownish (vs. colorless) proper exciple, larger ascospores, and different secondary metabolites. Lectotypes of Lecidea pringlei Tuck. and L. brandegeei Tuck. are designated here from the collection deposited in FH.
Article
Full-text available
We studied the biodiversity of Asterochloris photobionts found in Bolivian lichens to better understand their global spatial distribution and adaptation strategies in the context of a worldwide phylogeny of the genus. Based on nuclear ITS rDNA, the chloroplast rbcL gene and the actin type I gene we reconstructed a phylogenetic tree that recovered nine new Asterochloris lineages, while 32 Bolivian photobiont samples were assigned to 12 previously recognized Asterochloris lineages. We also show that some previously discovered Asterochloris photobiont species and lineages may occur in a broader spectrum of climatic conditions, and mycobiont species and photobionts may show different preferences along an altitude gradient. To reveal general patterns of of mycobiont specificity towards the photobiont in Asterochloris, we tested the influence of climate, altitude, geographical distance and effects of symbiotic partner (mycobiont) at the species level of three genera of lichen forming fungi: Stereocaulon, Cladonia and Lepraria. Further, we compared the specificity of mycobionts towards Asterochloris photobionts in cosmopolitan, Neotropical, and Pantropical lichen forming fungi. Interestingly, cosmopolitan species showed the lowest specificity to their photobionts, but also the lowest haplotype diversity. Neotropical and Paleotropical mycobionts, however, were more specific.
Article
Full-text available
Climatically extreme regions such as the polar deserts of the McMurdo Dry Valleys (78° S) in Continental Antarctica are key areas for a better understanding of changes in ecosystems. Therefore, it is particularly important to analyze and communicate current patterns of biodiversity in these sensitive areas, where precipitation mostly occurs in form of snow and liquid water is rare. Humidity provided by dew, clouds, and fog are the main water sources, especially for rock-dwelling crustose lichens as one of the most common vegetation-forming organisms. We investigated the diversity and interaction specificity of myco-/photobiont associations of 232 crustose lichen specimens, collected along an elevational gradient (171-959 m a.s.l.) within the McMurdo Dry Valleys. The mycobiont species and photobiont OTUs were identified by using three markers each (nrITS, mtSSU, RPB1, and nrITS, psbJ-L, COX2). Elevation, positively associated with water availability, turned out to be the key factor explaining most of the distribution patterns of the mycobionts. Pairwise comparisons showed Lecidea cancriformis and Rhizoplaca macleanii to be significantly more common at higher elevations and Carbonea vorticosa and Lecidea polypycnidophora at lower elevations. Lichen photobionts were dominated by the globally distributed Trebouxia OTU, Tr_A02 which occurred at all habitats. Network specialization resulting from myco-/photobiont bipartite network structure varied with elevation and associated abiotic factors. Along an elevational gradient, the spatial distribution, diversity, and genetic variability of the lichen symbionts appear to be mainly influenced by improved water relations at higher altitudes.
Article
Full-text available
Saxicolous, lecideoid lichenized fungi have a cosmopolitan distribution but, being mostly cold adapted, are especially abundant in polar and high-mountain regions. To date, little is known of their origin or the extent of their trans-equatorial dispersal. Several mycobiont genera and species are thought to be restricted to either the Northern or the Southern Hemisphere, whereas others are thought to be widely distributed and occur in both hemispheres. However, these assumptions often rely on morphological analyses and lack supporting molecular genetic data. Also unknown is the extent of regional differentiation in the southern polar regions. An extensive set of lecideoid lichens (185 samples) was collected along a latitudinal gradient at the southern end of South America. Subantarctic climate conditions were maintained by increasing the elevation of the collecting sites with decreasing latitude. The investigated specimens were placed in a global context by including Antarctic and cosmopolitan sequences from other studies. For each symbiont three markers were used to identify intraspecific variation (mycobiont: ITS, mtSSU, RPB1 ; photobiont: ITS, psbJ-L, COX2). For the mycobiont, the saxicolous genera Lecidea , Porpidia , Poeltidea and Lecidella were phylogenetically re-evaluated, along with their photobionts Asterochloris and Trebouxia . For several globally distributed species groups, the results show geographically highly differentiated subclades, classified as operational taxonomical units (OTUs), which were assigned to the different regions of southern South America (sSA). Furthermore, several small endemic and well-supported clades apparently restricted to sSA were detected at the species level for both symbionts.
Article
Full-text available
Lichens have high tolerance to harsh environmental conditions, where lichen symbiont interactions (e.g. myco- and photobionts) may play a crucial role. The characterization of fungal-algal association patterns is essential to understand their symbiotic interactions. This study investigated fungal-algal association patterns in Icelandic cetrarioid lichens using a multi-locus phylogenetic framework, including fungal nrITS, MCM7, mtSSU, RPB1 and RPB2 and algal nrITS, nrLSU, rbcL and mtCOXII data. Most Icelandic cetrarioid lichenized fungi were found to be specifically associated to the known Trebouxia clade “S” (Trebouxia simplex/suecica group), whereas the lichen-forming fungus Cetrariella delisei forms a symbiosis with a previously unrecognized lineage of Trebouxia, provisionally named as the “D” clade. This new Trebouxia lineage is supported by maximum likelihood and Bayesian phylogenetic analyses using all four included algal loci.
Article
Full-text available
Keystone mutualisms, such as corals, lichens or mycorrhizae, sustain fundamental ecosystem functions. Range dynamics of these symbioses are, however, inherently difficult to predict because host species may switch between different symbiont partners in different environments, thereby altering the range of the mutualism as a functional unit. Biogeographic models of mutualisms thus have to consider both the ecological amplitudes of various symbiont partners and the abiotic conditions that trigger symbiont replacement. To address this challenge, we here investigate 'symbiont turnover zones'-defined as demarcated regions where symbiont replacement is most likely to occur, as indicated by overlapping abundances of symbiont ecotypes. Mapping the distribution of algal symbionts from two species of lichen-forming fungi along four independent altitudinal gradients, we detected an abrupt and consistent β-diversity turnover suggesting parallel niche partitioning. Modelling contrasting environmental response functions obtained from latitudinal distributions of algal ecotypes consistently predicted a confined altitudinal turnover zone. In all gradients this symbiont turnover zone is characterized by approximately 12°C average annual temperature and approximately 5°C mean temperature of the coldest quarter, marking the transition from Mediterranean to cool temperate bioregions. Integrating the conditions of symbiont turnover into biogeographic models of mutualisms is an important step towards a comprehensive understanding of biodiversity dynamics under ongoing environmental change.
Article
The Lecanoromycetes includes most of the lichen-forming fungal species (>13 500) and is therefore one of the most diverse class of all Fungi in terms of phenotypic complexity. We report phylogenetic relationships within the Lecanoromycetes resulting from Bayesian and maximum likelihood analyses with complementary posterior probabilities and bootstrap support values based on three combined multilocus datasets using a supermatrix approach. Nine of 10 orders and 43 of 64 families currently recognized in Eriksson’s classification of the Lecanoromycetes (Outline of Ascomycota—2006 Myconet 12:1–82) were represented in this sampling. Our analyses strongly support the Acarosporomycetidae and Ostropomycetidae as monophyletic, whereas the delimitation of the largest subclass, the Lecanoromycetidae, remains uncertain. Independent of future delimitation of the Lecanoromycetidae, the Rhizocarpaceae and Umbilicariaceae should be elevated to the ordinal level. This study shows that recent classifications include several nonmonophyletic taxa at different ranks that need to be recircumscribed. Our phylogenies confirm that ascus morphology cannot be applied consistently to shape the classification of lichen-forming fungi. The increasing amount of missing data associated with the progressive addition of taxa resulted in some cases in the expected loss of support, but we also observed an improvement in statistical support for many internodes. We conclude that a phylogenetic synthesis for a chosen taxonomic group should include a comprehensive assessment of phylogenetic confidence based on multiple estimates using different methods and on a progressive taxon sampling with an increasing number of taxa, even if it involves an increasing amount of missing data.
Article
Lichens provide valuable systems for studying symbiotic interactions. In lichens, these interactions are frequently described in terms of availability, selectivity and specificity of the mycobionts and photobionts towards one another. The lichen-forming, green algal genus Trebouxia Puymaly is among the most widespread photobiont, associating with a broad range of lichen-forming fungi. To date, 29 species have been described, but studies consistently indicate that the vast majority of species-level lineages still lack formal description, and new, previously unrecognized lineages are frequently reported. To reappraise the diversity and the evolutionary relationships of species-level lineages in Trebouxia, we assembled DNA sequence data from over 1600 specimens, compiled from a range of sequences from previously published studies, axenic algal cultures, and lichens collected from poorly sampled regions. From these samples, we selected representatives of the currently known genetic diversity in the lichenized Trebouxia and inferred a phylogeny from multi-locus sequence data (ITS, rbcL, cox2). We demonstrate that the current formally described species woefully underrepresent overall species-level diversity in this important lichen-forming algal genus. We anticipate that an integrative taxonomic approach, incorporating morphological and physiological data from axenic cultures with genetic data, will be required to establish a robust, comprehensive taxonomy for Trebouxia. The data presented here provide an important impetus and reference dataset for more reliably characterizing diversity in lichenized algae and in using lichens to investigate the evolution of symbioses and holobionts.