Hostname: page-component-8448b6f56d-m8qmq Total loading time: 0 Render date: 2024-04-24T02:02:43.828Z Has data issue: false hasContentIssue false

Complex photobiont diversity in the marine lichen Lichina pygmaea

Published online by Cambridge University Press:  22 September 2021

Nathan A. M. Chrismas*
Affiliation:
Marine Biological Association of the UK, The Laboratory, Citadel Hill, Plymouth, UK
Ro Allen
Affiliation:
Marine Biological Association of the UK, The Laboratory, Citadel Hill, Plymouth, UK
Anita L. Hollingsworth
Affiliation:
Marine Biological Association of the UK, The Laboratory, Citadel Hill, Plymouth, UK University of Southampton, European Way, Southampton, UK
Joe D. Taylor
Affiliation:
University of Bradford, Bradford, West Yorkshire, UK
Michael Cunliffe*
Affiliation:
Marine Biological Association of the UK, The Laboratory, Citadel Hill, Plymouth, UK School of Biological and Marine Sciences, University of Plymouth, Plymouth, UK
*
Authors for correspondence: Nathan Chrismas, Michael Cunliffe, E-mail: natchr@mba.ac.uk, micnli@mba.ac.uk
Authors for correspondence: Nathan Chrismas, Michael Cunliffe, E-mail: natchr@mba.ac.uk, micnli@mba.ac.uk
Rights & Permissions [Opens in a new window]

Abstract

Lichens are a well-known symbiosis between a host mycobiont and eukaryote algal or cyanobacterial photobiont partner(s). Recent studies have indicated that terrestrial lichens can also contain other cryptic photobionts that increase the lichens’ ecological fitness in response to varying environmental conditions. Marine lichens live in distinct ecosystems compared with their terrestrial counterparts because of regular submersion in seawater and are much less studied. We performed bacteria 16S and eukaryote 18S rRNA gene metabarcoding surveys to assess total photobiont diversity within the marine lichen Lichina pygmaea (Lightf.) C. Agardh, which is widespread throughout the intertidal zone of Atlantic coastlines. We found that in addition to the established cyanobacterial photobiont Rivularia, L. pygmaea is also apparently host to a range of other marine and freshwater cyanobacteria, as well as marine eukaryote algae in the family Ulvophyceae (Chlorophyta). We propose that symbiosis with multiple freshwater and marine cyanobacteria and eukaryote photobionts may contribute to the ability of L. pygmaea to survive the harsh fluctuating environmental conditions of the intertidal zone.

Type
Research Article
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution, and reproduction in any medium, provided the original work is properly cited.
Copyright
Copyright © The Author(s), 2021. Published by Cambridge University Press on behalf of Marine Biological Association of the United Kingdom

Introduction

Lichen symbiosis is undergoing renewed scrutiny, in part because of the application of molecular ecology techniques such as metabarcoding (Hawksworth & Grube, Reference Hawksworth and Grube2020; Smith et al., Reference Smith, Dal Grande, Muggia, Keuler, Divakar, Grewe, Schmitt, Lumbsch and Leavitt2020). Historic assumptions about the ‘single fungus and single eukaryote alga or Cyanobacteria’ nature of the symbiosis do not take into account tripartite lichens (i.e., those containing cyanobacterial and algal photobionts) (Rikkinen et al., Reference Rikkinen, Oksanen and Lohtander2002; Henskens et al., Reference Henskens, Green and Wilkins2012), and have been further challenged by the more recent discovery of other fungi (Millanes et al., Reference Millanes, Diederich and Wedin2016; Spribille et al., Reference Spribille, Tuovinen, Resl, Vanderpool, Wolinski, Aime, Schneider, Stabentheiner, Toome-Heller, Thor, Mayrhofer, Johannesson and McCutcheon2016; Mark et al., Reference Mark, Laanisto, Bueno, Niinemets, Keller and Scheidegger2020), eukaryote algae (chlorobionts) and Cyanobacteria (cyanobionts) (Henskens et al., Reference Henskens, Green and Wilkins2012; Moya et al., Reference Moya, Molins, Martínez-Alberola, Muggia and Barreno2017) within lichens in addition to the primary myco- and photobionts (Smith et al., Reference Smith, Dal Grande, Muggia, Keuler, Divakar, Grewe, Schmitt, Lumbsch and Leavitt2020), alongside a complex diversity of heterotrophic bacteria (Grube et al., Reference Grube, Cardinale, de Castro, Müller and Berg2009). This increased diversity of organisms co-existing within the lichen thallus has added support to an ongoing reinterpretation of lichens as even more complex ecological associations comprising numerous interacting components (Honegger, Reference Honegger1991; Hawksworth & Grube, Reference Hawksworth and Grube2020).

Amongst this complex lichen community, photobionts are fundamental members responsible for providing sugars (and in the case of some cyanobacterial photobionts, available nitrogen) essential for the survival of the lichen complex (holobiont). Extended photobiont diversity is likely widespread in terrestrial lichens (Dal Forno et al., Reference Dal Forno, Lawrey, Sikaroodi, Gillevet, Schuettpelz and Lücking2021) and has been hypothesized to be associated with increased ecological adaptive capability (Casano et al., Reference Casano, del Campo, García-Breijo, Reig-Armiñana, Gasulla, del Hoyo, Guéra and Barreno2011; del Campo et al., Reference del Campo, Catalá, Gimeno, del Hoyo, Martínez-Alberola, Casano, Grube and Barreno2013; Muggia et al., Reference Muggia, Vancurova, Škaloud, Peksa, Wedin and Grube2013; Ruprecht et al., Reference Ruprecht, Brunauer and Türk2014). By recruiting a diverse range of photobionts, it has been proposed that lichens may benefit from enhanced ecological plasticity, facilitating the ability to survive a range of environments (Casano et al., Reference Casano, del Campo, García-Breijo, Reig-Armiñana, Gasulla, del Hoyo, Guéra and Barreno2011; del Campo et al., Reference del Campo, Catalá, Gimeno, del Hoyo, Martínez-Alberola, Casano, Grube and Barreno2013; Muggia et al., Reference Muggia, Vancurova, Škaloud, Peksa, Wedin and Grube2013). For example, the soil crust lichen Psora decipiens supports multiple lineages of Trebouxia and Asterochloris allowing it to dominate soil crusts from deserts to alpine sites (Ruprecht et al., Reference Ruprecht, Brunauer and Türk2014). Additionally, photobiont ‘switching’ may allow for a further increase in fitness where multiple mycobiont lineages share locally adapted photobiont genotypes (Piercey-Normore & DePriest, Reference Piercey-Normore and DePriest2001).

Changing environmental conditions are characteristic of the habitat occupied by marine lichens in the intertidal zone, with marine lichens being exposed to a uniquely dynamic range of pressures not experienced by terrestrial lichens including daily tidal cycles of seawater coverage. Lichina pygmaea (Lightf.) C. Agardh is a saxicolous fruticose lichen in the family Lichinaceae that inhabits the upper intertidal zone (Figure 1A). It can be found throughout the North Atlantic coastline, including the area surrounding Plymouth, UK (Naylor, Reference Naylor1930). L. pygmaea was historically thought to form a symbiosis with Calothrix (Cyanobacteria) (Whitton, Reference Whitton2012), which was updated through molecular characterization to Rivularia (Ortiz-Álvarez et al., Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015). Marine lichens such as L. pygmaea remain relatively under-studied compared with terrestrial taxa, and the degree to which complex photobiont diversity occurs in marine lichens is yet to be determined. The potential for multiple cyanobionts within members of the Lichinaceae was proposed by Bubrick & Galun (Reference Bubrick and Galun1984), but the extent of this has not yet been investigated using molecular ecology techniques. Here, we perform a broad metabarcoding survey of the bacterial (16S rRNA gene) and eukaryote (18S rRNA gene) phototrophic taxa associated with the L. pygmaea thallus to establish the potential for complex photobiont diversity in a marine lichen.

Fig. 1. (Ai–iii) Lichina pygmaea at Church Reef with close-up of thallus. Dotted line indicates approximate upper limit of the intertidal. Arrows point to L. pygmaea thalli. (B) Sampling locations of L. pygmaea near Plymouth (UK).

Methods

Sampling

Sampling and processing were carried out according to the protocol described in Parrot et al. (Reference Parrot, Antony-Babu, Intertaglia, Grube, Tomasi and Suzuki2015). In summary, L. pygmaea thalli were collected from three locations close to Plymouth in the south-west of the UK (Figure 1B) in spring 2016: Rame Head (lat = 50.3228, long = −4.2194), West Wembury (lat = 50.3159, long = 4.0856), and Church Reef (lat = 50.3159, long = −4.0844). At each site, a single thallus was sampled using sterile forceps, transported to the laboratory at 4°C and processed within 1 h. Each of the three thalli were inspected using a dissecting microscope to remove any debris or invertebrates and separated into 1 g subsamples (N = 5). To separate the intra- and extrathalline communities, thalli were washed twice by placing in a universal tube with 20 mL sterile seawater and rotating for 10 min. Wash water was filtered through a 0.45 μm cellulose nitrate filter. Both the washed thalli (intrathalline community) and filter (extrathalline community) were homogenized using a Mini-Beadbeater machine (Biospec Products) at 42 rpm for 40 s.

DNA extraction

DNA was extracted from the washed thalli (intrathalline) and the wash water filters (extrathalline) using the PowerSoil DNA Isolation Kit (MoBio, CA, USA) and stored at −20°C. The V4 region of the bacterial 16S rRNA gene was amplified using primers 515F and 926R (Caporaso et al., Reference Caporaso, Lauber, Walters, Berg-Lyons, Lozupone, Turnbaugh, Fierer and Knight2011) using Phusion® High-Fidelity DNA Polymerase (New England Biolabs) and PCR was performed under the following conditions: 94°C for 5 min, followed by 30 cycles of 94°C for 30 s, 53°C for 30 s and 72°C for 1 min, and a final elongation step at 5°C for 10 min. The V4 region of the eukaryotic 18S gene was targeted with the primer pair E572F and E1009R (Comeau et al., Reference Comeau, Li, Tremblay, Carmack and Lovejoy2011), and PCR was performed under the following conditions: 98°C for 30 s, followed by 30 cycles at 98°C for 10 s, 55°C for 30 s, 72°C for 30 s, with a final extension at 72°C for 5 min. Sequencing was performed on the Illumina MiSeq platform (Integrated Microbiome Resource, Dalhousie University, Canada).

Bioinformatics

16S rRNA and 18S rRNA gene reads were processed through the DADA2 pipeline (Callahan et al., Reference Callahan, McMurdie, Rosen, Han, Johnson and Holmes2016), which resolves amplicon sequence variants (ASVs) at single-nucleotide resolution (Callahan et al., Reference Callahan, McMurdie and Holmes2017). Reads were trimmed and truncated to remove primers and low-quality sequences respectively. Following error estimation, paired-end reads were merged, and chimeras were removed. Taxonomic classification of resolved ASVs was performed using the RDP naïve Bayesian classifier (Wang et al., Reference Wang, Garrity, Tiedje and Cole2007) using SILVA version 132 (Quast et al., Reference Quast, Pruesse, Yilmaz, Gerken, Schweer, Yarza, Peplies and Glöckner2013) for 16S rRNA gene ASVs and PR2 version 4.12.0 (Guillou et al., Reference Guillou, Bachar, Audic, Bass, Berney, Bittner, Boutte, Burgaud, de Vargas, Decelle, Del Campo, Dolan, Dunthorn, Edvardsen, Holzmann, Kooistra, Lara, Le Bescot, Logares, Mahé, Massana, Montresor, Morard, Not, Pawlowski, Probert, Sauvadet, Siano, Stoeck, Vaulot, Zimmermann and Christen2013) for 18S rRNA gene ASVs. Read counts, taxonomic classifications, and sample metadata were compiled as a phyloseq object prior to downstream analysis (McMurdie & Holmes, Reference McMurdie and Holmes2013). L. pygmaea reads were removed from analysis. A median of 15,115 16S rRNA gene reads and 1551 processed 18S rRNA gene reads per sample were recovered (Raw sequencing data are deposited in the European Read Archive under ENA study PRJEB43698). Alpha- (Observed, Shannon Index, Pielou's Evenness) and beta-diversity (Bray–Curtis dissimilarity) analyses were carried out using phyloseq and differential abundances of specific ASVs between the extra- and intrathalline samples were assessed using DESeq2 (Love et al., Reference Love, Huber and Anders2014), based on non-normalized read count data and a significance threshold of α = 0.05, with a value of +1 assigned to the matrix to allow for statistical comparison between samples where no reads were returned. This analysis allowed for a direct test of which ASVs are enriched within the thalli and therefore most likely to be acting as photobionts. All analyses were performed in the R environment (R Core Team, 2017).

Phylogenetic trees

Identities of cyanobacterial (accession numbers MZ169867–MZ169909) and eukaryote algal (accession numbers MZ190002–MZ190020) ASVs identified by DESeq2 were confirmed by performing BLAST searches (megablast, excluding uncultured/environmental sample sequences, e-value = 0.05) against the NCBI nt database (9 November 2020) (see Supplementary material). Two 18S rRNA gene sequences (ASV_433 and ASV_251) had indeterminate identities and were removed from further analysis. To build phylogenetic trees, best BLAST hits for each ASV and an additional 2–5 close matches were obtained. Where possible, information about source material was recorded. L. pygmaea derived Rivularia sequences from Ortiz-Álvarez et al. (Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015) were included in the cyanobacterial tree and photobionts of Hydropunctaria maura and Wahlenbergiella striatula from Thüs et al. (Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube, Friedl, Brodie, Andrew, Lücking, Lutzoni and Gueidan2011) in the algal tree. Sequences were aligned using MAFFT (using G-INS-i) and manually trimmed to remove gappy tails and poorly aligned regions. The 16S rRNA gene sequence ASV_170 was poorly aligned and excluded from further analysis. Phylogenies were reconstructed using the Cipres (Miller et al., Reference Miller, Pfeiffer and Schwartz2011) implementation of RAXML-HPC (Stamatakis, Reference Stamatakis2014) using the GTR-CAT model and 1000 bootstrap replicates. Outgroups were specified for both Cyanobacteria (Pseudanabaena) and Ulvophyceae (Trentepohlia) trees. Resulting trees were first checked in FigTree (https://github.com/rambaut/figtree), before being annotated in R using ggtree (Yu et al., Reference Yu, Smith, Zhu, Guan and Lam2017). All figures were manually edited for aesthetics in Inkscape (https://inkscape.org/).

Results

Cyanobacteria dominated the bacterial community within all three thalli accounting for up to 49% of 16S rRNA gene reads (Figure 2A), while the green algae family Ulvophyceae (Chlorophyta) dominated the intrathalline eukaryotic community, accounting for up to 82% of the 18S rRNA gene reads (Figure 2A). Analyses of alpha-diversity indicated that despite increased abundance, Cyanobacteria exhibited significantly lower diversity (Shannon, P ≤ 0.001) and evenness (Pielou, P ≤ 0.001) inside the sampled thalli (Figure 2B). Similarly, Ulvophyceae showed a non-significant (Shannon, P = 0.0781) reduction in diversity within the thalli and a significant reduction in evenness (Pielou, P ≤ 0.005) (Figure 2B). Analyses of beta-diversity revealed a clear distinction between the intrathalline and extrathalline communities of both Cyanobacteria (Adonis, P ≤ 0.001) and Ulvophyceae (Adonis, P ≤ 0.001) (Figure 2C).

Fig. 2. (A) Relative abundance of bacterial and eukaryotic intrathalline and extrathalline communities. (B) Comparison of alpha-diversity between intrathalline (red) and extrathalline (blue) communities of Cyanobacteria and Ulvophyceae, indicating Observed diversity, Shannon Index and Pielou's Evenness. (C) Bray–Curtis NMDS plots showing beta-diversity of Cyanobacteria and Ulvophyceae intrathalline (red) and extrathalline (blue) communities. Shape indicates sampling site.

A total of six dominant (average normalized read count >20) cyanobacterial ASVs were significantly enriched (P ≤ 0.005) within at least one location (Figure 3A). Three significantly enriched ASVs, ASV_6 (MZ169869), ASV_18 (MZ169873) and ASV_55 (MZ169883) accounting for up to 8.7, 4.6 and 1.5% of the intrathalline bacterial community respectively, and a fourth ASV_73 (MZ169887), which was only found at Rame Head where it comprised up to 6.1% of the community, all belonged to Rivularia, the established L. pygmaea photobiont (Ortiz-Álvarez et al., Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015). However, ASV_2 (MZ169867) and ASV_4 (MZ169868) (ASV_4 was only found at West Wembury) which made up to 20.6% and 18.1% of the intrathalline bacterial community respectively were both related to Pleurocapsa, a genus not previously associated with L. pygmaea. Other cyanobacterial ASVs were derived from the genera Acaryochloris, Phormidesmis and Synechocystis, some of which showed increased abundance in the extrathalline community. Within the phylogenetic tree (100 sequences, total alignment length = 1465 bp), all Rivularia ASVs occurred within a clade containing mostly marine and saline water derived sequences, while the highly abundant Pleurocapsa ASV_2 (MZ169867) fell within a freshwater dominated clade (Figure 4).

Fig. 3. Deseq2 plots indicating (A) Cyanobacteria and (B) Ulvophyceae ASVs enriched in the interior (red) and exterior (blue) of the L. pygmaea thallus. Prefix ‘endo-’ refers to intrathalline and ‘exo-’ refers to extrathalline. Circle size is the average of normalized read counts over all samples. Significant enrichments are labelled.

Fig. 4. 16S rRNA gene maximum likelihood phylogeny of Cyanobacteria associated with Lichina pygmaea. ASVs significantly enriched in the intrathalline samples are indicated with*. Prefix ‘endo-’ refers to intrathalline (red circles) and ‘exo-’ refers to extrathalline (blue circles). Circle size is relative abundance. Node labels indicate bootstrap support values for main clades.

The two abundant ASVs (average normalized read count > 20) belonging to the Ulvophyceae were significantly enriched (P ≤ 0.005) in the intrathalline community (Figure 3B), ASV_14 (MZ190003) (up to 16% intrathalline eukaryotes) and ASV_7 (MZ190002) (up to 50% intrathalline eukaryotes). Phylogenetic analysis (72 sequences, total alignment length = 1163 bp) placed these within clades of marine derived Paulbroadya and Blidingia respectively (Figure 5).

Fig. 5. 18S rRNA gene maximum likelihood phylogeny of Ulvophyceae associated with Lichina pygmaea ASVs significantly enriched in the intrathalline samples are indicated with *. Prefix ‘endo-’ refers to intrathalline (red circles) and ‘exo-’ refers to extrathalline (blue circles). Circle size is relative abundance. Node labels indicate bootstrap support values for main clades. Dashed lines next to clade names indicate paraphyletic clades.

Discussion

Here we report a considerable diversity of cyanobacterial and algal lineages associated with the interior of the L. pygmaea thallus. Complex diversity of lichen cyanobionts such as this has not been widely reported, despite early indications from the Lichinaceae family (Bubrick & Galun, Reference Bubrick and Galun1984). A previous molecular-based survey of L. pygmaea cyanobionts (Ortiz-Álvarez et al., Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015), which used conventional PCR and Sanger sequencing only amplified the dominant cyanobiont Rivularia. The whole community metabarcoding approach used here (which detects all sequence variants even at low abundance) suggests that the diversity of cyanobacteria contained within the thallus of L. pygmaea may be more complex than previously considered (Ortiz-Álvarez et al., Reference Ortiz-Álvarez, de los Ríos, Fernández-Mendoza, Torralba-Burrial and Pérez-Ortega2015), and the existence of high diversity of photobionts within the lichen thallus such as shown here has likely been overlooked in many lichens. Our findings mirror those of Onuț-Bränström et al. (Reference Onuț-Brännström, Benjamin, Scofield, Heiðmarsson, Andersson, Lindström and Johannesson2018) who discovered multiple Trebouxia genotypes in Thamnolia and Cetratia thalli following Ion Torrent sequencing, compared with a single genotype recovered by Sanger sequencing. More widespread use of community metabarcoding approaches to investigate photobiont diversity may therefore reveal a greater diversity of photobionts and hitherto unrecognized lichen partnerships.

While many terrestrial tripartite lichens are well recognized (Rikkinen et al., Reference Rikkinen, Oksanen and Lohtander2002; Henskens et al., Reference Henskens, Green and Wilkins2012), to the authors' knowledge the presence of Ulvophyceae within the L. pygmaea thallus is the first documented occurrence of a putative chlorobiont in this marine species. This potentially places L. pygmaea amongst several other cyanolichens shown to also contain chlorobionts (Henskens et al., Reference Henskens, Green and Wilkins2012). In particular, L. pygmaea appears to associate with Blidingia, a blade-like multicellular alga that forms an unusual association with Turgidosculum ulvae (Verrucariaceae) whereby the mycobiont colonizes algal filaments (Pérez-Ortega et al., Reference Pérez-Ortega, Miller and Ríos2018). The exact nature of the relationship of L. pygmaea and Blidingia is yet to be established, but hints toward a unique interaction that has been previously overlooked. The other potential photobiont, Paulbroadya is known to form symbioses with several other freshwater and marine lichens in the family Verrucariaceae (Thüs et al., Reference Thüs, Muggia, Pérez-Ortega, Favero-Longo, Joneson, O'Brien, Nelsen, Duque-Thüs, Grube, Friedl, Brodie, Andrew, Lücking, Lutzoni and Gueidan2011; Gasulla et al., Reference Gasulla, Guéra, de los Ríos and Pérez-Ortega2019) (Figure 5). The relationship here between the L. pygmaea-associated Paulbroadya and the same genus associated with the crustose marine lichen H. maura indicates the potential for sharing of photobionts between marine lichens. Photobiont ‘switching’ can occur between closely and distantly related lichens occupying the same habitat (Piercey-Normore & DePriest, Reference Piercey-Normore and DePriest2001; Dal Forno et al., Reference Dal Forno, Lawrey, Sikaroodi, Gillevet, Schuettpelz and Lücking2021), offering a fitness advantage through the acquisition of locally adapted photobionts, and is likely commonplace (Yahr et al., Reference Yahr, Vilgalys and DePriest2006, Reference Yahr, Vilgalys and Depriest2004; Magain et al., Reference Magain, Miadlikowska, Goffinet, Sérusiaux and Lutzoni2017). The potential for photobiont sharing in L. pygmaea is supported by the repeated occurrence of the same Ulvophyceae ASVs between sample sites, and comparisons between photobiont communities in L. pygmaea and H. maura from the same site are now necessary to establish whether the same algal genotypes occur within both species of lichen. Interestingly, this is in contrast to several of the cyanobacterial lineages, including Rivularia, which exhibited variation between sites indicating either reduced dispersal and/or vertical transmission of cyanobionts leading to subsequent localized population structuring.

The phylogenetic placement of many of the potential L. pygmaea photobionts identified in this study within clades dominated by marine-derived lineages suggests that the lichen draws from a pool of marine photobionts, supporting the concept that photobiont selection is influenced by prevailing environmental conditions (Piercey-Normore & Deduke, Reference Piercey-Normore and Deduke2011). Although the full ecological significance of the apparent complex photobiont diversity in L. pygmaea is yet to be determined, a selection of photobionts that enhance survival under different conditions could be a key factor allowing the lichen to occupy the marine environment. Other studies have proposed that the presence of multiple coexisting photobionts with different physiological properties may contribute to increased fitness of the lichen holobiont (Yahr et al., Reference Yahr, Vilgalys and DePriest2006; Casano et al., Reference Casano, del Campo, García-Breijo, Reig-Armiñana, Gasulla, del Hoyo, Guéra and Barreno2011). Since L. pygmaea is able to photosynthesize both in air and submerged in seawater (Raven et al., Reference Raven, Johnston, Handley and McINROY1990), the ability of L. pygmaea to host a variety of photobionts may provide it with an ecological advantage with multiple marine-derived photobionts promoting survival in the intertidal zone by photosynthesizing when in seawater, while the presence of some freshwater-derived photobionts may allow for increased rates of photosynthesis at low tide and during periods of high freshwater input (e.g. rain events).

Further research into the relationship between L. pygmaea and associated photobionts is now necessary, including determining biogeographic factors in symbiont recruitment, establishing relatedness of photobionts between different marine lichen species and the extent of photobiont sharing, and photobiont localization within the thallus. Determining how much these photobionts contribute to carbon and, in the case of the cyanobacteria, nitrogen fixation will be important for understanding their contribution to biogeochemical processes, and relevant physiological experiments may be carried out using growth chambers and other experimental approaches (Henskens et al., Reference Henskens, Green and Wilkins2012). RNA studies could help to quantify photobiont activity in relation to environmental parameters including seasonal, tidal and diel cycles. This work could contribute to our understanding of the evolution of L. pygmaea and adaptation of the Lichinaceae to marine environments.

Supplementary material

The supplementary material for this article can be found at https://doi.org/10.1017/S002531542100062X.

Acknowledgements

The authors wish to thank two anonymous reviewers whose comments significantly improved the quality of the manuscript.

Financial support

NC, RA and MC were supported by the European Research Council (ERC) MYCO-CARB project (grant agreement number 772584).

References

Bubrick, P and Galun, M (1984) Cyanobiont diversity in the Lichinaceae and Heppiaceae. The Lichenologist 16, 279287.CrossRefGoogle Scholar
Callahan, BJ, McMurdie, PJ, Rosen, MJ, Han, AW, Johnson, AJA and Holmes, SP (2016) DADA2: high-resolution sample inference from Illumina amplicon data. Nature Methods 13, 581583.CrossRefGoogle ScholarPubMed
Callahan, BJ, McMurdie, PJ and Holmes, SP (2017) Exact sequence variants should replace operational taxonomic units in marker-gene data analysis. ISME Journal 11, 2639.CrossRefGoogle ScholarPubMed
Caporaso, JG, Lauber, CL, Walters, WA, Berg-Lyons, D, Lozupone, CA, Turnbaugh, PJ, Fierer, N and Knight, R (2011) Global patterns of 16S rRNA diversity at a depth of millions of sequences per sample. Proceedings of the National Academy of Sciences USA 108, 45164522.CrossRefGoogle Scholar
Casano, LM, del Campo, EM, García-Breijo, FJ, Reig-Armiñana, J, Gasulla, F, del Hoyo, A, Guéra, A and Barreno, E (2011) Two Trebouxia algae with different physiological performances are ever-present in lichen thalli of Ramalina farinacea. Coexistence vs competition? Environmental Microbiology 13, 806818.CrossRefGoogle Scholar
Comeau, AM, Li, WKW, Tremblay, J-É, Carmack, EC and Lovejoy, C (2011) Arctic Ocean microbial community structure before and after the 2007 record Sea Ice Minimum. PLoS ONE 6, e27492.CrossRefGoogle ScholarPubMed
Dal Forno, M, Lawrey, J, Sikaroodi, M, Gillevet, PM, Schuettpelz, E and Lücking, R (2021) Extensive photobiont sharing in a rapidly radiating cyanolichen clade. Molecular Ecology 30, 17551776.CrossRefGoogle Scholar
del Campo, EM, Catalá, S, Gimeno, J, del Hoyo, A, Martínez-Alberola, F, Casano, LM, Grube, M and Barreno, E (2013) The genetic structure of the cosmopolitan three-partner lichen Ramalina farinacea evidences the concerted diversification of symbionts. FEMS Microbiology Ecology 83, 310323.CrossRefGoogle ScholarPubMed
Gasulla, F, Guéra, A, de los Ríos, A and Pérez-Ortega, S (2019) Differential responses to salt concentrations of lichen photobiont strains isolated from lichens occurring in different littoral zones. Plant and Fungal Systematics 64, 149162.CrossRefGoogle Scholar
Grube, M, Cardinale, M, de Castro, JV, Müller, H and Berg, G (2009) Species-specific structural and functional diversity of bacterial communities in lichen symbioses. ISME Journal 3, 11051115.CrossRefGoogle ScholarPubMed
Guillou, L, Bachar, D, Audic, S, Bass, D, Berney, C, Bittner, L, Boutte, C, Burgaud, G, de Vargas, C, Decelle, J, Del Campo, J, Dolan, JR, Dunthorn, M, Edvardsen, B, Holzmann, M, Kooistra, WHCF, Lara, E, Le Bescot, N, Logares, R, Mahé, F, Massana, R, Montresor, M, Morard, R, Not, F, Pawlowski, J, Probert, I, Sauvadet, A-L, Siano, R, Stoeck, T, Vaulot, D, Zimmermann, P and Christen, R (2013) The protist ribosomal reference database (PR2): a catalog of unicellular eukaryote small sub-unit rRNA sequences with curated taxonomy. Nucleic Acids Research 41, D597D604.CrossRefGoogle ScholarPubMed
Hawksworth, DL and Grube, M (2020) Lichens redefined as complex ecosystems. New Phytologist 227, 12811283.CrossRefGoogle ScholarPubMed
Henskens, FL, Green, TGA and Wilkins, A (2012) Cyanolichens can have both cyanobacteria and green algae in a common layer as major contributors to photosynthesis. Annals of Botany 110, 555563.CrossRefGoogle Scholar
Honegger, R (1991) Functional aspects of the lichen symbiosis. Annual Review of Plant Physiology and Plant Molecular Biology 42, 553578.CrossRefGoogle Scholar
Love, MI, Huber, W and Anders, S (2014) Moderated estimation of fold change and dispersion for RNA-seq data with DESeq2. Genome Biology 15, 550.CrossRefGoogle ScholarPubMed
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). Systematic Biology 66, 7499.Google Scholar
Mark, K, Laanisto, L, Bueno, CG, Niinemets, Ü, Keller, C and Scheidegger, C (2020) Contrasting co-occurrence patterns of photobiont and cystobasidiomycete yeast associated with common epiphytic lichen species. New Phytologist 227, 13621375.CrossRefGoogle ScholarPubMed
McMurdie, PJ and Holmes, S (2013) Phyloseq: an R package for reproducible interactive analysis and graphics of microbiome census data. PLoS ONE 8, e61217.CrossRefGoogle Scholar
Millanes, AM, Diederich, P and Wedin, M (2016) Cyphobasidium gen. nov., a new lichen-inhabiting lineage in the Cystobasidiomycetes (Pucciniomycotina, Basidiomycota, Fungi). Fungal Biology 120, 14681477.CrossRefGoogle Scholar
Miller, MA, Pfeiffer, W and Schwartz, T (2011) The CIPRES science gateway: a community resource for phylogenetic analyses. In Proceedings of the 2011 TeraGrid Conference: Extreme Digital Discovery, TG ‘11. New York, NY: Association for Computing Machinery, pp. 18. https://doi.org/10.1145/2016741.2016785.CrossRefGoogle Scholar
Moya, P, Molins, A, Martínez-Alberola, F, Muggia, L and Barreno, E (2017) Unexpected associated microalgal diversity in the lichen Ramalina farinacea is uncovered by pyrosequencing analyses. PLoS ONE 12, e0175091.CrossRefGoogle ScholarPubMed
Muggia, L, Vancurova, L, Škaloud, P, Peksa, O, Wedin, M and Grube, M (2013) The symbiotic playground of lichen thalli – a highly flexible photobiont association in rock-inhabiting lichens. FEMS Microbiology Ecology 85, 313323.CrossRefGoogle ScholarPubMed
Naylor, GL (1930) Note on the distribution of Lichina confinis and L. pygmaea in the Plymouth District. Journal of the Marine Biological Association of the United Kingdom 16, 909918.CrossRefGoogle Scholar
Onuț-Brännström, I, Benjamin, M, Scofield, DG, Heiðmarsson, S, Andersson, MGI, Lindström, ES and Johannesson, H (2018) Sharing of photobionts in sympatric populations of Thamnolia and Cetraria lichens: evidence from high-throughput sequencing. Scientific Reports 8, 4406.CrossRefGoogle ScholarPubMed
Ortiz-Álvarez, R, de los Ríos, A, Fernández-Mendoza, F, Torralba-Burrial, A and Pérez-Ortega, S (2015) Ecological specialization of two photobiont-specific maritime cyanolichen species of the genus Lichina. PLoS ONE 10, e0132718.CrossRefGoogle ScholarPubMed
Parrot, D, Antony-Babu, S, Intertaglia, L, Grube, M, Tomasi, S and Suzuki, MT (2015) Littoral lichens as a novel source of potentially bioactive Actinobacteria. Scientific Reports 5, 15839.CrossRefGoogle ScholarPubMed
Pérez-Ortega, S, Miller, KA and Ríos, ADL (2018) Challenging the lichen concept: Turgidosculum ulvae (Verrucariaceae) represents an independent photobiont shift to a multicellular blade-like alga. The Lichenologist 50, 341356.CrossRefGoogle Scholar
Piercey-Normore, MD and Deduke, C (2011) Fungal farmers or algal escorts: lichen adaptation from the algal perspective. Molecular Ecology 20, 37083710.CrossRefGoogle ScholarPubMed
Piercey-Normore, MD and DePriest, PT (2001) Algal switching among lichen symbioses. American Journal of Botany 88, 14901498.CrossRefGoogle ScholarPubMed
Quast, C, Pruesse, E, Yilmaz, P, Gerken, J, Schweer, T, Yarza, P, Peplies, J and Glöckner, FO (2013) The SILVA ribosomal RNA gene database project: improved data processing and web-based tools. Nucleic Acids Research 41, D590D596.CrossRefGoogle ScholarPubMed
R Core Team (2017) R: A Language and Environment for Statistical Computing. Vienna: R Foundation for Statistical Computing.Google Scholar
Raven, JA, Johnston, AM, Handley, LL and McINROY, SG (1990) Transport and assimilation of inorganic carbon by Lichina pygmaea under emersed and submersed conditions. New Phytologist 114, 407417.CrossRefGoogle ScholarPubMed
Rikkinen, J, Oksanen, I and Lohtander, K (2002) Lichen guilds share related cyanobacterial symbionts. Science 297, 357.CrossRefGoogle ScholarPubMed
Ruprecht, U, Brunauer, G and Türk, R (2014) High photobiont diversity in the common European soil crust lichen Psora decipiens. Biodiversity and Conservation 23, 17711785.CrossRefGoogle ScholarPubMed
Smith, HB, Dal Grande, F, Muggia, L, Keuler, R, Divakar, PK, Grewe, F, Schmitt, I, Lumbsch, HT and Leavitt, SD (2020) Metagenomic data reveal diverse fungal and algal communities associated with the lichen symbiosis. Symbiosis 82, 133147.CrossRefGoogle Scholar
Spribille, T, Tuovinen, V, Resl, P, Vanderpool, D, Wolinski, H, Aime, MC, Schneider, K, Stabentheiner, E, Toome-Heller, M, Thor, G, Mayrhofer, H, Johannesson, H and McCutcheon, JP (2016) Basidiomycete yeasts in the cortex of ascomycete macrolichens. Science 353, 488492.CrossRefGoogle ScholarPubMed
Stamatakis, A (2014) RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 13121313.CrossRefGoogle ScholarPubMed
Thüs, H, Muggia, L, Pérez-Ortega, S, Favero-Longo, SE, Joneson, S, O'Brien, H, Nelsen, MP, Duque-Thüs, R, Grube, M, Friedl, T, Brodie, J, Andrew, CJ, Lücking, R, Lutzoni, F and Gueidan, C (2011) Revisiting photobiont diversity in the lichen family Verrucariaceae (Ascomycota). European Journal of Phycology 46, 399415.CrossRefGoogle Scholar
Wang, Q, Garrity, GM, Tiedje, JM and Cole, JR (2007) Naïve Bayesian classifier for rapid assignment of rRNA sequences into the new bacterial taxonomy. Applied and Environmental Microbiology 73, 52615267.CrossRefGoogle ScholarPubMed
Whitton, BA (ed.) (2012) Ecology of Cyanobacteria II. Dordrecht: Springer. https://doi.org/10.1007/978-94-007-3855-3.CrossRefGoogle Scholar
Yahr, R, Vilgalys, R and Depriest, PT (2004) Strong fungal specificity and selectivity for algal symbionts in Florida scrub Cladonia lichens. Molecular Ecology 13, 33673378.CrossRefGoogle ScholarPubMed
Yahr, R, Vilgalys, R and DePriest, PT (2006) Geographic variation in algal partners of Cladonia subtenuis (Cladoniaceae) highlights the dynamic nature of a lichen symbiosis. New Phytologist 171, 847860.CrossRefGoogle ScholarPubMed
Yu, G, Smith, DK, Zhu, H, Guan, Y and Lam, TT-Y (2017) ggtree: an r package for visualization and annotation of phylogenetic trees with their covariates and other associated data. Methods in Ecology and Evolution 8, 2836.CrossRefGoogle Scholar
Figure 0

Fig. 1. (Ai–iii) Lichina pygmaea at Church Reef with close-up of thallus. Dotted line indicates approximate upper limit of the intertidal. Arrows point to L. pygmaea thalli. (B) Sampling locations of L. pygmaea near Plymouth (UK).

Figure 1

Fig. 2. (A) Relative abundance of bacterial and eukaryotic intrathalline and extrathalline communities. (B) Comparison of alpha-diversity between intrathalline (red) and extrathalline (blue) communities of Cyanobacteria and Ulvophyceae, indicating Observed diversity, Shannon Index and Pielou's Evenness. (C) Bray–Curtis NMDS plots showing beta-diversity of Cyanobacteria and Ulvophyceae intrathalline (red) and extrathalline (blue) communities. Shape indicates sampling site.

Figure 2

Fig. 3. Deseq2 plots indicating (A) Cyanobacteria and (B) Ulvophyceae ASVs enriched in the interior (red) and exterior (blue) of the L. pygmaea thallus. Prefix ‘endo-’ refers to intrathalline and ‘exo-’ refers to extrathalline. Circle size is the average of normalized read counts over all samples. Significant enrichments are labelled.

Figure 3

Fig. 4. 16S rRNA gene maximum likelihood phylogeny of Cyanobacteria associated with Lichina pygmaea. ASVs significantly enriched in the intrathalline samples are indicated with*. Prefix ‘endo-’ refers to intrathalline (red circles) and ‘exo-’ refers to extrathalline (blue circles). Circle size is relative abundance. Node labels indicate bootstrap support values for main clades.

Figure 4

Fig. 5. 18S rRNA gene maximum likelihood phylogeny of Ulvophyceae associated with Lichina pygmaea ASVs significantly enriched in the intrathalline samples are indicated with *. Prefix ‘endo-’ refers to intrathalline (red circles) and ‘exo-’ refers to extrathalline (blue circles). Circle size is relative abundance. Node labels indicate bootstrap support values for main clades. Dashed lines next to clade names indicate paraphyletic clades.

Supplementary material: File

Chrismas et al. supplementary material

Chrismas et al. supplementary material

Download Chrismas et al. supplementary material(File)
File 28.6 KB