Hostname: page-component-76fb5796d-9pm4c Total loading time: 0 Render date: 2024-04-25T15:26:18.535Z Has data issue: false hasContentIssue false

Resolving the phylogenetic relationship between Parmotrema crinitum and Parmotrema perlatum populations

Published online by Cambridge University Press:  29 July 2022

Ayoub Stelate*
Affiliation:
Department of Experimental Plant Biology, Charles University Faculty of Science, Viničná 5, 128 44 Prague, Czech Republic Departamento de Farmacología, Farmacognosia y Botánica, Facultad de Farmacia, Universidad Complutense de Madrid, Madrid 28040, Spain Department of Biology, Faculty of Sciences, Mohammed V University, Av Ibn Battouta 4, BP 1014 RP. Rabat, Morocco
Ruth Del-Prado
Affiliation:
Departamento de Farmacología, Farmacognosia y Botánica, Facultad de Farmacia, Universidad Complutense de Madrid, Madrid 28040, Spain
David Alors
Affiliation:
Microalgal Biotechnology Laboratory, Jacob Blaustein Institutes for Desert Research, Ben Gurion University of the Negev, Sede-Boker Campus 8499000, Israel
Hikmat Tahiri
Affiliation:
Department of Biology, Faculty of Sciences, Mohammed V University, Av Ibn Battouta 4, BP 1014 RP. Rabat, Morocco
Pradeep K. Divakar
Affiliation:
Departamento de Farmacología, Farmacognosia y Botánica, Facultad de Farmacia, Universidad Complutense de Madrid, Madrid 28040, Spain
Ana Crespo
Affiliation:
Departamento de Farmacología, Farmacognosia y Botánica, Facultad de Farmacia, Universidad Complutense de Madrid, Madrid 28040, Spain
*
Author for correspondence: Ayoub Stelate. E-mail: Ayoub.stelate@gmail.com

Abstract

The widespread species Parmotrema crinitum (Ach.) M. Choisy and Parmotrema perlatum (Huds.) M. Choisy are mainly distinguished by their reproductive strategies. While P. crinitum propagates by isidia, P. perlatum produces soredia. In this study, we aim to evaluate the phylogenetic relationship between both species and to critically examine their species boundaries. To this purpose, 46 samples belonging to P. crinitum and P. perlatum were used in our analysis, including 22 for which we studied the morphology and chemistry, before extracting their DNA. We used 35 sequences of the internal transcribed spacer region of nuclear ribosomal DNA (ITS) of Parmotrema perlatum from Europe and Africa (20 of which were newly generated), and 11 of Parmotrema crinitum from Europe, North America and North Africa (two newly generated). Additionally, 28 sequences of several species from Parmotrema were included in the ITS dataset. The ITS data matrix was analyzed using different approaches, such as traditional phylogeny (maximum likelihood and Bayesian analyses), genetic distances, automatic barcode gap discovery (ABGD) and the coalescent-based method poisson tree processes (PTP), in order to test congruence among results. Our results indicate that all samples referred to P. crinitum and P. perlatum nested in a well-supported monophyletic clade, but phylogenetic relationships among them remain unresolved. Delimitations inferred from PTP, ABGD and genetic distance analyses were comparable and suggested that P. crinitum and P. perlatum belong to the same lineage. Interestingly, two samples of P. perlatum separate in a different monophyletic clade, which is supported as a different lineage by all the analyses.

Type
Standard Paper
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 (https://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), 2022. Published by Cambridge University Press on behalf of the British Lichen Society

Introduction

Lichenized fungi form mutualistic relationships with photo-autotrophic organisms (photobionts), mainly green algae (Trebouxiophyceae and Ulvophyceae) and/or cyanobacteria. The lichen symbiosis has been highly successful within fungi, especially Ascomycota, with c. 19 400 currently accepted species (Lücking et al. Reference Lücking, Hodkinson and Leavitt2017) and an estimated diversity of more than 28 000 species (Lücking et al. Reference Lücking, Rivas Plata, Chaves, Umaña and Sipman2009; Leavitt et al. Reference Leavitt, Esslinger, Spribille, Divakar and Lumbsch2013). Additionally, lichens are commonly used to assess environmental disturbance, serving as bioindicators of air pollution, forest age and health, and climate change (Nimis et al. Reference Nimis, Scheidegger and Wolseley2002; Crespo et al. Reference Crespo, Divakar, Argüello, Gasca and Hawksworth2004; Giordani & Brunialti Reference Giordani, Brunialti, Upreti, Divakar, Shukla and Bajpai2015; Sujetovienė Reference Sujetovienė, Upreti, Divakar, Shukla and Bajpai2015; Sancho et al. Reference Sancho, Pintado and Green2019; Abas Reference Abas2021).

Recognizing phylogenetic relationships and delimiting species in lichens are crucial for ecological and conservation studies, assessing biotic diversity, and identifying factors driving diversification. They are also important for future investigations because phylogenetic differences may not be fully reflected in the phenotype.

Traditionally, to infer taxonomic boundaries in lichen-forming fungi, thin-layer chromatography (Culberson Reference Culberson1972), morphology and the expression of signature secondary metabolites, and isolation and identification of lichen substances have been used (Huneck & Yoshimura Reference Huneck and Yoshimura1996; Huneck Reference Huneck1999). However, these characters are highly variable and their homology has proved difficult to assess. For example, it has been shown that apothecial form and spore wall thickness have changed in parallel within Pertusaria. The Pertusaria-type ascus is plesiomorphic within the Pertusariaceae and thus cannot be used to circumscribe Pertusaria (Lumbsch & Schmitt Reference Lumbsch and Schmitt2001). However, with the use of molecular tools, it has been shown that phenotype-based taxonomy may not reflect natural groups, including cases in which morphologically distinct forms formerly recognized as distinct species are shown to represent a polymorphic species (see e.g. Boluda et al. Reference Boluda, Rico, Divakar, Nadyeina, Myllys, McMullin, Zamora, Scheidegger and Hawksworth2019), and other cases where several morphologically cryptic species are masked within a single nominal taxon (Del-Prado et al. Reference Del-Prado, Divakar, Lumbsch and Crespo2016).

Within lichen-forming Ascomycetes, Parmeliaceae (Lecanorales) constitutes one of the largest and best-studied families (Crespo et al. Reference Crespo, Lumbsch, Mattsson, Blanco, Divakar, Articus, Wiklund, Bawingan and Wedin2007, Reference Crespo, Divakar and Hawksworth2011; Thell et al. Reference Thell, Crespo, Divakar, Kearnefelt, Leavitt, Lumbsch and Seaward2012; Divakar et al. Reference Divakar, Crespo, Wedin, Leavitt, Hawksworth, Myllys, McCune, Randlane, Bjerke and Ohmura2015). This family is usually characterized morphologically by a certain type of ascoma ontogeny and the presence of an ascomatal structure known as a cupulate exciple (Henssen & Jahns Reference Henssen and Jahns1974). Parmeliaceae also comprises species which are frequently used in biomonitoring studies of atmospheric pollution, such as Parmelia sulcata, Flavoparmelia caperata, Parmotrema perlatum and Punctelia subrudecta (Hawksworth & Rose Reference Hawksworth and Rose1970; Crespo et al. Reference Crespo, Divakar, Argüello, Gasca and Hawksworth2004; De La Cruz et al. Reference De La Cruz, De La Cruz, Tolentino and Gioda2018). The application of phylogenetic analysis based on molecular (DNA) characters to delimit species allows us to determine a posteriori which types of phenotypic characters are good predictors of phylogenetic species and demonstrate how these characters evolve in this family and in lichenized fungi in general. These molecular data have led to the recognition of morphologically cryptic species, such as Parmelia serrana (Molina et al. Reference Molina, Crespo, Blanco, Lumbsch and Hawksworth2004), P. barrenoae (Divakar et al. Reference Divakar, Molina, Lumbsch and Crespo2005a), P. encryptata (Molina et al. Reference Molina, Divakar, Millanes, Sanchez, Del-Prado, Hawksworth and Crespo2011a) and Melanelixia californica (Divakar et al. Reference Divakar, Figueras, Hladun and Crespo2010), and conversely also to the union of species traditionally regarded as morphologically distinct (see e.g. Boluda et al. Reference Boluda, Rico, Divakar, Nadyeina, Myllys, McMullin, Zamora, Scheidegger and Hawksworth2019). Recently, coalescent-based species delimitation approaches have shown to be well suited to critically evaluate species boundaries in Parmeliaceae, as well as lichens in general (Leavitt et al. Reference Leavitt, Moreau, Lumbsch, Upreti, Divakar, Shukla and Bajpai2015). Furthermore, these methods can accurately display relationships when incomplete lineage sorting and gene tree heterogeneity hide phylogenetic relationships among species (Knowles & Carstens Reference Knowles and Carstens2007; Camargo et al. Reference Camargo, Morando, Avila and Sites2012). Commonly used methods to critically evaluate species boundaries include the poisson tree processes (PTP) model (Zhang et al. Reference Zhang, Kapli, Pavlidis and Stamatakis2013), the automatic barcode gap discovery (ABGD) (Puillandre et al. Reference Puillandre, Lambert, Brouillet and Achaz2012), the general mixed Yule coalescent model (GMYC) (Pons et al. Reference Pons, Barraclough, Gomez-Zurita, Cardoso, Duran, Hazell, Kamoun, Sumlin and Vogler2006; Monaghan et al. Reference Monaghan, Wild, Elliot, Fujisawa, Balke, Inward, Lees, Ranaivosolo, Eggleton and Barraclough2009), and SpedeSTEM (Ence & Carstens Reference Ence and Carstens2011).

Parmotrema (Massalongo Reference Massalongo1860) is one of the largest genera in the parmelioid group of the family Parmeliaceae. It includes c. 300 described species with an apparent centre of speciation in the Pacific Islands, tropical and subtropical regions of South America (Spielmann & Marcelli Reference Spielmann and Marcelli2020). The species of the genus are characterized by a pored epicortex, large thalli with broad lobes, a broad, naked marginal zone on the lower surface, and large, thick-walled, ellipsoid ascospores, sublageniform or filiform conidia (Elix Reference Elix1993), and (commonly) marginal cilia (Hale Reference Hale1974). Reproductive strategies vary among Parmotrema taxa. Sexual reproduction is restricted to characteristic fungal fruiting bodies (ascomata) producing ascospores. Ascospores are dispersed independently of the photosynthesizing partner (photobiont) and require reacquisition of the appropriate photobiont partner to re-establish the lichenized condition. Species within Parmotrema also commonly propagate asexually by means of vegetative diaspores, either isidia or soredia. These specialized vegetative reproductive propagules contain both fungal and algal symbionts, eliminating the need for independent acquisition of the appropriate photobiont.

In a molecular phylogeny of parmotremoid lichens (Blanco et al. Reference Blanco, Crespo, Divakar, Elix and Lumbsch2005), a single sample of both Parmotrema perlatum (Huds.) M. Choisy and Parmotrema crinitum (Ach.) M. Choisy, originally from Portugal, were included and these formed a well-supported monophyletic group, indicating that these specimens could be conspecific. Therefore, a critical evaluation of species boundaries is necessary.

Parmotrema crinitum is characterized by the presence of coralloid branched, apically ciliate isidia or often eciliate isidia (Divakar & Upreti Reference Divakar and Upreti2005), and the stictic acid complex in the medulla. According to Elix (Reference Elix1994), P. crinitum is a cosmopolitan species that is widespread in temperate, tropical regions and even in high humidity, sub-boreal forests (Elix Reference Elix1994; Kurokawa & Lai Reference Kurokawa and Lai2001). Many European countries have reported the presence of P. crinitum (Jablońska et al. Reference Jablońska, Oset and Kukwa2009) as have some Asian countries such as Japan (Yoshimura Reference Yoshimura1974), China (Wei Reference Wei1991) and Taiwan (Wang-Yang & Lai Reference Wang-Yang and Lai1976).

Parmotrema perlatum is a greenish grey foliose lichen, saxicolous or corticolous, loosely adnate, with rounded lobes. It can be recognized by its broad lobes, irregularly branched (7.0 cm), laterally overlapping, with frequent black cilia (0.20–)0.50–1.00 × 0.02(–0.10) mm, evenly distributed but less common in the lobe apices. Its soralia are marginal and linear, sometimes widely distributed or subcontinuous, concolorous with the thallus, and the medulla is white. The thallus contains atranorin, stictic acid, hypostictic acid, menegazziaic acid and norstictic acid (Spielmann & Marcelli Reference Spielmann and Marcelli2009). Parmotrema perlatum is a widespread species in temperate regions of the Northern and Southern Hemispheres: Asia (Hale Reference Hale1965; Kurokawa Reference Kurokawa1991; Kurokawa & Lai Reference Kurokawa and Lai2001), Europe (Hale Reference Hale1965), Africa (Hale Reference Hale1965; Swinscow & Krog Reference Swinscow and Krog1988), North America (Hale Reference Hale1965; Brodo et al. Reference Brodo, Sharnoff and Sharnoff2001; Nash & Elix Reference Nash, Elix, Nash, Ryan, Gries and Bungartz2002), Central America (Hale Reference Hale1965), South America (Hale Reference Hale1965), Australia (Kantvilas Reference Kantvilas2019) and New Zealand (Blanchon Reference Blanchon2013).

Morphologically, P. crinitum and P. perlatum can easily be separated based on the characteristics associated with their different dispersal strategies. Parmotrema crinitum has apically ciliate isidia or often eciliate isidia, and P. perlatum has marginal soralia. Parmotrema perlatum is one of the most common and widely utilized lichens in the Ayurvedic system of medicine and has been overexploited for uses in traditional medicine in India (Kumar & Upreti Reference Kumar and Upreti2001). In India, this species is currently considered threatened (Divakar & Upreti Reference Divakar and Upreti2005).

The aim of this study was to evaluate the phylogenetic relationships between P. crinitum and P. perlatum, and elucidate the possible monophyly between both species. Additionally, PTP, ABGD and genetic distance analyses were included to critically assess species boundaries. The ITS data matrix consisted of a total of 74 sequences of Parmotrema species, including 11 of P. crinitum and 35 of P. perlatum, plus two outgroup sequences of Crespoa carneopruinata. Morphological and chemical features of all samples were critically examined.

Material and Methods

Taxon sampling

Sequence data of the ITS locus were analyzed from 76 specimens, of which 11 sequences were of Parmotrema crinitum (two individuals were newly sequenced and nine sequences downloaded from GenBank) and 35 of Parmotrema perlatum (20 individuals were sequenced in this study and 15 sequences downloaded from GenBank). The new specimens were collected from distant geographical regions throughout the species distributions (Table 1). Additionally, 30 ITS sequences from 30 specimens, belonging to 14 different species (6 were newly sequenced and 24 sequences downloaded from GenBank), are included in this study to test the monophyly of both species within the genus. This total included Crespoa carneopruinata which was selected as outgroup since it has previously been shown to be the sister group of Parmotrema (Divakar et al. Reference Divakar, Crespo, Wedin, Leavitt, Hawksworth, Myllys, McCune, Randlane, Bjerke and Ohmura2015). Details of the material studied, including GenBank Accession numbers, are shown in Table 1. Species recognition was based mainly on morphological and chemical characters.

Table 1. Parmotrema species used in this study with GenBank Accession numbers of the ITS sequences and voucher information for the specimens. Crespoa carneopruinata was included in the phylogenetic analyses as outgroup. * = newly generated sequences from specimens for which morphological and chemical characters are given in Table S1 (available online).

Chemistry and morphology

Thallus morphology of all new specimens of Parmotrema crinitum and P. perlatum included in the molecular analyses was studied using a Nikon SMZ-1000 stereomicroscope to identify morphological characteristics of the thallus: lobe shape, isidia, soralia, cilia and rhizines. This is because P. crinitum and P. perlatum are traditionally differentiated based on these features. Photographs were taken with a Nikon 105 mm f/2.8D AF Micro-Nikkor lens coupled to a Nikon D90 camera.

Spot tests were carried out on the medulla with usual chemical reagents such as aqueous potassium hydroxide (K), Steiner's stable paraphenylenediamine (PD) and aqueous calcium hypochlorite (C).

Thin-layer chromatography (TLC) was carried out following Orange et al. (Reference Orange, James and White2010). We used TLC solvent system C (200 ml toluene/30 ml acetic acid) (Elix & Ernst-Russell Reference Elix and Ernst-Russell1993), with concentrated acetone extracts at 50 °C spotted onto silica gel 60 F254 aluminum sheets (Merck, Darmstadt). The aluminum sheets were dried for 10 min in an acetic acid atmosphere to maximize resolution.

DNA extraction, PCR and sequencing

Total DNA was extracted from a single, clean (under a dissecting microscope) lichen lobe using the DNeasy Plant Mini Kit (Qiagen, Barcelona) with a slight modification to the manufacturer's instruction (Crespo et al. Reference Crespo, Blanco and Hawksworth2001). Genomic DNA (5–25 ng, quantified using a quantitative PCR machine) was used for PCR amplifications of the ITS region. Standard PCR amplifications were conducted in 25 μl reaction volumes containing 12.5 μl of Master Mix (50 units/ml of Taq DNA polymerase supplied in a proprietary reaction buffer (pH 8.5), 400 μM dATP, 400 μM dGTP, 400 μM dCTP, 400 μM dTTP, 3 mM MgCl2), and 1.5 μl of each primer at 10 μM, 4.5 μl of water (H2O) and 5 μl of DNA template. Fungal nuclear internal transcribed spacer (ITS) rDNA was amplified using the primer pair ITS1F (5ʹ [CTT GGT CAT TTA GAG GAA GTA A] 3ʹ) (Gardes & Bruns Reference Gardes and Bruns1993)/ LR1 (5ʹ [GGT TGG TTT CTT TTC CT] 3ʹ) (Vilgalys & Hester Reference Vilgalys and Hester1990). We also tried primer pair ITS1-LM (5ʹ [GAA CCT GCG GAA GGA TCA TT] 3ʹ) (Myllys et al. Reference Myllys, Lohtander and Tehler2001) and ITS2-KL (5ʹ [ATG CTT AAG TTC AGC GGG TA] 3ʹ) (Lohtander et al. Reference Lohtander, Källersjö and Tehler1998), but we had better results with primer pair ITS1F/LR1. For any failed samples, we tried a nested PCR using the primer pair ITS1F/LR1 for the first PCR, and 5 μl of this PCR product as DNA template for the second PCR using ITS1-LM/ITS2-KL. With this nested PCR, we aimed to amplify a smaller region than with the previous primers. However, it did not yield any additional products.

Amplification was run in an automatic thermocycler (Techne Progene, Jepson Bolton & Co., Watford, Herts, UK) using the following parameters: initial denaturation at 94 °C for 5 min followed by 35 cycles at 94 °C for 1 min, 54 °C (ITS1F/LR1) for 1 min, and 72 °C for 1 min 30 s, with a final extension at 72 °C for 10 min. Amplification products were visualized on 1% agarose gel stained with SYBR green DNA (Life Technologies Corporation, Grand Island, New York, USA), and subsequently purified using ExoSAP-IT (GE Healthcare, Chalfont St Giles, UK) according to the manufacturer's instructions. Sequencing was performed using BigDye Terminator reaction kit (ABI PRISM, Applied Biosystems, Waltham, Massachusetts, USA). Cycle sequencing reactions were performed with the same sets of primers used in the amplification step. Sequencing reactions were electrophoresed on a 3730 DNA Analyzer (Applied Biosystems) at the Unidad de Genómica (Parque Científico de Madrid).

Sequence alignments and phylogenetic analyses

Sequence fragments generated for this study were assembled and edited using the program SeqMan v. 7 (Lasergene R, DNASTAR, Madison, Wisconsin, USA). Sequence identity was confirmed using the megaBLAST search function in GenBank. For the alignment we prepared a matrix of 480 base pairs (bp). Then we used the program MAFFT v. 7 (Katoh & Standley Reference Katoh and Standley2013) with the parameters set to default. If sequences had different lengths, only the part shared by all the sequences was used, and after manually removing ambiguously aligned nucleotide positions, we kept 451 unambiguously aligned base pairs for the final matrix that we used as input for the phylogenetic reconstruction. The alignment was analyzed using maximum likelihood (ML) and a Bayesian Markov chain Monte Carlo (B/MCMC) approach. The ML analysis was performed using an online version of the program RAxML-HPC BlackBox v. 8.2.8 (Stamatakis Reference Stamatakis2006; Stamatakis et al. Reference Stamatakis, Hoover and Rougemont2008), implemented on the Cipres science gateway (https://www.phylo.org/portal2/home.action; Miller et al. Reference Miller, Pfeiffer and Schwartz2010). We used a GTRGAMMA model, which includes a parameter (Γ) for rate heterogeneity among sites but chose not to include a parameter to estimate the proportion of invariable sites (Stamatakis Reference Stamatakis2006; Stamatakis et al. Reference Stamatakis, Hoover and Rougemont2008). Support values were assessed using the ‘rapid bootstrapping’ option with 1000 replicates.

For the Bayesian reconstruction, we used MrBayes v. 3.2.1 (Ronquist & Huelsenbeck Reference Ronquist and Huelsenbeck2003). The analysis was performed assuming a discrete gamma distribution with six rate categories (GTR + G). The nucleotide-substitution model parameters were selected using the Akaike Information Criterion as implemented in jModelTest (Posada Reference Posada2008), molecular clock not assumed. A run with 10 million generations, starting with a random tree and employing 12 simultaneous chains, was executed. Trees were saved to a file every 200th generation. The first 2 million generations (i.e. 20 000 trees) were deleted as the ‘burn-in’ of the chains. We plotted the log-likelihood scores of sample points against generations using Tracer v. 1.5 (Rambaut & Drummond Reference Rambaut and Drummond2007) and determined that stationarity had been achieved when the log-likelihood values of the sample points reached an equilibrium value (Huelsenbeck & Ronquist Reference Huelsenbeck and Ronquist2001). The trees obtained before stationarity were discarded. Posterior probabilities (PPs) were obtained from the 50% majority-rule consensus of sampled trees after excluding the initial 20% as burn-in. Only clades that received bootstrap support ≥ 70% in the ML analyses and PP ≥ 0.95 were considered strongly supported. The phylogenetic tree was drawn using FigTree v. 1.2.3 (Rambaut Reference Rambaut2009) and modified with CorelDRAW v. 8.

Candidate species identification

In order to establish candidate species limits in the phylogenetic tree, three computational approaches were used:

  1. 1) Poisson tree processes (PTP): this method does not require an ultra-metric tree, as the transition point between intra- and inter-specific branching rates is identified directly using the number of nucleotide substitutions (Zhang et al. Reference Zhang, Kapli, Pavlidis and Stamatakis2013). PTP incorporates the number of substitutions in the model of speciation and assumes that the probability of a substitution giving rise to a speciation event follows a Poisson distribution. The branch lengths of the input tree are supposed to be generated by two independent classes of Poisson events, one corresponding to speciation and the other to coalescence. The ML phylogeny of the ITS dataset obtained with RAxML was used as the input trees to run PTP species delimitation analysis on the PTP web server (http://species.h-its.org/ptp/). We ran the PTP analysis for 100 000 MCMC generations, with a thinning value of 100 and a burn-in of 25%. Outgroup taxa were removed for species delimitation.

  2. 2) Automatic barcode gap discovery (ABGD): this is an automatic procedure that sorts the sequences into hypothetical species based on the barcode gap. This method automatically finds the distance where the barcode gap is located (Puillandre et al. Reference Puillandre, Lambert, Brouillet and Achaz2012). The ABGD method was carried out on the ITS dataset using the web interface at http://wwwabi.snv.jussieu.fr/public/abgd/abgdweb.html. Default parameters were chosen using Kimura 2-parameter (K2P) distances that correct for transition rate bias (relative to transversions) in the substitution process (Kimura Reference Kimura1980). The default for the minimum relative gap width was set to different values between 0.1 and 0.15.

  3. 3) Genetic distances: pairwise ML distances (given as the number of nucleotide substitutions per site) among the ITS rDNA sequences of Parmotrema crinitum and P. perlatum were calculated. Genetic distances were calculated with TREE-PUZZLE v. 5.2 (Schmidt et al. Reference Schmidt, Strimmer, Vingron and von Haeseler2002) using the GTR model of nucleotide substitution, assuming a discrete gamma distribution with six rate categories. The program generates an output file which consists of a triangular matrix with all pairwise distances between all the samples included. This matrix was visualized with Microsoft Excel 2010 and genetic distances between different specimens of P. crinitum and P. perlatum were manually identified. Candidate species were proposed based on the threshold of 0.016 substitutions per site (s/s) which separates intra- and interspecific distances in parmelioid lichens (Del-Prado et al. Reference Del-Prado, Cubas, Lumbsch, Divakar, Blanco, de Paz G, Molina and Crespo2010). Distance values in the matrix ≤ 0.016 s/s have been considered as the values between samples of a single species. By using the Excel filter, we separated values ≤ 0.016, providing for every specimen included in the analysis a group of specimens that share the values characterizing its species range.

Results

Morphological and chemical analyses

Morphological and chemical characters of the specimens (P. crinitum and P. perlatum) newly generated and included in the phylogenetic analysis are presented in Supplementary Material Table S1, available online. Parmotrema crinitum and P. perlatum could not be distinguished based on the colour of the thalli, presence of cilia (Fig. 1A & B) and colour of the lower surface (Fig. 1E & F). However, the reproductive structures allowed unequivocal separation into the two species; Parmotrema perlatum has soralia (Fig. 1C) and P. crinitum has isidia (Fig. 1D).

Fig. 1. Main morphological characters of P. perlatum (A, C, E) and P. crinitum. (B, D, F), as indicated with arrows. A & B, black, simple cilia. C, marginal soralia concolorous with the thallus. D, isidia with apical cilia. E & F, black lower surface with brown naked zone peripherally. In colour online.

The TLC test showed that both species contained the stictic acid complex (stictic acid, menegazzic acid and hypostictic acid) and atranorin. The spot tests of P. crinitum were similar to those of P. perlatum (Fig. 2, Supplementary Material Table S1). However, the samples Parmotrema perlatum 34 and P. perlatum 35 lacked menegazzic acid (nos 10 and 12 in Fig. 2). The spot test was congruent with other samples of P. perlatum, and the morphology of both samples showed the same characteristics of P. perlatum, although the cilia were not as abundant as is usual in this species.

Fig. 2. Thin-layer chromatography profile of Parmotrema species (included in the DNA analysis) in solvent system C. P is the control, Pleurosticta acetabulum. Lanes 1 & 7, P. dilatatum. Lane 2, P. flavotinctum. Lane 3, P. mellissii. Lane 4, P. paulense. Lanes 5, 6, 8, 11 & 19, P. crinitum. Lane 9, empty. Lanes 10 & 12, Parmotrema perlatum. Lanes 13, 14, 15, 16, 17, 18 & 20, P. perlatum; these have the same TLC result as P. crinitum but were distinguished from each other based on the morphology. a = norstictic acid; b = atranorin; c = protocetraric acid; d = unknown; e = hypostictic acid; f = stictic acid; g = menegazzic acid. In colour online.

Phylogenetic analysis

The ITS data matrix consisted of a total of 74 sequences of Parmotrema species, including 46 of P. crinitum and P. perlatum (11 sequences of P. crinitum and 35 sequences of P. perlatum), with two sequences of Crespoa carneopruinata were included as outgroup (Table 1). The final matrix used as input for the phylogenetic reconstruction contained 451 unambiguously aligned base pairs (bp). The tree reconstruction (Fig. 3) comprised the following species of Parmotrema, included to test the monophyly of P. crinitum and P. perlatum: Parmotrema clavuliferum, P. dilatatum, P. flavotinctum, P. grayanum, P. haitiense, P. hypoleucinum, P. mellissii, P. paulense, P. perforatum, P. pseudoreticulatum, P. reticulatum, P. robustum, P. subtinctorium and P. tinctorum.

Fig. 3. Phylogenetic relationships between Parmotrema crinitum and P. perlatum. The green (thickened) line indicates the clade supported by maximum likelihood (ML) and Bayesian analyses; * = clades supported only by ML. ** = clades supported only by Bayesian analysis. Alphanumeric labels indicate clades and clusters. Species delimitation scenarios obtained from PTP, ABGD and genetic distances are indicated in columns to the right. In colour online.

The phylogenetic trees estimated from ML and Bayesian methods did not show any well-supported conflict; ML topologies are presented with bootstrap and B/MCMC analysis with posterior probability (ML bootstrap ≥ 70%; PP ≥ 0.95 in B/MCMC analysis) (Fig. 3). The LnL value was −1938.068210 for ML, and −2184.428 for the Bayesian analysis. The phylogenetic analysis showed that all samples of Parmotrema included in this analysis formed a monophyletic group. Within this genus the samples of Parmotrema crinitum and P. perlatum were grouped in one monophyletic group (clade A), with the exception of two samples of P. perlatum that were grouped in a separate clade B. While clade A comprised specimens of P. crinitum and P. perlatum collected from various geographical regions, clade B included only two specimens of P. perlatum collected from Tenerife (Canary Islands) and Lisbon (Portugal).

Although four well-supported monophyletic clusters can be recognized in clade A, the phylogenetic relationships among them remained unresolved (Fig. 3).

Cluster A1 (Fig. 3) contained 15 specimens of Parmotrema perlatum, most of them collected from Morocco, except four from the Canary Islands and two from Portugal and Turkey. Cluster A2 included eight specimens of Parmotrema crinitum from different geographical regions (Table 1). Cluster A3 grouped 18 samples of Parmotrema perlatum from the Canary Islands (Tenerife, Palma, Gomera), and cluster A4 included three samples of P. crinitum from the Canary Islands.

Identifying candidate species

We used the same RaxML tree obtained from the phylogenetic analysis to illustrate the delimitation of putative species recognized by the different approaches conducted with the ITS dataset. ABGD, PTP and genetic distance analyses applied to the ITS dataset detected two candidate species that corresponded to the well-supported clades A (including P. crinitum + P. perlatum) and B (including P. perlatum) obtained in the phylogenetic analysis (Fig. 3).

Discussion

Previous molecular phylogenetic studies have proposed the monophyly of Parmotrema crinitum and Parmotrema perlatum (Blanco et al. Reference Blanco, Crespo, Ree and Lumbsch2006; Crespo et al. Reference Crespo, Kauff, Divakar, del Prado, Pérez-Ortega, de Paz G, Ferencova, Blanco, Roca-Valiente and Núñez-Zapata2010). The primary goal of the present investigation was to evaluate the phylogenetic relationship between the two species using ML and Bayesian analyses. Additionally, we used different approaches to assess the species boundary, such as genetic distances based on the threshold of 0.015–0.017 s/s established by Del-Prado et al. (Reference Del-Prado, Cubas, Lumbsch, Divakar, Blanco, de Paz G, Molina and Crespo2010) to measure intra- and interspecific genetic distances in parmelioid lichens, and separate ranges of intra- and interspecific divergence. This threshold was established using both phylogenetically and morphologically well-delimited species (for details see Del-Prado et al. (Reference Del-Prado, Cubas, Lumbsch, Divakar, Blanco, de Paz G, Molina and Crespo2010)). We also used the poisson tree processes (PTP) model (Zhang et al. Reference Zhang, Kapli, Pavlidis and Stamatakis2013) and the automatic barcode gap discovery (ABGD; Puillandre et al. Reference Puillandre, Lambert, Brouillet and Achaz2012). Furthermore, we combined ecological, biogeographical, morphological and chemical data.

Most of the samples belonging to P. crinitum and P. perlatum were recovered in a well-supported monophyletic clade (clade A, Fig. 3), with the exception of two samples of P. perlatum, one from Tenerife (Canary Islands) and one from Lisbon (Portugal), that are separated in the monophyletic clade B, supported by ML and B/MCMC analysis. Morphologically, P. perlatum in clade A has more cilia compared to P. perlatum in clade B.

Thin-layer chromatography (TLC) revealed that the samples in clade B (nos 10 and 12 in Fig. 2) differ chemically from the samples of Parmotrema perlatum grouped in clade A. Several studies have shown that environmental factors, such as light, temperature, pH and culture media, can influence the secondary metabolism in lichens (BeGora & Fahselt Reference BeGora and Fahselt2001). However, we have samples from the same localities and same conditions as the samples in clade A. Furthermore, ABGD, PTP and genetic distance analyses supported clade B as a new sister group to clade A (Fig. 3), suggesting polyphyly, a common phenomenon in Parmeliaceae in general (Lumbsch & Leavitt Reference Lumbsch and Leavitt2011; Leavitt et al. Reference Leavitt, Divakar, Crespo and Lumbsch2016) and Parmotrema in particular (Divakar et al. Reference Divakar, Blanco, Hawksworth and Crespo2005b; Del-Prado et al. Reference Del-Prado, Divakar, Lumbsch and Crespo2016, Reference Del-Prado, Buaruang, Lumbsch, Crespo and Divakar2019; Widhelm et al. Reference Widhelm, Egan, Bertoletti, Asztalos, Kraichak, Leavitt and Lumbsch2016).

Clade A was split into 4 groups; however, the phylogenetic relationships among them remained unresolved (Fig. 3). Previous studies have shown that phylogenetic analyses alone are insufficient to explain phylogenetic relationships within Parmeliaceae. For example, based only on maximum parsimony and Bayesian analyses the sorediate P. sulcata was shown to belong to the same clade as the isidiate P. squarrosa (Molina et al. Reference Molina, Crespo, Blanco, Lumbsch and Hawksworth2004), leaving the authors unable to reach any conclusion regarding species boundaries. Similarly, phylogenetic analyses were insufficient to resolve genetic variability among Parmelia saxatilis specimens; whereas samples from distant geographical regions formed a monophyletic group, samples from neighbouring localities were separated (Crespo et al. Reference Crespo, Molina, Blanco, Schroeter, Sancho and Hawksworth2002; Molina et al. Reference Molina, Del-Prado, Divakar, Sánchez-Mata and Crespo2011b, Reference Molina, Divakar, Goward, Millanes, Lumbsch and Crespo2017). A study on Usnea perpusilla demonstrated that it was necessary to use a combined approach with molecular and morphological data to assess species boundaries in closely related and morphologically variable species (Wirtz et al. Reference Wirtz, Printzen and Lumbsch2008). For this reason, coalescent-based species delimitation analyses have been applied with the goal of explaining relationships among clades and delimiting species boundaries (Parnmen et al. Reference Parnmen, Rangsiruji, Mongkolsuk, Boonpragob, Nutakki and Lumbsch2012; Leavitt et al. Reference Leavitt, Esslinger, Spribille, Divakar and Lumbsch2013).

PTP, ABGD and genetic distance analyses supported clades A and B as putative distinct species. The values obtained were within the interspecific ranges (genetic distances as defined in Del-Prado et al. (Reference Del-Prado, Cubas, Lumbsch, Divakar, Blanco, de Paz G, Molina and Crespo2010)), which would support the two samples of P. perlatum in clade B as a separate species, different from clade A. Focusing on clade A, our species delimitation approaches (PTP, ABGD and genetic distance) supported P. crinitum and P. perlatum as one species. However, based only on one molecular marker (ITS) and the various phylogenetic analyses and species delimitation approaches used, can we consider Parmotrema perlatum conspecific with P. crinitum?

Similar cases have been reported in the Parmotrema perforatum species complex (Widhelm et al. Reference Widhelm, Egan, Bertoletti, Asztalos, Kraichak, Leavitt and Lumbsch2016), and authors have suggested that the phylogenetic relationships between sexual and asexual populations of this species group could be more complex than previously assumed. They also suggest that traditional tools based on reproductive mode and secondary metabolites (Culberson & Culberson Reference Culberson and Culberson1973) are no longer the key to identify species such as P. perforatum (Widhelm et al. Reference Widhelm, Egan, Bertoletti, Asztalos, Kraichak, Leavitt and Lumbsch2016).

A previous study on Umbilicaria (Ott et al. Reference Ott, Brinkmann, Wirtz and Lumbsch2004) focused on Umbilicaria kappenii and U. antarctica, which are distinguished only by their reproductive strategies. Umbilicaria antarctica propagates by thalloconidia and U. kappenii exhibits a variety of asexual propagules: soredia, adventive lobes and sorediate thallyles. To infer phylogenetic relationships between both species, the authors used molecular data from three loci. Results indicated that all samples morphologically referred to U. antarctica and U. kappenii form a monophyletic group and they proposed placing U. kappenii into synonymy with U. antarctica (Ott et al. Reference Ott, Brinkmann, Wirtz and Lumbsch2004). In Parmelia, Molina et al. (Reference Molina, Del-Prado, Divakar, Sánchez-Mata and Crespo2011b) rejected the previous hypothesis that P. sulcata and P. squarrosa form a monophyletic group (Molina et al. Reference Molina, Crespo, Blanco, Lumbsch and Hawksworth2004) and based on phylogenetic analyses and species delimitation approaches confirmed that P. sulcata is not conspecific with P. squarrosa. In addition, P. squarrosa is a reproductively isolated lineage and genetic distances clearly separate this from other Parmelia species (Del-Prado et al. Reference Del-Prado, Cubas, Lumbsch, Divakar, Blanco, de Paz G, Molina and Crespo2010). Within Usnea, studies have suggested that Usnea subfloridana was a secondary species derived from U. florida (Clerc Reference Clerc1984, Reference Clerc1987, Reference Clerc1997; Purvis et al. Reference Purvis, Coppins, Hawksworth, James and Moore1992). Usnea florida and U. subfloridana show many morphological similarities but they have different reproductive strategies. Usnea florida always displays many apothecia and produces no specialized asexual propagules. Usnea subfloridana has soralia, isidiomorphs and occasionally apothecia. Multilocus phylogenetic analyses based on sequences of the ITS, IGS and LSU regions of the nuclear ribosomal DNA and the gene coding for β-tubulin, Mcm7, RPB1 and RPB2 showed that specimens of the two morphospecies formed one monophyletic intermixed group, and not two groups corresponding to morphology (Articus et al. Reference Articus, Mattsson, Tibell, Grube and Wedin2002; Mark et al. Reference Mark, Cornejo, Keller, Flück and Scheidegger2016). This topology was further confirmed with a coalescent-based species delimitation approach (Mark et al. Reference Mark, Cornejo, Keller, Flück and Scheidegger2016). Authors have suggested that they could be conspecific but taxonomic conclusions must await further study. Moreover, a recent study using RADseq data suggests that closely related lichen species may need genome-wide data to test their species boundaries (Grewe et al. Reference Grewe, Lagostina, Wu, Printzen and Lumbsch2018).

Traditionally it was thought that asexually reproducing species in lichens, and in filamentous fungi in general, was an evolutionary dead end (Normark et al. Reference Normark, Judson and Moran2003). However, recent molecular studies have demonstrated that lineages with vegetative propagation can also present high genetic diversity (e.g. Parmelia sulcata (Molina et al. Reference Molina, Del-Prado, Divakar, Sánchez-Mata and Crespo2011b), Parmotrema reticulatum (Del-Prado et al. Reference Del-Prado, Divakar, Lumbsch and Crespo2016)). However, even in the absence of sexual reproduction, lichens can exchange photobionts and this process could provide opportunities for gene transfer (Piercey-Normore Reference Piercey-Normore2006; Nelsen & Gargas Reference Nelsen and Gargas2008).

While our study confirmed the monophyly of an intermixed clade of Parmotrema crinitum and P. perlatum, the taxonomic conclusion must await additional studies including more markers. The phylogenetic tree lacked ML or B/MCMC support for other widely accepted Parmotrema species, such as P. pseudoreticulatum and P. reticulatum. A more comprehensive taxon sampling and additional molecular markers will therefore be needed before making a formal taxonomic decision on the status of clade B.

Acknowledgements

This study was supported by the Spanish Ministerio de Ciencia e Innovacion (PID2019-105312GB-I00) and the Santander-Universidad Complutense de Madrid (PR87/19-22637 and G/6400100/3000).

Author ORCIDs

Ayoub Stelate, 0000-0001-8929-3046; David Alors, 0000-0001-7288-9521; Pradeep Kumar Divakar, 0000-0002-0300-0124; Ana Crespo, 0000-0002-5271-0157.

Supplementary Material

To view Supplementary Material for this article, please visit https://doi.org/10.1017/S0024282922000147.

References

Abas, A (2021) A systematic review on biomonitoring using lichen as the biological indicator: a decade of practices, progress and challenges. Ecological Indicators 121, 107197.CrossRefGoogle Scholar
Articus, K, Mattsson, JE, Tibell, L, Grube, M and Wedin, M (2002) Ribosomal DNA and B-tubulin data do not support the separation of the lichens Usnea florida and U. subfloridana as distinct species. Mycological Research 106, 412418.CrossRefGoogle Scholar
BeGora, MD and Fahselt, D (2001) Usnic acid and atranorin concentrations in lichens in relation to bands of UV irradiance. Bryologist 104, 134140.CrossRefGoogle Scholar
Blanchon, DJ (2013) Auckland lichens. Auckland Botanical Society Journal 68, 2127.Google Scholar
Blanco, O, Crespo, A, Divakar, PK, Elix, JA and Lumbsch, HT (2005) Molecular phylogeny of parmotremoid lichens (Ascomycota, Parmeliaceae). Mycologia 97, 150159.CrossRefGoogle Scholar
Blanco, O, Crespo, A, Ree, RH and Lumbsch, HT (2006) Major clades of parmelioid lichens (Parmeliaceae, Ascomycota) and the evolution of their morphological and chemical diversity. Molecular Phylogenetics and Evolution 39, 5269.CrossRefGoogle ScholarPubMed
Boluda, CG, Rico, VJ, Divakar, PK, Nadyeina, O, Myllys, L, McMullin, RT, Zamora, JC, Scheidegger, C and Hawksworth, DL (2019) Evaluating methodologies for species delimitation: the mismatch between phenotypes and genotypes in lichenized fungi (Bryoria sect. Implexae, Parmeliaceae). Persoonia 42, 75100.CrossRefGoogle ScholarPubMed
Brodo, IM, Sharnoff, SD and Sharnoff, S (2001) Lichens of North America. New Haven, Connecticut: Yale University Press.Google Scholar
Camargo, A, Morando, M, Avila, LJ and Sites, JW (2012) Species delimitation with ABC and other coalescent-based methods: a test of accuracy with simulations and an empirical example with lizards of the Liolaemus darwinii complex (Squamata: Liolaemidae). Evolution 66, 28342849.CrossRefGoogle Scholar
Clerc, P (1984) Contribution à la révision de la systématique des Usnées (Ascomycotina, Usnea) d'Europe. 1. Usnea florida (L.) Wigg. emend. Clerc. Cryptogamie, Bryologie et Lichénologie 5, 333360.Google Scholar
Clerc, P (1987) Systematics of the Usnea fragilescens aggregate and its distribution in Scandinavia. Nordic Journal of Botany 7, 479495.CrossRefGoogle Scholar
Clerc, P (1997) Notes on the genus Usnea Dill. ex Adanson. Lichenologist 29, 209215.CrossRefGoogle Scholar
Crespo, A, Blanco, O and Hawksworth, DL (2001) The potential of mitochondrial DNA for establishing phylogeny and stabilizing generic concepts in the parmelioid lichens. Taxon 50, 807819.CrossRefGoogle Scholar
Crespo, A, Molina, MC, Blanco, O, Schroeter, B, Sancho, LG and Hawksworth, DL (2002) rDNA ITS and β-tubulin gene sequence analyses reveal two monophyletic groups within the cosmopolitan lichen Parmelia saxatilis. Mycological Research 106, 788795.CrossRefGoogle Scholar
Crespo, A, Divakar, PK, Argüello, A, Gasca, C and Hawksworth, DL (2004) Molecular studies on Punctelia species of the Iberian Peninsula, with an emphasis on specimens newly colonizing Madrid. Lichenologist 36, 299308.CrossRefGoogle Scholar
Crespo, A, Lumbsch, HT, Mattsson, JE, Blanco, O, Divakar, PK, Articus, K, Wiklund, E, Bawingan, PA and Wedin, M (2007) Testing morphology-based hypotheses of phylogenetic relationships in Parmeliaceae (Ascomycota) using three ribosomal markers and the nuclear RPB1 gene. Molecular Phylogenetics and Evolution 44, 812824.CrossRefGoogle ScholarPubMed
Crespo, A, Kauff, F, Divakar, PK, del Prado, R, Pérez-Ortega, S, de Paz G, Amo, Ferencova, Z, Blanco, O, Roca-Valiente, B, Núñez-Zapata, J, et al. (2010) Phylogenetic generic classification of parmelioid lichens (Parmeliaceae, Ascomycota) based on molecular, morphological and chemical evidence. Taxon 59, 17351753.CrossRefGoogle Scholar
Crespo, A, Divakar, PK and Hawksworth, DL (2011) Generic concepts in parmelioid lichens, and the phylogenetic value of characters used in their circumscription. Lichenologist 43, 511535.CrossRefGoogle Scholar
Culberson, CF (1972) Improved conditions and new data for identification of lichen products by a standardized thin-layer chromatographic method. Journal of Chromatography A 72, 113125.CrossRefGoogle ScholarPubMed
Culberson, WL and Culberson, CF (1973) Parallel evolution in the lichen-forming fungi. Science 180, 196198.CrossRefGoogle Scholar
De La Cruz, ARH, De La Cruz, JKH, Tolentino, DA and Gioda, A (2018) Trace element biomonitoring in the Peruvian Andes metropolitan region using Flavoparmelia caperata lichen. Chemosphere 210, 849858.CrossRefGoogle ScholarPubMed
Del-Prado, R, Cubas, P, Lumbsch, HT, Divakar, PK, Blanco, O, de Paz G, Amo, Molina, MC and Crespo, A (2010) Genetic distances within and among species in monophyletic lineages of Parmeliaceae (Ascomycota) as a tool for taxon delimitation. Molecular Phylogenetics and Evolution 56, 125–33.CrossRefGoogle ScholarPubMed
Del-Prado, R, Divakar, PK, Lumbsch, HT and Crespo, A (2016) Hidden genetic diversity in an asexually reproducing lichen forming fungal group. PLoS ONE 11, e0161031.CrossRefGoogle Scholar
Del-Prado, R, Buaruang, K, Lumbsch, HT, Crespo, A and Divakar, PK (2019) DNA sequence-based identification and barcoding of a morphologically highly plastic lichen forming fungal genus (Parmotrema, Parmeliaceae) from the tropics. Bryologist 122, 281291.CrossRefGoogle Scholar
Divakar, PK and Upreti, DK (2005) Parmelioid Lichens in India (A Revisionary Study). Dehradun: Bishen Singh Mahendra Pal Singh.Google Scholar
Divakar, PK, Molina, MC, Lumbsch, HT and Crespo, A (2005 a) Parmelia barrenoae, a new lichen species related to Parmelia sulcata (Parmeliaceae) based on molecular and morphological data. Lichenologist 37, 3746.CrossRefGoogle Scholar
Divakar, PK, Blanco, O, Hawksworth, DL and Crespo, A (2005 b) Molecular phylogenetic studies on the Parmotrema reticulatum (syn. Rimelia reticulata) complex, including the confirmation of P. pseudoreticulatum. Lichenologist 37, 5565.CrossRefGoogle Scholar
Divakar, PK, Figueras, G, Hladun, N and Crespo, A (2010) Molecular phylogenetic studies reveal an undescribed species within the North American concept of Melanelixia glabra (Parmeliaceae). Fungal Diversity 42, 4755.CrossRefGoogle Scholar
Divakar, PK, Crespo, A, Wedin, M, Leavitt, SD, Hawksworth, DL, Myllys, L, McCune, B, Randlane, T, Bjerke, JW, Ohmura, Y, et al. (2015) Evolution of complex symbiotic relationships in a morphologically derived family of lichen-forming fungi. New Phytologist 208, 12171226.CrossRefGoogle Scholar
Elix, JA (1993) Progress in the generic delimitation of Parmelia sensu lato lichens (Ascomycotina: Parmeliaceae) and a synoptic key to the Parmeliaceae. Bryologist 96, 359383.CrossRefGoogle Scholar
Elix, JA (1994) Flora of Australia. Volume 55. Lichens: Lecanorales. 2. Parmeliaceae. Canberra: Australian Biological Resources Study.Google Scholar
Elix, JA and Ernst-Russell, KD (1993) A Catalogue of Standardized Thin Layer Chromatographic Data and Biosynthetic Relationships for Lichen Substances. 2nd Edn. Canberra: Australian National University.Google Scholar
Ence, DD and Carstens, BC (2011) SpedeSTEM: a rapid and accurate method for species delimitation. Molecular Ecology Resources 11, 473480.CrossRefGoogle ScholarPubMed
Gardes, M and Bruns, TD (1993) ITS primers with enhanced specificity for basidiomycetes – application to the identification of mycorrhizae and rusts. Molecular Ecology 2, 113118.CrossRefGoogle Scholar
Giordani, P and Brunialti, G (2015) Sampling and interpreting lichen diversity data for biomonitoring purposes. In Upreti, DK, Divakar, PK, Shukla, V and Bajpai, R (eds), Recent Advances in Lichenology, Vol. 1. New Delhi: Springer, pp. 1946.Google Scholar
Grewe, F, Lagostina, E, Wu, H, Printzen, C and Lumbsch, HT (2018) Population genomic analyses of RAD sequences resolves the phylogenetic relationship of the lichen-forming fungal species Usnea antarctica and Usnea aurantiacoatra. MycoKeys 43, 91113.CrossRefGoogle Scholar
Hale, ME (1965) A monograph of Parmelia subgenus Amphigymnia. Contributions from the United States National Herbarium 36, 193358.Google Scholar
Hale, ME (1974) Bulbothrix, Parmelina, Relicina, and Xanthoparmelia, four new genera in the Parmeliaceae. Phytologia 28, 479490.Google Scholar
Hawksworth, DL and Rose, L (1970) Qualitative scale of estimating sulfur dioxide air pollution in England and Wales using epiphytic lichens. Nature 227, 145148.CrossRefGoogle ScholarPubMed
Henssen, A and Jahns, HM (1974) Lichenes. Eine Einfuhrung in die Flechtenkunde. Stuttgart: Georg Thieme Verlag.Google Scholar
Huelsenbeck, JP and Ronquist, F (2001) MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17, 754755.CrossRefGoogle ScholarPubMed
Huneck, S (1999) The significance of lichens and their metabolites. Naturwissenschaften 86, 559570.CrossRefGoogle ScholarPubMed
Huneck, S and Yoshimura, I (1996) Identification of Lichen Substances. New York: Springer.CrossRefGoogle Scholar
Jablońska, A, Oset, M and Kukwa, M (2009) The lichen family Parmeliaceae in Poland. I. The genus Parmotrema. Acta Mycologica 44, 211222.CrossRefGoogle Scholar
Kantvilas, G (2019) An annotated catalogue of the lichens of Kangaroo Island, South Australia. Swainsona 32, 197.Google Scholar
Katoh, K and Standley, DM (2013) MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Molecular Biology and Evolution 30, 772780.CrossRefGoogle ScholarPubMed
Kimura, M (1980) A simple method for estimating evolutionary rate of base substitution through comparative studies of nucleotide sequences. Journal of Molecular Evolution 16, 111120.CrossRefGoogle Scholar
Knowles, LL and Carstens, BC (2007) Delimiting species without monophyletic gene trees. Systematic Biology 56, 887895.CrossRefGoogle ScholarPubMed
Kumar, K and Upreti, DK (2001) Parmelia spp. (Lichens) in ancient medicinal plant lore of India. Economic Botany 55, 458459.Google Scholar
Kurokawa, S (1991) Japanese species and genera of the Parmeliaceae. Journal of Japanese Botany 66, 152159.Google Scholar
Kurokawa, S and Lai, MJ (2001) Parmelioid lichen genera and species in Taiwan. Mycotaxon 77, 225284.Google Scholar
Leavitt, SD, Esslinger, TL, Spribille, T, Divakar, PK and Lumbsch, HT (2013) Multilocus phylogeny of the lichen-forming fungal genus Melanohalea (Parmeliaceae, Ascomycota): insights on diversity, distributions, and a comparison of species tree and concatenated topologies. Molecular Phylogenetics and Evolution 66, 138152.CrossRefGoogle Scholar
Leavitt, SD, Moreau, CS and Lumbsch, HT (2015) The dynamic discipline of species delimitation: progress toward effectively recognizing species boundaries in natural populations. In Upreti, DK, Divakar, PK, Shukla, V and Bajpai, R (eds), Recent Advances in Lichenology, Vol. 2. New Delhi: Springer, pp. 1144.CrossRefGoogle Scholar
Leavitt, SD, Divakar, PK, Crespo, A and Lumbsch, HT (2016) A matter of time – understanding the limits of the power of molecular data for delimiting species boundaries. Herzogia 29, 479492.CrossRefGoogle Scholar
Lohtander, K, Källersjö, M and Tehler, A (1998) Dispersal strategies in Roccellina capensis (Arthoniales). Lichenologist 30, 341350.CrossRefGoogle Scholar
Lücking, R. Rivas Plata, E, Chaves, JL, Umaña, L and Sipman, HJM (2009) How many tropical lichens are there… really? Bibliotheca Lichenologica 100, 399418.Google Scholar
Lücking, R, Hodkinson, BP and Leavitt, SD (2017) The 2016 classification of lichenized fungi in the Ascomycota and Basidiomycota – approaching one thousand genera. Bryologist 119, 361416.CrossRefGoogle Scholar
Lumbsch, HT and Leavitt, SD (2011) Goodbye morphology? A paradigm shift in the delimitation of species in lichenized fungi. Fungal Diversity 50, 5972.CrossRefGoogle Scholar
Lumbsch, HT and Schmitt, I (2001) Molecular data suggest that the lichen genus Pertusaria is not monophyletic. Lichenologist 33, 161170.CrossRefGoogle Scholar
Mark, K, Cornejo, C, Keller, C, Flück, D and Scheidegger, C (2016) Barcoding lichen-forming fungi using 454 pyrosequencing is challenged by artifactual and biological sequence variation. Genome 59, 685704.CrossRefGoogle ScholarPubMed
Massalongo, AB (1860) Esame comparativo di alcuni generi di licheni. Atti dell’ I. R. Istituto Veneto di Scienze, Lettere ed Arti, series 3 5, 247276.Google Scholar
Miller, MA, 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 November 2010, New Orleans, Louisiana, pp. 18.CrossRefGoogle Scholar
Molina, MC, Crespo, A, Blanco, O, Lumbsch, HT and Hawksworth, DL (2004) Phylogenetic relationships and species concepts in Parmelia s. str. (Parmeliaceae) inferred from nuclear ITS rDNA and β-tubulin sequences. Lichenologist 36, 3754.Google Scholar
Molina, MC, Divakar, PK, Millanes, AM, Sanchez, E, Del-Prado, R, Hawksworth, DL and Crespo, A (2011 a) Parmelia sulcata (Ascomycota: Parmeliaceae), a sympatric monophyletic species complex. Lichenologist 43, 585601.CrossRefGoogle Scholar
Molina, MC, Del-Prado, R, Divakar, PK, Sánchez-Mata, D and Crespo, A (2011 b) Another example of cryptic diversity in lichen-forming fungi: the new species Parmelia mayi (Ascomycota: Parmeliaceae). Organisms, Diversity and Evolution 11, 331342.CrossRefGoogle Scholar
Molina, MC, Divakar, PK, Goward, T, Millanes, AM, Lumbsch, HT and Crespo, A (2017) Neogene diversification in the temperate lichen-forming fungal genus Parmelia (Parmeliaceae, Ascomycota). Systematics and Biodiversity 15, 166181.CrossRefGoogle Scholar
Monaghan, MT, Wild, R, Elliot, M, Fujisawa, T, Balke, M, Inward, DJG, Lees, DC, Ranaivosolo, R, Eggleton, P, Barraclough, TG, et al. (2009) Accelerated species inventory on Madagascar using coalescent-based models of species delineation. Systematic Biology 58, 298311.CrossRefGoogle ScholarPubMed
Myllys, L, Lohtander, K and Tehler, A (2001) β-tubulin, ITS and group I intron sequences challenge the species pair concept in Physcia aipolia and P. caesia. Mycologia 93, 335343.CrossRefGoogle Scholar
Nash, TH III and Elix, JA (2002) Parmotrema. In Nash, TH III, Ryan, BD, Gries, C and Bungartz, F (eds), Lichen Flora of the Greater Sonoran Desert Region, Vol. I. Tempe, Arizona: Lichens Unlimited, Arizona State University, pp. 318329.Google Scholar
Nelsen, MP and Gargas, A (2008) Dissociation and horizontal transmission of codispersing lichen symbionts in the genus Lepraria (Lecanorales: Stereocaulaceae). New Phytologist 177, 264275.CrossRefGoogle Scholar
Nimis, PL, Scheidegger, C and Wolseley, P (2002) Monitoring with Lichens – Monitoring Lichens. Dordrecht: Kluwer Academics.CrossRefGoogle Scholar
Normark, BB, Judson, OP and Moran, NA (2003) Genomic signatures of ancient asexual lineages. Biological Journal of the Linnean Society 79, 6984.CrossRefGoogle Scholar
Orange, A, James, PW and White, FJ (2010) Microchemical Methods for the Identification of Lichens. 2nd Edn. London: British Lichen Society.Google Scholar
Ott, S, Brinkmann, M, Wirtz, N and Lumbsch, HT (2004) Mitochondrial and nuclear ribosomal DNA data do not support the separation of the Antarctic lichens Umbilicaria kappenii and Umbilicaria antarctica as distinct species. Lichenologist 36, 227234.CrossRefGoogle Scholar
Parnmen, S, Rangsiruji, A, Mongkolsuk, P, Boonpragob, K, Nutakki, A and Lumbsch, HT (2012) Using phylogenetic and coalescent methods to understand the species diversity in the Cladia aggregata complex (Ascomycota, Lecanorales). PLoS ONE 7, e52245.CrossRefGoogle Scholar
Piercey-Normore, MD (2006) The lichen-forming ascomycete Evernia mesomorpha associated with multiple genotypes of Trebouxia jamesii. New Phytologist 169, 331344.CrossRefGoogle Scholar
Pons, J, Barraclough, TG, Gomez-Zurita, J, Cardoso, A, Duran, DP, Hazell, S, Kamoun, S, Sumlin, WD and Vogler, AP (2006) Sequence-based species delimitation for the DNA taxonomy of undescribed insects. Systematic Biology 55, 595609.CrossRefGoogle ScholarPubMed
Posada, D (2008) jModelTest: phylogenetic model averaging. Molecular Biology and Evolution 25, 12531256.CrossRefGoogle ScholarPubMed
Puillandre, N, Lambert, A, Brouillet, S and Achaz, G (2012) ABGD, Automatic Barcode Gap Discovery for primary species delimitation. Molecular Ecology 21, 18641877.CrossRefGoogle ScholarPubMed
Purvis, OW, Coppins, BJ, Hawksworth, DL, James, PW and Moore, DM (1992) The Lichen Flora of Great Britain and Ireland. London: British Lichen Society.Google Scholar
Rambaut, A (2009) FigTree v. 1.2.3. [WWW resource] URL http://tree.bio.ed.ac.uk/software/figtree/ [Accessed 5 November 2010].Google Scholar
Rambaut, A and Drummond, J (2007) Tracer v. 1.5. [WWW resource] URL http://beast.bio.ed.ac.uk/Tracer.Google Scholar
Ronquist, F and Huelsenbeck, JP (2003) MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19, 15721574.CrossRefGoogle ScholarPubMed
Sancho, LG, Pintado, A and Green, TGA (2019) Antarctic studies show lichens to be excellent biomonitors of climate change. Diversity 11, 42.CrossRefGoogle Scholar
Schmidt, HA, Strimmer, K, Vingron, M and von Haeseler, A (2002) TREE-PUZZLE: maximum likelihood phylogenetic analysis using quartets and parallel computing. Bioinformatics 18, 502504.CrossRefGoogle ScholarPubMed
Spielmann, AA and Marcelli, MP (2009) Parmotrema s.l. (Parmeliaceae, lichenized Ascomycota) from Serra Geral slopes in central Rio Grande do Sul State, Brazil. Hoehnea 36, 551595.CrossRefGoogle Scholar
Spielmann, AA and Marcelli, MP (2020) Type studies on Parmotrema (Parmeliaceae, Ascomycota) with salazinic acid. Plant and Fungal Systematics 65, 403508.CrossRefGoogle Scholar
Stamatakis, A (2006) RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22, 26882690.CrossRefGoogle ScholarPubMed
Stamatakis, A, Hoover, P and Rougemont, J (2008) A rapid bootstrap algorithm for the RAxML webservers. Systematic Biology 75, 758771.CrossRefGoogle Scholar
Sujetovienė, G (2015) Monitoring lichen as indicators of atmospheric quality. In Upreti, DK, Divakar, PK, Shukla, V and Bajpai, R (eds), Recent Advances in Lichenology, Vol. 1. New Delhi: Springer, pp. 87118.Google Scholar
Swinscow, TDV and Krog, H (1988) Macrolichens of East Africa. London: British Museum of Natural History.Google Scholar
Thell, A, Crespo, A, Divakar, PK, Kearnefelt, I, Leavitt, SD, Lumbsch, HT and Seaward, MRD (2012) A review of the lichen family Parmeliaceae – history, phylogeny and current taxonomy. Nordic Journal of Botany 30, 641664.CrossRefGoogle Scholar
Vilgalys, R and Hester, M (1990) Rapid genetic identification and mapping of enzymatically amplified DNA from several Cryptococcus species. Journal of Bacteriology 172, 42384246.CrossRefGoogle ScholarPubMed
Wang-Yang, JR and Lai, MJ (1976) Additions and corrections to the lichen flora of Taiwan. Taiwania 21, 226228.Google Scholar
Wei, JC (1991) An Enumeration of Lichens in China. Beijing: International Academic Publishers.Google Scholar
Widhelm, TJ, Egan, RS, Bertoletti, FR, Asztalos, MJ, Kraichak, E, Leavitt, SD and Lumbsch, HT (2016) Picking holes in traditional species delimitations: an integrative taxonomic reassessment of the Parmotrema perforatum group (Parmeliaceae, Ascomycota). Botanical Journal of the Linnean Society 182, 868884.CrossRefGoogle Scholar
Wirtz, N, Printzen, C and Lumbsch, HT (2008) The delimitation of Antarctic and bipolar species of neuropogonoid Usnea (Ascomycota, Lecanorales): a cohesion approach of species recognition for the Usnea perpusilla complex. Mycological Research 112, 472484.CrossRefGoogle ScholarPubMed
Yoshimura, I (1974) Lichen Flora of Japan in Colour. Osaka: Hoikusha.Google Scholar
Zhang, J, Kapli, P, Pavlidis, P and Stamatakis, A (2013) A general species delimitation method with applications to phylogenetic placements. Bioinformatics 29, 28692876.CrossRefGoogle ScholarPubMed
Figure 0

Table 1. Parmotrema species used in this study with GenBank Accession numbers of the ITS sequences and voucher information for the specimens. Crespoa carneopruinata was included in the phylogenetic analyses as outgroup. * = newly generated sequences from specimens for which morphological and chemical characters are given in Table S1 (available online).

Figure 1

Fig. 1. Main morphological characters of P. perlatum (A, C, E) and P. crinitum. (B, D, F), as indicated with arrows. A & B, black, simple cilia. C, marginal soralia concolorous with the thallus. D, isidia with apical cilia. E & F, black lower surface with brown naked zone peripherally. In colour online.

Figure 2

Fig. 2. Thin-layer chromatography profile of Parmotrema species (included in the DNA analysis) in solvent system C. P is the control, Pleurosticta acetabulum. Lanes 1 & 7, P. dilatatum. Lane 2, P. flavotinctum. Lane 3, P. mellissii. Lane 4, P. paulense. Lanes 5, 6, 8, 11 & 19, P. crinitum. Lane 9, empty. Lanes 10 & 12, Parmotrema perlatum. Lanes 13, 14, 15, 16, 17, 18 & 20, P. perlatum; these have the same TLC result as P. crinitum but were distinguished from each other based on the morphology. a = norstictic acid; b = atranorin; c = protocetraric acid; d = unknown; e = hypostictic acid; f = stictic acid; g = menegazzic acid. In colour online.

Figure 3

Fig. 3. Phylogenetic relationships between Parmotrema crinitum and P. perlatum. The green (thickened) line indicates the clade supported by maximum likelihood (ML) and Bayesian analyses; * = clades supported only by ML. ** = clades supported only by Bayesian analysis. Alphanumeric labels indicate clades and clusters. Species delimitation scenarios obtained from PTP, ABGD and genetic distances are indicated in columns to the right. In colour online.

Supplementary material: File

Stelate et al. supplementary material

Table S1

Download Stelate et al. supplementary material(File)
File 40.4 KB