Gintarasia and Xalocoa, two new genera to accommodate temperate to subtropical species in the predominantly tropical Graphidaceae (Ostropales, Ascomycota)

August 8, 2017 | Autor: Ekaphan Kraichak | Categoría: Phylogeny, Lichens, Lichen Taxonomy
Share Embed


Descripción

CSIRO PUBLISHING

Australian Systematic Botany, 2013, 26, 466–474 http://dx.doi.org/10.1071/SB13038

Gintarasia and Xalocoa, two new genera to accommodate temperate to subtropical species in the predominantly tropical Graphidaceae (Ostropales, Ascomycota) Ekaphan Kraichak A,C, Sittiporn Parnmen A,B, Robert Lücking A and H. Thorsten Lumbsch A A

Science & Education, The Field Museum, 1400 South Lake Shore Drive, Chicago, IL 60605, USA. Department of Medical Sciences, Ministry of Public Health, Tivanon Road, Nonthaburi 11000, Thailand. C Corresponding author. Email: ekraichak@fieldmuseum.org B

Abstract. The phylogenetic placement of Chapsa lamellifera, C. megalophthalma and Diploschistes ocellatus was studied using a dataset of five genetic markers (mtSSU, nuLSU, RPB1, RPB2 and ITS). As extratropical species occurring in Australasia, C. lamellifera and C. megalophthalma differ from other species in that genus by having relatively large ascomata with muriform ascospores and complex chemistry of either the protocetraric or stictic acids chemosyndrome. D. ocellatus is unique within Diploschistes, in lacking lateral paraphyses and containing the norstictic acid chemosyndrome. Previous phylogenetic analysis gave inconclusive results regarding the phylogenetic position of these taxa, and hence in the present study, a larger sampling of molecular markers was employed. Our results demonstrated that the two Chapsa species and D. ocellatus are not part of their current genera. Consequently, the new genera Gintarasia Kraichak, Lücking & Lumbsch and Xalocoa Kraichak, Lücking & Lumbsch are described to accommodate these species. The new combinations Gintarasia lamellifera (Kantvilas & Vezda) Kraichak, Lücking & Lumbsch, G. lordhowensis (Mangold) Kraichak, Lücking & Lumbsch, G. megalophthalma (Müll. Arg.) Kraichak, Lücking & Lumbsch and Xalocoa ocellata (Vill.) Kraichak, Lücking & Lumbsch are also proposed. Received 31 August 2013, accepted 3 February 2014, published online 27 March 2014

Introduction Graphidaceae is one of the largest families of lichen-forming fungi, with over 1800 accepted species (Rivas Plata et al. 2013). It includes predominantly lichen-forming, crustose species mostly with a trentepohlioid photobiont, immersed to sessile, rounded to elongate–lirellate ascomata, and mostly distoseptate ascospores (Rivas Plata et al. 2012a). The family has recently been expanded to include the previously separate families Asterothyriaceae, Gomphillaceae, Solorinellaceae and Thelotremataceae (Mangold et al. 2008b; Baloch et al. 2010; Rivas Plata and Lumbsch 2011; Rivas Plata et al. 2012a, 2012b). Molecular phylogenies have dramatically changed our understanding of the evolution and classification of Graphidaceae, leading to a new classification of the family, with four subfamilies several tribes and over 50 currently accepted genera (Rivas Plata et al. 2012a; Lücking et al. 2013). Whereas traditional generic classifications relied on only a few characters, such as ascospores and excipula, recent circumscriptions used combinations of morphology, anatomy and chemistry of thallus and ascoma. Ascospore septation and pigmentation or excipular carbonisation are also employed in generic circumscription, but recently considered to be of minor taxonomic importance (Staiger 2002, pp. 34–36; Frisch et al. 2006, pp. 48–59; Mangold et al. 2009, pp. 195–196; Rivas Plata Journal compilation  CSIRO 2014

and Lumbsch 2011; Rivas Plata et al. 2012a; Parnmen et al. 2013). The large majority of species in the family are tropical and it represents an important component of corticolous rainforest microlichen biotas (Santesson 1952, pp. 31–37; Sipman and Harris 1989; Rivas Plata et al. 2008). However, extratropical species also exist in various clades throughout the family. Within the thelotremoid Graphidaceae with roundish ascomata, these include the genus Diploschistes Norman, with a centre of diversity in areas with Mediterranean climate (Lumbsch 1989; Guderley and Lumbsch 1996; Lumbsch and Elix 2003) and the small genera Melanotopelia Lumbsch & Mangold and Topeliopsis Kantvilas & Vezda, which mostly occur in subantarctic to temperate regions, but also extend into tropical areas at higher altitudes (Mangold et al. 2008a, 2009, pp. 195–196; Lumbsch et al. 2010). Also some larger tropical genera, such as Chapsa A.Massal. and Thelotrema Ach., extend with a few species into temperate regions (Purvis et al. 1995; Kantvilas and Vezda 2000; Lumbsch et al. 2010). Within Chapsa, a few species with fairly large chroodiscoid ascomata occur in temperate regions of Australasia and have been placed either in Chroodiscus Mull.Arg. or Chapsa (Kantvilas and Vezda 2000; Mangold et al. 2009, pp. 197–227; Lumbsch et al. 2010). However, the generic placement of these species www.publish.csiro.au/journals/asb

Gintarasia and Xalocoa – new lichen genera

Australian Systematic Botany

remains problematic because of several key differences in morphological and chemical characters. These extratropical species include Chapsa lamellifera Kantvilas & Vezda, C. lordhowensis Mangold and C. megalophthalma Mull.Arg., which are unique within Chapsa by having a combination of large ascomata with muriform ascospores and a complex chemistry, containing the protocetraric acid chemosyndrome (including fumarprotocetraric, protocetraric and succinprotocetraric acids), or stictic acid and biogenetically related substances (constictic, a-acetylhypoconstictic, cryptostictic, a-acetylconstictic, hyposalazinic and hypostictic acids) (Kantvilas and Vezda 2000; Mangold et al. 2009, pp. 206–207, 212–214; Table 1). Although the presence of lateral paraphyses readily distinguishes these species from a restricted circumscription of Chroodiscus, which includes strictly foliicolous lichens with chroodiscoid ascomata (Frisch

467

et al. 2006, pp. 126–134; Lücking et al. 2008; Papong et al. 2009), their relationships to the four currently accepted genera in the Chapsa clade (Parnmen et al. 2012) remain uncertain. A recent large-scale phylogeny of the family, which included 905 ingroup OTUs with three loci, suggested that Chapsa megalophthalma is most likely placed outside the Chapsa s.lat. clade (Rivas Plata et al. 2013; Fig. 1). The monophyly of the genus Diploschistes has also been discussed. While the majority of species in that genus have perithecioid to urceolate ascomata with well developed, dark brown pigmented lateral paraphyses, two species, D. bisporus (Bagl.) J.Steiner and D. ocellatus (Vill.) Norman, lack lateral paraphyses (Lumbsch 1989). The former species was separated at the generic level as Ingvariella bispora (Bagl.) Guderley & Lumbsch (Guderley et al. 1997), which has recently been shown to be unrelated to other Diploschistes and even to belong to an

Table 1. Comparison of morphological and chemical characters of the genera Chapsa s.str., Diploschistes s.str., Gintarasia and Xalocoa Genus

Ascomata

Chapsa s.str. Small- to mediumsized (0.5–1.7 mm in diameter), chroodiscoid Gintarasia Large (up to 4 mm in diameter), chroodiscoid

Hypothecium

Ascospores

Hyaline

Yellowish–brown to yellow

Diploschistes Small- to mediumHyaline s.str. sized (0.5–1.5 mm in diameter), perithecioid or urceolate Xalocoa Large (up to 4 mm Hyaline in diameter), lecanoroid

A. 72 OTUs, 2 loci

Lateral paraphyses

Chemistry

Distribution

Hyaline, transversely Present septate to muriform

Nil or stictic acid

Pantropical

Hyaline, muriform

Present

Australasian, extratropical

Brown, muriform,

Present

Succinprotocetraric acid chemosyndrome, and stictic acid chemosyndrome Diploschistesic, gyrophoric, and lecanoric acids

Brown, muriform

Absent

Norstictic acid

Mediterranean climate, subcosmopolitan

Cosmopolitan

B. 920 OTUs, 3 loci

Fig. 1. Schematic diagram illustrating two alternative hypotheses of the topology and placement of Diplochistes ocellatus and Chapsa megalophthalma within the lichenised fungi family Graphidaceae, after A. Fernández-Brime et al. (2013) and B. Rivas Plata et al. (2013).

468

Australian Systematic Botany

entirely different family (Fernández-Brime et al. 2011). The placement of D. ocellatus, however, remains controversial. Whereas some phylogenetic studies revealed a monophyletic Diploschistes, including D. ocellatus (Lumbsch and Tehler 1998; Martín et al. 2003; Fernández-Brime et al. 2013), others found the genus in its current circumscription to be nonmonophyletic (Parnmen et al. 2013; Rivas Plata et al. 2013). Similar to the case of C. megalophthalma, D. ocellatus appeared to be more closely related to other genera, such as Acanthotrema Frisch and Acanthothecis Clem., than its congeneric taxa in a large-scale phylogeny (Rivas Plata et al. 2013; Fig. 1). Such discrepancies in phylogenetic placements deserve a more thorough investigation with increased numbers of molecular markers. In the present study, we focused on two groups of temperate Graphidaceae to address their generic placement using DNAsequence data from three ribosomal and two protein-coding markers. Specifically, we addressed the following questions: first, whether Diploschistes as currently circumscribed is monophyletic; second, whether the chroodiscoid temperate Chapsa species with complex chemistries are congeneric with Chapsa sens.str.? Materials and methods Molecular methods Data matrices of 142 sequences of 28 taxa were assembled using sequences of mitochondrial small subunit rRNA (mtSSU), nuclear large subunit rRNA (nuLSU), the RNA polymerase II largest subunit (RPB1), the RNA polymerase II second largest subunit (RPB2), and the internal transcribed spacer (nuITS) sequences (Table 1). Our choice of taxa was based on the clades previously recovered in the large-scale phylogeny of the family (Rivas Plata et al. 2013; Fig. 1B). We included seven species of Chapsa sens. str. and rooted the resultant trees between this group and the rest of the sampled taxa, on the basis of previous studies (Mangold et al. 2008a; Rivas Plata et al. 2013). We did not include all species of Diploschistes sens.str. from the previous studies, because of their incomplete molecular data in the context of our current sampling scheme. All of the specimen and DNA extraction vouchers are housed in F. Total DNA was extracted from freshly collected material or herbarium specimens by using the DNeasy Plant Mini Kit (Qiagen, Valencia, CA, USA), E.Z.N.A. Fungal MiniPrep Kit (Omega-Biotech, Doraville, GA, USA), or REDExtract-N-Amp PCR Plant Kit (Sigma–Aldrich, St Louis, MO, USA), following the instructions of the manufacturer. Dilutions (from 10 1 up to 10–2) of DNA were used for PCR amplifications. Primers for amplification included (1) for mtSSU: mrSSU1 and mrSSU3R (Zoller et al. 1999), (2) for nuLSU: AL2R (Mangold et al. 2008b) and nu-LSU-1125-39 (= LR6) (Vilgalys and Hester 1990), (3) for RPB1: RPB1-AF and RPB1-CR (Matheny et al. 2002), (4) for RPB2: fRPB2-7cF and fRPB2-11aR (Liu et al. 1999) and (5) for nuITS: ITS1 and ITS4 (White et al. 1990). The 10-mL PCR reactions contained 2.5 mL of REDExtract-N-Amp PCR ReadyMix (Sigma–Aldrich), 0.5 mL of each primer (10 mM), 2-mL genomic DNA extract and 4.5-mL nuclease-free water. Amplification products were visualised on 1% agarose gels stained with ethidium bromide, and subsequently purified by

E. Kraichak et al.

using GELase Agarose Gel-Digesting Preparation (Epicentre, Madison, WI, USA). Fragments were sequenced using the Big Dye Terminator reaction kit (ABI PRISM, Applied Biosystems, Carlsbad, CA) and the same sets of primers from the amplification. Cycle sequencing was executed with the following program: 25 cycles of 95C for 30 s, 48C for 15 s and 60C for 4 min. Sequenced products were precipitated with 10 mL of sterile dH2O, 2 mL of 3 M Napa, and 50 mL of 95% EtOH before they were loaded on an ABI 3100 (Applied Biosystems) automatic sequencer. Sequence fragments obtained were assembled with Geneious 5.5.8 (Drummond et al. 2011) and manually adjusted. Sequence alignments and phylogenetic analysis Alignments were initially performed using Geneious (Drummond et al. 2011). Ambiguously aligned regions were then excluded by using Gblocks (Castresana 2000; Talavera and Castresana 2007), using options for a ‘less stringent’ selection on the Gblocks web server (http://molevol.cmima.csic.es/ castresana/Gblocks_server.html, accessed 10 September 2013). The alignments were then analysed using maximum-likelihood (ML) and Bayesian approaches (B/MCMC). The ML analysis was performed on the concatenated alignment with the program RAxML-HPC2 (version 7.3.1) on XSEDE (Stamatakis 2006), using the default rapid hillclimbing algorithm and the GTRGAMMA model of nucleotide substitution. The partitions were set for each of the loci. Rapid bootstrap estimates were carried out for 1000 pseudoreplicates (Stamatakis et al. 2008). For the B/MCMC analysis, the dataset was also partitioned into five parts (one for each locus) and then analysed using MrBAYES 3.1.2 (Huelsenbeck and Ronquist 2001). Model testing was performed for each locus, using the program jmodelltest2 (Guindon and Gascuel 2003; Darriba et al. 2012). On the basis of the likelihood values, GTR+I+G was chosen as the optimal substitution model for each of the loci. A run with 10 000 000 generations, starting with a random tree and employing four simultaneous chains, was executed. Heating of chains was set to 0.2. Posterior probabilities were approximated by sampling trees, using a variant of Markov Chain Monte Carlo (MCMC) method. Every 1000th tree was sampled to avoid sample autocorrelation. The first 4000 trees were discarded as burn-in. Of the remaining trees, a majority-rule consensus tree with average branch lengths was calculated using the sumt option of MrBAYES. Posterior probabilities were obtained for each clade. Only clades with bootstrap support equal or above 75% under ML and posterior probabilities equal or above 0.95 in a Bayesian framework were considered as supported. Phylogenetic trees were drawn using the program FigTree 1.4.0 (Rambaut 2012). The alignment and tree were submitted to the public database, TreeBase (http://purl.org/phylo/treebase/phylows/ study/TB2:S14970, accessed 14 November 2013). To test for alternative hypotheses of monophyly of the genera Diploschistes and Chapsa we respectively performed the (1) Shimodaira–Hasegawa (SH) (Shimodaira and Hasegawa 2001) and (2) expected likelihood weight (ELW) tests (Strimmer and Rambaut 2002), using the program Tree-PUZZLE v.5.2 (Schmidt et al. 2002). In particular, we chose SH as a test for alternative

Gintarasia and Xalocoa – new lichen genera

Australian Systematic Botany

hypothesis, because it appears to be more conservative than any other tests available (Buckley 2002). The program used the concatenated dataset to compare the best tree from the alternative hypotheses with the unconstrained ML tree. These trees were inferred in Tree-PUZZLE, using the GTR+I+G nucleotide substitution model. Results and discussion Seventy-two new sequences were generated in the present study and aligned with sequences obtained from GenBank (Table 2). A matrix of 2667 unambiguously aligned nucleotide-position characters was produced, and 1549 characters in the alignment were constant (Table 3). ML analysis yielded a tree with the final optimisation likelihood of –15 504.94. In the B/MCMC analysis, the likelihood parameters varied among different partitions (Table 4). Because the topologies of the ML and B/MCMC analyses did not show any strongly supported conflicts, only the topology of B/MCMC 50% majority-rule consensus tree is shown here. The nodes with strong support values in both the ML and Bayesian analyses (i.e. PP 0.95 in B/MCMC analysis and MP bootstrap of 75%) are indicated in Fig. 2. Diploschistes sens.str. formed a strongly supported monophyletic group (ML 100, B/MCMC 1.00), whereas

469

D. ocellatus was only distantly related to Diploschistes sens. str. and appeared to be a sister group to a strongly supported monophyletic Acanthotrema (ML 100, B/MCMC 1.00). However, this relationship lacked support from either analysis. Monophyly of Diploschistes as currently circumscribed was rejected in both alternative hypothesis tests (P < 0.001). D. ocellatus differs from other species in Diploschistes sens. str. in several morphological and chemical characters, and hence, it is not surprising that our data revealed D. ocellatus as a distinct lineage (Table 1). Whereas taxa in Diploschistes sens.str. have perithecioid to urcelate ascomata and a dark-pigmented, paraplectenchymatous exciple with lateral paraphyses, D. ocellatus has lecanoroid ascomata, a reduced exciple and lacks lateral paraphyses (Fig. 3A). Chemically, it differs in containing the norstictic acid chemosyndrome, whereas species in Diploschistes sens. str. contain orcinol depsides, such as diploschistesic, gyrophoric or lecanoric acids (Lumbsch 1989; Lumbsch and Elix 1989). Because of these differences, the species was previously classified in subgen. Thorstenia within Diploschistes (Fernández-Brime et al. 2013), but is here included in a new genus. Chapsa lamellifera and C. megalophthalma formed a well supported monophyletic group (ML 94, B/MCMC 1.00) that falls outside the strongly supported clade of Chapsa sens.str. (ML 100,

Table 2. Specimens used in the study, with their collecting data and GenBank accession numbers Newly generated sequences are indicated in bold. –, missing data Genus Chapsa Chapsa Chapsa Chapsa Chapsa Chapsa Chapsa Acanthothecis Acanthothecis Acanthotrema Acanthotrema Chapsa Chapsa Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes Diploschistes New sequences Total sequences

Species alborosella indica leprocarpa patens pulchra sublilacina thallotrema hololeucoides peplophora alboisidiatum brasilianum lamellifera megalophthalma actinostomus cinereocaesius cinereocaesius diacapsis diploschistoides elixii euganeus euganeus hensseniae muscorum ocellatus scruposus scruposus sticticus sticticus

Location Brazil Thailand Thailand Thailand Australia Mexico Venezuela Brazil USA Puerto Rica Peru Australia Australia Kenya Peru Costa Rica Spain Australia USA Switzerland Switzerland Australia Switzerland Greece Germany Germany Australia Kenya

Voucher number Lücking 31238a Parnmen 18486 Lücking 24004 Lücking 24003 Lumbsch 19129t Lücking RLD056 Lücking 32019 Lücking 31303 Common 1926A Mercado 290 Rivas Plata 08–08-Da Lumbsch 20009b Mangold 11S Lumbsch 3089 Lumbsch 19303 Lücking 15540 Unknown Lumbsch 19073b Elix 32450 Lumbsch 20603 Lumbsch 20605g Elix 32450 Lumbsch 20605d Lumbsch, s.n. Schmitt, s.n. Schmitt, s.n. Lumbsch 19109b Lumbsch 3056

Accession number RPB1 RPB2

mtSSU

nuLSU

ITS

JX420972 JX465280 JX420995 JX421003 EU075571 HQ639600 JX421013 JX420952 JX420954 KF688506 JX420958 JX420990 JX421456 KF688510 KF688502 JX421026 KF688498 KF688500 KF688504 KF688507 KF688508 KF688503 KF688509 KF688505 – KF688501 KF688499 KF688511

JX421439 JX465295 JX421455 JX421459 EU075619 JX421466 JX467681 JX421423 JX421424 KF688492 JX421429 JX421449 JX421456 KF688496 KF688490 KF688491 JX421481 AY605076 EU126644 KF688493 KF688494 NEW KF688495 AY605077 KF688488 KF688489 JX421482 KF688497

KF688521 KF688525 KF688523 KF688522 KC020292 KF688519 – – KF688524 KF688526 – – – KF688530 KF688516 KF688518 – KF688513 KF688520 KF688527 KF688528 KF688517 KF688529 DQ366252 KF688514 KF688515 KF688512 KF688531

JX420936 JX465322 JX420942 JX420941 EU075619 JX420842 JX420905 JX420938 – KF688542 JX420876 – KF688538 KF688546 KF688537 KF688540 KF688532 KF688534 KF688541 KF688543 KF688544 KF688539 KF688545 DQ366253 KF688535 KF688536 KF688533 KF688547

– – – – – – – – KF688483 KF688484 – – KF688480 KF688486 KF688479 KF688481 – – KF688482 – KF688485 – – KF688476 KF688477 KF688478 – KF688487

14 27

10 28

20 22

16 26

12 12

Total 72 115

470

Australian Systematic Botany

E. Kraichak et al.

Table 3. Nucleotide data, alignment information, and optimal model choice from model testing for molecular markers used in the current analysis Molecular Number of Number of Number of Optimal marker original nucleotide positions constant substitution positions after exclusion position model mtSSU nuLSU RPB1 RPB2 nuITS Total

780 936 496 884 272

438 663 496 884 186

303 437 248 489 72

3368

2667

1549

GTR+I+G GTR+I+G GTR+I+G GTR+I+G GTR+I+G

B/MCMC 1.00). The B/MCMC analysis showed a strong support (PP = 1) for this C. lamellifera–megalophthalma clade being sister to Acanthothecis spp., whereas the ML analysis did not provide strong support for this relationship. Alternative hypothesis tests rejected the inclusion of C. lamellifera and C. megalophthalma in Chapsa sens.str. in both tests (P < 0.001). This marked molecular difference is consistent with morphological and chemical differences, observed in previous studies (Kantvilas and Vezda 2000; Mangold et al. 2009). These two species and the phenotypically similar C. lordhowensis (Fig. 3B–D) differ from the species of Chapsa sens.str. by their larger ascomata and complex chemistry, including numerous depsidones of the protocetraric or stictic acid chemosyndrome. Even though the DNA sequences of C. lordhowensis could not be obtained for the current analysis, its phenotypic characters align more closely with C. lamellifera and C. megalophthalma than the rest of Chapsa, and it is therefore reasonable to place C. lordhowensis in this clade. Our results thus support the elevation of this unique Chapsa s.lat. clade as a separate genus. With increased taxon sampling, in both numbers of species and molecular markers, the current study provided evidence that D. ocellatus, C. lamellifera and C. megalophthalma are not closely related to their congenerics in their current generic circumscriptions. One solution would be to include all of the taxa outside Chapsa sens.str. (Diploschistes spp., Acanthotrema spp., Acanthothecis, Chapsa megalophthalma, C. lamellifera) in one genus. However, given the morphological and chemical disparity among these taxa, we instead propose two new generic names, Gintarasia and Xalacoa, to accommodate these two distinct lineages that do not fit the genera in which they are currently placed.

Table 4.

Taxonomic consequences Gintarasia Kraichak, Lücking & Lumbsch, gen. nov. MycoBank No.: MB 805426 A new genus in the family Graphidaceae, subfamily Graphidoidae, tribe Thelotremateae, characterised by a greyish-green to olive-green thallus covered by a cortex or epinecral layer, large chroodiscoid ascomata with exposed discs and thick thalline margins, a fused, hyaline to yellowish proper exciple with lateral paraphyses, a non-inspersed hymenium, hyaline, non-amyloid or amyloid ascospores, and containing depsidones of the protocetraric or stictic acid chemosyndrome. Type species: Gintarasia lamellifera (Kantvilas & Vezda) Kraichak, Lücking & Lumbsch. Etymology: this new genus is named after our friend and colleague, the Tasmanian lichenologist Gintaras Kantvilas, who has made important contributions to lichenology in Australia, especially Tasmania, and has worked on the taxonomy of Tasmanian thelotremoid Graphidaceae. Gintarasia lamellifera (Kantvilas & V ezda) Kraichak, Lücking & Lumbsch comb. nov. Mycobank no.: MB 805428 Chroodiscus lamelliferus Kantvilas & Vezda, Lichenologist 32: 336 (2000); Chapsa lamellifera (Kantvilas & Vezda) Mangold, in P.M. McCarthy (Ed.), Fl. Australia. 57: 653 (2009).

Type: Australia, Tasmania, Ben Ridge, E of Ben Nevis, alt. 750 m, G. Kantvilas 105/81 (holo: HO!). Gintarasia lordhowensis (Mangold) Kraichak, Lücking & Lumbsch comb. nov. Mycobank no.: MB 805429 Chapsa lordhowensis Mangold, in P.M. McCarthy (ed.), Fl. Australia. 57: 653 (2009).

Type: Australia, Lord Howe Island, Goat House Cave, J.A. Elix 42259 (holo: CANB!). Gintarasia megalophthalma (Müll. Arg.) Kraichak, Lücking & Lumbsch comb. nov. Mycobank no.: MB 805430 Thelotrema megalophthalmum Müll. Arg., Flora 65: 500 (1882);. Chroodiscus megalophthalmus (Müll. Arg.) Vezda & Kantvilas, in A. Vezda, Lich. Rar. Crit. Exs. 3: [25] (1992);. Chapsa megalophthalma (Müll. Arg.) Mangold, P.M. McCarthy (Ed.), Fl. Australia. 57: 654 (2009).

Type: Australia, Queensland, C. H. Hartmann s.n. (iso: BM!).

Toowoomba,

1882,

Parameter values (mean, variance in parentheses) obtained from the Bayesian inferences of the five-loci partitioned dataset of Diploschistes s. str. and their relatives

Locus

a

p(A)

p(C)

p(G)

p(T)

pinvar

mtSSU nuLSU RPB1 RPB2 nuITS

33.091 (2874.689) 0.926 (0.138) 2.202 (1.587) 1.535 (0.292) 25.152 (2386.885)

0.343 (0) 0.257 (0) 0.306 (0) 0.254 (0) 0.270 (0.001)

0.148 (0) 0.236 (0) 0.228 (0) 0.267 (0) 0.275 (0.001)

0.202 (0) 0.315 (0) 0.260 (0) 0.262 (0) 0.211 (0.001)

0.307 (0) 0.193 (0) 0.206 (0) 0.217 (0) 0.244 (0.001)

0.557 (0.025) 0.396 (0.007) 0.371 (0.003) 0.449 (0.01) 0.205 (0.011)

Gintarasia and Xalocoa – new lichen genera

Australian Systematic Botany

Fig. 2. Bayesian 50% majority rule consensus tree, illustrating relationships among Disploschistes sens.str. and their related species with a partitioned dataset of five loci (mtSSU, nuLSU, RPB1, RPB2 and nuITS). Branches with posterior probabilities 0.95 and ML-bootstrap values 75 are indicated as thickened branches, as well as the numbers at the node (ML bootstrap/posterior probability).

471

472

Australian Systematic Botany

E. Kraichak et al.

A

B

C

D

Fig. 3. A. Xalocoa ocellata (Lumbsch 10734). B. Gintarasia lamellifera (Tibell 18878). C. G. lordhowensis (Elix 42267). D. G. megalophthalma (Elix 30220). Scale bars: 1 mm.

Xalocoa Kraichak, Lücking & Lumbsch, gen. nov. MycoBank No.: MB 805427 Diploschistes subg. Thorstenia Fern.-Brime, Gaya & Llimona Taxon 63 (2): 227 (2013).

A new genus in the family Graphidaceae, subfamily Graphidoidae, tribe Thelotremateae, characterised by a greyish-white thallus covered by an epinecral layer, large apothecioid ascomata with exposed discs and thick, entire thalline margins, a thin, reduced, uncarbonised proper exciple lacking lateral paraphyses, a non-inspersed hymenium, pale brown, non-amyloid ascospores, bacilliform conidia, and containing the norstictic acid chemosyndrome. Type species: Xalocoa ocellata (Vill.) Kraichak, Lücking & Lumbsch. Etymology: the name is derived from the Catalan name ‘xaloc’ for scirocco, the warm Mediterranean wind originating in the Sahara. This refers to the distribution of the taxon in areas with Mediterranean climate. We have chosen the Catalan version of the word to honour our Catalan colleague Xavier Llimona for his

contributions to our understanding of Mediterranean lichens and the taxonomy of the genus Diploschistes sens.lat. Xalocoa ocellata (Vill.) Kraichak, Lücking & Lumbsch comb. nov. Mycobank no.: MB 805431 Lichen ocellatus Vill., Hist. pl. Dauph.: 988 (1789);. Urceolaria ocellata (Vill.) DC., in J.B.A.P.M. de Lamarck & A.P. de Candolle, Fl. fran¸c ., 3rd edn., 2: 372 (1805);. Parmelia ocellata (Vill.) Fr., Lichenogr. eur. reform.: 190 (1831);. Placodium ocellatum (Vill.) Link, Grundr. Kräuterk. 3: 189 (1832); Lagerheimina ocellata (Vill.) Kuntze, Revis. Gen. Pl.: 478 (1891);. Diploschistes ocellatus (Vill.) Norman, Nyt Mag. Naturvidensk. 7: 232 (1853).

Type: France, Vaucluse, St Martin, 1. Dec. 1962, G. Clauzade (neo: BM!, selected by Lumbsch (1989)). Supplementary material Supplementary material to this paper is available from the journal online (see http://www.publish.csiro.au/?act=view_file&file_ id=SB13038_AC.pdf). Contained therein are single-gene trees

Gintarasia and Xalocoa – new lichen genera

from maximum likelihood analyses of individual loci (Figs S1–S5). Acknowledgements This study was supported financially by a NSF grant (DEB-1025861) to The Field Museum (PI: HTL, co-PI: RL). Newly obtained DNA sequences were generated in the Pritzker Laboratory for Molecular Systematics and Evolution at the Field Museum.

References Baloch E, Lücking R, Lumbsch HT, Wedin M (2010) Major clades and phylogenetic relationships between lichenized and non-lichenized lineages in Ostropales (Ascomycota: Lecanoromycetes). Taxon 59, 1483–1494. Buckley T (2002) Model misspecification and probabilistic tests of topology: evidence from empirical data sets. Systematic Biology 51, 509–523. doi:10.1080/10635150290069922 Castresana J (2000) Selection of conserved blocks from multiple alignments for their use in phylogenetic analysis. Molecular Biology and Evolution 17, 540–552. doi:10.1093/oxfordjournals.molbev.a026334 Darriba D, Taboada GL, Doallo R, Posada D (2012) jModelTest 2: more models, new heuristics and parallel computing. Nature Methods 9, 772. doi:10.1038/nmeth.2109 Drummond AJ, Ashton B, Buxton S, Cheung M, Cooper A, Duran C, Field M, Heled J, Kearse M, Markowitz S, Moir R, Stones-Havas S, Sturrock S, Thierer TAW (2011) Geneious. Version 5.5.8. (Biomatters: Auckland New Zealand). Available at http://www.geneious.com [Verified April 2012] Fernández-Brime S, Llimona X, Molnar K, Stenroos S, Hognabba F, Bjoerk C, Lutzoni F, Gaya E (2011) Expansion of the Stictidaceae by the addition of the saxicolous lichen-forming genus Ingvariella. Mycologia 103, 755–763. doi:10.3852/10-287 Fernández-Brime S, Llimona X, Lutzoni F, Gaya E (2013) Phylogenetic study of Diploschistes (lichen-forming Ascomycota: Ostropales: Graphidaceae), based on morphological, chemical, and molecular data. Taxon 62, 267–280. doi:10.12705/622.10 Frisch A, Kalb K, Grube M (2006) Contributions towards a new systematics of the lichen family Thelotremataceae. Bibliotheca Lichenologica 92, 1–539. Guderley R, Lumbsch HT (1996) The lichen genus Diploschistes in South Africa (Thelotremataceae). Mycotaxon 58, 269–292. Guderley R, Lumbsch HT, Feige GB (1997) Ingvariella, a new genus in the Thelotremataceae (lichenized Ascomycotina). Nova Hedwigia 64, 147–154. Guindon S, Gascuel O (2003) A simple, fast and accurate method to estimate large phylogenies by maximum-likelihood. Systematic Biology 52, 696–704. doi:10.1080/10635150390235520 Huelsenbeck JP, Ronquist F (2001) MRBAYES: Bayesian inference of phylogenetic trees. Bioinformatics 17, 754–755. doi:10.1093/bioinformatics/17.8.754 Kantvilas G, Vezda A (2000) Studies on the lichen family Thelotremataceae in Tasmania. The genus Chroodiscus and its relatives. Lichenologist 32, 325–357. doi:10.1006/lich.2000.0274 Liu YJ, Whelen S, Hall BD (1999) Phylogenetic relationships among ascomycetes: evidence from an RNA polymerase II subunit. Molecular Biology and Evolution 16, 1799–1808. doi:10.1093/oxfordjournals.molbev.a026092 Lücking R, Papong K, Thammathaworn A, Boonpragob K (2008) Historical biogeography and phenotype-phylogeny of Chroodiscus (lichenized Ascomycota: Ostropales: Graphidaceae). Journal of Biogeography 35, 2311–2327. doi:10.1111/j.1365-2699.2008.01972.x

Australian Systematic Botany

473

Lücking R, Tehler A, Bungartz F, Rivas Plata E, Lumbsch HT (2013) Journey from the West: did tropical Graphidaceae (lichenized Ascomycota: Ostropales) evolve from a saxicolous ancestor along the American Pacific coast? American Journal of Botany 100, 844–856. doi:10.3732/ajb.1200548 Lumbsch HT (1989) Die holarktischen Vertreter der Flechtengattung Diploschistes (Thelotremataceae). The Journal of the Hattori Botanical Laboratory 66, 133–196. Lumbsch HT, Elix JA (1989) Taxonomy of some Diploschistes spp. (lichenized Ascomycetes, Thelotremataceae) containing gyrophoric acid. Plant Systematics and Evolution 167, 195–199. doi:10.1007/BF00936406 Lumbsch HT, Elix JA (2003) The lichen genus Diploschistes (Thelotremataceae) in Australia. Bibliotheca Lichenologica 86, 119–128. Lumbsch HT, Tehler A (1998) A cladistic analysis of the genus Diploschistes (Ascomycotina, Thelotremataceae). The Bryologist 101, 398–403. Lumbsch HT, Divakar PK, Messuti MI, Mangold A, Lucking R (2010) A survey of thelotremoid lichens (Ascomycota: Ostropales) in subantarctic regions excluding Tasmania. Lichenologist 42, 203–224. doi:10.1017/S002428290999048X Mangold A, Martin MP, Kalb K, Lücking R, Lumbsch HT (2008a) Molecular data show that Topeliopsis (Ascomycota, Thelotremataceae) is polyphyletic. Lichenologist 40, 39–46. doi:10.1017/S0024282908007366 Mangold A, Martin MP, Lücking R, Lumbsch HT (2008b) Molecular phylogeny suggests synonymy of Thelotremataceae within Graphidaceae (Ascomycota: Ostropales). Taxon 57, 476–486. Mangold A, Elix JA, Lumbsch HT (2009) Thelotremataceae. In ‘Flora of Australia’, Vol. 57. (Ed. PM McCarthy) pp. 195–420. (CSIRO Publishing: Melbourne) Martín MP, LaGreca S, Lumbsch HT (2003) Molecular phylogeny of Diploschistes inferred from ITS sequence data. Lichenologist 35, 27–32. doi:10.1006/lich.2002.0427 Matheny PB, Liu YJ, Ammirati JF, Hall BD (2002) Using RPB1 sequences to improve phylogenetic inference among mushrooms (Inocybe, Agaricales). American Journal of Botany 89, 688–698. doi:10.3732/ajb.89.4.688 Papong K, Corush J, Mangold A, Lücking R, Lumbsch HT (2009) Phylogenetic position of the foliicolous genus Chroodiscus (Ostropales, Ascomycota) inferred from nuclear and mitochondrial ribosomal DNA sequences. Fungal Diversity 38, 147–153. Parnmen S, Lücking R, Lumbsch HT (2012) Phylogenetic classification at generic level in the absence of distinct phylogenetic patterns of phenotypical variation: a case study in Graphidaceae (Ascomycota). PLoS ONE 7, e51392. doi:10.1371/journal.pone.0051392 Parnmen S, Cáceres MES, Lücking R, Lumbsch HT (2013) Myriochapsa and Nitidochapsa, two new genera in Graphidaceae (Ascomycota: Ostropales) for chroodiscoid species in the Ocellularia clade. The Bryologist 116, 127–133. doi:10.1639/0007-2745-116.2.127 Purvis OW, Jørgensen PM, James PW (1995) The lichen genus Thelotrema Ach. in Europe. Bibliotheca Lichenologica 58, 335–360. Rambaut A (2012) FigTree. Version 1.4. Available at http://tree.bio.ed.ac.uk/ software/figtree/ [Verified 5 December 2012] Rivas Plata E, Lumbsch HT (2011) Parallel evolution and phenotypic divergence in lichenized fungi: a case study in the lichen-forming fungal family Graphidaceae (Ascomycota: Lecanoromycetes: Ostropales). Molecular Phylogenetics and Evolution 61, 45–63. doi:10.1016/j.ympev.2011.04.025 Rivas Plata E, Lücking R, Lumbsch HT (2008) When family matters: an analysis of Thelotremataceae (lichenized Ascomycota: Ostropales) as bioindicators of ecological continuity in tropical forests. Biodiversity and Conservation 17, 1319–1351. doi:10.1007/s10531-007-9289-9 Rivas Plata E, Lücking R, Lumbsch HT (2012a) A new classification for the lichen family Graphidaceae sens. lat. (Ascomycota: Lecanoromycetes:

474

Australian Systematic Botany

E. Kraichak et al.

Ostropales). Fungal Diversity 52, 107–121. doi:10.1007/s13225-011-0135-8 Rivas Plata E, Lücking R, Lumbsch HT (2012b) Molecular phylogeny and systematics of the Ocellularia clade (Ascomycota: Ostropales: Graphidaceae). Taxon 61, 1161–1179. Rivas Plata E, Parnmen S, Staiger B, Mangold A, Frisch A, Weerakoon G, Hernández JE, Cáceres MES, Kalb K, Sipman HJM, Common RS, Lücking R, Lumbsch HT (2013) A molecular phylogeny of Graphidaceae (Ascomycota: Lecanoromycetes: Ostropales) including 437 species. MycoKeys 6, 55–94. doi:10.3897/mycokeys.6.3482 Santesson R (1952) Foliicolous lichens I. A revision of the taxonomy of the obligately foliicolous, lichenized fungi. Symbolae Botanicae Upsalienses 12, 1–590. Schmidt HA, Strimmer K, Vingron M, von Haeseler A (2002) TREEPUZZLE: maximum likelihood phylogenetic analysis using quartets and parallel computing. Bioinformatics 18, 502–504. Shimodaira H, Hasegawa M (2001) CONSEL: for assessing the confidence of phylogenetic tree selection. Bioinformatics 17, 1246–1247. doi:10.1093/bioinformatics/17.12.1246 Sipman HJM, Harris RC (1989) Lichens. In ‘Tropical Rain Forest Ecosystems’. (Eds H Lieth, MJA Werger) pp. 303–309. (Elsevier Science Publishers: Amsterdam) Staiger B (2002) Die Flechtenfamilie Graphidaceae: Studien in Richtung einer natürlicheren Gliederung. Bibliotheca Lichenologica 85, 1–526. Stamatakis A (2006) RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22, 2688–2690. doi:10.1093/bioinformatics/btl446

Stamatakis A, Hoover P, Rougemont J (2008) A rapid bootstrap algorithm for the RAxML web servers. Systematic Biology 57, 758–771. doi:10.1080/10635150802429642 Strimmer K, Rambaut A (2002) Inferring confidence sets of possibly misspecified gene tree. Proceedings of the Royal Society of London – B. Biological Sciences 269, 137–142. doi:10.1098/rspb.2001.1862 Talavera G, Castresana J (2007) Improvement of phylogenies after removing divergent and ambiguously aligned blocks from protein sequence alignments. Systematic Biology 56, 564–577. doi:10.1080/10635150701472164 Vilgalys R, Hester M (1990) Rapid genetic identification and mapping of enzymatically amplified ribosomal DNA from several Cryptococcus species. Journal of Bacteriology 172, 4238–4246. White TJ, Bruns T, Lee S, Taylor JW (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In ‘PCR Protocols: A Guide to Methods and Applications’. (Eds MA Innis, DH Gelfand, JJ Sninsky, TJ White) pp. 315–322. (Academic Press: New York) Zoller S, Scheidegger C, Sperisen C (1999) PCR primers for the amplification of mitochondrial small subunit ribosomal DNA of lichen-forming ascomycetes. Lichenologist 31, 511–516.

www.publish.csiro.au/journals/asb

Lihat lebih banyak...

Comentarios

Copyright © 2017 DATOSPDF Inc.