Research Article
Print
Research Article
Integrative taxonomy confirms three species of Coniocarpon (Arthoniaceae) in Norway
expand article infoAndreas Frisch, Victoria Stornes Moen§, Martin Grube|, Mika Bendiksby
‡ Norwegian University of Science and Technology, Trondheim, Norway
§ Norwegian University of Science and Technology, Trondgheim, Norway
| Karl-Franzens-University, Graz, Austria
¶ University of Science and Technology, Trondheim, Norway
Open Access

Abstract

We have studied the highly oceanic genus Coniocarpon in Norway. Our aim has been to delimit species of Coniocarpon in Norway based on an integrative taxonomic approach. The material studied comprises 120 specimens of Coniocarpon, obtained through recent collecting efforts (2017 and 2018) or received from major fungaria in Denmark, Finland, Norway and Sweden, as well as from private collectors. We have assessed (1) species delimitations and relationships based on Bayesian and maximum likelihood phylogenetic analyses of three genetic markers (mtSSU, nucITS and RPB2), (2) morphology and anatomy using standard light microscopy, and (3) secondary lichen chemistry using high-performance thin-layer chromatography. The results show three genetically distinct lineages of Coniocarpon, representing C. cinnabarinum, C. fallax and C. cuspidans comb. nov. The latter was originally described as Arthonia cinnabarina f. cuspidans and is herein raised to species level. All three species are supported by morphological, anatomical and chemical data.

Keywords

Arthoniales, Bayesian inference, maximum likelihood, morphology, mtSSU, nucITS, phylogeny, RPB2

Introduction

Species recognition is crucial for improved natural resource management and biodiversity conservation (Lumbsch and Leavitt 2011). Species recognizing can be challenging in lichenized fungi due to unclear species boundaries and/or cryptic diversity (e.g., Leavitt et al. 2011; Molina et al. 2011; Carlsen et al. 2012; Lücking et al. 2014; Bendiksby et al. 2015). One such example is Coniocarpon DC.

Coniocarpon belongs in the family Arthoniaceae Rchb. of the order Arthoniales. The Arthoniales is one of the largest orders of predominantly lichenized crustose and fruticose taxa (Sundin and Tehler 1998), with around 1500 lichenized, lichenicolous and saprotrophic fungi (Frisch et al. 2014; Van den Broeck et al. 2018). About 800 species belong to the family Arthoniaceae (Frisch and Thor 2010). The largest genus of the family is Arthonia Ach., with around 500 species (Grube 1995). As currently circumscribed, Arthonia is a polyphyletic genus with high diversity in morphology, chemistry, ecology, distribution and habitat preferences (Frisch et al. 2015). In the process of splitting Arthonia into monophyletic taxa, several genera have been described, emended or resurrected recently, of which Coniocarpon DC. is one (Frisch et al. 2014). Others include Bryostigma Poelt and Döbbeler, Coniarthonia Grube, Felipes (Ach.) Frisch and G.Thor, Inoderma (Ach.) Gray, Melarthonis Frisch and G.Thor, Pachnolepia A.Massal. and Synarthonia Müll.Arg. (Grube 2001; Frisch et al. 2014, 2015; Van den Broeck et al. 2018).

Coniocarpon is a small genus of four accepted species that are mainly distributed in humid tropical to warm-temperate regions of the world (Aptroot and Sparrius 2003; Coppins and Aptroot 2009; Van den Broeck et al. 2018), reaching higher latitudes, for instance, in the boreo-nemoral rainforests and other highly oceanic habitats in Norway (Blom et al. 2015). The genus is characterized within Arthoniaceae by crystalline orange, red and purple quinoid pigments in the ascomata that are dissolving with purple solution in K, by hyaline, macrocephalic, transversely septate ascospores that turn brownish with granular ornamentation in the epispore at late maturity, and by rounded to lirellate ascomata (Frisch et al. 2014, 2018; Van den Broeck et al. 2018). The widespread species C. cinnabarinum DC. and C. fallax (Ach.) Grube show marked morphological differentiation at the world level indicative for polyphyletic taxa.

The brightly pigmented ascomata in combination with their high value for nature conservation are two reasons why C. cinnabarinum and C. fallax are frequently collected in Norway. Both species are listed as vulnerable (VU) on the Norwegian Red List of Species 2015 (as A. cinnabarina and A. elegans; Henriksen and Hilmo 2015). The boreo-nemoral rainforests, located between 58–62°N (DellaSala et al. 2011), are also categorized as VU in the Norwegian Red List of Nature Types 2018 (Blom 2018). Coniocarpon and other taxa restricted to these forests are of particular importance for biodiversity conservation due to their limited distribution, which makes them vulnerable to extinction at a regional level.

Based on preliminary morphological and molecular evidence, the presence of three distinct taxa of Coniocarpon in Norway is hypothesized. This study aims to test this hypothesis by delimiting the taxa of Coniocarpon in Norway using an integrative taxonomic approach, including molecular, morphological and chemical data. All available specimens of Coniocarpon housed in fungaria in Denmark, Norway and Sweden have been revised. Species descriptions and an identification key are provided for all taxa and their distributions in Scandinavia are mapped.

Materials and methods

Collection methods and handling of fresh specimens

New specimens of Coniocarpon for this study were collected in boreo-nemoral rainforests on the west coast of Norway from Vest-Agder to Møre og Romsdal in 2017 and 2018. Specimens were placed in paper bags, allowed to air-dry and later stored at -20 °C due to prior knowledge of fast DNA degradation (e.g., Frisch et al. 2014). After DNA extraction, specimens were incorporated for long-term storage and access in the fungarium in Trondheim (TRH). Additional specimens were made available from fungaria in Bergen (BG), London (BM), Copenhagen (C), Edinburgh (E), Helsinki (H), Oslo (O), Paris (PC), Prag (PRA), Stockholm (S), Trondheim (TRH) and Uppsala (UPS).

Taxon sampling

In total, 26 specimens were used for the molecular data production; 25 specimens from Norway and 1 specimen from Great Britain. Outgroup taxa and additional sequences of Coniocarpon were downloaded from GenBank, in total 32 sequences. Eighteen of these (nine mtSSU and nine RPB2) represent the nine outgroup taxa, whereas 14 (eight mtSSU and six RPB2) were from eight specimens of Coniocarpon originating from Great Britain, Japan, Norway, Rwanda and Uganda. Outgroup taxa were selected based on their phylogenetic position in Frisch et al. (2014) and Van den Broeck et al. (2018).

DNA extraction and sequencing

DNA was isolated from specimens up to one year old. Genomic DNA was extracted following one of three methods. (1) Five to eight ascomata were sampled in 2 ml microcentrifuge tubes with two 3 mm diam. tungsten carbide beads each and crushed into a fine powder using a Retsch TissueLyser II. Subsequently, genomic DNA was extracted using the E.Z.N.A. SP Plant DNA Kit (Omega BIO-TEK, USA) following the manufacturer’s instructions. (2) Three to five ascomata were sampled directly in 0.2 ml Eppendorf PCR Tubes with 30 µl Dilution Buffer (Phire Plant Direct PCR Kit, ThermoFisher Scientific, Lithuania) and crushed with tweezers. (3) Small cuttings (ca. 50–100 µm × 50–100 µm in size) of the hymenium were sampled in 0.2 ml Eppendorf PCR Tubes and directly used for PCR amplification. The Phire Plant Direct PCR Kit (ThermoFisher Scientific, Lithuania) was used for PCR amplification in all three methods. Each PCR reaction contained 10 µl 2× Phire Plant PCR Buffer, 0.4 µl Phire Hot Start II DNA Polymerase, 1 µl of each primer for all genetic markers except RPB2 (1.5 µl of each primer), 1 µl genomic DNA (1:1) or the lichen sample, and was filled with H2O to the final volume of 20 µl. If PCR amplification resulted in weak products, 2 µl genomic DNA (1:10) was added. PCR amplification was done for the mtSSU, nucITS and the protein-coding gene RPB2 with the following primers: mtSSU1 + mtSSU3R (Zoller et al. 1999), ITS-1F + ITS4 (Larena et al. 1999, White et al. 1990) and RPB2-7cF + RPB2-11aR (Liu et al. 1999), respectively. PCR cycling conditions for mtSSU and nucITS started with an initial denaturation at 98 °C for 5 min, followed by 40 cycles of 98 °C for 5 s, 59 °C for 5 s, and 72 °C for 30 s, followed by a final extension of 72 °C for 1 min. For RPB2, annealing temperature was set to 57 °C. The PCR products were visualized on a 1% agarose gel stained with SYBR Safe DNA gel stain (ThermoFisher Scientific, USA) under UV light. Clean PCR products (i.e., those lacking visible contamination) were purified by adding 5 µl ExoSAP-IT Express PCR Cleanup (1:3 concentration; ThermoFisher Scientific, United Kingdom) to the PCR reactions. PCR reactions resulting in more than a single product were purified using the E.Z.N.A. Gel Extraction Kit (Omega BIO-TEK, USA) following the manufacturer’s instructions, except that we performed an additional wash buffer step. The PCR products were sent to Eurofins Genomics (Germany) for Sanger sequencing using the same primers as for the PCR reactions.

Sequence alignment and phylogenetic analyses

The sequences were edited and aligned using BioEdit v.7.0.5.3 (Hall 1999). The identity of the sequences was verified using the nucleotide BLAST search in GenBank. For the examination of topological incongruence among gene trees, maximum likelihood (ML) bootstrapping analyses was carried out on each of the data sets using the RAxML-HPC Blackbox ver. 8.2.10 (Stamatakis 2014). The standard settings without estimation of the proportion of invariable sites (GTRGAMMA) were used. Topological incongruence was assumed if conflicting tree topologies were supported by ≥ 70% bootstrap support. Since topological incongruence could not be observed, maximum likelihood (ML) bootstrapping analysis was carried out on the concatenated three-locus dataset of 43 accessions for Coniocarpon using default settings and adding single genes as parameter partition. The RAxML analysis was stopped automatically after 402 bootstrap replicates using the MRE-based bootstopping criterion (Pattengale et al. 2010).

The best-fit evolutionary model for each partition using the Bayesian information criterion (BIC; Schwarz 1979) was estimated using PartitionFinder2 ver. 2.1.1. (Lanfear et al. 2016). The input data included both ingroup and outgroup taxa. The pre-set partitions were mtSSU, RPB2/1st codon position, RPB2/2nd codon position, RPB2/3rd codon position, ITS1, 5.8S and ITS2. The Bayesian analysis was performed using MrBayes v.3.1.2 (Huelsenbeck et al. 2001; Ronquist and Huelsenbeck 2003). The MCMC run was using four independent chains and 10 million generations, sampling trees every 1000th generation. After removal of a burnin of 25%, 7500 trees were summarized in a final Bayesian 50% majority-rule consensus tree. All phylogenetic analyses were run on the CIPRES Science Gateway (Miller et al. 2010). Phylogenetic trees were visualized using FigTree ver. 1.4.4 (Rambaut 2018). Informative characters were estimated for ingroup taxa per locus using Winclada ver. 1.61 (Nixon 1999–2002).

Morphological and chemical investigations

The morphology of 120 specimens of Coniocarpon (Norway 87, Sweden 8, Denmark 16, Great Britain 7, Austria 1, Turkey 1) was studied. The morphology of all specimens was examined using a Leica M80 stereomicroscope and a Zeiss Standard Binocular microscope. Macroscopic photographs were taken with a Leica MZ16A stereomicroscope fitted with a Leica DFC420 camera. Microscopic photographs were taken with a Leica CTR6000 microscope fitted with a Leica DFC365 camera. Sections of ascomata and lichen thalli were cut by hand and mounted in water or lactic acid cotton blue (LCB). Length and width were measured for single ascomata as well as for aggregations composed of several ascomata. For the epithecium, exciple, hymenium and hypothecium, measurements were performed in LCB. Measurements of asci and ascospores were performed in water using squashed preparations. Only fully developed ascospores and commonly asci containing mature ascospores (sometimes asci without mature ascospores) were measured. Ascospore measurements are presented as (min.–)mean ± SD(–max.). The amyloid reaction of the apothecia was tested using 0.2% (Iodine diluted) and 1% (I), and 1% (I) solution after pretreatment with 10% potassium hydroxide (KOH) in water (KI). The quinoid pigments and Ca-oxalate crystals were measured in water and their shape studied. Later, the crystals reaction with KOH was observed. The quinoid pigments were identified by HPTLC (Arup et al. 1993) in solvent C. Quionid pigments are named according to Frisch et al. (2018) except for the newly identified A4.

Distribution maps

Based on occurrence information of all revised specimens, the distribution of Coniocarpon species in Scandinavia was illustrated by adding a delimited text layer to a Wikimedia map from QuickMap services in QGIS ver. 3.6.2. (QGIS Development Team 2019). The younger specimens (i.e., collected from the mid-80s and onwards) were placed on the maps by their geographical coordinates, while older specimens (collected 1870–1983) lacking geographical coordinates were placed on the maps based on locality information. The Earth Point coordinate converter (http://www.earthpoint.us/Convert.aspx) was used to convert coordinates.

Results

Molecular data

A total of 75 new sequences were obtained from the 26 included specimens of Coniocarpon (mtSSU 26, nucITS 25, RPB2 24; Table 1). The lengths of the alignments (all accessions included) and the phylogenetically informative characters for the ingroup taxa were for mtSSU 889/61, ITS1 569/160, 5.8S 129/0, ITS2 239/84 and RPB2 867/127.

The selected substitution models for the five subsets in PartitionFinder2 were: 1) GTR+G for mtSSU, 2) K80+1 for RPB2/1st, 3) F81+I for 5.8S and RPB2/2nd, 4) HKY+G for RPB2/3rd and 5) HKY+I for ITS1 and ITS2. The three gene-trees were congruent and a three-locus, concatenated dataset of 43 accessions were analyzed phylogenetically.

Table 1.

Vouchers and their GenBank accession numbers. New sequences are indicated in bold. An en dash indicates missing data.

Specimens Voucher Country mtSSU nucITS RPB2
Arthonia didyma Ertz 7587 (BR) Belgium EU704047 EU704010
Arthonia granithophlia Frisch 10/Se74 (UPS) Sweden KJ850981 KJ851107
Arthonia physcidiicola Frisch 11/Ug318 (UPS) Uganda KF707646 KF707657
Coniocarpon cinnabarinum 1 Frisch (TRH-L-29009) Norway MN733983 MN734118 MN719396
Coniocarpon cinnabarinum 2 Frisch (TRH-L-29000) Norway MN733980 MN734115 MN719393
Coniocarpon cinnabarinum 3 Frisch (TRH-L-29002) Norway MN733982 MN734117 MN719395
Coniocarpon cinnabarinum 4 Johnsen 111003 (UPS) Norway KJ850976 KJ851103
Coniocarpon cinnabarinum 5 Frisch (TRH-L-29008) Norway MN733984 MN734119 MN719397
Coniocarpon cinnabarinum 6 Frisch (TRH-L-29001) Norway MN733981 MN734116 MN719394
Coniocarpon cinnabarinum 7 Frisch 11/Ug297 (UPS) Uganda KJ850977 KJ851104
Coniocarpon cinnabarinum 8 Frisch 11/Ug296 (UPS) Uganda KP870158 KP870170
Coniocarpon cinnabarinum 9 Ertz 8730 (BR) Rwanda EU704046 EU704009
Coniocarpon cinnabarinum 10 Frisch 13/Jp128 (TNS) Japan MG201840
Coniocarpon cinnabarinum 11 Frisch 13/Jp127 (TNS) Japan MG201841
Coniocarpon cuspidans 1 Frisch (TRH-L-29026) Norway MN733977 MN734113 MN719390
Coniocarpon cuspidans 2 Acton, Malíček, Palice 25146 (PRA) Great Britain MN733979 MN719392
Coniocarpon cuspidans 3 Frisch (TRH-L-29013) Norway MN733970 MN734106 MN719384
Coniocarpon cuspidans 4 Frisch (TRH-L-29025) Norway MN733976 MN734112 MN719389
Coniocarpon cuspidans 5 Frisch (TRH-L-29014) Norway MN733971 MN734107 MN719385
Coniocarpon cuspidans 6 Klepsland (TRH-L-29017) Norway MN733973 MN734109 MN719387
Coniocarpon cuspidans 7 Frisch (TRH-L-29022) Norway MN733974 MN734110 MN719388
Coniocarpon cuspidans 8 Frisch (TRH-L-29015) Norway MN733972 MN734108 MN719386
Coniocarpon cuspidans 9 Frisch (TRH-L-29023) Norway MN733975 MN734111
Coniocarpon cuspidans 10 Frisch (TRH-L-29024) Norway MN733978 MN734114 MN719391
Coniocarpon fallax 1 Wågström 111123 (UPS) Norway MG201842 MG201850
Coniocarpon fallax 2 Frisch (TRH-L-29030) Norway MN733967 MN734103 MN719382
Coniocarpon fallax 3 Gaarder, Larsen (TRH-L-16791) Norway MN733961 MN734097 MN719376
Coniocarpon fallax 4 Frisch (TRH-L-29037) Norway MN733968 MN734104 MN719383
Coniocarpon fallax 5 Gaarder (TRH-L-16790) Norway MN733959 MN734095 MN719374
Coniocarpon fallax 6 Gaarder (TRH-L-16792) Norway MN733960 MN734096 MN719375
Coniocarpon fallax 7 Gaarder (TRH-L-16789) Norway MN733963 MN734099 MN719378
Coniocarpon fallax 8 Gaarder (TRH-L-15366) Norway MN733962 MN734098 MN719377
Coniocarpon fallax 9 Frisch (TRH-L-29029) Norway MN733966 MN734102 MN719381
Coniocarpon fallax 10 L10175 Great Britain KJ850979 KJ851101
Coniocarpon fallax 11 Frisch (TRH-L-29028) Norway MN733965 MN734101 MN719380
Coniocarpon fallax 12 Frisch (TRH-L-29036) Norway MN733969 MN734105
Coniocarpon fallax 13 Frisch (TRH-L-29027) Norway MN733964 MN734100 MN719379
Reichlingia leopoldii Ertz 13293 (BR) Belgium JF830773 HQ454722
Reichlingia syncesioides Frisch 11/Ug14 (UPS) Uganda KF707651 KF707656
Reichlingia zwackhii Thor 26800 (UPS) Sweden KF707652 KF707662
Synarthonia aurantiacopruinsa Van den Broeck 5764 (BR) DR Congo MH251874 MH271697
Synarthonia inconspicua Van den Broeck 6325 (BR) Uganda MH251880 MH271701
Synarthonia muriformis Ertz 19344 (BR) Madagascar MH251877 MH271699

Phylogeny

The phylogeny based on Bayesian and maximum likelihood analyses present Coniocarpon as monophyletic using the selected taxon and outgroup sampling. Three discrete, well-supported lineages are recovered within Coniocarpon. These are separated from each other by branches clearly exceeding the observed infraspecific branch-lengths. The three lineages represent C. cinnabarinum, C. cuspidans (Nyl.) Moen, Frisch and Grube and C. fallax (Fig. 1). Coniocarpon cinnabarinum is the supported sister taxon to C. cuspidans, while C. fallax is sister to these two taxa.

Coniocarpon cinnabarinum from Rwanda and Uganda form a well-supported clade and are sisters to C. cinnabarinum in Norway, while C. cinnabarinum from Japan is genetically distinct and sister to the remaining taxa of Coniocarpon. The two sampled specimens from Great Britain are genetically close to C. cuspidans and C. fallax in Norway, respectively.

Figure 1. 

The RAxML phylogenetic hypothesis of a concatenated three-locus data set (mtSSU, nucITS and RPB2) of 3 Coniocarpon and 9 outgroup taxa (marked by a dashed rectangle). The RAxML bootstrap proportions (first) and Bayesian posterior probabilities (second) are indicated. Place of origin provided behind taxon names.

Morphology and chemical characters

Forty-four specimens were identified as C. cinnabarinum, as C. cuspidans and 42 as C. fallax. Ascospore size (Fig. 2), ascospore septation (Fig. 3), ascoma shape and the distribution of pruina (Fig. 4) were identified as useful characters for species distinction. Moreover, differences were observed in the quinoid pigment patterns revealed by HPTLC (Fig. 5) and in the amyloidity of the ascomatal gels. Four quinoid pigments were identified showing different colors on the HPTLC plates prior to sulphuric acid treatment and charring; reddish (A1, A4), purple (A2), yellow (A3) (Fig. 5A). The color of the spots under UV365 light are deep purple (A1, A2), buff (A3) and dark salmon (A4) (Fig. 5B). The chemical results for all species are summarized in the Taxonomy section at the end of the Discussion.

Figure 2. 

Ascospore size (length x width) of C. cinnabarinum (n = 132; 20 ascospores from Denmark, 85 ascospores from Norway; 27 ascospores from Sweden), C. cuspidans (n = 100, all ascospores from Norway) and Coniocarpon fallax (n = 191, all ascospores from Norway).

Figure 3. 

Ascospore morphology. A Coniocarpon cinnabarinum (TRH-L-29007) B Coniocarpon fallax (TRH-L-29008) C Coniocarpon cuspidans (TRH-L-29036). Scale bar: 10 µm.

Figure 4. 

Morphological diversity of Coniocarpon in Norway. A Coniocarpon cinnabarinum (Frisch TRH-L-29000) B Coniocarpon cinnabarinum (Frisch TRH-L-29005) C Coniocarpon cuspidans (TRH-L-29022) D Coniocarpon cuspidans (TRH-L-29023) E Coniocarpon fallax (TRH-L-16793) F Coniocarpon fallax (Frisch TRH-L-29028). Scale bars: 1 mm.

Figure 5. 

HPTLC plate in solvent C of quinoid pigment patterns of C. cinnabarinum (left), C. fallax (middle) and C. cuspidans (right) A before treatment with sulfuric acid and charring, and B under UV365 light before treatment with sulfuric acid and charring. Note the absence of A3 in C. cuspidans. The position of atranorin and norstictic acid is indicated. Several additional spots on the HPTLC plates under UV365 light have not been identified.

Distribution and ecology

Distribution maps based on all revised specimens of Coniocarpon from Scandinavia confirm C. cinnabarinum for Denmark, Norway and Sweden, C. cuspidans for Norway, and C. fallax for Norway and Sweden (Fig. 6). Coniocarpon cinnabarinum has been collected in Norway in the boreo-nemoral rainforests in Rogaland and Hordaland, while the species occurs in other humid forests in Sweden (Skåne and Gotland) and Denmark (Sjælland and Jylland). Specimens of C. cuspidans have been seen in Norway from the boreo-nemoral rainforests in Vest-Agder, Rogaland and Hordaland. Coniocarpon fallax has been collected in Norway in the boreo-nemoral rainforests in Vest-Agder, Rogaland, Hordaland and Møre og Romsdal. This study further reports C. fallax from Sweden (Gotland) for the first time. Outside Scandinavia, C. cuspidans is confirmed for Great Britain and C. fallax for Austria, Great Britain, Switzerland and Turkey (not shown).

Figure 6. 

The distribution of 111 specimens of Coniocarpon in Norway, Sweden and Denmark based on material in BG, C, S, TRH and UPS. A C. cinnabarinum B C. cuspidans C C. fallax. Single dots may represent several collections.

Coniocarpon preferably grows on trees with smooth bark and the selected host tree species slightly follow a geographical pattern (see Specimens examined below Taxonomic conclusions). Most collections of C. cinnabarinum from Norway have been made from Corylus avellana L., including few from Fraxinus excelsior L. and Sorbus aucuparia L. The species is collected in Denmark from C. avellana, F. excelsior and Fagus sylvatica L., and in Sweden from F. excelsior. Most specimens of C. cuspidans have been collected from C. avellana, but the species has been seen from a rather wide range of trees including F. excelsior, Ilex aquifolium L., Quercus robur L. and S. aucuparia. Coniocarpon fallax has mainly been collected from F. excelsior from Vest-Agder to Hordaland (more rarely from C. avellana), while all specimens from Møre og Romsdal are from C. avellana. The species is further collected from F. excelsior on Gotland and Austria, and from Picea orientalis (L.) Link in Turkey.

Discussion

Species designations are hypotheses to be tested as new evidence becomes available. Recent molecular systematic studies have repeatedly revealed evolutionary lineages within phenotypically delimited lichenized fungi (e.g., Steinová et al. 2013; Lücking et al. 2014; Bendiksby et al. 2015; Alors et al. 2016; Hawksworth and Lücking 2017; Boluda et al. 2019). Whether or not such morphologically indistinguishable, or “cryptic”, evolutionary lineages should be recognized at species level may have critical implications for conservation biology and other fields of biology that use species as a fundamental unit (Lumbsch and Leavitt 2011).

In general, several factors should be considered in the process of assessing species status: proper selection of genetic markers (multiple, unlinked loci and from different genomic compartments), presence of statistically supported phylogenetic lineages, sufficiently large sample size, corroborating non-molecular character variation, and thorough review of the taxonomic and nomenclatural history (Grube and Kroken 2000; Printzen 2010; Lumbsch and Leavitt 2011). Our phylogenetic analyses of Coniocarpon are based on unlinked, multilocus DNA sequence data (mtSSU, nucITS, RPB2) showing high statistical support for three distinct genetic lineages (Fig. 1). Re-examination of morphology against the molecular phylogeny of 26 specimens revealed that the three lineages are further supported by differences in ascospore size (Fig. 2), ascospore septation (Fig. 3), distribution of pruina and ascoma shape (Fig. 4). Furthermore, the three lineages differ in the amyloidity of the ascomatal gels and pigment patterns revealed by HPTLC (Fig. 5). Finally, an additional 94 specimens, for which molecular data are not available, were revised using the same morphological and/or chemical characters. As such, this study fulfills the factors recommended for assessing species status.

Evolutionary lineages that remain intact when living in sympatry with close relatives might deserve species status (Coates et al. 2018). The distribution maps in Fig. 6 show sympatry for the three distinct genetic lineages of Coniocarpon in Norway, providing strong indirect evidence that there are mechanisms prohibiting exchange of genetic material among them, supporting their acceptance at species level. Hence, the integrated data gathered in this study jointly support the hypothesis of three distinct species of Coniocarpon in Norway, viz. C. cinnabarinum, C. fallax and C. cuspidans. The latter species was hidden in the extensive synonymy of C. cinnabarinum as Arthonia cinnabarina f. cuspidans Nyl. and is herein resurrected (see Taxonomic conclusions below).

The present data indicate a narrower distribution in Norway for C. cinnabarinum and C. cuspidans as compared to C. fallax (Fig. 6). All available collections from Møre og Romsdal were identified as C. fallax. Coniocarpon cinnabarinum, the only species reported from that county in the Norwegian Red List of Species 2015 could not be confirmed. However, not all collections of Coniocarpon in Norway were available for this study, and previous distributions were partly based on human observations as well. The distribution of the three Coniocarpon species in Norway needs further evaluation in light of the present investigation. Moreover, this study confirms C. cinnabarinum for Denmark and Sweden. Coniocarpon fallax is reported from Sweden (Gotland) for the first time (Fig. 6), based on specimens from S and UPS previously identified as C. cinnabarinum.

Species diversity and abundance generally correlate with habitat preference. Most collections of Coniocarpon in Norway were made in Hordaland (C. cinnabarinum 22, C. cuspidans 24, C. fallax 20), in the core area of the boreo-nemoral rainforests, having the highest levels of humidity and low average winter temperatures (Blom et al. 2015; Moen 1999). In comparison, the number of studied specimens from Vest-Agder (C. cinnabarinum 0, C. cuspidans 2, C. fallax 3) at the southern limit of the boreo-nemoral rainforests is distinctly lower, which might be partly explained by fewer rainforest localities (Blom et al. 2015). The occurrence of C. fallax in Møre og Romsdal indicates a wider ecological amplitude as compared to C. cinnabarinum and C. cuspidans in terms of oceanity and temperature. Coniocarpon cuspidans is currently only confirmed for Norway and Great Britain, but as the species has not been distinguished from C. fallax in the past, it might have a wider distribution in Western Europe.

This study provides a further step in the process of improving the biodiversity conservation in Norway by applying an integrative taxonomic approach for delimiting taxa in Coniocarpon. Coniocarpon cinnabarinum and C. fallax are designated as VU in the latest version of the Norwegian Red List for Species (Henriksen and Hilmo 2015). It is likely that the same status applies to C. cuspidans, growing with similar abundance in the boreo-nemoral rainforests as the other species of Coniocarpon. Moreover, all three species are facing the same threats such as algae growth on trees which is inhibiting the establishment and growth of lichens (Blom 2018).

Taxonomic conclusions

Coniocarpon DC

In Lamarck and de Candolle, Flore française 2: 323 (1805) [MB 1208]. Lectotype (selected in Santesson, Symbolae Botanicae Upsalienses 12(1): 68, 1952): Coniocarpon cinnabarinum DC., in Lamarck and de Candolle, Flore française 2: 323 (1805).

Key to Coniocarpon in Norway

1 Ascospores mostly > 20 µm long; ascomata typically rounded to weakly lobate, rarely lirellate; orange–red pruina present C. cinnabarinum
Ascospores mostly ≤ 20 µm long; ascomata typically irregularly lirellate; orange–red pruina present or absent 2
2 Orange–red pruina present; ascospores (15–)17–20(–22) × (6–)7–9(–10) µm, (1–)3–4(–5) transversely septate C. fallax
Orange–red pruina absent; ascospores (15–)16–18(–20) × (6–)7–8(–9) µm, (2–)3(–4) transversely septate C. cuspidans

Coniocarpon cinnabarinum DC.

MycoBank No: 383614
Figs 2, 3A, 4A, B, 5, 6A

Coniocarpon cinnabarinum DC.: Lamarck and de Candolle, Flore française 2: 323 (1805). Type: not selected [see under "Notes" below].

= Spiloma ? tumidulum Ach., Methodus qua omnes detectos lichenes: 11 (1803) [MB 405550]. Type: Hispania, Schousboe (H-Ach. 3 c!, holotype).

= Spiloma tumidulum var. rubrum Ach., Lichenographia universalis: 137 (1810) – nom. illegit. Type: Gallia (H-Ach. 2 c!).

Description

Thallus pale olive gray to brown, weakly glossy to matt, smooth, endophloeodal to partly epiphloeodal, continuous; prothallus line dark gray to brown, sometimes present when in contact with other lichens; photobiont trentepohlioid, the cells rounded to elliptical, 7–12 × 4–8 μm, forming short chains. Ascomata irregularly rounded to elliptical to weakly lobed, rarely distinctly lirellate, with steep flanks, emergent from thallus, 0.1–0.4 × 0.1–0.3 mm, 95–140 µm tall, solitary or forming loose to dense aggregations of 3–15 ascomata, (0.3–)0.5–2.0(–3.5) × 0.3–1.6 mm; disc dark purple black, flat to weakly convex, matt to weakly glossy, white pruinose, a layer of orange–red pruina sometimes present above the white pruina; old ascomata sometimes epruinose; margins level with the disc, typically orange–red pruinose, sometimes with additional patches of white pruina; proper exciple brown, 8–15 µm wide, composed of compressed and vertically oriented paraphysoidal hyphae, the hyphae 1–2 µm thick, branched and netted, often forming short hairs up to 15 µm long at the outer margin; old bark cells often attached to the exciple; epithecium brown, 10–25 µm tall, conglutinated only in the lower parts, composed of branched tips of the paraphysoidal hyphae extending horizontally above the asci; the tips slightly widened to 3(–4) µm, sometimes extending from the epithecium as sparsely branched anticlinal hairs up to 22 µm long; hymenium hyaline, strongly conglutinated, (45–)65–90 µm tall, paraphysoids densely branched and netted, 1–2 µm thick; hypothecium hyaline, conglutinated, 20–35 µm tall, formed of irregular prosoplectenchymatic hyphae 1–2 µm diam.; crystals common in epithecium and proper exciple, of two types: hyaline, leafy crystals, 1–5(–8) µm, and orange, red or purple, granular crystals, 1–2(–4) µm; a weak amorphous, red to purple pigmentation present in exciple, epithecium and patchily distributed in the hymenium. ASCI of the Arthonia-type, long obpyriform to clavate, 62–84 µm × 24–35 µm (n = 34), 8-spored, the ascospores stacked; tholus 8–11 µm thick, lateral ascospore wall 1–2 µm thick. ASCOSPORES hyaline, (3–)4–5(–8) transversely septate, (19–)23–28(–30) × (8–)10–11(–12) µm (l: mean = 25.7, STD = 2.3; w: mean = 10.5, STD = 0.9; n = 132), obovate, with enlarged apical cell, getting pale brown with granular ornamentation in the epispore at late maturity; development macrocephalic.

Chemistry

Pigments A1, A3 and A4 in variable amounts detected by HPTLC. Proper exciple Idil+ blue, I+ blue, KI+ blue; epithecium Idil+ blue, I+ blue, KI+ blue; hymenium Idil+ red, I+ red, KI+ blue; hypothecium Idil+ red, I+ red, KI+ blue. A hemiamyloid ring present in the tholus of the asci. Hyaline crystals dissolve in K. Orange, red and purple crystals dissolve in K with a clear, fleeting, purplish solution.

Specimens examined

Norway – Rogaland • Rennesøy, Berge; 59°05.868'N, 05°42.320'E; on C. avellana; 30–40 m a.s.l.; 12. Jan. 2008; J. I. Johnsen leg.; BG L-86128. – Hordaland • Askøy, close to Ask farm; on S. aucuparia; 10–30 m a.s.l.; 31. Aug. 1909; J. J. Havaas leg.; UPS-L-277202. • Bømlo, Børøy, Storavatnet; 59°42.9420'N, 05°15.7680'E; on C. avellana; 30. Apr. 2018; G. Gaarder leg.; TRH-L-18030 • ibid.; Lykling; 59°42.6780'N, 05°12.3540'E; on C. avellana; 30. Apr. 2018; G. Gaarder leg.; TRH-L-18036 • ibid.; S of Liarnuten; on C. avellana; 21. Jun. 1997; T. Knutsson leg.; UPS-L-737333 • ibid.; Skogafjellet; 59°38.812'N, 05°12.098'E; on C. avellana; 35 m a.s.l., 19. Jul. 2017; A. Frisch leg.; TRH-L-29000 • ibid.; 59°38.818'N, 05°12.082'E; on C. avellana; 10 m a.s.l.; 19. Jul. 2017; A. Frisch leg.; TRH-L-29006, TRH-L-29007, TRH-L-29008 • ibid.; on F. excelsior; A. Frisch leg.; TRH-L-29009 • ibid.; 59°38.833'N, 05°12.153'E; on C. avellana; 50 m a.s.l.; 19. Jul. 2017; A. Frisch leg.; TRH-L-29001, TRH-L-29002. • Kvam, Gravdal SW, Geitaknottane Nat. Reserve, NE of Lønningshaugen; 60°06.690'N, 05°51.068'E; on C. avellana; 150–250 m a.s.l.; 28. Aug. 1997; P. G. Ihlen leg.; BG-L-35863. • Lindås, Kvalvika-Røyldalane; 60°38.338'N, 05°26.258'E; on C. avellana; 35 m a.s.l.; 14. May 2018; A. Frisch leg.; TRH-L-29010, TRH-L-29011, TRH-L-29012. • Os, Innerøya, Halhjem; 60°08.5020'N, 05°24.8520'E; on S. aucuparia; 10. May 2018; G. Gaarder leg.; TRH-L-18042 • ibid.; 60°08.5920'N, 05°25.3440'E; on C. avellana; 10. May 2018; G. Gaarder leg.; TRH-L-18043. • Stord, Digernes, Geitåsen; 59°45.402'N, 05°25.092'E; on C. avellana; 28. Apr. 2018; G. Gaarder leg.; TRH-L-18033 • ibid.; Valavåg, Nes-Åsen; 59°46.0740'N, 05°24.8040'E; on C. avellana; 27. Apr. 2018; G. Gaarder, U. Hanssen leg.; TRH-L-18087. • Tysnes, Beltestad, Beltestadknappen; 59°59.883'N, 05°27.543'E; on C. avellana; 13 m a.s.l.; 9. May 2018; A. Frisch leg.; TRH-L-29003, TRH-L-29004 • ibid.; 59°59.900'N, 05°27.555'E; on C. avellana; 5 m a.s.l.; 9. May 2018; A. Frisch leg.; TRH-L-29005. SWEDEN – Gotland • Stenkumla, Myrsö; 1869; Laurer leg.; UPS-L-002825. – Skåne • Dalby, Dalby Söderskog; on F. excelsior; 23. Jul. 1947; R. Santesson leg.; UPS-L-118296 • ibid.; Ottarp, Bälteberga; on F. excelsior; 20. Aug. 1946; O. Almborn leg.; S-F-71116, UPS-L-60625 • ibid.; 16. Sep. 1959; G. Degelius, O. Almborn leg.; UPS-L-60624. DENMARK – Jylland • Horsens, Elling Skov; on F. excelsior; 26. Mar. 1887; J. Jeppesen leg.; C-L-28996 • ibid.; on F. sylvatica; 26. Mar 1887; J. Jeppesen leg.; C-L-28993 • ibid.; on F. excelsior; 26. Feb. 1887; J. Jeppesen leg.; S-F-71202, C-L-28992 • ibid.; on F. excelsior; 20. Feb. 1887; J. P. Pedersen leg.; C-L-28991, C-L-28994 • ibid.; Hansted Skov; on F. excelsior; 5. Des. 1886; J. Jeppesen leg.; C-L-29000 • ibid.; 1. Feb. 1887; J. Jeppesen leg.; C-L-28999 • ibid.; on F. sylvatica; 6. Mar. 1887; J. Jeppesen leg.; C-L-28998. • Lihme, Kås skov; on C. avellana; 6. Aug. 1979; G. Thor leg.; UPS-L-165392 • ibid.; on F. excelsior; 25. May. 1976; S. Svane leg.; C-L-28997, C-L-28988 • ibid.; on C. avellana; 25. May 1976; M. S. Christansen leg.; C-L-28990 • ibid.; Bringsbjerg Krat; 56°37.129'N, 08°41.423'E; on C. avellana; 21. Oct. 2002; R. S. Larsen leg.; C-L-17076. – Sjælland • Haslev; 29. Jul. 1887; Taussieng leg.; C-L-28995 • ibid.; Skarresø; 4. Nov. 1870; C. Grönlund leg.; UPS-L-002896.

Notes

Coniocarpon cinnabarinum differs from the other Coniocarpon species in Norway by distinctly larger ascospores and in ascospore septation: (19–)23–28(–30) × (8–)10–11(–12) µm, (3–)4–5(–8) transversely septate vs (15–)16–18(–20) × (6–)7–8(–9) µm, (2–)3(–4) transversely septate in C. cuspidans vs (15–)17–20(–22) × (6–)7–9(–10) µm, (1–)3–4(–5) transversely septate in C. fallax. Further, the ascomata in C. cinnabarinum are mostly irregularly rounded to elliptical and only rarely lirellate as in C. cuspidans and C. fallax. The ascomatal disc in C. cinnabarinum is typically covered by a thick layer of white pruina which may be overlaid by orange–red pruina, and the ascomatal margin is orange–red pruinose. In C. cuspidans, the ascomata completely lack orange-red pruina, while a thin white pruina may be occasionally present. In C. fallax, the distribution of pruina is similar to C. cinnabarinum, but the white pruina is less pronounced and may even be lacking. Additional differences have been observed in the reaction of proper exciple and epithecium to iodine: Idil/I+ blue in C. cinnabarinum and C. fallax vs Idil/I+ red in C. cuspidans. The quinoid pigments A1, A3 and A4 have been identified in C. cinnabarinum in various amounts. The quinoid patterns in C. fallax are similar, while in C. cuspidans A3 is absent or occurs in trace amounts only. The pigment A2 has only been found in C. cuspidans.

Coniocarpon cinnabarinum is the selected type species of Coniocarpon (Santesson 1952), but the name is preceded by Spiloma tumidulum (Acharius 1803) and possibly Sphaeria gregaria Weigel (1772). The situation is further complicated by the fact that the only specimen of C. cinnabarinum in PC that is unequivocally linked to the publication of Flore française and donated by de Candolle, represents C. fallax. Since C. cinnabarinum is a well established species and often cited in literature, we intend to propose this name for conservation. The typification of the species will be discussed in the proposal, which is currently under preparation.

Coniocarpon cuspidans (Nyl.) Moen, Frisch & Grube, comb. nov.

MycoBank No: 833812
Figs 2, 3C, 4C, D, 5, 6B

Arthonia cinnabarina f. cuspidans Nyl., Flora 59: 310 (1876) [MB 372360]. Type: Ilicicola in Hibernia, n. 6, Dough[ruagh Mountain], 1875, Larbalestier (H-Nyl 5607! lectotype, here selected).

Description

Thallus pale brown to pale fawn to off white, matt to weakly glossy, smooth, endophloeodal to partially epiphloeodal, continuous; prothallus line dark gray to brown to black, sometimes present when in contact with other lichens; photobiont trentepohlioid, the cells rounded to elliptical, 6–13 × 5–11 µm forming short chains. Ascomata weakly elongate to irregularly lirellate, with steep flanks, emergent from thallus, 0.2–0.6 × 0.1–0.2 mm, 60–105 µm tall, typically forming loose to dense aggregations of 3–15 ascomata, weakly elongated to irregularly lirellate to stellate, 0.4–1.8(–2.5) × (0.1)0.3–1.0(–2.0) mm; disc black to dark purple black, flat to weakly convex, weakly glossy to matt, epruinose, rarely with patches of a thin white pruina; margins level with the disc, epruinose, rarely with patches of a thin white pruina; proper exciple brown, 7–20 µm wide, composed of compressed and vertically oriented paraphysoidal hyphae, the hyphae 1–2 µm thick, branched and netted, sometimes forming short hairs up to 16 µm long at the outer margin; old bark cells sometimes attached to the exciple; epithecium brown, 8–20 µm tall, conglutinated only in the lower parts, composed of branched tips of the paraphysoidal hyphae extending horizontally above the asci; the tips slightly widened to 3(–4) µm, sometimes extending from the epithecium as sparsely branched anticlinal hairs up to 12 µm long; hymenium hyaline, strongly conglutinated, 41–73 µm tall, paraphysoids densely branched and netted, 1–2 µm thick; hypothecium hyaline, conglutinated, 15–30 µm tall, formed of irregular prosoplectenchymatic hyphae 1–2 µm diam.; crystals common in epithecium and proper exciple, of two types: hyaline, leafy crystals, 1–5 µm, and red or purple granular crystals, 1–3 µm; a weak amorphous, red to purple pigmentation present in exciple, epithecium and patchily distributed in the hymenium. ASCI of the Arthonia-type, long obpyriform to clavate, 45–70 × 19–28 µm (n = 31), 8-spored, the ascospores stacked; tholus 4–8 µm thick, lateral ascospore wall 1–2 µm thick. ASCOSPORES hyaline, (2–)3(–4) transversely septate, (15–)16–18(–20) × (6–)7–8(–9) µm (l: mean = 17.4, STD = 1.2; w: mean = 7.5, STD = 0.7; n = 100), obovate, with enlarged apical cell, getting pale brown with granular ornamentation in the epispore at late maturity; development macrocephalic.

Chemistry

Pigments A1, A2, A3 and A4 in variable amounts detected by HPTLC. Proper exciple Idil+ red, I+ red, KI+ blue; epithecium Idil+ red, I+ red, KI+ blue; hymenium Idil+ red, I+ red, KI+ blue; hypothecium Idil+ red, I+ red, KI+ blue. A hemiamyloid ring in the tholus of the asci not observed. Hyaline crystals dissolve in K. Purple crystals dissolve in K with hyaline solution. Red and purple crystals dissolve in K with purplish solution.

Specimens examined

Norway – Vest-Agder • Flekkefjord, Hidra, Høgåsen; 58°13.585'N, 06°33.370'E; on S. aucuparia; 35 m a.s.l.; 15. Jul. 2017; J. Klepsland leg.; TRH-L-29017 • ibid.; Nonfjell; 58°13.445'N, 06°33.550'E; 5–25 m a.s.l.; 15. Jul. 2017; A. Frisch TRH-L-29022. – Rogaland • Tysvær, Svinali W; 59°25.098'N, 05°34.104'E; on C. avellana; 19. Oct. 2017; G. Gaarder leg.; TRH-L-18034. – Hordaland • Austevoll, Huftaøy, Bjelland farm NE; 60°05.000'N, 05°16.000'E; on C. avellana; 0–40 m a.s.l.; 6. Jun. 1996; T. Tønsberg leg.; BG-L-32077, BG-L-34115. • Bømlo, Børøy, Masterhaugane nord; 59°42.840'N, 05°15.5040'E; on S. aucuparia; 30. Apr. 2018; G. Gaarder leg.; TRH-L-18038, TRH-L-18078 • ibid.; 59°42.6420'N, 05°15.3540'E; on C. avellana; 30. Apr. 2018; G. Gaarder leg.; TRH-L-18040 • ibid.; Kuhillerdalen; 59°45.36'N, 05°16.82'E; on C. avellana; 70 m a.s.l.; 11. May 2015; J. B. Jordal, H. H. Blom leg.; TRH-L-16794 • ibid.; Lykling, Lyklingfjorden N; 59°42.321'N, 05°10.569'E; on C. avellana; 10–20 m a.s.l.; 13. May 1996; T. Tønsberg leg.; BG-L-31539 • ibid.; 59°42.300'N, 05°10.600'E; on C. avellana; 40–60 m a.s.l.; 1. Jun. 1997; S. Ekman leg.; BG-L-38200 • ibid.; Skogafjellet; 59°38.833'N, 05°12.153'E; on C. avellana; 50 m a.s.l; 19. Jul. 2017; A. Frisch TRH-L-29015 • ibid.; 59°38.812'N, 05°12.098'E; on C. avellana; 35 m a.s.l.; 19. Jul. 2017; A. Frisch leg.; TRH-L-29013, TRH-L-29014 • ibid.; 59°38.818'N, 05°12.082'E; on C. avellana; 10 m a.s.l.; 19. Jul. 2017; A. Frisch leg.; TRH-L-29023 • ibid.; on S. Aucuparia; 10 m a.s.l.; 19. Jul. 2017; A. Frisch leg.; TRH-L-29024 • ibid.; 59°38.972'N, 05°12.432'E; on C. avellana; 12 m a.s.l.; 19. Jul. 2017; A. Frisch leg.; TRH-L-29025, TRH-L-29026 • ibid.; S of Totlandstjørna; 59°41.172'N, 05°20.940'E; on C. avellana; 28. Jun. 2017; G. Gaarder leg.; TRH-L-18035 • ibid.; Totsida; 59°40.8180'N, 05°19.5540'E; on I. aquifolium; 27. Jun. 2017; G. Gaarder, M. Lorentzen leg.; TRH-L-18037 • ibid.; Våge; 59°43.6260'N, 05°13.5060'E; on C. avellana; 30. Apr. 2018; G. Gaarder leg.; TRH-L-18031. • Fusa, Holmefjord, Eikhaugen; 60°17.952'N, 05°39.867'E; on Q. robur; 30 m a.s.l.; 8. May 2018; A. Frisch leg.; TRH-L-29021. • Kvam, Nes N; 60°10.085'N, 05°55.535'E; on F. excelsior; 4. Jun. 2018; G. Gaarder leg.; TRH-L-18608. • Stord, Åsen SW of Sagvåg; 59°46.043'N, 05°24.593'E; on C. avellana; 50 m a.s.l.; 28. Apr. 2018; A. Frisch leg.; TRH-L-29019 • ibid.; 59°46.088'N, 05°24.757'E; on C. avellana; 33 m a.s.l.; 28. Apr. 2018; A. Frisch leg.; TRH-L-29020. • Sund, Steinsland; 60°12.199'N, 05°04.929'E; on C. avellana; 20–40 m a.s.l.; 9. Mar. 1997; T. Tønsberg leg.; BG-L-34117. • Tysnes, Beltestad, Beltestadknappen; 59°59.883'N, 05°27.543'E; on C. avellana; 13 m a.s.l.; 9. May 2018; A. Frisch leg.; TRH-L-29018.

Additional examined specimens

Great Britain – Scotland • Argyll, Appin, Glen Stockdale; 56°34.383'N, 05°21.583'W; on C. avellana; 65–90 m a.s.l.; 5. Jun. 2018; A. Acton, J. Malíček, Z. Palice leg. 25146; PRA • ibid.; Westerness Locli Surnart, Reripole Ravine; on Corylus sp.; 10. Mar. 1983; B. J. Coppins leg.; BG-L-58163 • ibid.; Benderloch, Lochnell House; 17/88.38; on Corylus; 15–45 m a.s.l.; 4. Aug. 1980; B. J. Coppins leg. 8056; E • Mid Perth (VC 88), S side of Loch Earn, west of Ardvorlich cottage; 27/ 61.22(-3); 105–130 m a.s.l.; 8. Aug. 1980; B. J. Coppins & P. W. James leg.; E • Clyde Islands, Arran (VC 100), Glenashdale; 26/03.25; on Corylus; 4. Apr. 1984; B. J. Coppins leg. 10169; E • Westerness (VC 97), S side of Loch Sunart, Laudale Woods ravine 0.5 km W of Liddesdale; 17/77.59; 0–75 m a.s.l.; 9. Mar. 1983; B. J. Coppins & P. M. Jørgensen leg.; E • Devon, SW of Bridge Reeve; 21/65.13; on Quercus petraea; 1. Aug. 1972; P. Harrold leg.; E.

Notes

Coniocarpon cuspidans is characterized by lirellate ascomata lacking orange–red pruina, while red and purple pigment crystals and a weak amorphous red to purple pigmentation is present in proper exciple and epithecium. The quinoid pigment A3 is absent or occurs in possible trace amounts only. A3 may correspond to orange pigment crystals observed in microscopical preparations of C. cinnabarinum and C. fallax but not in C. cuspidans. These crystals are located in the pruina of the ascomatal margin and disc. The ascospores in C. cuspidans are the smallest observed for the genus in Norway, (15–)16–18(–20) × (6–)7–8(–9) µm and (2–)3(–4) transversely septate. A hemiamyloid ring in the tholus of the asci, which is present in C. cinnabarinum and C. fallax, could not be observed in the investigated material. Further differences to C. cinnabarinum and C. fallax are discussed under those species.

The protologue of Arthonia cinnabarina f. cuspidans (Nylander 1876) cites specimens from Ireland and Cuba as original material: “Ilicicola in Hibernia (Larbalestier)” and “Exotica eadem datur in C. Wright. Cub. no. 123 a et b”. Nylander obviously considered the material from Ireland as the factual type collection. We have selected a specimen from the Nylander herbarium in Helsinki as the lectotype, which is the only specimen from Ireland that undoubtedly has been seen by Nylander. A possible syntype exists in BM: “Derryclare, Connemara, ilicicola, 1876 [BM000974345]”. Another specimen [BM000974347] from Larbalestier’s herbarium has a printed later label that only states “On young trees. Doughruagh Mountain and other places in Connemara”. The type status of this specimen is unclear. Both specimens have been seen by us as high-resolution pictures obtained from the data portal of BM.

Coniocarpon fallax (Ach.) Grube

MycoBank No: 808766
Figs 2, 3B, 4E, F, 5, 6C

Coniocarpon fallax (Ach.) Grube: Frisch et al., Taxon 63: 737 (2014). Spiloma fallax Ach., Methodus qua omnes dectectos lichenes: 10 (1803) [MB 405518]. Type: Germania (H-Ach. 2 a!, lectotype, selected in Frisch et al., Taxon 63: 737, 2014).

= Spiloma elegans Ach., Lichenographia universalis: 135 (1810) [MB 405516]. Coniocarpon elegans (Ach.) Duby, Aug. Pyrami de Candolle Botanicon Gallicum: 675 (1830) [MB 383617]. Lichen elegans (Ach.) Lam. in Lamarck and Poiret, Encyclopédie méthodique, botanique, suppl. 3(1): 352 (1813) [MB 122540]. Arthonia elegans (Ach.) Almq., Kongliga Svenska vetenskaps-akademiens handlingar 17(6): 19 (1880) [MB 118959]. Type: Schleicher, Plantae Cryptogamicae Helvetiae exsiccatae, centuria5 no. 54 (S, lectotype, selected in Van den Broeck et al., Plant Ecology and Evolution 151: 346, 2018).

Description

Thallus pale fawn to gray brown, weakly glossy to matt, smooth, endophloeodal to partly epiphloedal, continuous; prothallus line dark grey to brown, sometimes present when in contact with other lichens; photobiont trentepohlioid, the cells rounded to elliptical, 8–13 × 5–10 µm, forming short chains. Ascomata weakly elongate to irregularly lirellate, with steep flanks, emergent from thallus, 0.2–0.4 mm × 0.1–0.2 mm, 65–110 µm tall, typically forming loose to dense aggregations of 3–15 ascomata, weakly elongated to irregularly lirellate to stellate, 0.2–1.5(–2.3) × 0.1–1.8 mm; disc black to dark purple black, flat to weakly convex, weakly glossy to matt, epruinose or with a thin layer of white pruina; margins level with the disc, orange–red pruinose, sometimes with additional patches of white pruina; proper exciple brown, 7–20 µm wide, composed of compressed and vertically oriented paraphysoidal hyphae, the hyphae 1–2 µm thick, branched and netted, often forming short hairs up to 17 µm long at the outer margin; old bark cells sometimes attached to the exciple; epithecium brown, 10–20 µm tall, conglutinated only in the lower parts, composed of branched tips of the paraphysoidal hyphae extending horizontally above the asci; the tips slightly widened to 3(–4) µm, sometimes extending from the epithecium as sparsely branched anticlinal hairs up to 24 µm long; hymenium hyaline, strongly conglutinated, 35–70 µm tall, paraphysoids densely branched and netted, 1–2 µm thick; hypothecium hyaline, conglutinated, 15–30 µm tall, formed of irregular prosoplectenchymatic hyphae 1–2 µm diam.; crystals common in epithecium and proper exciple, of two types: hyaline, leafy crystals, 1–5(–7) µm, and orange, red or purple granular crystals, 1–2(–4) µm; a weak amorphous, red to purple pigmentation present in exciple, epithecium and patchily distributed in the hymenium. ASCI of the Arthonia-type, long obpyriform to clavate, 50–75 × 20–32 µm (n = 33), 8-spored, the ascospores stacked; tholus 5–8 µm thick, lateral ascospore wall 1–2 µm thick. Ascospores hyaline, (1–)3–4(–5) transversely septate, (15–)17–20(–22) × (6–)7–9(–10) µm (l: mean = 18.5, STD = 1.9; w: mean = 8.2, STD = 0.9; n = 191), obovate, with enlarged apical cell, getting pale brown with granular ornamentation in the epispore at late maturity; development macrocephalic.

Chemistry

Pigments A1, A3 and A4 in variable amounts detected by HPTLC. Proper exciple Idil+ blue, I+ dark blue, KI+ dark blue; epithecium Idil+ blue, I+ dark blue, KI+ dark blue; hymenium Idil+ red, I+ red brown, KI+ dark blue; hypothecium Idil+ blue, I+ dark blue, KI+ dark blue. A hemiamyloid ring present in the tholus of the asci. Hyaline, crystals dissolve in K. Orange, red and purple crystals dissolve in K with purplish solution.

Specimens examined

Norway – Vest-Agder • Lyngdalsfjord; on F. excelsior; 8. Apr. 1905; A. H. Magnusson leg.; S-F-71115 • ibid.; 1925; A. H. Magnusson leg.; S-F-71114 • ibid.; on F. excelsior; 25. Jan. 1939; A. H. Magnusson leg.; UPS-L-002899. – Rogaland • Gjesdal, Dirdal NE; on F. excelsior; 6. Oct. 1984; S. Hultengren leg.; UPS-L-654296 • ibid.; 58°49.810'N, 06°11.970'E; on F. excelsior; 160 m a.s.l.; 12. Jul. 2017; A. Frisch leg.; TRH-L-29036, TRH-L-29027. – Hordaland • Askøy; on C. avellana and F. excelsior; 1909; J. J. Havaas leg.; UPS-L-137313 • Fusa, Tveitane; 13 km S of Mundheim; 60°03'N, 05°52'E; on C. avellana; 150 m a.s.l.; 18. Aug. 1995; A. Nordin leg.; UPS-L-61739 • ibid.; Øvre Hålandsdalen, W of Orra; 60°15.5065'N, 05°55.137'E; on F. excelsior; 120 m a.s.l.; 24. Feb. 2015; S. Vatne leg.; TRH-L-16793. • Kvam, Daleelva N; 60°07.453'N, 05°52.721'E; on F. excelsior; 10. Jun. 2018; G. Gaarder leg.; TRH-L-18605 • ibid.; 60°07.668'N, 05°52.898'E; on F. excelsior; 13. Jun. 2018; G. Gaarder leg.; TRH-L-18606 • ibid.; Furhovda; 60°09.243'N, 05°53.910'E; on F. excelsior; 5. Jun. 2018; G. Gaarder, M. Lorentzen leg.; TRH-L-18604 • ibid.; Hovden; 60°13.735'N, 05°59.633'E; on F. excelsior; 11. Jun. 2018; G. Gaarder leg.; TRH-L-18607. • Lindås, Helltveit W; 60°37.872'N, 05°26.080'E; on F. excelsior; 24. Jul. 1980; T. Tønsberg leg.; BG-L-26222 • ibid.; Kvalvika-Røyldalane; 60°38.245'N, 05°26.345'E; on F. excelsior; 35 m a.s.l.; 14. May 2018; A. Frisch, J. Klepsland leg.; TRH-L-29035. • Os, Li; 60°10.38'N, 05°26.45'E; on F. excelsior; 60–120 m a.s.l.; 22. Jul. 1979; T. Tønsberg leg.; BG-L-26221, BG-L-26221. • Stord, Valavåg, Nes-Åsen; 59°46.116'N, 05°24.660'E; on F. excelsior; 27. Apr. 2018; G. Gaarder, U. Hanssen leg.; TRH-L-18041. • Tysnes, SE slope of Skardnipa near Teigen; 59°58.360'N, 05°39.008'E; on F. excelsior; 23 m a.s.l.; 9. May 2018; A. Frisch, J. Klepsland leg.; TRH-L-29031 • ibid.; 59°58.361'N, 05°39.023'E; on F. excelsior; 19 m a.s.l.; 9. May 2018; A. Frisch, J. Klepsland leg.; TRH-L-29032, TRH-L-29033 • ibid.; 59°58.350'N, 05°38.992'E; on F. excelsior; 24 m a.s.l.; 9. May 2018; A. Frisch, J. Klepsland leg.; TRH-L-29034 • ibid.; Sunde, Loksund; on F. excelsior; 27. Aug. 1910; J. J. Havaas leg.; UPS-L-137512. • Tysnes, Tysnesøy, N of Onarheim; 59°58.377'N, 05°39.047'E; on F. excelsior; 30 m a.s.l.; 21. Jul. 2017; A. Frisch leg.; TRH-L-29028, TRH-L-29029. – Møre og Romsdal • Fræna, S of Hustad, Lunheim; 62°55.3871'N, 07°06.892'E; on C. avellana; 60 m a.s.l.; 15. Apr. 2016; H. Holien leg.; TRH-L-17089 • ibid.; Nordmark E; 62°54.998'N, 07°07.068'E; on C. avellana; 80 m a.s.l.; 26. Apr. 1998; G. Gaarder leg.; BG-L-39619 • ibid.; Tverrfjell; 62°54.872'N, 07°16.243'E; on C. avellana; 60 m a.s.l.; 5. Jul. 2017; A. Frisch leg.; TRH-L-29037, TRH-L-29030. • Skodje, Igletjønna; 62°28.6565'N, 06°34.4845'E; on C. avellana; 20. Apr. 2014; G. Gaarder, P. Larsen leg.; TRH-L-16791. • Tingvoll, Kamsvågtrøa; 63°01.8998'N, 08°08.150'E; on C. avellana; 16. Feb. 2014; G. Gaarder leg.; TRH-L-15366 • ibid.; Langvatnet NE; 63°02.9460'N, 08°04.3020'E; on C. avellana; 8. Nov. 2014; G. Gaarder leg.; TRH-L-16792 • ibid.; Skjelberget; 63°03.2040'N, 08°03.8640'E; on C. avellana; 25. Apr. 2014; G. Gaarder leg.; TRH-L-16789 • ibid.; Årøyvatnet; 63°02.8320'N, 08°02.3940'E; on C. avellana; 8. Nov. 2014; G. Gaarder leg.; TRH-L-16790. SWEDEN – Gotland • Bunge, Hägur, Mörku, 1 km NE of the church; 57°51.00'N, 19°00.00'E; on F. excelsior; 15 m a.s.l.; 27. Apr. 1996; A. Nordin leg.; UPS-L-74225 • 600 m from Bunge church; on F. excelsior; 27. Apr 1996; G. Westling leg.; S L-52399; • Bäl, Gute, 1 km E Bäl church; 57°39.00'N, 18°40.00'E; on F. excelsior; 25. Nov. 1996; P. Johanson leg.; UPS-L-98398.

Additional examined specimens

Austria – Upper Austria • Totes Gebirge, Lake Almsee SSE; 47°44.600'N, 13°57.400'E; on F. excelsior; 600 m a.s.l.; 31. May 1998; T. Tønsberg leg.; BG-L-66299. Turkey – Trabazon • Trabazon Vilayet, Uzungöl c. 14 km SSE of Caykara; 40°36.8700'N, 40°18.8500'E; on P. orientalis; 24. Jun. 2001; C. Printzen, B. Kanz leg.; BG-L-77481.

Notes

Coniocarpon fallax resembles C. cinnabarinum by ascomata that are covered in an orange–red and white pruina. However, the white pruina is less pronounced in C. fallax and may be absent. The ascomata of C. fallax further are elongate to clearly lirellate and the ascospores are distinctly smaller with less septa: (15–)17–20(–22) × (6–)7–9(–10) µm, (1–)3–4(–5) transversely septate vs (19–)23–28(–30) × (8–)10–11(–12) µm, (3–)4–5(–8) transversely septate in C. cinnabarinum. Mature, apparently well-developed elliptical ascospores with only a single septum were found in one specimen (TRH-L-17089) of C. fallax. Coniocarpon cuspidans has ascospores of similar size which, however, are predominantly 3-septate. The elongate to lirellate ascomata of that species lack an orange–red pruina, and proper exciple and epithecium react Idil/I+ red in iodine.

Acknowledgements

The curators of the herbaria BG, C, H, O, PC, S, TRH and UPS are acknowledged for the loan of specimens and permission to study the material. Zdenek Palice kindly sent specimens of Coniocarpon from Great Britain (Scotland). The authors acknowledge financial support from the Norwegian Biodiversity Information Centre through the Species Project (pnr.70184237) Three storied diversity: Mapping and barcoding crustose lichens and lichenicolous fungi in the Norwegian rainforests.

References

  • Acharius E (1803) Methodus qua omnes detectos lichenes. C. F. Marquard, Stockholmiae, iv, 1–393.
  • Acharius E (1810) Lichenographia universalis. I. F. Danckwerts, Gottingae: viii, 1–696.
  • Almquist S (1880) Monographia Arthoniarum Scandinaviae. Kongliga Svenska Vetenskaps-Akademiens Handlingar 17(6): 3–69.
  • Alors D, Lumbsch HT, Divakar PK, Leavitt SD, Crespo A (2016) An integrative approach for understanding diversity in the Punctelia rudecta species complex (Parmeliaceae, Ascomycota). PLoS ONE 11(2): e0146537. https://doi.org/10.1371/journal.pone.0146537
  • Arup U, Ekman S, Lindblom L, Mattsson J-E (1993) High performance thin layer chromatography (HPTLC), an improved technique for screening lichen substances. The Lichenologist 25: 61–71. https://doi.org/10.1006/lich.1993.1018
  • Bendiksby M, Haugan R, Spribille T, Timdal E (2015) Molecular phylogenetics and taxonomy of the Calvitimela aglaea complex (Tephromelataceae, Lecanorales). Mycologia 107: 1172–1183. https://doi.org/10.3852/14-062
  • Blom HH (2018) Ikke eller svært lite uttørkningseksponert boreonemoral og sørboreal skog i klart til sterkt oseanisk seksjon, Skog. Norsk rødliste for naturtyper 2018. Artsdatabanken, Trondheim. https://artsdatabanken.no/RLN2018/84 [accessed 6 May 2019]
  • Blom HH, Gaarder G, Ihlen PG, Jordal JB, Evju M (2015) Poor boreonemoral rain-forest – a hotspot-habitat. Final report from the third period of the ARKO project. NINA Report 1169. Norsk institutt for naturforskning (NINA), Trondheim 1–97.
  • Boluda CG, Rico VJ, Divakar PK, Nadyeina O, Myllys L, McMullin RT, Zamora JC, Scheidegger C, Hawksworth DL (2019) Evaluating methodologies for species delimitation: the mismatch between phenotypes and genotypes in lichenized fungi (Bryoria sect. Implexae, Parmeliaceae). Persoonia 42: 75–100. https://doi.org/10.3767/persoonia.2019.42.04
  • Carlsen T, Bendiksby M, Hofton TH, Reiso S, Bakkestuen V, Haugan R, Kauserud H, Timdal E (2012) Species delimitation, bioclimatic range, and conservation status of threatened lichen Fuscopannaria confusa. The Lichenologist 44: 565–575. https://doi.org/10.1017/S0024282912000199
  • Coates DJ, Byrne M, Moritz C (2018) Genetic diversity and conservation units: Dealing with the species-population continuum in the age of genomics. Frontiers in Ecology and Evolution 6: 165. https://doi.org/10.3389/fevo.2018.00165
  • Coppins BJ, Aptroot A (2009) Arthonia Ach. (1806). In: Smith CW, Aptroot A, Coppins BJ, Fletcher OL, James PA, Wolsely PA (Eds) The Lichens of Great Britain and Ireland. 6th edition. The British Lichen Society, 153–171.
  • DellaSala DA, Alaback P, Drescher A, Holien H, Spribille T, Ronnenberg K (2011) Temperate and boreal rainforest relicts of Europe. In: DellaSala AD (Eds) Temperate and Boreal Rainforests of the World: Ecology and Conservation. Island Press, Washington DC, 154–160. https://doi.org/10.5822/978-1-61091-008-8_6
  • Duby JE (1830) Aug. Pyrami de Candolle Botanicon Gallicum. Editio secunda. Pars secunda, planta cellulares continens, Desray, Paris, 545–1068.
  • Frisch A, Ohmura Y, Ertz D, Thor G (2015) Inoderma and related genera in Arthoniaceae with elevated white pruinose pycnidia or sporodochia. The Lichenologist 47: 233–256. https://doi.org/10.1017/S0024282915000201
  • Grube M (1995) A taxonomic survey of arthonioid fungi with reddish K+ reactive pigments. Doctoral dissertation, Karl-Franzens-Universität, Graz.
  • Hawksworth DL, Lücking R (2017) Fungal diversity revisited: 2.2 to 3.8 million species. Microbiology Spectrum 5(4): FUNK-0052-2016.
  • Henriksen S, Hilmo O (2015) The Norwegian red list for species v.1.2. Artsdatabanken, Norway, 1–193.
  • Huelsenbeck JP, Ronquist F, Nielsen R, Bollback JP (2001) Bayesian inference of phylogeny and its impact on evolutionary biology. Science 294: 2310–2314. https://doi.org/10.1126/science.1065889
  • Lamarck MM, de Candolle AP (1805) Flore française. Troisieme edition. Tome second. H. Agasse, Paris, xii, 1–600.
  • Lamarck MM, Poiret JLM (1813) Encyclopédie méthodique. Botanique, suppl. 3(1) H. Agasse, Paris, 1–368.
  • Lanfear R, Frandsen PB, Wright AM, Senfeld T, Calcott B (2016) PartitionFinder 2: new methods for selecting partitioned models of evolution for molecular and morphological phylogenetic analyses. Molecular Biology and Evolution 34: 772–773. https://doi.org/10.1093/molbev/msw260
  • Larena I, Salazar O, González V, Julián MC, Rubio V (1999) Design of a primer for ribosomal DNA internal transcribed spacer with enhanced specificity for ascomycetes. Journal of Biotechnology 75: 187–194. https://doi.org/10.1016/S0168-1656(99)00154-6
  • Leavitt SD, Johnson L, St. Clair LL (2011) Species delimitation and evolution in morphologically and chemically diverse communities of the lichen forming genus Xanthoparmelia (Parmeliaceae, Ascomycota) in Western North America. American Journal of Botany 98: 175–188. https://doi.org/10.3732/ajb.1000230
  • Lücking R, Dal-Forno M, Sikaroodi M, Gillevet PM, Bungartz F, Moncada B, Yánez-Ayabaca A, Chaves JL, Coca FL, Lawrey JD (2014) A single macrolichen constitutes hundreds of unrecognized species. Proceedings of the National Academy of Sciences of the United States of America 111: 11091–11096. https://doi.org/10.1073/pnas.1403517111
  • Miller MA, Pfeiffer W, Schwartz T (2010) Creating the CIPRES Science Gateway for inference of large phylogenetic trees. Proceedings of the Gateway Computing Environments Workshop (GCE), New Orleans, LA, 1–8. https://doi.org/10.1109/GCE.2010.5676129
  • Moen A (1999) National atlas of Norway: vegetation. Norwegian Mapping Authority, 1–200.
  • Molina MC, Del-Prado R, Divakar PK, Sánchez-Mata D, Crespo A (2011) Another example of cryptic diversity in lichen-forming fungi: The new species Parmelia mayi (Ascomycota: Parmeliaceae). Organism, Diversity & Evolution 11: 331–342. https://doi.org/10.1007/s13127-011-0060-4
  • Nixon KC (1999–2002) WinClada ver. 1.0000. Published by the author, Ithaca, NY, USA.
  • Nylander W (1876) Addenda nova ad Lichenographiam europaeam XXV. Flora 59: 305–311.
  • Pattengale ND, Alipour M, Binida-Emonds OR, Moret BM, Stamatakis A (2010) How many bootstrap replicates are necessary? Journal of Computational Biology 17: 337–354. https://doi.org/10.1089/cmb.2009.0179
  • Printzen C (2010) Lichen systematics: the role of morphological and molecular data to reconstruct phylogenetic relationships. In: Lüttge U (Ed.) Progress in Botany. Berlin, Springer, 233–275. https://doi.org/10.1007/978-3-642-02167-1_10
  • QGIS Development Team (2019) QGIS Geographic Information System. Open Source Geospatial Foundation Project.
  • Santesson R (1952) Foliicolous lichens I. A revision of the taxonomy of the obligately foliicolous, lichenized fungi. Symbolae Botanicae Upsalienses 12(1): 1–590.
  • Steinová J, Steroos S, Grube M, Škaloud P (2013) Genetic diversity and species delimitation of the zeorin-containing red-fruited Cladonia species (lichenized Ascomycota) assessed with ITS rDNA and b-tubulin data. The Lichenologist 45(5): 665–684. https://doi.org/10.1017/S0024282913000297
  • Van den Broeck D, Frisch A, Razafindrahaja T, Van de Vijver B, Ertz D (2018) Phylogenetic position of Synarthonia (lichenized Ascomycota, Arthoniaceae), with the description of six new species. Plant Ecology and Evolution 151: 327–351. https://doi.org/10.5091/plecevo.2018.1506
  • White TJ, Bruns T, Lee S, Taylor J (1990) Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, White TJ (Eds) PCR Protocols: A Guide to Methods and Applications, Academic Press, New York, 315–322. https://doi.org/10.1016/B978-0-12-372180-8.50042-1
  • Zoller S, Scheidegger C, Sperisen C (1999) PCR primers for the amplification of mitochondrial small subunit ribosomal DNA of lichen-forming ascomycetes. The Lichenologist 31: 511–516. https://doi.org/10.1006/lich.1999.0220
login to comment