Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


Based on DNA sequence data, the genus Leptosillia is shown to belong to the Xylariales. Molecular phylogenetic analyses of ITS-LSU rDNA sequence data and of a combined matrix of SSU-ITS-LSU rDNA, rpb1, rpb2, tef1 and tub2 reveal that the genera Cresporhaphis and Liberomyces are congeneric with Leptosillia. Coelosphaeria fusariospora, Leptorhaphis acerina, Leptorhaphis quercus f. macrospora, Leptorhaphis pinicola, Leptorhaphis wienkampii, Liberomyces pistaciae, Sphaeria muelleri and Zignoëlla slaptonensis are combined in Leptosillia, and all of these taxa except for C. fusariospora, L. pinicola and L. pistaciae are epitypified. Coelosphaeria fusariospora and Cresporhaphis rhoina are lectotypified. Liberomyces macrosporus and L. saliciphilus, which were isolated as phloem and sapwood endophytes, are shown to be synonyms of Leptosillia macrospora and L. wienkampii, respectively. All species formerly placed in Cresporhaphis that are now transferred to Leptosillia are revealed to be non-lichenized. Based on morphology and ecology, Cresporhaphis chibaensis is synonymised with Rhaphidicyrtis trichosporella, and C. rhoina is considered to be unrelated to the genus Leptosillia, but its generic affinities cannot be resolved in lack of DNA sequence data. Phylogenetic analyses place Leptosillia as sister taxon to Delonicicolaceae, and based on morphological and ecological differences, the new family Leptosilliaceae is established. Furfurella, a new genus with the three new species, F. luteostiolata, F. nigrescens and F. stromatica, growing on dead branches of mediterranean fabaceous shrubs from tribe Genisteae, is revealed to be the closest relative of Delonicicola in the family Delonicicolaceae, which is emended. ITS rDNA sequence data retrieved from GenBank demonstrate that the Leptosilliaceae were frequently isolated or sequenced as endophytes from temperate to tropical regions, and show that the genus Leptosillia represents a widely distributed component of endophyte communities of woody plants.

Free full text 


Logo of persooniaLink to Publisher's site
Persoonia. 2019 Jun; 42: 228–260.
PMCID: PMC6712540
PMID: 31551620

Lichens or endophytes? The enigmatic genus Leptosillia in the Leptosilliaceae fam. nov. (Xylariales), and Furfurella gen. nov. (Delonicicolaceae)

Abstract

Based on DNA sequence data, the genus Leptosillia is shown to belong to the Xylariales. Molecular phylogenetic analyses of ITS-LSU rDNA sequence data and of a combined matrix of SSU-ITS-LSU rDNA, rpb1, rpb2, tef1 and tub2 reveal that the genera Cresporhaphis and Liberomyces are congeneric with Leptosillia. Coelosphaeria fusariospora, Leptorhaphis acerina, Leptorhaphis quercus f. macrospora, Leptorhaphis pinicola, Leptorhaphis wienkampii, Liberomyces pistaciae, Sphaeria muelleri and Zignoëlla slaptonensis are combined in Leptosillia, and all of these taxa except for C. fusariospora, L. pinicola and L. pistaciae are epitypified. Coelosphaeria fusariospora and Cresporhaphis rhoina are lectotypified. Liberomyces macrosporus and L. saliciphilus, which were isolated as phloem and sapwood endophytes, are shown to be synonyms of Leptosillia macrospora and L. wienkampii, respectively. All species formerly placed in Cresporhaphis that are now transferred to Leptosillia are revealed to be non-lichenized. Based on morphology and ecology, Cresporhaphis chibaensis is synonymised with Rhaphidicyrtis trichosporella, and C. rhoina is considered to be unrelated to the genus Leptosillia, but its generic affinities cannot be resolved in lack of DNA sequence data. Phylogenetic analyses place Leptosillia as sister taxon to Delonicicolaceae, and based on morphological and ecological differences, the new family Leptosilliaceae is established. Furfurella, a new genus with the three new species, F. luteostiolata, F. nigrescens and F. stromatica, growing on dead branches of mediterranean fabaceous shrubs from tribe Genisteae, is revealed to be the closest relative of Delonicicola in the family Delonicicolaceae, which is emended. ITS rDNA sequence data retrieved from GenBank demonstrate that the Leptosilliaceae were frequently isolated or sequenced as endophytes from temperate to tropical regions, and show that the genus Leptosillia represents a widely distributed component of endophyte communities of woody plants.

Keywords: Ascomycota, Diaporthales, eight new combinations, five new taxa, phylogenetic analysis, pyrenomycetes, Sordariomycetes

INTRODUCTION

The monotypic genus Leptosillia, based on L. notha, was posthumously described by Höhnel (1928) in a manuscript edited by J. Weese, with Harpostroma notha as its asexual morph. As the genus name suggests, Leptosillia was considered to be closely related to the diaporthalean genus Sillia. Oddly enough, it was, however, classified in Botryosphaeriaceae (‘Melanopsoideae’), which was probably added by J. Weese. Since its original description, Leptosillia notha has apparently never been recorded again, although it is growing on bark of Acer pseudoplatanus, which is a common and widespread tree in many parts of Europe. Due to the vague original description and the lack of illustrations, its systematic placement could so far not be critically evaluated, and the few references in the literature made it even more mysterious. Hawksworth (in Eriksson & Hawksworth 1987) noted that the type of Leptosillia was based on a specimen of Cryptospora (= Sillia) cinctula distributed by Rehm (Ascomyceten, no. 2047; Rehm 1913), and after studying a slide of the type at FH, the fungus was tentatively referred to Valsaceae. However, it is unclear how Hawksworth came to that conclusion, as the original description of L. notha was based on a German collection made by H. Diedicke, and neither in the original description nor on the labels of the type collection, neither Cryptospora (= Sillia) cinctula nor Rehm’s Ascomyceten are mentioned. This misapplication was perpetuated in the latest edition of the Dictionary of the Fungi (Kirk et al. 2008), and Leptosillia is currently placed in Valsaceae in Index Fungorum (http://www.indexfungorum.org/Names/Names.asp; accessed in Feb. 2019).

In the course of an ongoing research project on phylogenetics of Diaporthales, the first author successfully recollected Leptosillia notha to clarify its systematic affiliation by morphology and DNA sequence data. We also collected, cultured and sequenced a small pyrenomycete from the corky bark strips of Ulmus minor, which we identified as Cresporhaphis ulmi (Calatayud & Aguirre-Hudson 2001). To our surprise, the ITS-LSU rDNA sequences of Leptosillia notha and Cresporhaphis ulmi turned out to be highly similar, raising the question whether both are congeneric. Nucleotide BLAST searches of the ITS also revealed a high similarity to Liberomyces, an endophytic coelomycetous asexual morph genus of xylarialean affinities that was isolated from the inner bark and sapwood of Salix and Ulmus species (Pažoutová et al. 2012). In addition, we collected several specimens of a pyrenomycete with a yellow scurf and valsa-like ascospores on dead branches of fabaceous mediterranean shrubs, which could not be identified but later turned out to be closely related to the isolates mentioned above as well. The monotypic genus Delonicicola, which was recently described from seed pods of Delonix regia in Thailand (Perera et al. 2017), also showed high sequence similarities to our isolates. This prompted us to recollect several other Cresporhaphis species. These were isolated in pure culture; the morphology of their sexual and asexual morphs was studied and their ecology was investigated to ascertain if these are truly lichenised as previously postulated. In addition, multi-gene analyses were performed with a matrix of SSU-ITS-LSU, rpb1, rpb2, tef1 and tub2 sequences to reveal their phylogenetic affiliation, to clarify genus, species and family boundaries and to settle their taxonomy in a polyphasic approach.

MATERIALS AND METHODS

Sample sources

All isolates included in this study originated from ascospores of freshly collected specimens on bark of living or recently dead branches or trunks; typical habitats of Leptosillia species are illustrated in Fig. 1. Details of the strains including NCBI GenBank accession numbers of gene sequences used to compute the phylogenetic trees are listed in Table 1. Strain acronyms other than those of official culture collections are used here primarily as strain identifiers throughout the work. Representative isolates have been deposited at the Westerdijk Fungal Biodiversity Centre (CBS-KNAW), Utrecht, The Netherlands. Details of the specimens used for morphological investigations are listed in the Taxonomy section under the respective descriptions. Herbarium acronyms are according to Thiers (2018), and citation of exsiccatae follows Triebel & Scholz (2018). Freshly collected specimens have been deposited in the Fungarium of the Department of Botany and Biodiversity Research, University of Vienna (WU).

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g001.jpg

Typical habitats of the Leptosillia species sampled; arrows denoting ascomata on cork wings (b, j), bark furrows (d, h) or bark scales (f). a–b. Leptosillia acerina on branches of Acer campestre; c–d. Leptosillia macrospora on bark of living trunks of Quercus robur; e–f. Leptosillia muelleri on bark of living trunks of Acer pseudoplatanus; g–h. Leptosillia wienkampii on bark of living trunks of Salix sp.; i–j. Leptosillia slaptonensis on branches of Ulmus minor.

Morphology

Microscopic observations were made in tap water except where noted. Methods of microscopy included stereomicroscopy using a Nikon SMZ 1500 equipped with a Nikon DS-U2 digital camera or a Keyence VHX-6000 system, and Nomarski differential interference contrast (DIC) using a Zeiss Axio Imager.A1 compound microscope equipped with a Zeiss Axiocam 506 colour digital camera. Images and data were gathered using the NIS-Elements D v. 3.22.15 or Zeiss ZEN Blue Edition software. For certain images of ascomata the stacking software Zerene Stacker v. 1.04 (Zerene Systems LLC, Richland, WA, USA) was used. Measurements are reported as maxima and minima in parentheses and the range representing the mean plus and minus the standard deviation of a number of measurements given in parentheses.

Culture preparation, DNA extraction, PCR and sequencing

Ascospore isolates were prepared and grown on 2 % corn meal dextrose agar (CMD; CMA: Sigma, St Louis, Missouri; supplemented with 2 % (w/v) D(+)-glucosemonohydrate) or 2 % malt extract agar (MEA; 2 % w/v malt extract, 2 % w/v agar-agar; Merck, Darmstadt, Germany). Growth of liquid cultures and extraction of genomic DNA was performed as reported previously (Voglmayr & Jaklitsch 2011, Jaklitsch et al. 2012) using the DNeasy Plant Mini Kit (QIAgen GmbH, Hilden, Germany).

The following loci were amplified and sequenced: the complete internal transcribed spacer region (ITS1-5.8S-ITS2) and a c. 900–1200 bp fragment of the large subunit nuclear ribosomal DNA (nuLSU rDNA), amplified and sequenced as a single fragment with primers V9G (De Hoog & Gerrits van den Ende 1998) and LR5 (Vilgalys & Hester 1990); a c. 1.2 kb fragment of the RNA polymerase II subunit 1 (rpb1) gene with primers RPB1-Af (Stiller & Hall 1997) and RPB1-6R1asc (Hofstetter et al. 2007); a c. 1.2 kb fragment of the RNA polymerase II subunit 2 (rpb2) gene with primers fRPB2-5f and fRPB2-7cr (Liu et al. 1999) or dRPB2-5f and dRPB2-7r (Voglmayr et al. 2016a); a c. 1.3–1.5 kb fragment of the translation elongation factor 1-alpha (tef1) gene with primers EF1-728F (Carbone & Kohn 1999) and TEF1LLErev (Jaklitsch et al. 2005) or EF1-2218R (Rehner & Buckley 2005); and a c. 1.6 kb fragment of the beta tubulin (tub2) gene with primers T1 and T22 (O’Donnell & Cigelnik 1997) or T1D and T22D (Voglmayr et al. 2019). PCR products were purified using an enzymatic PCR cleanup (Werle et al. 1994) as described in Voglmayr & Jaklitsch (2008). DNA was cycle-sequenced using the ABI PRISM Big Dye Terminator Cycle Sequencing Ready Reaction Kit v. 3.1 (Applied Biosystems, Warrington, UK) and the PCR primers; in addition, primers ITS4 (White et al. 1990), LR2R-A (Voglmayr et al. 2012) and LR3 (Vilgalys & Hester 1990) were used for the ITS-LSU region, TEF1_INTF (Jaklitsch 2009) and TEFD_iR (Voglmayr et al. 2018) for tef1, and BtHVf (Voglmayr & Mehrabi 2018) and BtHV2r (Voglmayr et al. 2016b) for tub2. Sequencing was performed on an automated DNA sequencer (3730xl Genetic Analyzer, Applied Biosystems).

Data analysis

Following the results of nucleotide BLAST searches of ITS and LSU sequences generated during the present study, a phylogenetic analysis was performed with an ITS-LSU rDNA sequence matrix of a representative selection of Xylariales. Taxon and sequence selection was based on Jaklitsch et al. (2016b), with some recent additions (Perera et al. 2017, Voglmayr et al. 2018, Wendt et al. 2018). For rooting the tree, LSU sequences of four taxa of Sordariomycetes (Calosphaeria pulchella, Chaetosphaeria innumera, Diaporthe eres, Ophiostoma piliferum) were included as outgroups. For detailed investigations of species relationships and delimitation within and between the genera and families, a combined matrix of five loci (partial SSU-ITS-LSU rDNA, rpb1, rpb2, tef1 and tub2) was produced. Four taxa of Sordariomycetes (Calosphaeria pulchella, Caudospora taleola, Juglanconis juglandina, Lasiosphaeria ovina) were selected as outgroup taxa; due to alignment issues, their ITS and tef1 introns were not included in the matrix. The GenBank accession numbers of sequences used in these analyses are given in Table 1. For some strains for which whole genome data are available, sequences were retrieved from JGI-DOE (http://genome.jgi.doe.gov/).

Sequence alignments for phylogenetic analyses were produced with server versions of MAFFT (www.ebi.ac.uk/Tools/mafft or http://mafft.cbrc.jp/alignment/server/), checked and refined using BioEdit v. 7.2.6 (Hall 1999). For tef1 and ITS-LSU rDNA, the localpair and for tub2 the globalpair options were selected for performing fast Fourier transform (FFTS), with a gap open penalty of 1.0 for tef1 and tub2; for all other markers, the default settings were used. Poorly aligned and gappy regions were removed from the ITS and the introns of tef1 and tub2, and the terminal intron of the rpb2 was entirely removed. The final ITS-LSU matrix used for phylogenetic analyses contained 1 345 and the combined five loci data matrix 7 052 nucleotide characters; viz. 1 626 of SSU-ITS-LSU, 1 210 of rpb1, 1 104 of rpb2, 1 516 of tef1 and 1 596 of tub2. Prior to phylogenetic analyses, the approach of Wiens (1998) was applied to test for significant levels of localised incongruence among the markers used for the combined analysis, using the level of bootstrap support (Sung et al. 2007) as described in Jaklitsch & Voglmayr (2014). For this, the 70 % maximum parsimony (MP) bootstrap consensus trees calculated for each individual partition, using the same parameters as given below, were compared. Except for some nodes within the same species, no topological conflicts were observed between these bootstrap trees of the various genes, indicating the absence of significant incongruence and combinability of the five loci (Wiens 1998).

Maximum likelihood (ML) analyses were performed with RAxML (Stamatakis 2006) as implemented in raxmlGUI v. 1.5 (Silvestro & Michalak 2012), using the ML + rapid bootstrap setting and the GTRGAMMA substitution model with 1 000 bootstrap replicates. The matrix was partitioned for the different gene regions included in the combined multilocus analyses.

Maximum parsimony (MP) analyses were performed with PAUP v. 4.0a163 (Swofford 2002). All molecular characters were unordered and given equal weight; analyses were performed with gaps treated as missing data; the COLLAPSE command was set to MINBRLEN. For the ITS-LSU matrix, first a parsimony ratchet approach was used. For this, nexus files were prepared using PRAP v. 2.0b3 (Müller 2004), implementing 1 000 ratchet replicates with 25 % of randomly chosen positions upweighted to 2, which were then run with PAUP. In a second step, the best trees obtained by the parsimony ratchet analyses were loaded in PAUP and subjected to heuristic search using TBR branch swapping (MULTREES option in effect, steepest descent option not in effect). MP analysis of the combined multilocus matrix was done using 1 000 replicates of heuristic search with random addition of sequences and subsequent TBR branch swapping (MULTREES option in effect, steepest descent option not in effect). Bootstrap analyses with 1 000 replicates were performed with the same settings, but using 5 rounds of random sequence addition and subsequent branch swapping during each bootstrap replicate; in addition, each replicate was limited to 1 million rearrangements in the ITS-LSU matrix.

RESULTS

Molecular phylogeny

Of the 1 345 characters included in the ITS-LSU analyses, 516 were parsimony informative. The best ML tree (lnL = -19 991.4865) revealed by RAxML is shown as Fig. 2. MP analyses revealed 4 598 MP trees 4 041 steps long (not shown). Most of the tree backbone was identical in all MP trees; differences were mainly present within the clade containing the Amphisphae riaceae, Apiosporaceae, Beltraniaceae, Melogrammataceae, Phlogicylindriaceae, Pseudomassariaceae, Sporocadaceae and Vialaeaceae (AABMPPSV clade; not shown). The MP strict consensus tree was mostly compatible with the ML tree; notable exceptions were a placement of the Coniocessiaceae-Microdochiaceae clade basal to the AABMPPSV clade; an interchanged position of the Hypoxylaceae with the Barrmaeliaceae-Graphostromataceae clade, and within the Leptosillia clade, a position of the Calocedrus macrolepis endophyte as sister to Leptosillia macrospora and of the Nothofagus fusca endophyte as basal to the clade containing, amongst various endophyte accessions, Leptosillia pistaciae, L. macrospora, L. slaptonensis and L. wienkampii (not shown). The clade containing Delonicicola, Furfurella, Leptosillia and numerous unclassified endophytes received high support in both analyses (96 % ML, 97 % MP), and the clade containing Delonicicola, Furfurella gen. nov. and the Phoradendron endophyte medium (89 % ML) and high (96 % MP) support. The Leptosillia clade, however, was resolved as monophyletic only in the ML analyses, where it received moderate support (78 %); besides the six Leptosillia species, this clade contained numerous ITS sequence accessions of endophytes from various geographic areas and hosts, which were scattered throughout the clade (Fig. 2). In the strict consensus of the MP trees, three subclades were placed in a polytomy:

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g002.jpg

Phylogram of the ML tree (lnL = -19 991.4865) revealed by RAxML from an analysis of the ITS-LSU rDNA matrix of selected Xylariales, showing the phylogenetic position of Furfurella and Leptosillia. Strain/culture numbers or GenBank accession numbers are given following the taxon names; for the endophyte isolates, the host is given in brackets. ML and MP bootstrap support above 50 % are given at the first and second position, respectively, above or below the branches. Accessions in bold were isolated and sequenced in the present study; those in green were generated in endophyte studies, those in red represent plant pathogens, and those in blue were isolated from ascomata growing on dead plant tissues (bark, wood, seed pods).

  • i. the Delonicicola clade;

  • ii. a highly supported Leptosillia acerina-L. muelleri clade (including various endophyte isolates); and

  • iii. a weakly supported clade containing the residual Leptosillia species plus the rest of endophyte isolates (not shown).

Of the 7 052 characters included in the combined five locus analyses, 3 093 were parsimony informative (476 from SSU-ITS-LSU, 613 from rpb1, 579 from rpb2, 656 from tef1 and 769 from tub2). The best ML tree (lnL = -84 566.4626) revealed by RAxML is shown as Fig. 3. The MP analysis revealed 6 MP trees 19 319 steps long (not shown); tree topologies of all MP trees were identical except for slightly different positions of Calceomyces lacunosus. Tree topologies of the MP trees were similar to the ML tree, except for a sister group relationship of Diatrypaceae and Lopadostomataceae, a basal position of Requienella to the other Xylariaceae s.lat., a sister group relationship of Graphostromataceae to Xylariaceae s.str., and an interchanged position of Microdochium and Calceomyces in some of the MP trees (not shown). In both analyses, the clade containing Delonicicola, Furfurella and Leptosillia and the Delonicicola-Furfurella subclade received maximum support (Fig. 3), while the Leptosillia subclade was highly supported (99 % ML, 93 % MP). Given the marked morphological differences (see below) and the highly supported phylogenetic subdivision in the multigene analyses, the new family Leptosilliaceae is established for the genus Leptosillia.

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g003.jpg

Phylogram of the ML tree (lnL = -84 566.4626) revealed by RAxML from an analysis of the combined SSU-ITS-LSU-rpb1-rpb2-tef1-tub2 matrix of selected Xylariales, showing the phylogenetic position of Furfurella and Leptosillia. ML and MP bootstrap support above 50 % are given at the first and second position, respectively, above or below the branches. Strain/culture numbers are given following the taxon names; accessions in bold were isolated and sequenced in the present study.

Culture characteristics

Culture images of two Furfurella and five Leptosillia species grown on CMD are shown in Fig. 4. Detailed culture descriptions are given under the respective species.

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g004.jpg

Cultures on CMD at 15–17 °C. a. Furfurella nigrescens (CE); b. Furfurella stromatica (CE4); c–e. Leptosillia acerina (c, e: CRA1, d: CRA); f–h. Leptosillia macrospora (f: CRM1, g: CRM4, h: CRM2); i–k. Leptosillia muelleri (i, k: CRM3, j: CRM6); l–m. Leptosillia slaptonensis (l: CRU1, m: CRU2); n–p. Leptosillia wienkampii (n: CRW, o: CRW1, p: CRU). a, c, e–g, i, k–l, n–o. Surface view; b, d, h, j, m, p. Reverse. All after 58 d; except b after 27 d and e, k after 7.5 mo.

TAXONOMY

Delonicicolaceae R.H. Perera et al., emend. Voglmayr & Jaklitsch

Type genus. Delonicicola R.H. Perera et al., Cryptog. Mycol. 38: 334. 2017.

Family of Xylariales. Pseudostromata variable, from conspicuously pulvinate to virtually absent, immersed in host tissue, erumpent to rarely superficial, variously coloured, ranging from yellowish, brown to black; visible as raised, dark spots on the host surface, as black, more or less elevated patches on wood or erumpent through bark, occasionally covered by bright turquoise, yellow to yellow-green scurf. Ascomata perithecial, immersed in pseudostroma, aggregated, globose, subglobose to conical or irregular, subhyaline to pale brown, with an apical ostiole. Peridium subhyaline to medium brown, KOH-, of textura angularis to prismatica. Ostioles papillate. Hamathecium composed of hyaline, septate or aseptate, unbranched or occasionally branched paraphyses. Asci arising from the base or margins of the ascomata, clavate to cylindrical, straight, curved to sinuous, thin-walled, containing 8 biseriately arranged ascospores, inamyloid and without a distinct apical apparatus. Ascospores ellipsoid or allantoid, equilateral or inequilateral, aseptate or septate, not constricted at the septa, hyaline, thin-walled, smooth, with rounded apices, without appendages or gelatinous sheath. Asexual morph unknown.

Notes — We provide an emended familial description here, as the Delonicicolaceae in the sense of the original authors also include the taxa here segregated in a family of their own, Leptosilliaceae (see below), and the new genus Furfurella with allantoid ascospores and variously developed, immersed, erumpent or superficial pseudostromata usually covered by a bright greenish yellow scurf, is here added to Delonicicolaceae. The illustrations of the type species, Delonicicola siamense, in Perera et al. (2017), unfortunately do not allow for evaluation of stromatic configuration and ascoma morphology in detail. We also propose that other morphological features should be re-checked (e.g., the authors surprisingly reported that the paraphyses were lacking septa!).

Furfurella Voglmayr & Jaklitsch, gen. nov. — MycoBank MB829925

Etymology. Referring to the bright greenish to yellow scurf on its stromata.

Type species. Furfurella stromatica Voglmayr & Jaklitsch.

Pseudostromata variously developed, from pulvinate to virtually absent, erumpent through bark cracks or embedded in bark or wood, commonly blackening the substrate surface, usually covered by a bright yellow, yellow-green to turquoise scurf dissolving a bright yellow pigment in KOH. Ascomata perithecial, 120–460 μm diam, immersed in pseudostroma, usually densely aggregated in groups of 2–25, rarely scattered singly, lenticular, subglobose to pyriform, horizontally compressed when dry, with a central apical ostiole, perithecial content dull orange to brown and waxy when dry. Peridium light to medium brown, KOH-, becoming hyaline towards the centrum, pseudoparenchymatous to prosenchymatous, consisting of thin- or thick-walled, hyaline to brown, isodiametric to elongated cells forming a textura angularis or prismatica. Ostioles variously developed, from inconspicuous and not protruding to long cylindrical and protruding; ostiolar canal with c. 1 μm wide hyaline periphyses embedded in a gelatinous matrix. Hamathecium composed of elongate, hyaline, septate, occasionally branched, basally broad and apically tapering paraphyses. Asci arising from the base and the margins of the ascomata, sequentially produced; fusoid, clavate to cylindrical, straight, curved or sinuous, thin-walled, with marginal fissurate dehiscence, containing 8 biseriately or fasciculately arranged ascospores, without a stipe and an apical apparatus, inamyloid but appearing bitunicate with a distinct ocular chamber in Lugol after treatment with 3 % KOH. Ascospores allantoid, aseptate, hyaline, thin-walled, smooth, with rounded apices, without appendages or gelatinous sheath. Asexual morph unknown.

Notes — Furfurella can be easily discriminated from its closest relative, Delonicicola, by its large, allantoid, aseptate ascospores, a bright yellow, yellow-green to turquoise scurf on the stromata and ostioles, a medium brown ascoma wall, and by growth on dead branches of mediterranean fabaceous shrubs from tribe Genisteae.

In all species, the ascospore contours are only faintly seen in asci mounted in water, but become distinct in KOH and Lugol. Ascospores and asci shrink considerably in Lugol, therefore measurements were done in water to ensure comparability of the data.

Furfurella luteostiolata Voglmayr & Jaklitsch, sp. nov. — MycoBank MB829926; Fig. 5

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g005.jpg

Furfurella luteostiolata (WU 39989, holotype). a–c. Ostioles erumpent through bark with sulphur yellow scurf in surface or side view; d. pseudostroma with two perithecia in vertical section; e–f. vertical sections of perithecial walls (e. from upper part of two adjacent perithecia, f. lateral); g. vertical section of pseudostroma around ostioles, yellow brown colour originating from yellow scurf dissolved in KOH; h–j. asci; k. ascus apex; l. septate paraphysis; m–x. ascospores. All in water, except e–g in 3 % KOH, j–l in Lugol after KOH pre-treatment. — Scale bars: a–d = 200 μm; e–x = 10 μm.

Etymology. Referring to the yellow scurf around its ostioles.

Holotype. GREECE, Crete, Chania, Omalos, 920 m a.s.l., N35.37° E23.897°, in bark of thin dead branches of Genista acanthoclada, soc. Microthyrium sp., Diaporthe sp., 5 June 2015, W. Jaklitsch & H. Voglmayr (WU 39989; ex-holotype culture CBS 143620 = CE3).

Pseudostromata immersed in the woody substrate and erumpent through the bark, reduced mostly to the region around the apical parts of the ascomata and covered by a bright sulphur yellow scurf, slightly blackening the bark surface around the erumpent stromata. Ascomata perithecial, c. 200–250 μm diam, embedded in bark or wood, solitary or in groups of up to 5, irregularly subglobose to pyriform, horizontally compressed when dry, with a central apical ostiole; perithecial contents dull brown, waxy when dry. Peridium 16–26 μm thick, brown, KOH-, becoming hyaline towards the centrum, pseudoparenchymatous to prosenchymatous, consisting of rather thick-walled, brown, isodiametric to elongated cells 3–16 × 2–5 μm forming a textura angularis, thin-walled and hyaline towards the centrum. Ostioles c. 110–130 μm long, 30–60 μm wide, not protruding above the stroma surface, apically black, surrounded by stromatic tissues covered by sulphur yellow scurf. Hamathecium composed of elongate, hyaline, septate, occasionally branched paraphyses up to 6 μm wide at the base, gradually tapering to 1.7 μm towards the distal ends. Asci (83–)88–107(–115) × (12.0–)12.8–14.7(–15.3) μm (n = 20), fusoid to cylindrical, straight or slightly curved, thin-walled, with fissurate dehiscence, containing 8 biseriately arranged ascospores, without a stipe and an apical apparatus, inamyloid. Ascospores (24–)27–32(–34) × (5.5–)6.5–7.5(–8.2) μm, l/w = (3.1–)3.8–4.8(–5.3) (n = 75), allantoid, aseptate, hyaline, thin-walled, smooth, with rounded apices, without appendages or gelatinous sheath. Asexual morph unknown.

Culture characteristics — On CMD colony radius 32 mm after 23 d at 22 °C. Colony whitish, very dense, turning cream with age, with abundant white aerial mycelium in the centre.

Habitat & Host range — Only known from corticated dead branches of Genista acanthoclada.

Distribution — Only known from the type collection in Crete (Greece).

Notes — Furfurella luteostiolata differs from the other two known Furfurella species by its broader and stouter ascospores and by the bright sulphur yellow scurf around the ostioles.

Furfurella nigrescens Voglmayr & Jaklitsch, sp. nov. — MycoBank MB829927; Fig. 6

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g006.jpg

Furfurella nigrescens. a–d. Black pseudostromata in wood or bark (a, b. with yellow-green scurf; c. clypeus-like black discoloration of bark); e, h. transverse section through pseudostromata and perithecia with dull orange, waxy perithecial contents; f. bright sulphur yellow scurf; g. black protruding ostioles laterally covered by yellow-green scurf; i. transverse section of perithecial wall and pseudostroma (bottom left); j. vertical section of pseudostroma embedded in bark with single perithecium and clypeus-like black discoloration of bark surface; k. section of peridium and pseudostroma (bottom); l. septate paraphyses; m–o. asci; p–x. ascospores. All in water, except l, o in Lugol after KOH pre-treatment, n in 3 % KOH (a–b, e–g, m–n, p–r: WU 39990 (holotype); c, h–l, o, s–v: WU 39992; d, w, x: WU 39991). — Scale bars: a–c, e = 300 μm; d, f = 500 μm; g–j = 100 μm; k–x = 10 μm.

Etymology. Referring to the blackening of the host surface around the pseudostromata.

Holotype. SPAIN, Andalucía, at km 26 between La Sauceda and Puerto Galiz, 500 m a.s.l., in bark and wood of dead branches of Calicotome villosa, soc. Valsaria spartii, 2 Apr. 2014, W. Jaklitsch (WU 39990; ex-holotype culture CBS 143622 = CE1).

Pseudostromata embedded in bark or wood, reduced to substrate blackening and scurf, sometimes slightly elevating the substrate, distinctly blackening the host surface and commonly covered by a bright yellow, yellowish green to turquoise scurf. Ascomata perithecial, 130–420 μm diam, c. 230–270 μm high, embedded in bark or wood, solitary or aggregated in groups, lenticular, subglobose to pyriform, horizontally compressed when dry, with a central apical ostiole. Peridium 11–22 μm thick, light brown, KOH-, becoming hyaline towards the centrum, pseudoparenchymatous to prosenchymatous, consisting of thin-walled, light brown, isodiametric to elongated cells 5–11 × 1–3.5 μm forming a textura angularis to prismatica, becoming hyaline towards the centrum; perithecial contents dull orange, waxy when dry. Ostioles either flat, non-protruding, or distinctly cylindrical to conical and projecting up to 200 μm, 80–100 μm wide, black, of thick-walled, dark brown cells with narrow lumen; when protruding ostiole laterally covered by a sulphur yellow to yellowish green scurf. Hamathecium composed of elongate, hyaline, septate, occasionally branched paraphyses up to 5.5 μm wide at the base, gradually tapering to 1.3 μm towards the distal ends. Asci (64–)72–89(–99) × (10.5–)11.5–13.2(–14.5) μm (n = 35), fusoid, clavate to cylindrical, straight or slightly curved, thin-walled, with fissurate dehiscence, containing 8 ascospores biseriately arranged or in two fascicles, without a stipe and an apical apparatus, inamyloid. Ascospores variable in length, (18–)20–29(–35) × (4.0–)4.5–5.2(–6.0) μm, l/w = (3.7–)4.2–5.9(–7.2) (n = 157), allantoid, aseptate, hyaline, thin-walled, smooth, with rounded apices, without appendages or gelatinous sheath. Asexual morph unknown.

Culture characteristics — On CMD colony radius up to 34 mm after 29 d at 22 °C. Colony whitish, soon cream, very dense, with abundant bright yellow mycelium in the centre; odour fruity to yeast-like.

Habitat & Host range — On dead branches of Calicotome villosa and Chamaecytisus creticus.

Distribution — Mediterranean; known from Spain and Greece (Crete).

Additional specimens examined. GREECE, Crete, Chania, SW Lakki, 580 m a.s.l., N35.392° E23.928°, in wood of thin decorticated branches of Chamaecytisus creticus, 5 June 2015, W. Jaklitsch & H. Voglmayr (WU 39991; culture CBS 143621 = CE2). – SPAIN, Andalucía, near Puerto Galiz, 450 m a.s.l., in bark of thin dead branches of Calicotome villosa, 2 Apr. 2014, W. Jaklitsch (WU 39992; culture CE).

Notes — Compared to the other two species of the genus, Furfurella nigrescens is more inconspicuous as its scurf is less prominent and sometimes even entirely absent. However, it is distinctly blackening the host surface, ranging from circular and clypeus-like around single ascomata in bark to extensive irregular patches around aggregated ascomata embedded in wood. In addition, its cultures develop a bright yellow aerial mycelium on CMD (Fig. 4a).

Furfurella stromatica Voglmayr & Jaklitsch, sp. nov. — MycoBank MB829928; Fig. 7

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g007.jpg

Furfurella stromatica. a–e. Pseudostromata covered by bright yellow-green scurf and erumpent through bark (a–d) or superficial on wood (e); f. transverse section through pseudostroma and perithecia; g. side view of pulvinate stroma on wood, with protruding ostioles; h. vertical section of pseudostroma and perithecium with orange, waxy perithecial content; i. vertical section of pseudostroma and perithecia, with yellow scurf dissolving in KOH; j. black erumpent ostioles laterally covered by bright yellow-green scurf; k. section of perithecial wall and pseudostroma (lower half); l–m. asci; n–y. ascospores. All in water, except i, l in 3 % KOH, m in Lugol after KOH pre-treatment (a–d, f, i–j, l, t–y: WU 39993 (holotype); e, g–h, k, m–s: WU 39994). — Scale bars: a–f, i–j = 300 μm; g–h 100 μm; k–y = 10 μm.

Etymology. Referring to the well-developed pseudostromata.

Holotype. SPAIN, Andalucía, Jaén, Valdepeñas de Jaén, El Parrizoso, 1025 m a.s.l., N37°36’50.26" W3°43’12.34", on dead corticated branch of Genista cinerea, 29 Feb. 2016, S. Tello S.T.29021601 (WU 39993; ex-holotype culture CBS 144409 = CE4).

Pseudostromata conspicuous, 0.25–2.1 mm long, 0.15–1.2 mm wide, pulvinate, superficial on wood or erumpent through bark cracks, exterior black and covered by a bright sulphur yellow, yellowish green to turquoise scurf, interior light brown. Ascomata perithecial, 240–460 μm diam, c. 250–280 μm high, embedded in a pseudostroma, gregarious in groups up to 25, subglobose, globose to pyriform, horizontally compressed when dry, with a central apical ostiole; perithecial content dull orange, waxy when dry. Peridium 21–29 μm thick, light brown, KOH-, becoming hyaline towards the centrum, pseudoparenchymatous to prosenchymatous, consisting of thin-walled, light brown, isodiametric to elongated cells 2–11.5 × 1–3.5 μm forming a textura angularis, becoming hyaline towards the centrum. Ostioles cylindrical to conical, protruding above stromata up to 250 μm, 80–160 μm wide, black, laterally covered by a sulphur yellow to yellowish green scurf. Hamathecium composed of elongate, hyaline, septate, occasionally branched paraphyses up to 5 μm wide at the base, gradually tapering to 1.7 μm towards the distal ends, deliquescent at maturity. Asci (78–)89–122(–139) × (10.7–)11.3–13.5(–14.5) μm (n = 28), clavate to cylindrical, usually slightly curved, thin-walled, with fissurate dehiscence, containing 8 biseriately arranged ascospores, without a stipe and an apical apparatus, inamyloid, easily detached at maturity. Ascospores variable in length, (23–)29–38(–47) × (3.7–)4.7–5.5(–6.5) μm, l/w = (5.1–)5.7–7.1(–8.1) (n = 103), allantoid, aseptate, hyaline, thin-walled, smooth, with rounded apices, without appendages or gelatinous sheath. Asexual morph unknown.

Culture characteristics — On CMD colony radius 40 mm after 29 d at 22 °C, covering almost the entire plate. Colony whitish, dense, thin, becoming yellowish brown to brown from the centre, with white aerial mycelium in the centre; odour sweetish.

Habitat & Host range — On dead branches of Genista cinerea.

Distribution — Only known from southern Spain (Andalucía).

Additional specimen examined. SPAIN, Andalucía, Jaén, Valdepeñas de Jaén, Cañón de Pitillos, 790 m a.s.l., N37°37’5.22" W3°41’32.14", on dead decorticated branch of Genista cinerea, 15 Mar. 2018, S. Tello S.T.15031803 (WU 39994; culture CE5).

Notes — Furfurella stromatica is well distinct from the other known species of the genus by its conspicuous elongate pulvinate pseudostromata containing up to 25 perithecia with distinctly protruding black ostioles. This overall appearance, and in particular the fact that the asci become easily detached at maturity, the deliquescent paraphyses and the allantoid hyaline aseptate ascospores, point towards a placement of this fungus in the Diaporthales. Similar cases of misleading morphological evidence for taxa phylogenetically recently reclassified in Xylariales include e.g., Melogramma (previously classified in Diaporthales; Jaklitsch & Voglmayr 2012), Acrocordiella and Requienella (previously classified in Pyrenulales; Jaklitsch et al. 2016b) and Strickeria (previously classified in Dothideomycetes; Jaklitsch et al. 2016b). Of all three Furfurella species, F. stromatica has the most conspicuous bright yellow to yellowish green scurf.

Key to species of Furfurella

  • 1. Pseudostromata conspicuous, erumpent to superficial, pulvinate, exterior black and covered by a bright sulphur yellow, yellowish green to turquoise scurf . . . . . . . . . .F. stromatica

  • 1. Pseudostromata inconspicuous, reduced to virtually absent, mostly in host tissue . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2

  • 2. Pseudostromata concentrated around the erumpent ostioles, at margins covered by bright sulphur yellow scurf, ascospores (5.5–)6.5–7.5(–8.2) μm wide . . . . . F. luteostiolata

  • 2. Pseudostromata embedded in substrate, not to slightly elevating but blackening the substrate surface, ascospores (3.7–) 4.7–5.5(–6.5) μm wide . . . . . . . . . . . . . . . . . F. nigrescens

Leptosilliaceae Voglmayr & Jaklitsch, fam. nov. — MycoBank MB829929

Etymology. Referring to the name of the type genus.

Type genus. Leptosillia Höhn.

Family of Xylariales. Ascomata perithecial, superficial to partly immersed in bark, scattered, gregarious or confluent, black, sometimes collapsed, with a central apical ostiolar papilla. Peridium melanized, KOH-, of textura angularis or prismatica. Ostioles papillate, sometimes sulcate, base of the ostiolar canal sometimes with hyaline periphyses. Hamathecium composed of hyaline, septate, occasionally branched paraphyses embedded in a gelatinous matrix. Asci arising from the base of the ascomata, sequentially produced; clavate to cylindrical, curved to sinuous, thin-walled, containing 8 bi-, triseriately or fasciculately arranged ascospores, inamyloid and without a distinct apical apparatus. Ascospores ranging in shape from nearly straight, falcate, lunate, sinuous, sigmoid to hook-shaped, aseptate or septate, not constricted at the septa, hyaline, thin-walled, smooth, with rounded to subacute apices, without appendages or gelatinous sheath.

Conidiomata pycnidial, superficial to partly immersed in bark, globose to pyriform, black, scattered, aggregated or confluent, uni- or irregularly plurilocular. Peridium more or less melanized, of textura globulosa to angularis. Conidiophores short, hyaline, arising from the inner layer of the peridium. Conidiogenous cells cylindrical to lageniform. Conidiogenesis either enteroblastic-phialidic or holoblastic with sympodial proliferation, both types sometimes found within the same conidioma. Conidia commonly of two types according to their formation, allantoid, falcate or filiform, aseptate, hyaline, thin-walled.

Notes — Leptosilliaceae is closely related to Delonicicolaceae, from which it differs significantly by semi-immersed to superficial, black ascomata and, when present (L. muelleri), by different stroma structure.

Leptosillia Höhn., Mitt. Bot. Inst. Tech. Hochsch. Wien 5: 111. 1928

Synonyms. Cresporhaphis M.B. Aguirre, Bull. Brit. Mus. (Nat. Hist.), Bot. 21: 146. 1991.

Harpostroma Höhn., Mitt. Bot. Inst. Tech. Hochsch. Wien 5: 112. 1928.

Liberomyces Pažoutová et al., Mycologia 104: 201. 2012.

Type species. Leptosillia notha Höhn., a synonym of L. muelleri (Duby) Voglmayr & Jaklitsch.

Ascomata perithecial, 100–400 μm diam, superficial to partly immersed in bark, scattered singly, gregarious or confluent, black, smooth, sometimes collapsed, (sub)globose to pyriform, with a central apical ostiole. Peridium melanized, KOH-, becoming subhyaline towards the centrum, pseudoparenchymatous to prosenchymatous, consisting of thick-walled, dark brown, isodiametric to elongated cells forming a textura angularis or prismatica. Ostioles papillate, sometimes sulcate; base of the ostiolar canal sometimes with hyaline periphyses. Hamathecium composed of elongate, hyaline, filiform, septate, occasionally branched paraphyses embedded in an inamyloid gelatinous matrix; in some species with hyaline, refractive, dextrinoid granular exudates turning amber-red in Lugol. Asci arising from the base of the ascomata, sequentially produced; clavate to cylindrical, curved to sinuous, thin-walled, containing 8 bi-, triseriately or fasciculately arranged ascospores, with a short stipe, without a distinct apical apparatus, inamyloid but sometimes a narrow, short pore visible in Lugol. Ascospores from nearly straight, hooked, falcate, lunate, sinuous to sigmoid, aseptate or up to 11-septate, not constricted at the septa, hyaline, thin-walled, smooth, with rounded to subacute apices, without appendages or gelatinous sheath.

Conidiomata pycnidial, superficial to partly immersed in bark, globose to pyriform, black, smooth, scattered, aggregated or confluent, uni- or irregularly plurilocular. Peridium light to dark brown, continuous, composed of thin-walled, more or less iso-diametric cells, forming a textura globulosa to angularis. Conidiophores short, hyaline, thin-walled, smooth, branched up to three times, arising from the inner layer of the peridium. Conidiogenous cells cylindrical to lageniform. Conidiogenesis either enteroblastic-phialidic and bearing usually curved, filiform, sometimes narrowly falcate conidia, or holoblastic with sympodial proliferation and bearing allantoid to falcate conidia; in some species both types of conidiogenous cells and conidia produced in the same conidioma. Conidia allantoid, falcate, lunate or filiform, aseptate, hyaline, thin-walled, smooth.

Notes — Leptosillia was posthumously described (Höhnel 1928) in a manuscript edited by J. Weese, based on a holomorphic specimen collected on bark of Acer pseudoplatanus in Germany. While Höhnel is given as the author of this publication, it is not clear which additions were provided by Weese.

The comment of Hawksworth (in Eriksson & Hawksworth 1987) that the type of Leptosillia was based on a specimen of Sillia cinctula distributed by Rehm in his Ascomyceten no. 2047 is erroneous. Rehm’s Ascomyceten no. 2047 of Cryptospora (= Sillia) cinctula represents a North American collection from Castanea dentata (Rehm 1913), which conforms to the original description of that species and has nothing to do with Leptosillia.

All Cresporhaphis species currently accepted in Index Fungorum (accessed Feb. 2019) are here combined in Leptosillia except C. chibaensis and C. rhoina; for further details see below. Although no DNA data are available for C. fusariospora and C. pinicola, their morphology and habitat support inclusion in the genus.

Leptosillia acerina (Rehm) Voglmayr & Jaklitsch, comb. nov. — MycoBank MB829930; Fig. 8

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g008.jpg

Leptosillia acerina. a–b. Perithecia on bark (right ascoma in a. laterally collapsed, in b. horizontally collapsed and cupulate); c. side view of perithecium with apical papilla; d. paraphyses; e–f. asci; g. ascus tip in Lugol; h–r. vital ascospores; s. pycnidia in culture (CMD, isolation plate, 40 d); t–u. conidiophores, conidiogenous cells and conidia from pycnidia on natural substrate; v–c1. conidia from natural substrate; d1–f1. conidia from pycnidia in culture (CMD, isolation plate, 21 d). All in water, except where noted (a–b: WU 39995 (epitype); c, g, m–r, t–c1: WU 39998; d–f, h–l: WU 39997; s: CRA; d1–f1: CRA3). — Scale bars: a–c = 100 μm; d–f, t–u = 10 μm; g–r, v–f1 = 5 μm; s = 500 μm.

Basionym. Leptorhaphis acerina Rehm, Ber. Naturhist. Vereins Augsburg 26: 51. 1881.

Synonyms. Cresporhaphis acerina (Rehm) M.B. Aguirre, Bull. Brit. Mus. (Nat. Hist.), Bot. 21 (2): 147. 1991.

Metasphaeria robergia Schulzer & Sacc., Rev. Mycol. (Toulouse) 6 (no. 22): 70. 1884.

Typification. GERMANY, Bayern, Franken, near Sugenheim, young deciduous forest, on cortex of living branches of Acer campestre, 1870, H. Rehm, Ascomyc. no. 197 (S-L1668, lectotype of Leptorhaphis acerina selected by Aguirre-Hudson 1991; K(M) 111821, W 1923-1578, W 2009-00424: isolectotypes). – AUSTRIA, Burgenland, Breitenbrunn, Thenauriegel, on cork wings of branches of Acer campestre, 23 July 2016, H. Voglmayr (WU 39995, epitype of Leptorhaphis acerina here designated (MBT 385916); ex-epitype culture CRA1 = CBS 143939).

Ascomata perithecial, immersed in bark to half of their height, (105–)140–200(–240) μm diam (n = 46), black, shiny, smooth, scattered singly, subglobose to hemispherical, circular from above, often laterally or horizontally collapsed and then cupulate, with a central apical papilla. Peridium continuous, of a textura angularis, composed of an outer dark brown, 12–24 μm thick layer of thin-walled cells 2.5–7.5 μm diam with dark brown walls, and an inner, 12–16 μm thick hyaline to pale brown layer of (sub)hyaline cells slightly smaller than those of the outer layer. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched paraphyses 2–3 μm wide, embedded in an inamyloid gelatinous matrix; periphyses not observed. Asci (57–)68–82(–89) × (6.0–)6.3–7.2(–7.5) μm (n = 34), unitunicate, cylindrical, slightly curved to sinuous, thin-walled, containing 8 bi- or triseriately arranged ascospores, with a short stipe, inamyloid but a narrow, short pore visible in Lugol. Ascospores (21–)24–30(–32) × (2.5–)2.8–3.2(–3.5) μm, l/w = (5.9–)8.2–10.1(–10.9) (n = 115), falcate, aseptate, hyaline, thin-walled, smooth, with subacute tapering ends, when vital containing numerous guttules especially towards the ends.

Pycnidia scattered on bark, black, very similar to and practically indistinguishable from ascomata except for their slightly smaller size. Conidiophores short, hyaline, smooth, branched up to three times, arising from the inner wall of the pycnidium. Conidiogenous cells (7.0–)7.8–12.8(–17.5) × (1.9–)2.1–3.2(–3.7) μm (n = 29), enteroblastic, phialidic, lageniform to cylindrical, hyaline, smooth, arranged in dense terminal whorls of up to 6. Conidia (23–)26–29(–32) × (1.9–)2.0–2.4(–2.7) μm, l/w = (10.4–)11.4–13.5(–16.0) (n = 65), falcate, aseptate, hyaline, thin-walled, smooth, with subacute tapering ends, containing few guttules close to the wall.

Culture characteristics and asexual morph in culture — Colony on CMD at 16 °C reaching 37–58 mm diam after 58 d; first white, turning cream to greyish brown in the centre, with woolly aerial mycelium mostly in the colony centre, margin uneven, wavy, reverse light brown with darker brown centre, often with radial, irregularly wavy, lighter or darker lines, with age secreting a bright yellow diffusible pigment in agar. Pycnidia (180–)230–345(–400) μm diam (n = 20), immersed to almost superficial, black, single, aggregated to confluent, opening by an ostiole or by irregular rupture and exuding white masses of conidia. Conidiophores and conidiogenous cells similar to those from natural substrate. Conidia (18–)20–25(–28) × (2.0–)2.3–2.7(–3.0) μm, l/w = (7.0–)8.1–10.2(–12.5) (n = 54), falcate to lunate, aseptate, hyaline, thin-walled, smooth, with subacute tapering ends, containing numerous guttules especially towards the ends.

Habitat & Host range — Only known from cork wings and outgrowths (the rhytidome) of living or dead branches of Acer campestre.

Distribution — Europe; known from Austria, Croatia and Germany (Aguirre-Hudson 1991, this study).

Additional specimens examined (all on cork wings of branches of Acer campestre). Austria, Niederösterreich, SE Gaaden, Am Tenneberg, 28 Jan. 2017, H. Voglmayr & I. Greilhuber (WU 39996, culture CRA3); Mödling, Richardshof, 5 Nov. 2016, H. Voglmayr & I. Greilhuber (WU 39997, culture CRA2); Pfaffstätten, near Heberlberg, 23 Apr. 2016, H. Voglmayr (WU 39998, culture CRA).

Notes — Leptosillia acerina is well characterised by its host, Acer campestre. It has so far been only found on cork wings of young living or recently dead branches, which are formed by young trees in open stands. Although the species has apparently not been recorded since the late 19th century and was only known from the type localities of the heterotypic synonyms (Aguirre-Hudson 1991), the current observations in eastern Austria indicate that, at least in Central Europe, it may be rather common in suitable habitats. This species has most likely been overlooked in mycological field studies.

Leptosillia fusariospora (Ellis & Everh.) Voglmayr & Jaklitsch, comb. nov. — MycoBank MB829931; Fig. 9

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g009.jpg

a–y. Leptosillia fusariospora (GZU 000335714, isotype). a–c. Perithecia on bark (a, c. horizontally collapsed and cupulate); d–e. side view of perithecia with apical papilla; f. peridium in section; g–j. asci (g. with paraphysis, j. in Lugol after KOH pre-treatment); k, l. ascus tips (l. in Lugol after KOH pre-treatment); m–y. ascospores; arrows denoting septa. — z–h1. Leptosillia aff. fusariospora (NY 00270482). z, a1. Cupulate perithecia on bark in side (z) and surface (a1) view; b1–h1. ascospores. All in 3 % KOH, except where noted. — Scale bars: a–e, z–a1 = 100 μm; f–j = 10 μm; k–y, b1–h1 = 5 μm.

Basionym. Coelosphaeria fusariospora Ellis & Everh., J. Mycol. 4: 65. 1888.

Synonym. Leiosphaerella fusariospora (Ellis & Everh.) M.E. Barr, Mycotaxon 46: 62. 1993.

Typification. USA, Kansas, on bark of (living)? cottonwood trees (Populus deltoides), soc. Teichospora kansensis, without place and date, G. Egeling, comm. J.W. Eckfeldt, in Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 1957 (NY 00883560, lectotype of Coelosphaeria fusariospora here selected, MBT 385917; GZU 000335714, K(M) 252230, K(M) 252231, NY 00883561, NY 00883562, NY 00883563 isolectotypes).

Ascomata perithecial, superficial to basally immersed in bark, (135–)160–205(–230) μm diam (n = 30), 90–190 μm high, black, shiny, smooth, scattered singly to gregarious, subglobose to hemispherical, circular from above, commonly horizontally collapsed and then cupulate, with a distinct central apical papilla c. 30–55 μm wide, 30–50 μm high. Peridium continuous, dark brown, becoming light brown to hyaline towards the centrum, 17–40 μm thick, of textura angularis composed of thick-walled, isodiametric to elongated cells 4–15 × 2–4 μm with dark brown walls, becoming thin-walled and subhyaline towards the centrum. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched, 1.7–3.5 μm wide paraphyses embedded in an inamyloid gelatinous matrix; periphyses not observed. Asci (46–)51–68(–83) × (7.0–)8.0–9.5(–10.3) μm (n = 50), unitunicate, cylindrical, straight, curved to sinuous, thin-walled, containing 8 ascospores arranged biseriately or in two fascicles, with a short stipe, inamyloid and without a distinct apical apparatus. Ascospores (20–)24–30(–33) × (2.3–)2.7–3.3(–3.6) μm, l/w = (6.4–)7.7–10.2(–12.0) (n = 90), mostly fusiform to slightly curved with strongly curved to hooked ends, sometimes falcate to lunate, aseptate, occasionally becoming uniseptate at maturity, hyaline, thin-walled, smooth, with subacute tapering ends. Asexual morph not observed.

Habitat & Host range — With certainty known only from bark of living trunks of Populus deltoides; probably also on Celtis occidentalis.

Distribution — North America; with certainty only known from Kansas, the USA.

Additional specimens examined. USA, Kansas, Rockport, on bark of Celtis orientalis, Nov. 1893, E. Bartholomew, in Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 3016 (K(M) 252322, NY 00270482).

Notes — Leptosillia fusariospora is well characterised by an ascospore shape similar to macroconidia of Fusarium, from which its species epithet was derived. It is similar to the European L. acerina in its horizontally collabent, cupulate ascomata and has aseptate ascospores of similar size; however, it differs by differently shaped ascospores occasionally becoming uniseptate at maturity, different hosts and distribution (North America vs Europe). Unfortunately, no cultures and DNA sequences are available for L. fusariospora, but both the morphological characteristics and the ecology of the species match with the genus Leptosillia, into which it is therefore combined.

Numerous copies of the type collection were distributed as Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 1957, but to our knowledge no lectotype has yet been selected. In NY, where the Ellis collection is kept, there are four collections corresponding to the protologue, one (NY 00883560) labelled as holotype, two, bearing the label of Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 1957 (NY 00883561, NY 00883562), as isotypes, and one (NY 00883563) without a type label but with the same data given for the other collections, but with a collection date Oct. 1887. The latter also represents an isotype as it was collected ahead of the publication of the taxon, and it includes an original note with exactly the same ascospore and ascus measurements as given in the protologue. Based on preservation, the isotype specimen NY 00883560 of the Ellis collection is here selected as lectotype. Most of the isotype specimens investigated also contain ascomata of Teichospora kansensis.

The collection from Celtis occidentalis distributed as Coelosphaeria fusariospora in Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 3016 is here only tentatively attributed to the species. It has distinctly longer ascospores ((28–)33–41(–44) × (2.2–)2.5–2.9(–3.1) μm, l/w = (10.8–)12.1–15.3(–16.7) (n = 25)) of a different shape, being narrowly fusiform to falcate but lacking strongly curved to hooked ends (Fig. 9b1–h1), and also the ascomata are distinctly superficial (Fig. 9z, a1) and somewhat larger ((183–)206–276(–361) μm diam (n = 60)). The specimen may therefore represent a distinct Leptosillia species, but fresh collections and sequences are necessary for a detailed evaluation.

The treatment of Coelosphaeria fusariospora by Barr (1993) is confusing: first she combined it in Leiosphaerella, but a few pages later she considered the species to be conspecific with Cresporhaphis rhoina. Our detailed re-examination of type specimens of Coelosphaeria fusariospora and Cresporhaphis rhoina did not confirm this synonymy, but revealed them as two different, unrelated species. While asci, ascospores and also the corticolous ecology of Coelosphaeria fusariospora are in full agreement with Leptosillia, Cresporhaphis rhoina differs by an amyloid apical ascus ring, mostly fusoid to curved ascospores of irregular shapes and by growth on dead wood. The latter is therefore not considered to be congeneric with Leptosillia (see notes under C. rhoina below).

Leptosillia macrospora (Eitner) Voglmayr & Jaklitsch, comb. & stat. nov. — MycoBank MB829932; Fig. 10

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g010.jpg

Leptosillia macrospora. a–d. Perithecia on bark; note the stellate or sulcate structures on the apical papillae; e–f. asci in 3 % KOH; g. ascus tip in Lugol; h. paraphyses; i–t. vital ascospores; arrows denoting septa; u–v. pycnidia and conidial drops in culture (CMD, isolation plate, 7 d); w–y. conidiophores, conidiogenous cells and conidia from pycnidia on natural substrate; z–f1. conidia from natural substrate; g1–r1. conidia from pycnidia in culture (CMD, 6 mo). All in water, except where noted (a–b, i–l, n–p, w–f1: WU 39999 (epitype); c–f: WU 40004; g–h, m, q–t: WU 40000; u–v, g1–r1: CRM2). — Scale bars: a–b, u = 200 μm; c–d, v = 100 μm; e–f, h–t, w–y = 10 μm; g, z–r1 = 5 μm.

Basionym. Leptorhaphis quercus f. macrospora Eitner, Jahresber. Schles. Ges. Vaterl. Cult. 78: 25. 1901 ‘1900’.

Synonyms. Cresporhaphis macrospora (Eitner) M.B. Aguirre, Bull. Brit. Mus. (Nat. Hist.), Bot. 21 (2): 149. 1991.

Liberomyces macrosporus Pažoutová et al., Mycologia 104: 201. 2012.

Typification. POLAND, Silesia, Nimptsch, Klein Ellguth, on Quercus robur, 12 Apr. 1892, E. Eitner (W 19701, lectotype of Leptorhaphis quercus f. macrospora selected by Aguirre-Hudson 1991). – AUSTRIA, Niederösterreich, Schönfeld, Wacholderheide NE of the golf course, on bark of living trunks of Quercus petraea, 5 May 2016, H. Voglmayr & I. Greilhuber (WU 39999, epitype of Leptorhaphis quercus f. macrospora here designated (MBT 385918); ex epitype culture CRM2 = CBS 143627).

Ascomata perithecial, half-immersed in bark to superficial, (170–)200–255(–275) μm wide (n = 46), (210–)220–270(–280) μm high (n = 6), black, smooth, scattered singly, sometimes gregarious, pyriform, circular from above, sometimes laterally collapsed, with a central apical papilla laterally enlarged by stellate or sulcate structures on the surface. Peridium continuous, dark brown, becoming hyaline towards the centrum, 20–30 μm thick, of a textura angularis composed of thin-walled, isodiametric to elongated cells 4–8 μm diam with subhyaline to dark brown walls. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched, 1.5–3.3 μm wide paraphyses embedded in an inamyloid gelatinous matrix; periphyses not observed. Asci (94–)104–120(–135) × (8.8–)9.5–11.5(–13.0) μm (n = 46), unitunicate, clavate to cylindrical, curved, thin-walled, containing 8 ascospores arranged in fascicles, with a short stipe, inamyloid and without a distinct apical apparatus. Ascospores (53–)65–93(–109) × (3.0–)3.5–4.5(–5.0) μm, l/w = (12.8–)16.1–24.2(–31.3) (n = 82), variable in shape from sinuous, sigmoid, semicircular to hook-shaped, at maturity 1–3 septate, hyaline, thin-walled, smooth, with narrowly rounded ends, multiguttulate when vital.

Pycnidia scattered on bark, black, similar to ascomata except for smaller size, c. 100–150 μm diam. Peridium continuous, dark brown, c. 10 μm thick, composed of few layers of thin-walled dark brown cells 3–5.5 μm diam. Conidiophores short, reduced, hyaline, smooth, branched up to two times, composed of short, cylindrical to almost isodiametric cells arising from the inner wall of the pycnidium. Conidiogenous cells (2.0–)7.3–12.0(–17.5) × (0.9–)1–1.9(–2.5) μm (n = 83), enteroblastic, phialidic, lageniform to cylindrical, hyaline, smooth, arranged in dense terminal whorls. Conidia (16–)18–23(–24) × (0.8–)0.9–1.1(–1.2) μm, l/w = (16.3–)17.6–23.7(–26.1) (n = 20), filiform, curved, aseptate, hyaline, thin-walled, smooth, containing few guttules when vital.

Culture characteristics and asexual morph in culture — Colony on CMD at 16 °C reaching 38–56 mm diam after 58 d; variable in colour and growth depending on the strain, greyish brown to black, sometimes with cream sectors, with sparse, short to woolly aerial mycelium, margin often strongly uneven, wavy, reverse light brown to black. Pycnidia (85–)100–160(–210) μm diam (n = 24), immersed to almost superficial, dark brown to black, single, aggregated to confluent, opening by an ostiole or irregular rupture and exuding white masses of conidia. Ostiole circular or oval, 20–40 μm wide; ostiolar neck 10–30 μm high. Peridium of an outer layer of textura intricata of dark brown, thick-walled cells, and an inner layer of textura angularis of lighter brown cells, basal parts of the peridium subhyaline, consisting of thin-walled cells. Conidiophores simple or irregularly branched, hyaline, smooth, arising from the inner wall of the entire conidioma. Conidiogenous cells of two types: a) (7–)8(–10) × 1.5–1.7 (–2) μm, holoblastic with sympodial proliferation, bearing allantoid conidia; b) similar to those observed on the natural substrate, enteroblastic, phialidic, bearing filiform conidia. Conidia of two types: a) holoblastic, (7.0–)9.7–12.5(–15.2) × (1.3–)1.6–2.0(–2.2) μm, l/w = (4.5–)5.4–7.0(–8.5) (n = 150), allantoid, often typically curved on the proximal end, hyaline, smooth; b) enteroblastic similar to those observed on the natural substrate, (14–)17(–21) × 1 μm, filiform, curved.

Habitat & Host range — On bark of living trunks of various Quercus species (Aguirre-Hudson 1991, this study); in one occasion isolated from healthy phloem of living Ulmus laevis (Pažoutová et al. 2012).

Distribution — Europe; known from Austria, Croatia, Czech Republic, Germany, Hungary, Poland, Sweden, Switzerland (Aguirre-Hudson 1991, Otte et al. 2017, this study).

Additional specimens examined (all on bark of living trunks of Quercus spp.). Austria, Burgenland, Purbach, Purbacher Heide, on Quercus pubescens, 1 Apr. 2017, H. Voglmayr & I. Greilhuber (WU 40000); Niederösterreich, Stopfenreuth, Donauauen, on Quercus robur, 25 Mar. 2017, H. Voglmayr & I. Greilhuber (WU 40001); Oberösterreich, St. Willibald, between Oberantlang and Landersberg, on Quercus robur, 22 May 2016, H. Voglmayr (WU 40002, culture CRM4). – GERMANY, Bayern, Bernried am Starnberger See, park of Schloss Hohenried, on Quercus robur, 12 Sept. 2016, H. Voglmayr & W. Jaklitsch (WU 40003, culture CRM7); Niedersachsen, Hamburg (near), Buxtehude, by roadside, on Quercus robur, 30 Aug. 2015, H.G. Wagner (K(M) 199846); Hannover, Bückeburger Allee, on Quercus robur, 26 Apr. 2016, H.G. Wagner (WU 40004, culture CRM1); Thüringen, NE of Eisenach, W of Wolfsbehringen, on the wayside of an old oak forest, on young Quercus robur, 13 June 2008, H.G. Wagner (K(M) 158044); Zechsteingürtel, Kyffhäuser, N Bad Frankenhausen, Georg-Höhe, on Quercus sp., 12 Mar. 2015, J. Eckstein 38831 (K(M) 201601).

Notes — At least in Central Europe, Leptosillia macrospora seems to be widely distributed and probably not uncommon on young oak trees, sometimes even found on trees planted by roadsides in towns and cities. All our recent collections were from bark crevices of oak trunks of 10–30 cm diam.

Leptosillia macrospora might be confused with several other unrelated fungi (lichenised or not) also with colourless, multiseptate, filiform ascospores; e.g., Rhaphidicyrtis trichosporella, also found on oaks, which differs by bitunicate asci and hamathecium gel turning deep blue in Lugol’s Iodine pretreated with 10 % KOH. There are also similarities to some species of the genus Pseudosagedia, and in particular with P. leptospora, but in this the ascospores are at least 7-septate, the ascomata present an additional involucrellum over the exciple, and the thallus is clearly lichenized with Trentepohlia. Berger & Priemetzhofer (2000) reported the species from Austria growing on Tilia cordata (Donautal, Oberösterreich, Berger 9578). We requested the material for study, but instead we received another collection from the same area, also labelled as Creporhaphis macrospora, but growing on Malus sp. (Berger 12951). Examination of this voucher revealed a fungus with filiform, multiseptate ascospores (75–90 × 3.5 μm) but with distinct fissitunicate asci. This collection is probably a new species of the genus Lophiostoma (sensu Hirayama & Tanaka 2011), related to Lophiostoma subcutanea (see Huhndorf 1992: 503–504; Fig. 2), which is also found on bark of Rosaceae but has smaller ascospores (25–29 × 3–3.5 μm).

Based on sequence data and morphology, Liberomyces macrosporus represents the asexual morph of Leptosillia macrospora, and is therefore a synonym of the latter. The description of the asexual morph in pure culture was modified from the description of Pažoutová et al. (2012), and that of the pycnidia from natural substrate was adapted from the description in Aguirre-Hudson (1991). In the present study, pycnidia on the natural substrate could be found in only one occasion, and the pycnidium investigated only produced enteroblastically formed, filiform conidia; however, their size agrees well with those recorded from culture and given in Aguirre-Hudson (1991).

Leptosillia muelleri (Duby) Voglmayr & Jaklitsch, comb. nov. — MycoBank MB829933; Fig. 11, ,1212

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g011.jpg

Leptosillia muelleri, sexual morph. a–i. Single and confluent perithecia on bark in surface (a–f) and side (g–i) view; j. vertical section through pseudo-stroma with perithecia; k. strongly dextrinoid granular hamathecial exudates in Lugol after KOH pre-treatment; l–m. asci; n. ascus tip in Lugol; o. paraphyses; p–g1. ascospores (p–t. dead, u–g1. vital). All in water, except where noted (a, e, j, p–t: FH 00304540 (holotype of Leptosillia notha); b–c, f–g, l, o, u, y–z: WU 40005 (epitype); d, h, n, v–x: WU 40006; i, m, a1–b1: WU 40007; k, c1–g1: WU 40008). — Scale bars: a–b = 500 μm; c–e = 200 μm; f–i = 100 μm; j = 50 μm; k–n = 10 μm; o–g1 = 5 μm.

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g012.jpg

a–d1. Leptosillia muelleri, asexual morph. a–c. Pycnidia on CMD isolation plates (a–b. 7 d; c. 37 d); d–h. conidiophores, conidiogenous cells and conidia from pycnidia on natural substrates; i–k. conidiophores, conidiogenous cells and conidia from pycnidia in pure culture (i, k. CMD 8 d; j. CMD 19 d); l–r, z–d1. conidia from natural substrate (l–n. dead; o–r, z–d1. vital); s–y. vital conidia from pycnidia in culture on CMD (s–u. 19 d; v–y. 8 d). All in water, except d, l–n from permanent slide (a–b, i, v–x: CRM3 (ex-epitype culture); c, j, s–u: CRM; d, l–n: FH 00304540 (holotype of Leptosillia notha); e–h, o–r, z–d1: WU 40005 (epitype); k, y: CRM6). e1. Septoria notha (holotype, PAD) — Scale bars: a = 1 mm; b–c = 200 μm; d–k = 10 μm; l–d1 = 5 μm; e1 = 3 mm.

Basionym. Sphaeria muelleri Duby, in Rabenhorst, Klotzsch. Herb. Vivum. Mycol., Edn 2: no. 642. 1858.

Synonyms. Cresporhaphis muelleri (Duby) M.B. Aguirre, Bull. Brit. Mus. (Nat. Hist.), Bot. 21: 151. 1991.

Leptosphaeria muelleri (Duby) Auersw., in Gonnermann & Rabenhorst, Mycol. Eur. Pyren. 5–6: t. 12, f. 167. 1869.

Psilosphaeria muelleri (Duby) Cooke (as ‘mulleri’), Grevillea 16 (no. 78): 50. 1887.

Zignoëlla muelleri (Duby) Sacc. & Traverso, Syll. Fung. (Abellini) 20: 1170. 1911.

Cytosporina notha (Sacc.) Died., Krypt.-Fl. Brandenburg (Leipzig) 9: 545. 1914.

Harpostroma nothum (Sacc.) Höhn. (as ‘notha’), Mitt. Bot. Inst. Tech. Hochsch. Wien 5: 112. 1928.

Leptosillia notha Höhn., Mitt. Bot. Inst. Tech. Hochsch. Wien 5: 111. 1928.

Leptorhaphis aggregata Eitner, Jahresber. Schles. Ges. Vaterl. Cult. 78 (2b Abth.): 25. 1901 [1900].

Leptorhaphis wienkampii var. aggregata (Eitner) Keissl., Rabenh. Krypt.-Fl. 9, 1: 246, f. 82. 1937.

Typification. FRANCE, Haute Savoie, Les Contamines, on bark of Acer, without date, J. Müller Argoviensis, in Rabenhorst, Klotzsch. Herb. Vivum. Mycol. Ed. II no. 642 (K(M) 252333, lectotype of Sphaeria muelleri selected by Aguirre-Hudson 1991; K(M) 252334, isolectotype). – GERMANY, Erfurt, on bark of Acer pseudoplatanus, 15 Apr. 1905, H. Diedicke, Herb. A. Höhnel no. 4269 (FH 00304540, holotype of Leptosillia notha). – AUSTRIA, Oberösterreich, St. Willibald, Oberantlang N Siegl, on bark of Acer pseudoplatanus, 22 May 2016, H. Voglmayr (WU 40005, epitype of Sphaeria muelleri (MBT 385919) and Leptosillia notha (MBT 385920) here designated; ex epitype culture CRM3 = CBS 143628). – FRANCE, Saintes, on Acer pseudoplatanus, without date, P. Brunaud no. 67, Herb. Saccardo (PAD, holotype of Septoria notha).

Ascomata perithecial, embedded in a pseudostroma, emerging from cracks on the surface of bark scales, (100–)140–210(–260) μm diam (n = 67), black, matt, smooth, rarely scattered singly and pyriform, but usually confluent and then irregular in shape and c. 1 mm long, immersed in bark to half of their height, not collapsing, with an indistinct to distinct central apical papilla. Peridium continuous, of a textura angularis, composed of an outer dark brown to black, 10–30(–45) μm thick layer of thin-walled isodiametric cells 2.5–5.5 μm diam with dark brown walls, forming a pseudostroma surrounding the inner wall, and an inner, 12–20 μm thick subhyaline to pale brown layer corresponding to the perithecium wall of (sub)hyaline to light brown cells similar to those of the outer layer but slightly smaller and sometimes radially compressed changing into a textura prismatica. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched, 1–3 μm wide paraphyses embedded in an inamyloid gelatinous matrix; with hyaline, refractive, strongly dextrinoid granular exudates turning amber-red in Lugol; periphyses not observed. Asci (65–)75–100(–107) × (7.2–)8.0–9.5(–11.3) μm (n = 25), unitunicate, cylindrical, slightly curved to sinuous, thin-walled, containing 8 bi- or tri seriately arranged ascospores, with a short stipe, without a distinct apical apparatus, inamyloid but a narrow, short pore visible in Lugol. Ascospores (20–)25–33(–38) × (2.0–)2.8–3.7(–4.7) μm, l/w = (5.6–)7.7–10.1(–12.0) (n = 98), fusoid, lunate to falcate, aseptate, hyaline, thin-walled, smooth, with subacute to narrowly rounded tapering ends, when vital containing 2–3 large and numerous small guttules especially towards the ends.

Pycnidia on bark usually confluent, black, practically indistinguishable from ascomata. Conidiophores short, hyaline, smooth, simple or irregularly branched, arising from the inner wall of the pycnidium. Conidiogenous cells (6.8–)8.0–15.2(–27.5) × (1.2–)1.6–3.6(–5.3) μm (n = 72), lageniform to cylindrical, of two types interspersed within the same pycnidium: a) holoblastic with sympodial proliferation, bearing falcate conidia; b) enteroblastic, phialidic, bearing narrower, filiform conidia. Conidia of two types: a) holoblastic, 21–27(–32) × 2.0–2.5(–3.0) μm, l/w = (8.8–)9.5–11.8(–13.0) (n = 27), falcate, hyaline, smooth, with narrowly rounded ends, with few small guttules when vital; b) enteroblastic, (19–)23–28(–31) × (0.8–)0.9–1.2(–1.4) μm, l/w = (16.5–)19.9–29.0(–34.6) (n = 33), filiform, curved to semi-circular.

Culture characteristics and asexual morph in culture — Colony on CMD at 16 °C reaching 50–55 mm diam after 58 d; first white, turning cream to greyish brown in the centre, with white woolly aerial mycelium, reverse cream, dark greyish brown in the centre, with age secreting a deep yellow diffusible pigment in agar. Pycnidia (115–)145–330(–500) μm diam (n = 114), immersed to almost superficial, black, single, aggregated to confluent, opening by an ostiole and exuding white masses of conidia. Conidiophores short, hyaline, similar to those from natural substrate. Conidiogenous cells holoblastic, similar to those described from natural substrate. Conidia (20–)24–29(–31) × (2.0–)2.5–3.3(–3.5) μm, l/w = (7.0–)8.0–10.1(–12.0) (n = 88), nearly straight, falcate, lunate to hook-shaped, aseptate, hyaline, thin-walled, smooth, with narrowly rounded tapering ends, containing numerous large guttules.

Habitat & Host range — Only known from bark scales of mature living trees of Acer pseudoplatanus.

Distribution — Europe; known from Austria, Czech Republic, France, Germany, Poland, Switzerland (Aguirre-Hudson 1991, this study).

Additional specimens examined (all on bark scales of mature living trunks of Acer pseudoplatanus). AUSTRIA, Kärnten, St. Margareten im Rosental, Oberdörfl, at Nagu, 10 Apr. 2016, H. Voglmayr & W. Jaklitsch (WU 40006; culture CRM); Niederösterreich, Lunz am See, at Mittersee, 10 May 2016, H. Voglmayr (WU 40007; culture CRM6); Puchberg am Schneeberg, Sonnleiten, Wasserfallweg, 5 Aug. 2017, H. Voglmayr & I. Greilhuber (WU 40008). – CZECH REPUBLIC, Bohemia, Petrovice u Sušice, E Chamutice, 1 June 2018, H. Voglmayr & M. Greilhuber (WU 40009).

Notes — The holotype of Leptosillia notha, a holomorphic collection, from the Höhnel herbarium deposited in FH morphologically resembles our recent collections and the type of the earlier name Sphaeria muelleri. Remarkably, in the conidiomata observed on the natural substrate of the epitype two types of conidia are present: falcate and filiform ones, which are formed holoblastically and enteroblastically, respectively. However, in pure culture only holoblastical multiguttulate conidia were found; these were somewhat wider than those observed on the natural substrate. In a permanent mount of conidiomatal sections attached to the holotype, only phialides with filiform conidia were seen, with conidial sizes only slightly wider ((22–)26–31(–34) × (1.3–)1.4–1.7(–1.8) μm (n = 25)); this, however, may be due to the mounting medium. To preserve the holotype, we did not make new preparations of the asexual morph.

Diedicke (1915) identified the asexual morph on the holotype collection of Leptosillia notha as Septoria notha and recombined the species as Cytosporina notha. Subsequently, Höhnel (1928) established the monotypic genus Harpostroma for the latter, but challenged the conspecificity with Saccardo’s Septoria notha. We agree that this conspecificity is doubtful. The type specimen of Septoria notha is extant in PAD, and although it could not be microscopically investigated, no structures resembling Leptosillia notha were seen on the specimen under the stereomicroscope. Also the ecology does not quite fit, as the substrate is a thin, corticated branch of c. 6 mm diam (Fig. 12e1), while L. notha is confined to bark scales of old living trunks. In the original description (Saccardo 1880), the host of Septoria notha is erroneously given as Acer platanoides; it is here re-identified as Acer pseudoplatanus based on bark and wood characters of the type specimen. This is in line with the fact that Saccardo (1880) assumed a connection with Diaporthe hystrix, a species commonly known from Acer pseudoplatanus but not from A. platanoides (Wehmeyer 1933).

Leptosillia pinicola (Samp.) Voglmayr & Jaklitsch, comb. nov. — MycoBank MB829934; Fig. 13

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g013.jpg

Leptosillia pinicola (UPS L-074953, isoneotype). a–f. Single and gregarious perithecia on bark in surface (a–d) and side (e–f) view; g. peridium in section; h. paraphyses; i–k. asci (k in Lugol); l–o. ascospores. All in 3 % KOH, except where noted. — Scale bars: a = 1 mm; b–c = 200 μm; d–f = 100 μm; g, l–o = 5 μm; h–k = 10 μm.

Basionym. Leptorhaphis pinicola Samp., Bolm Soc. Broteriana, Coimbra, sér. 2 2: 163. 1924 (1923).

Synonym. Cresporhaphis pinicola (Samp.) M.B. Aguirre, Bull. Brit. Mus. (Nat. Hist.), Bot. 21: 152. 1991.

Typification. PORTUGAL, Estremadura, Sierra de Sintra, Castelo dos Mouros, on bark of Pinus sp., 11 Apr. 1943, C. Tavares (LISU 511, neotype designated by Aguirre-Hudson 1991; UPS L-074953!, isoneotype).

Ascomata perithecial, superficial on bark, (150–)180–230(–280) μm wide (n = 32), black, shiny, smooth, scattered singly, sometimes gregarious, globose to pyriform, circular from above, with an indistinct central apical papilla. Peridium continuous, dark brown, becoming hyaline towards the centrum, c. 20 μm thick, of textura angularis composed of thick-walled, isodiametric to slightly elongated cells 3.5–5.5 μm diam with dark brown walls, towards the centrum becoming a textura angularis-prismatica of thinner-walled pale brown to subhyaline cells. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched, 1.5–2.3 μm wide paraphyses embedded in an inamyloid gelatinous matrix; periphyses of hyaline, smooth, thin-walled, 1.5–2 μm wide hyphae. Asci (74–)78–89(–95) × (9.5–)9.8–11.0(–11.7) μm (n = 24), unitunicate, cylindrical to fusoid, usually slightly curved, thin-walled, containing 8 ascospores arranged in fascicles, with a short stipe, inamyloid and without a distinct apical apparatus. Ascospores (35–)44–58(–65) × (1.8–)2.4–3.0(–3.5) μm, l/w = (14.1–)16.2–21.7(–23.7) (n = 20), acicular, often slightly curved, 5–11-septate, not constricted at the septa, hyaline, thin-walled, smooth.

Notes — Only two collections from the type locality in Portugal (Sintra, near Lisbon; Aguirre-Hudson 1991), dating back to the first half of the 20th century, are confirmed here as belonging to Leptosillia pinicola. Unfortunately, the species could not be recollected by the first author despite extensive search on the bark of various pine species at and near the type locality. Despite the lack of fresh material for sequencing, morphologically the species fits well in the genus Leptosillia. The current description and illustrations are based on the isoneotype specimen from UPS, with few additions from Aguirre-Hudson (1991).

The species (as Cresporhaphis pinicola) has been cited from Austria by Berger et al. (1998) from bark of Prunus avium, and from Lithuania by Motiejûnaitë (2007) from bark of Berberis sp. Re-examination of the latter has confirmed that the material (K(M) 117899) is not conspecific with the type of Leptosillia pinicola because the ascospores are longer and more slender (62–78 × 2–3 μm), and arranged in a single fascicle in the ascus. This collection might yet represent a new species of Leptosillia, but DNA studies will be needed to confirm this. It is also unlikely that the material recorded from Austria is conspecific to L. pinicola due to the unrelatedness of the host, but we had no opportunity to study the collection.

Leptosillia pistaciae (Voglmayr et al.) Voglmayr, comb. nov. — MycoBank MB829935

Basionym. Liberomyces pistaciae Voglmayr et al., Mycokeys 40: 41. 2018.

Notes — In the current phylogenetic analyses this recently described serious canker pathogen of pistachio (Pistacia vera) is placed within Leptosillia, which necessitates a generic transfer. So far, no sexual morph is known for this species. For morphological description, illustrations and pathogenicity, see Vitale et al. (2018).

Leptosillia slaptonensis (P.F. Cannon) Voglmayr, M.B. Aguirre & Jaklitsch, comb. nov. — MycoBank MB829936; Fig. 14

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g014.jpg

Leptosillia slaptonensis. a–d. Perithecia on bark in surface (a–b) and side (c–d) view; e–f. asci (f. in Lugol); g. paraphyses; h–m, o–t. vital ascospores; arrows denoting septa; n. pale, translucent pycnidia in culture (CMD, isolation plate, 42 d); u–v. conidiophores, conidiogenous cells and conidia from pycnidia on natural substrate; w–a1. conidia from natural substrate; b1–c1. conidiophores and conidiogenous cells from pycnidia in culture (CMD, isolation plate, 42 d); d1–j1. conidia from pycnidia in culture (CMD, isolation plate; d1. 20 d; e1–j1. 40 d). All in water, except where noted (a–f, h–j: WU 40010 (epitype); g, o–p: WU 40012; k–m: WU 40014; q–a1: WU 40015; n, b1–c1, e1–j1: CRU3; d1: CRU2). — Scale bars: a–d = 100 μm; e–g, u–v, b1–c1 = 10 μm; h–m, o–t, w–a1, d1–j1 = 5 μm; n = 400 μm.

Basionym. Zignoëlla slaptonensis P.F. Cannon, Syst. Ascomycetum 15: 129. 1997.

Synonym. Cresporhaphis ulmi Calat. & M.B. Aguirre, Mycol. Res. 105: 123. 2001.

Typification. GREAT BRITAIN, England, South Devon, Slapton, near Kingsbridge, Slapton Ley National Nature Reserve, Marsh Lane, on dead cankered branches of Ulmus minor, 6 May 1994, P.F. Cannon (IMI 362466c, holotype of Zignoëlla slaptonensis). – SPAIN, Aragón, Teruel, between Puebla de Arenoso and Olba, close to Los Lucas, c. 2 km E of Olba, c. 700 m a.s.l., on suberose outgrowths of Ulmus minor twigs, 14 Mar. 1999, V. Calatayud (MA-Fungi 41352, holotype of Cresporhaphis ulmi). – AUSTRIA, Niederösterreich, Mödling, Eichkogel, on cork wings and outgrowths of branches of Ulmus minor, 29 Apr. 2015, W. Jaklitsch & H. Voglmayr (WU 40010, epitype of Zignoëlla slaptonensis (MBT 385921) and Cresporhaphis ulmi (MBT 385922) here designated; ex epitype culture NAD = CBS 145296).

Ascomata perithecial, superficial to partly immersed in bark, (115–)145–190(–250) μm diam (n = 77), black, shiny, smooth, scattered singly to aggregated and occasionally confluent, pyriform, circular from above, commonly laterally collapsed, with a central apical papilla. Peridium continuous, c. 25–30 μm thick, a textura angularis of thin-walled, isodiametric or somewhat elongated dark brown cells 6–10 μm diam with dark brown walls, becoming paler towards the centrum. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched, 2–4 μm wide paraphyses embedded in an inamyloid gelatinous matrix; periphyses 2–3 μm wide, unbranched, thin-walled, smooth. Asci (67–)79–98(–114) × (9.5–)10.2–12.3(–14.5) μm (n = 64), unitunicate, clavate to fusiform, curved, thin-walled, containing 8 ascospores arranged in two fascicles, with a short stipe, inamyloid and without a distinct apical apparatus; with fissurate dehiscence. Ascospores (31–)37–46(–55) × (3.2–)3.5–4.0(–4.8) μm, l/w = (7.4–)9.4–12.6(–14.0) (n = 116), falcate, 1- or 3-septate, hyaline, thin-walled, smooth, with narrowly to broadly rounded ends, multiguttulate when vital.

Pycnidia scattered on bark, black, practically indistinguishable from ascomata except for slightly smaller size. Conidiophores short, hyaline, smooth, branched up to two times, arising from the inner wall of the pycnidium. Conidiogenous cells (5.7–)7.3–10.0(–13.5) × (1.4–)1.6–2.2(–3.1) μm (n = 75), holoblastic with sympodial proliferation, lageniform to cylindrical, hyaline, smooth, disposed in dense terminal whorls of up to 5. Conidia (15–)19–23(–25) × (1.5–)1.7–2.2(–2.7) μm, l/w = (7.4–)9.3–12.0(–14.0) (n = 90), falcate to lunate, aseptate, hyaline, thin-walled, smooth, with narrowly rounded ends, containing small guttules when vital.

Culture characteristics and asexual morph in culture — Colony on CMD at 16 °C reaching 45–51 mm diam after 58 d; first cream, turning dark grey brown to black in the centre, with sparse aerial mycelium mostly in the centre, margin even, reverse medium to dark grey brown at least in the centre. Pycnidia (230–)250–370(–410) μm diam (n = 10), partly immersed to almost superficial, pale whitish translucent, aggregated to confluent, opening by irregular apical ruptures. Conidiophores and conidiogenous cells similar to those from the natural substrate but less regular and more variable in shape; often producing a single conidium; sympodial conidiation rarely seen. Conidia (13–)15–23(–29) × (2.1–)2.3–2.7(–3.1) μm, l/w = (4.5–)6.0–9.4(–12.6) (n = 50), similar to those from the natural substrate but more irregular and variable in shape, varying from allantoid, falcate to sigmoid, aseptate, rarely becoming 1-septate, hyaline, thin-walled, smooth, with mostly broadly rounded ends, sometimes containing numerous guttules especially towards the ends.

Habitat & Host range — Only known from cork wings and outgrowths of living or dead branches of Ulmus minor.

Distribution — Europe; known from Austria, UK, Spain (Cannon 1997, Calatayud & Aguirre-Hudson 2001, this study).

Additional specimens examined (all on cork wings of branches of Ulmus minor). AUSTRIA, Niederösterreich, Marchauen E Markthof, 8 Sept. 2018, H. Voglmayr & I. Greilhuber (WU 40011, culture CRU3); Marchauen E Schloß-hof, 17 June 2017, H. Voglmayr & I. Greilhuber (WU 40012); Orth an der Donau, Donauauen near Uferhaus, 10 Mar. 2018, H. Voglmayr & I. Greilhuber (WU 40013); Neunkirchen, Mollram, 24 Nov. 2018, H. Voglmayr & I. Greilhuber (WU 40022); Prellenkirchen, Spitzerberg SW Edelstal, 12 Mar. 2017, H. Voglmayr & I. Greilhuber (WU 40014, culture CRU2); Wien, 21 distr., Stammersdorf, Marchfeldkanalweg near Heeresspital, 12 June 2016, H. Voglmayr & W. Jaklitsch (WU 40015, culture CRU1).

Notes — The types of Zignoëlla slaptonensis and Cresporhaphis ulmi match in all respects, including the host, with the former name having priority. A recent Austrian collection, for which a culture and sequences are available, is here selected as epitype for both Z. slaptonensis and C. ulmi to stabilise the species concepts and the nomenclatural connection of both names. Leptosillia slaptonensis and L. acerina resemble in habitus and share a similar ecology, both growing on cork wings and outgrowths (rhytidome) of living or recently dead branches, but they differ in their hosts and by ascospore characters.

The ITS GenBank accession FJ025239, derived from an endophyte isolated by Sun et al. (2012) from twigs of Ulmus macrocarpa in China, represents a distinct Leptosillia species, apparently closely related to L. wienkampii and L. slaptonensis, which both occur on Ulmus species in Europe. Leptosillia slaptonensis is so far only known from Ulmus minor, a host on which also L. wienkampii has been found in the present study; however, the latter only occurred on bark of living trunks, while all collections of L. slaptonensis were found on cork wings of thin branches. Leptosillia slaptonensis and L. wienkampii are closely related (Fig. 1, ,2)2) and have a similar ascospore shape and overlapping ascospore sizes, but can be reliably distinguished by the 1–3 septate vs aseptate ascospores, respectively.

Leptosillia wienkampii (J. Lahm ex Hazsl.) Voglmayr & Jaklitsch, comb. nov. — MycoBank MB829937; Fig. 15

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g015.jpg

Leptosillia wienkampii. a–d. Perithecia on bark in surface (a–b) and side (c–d) view; note the sulcate structures on the apical papillae (c–d); e–f. asci (f. in Lugol); g. paraphyses; h–t. vital ascospores; u. strongly dextrinoid granular hamathecial exudates in Lugol after KOH pre-treatment; v. pycnidia and conidial drops in culture (CMD, 16 d); w–z, g1. conidiophores, conidiogenous cells and conidia from pycnidia on natural substrate; a1–f1. conidiophores, conidiogenous cells and conidia from pycnidia in culture (CMD, isolation plate, 40 d). All in water, except where noted (a, g, r: WU 40021; b: WU 40018; c, s–t, w–z, g1: WU 40020; d–e, h–p: WU 40017; f, q: WU 40016 (epitype); u: WU 40023; v, a1, e1: CRW3; b1–d1, f1: CRW1). — Scale bars: a–b, v = 200 μm; c–d = 100 μm; e–f, u = 10 μm; g–t, w–g1 = 5 μm.

Basionym. Leptorhaphis wienkampii J. Lahm ex Hazsl., Verh. Ver. Nat., Heilk. Pressb. 5: 12. 1861 (1860–1861).

Synonyms. Cresporhaphis wienkampii (J. Lahm ex Hazsl.) M.B. Aguirre, Bull. Brit. Mus. (Nat. Hist.), Bot. 21: 154. 1991.

Liberomyces saliciphilus Pažoutová et al., Mycologia 104: 201. 2012.

Typification. GERMANY, Westfalen, Münster, Handorf, on bark of Salix fragilis, Nitschke, in Rabenhorst, Lich. Eur. Exs. no. 651 (L, neotype of Leptorhaphis wienkampii designated by Aguirre-Hudson 1991; WU s.n., W 2009-00420 isoneotypes). – UNITED KINGDOM, England, Surrey, Richmond, Royal Botanic Gardens Kew, on bark of living trunk of Salix fragilis var. russelliana, 24 Mar. 2016, E. Rangel & M.B. Aguirre-Hudson (WU 40016, epitype of Leptorhaphis wienkampii (MBT 385923) here designated; ex-epitype culture CRW = CBS 143630).

Ascomata perithecial, superficial on bark, (120–)170–260(–320) μm diam (n = 65), black, matt, smooth to areolate, scattered singly, pyriform, circular from above, with a central apical papilla laterally slightly enlarged by stellate or sulcate structures on the surface. Peridium continuous, of a textura angularis, composed of an outer dark brown, 25–40 μm thick layer of thin-walled isodiametric to laterally compressed cells 4–8 × 2–4 μm with dark brown walls, and an inner hyaline to pale brown layer of (sub)hyaline to light brown cells slightly smaller than those of the outer layer. Hamathecium composed of hyaline, smooth, thin-walled, septate, occasionally branched, 1.3–4 μm wide paraphyses embedded in an inamyloid gelatinous matrix; with hyaline, refractive, strongly dextrinoid granular exudates turning amber-red in Lugol; periphyses smooth, thin-walled, unbranched, less than 2 μm wide. Asci (76–)82–110(–134) × (7.8–)8.5–10.3(–11.0) μm (n = 43), unitunicate, cylindrical to clavate, strongly sinuous, thin-walled, containing 8 bi- or triseriately arranged ascospores, with a short stipe, inamyloid and without a distinct apical apparatus. Ascospores (22–)26–39(–48) × (3.0–)3.5–4.2(–5.0) μm, l/w = (5.1–)6.3–11.1(–15.6) (n = 203), falcate to lunate, aseptate, hyaline, thin-walled, smooth, with broadly rounded ends, multiguttulate when vital, often with 2–3 large and numerous small guttules.

Pycnidia scattered on bark, black, similar to ascomata except for slightly smaller sizes. Conidiophores short, hyaline, smooth, densely branched up to three times, arising from the inner wall of the pycnidium. Conidiogenous cells (5.0–)7.5–11.5(–15.8) × (1.2–)1.5–2.0(–2.2) μm (n = 50), holoblastic with sympodial proliferation, lageniform to cylindrical, hyaline, smooth, arranged intercalarily or in dense terminal whorls on the conidiophore. Conidia (5.0–)5.5–6.2(–7.0) × (1.4–)1.6–1.9(–2.1) μm, l/w = (2.5–)3.1–3.7(–4.3) (n = 101), falcate, aseptate, hyaline, thin-walled, smooth, with narrowly rounded ends, containing few small guttules when vital.

Culture characteristics and asexual morph in culture — Colony on CMD at 16 °C reaching 50–66 mm diam after 58 d; variable in colour depending on the strain, cream, often turning dark grey brown to black with age, with sparse to abundant aerial mycelium, margin even or irregularly wavy, reverse cream with medium to dark grey brown patches in the centre or entirely dark grey brown to black. Pycnidia (165–)205–440(–655) μm diam (n = 66), immersed to almost superficial, black, single, aggregated to confluent, uni- or irregularly plurilocular, opening by an irregular rupture and exuding white masses of conidia. Conidiophores simple or irregularly branched, hyaline, smooth, arising from the inner wall of the entire conidioma. Conidiogenous cells of two types: a) holoblastic with sympodial proliferation similar to those from the natural substrate, (3.3–)5.5–8.7(–10.7) × (1.8–)2.0–2.7(–3.0) μm (n = 35), bearing allantoid conidia; b) enteroblastic, phialidic, (4.3–)8.5–12.5(–14.5) × (1.1–)1.4–1.9(–2.2) μm (n = 57), bearing filiform to narrowly falcate conidia. Conidia of two types: when holoblastically formed similar to those recorded from natural substrate, (5.0–)5.5–7.0(–7.8) × (1.2–)1.5–1.8(–2) μm, l/w = (2.9–)3.1–4.7(–6.4) (n = 34), allantoid, hyaline, smooth; when enteroblastically formed (12.7–)16.5–23.8(–29) × (0.8–)1.0–1.4(–1.7) μm, l/w = (9.4–)13.4–22.0(–25.3) (n = 57), filiform to narrowly falcate, hyaline.

Habitat & Host range — On bark of trunks of various deciduous trees; recorded from Populus spp., Pyrus communis, Robinia pseudoacacia, Salix spp. (mostly S. alba and S. fragilis), Ulmus glabra, U. laevis, U. minor (Aguirre-Hudson 1991, Aguirre-Hudson et al. 2005, Pažoutová et al. 2012, this study).

Distribution — Europe; known from Austria, Bulgaria, Czech Republic, Germany, Italy, Norway, Poland, Slovakia, Sweden, UK (Aguirre-Hudson 1991, Aguirre-Hudson et al. 2005, Pažoutová et al. 2012, this study).

Additional specimens examined. AUSTRIA, Niederösterreich, Baden, Helenental, on bark of living trunk of Salix fragilis, 23 Apr. 2016, H. Voglmayr & I. Greilhuber (WU 40017; culture CRW1); Hohenau, Marchauen near sugar refinery, on bark of living trunk of Ulmus laevis, 5 June 2016, H. Voglmayr & I. Greilhuber (WU 40018; culture CRM5); Neunkirchen, Mollram, on bark of living trunk of Ulmus minor, 1 Nov. 2018, H. Voglmayr & I. Greilhuber (WU 40019; culture CRW3); Puchberg am Schneeberg, Sonnleiten, Wasser-fallweg, on bark of living trunk of Salix sp., 24 Nov. 2018, H. Voglmayr & I. Greilhuber (WU 40023); Steiermark, Ardning, riverine forest of the Enns adjacent to Pürgschachener Moor, on bark of living trunk of Ulmus glabra, 26 May 2016, H. Voglmayr & I. Greilhuber (WU 40020; culture CRU). – ITALY, Sicily, Graniti, Casa delle Monache, on bark of living trunk of Ulmus minor, 16 June 2016, H. Voglmayr & W. Jaklitsch (WU 40021; culture CRW2). – UK, England, Surrey, Kew, Royal Botanic Gardens, Lake (NW side of), on bark crevices of Populus lasiocarpa, 20 Aug. 2007, M.B. Aguirre-Hudson & T. Kokubun (K(M) 154226); ibid., on bark of Salix fragilis var. russelliana, 20 Aug. 2007, M.B. Aguirre-Hudson & T. Kokubun (K(M) 154239); South Essex, VC18, Southend-on-Sea, Chalkwell Park, by pond, on bark furrows of Salix sp., 1 July 2014, P.M. Earland-Bennett (K(M) 199631); ibid., Southchurch Park, by lake in park, on bark furrows of Salix sp., 5 June 2014, P.M. Earland-Bennett (K(M) 199632).

Notes — Based on sequence data and morphology, Liberomyces saliciphilus represents the asexual morph of Leptosillia wienkampii, and is therefore a synonym of the latter. Most of the description of the asexual morph in pure culture was based on own observations, with a few additions from the description of Liberomyces saliciphilus by Pažoutová et al. (2012). In the present study, pycnidia on natural substrate could be found on only two specimens, and they produce holoblastically formed allantoid conidia matching those from pure culture. When describing L. saliciphilus, Pažoutová et al. (2012) recorded only the holoblastically formed conidia from pycnidia in pure culture; yet, in some of our pure cultures, both holoblastically and enteroblastically formed conidia were occasionally produced within the same pycnidia. Remarkably, Aguirre-Hudson (1991) recorded pycnidia on the natural substrate with enteroblastically produced, cylindrical to filiform conidia 20–25 μm × 1 μm in size, indicating that on the natural substrate the two different conidial types may be formed in different pycnidia.

Key to accepted species of Leptosillia with sexual morphs

  • 1. Ascospores aseptate, occasionally 1-septate . . . . . . . . . .2

  • 1. Ascospores consistently 1- to multiseptate . . . . . . . . . . . .5

  • 2. On bark of Acer spp.; only known from Europe; ascospores always aseptate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3

  • 2. On bark of other hosts; in Europe or North America; ascospores occasionally 1-septate . . . . . . . . . . . . . . . . . . . . . .4

  • 3. Ascomata commonly confluent, in a pseudostroma, not collapsed; on bark scales of mature trunks of Acer pseudoplatanus; on the natural substrate conidia of two types (enteroblastic-filiform, holoblastic-falcate) formed within the same pycnidium . . . . . . . . . . . . . . . . . . . . . . . . L. muelleri

  • 3. Ascomata solitary, often collapsed; on cork wings and outgrowths of branches of Acer campestre; only enteroblastic-falcate conidia known . . . . . . . . . . . . . . . . . . . . .L. acerina

  • 4. Ascospores falcate to lunate, with broadly rounded ends; ascomata not horizontally collapsed, with an apical papilla laterally slightly enlarged by stellate or sulcate structures; asci strongly sinuous; on various broadleaf trees (mostly Salix and Ulmus spp.) in Europe . . . . . . . . . .L. wienkampii

  • 4. Ascospores straight to slightly curved, usually with distinctly hooked, narrowly rounded ends (similar to Fusarium macroconidia); ascomata commonly horizontally collapsed and cupulate, with an apical papilla without stellate or sulcate structures; asci straight, curved to slightly sinuous; on Populus deltoides and (probably) Celtis occidentalis in North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .L. fusariospora

  • 5. Ascospores multiseptate; on bark of trunks of Pinus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .L. pinicola

  • 5. Ascospores 1–3-septate; on various broadleaf trees . . . .6

  • 6. Ascospores 50–110 μm long; on trunks of Quercus spp. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . L. macrospora

  • 6. Ascospores 30–55 μm long; on cork wings and outgrowths of branches of Ulmus minor . . . . . . . . . . . .L. slaptonensis

EXCLUDED CRESPORHAPHIS SPECIES

Based on morphology and ecology, the following two species are considered not to be congeneric with Cresporhaphis, and they are therefore not transferred to Leptosillia.

Cresporhaphis rhoina M.E. Barr, Mycotaxon 46: 64. 1993; Fig. 16

An external file that holds a picture, illustration, etc.
Object name is per-42-228-g016.jpg

Cresporhaphis rhoina. a–d. Horizontally collapsed, cupulate perithecia on wood in surface (a–c) and side (d) view; e. ascoma in vertical section; f. peridium in section; g–h, k. asci and paraphyses (k. in Lugol after KOH pre-treatment); i–j, l–m. ascus tips (l–m. in 3 % KOH followed by Lugol, showing the shallow amyloid apical ring); n. paraphyses tips; o–z. ascospores. All in 3 % KOH, except where noted (a–b, d–f, i–j: GZU 000335638 (isotype); c, g–h, k–w: GZU 000335637 (isotype); x–z: NY 00875175 (lectotype)). — Scale bars: a = 400 μm; b–c = 200 μm; d = 100 μm; e = 50 μm; f–h, k, o–z = 10 μm; i–j, l–n = 5 μm.

Replaced synonym. Sphaeria rhoina Ellis & Everh., J. Mycol. 1 (7): 92. 1885, non Sphaeria rhoina Schwein., Trans. Amer. Philos. Soc., New Series 4 (2): 218. 1832 ‘1834’.

Typification. USA, New Jersey, Gloucester Co., Newfield, weather-beaten dead wood of Rhus copallinum, May 1885, without collector, in Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 1669 (NY 00875175, lectotype of Sphaeria rhoina Ellis & Everh. (MBT 385924) here selected; GZU 000335637, GZU 000335638, isotypes).

Ascomata perithecial, superficial on dead wood, (140–)170–300(–460) μm diam (n = 101), 60–140 μm high (n = 20), black, matt, smooth, scattered to gregarious, lenticular, horizontally collapsed and distinctly cupulate when dry, circular from above, with or without a small central apical papilla. Peridium continuous, of a textura angularis, 23–38 μm thick, composed of an outer blackish brown, 12–28 μm thick layer of very thick-walled, more or less isodiametric cells with dark brown walls, and an inner brown, 5–20 μm thick layer of brown, elongate, thin-walled cells 5–19 × 2–5 μm. Hamathecium composed of hyaline, smooth, thin-walled, septate, mostly unbranched, 1.5–3 μm wide paraphyses embedded in an inamyloid gelatinous matrix; periphyses not observed. Asci (63–)74–87(–102) × (6.3–)7.7–9.0(–10.0) μm (n = 49), unitunicate, fusiform, straight to curved, thin-walled, containing 8 irregularly biseriately arranged ascospores, with a short stipe and a small, shallow, amyloid, c. 1.8 μm wide and 0.5 μm high apical ring. Ascospores (19–)23–30(–39) × (2.5–)2.8–3.2(–3.5) μm, l/w = (6.7–)7.5–10.1(–13.4) (n = 91), variously shaped from straight and fusiform, falcate, hook-shaped to sinuous, aseptate, hyaline, thin-walled, smooth, with narrowly rounded to subacute ends, containing few guttules. Asexual morph unknown.

Notes — Barr (1993) established Cresporhaphis rhoina as a new name for Sphaeria rhoina Ellis & Everh., a later homonym of Sphaeria rhoina Schwein. Based on similar ascomata, asci and ascospores, Barr (1993) considered C. rhoina to be closely related to the generic type, C. wienkampii. The asci were described as unitunicate with a shallow inamyloid apical ring. However, re-examination of the type collection showed the presence of a small but distinct amyloid apical ring in Lugol after KOH pre-treatment, which indicates xylarialean affinities but excludes the species from Cresporhaphis. Also the growth on dead wood differs from all confirmed species of Cresporhaphis, which are all corticolous. It is therefore not congeneric with Cresporhaphis (and accordingly Leptosillia), but its morphological characters are insufficient to allow a well-founded generic reclassification within Xylariales.

The synonymy of C. rhoina and Coelosphaeria fusariospora proposed by Barr (1993) could not be confirmed after re-examination of type material of both species; the latter lacks a distinct amyloid apical ascus ring and has ascospores and a corticolous ecology in line with Leptosillia, into which it is thus combined (see also notes of L. fusariospora above).

Numerous copies of the type collection were distributed as Ellis & Everhart, N. Amer. Fungi Ser. 2 no. 1669, but to our knowledge no lectotype has been selected. The specimen from NY bears the original notes; the ascospore size range is somewhat smaller than in the copies from GZU and exactly matches the range (20–25 μm) given in the original description in Ellis & Everhart (1885). In all other characters, the specimens from NY and GZU fully agree. Therefore, the investigated isotype specimen from NY is here selected as lectotype.

Rhaphidicyrtis trichosporella (Nyl.) Vain., Acta Soc. Fauna Fl. Fenn. 49: 217. 1921

Synonym. Cresporhaphis chibaensis H. Harada, Lichenology 12: 32. 2014.

Holotype of Cresporhaphis chibaensis. JAPAN, Honshu, Chiba-ken, Inzai-shi, Muzai, 15 m elev., on trunk of Alnus japonica, 4 Dec. 2007, H. Harada 25172 (CBM-FL-23891).

For descriptions and illustrations of Cresporhaphis chibaensis, see Harada (2014).

Notes — Cresporhaphis chibaensis differs substantially in several respects from the generic type and the other confirmed species of Cresporhaphis: It is clearly lichenised with a distinct crustose lichen thallus, has perithecial ascomata with lateral, slightly sunken ostioles not situated on an apical papilla, a hamathecium composed of very thin, at least apically anastomosing threads and very long, filiform ascospores with numerous septa. It is therefore not considered to be related to the Cresporhaphis species here transferred to Leptosillia. Harada (2014) failed to note the hamathecial gel Lugol’s iodine reaction in the protologue of the species. Nevertheless, and considering the other morphological features, this species is a later synonym of Rhaphidicyrtis trichosporella, a species known from various substrates in Northern Europe, including Alnus sp. (Ekman et al. 2013).

DISCUSSION

Molecular phylogeny

The molecular phylogenetic analyses confirm a close relationship of all Cresporhaphis species for which DNA data are available with the type species of Leptosillia, and further, these are closely related to Delonicicola. Within this clade, two highly supported lineages are evident in the multigene analyses: the Delonicicola-Furfurella subclade and the Leptosillia subclade, which we recognise as two distinct families, Delonicicolaceae and Leptosilliaceae, based also on marked morphological differences between those genera. In the ITS-LSU rDNA analyses, the Leptosilliaceae are resolved only in the ML analyses (Fig. 2) but not in the MP analyses. This shows that, within Xylariales, the ITS-LSU alone does not always resolve generic and family affiliations well, which is also known from previous studies (e.g., Voglmayr & Yule 2006, Jaklitsch & Voglmayr 2012, Jaklitsch et al. 2016b). This conflict is, e.g., also seen in the Pseudomassariaceae, a morphologically and ecologically well-characterised family, which is also monophyletic in the ML analyses, albeit with low support (Fig. 2), but not resolved in the MP analyses. Insufficient phylogenetic resolution may be the result of rearrangements and length differences of the ITS, causing problems in producing a reliable alignment, in combination with insufficient phylogenetically informative and/or homoplastic characters. Therefore, multigene phylogenies are necessary for an improved phylogenetic resolution within Xylariales (Voglmayr et al. 2018, Wendt et al. 2018).

Classification

The taxa here classified in Leptosillia are a case example how the historical divide of the mycological and lichenological communities led to multiple separate, independent descriptions of the same species within different classification frames, and how this also influenced the hypotheses about their ecology.

Being bark inhabitants, most of the species here classified as Leptosillia were first encountered and described by lichenologists, and based on ascoma and ascospore characters, most of them were originally placed in the heterogeneous genus Leptorhaphis. In her detailed monograph, Aguirre-Hudson (1991) confined Leptorhaphis to bark saprotrophs with affinities to Arthopyreniaceae (Dothideomycetes), and she transferred putatively lichenised species with thin-walled, unitunicate asci, true paraphyses and perithecial ascomata to the new genus Cresporhaphis, which she tentatively classified within the Trichosphaeriales (Sordariomycetes). This classification was mostly accepted up to date (e.g., Lücking et al. 2017), but challenged in Jaklitsch et al. (2016a) who considered this placement doubtful. However, there was consensus that its phylogenetic placement required further detailed studies.

Until our present study, the then monotypic genus Leptosillia was classified within the Diaporthales (Kirk et al. 2008), with a presumed familial affiliation to the Valsaceae (Index Fungorum, accessed Feb. 2019). This classification was primarily based on the original description (Höhnel 1928), which hypothesised a close relationship to the diaporthalean genus Sillia, and was perpetuated in Eriksson & Hawksworth (1987). However, after its description the taxon was never recorded again, and the original material was never critically re-examined. Therefore, it is not surprising that no connection of the little-known Leptosillia notha was ever made to species classified in Leptorhaphis, and later Cresporhaphis.

As a result of our study, the comparison of the type specimens of Cresporhaphis muelleri and Leptosillia notha confirmed them to represent the same species, requiring a name change to L. muelleri, based on priority. As the genus Cresporhaphis has a different generic type species, C. wienkampii, the question arises whether the two genera should be kept separate or classified within the same genus, which in the latter case should be Leptosillia due to priority. The results of the phylogenetic analyses (Fig. 1, ,2)2) revealed both options as tenable, as the L. acerina-L. muelleri and L. macrospora-L. slaptonensis-L. wienkampii lineages formed two distinct subclades within the Leptosilliaceae. However, after critical consideration we prefer a classification of all species under a single genus Leptosillia, as we did not find any morphological or ecological characters diagnostic for the two lineages. In addition, if Cresporhaphis were maintained, also L. pistaciae would need another generic name, as would several other lineages now only known as endophyte isolates. It would also be impossible to generically place L. fusariospora and L. pinicola, which morphologically belong to Leptosilliaceae but for which no DNA sequence data are available. All these arguments favour a classification within a single genus.

When describing Delonicicola and Delonicicolaceae, Perera et al. (2017) also established a new order Delonicicolales. However, in their phylogenetic analyses the placement of Delonicicolaceae as sister group to Xylariales did not receive statistical support. In our phylogenetic analyses of the ITS-LSU matrix the Delonicicolaceae-Leptosilliaceae clade was embedded within Xylariales (Fig. 2), while in the multi-gene analyses a sister group relationship to the other Xylariales was highly supported (Fig. 3). However, the latter analyses contain only a small subset of Xylariales, as most xylarialean lineages lack multigene sequence data. Considering these uncertainties, we do not accept a separate order Delonicicolales here.

Morphology of the asexual morph

Pycnidial asexual morphs were produced in culture in all Leptosillia species investigated so far. The asexual morph of the genus Leptosillia is remarkable by the common presence of two morphologically different types of conidia, which are also differently produced, i.e., enteroblastic phialidic and holoblastic with sympodial proliferation. In several species, these two types have been observed within the same conidiomata (e.g., L. macrospora, L. muelleri, L. wienkampii), but apparently both types are not always produced. For instance, Pažoutová et al. (2012) observed two types in L. wienkampii, but only a single type in L. macrospora, while in our investigations it was the other way round. Therefore, it cannot be excluded that both types are also formed in species for which so far only a single type has been observed (L. acerina, L. slaptonensis). Interestingly, pycnidia were commonly produced in the isolation plates, while in several species only few or no pycnidia were formed after subculturing. Apart from the species treated in our manuscript, holoblastically formed falcate conidia have been reported by Kolařík et al. (2012) for one of the endophyte isolates (VegaE4-79 from Coffea arabica).

In our fresh collections, pycnidia were rarely seen on the natural substrate; however, as they are very similar to ascomata except for their smaller sizes, they could have been overlooked. In these, two conidial types have only been observed in L. muelleri, while in the other species either the enteroblastic phialidic (L. acerina, L. macrospora) or the holoblastic type with sympodial proliferation (L. slaptonensis, L. wienkampii) was present. So far, no asexual morphs were observed for the species only known from herbarium specimens, L. fusariospora and L. pinicola.

Ecology

Based on the association with corticolous algae on bark, most of the species here classified as Leptosillia were commonly considered to be facultatively lichenised, which may be due to the fact that they were mainly studied by lichenologists. When establishing the genus Cresporhaphis, a synonym of Leptosillia, Aguirre-Hudson (1991) described the thallus as crustose, smooth to pulverulent, greyish white and immersed in the bark but associated with an unidentified globose chlorococcoid photobiont. In the notes to the various species included, she described them as ‘probably lichenized’, and later, Calatayud & Aguirre-Hudson (2001) considered Cresporhaphis ulmi as not lichenised. Detailed investigations of numerous fresh specimens collected during the present study as well as of herbarium specimens did not confirm the presence of a lichen thallus in the former Cresporhaphis species here reclassified in Leptosillia. Although under certain environmental conditions the ascomata may be associated with chlorococcoid algae, this association is not constantly observed and entirely missing in some collections of all species examined. In addition, all species studied in fresh condition germinate and grow easily in pure culture. Therefore, the current investigations do not support that the former Cresporhaphis species are lichenised, with the exception of the recently described Cresporhaphis chibaensis, which, however, is not considered to be congeneric with the type of Cresporhaphis but conspecific with the lichen Rhaphidicyrtis trichosporella.

The publication of Pažoutová et al. (2012) shed a new light on the ecology of Leptosillia. They isolated and described two asexual morph species, Liberomyces macrosporus and L. saliciphilus, as endophytes from phloem and sapwood of various, usually symptomless broadleaf trees. In our investigations, morphology and sequence data revealed the former to be synonymous with Leptosillia macrospora and L. wienkampii, respectively. This points to a primary ecology of Leptosillia as endophytic, which is also in line with the formation of ascomata on bark of living trees, and further supported by the numerous ITS GenBank accessions of endophytes from various hosts and geographic origins which are embedded within the Leptosillia clade (Fig. 2). Therefore, this indicates that the Leptosilliaceae comprise widespread and important components of the endophyte communities of woody hosts, and they may harbour numerous undescribed species especially in understudied tropical and subtropical areas. It is interesting that Leptosillia pistaciae, a recently described canker pathogen of Pistacia vera (Vitale et al. 2018), is also embedded within the Leptosillia clade, which indicates that pathogenicity may have secondarily evolved from an endophytic lifestyle. However, it also cannot be excluded that some of the strains isolated as endophytes may actually represent latent pathogens.

Table 1

Isolates and accession numbers used in the phylogenetic analyses. Isolates/sequences in bold were isolated/sequenced in the present study.

TaxonStrain1Host2Type3Substrate/Isolation sourceCountryGenBank accession no.
References2
ITSLSUrpb1rpb2tef1tub2
Acrocordiella occultaRS9 = CBS 140500EKT949893KT949893
Amphibambusa bambusicolaMFLUCC 11–0617HKP744433KP744474
Amphisphaeria umbrinaHKUCC 994AF009805AF452029
Annulohypoxylon truncatumCBS 140778EKY610419KY610419KY624277KX376352
Anthostoma decipiensCD = CBS 133221KC774565KC774565
Anthostomella rubicolaMFLUCC 16-0479KX533455KX533456KX789493KX789494
Arthrinium arundinisCBS 133509 = NRRL 25634KF144886KF144930genome6genome6genome6genome6
Arthrinium phragmitisCBS 135458HKF144909KF144956
Arthrinium saccharicolaCBS 831.71KF144922KF144969
Barrmaelia rhamnicolaBR = CBS 142772EMF488990MF488990MK523257MF488999MF489009MF489018
Bartalinia robillardoidesCBS 122705EKJ710460KJ710438
Basiseptospora fallaxPSC = CBS 129020EJF440983JF440983
Beltrania rhombicaCPC 27482KX306749KX306778
Beltraniopsis neolitseaeCBS 137974HKX306749KJ869183
Biscogniauxia nummulariaMUCL 51395EKY610382KY610427KY624236KX271241
Cainia graminisCBS 136.62KR092793AF431949
Calceomyces lacunosusCBS 633.88HKY610397KY610476KY624293KX271265
Calosphaeria pulchellaCBS 115999NG_058734genome6GU180661FJ238421KT716476
Camillea obulariaATCC28093AF201714KY610429
Caudospora taleolaCBS 143508NMG495961MG495980MG495989MG495998MG496005
Chaetosphaeria innumeraMR 1175AF178551
Collodiscula japonicaCBS 124266JF440974JF440974
Coniocessia maximaCBS 593.74HGU553332GU553344
Coniocessia nodulisporioidesCBS 281.77IGU553333GU553352
Creosphaeria sassafrasST.MA. 14087KY610411KY610468KY624265KX271258
Cryptovalsa rabenhorstiiCreI = CBS 125574KC774567KC774567
Daldinia concentricaCBS 113277AY616683KY610434KY624243KC977274
Delonicicola siamenseMFLUCC 15-0670Delonix regiaHdry seed podsThailandNR_156345NG_059172MF158346Perera et al. (2017)
Delonicicolaceae sp.MYCO-ARIZ SNP360Phoradendron californicumliving stem tisueUSAKP335540Massimo et al. (2015)
MYCO-ARIZ SNP402Phoradendron californicumliving stem tisueUSAKP335578Massimo et al. (2015)
Diaporthe eresCBS 109767AF408350
Diatrype disciformisCBS 197.49DQ470964DQ471158DQ470915DQ471085
Entosordaria perfidiosaCBS 142773EMF488993MF488993MK523258MF489003MF489012MF489021
Eutypa lataUCR-EL1genome6genome6genome6genome6genome6genome6
Funiliomyces biseptatusCBS 100373HAY772015
Furfurella luteostiolataCE3 = CBS 143620Genista acanthocladaHbarkGreeceMK527842MK527842MK523259MK523273MK523302MK523330
Furfurella nigrescensCECallicotome villosabarkSpainMK527843MK527843MK523274MK523303MK523331
CE1 = CBS 143622Callicotome villosaHbarkSpainMK527844MK527844MK523260MK523275MK523304MK523332
CE2 = CBS 143621Chamaecytisus creticusbarkGreeceMK527845MK527845MK523276MK523305MK523333
Furfurella stromaticaCE4 = CBS 144409Genista cinereaHbarkSpainMK527846MK527846MK523261MK523277MK523306MK523334
CE5Genista cinereabarkSpainMK527847MK527847MK523278MK523335
Graphostroma platystomumCBS 270.87HG934115AY083827KY624296DQ836915HG934108
Hymenopleella hippophaëicolaLH = CBS 140410EKT949901KT949901MK523262MK523279MK523307MK523336
Hyponectria buxiUME 31430AY083834
Hypoxylon fragiformeMUCL 51264EKC477229KM186295KM186296KX271282
Idriella lunataCBS 204.56HKP859044KP858981
Juglanconis juglandinaCBS 133343KY427149KY427181KY427199KY427218KY427234
Kretzschmaria deustaCBS 163.93KC477237KY610458KY624227KX271251
Lasiosphaeria ovinaCBS 958.72AY587946genome6genome6genome6genome6
Leiosphaerella praeclaraCBS 125586JF440976JF440976
Lepteutypa fuckeliiLEF = CBS 140409NKT949902KT949902MK523263MK523280MK523308MK523337
Leptosillia acerinaCRAAcer campestrebarkAustriaMK527848MK527848MK523281MK523309MK523338
CRA1 = CBS 143939Acer campestreEbarkAustriaMK527849MK527849MK523264MK523282MK523310MK523339
CRA2Acer campestrebarkAustriaMK527850MK527850MK523283MK523311MK523340
CRA3Acer campestrebarkAustriaMK527851MK527851MK523284MK523312MK523341
Leptosillia macrosporaCCF 4028Ulmus laevisH4living bark/sapwood tissueCzech RepublicFR715522FR715522FR715509FR715498Pažoutová et al. (2012)
CRM1Quercus roburbarkGermanyMK527852MK527852MK523285MK523313MK523342
CRM2 = CBS 143627Quercus petraeaEbarkAustriaMK527853MK527853MK523265MK523286MK523314MK523343
CRM4Quercus roburbarkAustriaMK527854MK527854MK523287MK523315MK523344
CRM7Quercus roburbarkGermanyMK527855MK527855MK523288MK523316MK523345
Leptosillia muelleriCRMAcer pseudoplatanusbarkAustriaMK527856MK527856MK523289MK523317MK523346
CRM3 = CBS 143628Acer pseudoplatanusEbarkAustriaMK527857MK527857MK523266MK523290MK523318MK523347
CRM6Acer pseudoplatanusbarkAustriaMK527858MK527858MK523291MK523319MK523348
Leptosillia pistaciaeISPaVe 1958 = CBS 128196Pistacia veraHcanker tissue (parasite)ItalyMH798901MH798901MK523267MH791334MK523320MH791335Vitale et al. (2018)
ISPaVe 2105Pistacia veracanker tissue (parasite)ItalyFR681904Vitale et al. (2018)
ISPaVe 2106Pistacia veracanker tissue (parasite)ItalyFR681905Vitale et al. (2018)
Leptosillia slaptonensisCRU1 = CBS 143629Ulmus minorbarkAustriaMK527859MK527859MK523292MK523321MK523349
CRU2Ulmus minorbarkAustriaMK527860MK527860MK523293MK523350
CRU3Ulmus minorbarkAustriaMK527861MK527861
NAD = CBS 145296Ulmus minorEbarkAustriaMK527862MK527862MK523268MK523294MK523322MK523351
Leptosillia sp.A23Annona squamosaendophyteChinaEF488447unpublished
AWB8Aquilaria malaccensisliving wood tissueIndiaJX448359Premalatha & Kalra (2013)
PPM8003Calocedrus macrolepis var. formosanaliving host tissueTaiwanKX227618KX227617unpublished
PPM8004Calocedrus macrolepis var. formosanaliving host tissueTaiwanKX242164KX242162unpublished
E8520CCasearia prunifolialiving stem tisueEcuadorHQ117861unpublished
VegaE4-79Coffea arabicaliving petiole tissueUSA (Hawaii)EU009996Vega et al. (2010)
OTU173Coffea sp.leaf disk tissuePuerto RicoKT328745James et al. (2016)
INBio 573BErythroxylum macrophyllumliving host tissueCosta RicaKU204602unpublished
CXEugenia uruguayensisleaf petiole tissueUruguayKU212366García-Laviña et al. (2016)
MX17Hevea brasiliensisliving sapwood tissueMexicoJQ905737unpublished
MX194Hevea brasiliensisliving sapwood tissueMexicoJQ905738unpublished
E9226aIlex guayusaliving host tissueEcuadorJN662478unpublished
HS52living unidentified plantsliving host tissueChinaKY496833unpublished
MIB07Madhuca indicaliving bark tissueIndiaJN604095Verma et al. (2014)
clone OTU_F75_R46Nothofagus fuscaliving leavesNew ZealandMF976713Johnston et al. (2017)
E15610EPsammisia sodiroiliving stem tisueEcuadorKM266133unpublished
E11-3111Ulmus macrocarpaliving bark/sapwood tissueChinaFJ025239unpublished
M36unknownunknown (mangrove)unknownKT336540unpublished
E14625AVirola calophyllaliving stem tisueEcuadorKM265634unpublished
Leptosillia wienkampiiAK8/09Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715513FR715513Pažoutová et al. (2012)
CCF 4020Ulmus laevisH5living bark/sapwood tissueCzech RepublicFR715515FR715515Pažoutová et al. (2012)
CCF 4021Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715519FR715519Pažoutová et al. (2012)
CCF 4022Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715516FR715516Pažoutová et al. (2012)
CCF 4023Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715521FR715521Pažoutová et al. (2012)
CCF 4024Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715520FR715520Pažoutová et al. (2012)
Leptosillia wienkampii (cont.)CCF 4025Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715514FR715514Pažoutová et al. (2012)
CCF 4026Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715518FR715518Pažoutová et al. (2012)
CCF 4027Ulmus laevisliving bark/sapwood tissueCzech RepublicFR715517FR715517Pažoutová et al. (2012)
CRM5Ulmus laevisbarkAustriaMK527863MK527863MK523295MK523323
CRUUlmus glabrabarkAustriaMK527864MK527864MK523296MK523324MK523352
CRW = CBS 143630Salix fragilis var. russelianaEbarkUKMK527865MK527865MK523269MK523297MK523325MK523353
CRW1Salix fragilisbarkAustriaMK527866MK527866MK523298MK523326MK523354
CRW2Ulmus minorbarkItalyMK527867MK527867MK523299MK523327
CRW3Ulmus minorbarkAustriaMK527868MK527868
H041Salix albaliving bark/sapwood tissueCzech RepublicFR715510FR715510FR715507FR715496Pažoutová et al. (2012)
H077Salix albaliving bark/sapwood tissueCzech RepublicFR715511FR715511FR715508FR715497Pažoutová et al. (2012)
H133Salix albaliving bark/sapwood tissueCzech RepublicFR715512FR715512Pažoutová et al. (2012)
Lopadostoma gastrinumCBS 134632NKC774584KC774584
Lopadostoma turgidumCBS 133207EKC774618KC774618MK523270KC774563MF489024
Melogramma campylosporumMBU = CBS 141086JF440978JF440978
Microdochium lycopodinumCBS 125585HJF440979JF440979KP859125KP859080
Microdochium phragmitisCBS 285.71EKP859013KP858949KP859122KP859076
Obolarina dryophilaMUCL 49882GQ428316GQ428316KY624284GQ428322
Ophiostoma piliferumCBS 158.74DQ470955
Pestalotiopsis knightiaeCBS 114138HKM199310KM116227
Phlogicylindrium eucalyptorumCBS 111689KF251205KF251708
Phlogicylindrium uniformeCBS 131312HJQ044426JQ044445
Polyancora globosaCBS 118182HDQ396469DQ396466
Pseudapiospora corniPCO = CBS 140736NKT949907KT949907
Pseudoanthostomella delitescensMFLUCC 16-0477KX533451KX533452KX789491KX789490
Pseudomassaria chondrosporaCBS 125600JF440981JF440981
Pseudomassariella vexataLVE = CBS 129021EJF440977JF440977genome6genome6genome6genome6
Requienella seminudaRS12 = CBS 140502EKT949912KT949912MK523271MK523300MK523328
Robillarda sessilisCBS 114312EKR873256KR873284
Rosellinia aquilaMUCL 51703KY610392KY610460KY624285KX271253
Seiridium marginatumBLO = CBS 140403EKT949914KT949914MK523272MK523301MK523329LT853249
Strickeria kochiiC143 = CBS 140411EKT949918KT949918
Truncatella angustataICMP 7062AF405306AF382383
Vialaea insculptaDAOM 240257JX139726JX139726
Vialaea minutellaBRIP 56959KC181926KC181924
Xylaria hypoxylonCBS 122620EKY610407KY610495KY624231KX271279

1 Abbreviations: ATCC: American Type Culture Collection, Manassas, VA, USA; BRIP: Queensland Plant Pathology Herbarium, Brisbane, Australia; CBS: Culture collection of the Westerdijk Fungal Biodiversity Institute, Utrecht, The Netherlands; CCF: Culture collection of the Dept. of Botany, Charles University, Prague, Czech Republic; CPC: Culture collection of Pedro Crous, housed at CBS; DAOM: Canadian National Mycological Herbarium, Ottawa, Candada; HKUCC: The University of Hong Kong Culture Collection, Hong Kong, China; ICMP: International Collection of Microorganisms from Plants, Auckland, New Zealand; INBio: Instituto Nacional de Biodiversidad, Costa Rica; ISPaVe: Culture collection of the Consiglio per la Ricerca in Agricoltura e l’Analisi dell’Economia Agraria, Roma, Italy (CREA-DC); MFLUCC: Mae Fah Luang University Culture Collection, Chiang Rai, Thailand; MR: Culture collection of Martina Réblová, Department of Taxonomy, Institute of Botany of the Czech Academy of Sciences, Průhonice, Czech Republic; MUCL: BCCM/MUCL Agro-food & Environmental Fungal Collection, Louvain-la-Neuve, Belgium; MYCO-ARIZ: Gilbertson Mycological Herbarium, University of Arizona, Tucson, USA; NRRL: Agrigultural Research Service Culture Collection, Peoria, IL, USA; ST.MA.: Culture collection of Mark Stadler, Helmholtz-Zentrum für Infektionsforschung, Braunschweig, Germany; UCR: University of California, Riverside, USA; UME: Herbarium of the Department of Ecology and Environmental Science, Umeå University, Umeå, Sweden.

2 Hosts and References only given for GenBank sequence accessions within the Delonicicolacaeae-Leptosilliaceae clade.

3 E ex-epitype strain; H ex-holotype strain; I ex-isotype strain; N ex-neotype strain.

4 Ex-holotype strain of Liberomyces macrosporus.

5 Ex-holotype strain of Liberomyces saliciphilus.

6 Sequence retrieved from genome deposited at JGI-DOE (http://genome.jgi.doe.gov/).

Acknowledgments

We thank Genevieve E. Tocci (FH) for locating, providing details and sending the type collection of Leptosillia notha, Don Pfister (FH) for permission to investigate it, Rosella Marcucci (PAD) for access to the holotype specimen of Septoria notha, the curators of GZU and NY for loan of additional type material and Walter Till (WU) for handling the herbarium loans. The financial support by the Austrian Science Fund (FWF; project P27645-B16) to HV is gratefully acknowledged.

REFERENCES

  • Aguirre-Hudson B. 1991. A taxonomic study of the species referred to the ascomycete genus Leptorhaphis. Bulletin of the British Museum (Natural History), Botany Series 21: 85–192. [Google Scholar]
  • Aguirre-Hudson B, Farkas E, Lőkös L. 2005. New records of Leptorhaphis and other ascomycete genera from the Carpathian basin (Europe). Herzogia 18: 47–50. [Google Scholar]
  • Barr ME. 1993. Redisposition of some taxa described by J.B. Ellis. Mycotaxon 46: 45–76. [Google Scholar]
  • Berger F, Priemetzhofer F. 2000. Neue und seltene Flechten und lichenicole Pilze aus Oberösterreich, Österreich III. Herzogia 14: 59–84. [Google Scholar]
  • Berger F, Priemetzhofer F, Türk R. 1998. Neue und seltene Flechten und lichenicole Pilze aus Oberösterreich, Österreich IV. Beiträge zur Naturkunde Oberösterreichs 6: 397–416. [Google Scholar]
  • Calatayud V, Aguirre-Hudson B. 2001. Observations on the genus Cresporhaphis (Trichosphaeriaceae), with a key to the known species, and C. ulmi sp. nov. Mycological Research 105: 122–126. [Google Scholar]
  • Cannon PF. 1997. Two new genera of Ascomycota, and other new or interesting fungi from Slapton Ley National Nature Reserve and its environs. Systema Ascomycetum 15: 121–138. [Google Scholar]
  • Carbone I, Kohn LM. 1999. A method for designing primer sets for speciation studies in filamentous ascomycetes. Mycologia 91: 553–556. [Google Scholar]
  • De Hoog GS, Gerrits van den Ende AHG. 1998. Molecular diagnostics of clinical strains of filamentous basidiomycetes. Mycoses 41: 183–189. [Abstract] [Google Scholar]
  • Diedicke H. 1915. Pilze VII. Kryptogamenflora der Mark Brandenburg und angrenzender Gebiete 9: 1–962. [Google Scholar]
  • García-Laviña CX, Bettucci L, Tiscornia S. 2016. Fungal communities associated with Eugenia uruguayensis (Myrtaceae) leaf litter in early stages of decomposition in Uruguay. Sydowia 68: 139–150. [Google Scholar]
  • Ekman S, Aguirre-Hudson B, Arup U, et al. 2013. Rhaphidicyrtis trichosporella new to Sweden. Graphis Scripta 25: 6–11. [Google Scholar]
  • Ellis JB, Everhart BM. 1885. New species of fungi. Journal of Mycology 1: 88–93. [Google Scholar]
  • Eriksson OE, Hawksworth DL. 1987. Notes on ascomycete systematics. Nos 225–463. Systema Ascomycetum 6: 111–165. [Google Scholar]
  • Hall TA. 1999. BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95–98. [Google Scholar]
  • Harada H. 2014. Cresporhaphis chibaensis sp. nov. (lichenized Ascomycota, Trichosphaeriaceae) from Chiba-ken, central Japan. Lichenology 12: 31–36. [Google Scholar]
  • Hirayama K, Tanaka K. 2011. Taxonomic revision of Lophiostoma and Lophiotrema based on reevaluation of morphological characters and molecular analyses. Mycoscience 52: 401–412. [Google Scholar]
  • Hofstetter V, Miądlikowska J, Kauff F, et al. 2007. Phylogenetic comparison of protein-coding versus ribosomal RNA-coding sequence data: A case study of the Lecanoromycetes (Ascomycota). Molecular Phylogenetics and Evolution 44: 412–426. [Abstract] [Google Scholar]
  • Höhnel F. von. 1928. Über Septoria notha Sacc. Mitteilungen aus dem Botanischen Institut der Technischen Hochschule in Wien 5: 108–112. [Google Scholar]
  • Huhndorf S. 1992. Systematics of Leptosphaeria species found on the Rosaceae. Illinois Natural History Survey Bulletin 34: 479–534. [Google Scholar]
  • Jaklitsch WM. 2009. European species of Hypocrea Part I. The greenspored species. Studies in Mycology 63: 1–91. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jaklitsch WM, Baral HO, Lücking R, et al. 2016a. Syllabus of plant families – A. Engler’s Syllabus der Pflanzenfamilien Part 1/2: Ascomycota, 13th edn Borntraeger, Stuttgart. [Google Scholar]
  • Jaklitsch WM, Gardiennet A, Voglmayr H. 2016b. Resolution of morphology-based taxonomic delusions: Acrocordiella, Basiseptospora, Blogiascospora, Clypeosphaeria, Hymenopleella, Lepteutypa, Pseudapiospora, Requienella, Seiridium and Strickeria. Persoonia 37: 82–105. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jaklitsch WM, Komon M, Kubicek CP, et al. 2005. Hypocrea voglmayrii sp. nov. from the Austrian Alps represents a new phylogenetic clade in Hypocrea/Trichoderma. Mycologia 97: 1365–1378. [Abstract] [Google Scholar]
  • Jaklitsch WM, Stadler M, Voglmayr H. 2012. Blue pigment in Hypocrea caerulescens sp. nov. and two additional new species in sect. Trichoderma. Mycologia 104: 925–941. [Europe PMC free article] [Abstract] [Google Scholar]
  • Jaklitsch WM, Voglmayr H. 2012. Phylogenetic relationships of five genera of Xylariales and Rosasphaeria gen. nov. (Hypocreales). Fungal Diversity 52: 75–98. [Google Scholar]
  • Jaklitsch WM, Voglmayr H. 2014. Persistent hamathecial threads in the Nectriaceae, Hypocreales: Thyronectria revisited and re-instated. Persoonia 33: 182–211. [Europe PMC free article] [Abstract] [Google Scholar]
  • James TY, Marino JA, Perfecto I, et al. 2016. Identification of putative coffee rust mycoparasites via single-molecule DNA sequencing of infected pustules. Applied and Environmental Microbiology 82: 631–639. [Europe PMC free article] [Abstract] [Google Scholar]
  • Johnston PR, Park D, Smissen RD. 2017. Comparing diversity of fungi from living leaves using culturing and high-throughput environmental sequencing. Mycologia 109 (4): 643–654. [Abstract] [Google Scholar]
  • Kirk PM, Cannon PF, Minter DW, et al. 2008. Ainsworth & Bisby’s dictionary of the fungi, 10th edn. CABI, Wallingford. [Google Scholar]
  • Kolařík M, Stodůlková E, Kubátová A, et al. 2012. New endophytic species from the phloem of broadleaf trees. In: Schneider C, Leifert C, editor; , Feldmann F, editor. (eds), Endophytes for plant protection: the state of the art: 47–52. Deutsche Phytomedizinische Gesellschaft, Braunschweig. [Google Scholar]
  • Liu YL, Whelen S, Hall BD. 1999. Phylogenetic relationships among ascomycetes: evidence from an RNA polymerase II subunit. Molecular Biology and Evolution 16: 1799–1808. [Abstract] [Google Scholar]
  • Lücking R, Hodkinson BP, Leavitt SD. 2017. Corrections and amendments to the 2016 classification of lichenized fungi in the Ascomycota and Basidiomycota. The Bryologist 120: 58–69. [Google Scholar]
  • Massimo NC, Nandi Devan MM, Arendt KR, et al. 2015. Fungal endophytes in aboveground tissues of desert plants: infrequent in culture, but highly diverse and distinctive symbionts. Microbial Ecology 70: 61–76. [Europe PMC free article] [Abstract] [Google Scholar]
  • Motiejûnaitë J. 2007. Lichenized, lichenicolous and allied fungi of Žemaitija National Park (Lithuania). Herzogia 20: 179–188. [Google Scholar]
  • Müller K. 2004. PRAP – calculation of Bremer support for large data sets. Molecular Phylogenetics and Evolution 31: 780–782. [Abstract] [Google Scholar]
  • O’Donnell K, Cigelnik E. 1997. Two divergent intragenomic rDNA ITS2 types within a monophyletic lineage of the fungus Fusarium are nonorthologous. Molecular Phylogenetics and Evolution 7: 103–116. [Abstract] [Google Scholar]
  • Otte V, Wagner HG, Fürstenow J, et al. 2017. Bemerkenswerte Flechtenfunde aus Brandenburg XIV. Verhandlungen des Botanischen Vereins von Berlin und Brandenburg 149: 153–171. [Google Scholar]
  • Pažoutová S, Šrůtka P, Holuša J, et al. 2012. Liberomyces gen. nov. with two new species of endophytic coelomycetes from broadleaf trees. Mycologia 104: 198–210. [Abstract] [Google Scholar]
  • Perera RH, Maharachchikumbura SS, Jones EG, et al. 2017. Delonicicola siamense gen. & sp. nov. (Delonicicolaceae fam. nov., Delonicicolales ord. nov.), a saprobic species from Delonix regia seed pods. Cryptogamie, Mycologie 38: 321–340. [Google Scholar]
  • Premalatha K, Kalra A. 2013. Molecular phylogenetic identification of endophytic fungi isolated from resinous and healthy wood of Aquilaria malaccensis, a red listed and highly exploited medicinal tree. Fungal Ecology 6: 205–211. [Google Scholar]
  • Rehm H. 1913. Ascomycetes exs. Fasc. 52. Annales Mycologici 11: 166–171. [Google Scholar]
  • Rehner SA, Buckley E. 2005. A Beauveria phylogeny inferred from nuclear ITS and EF1-α sequences: evidence for cryptic diversification and links to Cordyceps teleomorphs. Mycologia 97: 84–98. [Abstract] [Google Scholar]
  • Saccardo PA. 1880. Fungi Gallici lecti a cl. viris P. Brunaud, C.C. Gillet, Abb. Letendre, A. Malbranche, J. Therry vel editi in Mycotheca Gallica C. Roumeguèri. Series III. Michelia 2: 302–376. [Google Scholar]
  • Silvestro D, Michalak I. 2012. raxmlGUI: a graphical front-end for RAxML. Organisms Diversity & Evolution 12: 335–337. [Google Scholar]
  • Stamatakis E. 2006. RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688–2690. [Abstract] [Google Scholar]
  • Stiller JW, Hall BD. 1997. The origin of red algae: implications for plastid evolution. Proceedings of the National Academy of Sciences, USA 94: 4520–4525. [Europe PMC free article] [Abstract] [Google Scholar]
  • Sun X, Ding Q, Hyde KD, et al. 2012. Community structure and preference of endophytic fungi of three woody plants in a mixed forest. Fungal Ecology 5: 624–632. [Google Scholar]
  • Sung GH, Sung JM, Hywel-Jones NL, et al. 2007. A multigene phylogeny of Clavicipitaceae (Ascomycota, Fungi): identification of localised incongruence using a combinational bootstrap approach. Molecular Phylogenetics and Evolution 44: 1204–1223. [Abstract] [Google Scholar]
  • Swofford DL. 2002. PAUP* 4.0b10: phylogenetic analysis using parsimony (*and other methods). Sinauer Associates, Sunderland, Massachusetts. [Google Scholar]
  • Thiers B. 2018. Index Herbariorum: A global directory of public herbaria and associated staff. New York Botanical Garden’s Virtual Herbarium. http://sweetgum.nybg.org/ih/. [Google Scholar]
  • Triebel D, Scholz P. 2018. IndExs – Index of Exsiccatae. Botanische Staats-sammlung München. http://indexs.botanischestaatssammlung.de/. [Google Scholar]
  • Vega FE, Simpkins A, Aime C, et al. 2010. Fungal endophyte diversity in coffee plants from Colombia, Hawai’i, Mexico and Puerto Rico. Fungal Ecology 3: 122–138. [Google Scholar]
  • Verma SK, Gond SK, Mishra A, et al. 2014. Impact of environmental variables on the isolation, diversity and antibacterial activity of endophytic fungal communities from Madhuca indica Gmel. at different locations in India. Annals of Microbiology 64: 721–734. [Google Scholar]
  • 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. [Europe PMC free article] [Abstract] [Google Scholar]
  • Vitale S, Aiello D, Guarnaccia V, et al. 2018. Liberomyces pistaciae sp. nov., the causal agent of pistachio cankers and decline in Italy. Mycokeys 40: 29–51. [Europe PMC free article] [Abstract] [Google Scholar]
  • Voglmayr H, Akulov OY, Jaklitsch WM. 2016a. Reassessment of Allantonectria, phylogenetic position of Thyronectroidea, and Thyronectria caraganae sp. nov. Mycological Progress 15: 921. [Europe PMC free article] [Abstract] [Google Scholar]
  • Voglmayr H, Friebes G, Gardiennet A, et al. 2018. Barrmaelia and Entosordaria in Barrmaeliaceae (fam. nov., Xylariales) and critical notes on Anthostomella-like genera based on multi-gene phylogenies. Mycological Progress 17: 155–177. [Europe PMC free article] [Abstract] [Google Scholar]
  • Voglmayr H, Gardiennet A, Jaklitsch WM. 2016b. Asterodiscus and Stigmatodiscus, two new apothecial dothideomycete genera and the new order Stigmatodiscales. Fungal Diversity 80: 271–284. [Europe PMC free article] [Abstract] [Google Scholar]
  • Voglmayr H, Jaklitsch WM. 2008. Prosthecium species with Stegonsporium anamorphs on Acer. Mycological Research 112: 885–905. [Abstract] [Google Scholar]
  • Voglmayr H, Jaklitsch WM. 2011. Molecular data reveal high host specificity in the phylogenetically isolated genus Massaria (Ascomycota, Massariaceae). Fungal Diversity 46: 133–170. [Google Scholar]
  • Voglmayr H, Jaklitsch WM, Mohammadi H, et al. 2019. The genus Juglanconis (Diaporthales) on Pterocarya. Mycologial Progress 18: 425–437. [Europe PMC free article] [Abstract] [Google Scholar]
  • Voglmayr H, Mehrabi M. 2018. Molecular phylogeny and a new Iranian species of Caudospora (Sydowiellaceae, Diaporthales). Sydowia 70: 67–80. [Europe PMC free article] [Abstract] [Google Scholar]
  • Voglmayr H, Rossman AY, Castlebury LA, et al. 2012. Multigene phylogeny and taxonomy of the genus Melanconiella (Diaporthales). Fungal Diversity 57: 1–44. [Google Scholar]
  • Voglmayr H, Yule C. 2006. Polyancora globosa gen. et sp. nov., an aero-aquatic fungus from Malaysian peat swamp forests. Mycological Research 110: 1242–1252. [Abstract] [Google Scholar]
  • Wehmeyer LE. 1933. The genus Diaporthe Nitschke and its segregates. University of Michigan Studies Scientific Series 9: 1–349. [Google Scholar]
  • Wendt L, Sir EB, Kuhnert E, et al. 2018. Resurrection and emendation of the Hypoxylaceae, recognised from a multigene phylogeny of the Xylariales. Mycological Progress 17: 115–154. [Google Scholar]
  • Werle E, Schneider C, Renner M, et al. 1994. Convenient single-step, one tube purification of PCR products for direct sequencing. Nucleic Acids Research 22: 4354–4355. [Europe PMC free article] [Abstract] [Google Scholar]
  • White TJ, Bruns T, Lee S, et al. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, editor; , Gelfand DH, editor; , Sninsky JJ, editor. , et al. (eds), PCR protocols: A guide to methods and applications: 315–322. Academic Press, San Diego. [Google Scholar]
  • Wiens JJ. 1998. Combining datasets with different phylogenetic histories. Systematic Biology 47: 568–581. [Abstract] [Google Scholar]

Articles from Persoonia : Molecular Phylogeny and Evolution of Fungi are provided here courtesy of Naturalis Biodiversity Center & Centraalbureau voor Schimmelcultures

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/67197318
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/67197318

Article citations


Go to all (11) article citations

Data 


Data behind the article

This data has been text mined from the article, or deposited into data resources.

Similar Articles 


To arrive at the top five similar articles we use a word-weighted algorithm to compare words from the Title and Abstract of each citation.


Funding 


Funders who supported this work.

Austrian Science Fund FWF (1)