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 


Although the genus Pseudocercospora has a worldwide distribution, it is especially diverse in tropical and subtropical countries. Species of this genus are associated with a wide range of plant species, including several economically relevant hosts. Preliminary studies of cercosporoid fungi from Brazil allocated most taxa to Cercospora, but with the progressive refinement of the taxonomy of cercosporoid fungi, many species were relocated to or described in Pseudocercospora. Initially, species identification relied mostly on morphological features, and thus no cultures were preserved for later phylogenetic comparisons. In this study, a total of 27 Pseudocercospora spp. were collected, cultured, and subjected to a multigene analysis. Four genomic regions (LSU, ITS, tef1 and actA) were amplified and sequenced. A multigene Bayesian analysis was performed on the combined ITS, actA and tef1 sequence alignment. Our results based on DNA phylogeny, integrated with ecology, morphology and cultural characteristics revealed a rich diversity of Pseudocercospora species in Brazil. Twelve taxa were newly described, namely P. aeschynomenicola, P. diplusodonii, P. emmotunicola, P. manihotii, P. perae, P. planaltinensis, P. pothomorphes, P. sennae-multijugae, P. solani-pseudocapsicicola, P. vassobiae, P. wulffiae and P. xylopiae. Additionally, eight epitype specimens were designated, three species newly reported, and several new host records linked to known Pseudocercospora spp.

Free full text 


Logo of persooniaLink to Publisher's site
Persoonia. 2016 Dec; 37: 142–172.
Published online 2016 Feb 12. https://doi.org/10.3767/003158516X691078
PMCID: PMC5315286
PMID: 28232763

Exploring fungal mega-diversity: Pseudocercospora from Brazil

Abstract

Although the genus Pseudocercospora has a worldwide distribution, it is especially diverse in tropical and subtropical countries. Species of this genus are associated with a wide range of plant species, including several economically relevant hosts. Preliminary studies of cercosporoid fungi from Brazil allocated most taxa to Cercospora, but with the progressive refinement of the taxonomy of cercosporoid fungi, many species were relocated to or described in Pseudocercospora. Initially, species identification relied mostly on morphological features, and thus no cultures were preserved for later phylogenetic comparisons. In this study, a total of 27 Pseudocercospora spp. were collected, cultured, and subjected to a multigene analysis. Four genomic regions (LSU, ITS, tef1 and actA) were amplified and sequenced. A multigene Bayesian analysis was performed on the combined ITS, actA and tef1 sequence alignment. Our results based on DNA phylogeny, integrated with ecology, morphology and cultural characteristics revealed a rich diversity of Pseudocercospora species in Brazil. Twelve taxa were newly described, namely P. aeschynomenicola, P. diplusodonii, P. emmotunicola, P. manihotii, P. perae, P. planaltinensis, P. pothomorphes, P. sennae-multijugae, P. solani-pseudocapsicicola, P. vassobiae, P. wulffiae and P. xylopiae. Additionally, eight epitype specimens were designated, three species newly reported, and several new host records linked to known Pseudocercospora spp.

Keywords: biodiversity, Capnodiales, cercosporoid, Dothideomycetes, multigene phylogeny, Mycosphaerellaceae, plant pathogen, systematics

INTRODUCTION

The genus Pseudocercospora was described by Spegazzini (1910) with P. vitis as type species. Pseudocercospora belongs to the Mycosphaerellaceae (Capnodiales, Dothideomycetes), and several species have mycosphaerella-like sexual morphs (Crous et al. 2013a). With the amendment of Article 59 of the International Code of Nomenclature for algae, fungi and plants (ICN), a single generic name is now used for Pseudocercospora spp. (Hawksworth et al. 2011, Wingfield et al. 2012, Crous et al. 2015). This has led to changes in the holomorphic name of some important fungal pathogens such as the etiological agent of South American leaf blight of rubber, P. ulei (≡ Microcyclus ulei, Hora Júnior et al. 2014) and leaf and fruit spot of pistachio, P. pistacina (≡ Septoria pistacina, Crous et al. 2013b).

Pseudocercospora is a cosmopolitan genus of phytopathogenic fungi that is associated with a wide range of plant species, including several economically relevant hosts (Crous et al. 2013a, Bakhshi et al. 2014). Furthermore, some of the species, e.g. P. angolensis and P. fijiensis are regarded as being of quarantine significance (Churchill 2011, Crous et al. 2013a).

Several important plant pathogenic Pseudocercospora spp. are known from Brazil. Besides P. fijiensis (black leaf streak of Musa), P. griseola (angular leaf spot of Phaseolus vulgaris) and P. ulei (South American leaf blight of Hevea brasiliensis), other economically relevant species include P. abelmoschi (leaf spot of Abelmoschus esculentus), P. anacardii (leaf spot of Anacardium occidentale), P. bixae (leaf spot of Bixa orellana), P. cruenta (leaf spot of Vigna unguiculata ssp. sesquipedalis), P. kaki (leaf spot of Diospyros kaki), P. musae (yellow Sigatoka of Musa), P. paraguayensis (leaf spot of Eucalyptus) and P. vitis (leaf spot of Vitis) (Chupp 1954, Crous & Braun 2003, Kimati et al. 2005, Hunter et al. 2006, Crous et al. 2006, 2013a, Arzanlou et al. 2007, 2008, 2010, Churchill 2011, Braun et al. 2013, Kirschner 2014).

Among the Pseudocercospora spp. described from Brazil, several have also been recognised as having potential for use as biological control agents of invasive weeds. For example, P. borreriae could be used for the biocontrol of Mitracarpus hirtus (Pereira & Barreto 2005), P. cryptostegiae-madagascariensis for Cryptostegia madagascariensis (Silva et al. 2008), P. palicourea for Palicourea marcgravii (Pereira & Barreto 2006), P. pereskiae as a classical biocontrol agent against Pereskia aculeata (Pereira & Barreto 2007) and P. subsynnematosa for Tibouchina herbacea (Parreira et al. 2014).

Surveys of the biodiversity of Brazilian cercosporoid fungi in native and cultivated plants date back as far as 1929, when A.S. Muller collected and described many species from the State of Minas Gerais (Muller & Chupp 1934). Later, A.P. Viégas dedicated particular attention to this group of fungi in Brazil, describing more than 90 species in a single publication (Viégas 1945). A.C. Batista also investigated and described several additional species (Batista et al. 1960). Some publications have dealt with the re-examination of the species described by Viégas (Crous et al. 1997, 1999); these studies resulted in several cercosporoid fungi being allocated to other genera, including Pseudocercospora. During the last decades numerous Pseudocercospora spp. have been described from Brazilian biomes such as the Caatinga (semi-arid) (Braun et al. 1999, Braun & Freire 2002, 2004, 2006), the Atlantic rainforest - Mata Atlântica (Rocha et al. 2008, Soares & Barreto 2008, Parreira et al. 2014), and especially from the Cerrado (Furlanetto & Dianese 1999, Hernández-Gutiérrez & Dianese 2009, 2014, Hernández-Gutiérrez et al. 2014). With a few exceptions (e.g., Crous et al. 2013a, Rocha et al. 2013, Parreira et al. 2014), publications dealing with Brazilian Pseudocercospora spp. lack molecular data and rely solely on morphological characteristics, making phylogenetic comparisons to species from other countries impossible. The genus Pseudocercospora accommodates several synnematal and non-synnematal cercospora-like species that produce pigmented conidiophores and conidia with unthickened (or slightly thickened), non-darkened conidial scars and hila (Deighton 1976, Braun 1995). However, the application of DNA phylogenetic analyses to species in the Mycosphaerella complex (Stewart et al. 1999, Crous et al. 2000, 2001) demonstrated that Pseudocercospora is heterogeneous. Indeed, Crous et al. (2001) regarded the unthickened (or slightly thickened) conidial scars to be a synapomorphy shared among several cercosporoid genera. Recently, multigene DNA analyses revealed that the morphological characteristics previously ascribed solely to Pseudocercospora evolved more than once within the Mycosphaerellaceae (Frank et al. 2010, Crous et al. 2013a).

Pseudocercospora s.str. was circumscribed as having species with conidiophores that are solitary, fasciculate, synnematal, or arranged in sporodochia, giving rise to conidia that are pigmented with unthickened or slightly thickened and darkened scars (Braun et al. 2013, Crous et al. 2013a). However, some species with characteristics that are not typical of Pseudocercospora s.str. were placed in Pseudocercospora until more sequences became available, and the clades these species belong to become better resolved (Minnis et al. 2011, Crous et al. 2013a). Additionally, Crous et al. (2013b) recently included Septoria pistacina, which only has pycnidial conidiomata, in Pseudocercospora s.str., highlighting the morphological plasticity occurring within this genus. Hora Júnior et al. (2014) employed multigene DNA data to reconstruct the molecular phylogeny of the fungus causing South American leaf blight of rubber (P. ulei), and showed that it was firmly located within Pseudocercospora s.str. Moreover, the associated conidiomatal Aposphaeria morph was shown to possess a spermatial function. All of these cases suggest that the present generic circumscription of Pseudocercospora s.str. has changed with time as more DNA phylogenetic data became available (Crous et al. 2013a, Bakhshi et al. 2014, Nguanhom et al. 2015), and may continue to be further refined in future years.

The aim of the present study was therefore to initiate a re-evaluation of Pseudocercospora spp. occurring in Brazil, based on a combination of morphological, cultural and molecular data using the Consolidated Species Concept proposed by Quaedvlieg et al. (2014). Whenever possible, epitypes for known species were designated and DNA sequences deposited in NCBIs GenBank nucleotide database.

MATERIAL AND METHODS

Sample collection and isolates

Surveys were conducted between 2013 and 2014 in the Reserva Florestal Mata do Paraíso (Viçosa, Minas Gerais), the campus of the Universidade Federal de Viçosa (Viçosa, Minas Gerais) and neighbouring areas in the municipality of Viçosa, Floresta Nacional de Paraopeba (Paraopeba, Minas Gerais), Estação Ecológica de Águas Emendadas (Distrito Federal, Brasília), Parque Nacional da Chapada dos Veadeiros (Alto Paraíso de Goiás, Goiás), Instituto Agronômico de Campinas (Campinas, São Paulo), municipality of Lavras (Minas Gerais) and Nova Friburgo (Rio de Janeiro). Samples with cercosporoid leaf spot symptoms were collected, dried in a plant press, and taken to the laboratory. Fungal isolations were performed by direct transfer of fungal structures onto plates containing vegetable broth agar (VBA) as described by Pereira et al. (2003) or 2 % potato-dextrose agar (PDA; HiMedia). Axenic cultures were preserved on potato-carrot agar (PCA) slants or on silica gel and were deposited in the culture collection of the Universidade Federal de Viçosa, Coleção Oswaldo Almeida Drummond (COAD). Representative specimens were deposited at the Fungarium of the Universidade Federal de Viçosa (VIC) and CBS Fungarium (CBS H).

Morphology

Taxonomic descriptions were based on observations of fungal structures present on plant specimens. Samples with cercosporoid leaf spot symptoms were viewed under a Nikon® SMZ 1 000 dissecting microscope. Morphological structures were removed from the lesions with a sterile dissecting needle and mounted in clear lactic acid. Measurements were made at 1 000× magnification using a Carl Zeiss® Axioskop 2 compound microscope. High-resolution photographic images of diseased material, leaf lesions and microscopic fungal structures were captured with a Nikon® digital sight DS-fi1 high definition colour camera. Images of fungal structures were captured and measurements were taken using the Nikon® software NIS-Elements v. 2.34. Adobe Photoshop CS5 was used for the final editing of the acquired images and photographic preparations. Culture descriptions were based on observations of colonies formed in plates containing 2 % malt extract agar (MEA) following incubation at 24 °C for 2–4 wk in the dark in duplicate. Colour terminology followed Rayner (1970). Nomenclatural novelties and descriptions were deposited in MycoBank (www.MycoBank.org, Crous et al. 2004).

DNA isolation, PCR amplification and sequencing

Genomic DNA was extracted from mycelium growing on MEA plates at 25 °C for up to 4 wk depending on their growth rate, using the CTAB extraction protocol as outlined by Crous et al. (2009). Four nuclear gene regions were targeted for Polymerase Chain Reaction (PCR) amplification and subsequent sequencing. The Internal Transcribed Spacer (ITS) region was amplified using primers ITS-5 and ITS-4 (White et al. 1990), the Large Subunit (28S nrDNA, LSU) with LR0R (Rehner & Samuels 1994) and LR5 (Vilgalys & Hester 1990), the translation elongation factor 1-alpha (tef1) with EF1-728F (Carbone & Kohn 1999) and EF-2 (O’Donnell et al. 1998) and actin (actA) with ACT-512F and ACT-783R (Carbone & Kohn 1999). PCR mixtures included the following ingredients for each 12.5 μL reaction: 10–20 ng of template DNA, 1× PCR buffer, 0.63 μL DMSO (99.9 %), 1.5 mM MgCl2, 0.5 μM of each primer, 0.25 mM of each dNTP, 1.0 U BioTaq® DNA polymerase (Bioline GmbH Luckenwalde, Germany). The PCRs were carried out with a MyCyclerTM Thermal Cycler (Bio-Rad Laboratories B.V., Veenendal, The Netherlands). Conditions for the PCR amplification consisted of an initial denaturation at 95 °C for 5 min; followed by 40 cycles of denaturation at 95 °C for 30 s; annealing at 52 °C for ITS and LSU, 54 °C for tef1 or 55 °C for actA for 30 s; extension at 72 °C for 1 min and a final extension step at 72 °C for 7 min. Following PCR amplification, amplicons were visualised on 1 % agarose gels to check for product size and purity. The PCR products were sequenced in both directions using the PCR primers and the BigDye® Terminator Cycle Sequencing Kit v. 3.1 (Applied Biosystems, Foster City, CA, USA), following the protocol of the manufacturer. DNA sequencing amplicons were purified through Sephadex® G-50 Superfine columns (Sigma Aldrich, St. Louis, MO) in MultiScreen HV plates (Millipore, Billerica, MA). Purified sequence reactions were run on an ABI Prism 3730xl DNA Analyser (Life Technologies, Carlsbad, CA, USA). The consensus sequences were generated using MEGA v. 6.0.6 (Molecular Evolutionary Genetics Analyses) (Tamura et al. 2013). All sequences were checked manually, and nucleotides with ambiguous positions were clarified using both primer direction sequences.

Phylogenetic analyses

Consensus sequences were compared against NCBIs GenBank nucleotide database using their megaBLAST algorithm. The most similar sequences were downloaded in FASTA format and the sequence datasets for the four genomic loci were aligned individually using the MAFFT v. 7 online portal (http://mafft.cbrc.jp/alignment/server/index.html) (Katoh & Standley 2013). In addition, the combined sequence alignment of Crous et al. (2013a) was downloaded from TreeBASE (Study S12805) and used as an initial reference alignment for species identification. Resulting sequence alignments were manually checked and adjusted in MEGA v. 6.06 and were concatenated with Mesquite v. 2.75 (Maddison & Maddison 2011). A phylogenetic re-construction was conducted on the aligned LSU dataset to determine generic relationships. For the LSU alignment, MrModeltest v. 2.2 (Nylander 2004) was used to select the optimal model of nucleotide substitution prior to the Bayesian Inference (BI) analysis using MrBayes v. 3.2.1 (Ronquist & Huelsenbeck 2003). The general time-reversible model of evolution (Rodriguez et al. 1990), including estimation of invariable sites and assuming a discrete gamma distribution with six rate categories (GTR+I+G) was used. Subsequently, a species-level phylogeny was derived from a concatenated ITS (alignment position 1–482), actA (alignment position 510–714) and tef1 (alignment position 720–1270) dataset using MrModeltest v. 2.2 to select the optimal model of nucleotide substitution for each locus based on the Akaike Information Criterion prior to the BI analysis. Gaps longer than 10 nucleotides were excluded from the analyses (tef1 only, see alignment in TreeBASE). The results of MrModeltest recommended a HKY85 model for tef1, and a GTR model for ITS and actA. For actA and tef1, a dirichlet (1,1,1,1) state frequency distribution was set and for ITS a fixed (equal) state frequency distribution, and for all three loci an inverse gamma distributed rate variation. Two sets of four MCMC (Markov Chain Monte Carlo) chains were run simultaneously, starting from random trees and lasting until the critical value for the topological convergence diagnostic reached 0.01. Trees were sampled every 1 000 generations and the first 25 % of the trees were discarded as the burn-in phase for each analysis and posterior probabilities (Rannala & Yang 1996) were determined from the remaining trees and are presented on the left of each node (Fig. 1). Sequences derived from this study were deposited in GenBank (http://www.ncbi.nlm.nih.gov/genbank) (Table 1), the alignments and trees in TreeBASE (www.treebase.org) (S17995). A parsimony analysis was also performed on the combined alignment as described by Arzanlou et al. (2008). The resulting phylogenetic tree was printed with Geneious v. 7.1.8 (http://www.geneious.com, Kearse et al. 2012), and the layout of the tree for publication was carried out using Adobe Illustrator v. CS5.

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g001a.jpg
An external file that holds a picture, illustration, etc.
Object name is per-37-142-g001b.jpg

The Bayesian phylogenetic tree inferred from DNA sequence data from the multigene alignment (ITS, actA and tef1) of Pseudocercospora species. Species from Brazil are in bold face and in coloured blocks with clade numbers for reference in the species notes. Novel species are indicated in red colour and the type status of strains are indicated next to the culture collection number. Bayesian posterior probabilities (BPP, > 0.60) and parsimony bootstrap support (PBS, > 60) values are indicated at the nodes (BPP/PBS). Thickened black branches represent nodes which are fully supported in both analyses (BPP = 1.00 / PBS = 100), while thickened blue branches were highly supported in both analyses (BPP = > 0.94 / PBS = > 94). The tree was rooted to Passalora eucalypti CBS 111318.

Table 1

Collection details and GenBank accession numbers of isolates included in this study.

SpeciesCulture accession numbers1CollectorHostFamilyCountryGenBank accession numbers2
LSUITStef1actA
Passalora eucalyptiCBS 111318; CPC 1457 (ex-type)P.W. CrousEucalyptus salignaMyrtaceaeBrazilGU253860GU269845GU384558GU320548
Pseudocercospora acericolaCBS 122279R. KirschnerAcer albopurpurascensAceraceaeTaiwanGU253699GU269650GU384368GU320358
P. aeschynomenicolaCPC 25227; COAD 1972 (ex-type)M. SilvaAeschynomene falcataFabaceaeBrazilKT290173KT290146KT290200KT313501
P. angolensisCBS 112933; CPC 4118M.C. PretoriusCitrus sp.RutaceaeZimbabweGU214470AY260063/GU269836GU384548JQ325010
CBS 149.53 (ex-type)T. de Carvalho & O. MendesCitrus sinensisRutaceaeAngolaJQ324941JQ324975JQ324988JQ325011
P. assamensisCBS 122467 (ex-type)I. BuddenhagenMusa cultivarMusaceaeIndiaGU253705GU269656GU384374GU320364
P. atromarginalisCBS 114640C.F. HillSolanum sp.SolanaceaeNew ZealandGU253706GU269658GU384376GU320365
CBS 132010; CPC 11372H.D. ShinSolanum nigrumSolanaceaeSouth KoreaGU214671GU269657GU384375
CPC 25230; COAD 1975M. SilvaSolanum americanumSolanaceaeBrazilKT290176KT290149KT290203KT313504
P. basitruncataCBS 114664; CPC 1202 (ex-type)M.J. WingfieldEucalyptus grandisMyrtaceaeColombiaGU253710/DQ204759DQ267600/GU269662DQ211675DQ147622
P. bixaeCPC 25244; COAD 1563 (ex-epitype)R.W. BarretoBixa orellanaBixaceaeBrazilKT290180KT290153KT290207KT313508
P. boehmeriigenaCPC 25243; COAD 1562R.W. BarretoBohemia niveaUrticaceaeBrazilKT290179KT290152KT290206KT313507
P. catalpigenaMUCC 743C. Nakashima & I. ArakiCatalpa ovataBignoniaceaeJapanGU253731GU269690GU384406GU320395
P. cercidis-chinensisCBS 132109; CPC 14481 (ex-epitype)H.D. ShinCercis chinensisFabaceaeSouth KoreaGU253718GU269670GU384387GU320376
P. chamaecristaeCPC 25228; COAD 1973 (ex-epitype)M. SilvaChamaecrista sp.FabaceaeBrazilKT290174KT290147KT290201KT313502
P. chengtuensisCBS 131924; CPC 10696H.D. ShinLycium chinenseSolanaceaeSouth KoreaJQ324942GU269673GU384390GU320379
P. contrariaCBS 132108; CPC 14714H.D. ShinDioscorea quinquelobaDioscoreaceaeSouth KoreaJQ324945GU269677GU384394GU320385
P. cordianaCBS 114685; CPC 2552 (ex-type)P.W. Crous & R.L. BenchimolCordia goeldianaBoraginaceaeBrazilGU214472AF362054/GU269681GU384398GU320387
P. corylopsidisMUCC 874T. Kobayashi & C. NakashimaHamamelis japonicaHamamelidaceaeJapanGU253757GU269721GU384437GU320425
MUCC 908 (ex-epitype)C. Nakashima & E. ImaizumiCorylopsis spicataHamamelidaceaeJapanGU253727GU269684GU384401GU320390
P. cotoneastriMUCC 876T. Kobayashi & C. NakashimaCotoneaster salicifoliusRosaceaeJapanGU253728GU269685GU384402GU320391
P. crousiiCBS 119487C.F. HillEucalyptus sp.MyrtaceaeNew ZealandGU253729GU269686GU384403GU320392
P. cruentaCBS 132021; CPC 10846H. BookerVigna sp.FabaceaeTrinidadGU214673GU269688GU384404JQ325012
P. diplusodoniiCPC 25179; COAD 1476 (ex-type)M. SilvaDiplusodon sp.LythraceaeBrazilKT290162KT290135KT290189KT313490
P. elaeocarpiMUCC 925C. NakashimaElaeocarpus sp.ElaeocarpaceaeJapanGU253740GU269701GU384417GU320405
P. emmotunicolaCPC 25187; COAD 1491 (ex-type)M. SilvaEmmotum nitensIcacinaceaeBrazil KT290163KT290136KT290190KT313491
P. euphorbiacearumCPC 25222; COAD 1537M. SilvaDalechampia sp.EuphorbiaceaeBrazil KT290172KT290145KT290199KT313503
P. eustomatisCBS 110822G. Dal BelloEustroma grandiflorumGentianaceaeArgentinaGU253744GU269705GU384421GU320409
P. exilisCPC 25193; COAD 1501 (ex-epitype)M. SilvaChamaecrista orbiculataFabaceaeBrazil KT290166KT290139KT290193 KT313494
P. fijiensisCBS 120258; CIRAD 86 (ex-epitype)J. CarlierMusa sp.MusaceaeCameroonJQ324952EU514248Genome3Genome3
MUCC 792T. Kobayashi & C. NakashimaMusa cultivarMusaceaeJapanGU253776GU269748JQ324994GU320450
P. fukuokaensisCBS 132111; CPC 14689H.D. ShinStyrax japonicusStyracaceaeSouth KoreaGU253750GU269713GU384429GU320417
MUCC 887 (ex-epitype)T. KobayashiStyrax japonicusStyracaceaeJapanGU253751GU269714GU384430GU320418
P. fuligenaCBS 132017; CPC 12296Z. MershaLycopersicon sp.SolanaceaeThailandJQ324953GU269711GU384427GU320415
MUCC 533C. NakashimaLycopersicon esculentumSolanaceaeJapanGU253749GU269712GU384428GU320416
P. glaucaCBS 131884; CPC 10062H.D. ShinAlbizzia julibrissinFabaceaeSouth KoreaGU253752GU269715GU384431GU320419
P. guianensisMUCC 855C. Nakashima & T. AkashiLantana camaraVerbenaceaeJapanGU253755GU269719GU384435GU320423
MUCC 879C. NakashimaLantana camaraVerbenaceaeJapanGU253756GU269720GU384436GU320424
P. latensMUCC 763C. Nakashima & T. AkashiLespedeza wilfordiiFabaceaeJapanGU253763GU269732GU384445GU320434
P. lonicericolaMUCC 889 (ex-neotype)T. KobayashiLonicera gracilipes var. glabraCaprifoliaceaeJapanGU253766GU269736JQ324999GU320438
P. luzardiiCPC 2556A.C. AlfenasHancornia speciosaApocynaceaeBrazilGU214477AF362057/GU269738GU384450GU320440
CPC 25196; COAD 1505 (ex-epitype)M. SilvaHarcornia speciosaApocynaceaeBrazilKT290167KT290140KT290194KT313495
P. lythriCBS 132115; CPC 14588 (ex-epitype)H.D. ShinLythrum salicariaLythraceaeSouth KoreaGU253771GU269742GU384454GU320444
MUCC 865I. Araki & M. HaradaLythrum salicariaLythraceaeJapanGU253772GU269743GU384455GU320445
P. macrosporaCBS 114696; CPC 2553P.W. Crous & R.L. BenchimolBertholletia excelsaLecythidaceaeBrazilGU214478AF362055/GU269745GU384457GU320447
P. maliMUCC 886T. KobayashiMalus sieboldiiRosaceaeJapanGU253773GU269744GU384456GU320446
P. manihotiiCPC 25219; COAD 1534 (ex-type)M. SilvaManihot sp.EuphorbiaceaeBrazilKT290171KT290144KT290198KT313499
P. nephrolepidisCBS 119121R. KirschnerNephrolepis auriculataOleandraceaeTaiwanGU253779GU269751GU384462GU320453
P. nogalesiiCBS 115022C.F. HillChamaecytisus proliferusFabaceaeNew ZealandJQ324960GU269752GU384463GU320454
P. norchiensisCBS 114641C.F. HillRubus sp.RosaceaeNew ZealandGU253794GU269772GU384484GU320475
CBS 120738; CPC 13049 (ex-type)W. GamsEucalyptus sp.MyrtaceaeItalyGU253780EF394859/GU269753GU384464GU320455
P. oenotheraeCBS 131885; CPC 10290H.D. ShinOenothera odorataOnagraceaeSouth KoreaJQ324961GU269856GU384567GU320559
CBS 131920; CPC 10630H.D. ShinOenothera odorataOnagraceaeSouth KoreaGU253781GU269755GU384466GU320457
P. pallidaCBS 131889; CPC 10776H.D. ShinCampsis grandifloraBignoniaceaeSouth KoreaGU214680GU269758GU384469GU320459
P. paraguayensisCBS 111286; CPC 1459P.W. CrousEucalyptus nitensMyrtaceaeBrazilGU214479/DQ204764DQ267602DQ211680DQ147606
CBS 111317; CPC 1458P.W. CrousEucalyptus nitensMyrtaceaeBrazilGQ852634JQ324978GU384522JQ325021
P. peraeCPC 25171, COAD 1465 (ex-type)M. SilvaPera glabrataEuphorbiaceaeBrazilKT290159KT290132KT290186KT313487
P. pini-densifloraeMUCC 534Y. TokushigePinus thunbergiiPinaceaeJapanGU253785GU269760GU384471GU320461
P. piperisFBR 151R.E. HanadaPiper aduncumPiperaceaeBrazilJX875063JX875062JX896123
P. planaltinensisCPC 25189; COAD 1495 (ex-type)M. SilvaChamaecrista sp.FabaceaeBrazilKT290164KT290137KT290191KT313492
P. plumeriifoliiCPC 25191; COAD 1498 (ex-epitype)M. SilvaHimatanthus obovatusApocynaceaeBrazilKT290165 KT290138 KT290192KT313493
P. plunkettiiCPC 26081; COAD 1548R.W. BarretoMikania hirsutissimaAsteraceaeBrazilKT290178 KT290151KT290205KT313506
P. pothomorphesCPC 25166; COAD 1450 (ex-type)O.L. PereiraPothomorphe umbellataPiperaceaeBrazil KT290158KT290131KT290185KT313486
P. pouzolziaeCBS 122280R. KirschnerGonostegia hirtaUrticaceaeTaiwanGU253786GU269761GU384472GU320462
P. prunicolaCBS 132107; CPC 14511H.D. ShinPrunus x yedoensisRosaceaeSouth KoreaGU253723GU269676GU384393GU320382
P. purpureaCBS 114163; CPC 1664P.W. CrousPersea americanaLauraceaeMexicoGU253804GU269783GU384494GU320486
P. pyracanthaeMUCC 892T. Kobayashi & C. NakashimaPyracantha angustifoliaRosaceaeJapanGU253792GU269767GU384479GU320470
P. pyracanthigenaCBS 131589; CPC 10808 (ex-type)H.D. ShinPyracantha angustifoliaRosaceaeSouth KoreaGU269766GU384478GU320469
P. rhamnellaeCBS 131590; CPC 12500 (ex-type)H.D. ShinRhamnella frangulioidesRhamnaceaeSouth KoreaGU253813GU269795GU384505GU320496
P. rhapisicolaCBS 282.66K. TubakiRhapis flabellifornisArecaceaeJapanGU253793GU269770GU384482GU320473
P. richardsoniicolaCPC 25248; COAD 1568 (ex-epitype)R.W. BarretoRichardia brasiliensisRubiaceaeBrazilKT290181KT290154KT290208KT313509
P. rigidaeCPC 25175; COAD 1472 (ex-epitype)M. SilvaPalicourea rigidaRubiaceaeBrazilKT290161 KT290134KT290188KT313489
P. rubiMUCC 875T. Kobayashi & C. NakashimaRubus allegheniensisRosaceaeJapanGU253795GU269773GU384485GU320476
P. sawadaeCBS 115024C.F. HillPsidium guajavaMyrtaceaeNew ZealandJQ324967GU269775GU320478
P. sennae-multijugaeCPC 25206; COAD 1519 (ex-type)M. SilvaSenna multijugaFabaceaeBrazilKT290169KT290142KT290196KT313497
P. solani-pseudocapsicicolaCPC 25229; COAD 1974 (ex-type)M. SilvaSolanum pseudocapsicumSolanaceaeBrazilKT290175KT290148KT290202KT313503
P. sordidaMUCC 913C. Nakashima & E. ImaizumiCampsis radicansBignoniaceaeJapanGU253798GU269777GU384488GU320480
Pseudocercospora sp.CBS 110998; CPC 1054M.J. WingfieldEucalyptus grandisMyrtaceaeSouth AfricaGU253799GU269778GU384489GU320481
CBS 111373; CPC 1493M.J. WingfieldEucalyptus globulusMyrtaceaeUruguayGU253803GU269782GU384493GU320485
CBS 113387A. den BreeyenLantana camaraVerbenaceaeJamaicaGU253754GU269718GU384434GU320422
CBS 131922; CPC 10645P.W. CrousBrazilGU253700GU269651GU384369GU320359
P. stephanandraeMUCC 914 (ex-epitype)C. Nakashima & E. ImaizumiStephanandra incisaRosaceaeJapanGU253831GU269814GU384526GU320516
P. stizolobiiCPC 25217; COAD 1532M. SilvaMucuna aterrimaFabaceaeBrazilKT290170KT290143 KT290197KT313498
P. struthanthiCPC 25199; COAD 1512 (ex-epitype)M. SilvaStruthanthus flexicaulisLoranthaceaeBrazilKT290168 KT290141 KT290195KT313496
P. subsessilisCBS 136.94R.F. CastanedaCubaGU253832GU269815GU384527GU320517
P. subtorulosaCBS 117230R. KirschnerMelicope sp.RutaceaeTaiwanGU253833GU269816GU384528GU320518
P. tecomicolaCPC 25260; COAD 1585R.W. BarretoTecoma stansBignoniaceaeBrazilKT290183KT290156KT290209KT313511
P. trinidadensisCPC 26082; COAD 1756R.W. BarretoCroton urucuranaEuphorbiaceaBrazilKT290184 KT290157KT290210
P. udagawanaCBS 131931; CPC 10799H.D. ShinHovenia dulcisRhamnaceaeSouth KoreaGU269824GU384537GU320527
P. variicolorMUCC 746C. Nakashima & I. ArakiPaeonia lactiflora var.trichocarpaPaeoniaceaeJapanGU253843GU269826GU384538GU320530
P. vassobiaeCPC 25251; COAD 1572 (ex-type)R.W. BarretoVassobia brevifloraSolanaceaeBrazilKT290182 KT290155KT313510
P. viburnigenaCBS 125998; CPC 15249 (ex-epitype)M.K. CrousViburnum davidiiCaprifoliaceaeNetherlandsGU253827GU269809GU384520GU320512
P. weigelaeMUCC 899T. Kobayashi & Y. KobayashiWeigela coraeensisCaprifoliaceaeJapanGU253847GU269831GU384543GU320535
P. wulffiaeCPC 25232; COAD 1976 (ex-type)M. SilvaWulffia stenoglossaAsteraceaeBrazilKT290177KT290150KT290204 KT313505
P. xylopiaeCPC 25173; COAD 1469 (ex-type)M. SilvaXylopia aromaticaAnnonaceaeBrazil KT290160 KT290133KT290187KT313488
P. zelkovaeCBS 132118; CPC 14717H.D. ShinZelkova serrataUlmaceaeSouth KoreaGU253850GU269834GU384546JQ325028
MUCC 872T. Kobayashi & C. NakashimaZelkova serrataUlmaceaeJapanGU253851GU269835GU384547GU320537

1 CBS: CBS-KNAW Fungal Biodiversity Centre, Utrecht, The Netherlands; COAD: Coleção de Cultura Octávio Almeida Drummond, Universidade Ferderal de Viçosa, Viçosa, Brazil; CPC: Culture collection of Pedro Crous, housed at CBS; MUCC: Culture Collection, Laboratory of Plant Pathology, Mie University, Tsu, Mie Prefecture, Japan.

2 LSU: partial 28S nrRNA gene; ITS: internal transcribed spacer regions 1 & 2 including 5.8S nrRNA gene; tef1: partial translation elongation factor 1-alpha gene; actA: partial actin gene.

3 Sequence for this locus obtained from: http://genome.jgi-psf.org/Mycfi1/Mycfi1.home.html.

RESULTS

Isolates

A total of 42 specimens bearing Pseudocercospora colonies were obtained in the surveys. Twenty-seven species of Pseudocercospora were recognised as being present in these samples. Hosts belonged to the following families: Annonaceae, Apocynaceae, Asteraceae, Bignoniaceae, Bixaceae, Euphorbiaceae, Fabaceae, Icacinaceae, Loranthaceae, Lythraceae, Piperaceae, Rubiaceae, Solanaceae and Urticaceae. These hosts included weeds, agricultural species, forestry species and native plants from the Mata Alântica and the Cerrado.

Phylogeny

The LSU alignment consisted of 69 strains (including the outgroup sequence) and 713 characters were included in the analysis. The alignment had 97 unique site patterns. The LSU phylogeny (TreeBASE S17995), revealed that all strains obtained from the survey and recognised as having the morphological features of members of Pseudocercospora clustered within Pseudocercospora s.str. (data not shown, see TreeBASE). These were subsequently included in the combined actA, tef1 and ITS alignment for species level identification (Fig. 1).

For the species level analysis of the 27 Pseudocercospora isolates from Brazil, DNA sequence data from the actA, tef1 and ITS gene regions were combined for the Bayesian analyses. The concatenated alignment contained a total of 97 strains (70 strains from NCBI and 27 strains from this study) (Table 1). Passalora eucalypti (CBS 111318) served as the outgroup taxon. The final aligned sequences of the ITS (482 characters), actA (205 characters) and tef1 (373 characters) gene regions had a total length of 1 060 characters (including alignment gaps) which were included in the analyses. The gaps in the alignment were treated as fifth base for the parsimony analyses and from the analysed characters 504 were constant (ITS: 335, actA: 90, tef1: 79), 167 were variable and parsimony-uninformative (ITS: 72, actA: 23, tef1: 72) and 389 were parsimony informative (ITS: 75, actA: 92, tef1: 222). All genes were also assessed individually using Bayesian analyses (data not shown, see TreeBASE). The Bayesian analysis of the combined alignment, based on 543 unique site patterns (ITS: 141, actA: 120, tef1: 282) lasted 7 055 000 generations and the consensus trees and posterior probabilities (PP) were calculated from the 10 584 trees left after discarding 3 528 trees (the first 25 % of the generations) for burn-in (Fig. 1). A maximum of 1 000 equally most parsimonious trees (Tree Length = 2 288, CI = 0.481, RI = 0.817, RC = 0.393) were saved from the parsimony analysis (data not shown, see TreeBASE). Overall, the same terminal clades were found and the biggest differences between the parsimony tree and Bayesian tree were observed as rearrangements in the backbone of the tree, affecting the order of clades and not the species delimitation. Parsimony bootstrap support values (PBS) are plotted at the nodes, which are congruent between the parsimony bootstrap tree and the Bayesian phylogeny (Fig. 1).

The ITS region had limited resolution for differentiating species, resolving only 12 of the included 82 species, whereas the Bayesian trees based on the actA and tef1 regions resolved 41 and 38 out of 80 (for two species of each locus sequence data were missing) species respectively (data not shown, see TreeBASE). Only 11 species were supported as being distinct by all three loci in the individual Bayesian analyses, whereas 32 species were not distinct based on any of the individual loci. Details about the performance of the different loci are provided under the species notes below.

Taxonomy

Based on phylogenetic analyses, host data and morphological comparisons (Consolidated Species Concept), the Pseudocercospora isolates from Brazil could be assigned to 27 different taxa (Fig. 1), revealing a rich diversity among the Pseudocercospora spp. in this country. Among these, 12 species namely P. aeschynomenicola, P. diplusodonii, P. emmotunicola, P. manihotii, P. perae, P. planaltinensis, P. pothomorphes, P. sennae-multijugae, P. solani-pseudocapsicicola, P. vassobiae, P. wulffiae and P. xylopiae were treated as new and are described below. Epitypes were designated for a further eight species namely P. bixae, P. chamaecristae, P. exilis, P. luzardii, P. plumeriifolii, P. richardsoniicola, P. rigidae and P. struthanthi, and three species namely P. boehmeriigena, P. euphorbiacearum and P. tecomicola were found to represent new reports for Brazil, and three species represented new host associations. Additionally four isolates were shown to belong to known species. Brazilian isolates were distributed across the whole phylogeny and therefore did not cluster following a common geographic origin. The clades containing the Brazilian Pseudocercospora isolates are highlighted in the phylogenetic tree (Fig. 1). The phylogenetic relation of the various isolates is discussed in the species notes, where applicable.

Pseudocercospora aeschynomenicola Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813624; Fig. 2

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g002.jpg

Pseudocercospora aeschynomenicola (VIC 42805). a. Aeschynomene falcata with leaf spots; b. leaf spots on upper and lower leaf surface; c, d. conidiophores emerging through stomata; e. conidiogenous cells; f–i. conidia. — Scale bars: c–i = 10 μm.

Etymology. Name derived from the plant host genus Aeschynomene, from which it was collected.

Leaf spots amphigenous, irregular, scattered, grey-brown surrounded by a chlorotic halo, 1–5 mm diam. Internal mycelium, subhyaline, branched, septate, smooth, 2–2.5 μm diam. External mycelium absent. Stromata absent or small, substomatal, composed of brown textura angularis. Conidiophores hypophyllous, solitary or in small fascicles, loose, emerging through stomata, cylindrical, 12–42.5 × 3–5 μm, 0–4-septate, straight to geniculate-sinuous, unbranched, pale to medium brown, smooth. Conidiogenous cells terminal, integrated, proliferating sympodially and percurrently, subcylindrical, 8–21 × 3–5 μm, pale brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, finely guttulate, brown, smooth, subcylindrical-filiform, straight to sigmoid, 35–167 × 2–3.5 μm, apex obtuse to subacute, base obconically truncate, 2.5–3 μm wide, 4–14-septate; hila unthickened, not darkened, 1–2 μm diam.

Culture characteristics — Very slow-growing (16–18 mm diam after 20 d), convex with smooth to slightly irregularly lobate margins, aerial mycelium velvety, olivaceous grey centrally, olivaceous black periphery, iron-grey to green-black reverse, sterile.

Specimens examined. BRAZIL, Minas Gerais, Viçosa, Reserva Florestal Mata do Paraíso, on leaves of Aeschynomene falcata (Fabaceae), 22 Jan. 2014, M. Silva (holotype VIC 42805, culture ex-type COAD 1972; isotype CBS H-22164, culture ex-isotype CPC 25227).

Notes — Only one cercosporoid fungus is thus far known to occur on Aeschynomene falcata, namely Semipseudocercospora aeschynomenes from Brazil (Crous & Braun 2003). The genus Semipseudocercospora is distinguished from Pseudocercospora by having “short cylindrical pegs on which the conidia are borne, aggregated towards the tip of the conidiophores” (Yen 1983) and having ellipsoid-ovoid, short conidia with attenuated bases (Yen 1983, Crous & Braun 2003). The morphology of the fungus collected on A. falcata clearly places it in Pseudocercospora. Phylogenetically, P. aeschynomenicola clustered between Pseudocercospora sp. from an unknown host (CPC 10645) and P. eustomatis on Eustroma glandiflorum (Gentianaceae) (Fig. 1, clade 8). It is not possible to distinguish P. aeschynomenicola from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and in the tef1 phylogeny it cannot be distinguished from Pseudocercospora sp. CPC 10645, P. piperis (strain FBR 151) and P. struthanthi.

Pseudocercospora atromarginalis (G.F. Atk.) Deighton, Mycol. Pap. 140: 139. 1976

Basionym. Cercospora atromarginalis G.F. Atk., J. Elisha Mitchell Sci. Soc. 8: 59. 1892.

Descriptions & Illustrations — Deighton (1976: 139, f. 237), Hsieh & Goh (1990: 313, f. 237).

Specimen examined. BRAZIL, Minas Gerais, Viçosa, Reserva Florestal Mata do Paraíso, on leaves of Solanum americanum (Solanaceae), 23 Jan. 2014, M. Silva (CBS H-22167, VIC 42808, cultures COAD 1975, CPC 25230).

Notes — Pseudocercospora atromarginalis and P. chengtuensis, both described on Solanaceae, could not be distinguished based on the phylogenetic analysis of the combined alignment (Fig. 1, clade 9). This was also observed by Crous et al. (2013a) and Bakhshi et al. (2014). Furthermore, these species are morphologically similar (Crous et al. 2013a). To confirm whether they are synonymous or distinct species it is necessary to re-collect samples from the type localities of both species. It is not possible to distinguish P. atromarginalis from P. chengtuensis, P. fuligena or P. stizolobii based solely on ITS data, or from P. chengtuensis, P. cruenta or P. fuligena based solely on a tef1 phylogeny. In the actA phylogeny it cannot be distinguished from P. chengtuensis, and is it very closely related to P. fuligena.

Pseudocercospora bixae (Allesch. & F. Noack) Crous et al., Mycotaxon 64: 418. 1997 — Fig. 3

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g003.jpg

Pseudocercospora bixae (VIC 41563). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of leaf spot with fruiting; d, e. fasciculate conidiophores and conidiogenous cells; f–h. conidia. — Scale bars: d–h = 10 μm.

Basionym. Cercospora bixae Allesch. & F. Noack, Bol. Inst. Agron. São Paulo 85. 1898.

Leaf spots amphigenous, irregular, pale brown surrounded by an ill-defined black margin followed by a chlorotic halo, 4–12 mm diam. Internal mycelium, subhyaline, septate, branched, smooth, 3–4 μm diam. External mycelium absent. Stromata well-developed, semi-immersed, 12–32 × 22–50 μm, composed of medium brown textura angularis. Conidiophores amphigenous, in loose to dense fascicles arising from the upper cells of the stroma, subcylindrical, 12–50 × 2.5–4 μm, 0–3-septate, straight to variously curved, unbranched, medium brown, smooth. Conidiogenous cells terminal, integrated, subcylindrical, proliferating sympodially and percurrently, 5–31 × 2.5–4 μm. Conidiogenous loci inconspicuous, unthickened, not darkened, somewhat refractive. Conidia solitary, finely guttulate, pale brown, smooth, obclavate, straight to slightly curved, 34–99 × 3–4 μm, apex subobtuse, base obconically truncate, 2–3.5 μm wide, 2–7-septate; hila unthickened, not darkened, 1.5–2.5 μm diam.

Culture characteristics — Slow-growing (23–26 mm diam after 20 d); circular, raised, convex, margin smooth, irregular, aerial mycelium velvety, olivaceous grey, reverse olivaceous black, sterile.

Specimens examined. BRAZIL, São Paulo, Instituto Agronômico de Campinas, on leaves of Bixa orellana (Bixaceae), Sept. 1897, F. Noack (holotype IACM); Minas Gerais, Viçosa, Universidade Federal de Viçosa, on leaves of Bixa orellana, 21 May 2013, R.W. Barreto (epitype designated here VIC 41563, MBT202072, culture ex-epitype COAD 1563; iso-epitype CBS H-22171, culture ex-isoepitype CPC 25244).

Notes — The epitype of P. bixae, designated here, is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same country as the type. No DNA sequence data were available for P. bixae until now. Phylogenetically, P. bixae is most similar to P. sordida (Fig. 1, clade 5). Pseudocercospora sordida occurs on hosts in the Bignoniaceae, while P. bixae occurs on hosts in the Bixaceae (Crous & Braun 2003). Morphologically, the two species are quite distinct. Pseudocercospora sordida has longer and wider conidiophores (20–120 × 3.5–5 μm) and longer and wider conidia (20–200 × 3–5.5 μm) than those of P. bixae (Deighton 1976). It is not possible to distinguish P. bixae from P. sordida and P. luzardii based solely on ITS data, and it is close to, but distinct from, P. purpurea based on the tef1 phylogeny. In the actA phylogeny it is distinct from all other species.

Pseudocercospora boehmeriigena U. Braun, Trudy Bot. Inst. Komarova 20: 42. 1997 — Fig. 4

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g004.jpg

Pseudocercospora boehmeriigena (VIC 41562). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of leaf spot with fruiting; d, e. conidiophores in a loose fascicle; f. conidiogenous cells; g–j. conidia. — Scale bars: d–j = 10 μm.

Basionym. Cercospora boehmeriae Peck, Ann. Rep. N.Y. State Mus. Nat. Hist. 34: 48. 1881.

Pseudocercospora boehmeriae (Peck) Y.L. Guo & X.L. Liu, Mycosystema 2: 229. 1989. Nom. Illegit., Art. 53.1.

Leaf spots amphigenous, irregular to angular, pale brown to brown, 4–13 mm diam, vein-delimited. Internal mycelium indistinct. External mycelium absent. Stromata poorly developed, consisting of a few brown cells. Conidiophores epiphyllous, aggregated in loose fascicles, cylindrical, 13–26.5 × 2.5–3.5 μm, 0–2-septate, straight or variously curved, unbranched, pale to brown, smooth. Conidiogenous cells terminal, subcylindrical, proliferating sympodially, 6–20 × 2.5–3 μm, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, pale to pale brown, smooth, cylindrical, straight to curved, 50–102 × 3–4.5 μm, apex subobtuse or bluntly rounded, base truncate, 2–4 μm wide, 3–12-septate; hila neither thickened nor darkened, 2–3 μm diam.

Culture characteristics — Very slow-growing (12–14 mm diam after 20 d); corrugated, compressing the medium, raised, erumpent, aerial mycelium sparse, irregularly lobate margins, white and grey, reverse iron-grey, sterile.

Specimen examined. BRAZIL, Minas Gerais, Viçosa, Universidade Federal de Viçosa (Avicultura), on leaves of Boehmeria nivea (Urticaceae), 21 May 2013, R.W. Barreto (CBS H-22170, VIC 1562, cultures COAD 41562, CPC 25243).

Notes — The morphology of the Brazilian collection on Boehmeria nivea (ramie) fits well with the description of P. boehmeriigena (Braun & Mel’nik 1997). This species was previously reported from several countries, including Cambodia, China, Cuba, India and Indonesia (Crous & Braun 2003). This is the first report of P. boehmeriigena associated with leaf spots of B. nivea in Brazil. Phylogenetically, P. boehmeriigena is distinct from other species (Fig. 1, clade 3) and it has a position basal to a clade containing P. nephrolepidis and P. pouzolziae. It is not possible to distinguish P. boehmeriigena from P. nephrolepidis and P. pouzolziae based solely on ITS data. In the actA and tef1 phylogenies it is distinct from all other species.

Pseudocercospora chamaecristae U. Braun & F.O. Freire, Cryptog. Mycol. 23: 305. 2002 — Fig. 5

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g005.jpg

Pseudocercospora chamaecristae (VIC 42806). a, b. Leaf spots on upper and lower leaf surface; c. close-up of lesion with fruiting; d. synnematous conidiophores; e. conidiogenous cells; f–h. conidia. — Scale bars: d–h = 10 μm.

Leaf spots amphigenous, irregular, scattered, reddish centrally surrounded by a dark brown border, 1–3 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata immersed, substomatal, 24–46 μm diam, composed of dark brown textura angularis. Conidiophores hypophyllous, aggregated in dense synnematous conidiomata, subcylindrical, 126–278.5 × 3–4 μm, multiseptate, straight, variously curved or geniculate-sinuous, unbranched, individual conidiophores, brown to medium brown, smooth. Conidiogenous cells integrated, terminal, subcylindrical, proliferating sympodially and percurrently, 21–34 × 3–4 μm, pale brown, smooth. Conidiogenous loci inconspicuous to subinconspicuous, somewhat refractive. Conidia solitary, guttulate, pale brown, smooth, subcylindrical to ellipsoid-fusoid, obclavate, straight to curved, 30–38 × 4–6 μm, apex obtuse, base obconically truncate, 4–5 μm wide, 0–4-septate; hila unthickened, not darkened, 2–3 μm diam.

Culture characteristics — Very slow-growing (6 mm diam after 20 d), raised, stromatic, compressing and cracking the medium, iron-grey, reverse olivaceous black, sterile.

Specimens examined. BRAZIL, Ceará, Preaoca, Cascavel, on leaves of Chamaecrista setosa (Fabaceae), 9 Nov. 2000, F. Freire (holotype HAL 1718); Minas Gerais, Viçosa, Reserva Florestal Mata do Paraíso, on leaves of Chamaecrista sp. (Fabaceae), 22 Jan. 2014, M. Silva (epitype designated here VIC 42806, MBT202015, culture ex-epitype COAD 1973; isoepitype CBS H-22165, culture ex-isoepitype CPC 25228).

Notes — The epitype of P. chamaecristae designated here is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same country as the type. Phylogenetically, Pseudocercospora chamaecristae described from Chamaecrista sp. clustered in the same clade with P. exilis described from Chamaecrista orbiculata (Fig. 1, clade 1). Although both species form synnemata and occur on the same host genus, they were considered to be morphologically distinct by Hernández-Gutiérrez & Dianese (2009). Pseudocercospora exilis has percurrently proliferating conidiogenous cells, longer conidiophores (149–332 μm) and longer conidia (38–103 μm) (Hernández-Gutiérrez & Dianese 2009). Our molecular data support their view and confirm that P. chamaecristae and P. exilis are in fact distinct species. In the ITS and tef1 phylogenies P. chamaecristae is distinct from all other species, while it is distinct from but related to P. exilis in the actA phylogeny.

Pseudocercospora diplusodonii Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813581; Fig. 6

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g006.jpg

Pseudocercospora diplusodonii (VIC 42730). a. Diplusodon sp. with leaf spots on field; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. cross-section showing the internal mycelium; e. conidiophore in a small fascicle; f. conidiogenous cells; g–j. conidia. — Scale bars: d–j = 10 μm.

Etymology. Name derived from the plant host genus Diplusodon.

Leaf spots amphigenous, irregular, scattered, initially chlorotic, becoming brown with age, angular and vein-delimited, 3–8 mm diam. Internal mycelium, intra- and intercellular, 2.5–4.5 μm diam, branched, subhyaline, septate, smooth. External mycelium absent. Stromata well-developed, emerging through stomata, subglobose to irregular, brown, 17–27 × 17–39 μm, composed of dark brown textura subglobosa. Conidiophores hypophyllous, aggregated in fascicles arising from the upper cells of the stroma, subcylindrical, 12–39 × 3–5 μm, 0–4-septate, straight or geniculate, unbranched, brown, smooth. Conidiogenous cells terminal, subcylindrical, proliferating sympodially, 7.5–25 × 3.0–4.5 μm, brown, smooth to finely verruculose. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, subhyaline to pale brown, smooth, subcylindrical, straight to gently curved, 46–105 × 3–4 μm, apex obtuse, base truncate, 2.5–3 μm wide, 3–8-septate; hila unthickened, neither darkened nor refractive, 1.5–2 μm diam.

Culture characteristics — Slow-growing (18–20 mm diam after 20 d), raised, convex, corrugate, margins lobate, with aerial mycelium sparse, pale olivaceous grey, reverse iron-grey, sterile.

Specimen examined. BRAZIL, Minas Gerais, Paraopeba, Floresta Nacional de Paraopeba (FLONA), on leaves of Diplusodon sp. (Lythraceae), 31 Mar. 2013, M. Silva (holotype VIC 42730, culture ex-type COAD 1476; isotype CBS H-22151, culture ex-isotype CPC 25179).

Notes — No species of Pseudocercospora seem to have been recorded on Diplusodon (Crous & Braun 2003, Farr & Rossman 2015). Among the Pseudocercospora spp. described on plants in the Lythraceae, only P. cupheae, P. lagerstroemiae-lanceolatae and P. lythri are morphologically similar to P. diplusodonii. Pseudocercospora cupheae has shorter and narrower conidiophores (5–15 × 2–3 μm) and longer conidia (40–130 μm) than the newly described species (Braun 1999). In contrast to P. lagerstroemiae-lanceolatae, P. diplusodonii has no external mycelium with solitary conidiophores and longer and wider fasciculate conidiophores (10–100 × 3–6 μm) (Crous & Braun 2003), and is also distinguished from P. lythri by lacking external mycelium, longer conidiophores (10–90 × 2.5–5.5 μm), and wider conidia (20–110 × 3–5 μm) (Shin & Braun 2000). Pseudocercospora diplusodonii is clearly distinct from all other species of Pseudocercospora included in the phylogenetic analysis (Fig. 1, clade 16), including P. lythri (which is located between clades 2 and 3 in Fig. 1), which is also associated with a member of the Lythraceae. It is not possible to distinguish P. diplusodonii from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, but it is distinct in the tef1 phylogeny.

Pseudocercospora emmotunicola Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813583; Fig. 7

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g007.jpg

Pseudocercospora emmotunicola (VIC 42744). a. Emmotum nitens with leaf spots; b. leaf spots on upper and lower leaf surface; c. cross-section showing the internal mycelium; d. sporodochial conidiophores; e. conidiogenous cells; f. conidia. — Scale bars: c–f = 10 μm.

Etymology. Name derived from the host genus Emmotum.

Leaf spots amphigenous, scattered, chlorotic becoming ochraceous-yellow, poorly delimited, diffuse, 5–15 mm diam. Internal mycelium, subhyaline, septate, smooth, 2–2.5 μm diam. External mycelium absent. Stromata well-developed, 12–22 × 20–38 μm, erumpent, angular, composed of dark brown textura angularis. Conidiophores hypophyllous, sporodochial arising from the stroma, subcylindrical, 8–29 × 2–3 μm, 0–1-septate, straight or geniculate, pale brown, unbranched, becoming subhyaline towards the apex, smooth. Conidiogenous cells terminal, integrated, proliferating sympodially, 9–16 × 2–3.5 μm, subhyaline to pale brown, subcylindrical, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, pale brown, smooth, subcylindrical, straight to curved, 24–99 × 2–3.5 μm, apex obtuse, base truncate, 1.5–2.5 μm wide, 1–12-septate; hila unthickened, not darkened, 1.5–2 μm diam.

Culture characteristics — Slow-growing (21–24 mm diam after 20 d), raised with smooth to slightly irregular margins, aerial mycelium velvety, olivaceous grey, grey-sepia centrally, olivaceous black periphery, reverse iron-grey to greenish black, sterile.

Specimen examined. BRAZIL, Distrito Federal, Brasília, Estação Ecológica de Águas Emendadas, on leaves of Emmotum nitens (Icacinaceae), 16 Apr. 2013, M. Silva (holotype VIC 42744, culture ex-type COAD 1491; isotype CBS H-22152, culture ex-isotype CPC 25187).

Notes — No species of Pseudocercospora are known to occur on Emmotum (Icacinaceae) (Farr & Rossman 2015). In the multigene phylogenetic analysis, P. emmotunicola is basal in a clade containing P. perae and P. guianensis (Fig. 1, clade 15). It is not possible to distinguish P. emmotunicola from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and it cannot be distinguished from P. perae in the tef1 phylogeny.

Pseudocercospora euphorbiacearum U. Braun, Biblioth. Lichenol. 86: 89. 2003 — Fig. 8

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g008.jpg

Pseudocercospora euphorbiacearum (VIC 42797). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion with fruiting; c. fasciculate conidiophores; d. conidiogenous cells; e–h. conidia. — Scale bars: c–h = 10 μm.

Leaf spots amphigenous, circular to irregular, chlorotic with a white centre, 4–12 mm diam. Internal mycelium intercellular, 2–3.5 μm, branched, subhyaline, septate, smooth. External mycelium absent. Stromata hypophyllous, erumpent, well-developed, erumpent, 17–31.5 × 17–47 μm, composed of brown textura angularis. Conidiophores aggregated in dense fascicles arising from the upper cells of the stromata, subcylindrical, 17–42 × 2.5–4 μm, 0–4-septate, straight to geniculate-sinuous, unbranched, pale olivaceous to olivaceous brown, smooth. Conidiogenous cells terminal, integrated, subcylindrical, proliferating sympodially, 10–27 × 2.5–4 μm, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, finely guttulate, subhyaline to pale olivaceous, smooth, subcylindrical, straight to curved, 49–94 × 3–4 μm, apex obtuse, base obconically to truncate, 2.5–3.5 μm wide, 3–14-septate; hila unthickened, not darkened, 1–2 μm diam.

Culture characteristics — Slow-growing (25–28 mm diam after 20 d), convex, circular with smooth to slightly irregularly lobate margins, aerial mycelium velvety, pale olivaceous grey, reverse olivaceous black, sterile.

Specimen examined. BRAZIL, Minas Gerais, Viçosa, Reserva Florestal Mata do Paraíso, on leaves of Dalechampia sp. (Euphorbiacea), 5 Aug. 2013, M. Silva (CBS H-22163, VIC 42797, cultures COAD 1537, CPC 25222).

Notes — The morphology of the Brazilian specimen fits well within the original description of P. euphorbiacearum described on Dalechampia scandens from the Dominican Republic (Braun 2003). This is the first report of P. euphorbiacearum in Brazil, and the first time molecular data is generated for this species. Phylogenetically, P. euphorbiacearum (on Euphorbiaceae) is closely related to P. pini-densiflorae (on Pinaceae) based on the multigene alignment (Fig. 1, clade 12). Pseudocercospora pini-densiflorae is a pathogen of a distantly related host family (Pinaceae) and is morphologically distinct from P. euphorbiacearum (Chupp 1954, Crous & Braun 2003). It is not possible to distinguish P. euphorbiacearum from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and it can barely be distinguished from P. pini-densiflorae and P. trinidadensis in the tef1 phylogeny.

Pseudocercospora exilis A. Hern.-Gut. & Dianese, Mycotaxon 108: 17. 2009 — Fig. 9

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g009.jpg

Pseudocercospora exilis (VIC 42754). a. Chamaecrista orbiculata with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of circular lesion; d. close-up of lesion with fruiting; e, f. synnematous conidiophores; g–k. conidia. — Scale bars: e–k = 10 μm.

Leaf spots amphigenous, circular or irregular, scattered, grey-brown centrally with a dark brown to black margin, 1–6 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata small to well-developed, substomatal, 15–42 μm diam, composed of brown textura globosa. Conidiophores amphigenous, aggregated in synnemata, subcylindrical, 115–306 × 5–6.5 μm, 4–15-septate, straight, curved or geniculate-sinuous at the upper portion, unbranched, brown, smooth. Conidiogenous cells integrated, terminal, proliferating percurrently, 18–32 × 5–6.5 μm, pale brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened, somewhat refractive. Conidia solitary, finely guttulate, pale brown, smooth, obclavate or fusoid, straight to slightly curved, 42–78.5 × 5–6.5 μm, apex rounded, base obconically truncate, 4.5–6 μm wide, 1–7-septate; hila unthickened, not darkened, 2.5–4 μm diam.

Culture characteristics — Very slow-growing (12–15 mm diam after 20 d), raised, corrugated, with smooth, irregular margins, green-black centrally with shiny black margins, reverse olivaceous black, sterile.

Specimens examined. BRAZIL, Distrito Federal, Brasília, on leaves of Chamaecrista orbiculata (Fabaceae), 9 Aug. 1992, J.C. Dianese (holotype UB Mycol. Col. 1477); Estação Ecológica de Águas Emendadas, on leaves of Chamaecrista orbiculata, 21 Apr. 2013, M. Silva (epitype designated here VIC 42754, MBT202016, culture ex-epitype COAD 1501; isoepitype CBS H-22155, culture ex-isoepitype CPC 25193).

Notes — The epitype of P. exilis, designated here, is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same biome and country as the holotype. Also see the notes under P. chamaecristae. In the multigene phylogenetic analysis, P. exilis groups with P. chamaecristae (Fig. 1, clade 1). In the ITS and tef1 phylogenies P. exilis is distinct from all other species, while it is distinct from but related to P. chamaecristae in the actA phylogeny.

Pseudocercospora luzardii Furlan. & Dianese, Mycol. Res. 103: 1207. 1999 — Fig. 10

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g010.jpg

Pseudocercospora luzardii (VIC 42758). a. Harconia speciosa with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. close-up of lesion with fruiting; e. fasciculate conidiophores; f–h. conidia. — Scale bars: e–h = 10 μm.

Leaf spots amphigenous, distinct, oval to irregular, pale grey in the centre surrounded by a purple brown to dark brown margin, 2–7 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata epiphyllous, well-developed, subimmersed, 34–53.5 × 43–82 μm, compose of dark brown textura angularis. Conidiophores aggregated in dense fascicles, cylindrical, 19–84 × 3–6 μm, 1–6-septate, straight or sinuous, unbranched, brown, smooth. Conidiogenous cells integrated, terminal, polyblastic, proliferating percurrently, 6–25 × 3–6 μm, pale brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, finely guttulate, pale brown to brown, smooth, cylindrical, straight to variously curved, 19–84 × 3–5 μm, apex subobtuse, base obconic to subtruncate, 3–4.5 μm wide, 1–8-septate; hila neither thickened nor darkened, 1.5–2 μm diam.

Culture characteristics — Very slow-growing (18 mm diam after 20 d), raised, corrugated, with smooth, lobate margins, aerial mycelium sparse, velvety, grey with patches of olivaceous grey, reverse iron-grey, sterile.

Specimens examined. BRAZIL, Goiás, Cristalina, Fazenda Nova Índia, on leaves of Harcornia speciosa (Apocynaceae), 10 Apr. 1993, J.C. Dianese (holotype, UB Mycol. Col. 4149); Distrito Federal, Brasília, Estação Ecológica de Águas Emendadas, on leaves of Harcornia speciosa, 19 Apr. 2013, M. Silva (epitype designated here VIC 42758, MBT202017, culture ex-epitype COAD 1505; isoepitype CBS H-22156, culture ex-isoepitype CPC 25196).

Notes — The epitype of P. luzardii, designated here, is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same locality as the holotype. The DNA sequence data place the culture from this study together with strain CPC 2556, identified by Crous et al. (2013a) as P. luzardii (Fig. 1, clade 4). The phylogenetic placement is in agreement with the morphological data, confirming this species as P. luzardii. It is not possible to distinguish P. luzardii from P. bixae and P. sordida based solely on an ITS phylogeny, but it can be distinguished from all other Pseudocercospora spp. based on the individual tef1 and actA phylogenies.

Pseudocercospora manihotii Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813584; Fig. 11

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g011.jpg

Pseudocercospora manihotii (VIC 42793). a. Manihot sp. with leaf spots; b. leaf spots on upper and lower leaf surface; c, d. fasciculate conidiophores; e–j. conidia. — Scale bars: c–e = 10 μm.

Etymology. Name derived from the plant host genus Manihot.

Leaf spots amphigenous, irregular, scattered, reddish brown surrounded by a dark brown to black margin, 10–35 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata well-developed, subimmersed or erumpent, 23–46 × 38–64 μm, composed of brown textura angularis. Conidiophores epiphyllous, aggregated in dense fascicles arising from the upper cells of the stroma, cylindrical, 15–56 × 3–6 μm, 0–3-septate, straight to slightly geniculate-sinuous, unbranched, pale brown, smooth. Conidiogenous cells terminal, sometimes intercalary, cylindrical, proliferating sympodially, 12.5–29 × 3–5.5 μm, pale brown, smooth. Conidiogenous loci slightly conspicuous, slightly thickened, not darkened. Conidia solitary, finely guttulate, pale brown, smooth, cylindrical to narrowly obclavate, straight to curved, 19–97 × 2–4 μm, apex rounded to subacute, base obconically truncate, 2–3 μm wide, 0–10-septate; hila unthickened, not darkened, 1.5–2.5 μm diam.

Culture characteristics — Very slow-growing (15–18 mm diam after 20 d); convex, with smooth, lobate margins, and sparse aerial mycelium, olivaceous grey, reverse iron-grey, sterile.

Specimen examined. BRAZIL, Minas Gerais, Paraopeba, Floresta Nacional de Paraopeba (FLONA), on leaves of Manihot sp. (Euphorbiaceae), 29 Apr. 2013, M. Silva (holotype VIC 42793, culture ex-type COAD 1534; isotype CBS H-22161, culture ex-isotype CPC 25219).

Notes — No Pseudocercospora spp. are known to be associated with the genus Manihot. Several species of Pseudocercospora are known to occur on Euphorbiaceae, but all are dissimilar to the fungus collected on Manihot (Crous & Braun 2003, Farr & Rossman 2015). Pseudocercospora hurae is the species having the most similar morphology to that of P. manihotii among those described on members of the Euphorbiaceae (Deighton 1976). It also has well-developed stromata with conidiophores forming dense fascicles, but differs from the newly proposed species in having smaller and narrower conidiophores (5–40 × 3–4.5 μm) (Deighton 1976). Pseudocercospora manihotii clusters together with P. wulffiae in the phylogeny derived from the combined alignment (Fig. 1, clade 6). The DNA sequences generated here (ITS, actA and tef1) did not allow for a clear distinction between P. manihotii and P. wulffiae (Fig. 1, clade 6). However, P. wulffiae is a pathogen of plants belonging to a different host family (Asteraceae), and it has a clearly distinct mor-phology (shorter and narrower conidiophores (14–21 × 2–3 μm) and shorter conidia (37.5–87 μm) indicating that these are distinct taxa for which additional gene regions will be required to resolve the species boundaries. It is not possible to distinguish P. manihotii from several other Pseudocercospora spp. based solely on an ITS phylogeny, and it cannot be distinguished from P. wulffiae in the tef1 phylogeny. In the actA phylogeny it is more distinct from closely related species.

Pseudocercospora perae Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813589; Fig. 12

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g012.jpg

Pseudocercospora perae (VIC 42721). a. Pera glabrata with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. close-up of lesion with fruiting; e. cross-section showing the internal mycelium; f, g. conidiophores in sporodochial; h. conidiogenous cells; i–k. conidia. — Scale bars: e–k = 10 μm.

Etymology. Name derived from the plant host genus Pera.

Leaf spots amphigenous, circular to irregular, pale brown to brown, on upper surface white centrally, 3–6 mm diam, surrounded by a black margin. Internal mycelium, subhyaline, septate, branched, smooth, 3.5–4 μm diam. External mycelium absent. Stromata well-developed, 14–35 × 23–42 μm, subimmersed or erumpent, brown, composed of dark brown textura angularis. Conidiophores hypophyllous, aggregated in loose to dense fascicles, arising from the upper cells of the stroma, cylindrical, 9–68.5 × 3–4 μm, 0–3-septate, straight or geniculate, unbranched, brown, smooth. Conidiogenous cells terminal, integrated, subcylindrical, proliferating percurrently, 7–17 × 3–3.5 μm, brown, smooth to finely verruculose. Conidiogenous loci inconspicuous, slightly thickened, not darkened. Conidia solitary, finely guttulate, subhyaline to pale brown, smooth, subcylindrical, straight to curved at the apex, 27–102 × 3–5 μm, apex obtuse, base truncate, 2.5–3.5 μm wide, 5–6-septate; hila unthickened, neither darkened nor refractive, 1.5–2 μm diam.

Culture characteristics — Slow-growing (25–28 mm diam after 20 d), raised, circular with smooth to slightly irregular margins, aerial mycelium velvety, pale olivaceous grey with olivaceous black periphery, reverse greenish black, sterile.

Specimen examined. BRAZIL, Minas Gerais, Paraopeba, Floresta Nacional (FLONA), on leaves of Pera glabrata (Euphorbiaceae), 3 Jan. 2013, M. Silva (holotype VIC 42721, culture ex-type COAD 1465; isotype CBS H-22148, culture ex-isotype CPC 25171).

Notes — No species of Pseudocercospora or other cercosporoid fungi and mycosphaerella-like sexual morphs are presently known to occur on species of Pera, but numerous Pseudocercospora spp. have been described from hosts in the Euphorbiaceae (Farr & Rossman 2015). Among these P. crotoniphila is morphologically similar but distinguishable from P. perae by having shorter and wider conidiophores (20–40 × 4–5 μm) and shorter conidia (20–90 μm) (Crous et al. 1999). Another species similar to P. perae is P. hieronymae that differs by having narrower conidia (2.5–4 μm) (Chupp 1954, Crous & Braun 2003), while P. hurae has shorter conidiophores (5–40 × 3–4.5 μm) and narrower conidia (2–4.5 μm) (Chupp 1954). In the multigene phylogenetic analysis, P. perae is in a clade containing P. emmotunicola and P. guianensis (Fig. 1, clade 15). It is not possible to distinguish P. perae from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and it cannot be distinguished from P. emmotunicola in the tef1 phylogeny.

Pseudocercospora planaltinensis Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813591; Fig. 13

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g013.jpg

Pseudocercospora planaltinensis (VIC 42748). a. Chamaecrista sp. with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. close-up of lesion with fruiting; e. cross-section showing the sporodochial conidioma; f. conidiogenous cells; g–k. conidia. — Scale bars: e–k = 10 μm.

Etymology. Name derived from Planaltina, the Brazilian municipality where the fungus was first found.

Leaf spots amphigenous, brown, surrounded by a dark brown to black defined margin, irregular, 2–11 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata well-developed, immersed, 128–147.5 μm diam, composed of brown textura porrecta. Conidiophores amphigenous, mostly epiphyllous, sporodochial, arising from the stromata, cylindrical, 11–68 × 3–5.5 μm, 0–3-septate, straight, unbranched, brown, smooth. Conidiogenous cells terminal, cylindrical, proliferating percurrently, 5–31 × 3–5 μm, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, pale brown, smooth, cylindrical to obclavate, straight to curved, 49–129 × 3–5 μm, apex obtuse or acute, base obconically truncate, 2.5–4.5 μm wide, 1–8-septate; hila not thickened, not darkened, 1.5–2.5 μm diam.

Culture characteristics — Very slow-growing (16–18 mm diam after 20 d), raised, margins lobate, aerial mycelium velvety, pale olivaceous grey, reverse iron-grey, sterile.

Specimen examined. BRAZIL, Distrito Federal, Brasília, Estação Ecológica de Águas Emendadas, on leaves of Chamaecrista sp. (Fabaceae), 17 Apr. 2013, M. Silva (holotype VIC 42748, culture ex-type COAD 1495; isotype CBS H-22153, culture ex-isotype CPC 25189).

Notes — There are five Pseudocercospora spp. known to occur on the host genus Chamaecrista, namely P. chamaecristae, P. chamaecristigena, P. exilis, P. luzianiensis and P. nigricans (Farr & Rossman 2015). Pseudocercospora chamaecristae, P. chamaecristigena, P. exilis and P. luzianiensis are easily separated on morphological basis from P. planaltinensis by having different conidial shapes and wider conidia with longer synnematous conidiophores (Braun & Freire 2002, Hernández-Gutiérrez & Dianese 2009). Pseudocercospora nigricans has conidia similar to those of P. planaltinensis. However, conidia of P. nigricans are smaller (18–80 × 3–5 μm), its conidiophores are not arranged in sporodochia and the stromata are either absent or reduced to a few cells (Chupp 1954, Brown & Morgan-Jones 1977). Genetically, P. planaltinensis is very distinct from all other species of Pseudocercospora included in the phylogenetic analysis (Fig. 1, clade 13), and is somewhat related to P. subsessilis, a species known to cause leaf spots on Azadirachta indica, Melia azadirachta and Swietenia macrophylla (Meliaceae) (Braun & Castañeda-Ruiz 1991, Braun & Freire 2006, Farr & Rossman 2015). Morphologically, P. subsessilis differs from P. planaltinensis by having smaller and narrower conidia (25–80 × 2–4 μm) (Chupp 1954). The species is distinct from all other included Pseudocercospora spp. based on individual gene trees of all three loci, ITS, actA and tef1.

Pseudocercospora plumeriifolii (Bat. & Peres) U. Braun et al., Cryptog. Mycol. 20: 102. 1999 — Fig. 14

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g014.jpg

Pseudocercospora plumeriifolii (VIC 42751). a. Himatanthus obovatus with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. sporodochial conidioma; e. conidiogenous cells; f–i. conidia. — Scale bars: d–i = 10 μm.

Basionym. Cercospora plumeriifolii Bat. & Peres, Pub. Inst. Micol. Recife 262: 23. 1960.

Leaf spots amphigenous, scattered, irregular, greyish, delimited by a dark brown to black margin, 4–12 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata amphigenous, well-developed, 55–92 × 99–121 μm, immersed to partly erumpent, angular to globose, composed of dark brown textura angularis. Conidiophores sporodochial, arising from a stroma, cylindrical, 13–45 × 2.5–4 μm, 0–4-septate, straight to geniculate-sinuous, unbranched, brown, smooth. Conidiogenous cells terminal, proliferating sympodially, 7–19 × 3–4 μm, subcylindrical to sinuous, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, olivaceous to olivaceous brown, smooth, obclavate, straight to curved, 25–110 × 3–5 μm, apex obtuse, base obconically truncate, 2.5–4.5 μm wide, 2–9-septate; hila unthickened, not darkened, 1.5–2.5 μm diam.

Culture characteristics — Very slow-growing (20 mm diam after 20 d), raised with smooth margins, aerial mycelium velvety, centre olivaceous grey, olivaceous black periphery, reverse green-black, sterile.

Specimens examined. BRAZIL, Minas Gerais, Paraopeba, Horto Florestal, on leaves of Himatanthus obovatus (Apocynaceae), 1960, Batista (holotype, IMUR 19074); Distrito Federal, Brasília, Estação Ecológica de Águas Emendadas, on leaves of Himatanthus obovatus, 19 Apr. 2013, M. Silva (epitype designated here VIC 42751, MBT202067, culture ex-epitype COAD 1498; isoepitype CBS H-22154, culture ex-isoepitype CPC 25191).

Notes — The epitype of P. plumeriifolii, designated here, is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same biome and country as the type. No DNA sequence data were available for P. plumeriifolii until now. Phylogenetically, P. plumeriifolii clusters in a clade with P. catalpigena, P. pallida, P. rhapisicola and P. rigidae (Fig. 1, clade 17). Pseudocercospora catalpigena differs from P. plumeriifolii by having shorter and wider conidiophores (5–35 × 3–6 μm) (Braun et al. 2003), while P. rigidae has longer and wider conidiophores (21–85 × 3–5 μm). Pseudocercospora pallida and P. rhapisicola are morphologically similar, but they are described from hosts in different families, Bignoniaceae and Arecaceae, respectively (Goh & Hsieh 1989, Shin & Braun 2000). It is not possible to distinguish P. plumeriifolii from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and it can barely be distinguished from P. catalpigena, P. pallida and P. rhapisicola in the tef1 phylogeny.

Pseudocercospora plunkettii (Chupp) R.F. Castañeda & U. Braun, Cryptog. Bot. 2: 295. 1991 — Fig. 15

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g015.jpg

Pseudocercospora plunkettii (VIC 42644). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. sporodochial conidiophores; d. close-up of conidiophores and conidiogenous cells; e–i. conidia. — Scale bars: c, e–i = 10 μm, d = 20 μm.

Basionym. Cercospora plunkettii Chupp, A monograph of the fungus Cercospora: 154. 1954.

Leaf spots amphigenous, irregular, grey-brown surrounded by a black border, 3–12 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata amphigenous, well-developed, 32–39 × 48–53 μm, angular to irregular, composed of dark brown textura angularis. Conidiophores aggregated in dense fascicles, emerging through stromata, 20–85 × 3.5–5 μm, 3–8-septate, straight to strongly geniculate-sinuous, unbranched, pale brown, smooth. Conidiogenous cells terminal, 6–31 × 3.5–5 μm, pale brown, proliferating sympodially, rarely percurrently, smooth. Conidia solitary, guttulate, pale brown, smooth, subcylindrical to obclavate, straight to curved, 49–81 × 3–5 μm, apex obtuse to subacute, base obconically truncate, 3–5 μm, 6–10-septate; hila unthickened, not darkened, 2.5–5 μm diam.

Culture characteristics — Slow-growing (23 mm diam after 20 d), raised with smooth to slightly irregular margins, aerial mycelium velvety, olivaceous grey, reverse olivaceous black, sterile.

Specimen examined. BRAZIL, Rio de Janeiro, Nova Friburgo, Fazenda Barreto II, on leaves of Mikania sp. (Asteraceae), 10 Feb. 2013, R.W. Barreto (CBS H-22169, VIC 42644, COAD 1548, CPC 26081).

Notes — Pseudocercospora plunkettii was previously recorded on Mikania cordifolia in Cuba and Mexico (Chupp 1954, Braun & Castañeda-Ruiz 1991) and on Mikania micrantha in Venezuela and Brazil (Barreto & Evans 1995, Crous & Braun 2003). Our fungus compared well with the description of P. plunkettii, and the present study represents the first sequence data for this species. The species clusters with P. basitruncata and P. richardsoniicola (Fig. 1, clade 2). Pseudocercospora basitruncata is morphologically distinct from P. plunkettii by having shorter conidiophores (12–60 μm) and longer conidia (25–90 μm), while P. richardsoniicola has longer conidiophores and conidia (90–192 μm, 36–97 μm, respectively) (Crous 1998, Crous & Câmara 1998). Pseudocercospora plunkettii is distinct from other species in the ITS phylogeny, and closely related to P. basitruncata and P. richardsoniicola in the tef1 and actA phylogenies.

Pseudocercospora pothomorphes Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB814904; Fig. 16

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g016.jpg

Pseudocercospora pothomorphes (VIC 42705). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of leaf spot with fruiting; d, e. fasciculate conidiophores; f–i. conidia. — Scale bars: d–i = 10 μm.

Etymology. Name derived from the plant host genus Pothomorphe.

Leaf spots amphigenous, irregular or angular, scattered, brown, vein-delimited, 1–8.5 mm diam. Internal mycelium subhyaline, septate, branched, smooth, 2.5–4 μm diam. External mycelium absent. Stromata lacking or reduced to only a few cells. Conidiophores hypophyllous, aggregated in small to moderately large fascicles, loose, arising from stromata, emerging through stomata, cylindrical, 15–90 × 3.5–6 μm, 0–5-septate, straight or sinuous, rarely branched, brown, becoming paler towards the apex, smooth. Conidiogenous cells terminal, pale brown, subcylindrical, smooth, proliferating sympodially and percurrently, 7–19 × 3–5.5 μm, apical loci indistinct, unthickened and not darkened. Conidia solitary, guttulate, subhyaline to pale brown, smooth, subcylindrical to narrowly obclavate, straight to curved, 26–68.5 × 3.5–5 μm, apex rounded to subacute, base truncate, 2.5–4 μm wide, 1–7-septate; hila neither thickened nor darkened, 2–2.5 μm diam.

Culture characteristics — Slow-growing (19–22 mm diam after 20 d), convex, somewhat folded, with smooth to slightly irregular margins, aerial mycelium velvety, olivaceous grey, green-black reverse, sterile.

Specimen examined. BRAZIL, Minas Gerais, Viçosa, Reserva Florestal Mata do Paraíso, on leaves of Pothomorphe umbellata (Piperaceae), 15 Nov. 2012, O.L. Pereira (holotype VIC 42705, culture ex-type COAD 1450; isotype CBS H-22147, culture ex-isotype CPC 25166).

Notes — One species of Pseudocercospora is known on Pothomorphe, namely Pseudocercospora piperis reported on Pothomorphe peltata in Panama and on Po. umbellata in Brazil (Crous & Braun 2003, Farr & Rossman 2015). Morphologically, P. piperis differ from P. pothomorphii by having conidiophores that are branched and shorter (20–80 μm), as well as longer conidia (25–130 μm) (Deighton 1976). Rocha et al. (2013) deposited sequences in GenBank for P. piperis on Piper aduncum (tef1: JX896123; ITS: JX875062) that differ from the sequences generated for P. pothomorphes on Pothomorphe umbellata collected during this study (Table 1). Based on DNA sequence data, these species possess only 87 % similarity in the partial gene region of tef1; unfortunately no actA sequences of strain FBR1 are available for comparison. In the molecular phylogeny derived from the multigene alignment, the two isolates cluster in two different clades (Fig. 1, clade 8 for strain FBR 151 and clade 11 for P. pothomorphes). It is not possible to distinguish strains FBR 151 and COAD 1450 from numerous other Pseudocercospora spp. based solely on an ITS phylogeny. In the tef1 phylogeny, P. pothomorphes cannot be distinguished from Pseudocercospora sp. CBS 110998 and P. cordiana, whereas strain FBR 151 cannot be distinguished from Pseudocercospora sp. CPC 10645, P. aeschynomenicola and P. struthanthi. In the actA phylogeny, P. pothomorphes is close to but distinct from Pseudocercospora sp. CPC 10645.

Pseudocercospora richardsoniicola Crous & M.P.S. Câmara, Mycotaxon 68: 307. 1998 — Fig. 17

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g017.jpg

Pseudocercospora richardsoniicola (VIC 42661). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of leaf spot with fruiting; d. fasciculate conidiophores; e–i. conidia. — Scale bars: d–i = 10 μm.

Basionym. Cercospora richardsoniae Henn., Hedwigia 41: 117. 1902 (non C. richardsoniae Ellis & Everh.).

Leaf spots amphigenous, irregular to circular, scattered, pale brown, surrounded by a dark brown border, 4–14 mm diam. Internal and external mycelium pale brown, 3–4 μm diam. Stromata amphigenous, well-developed, 45–61 × 54–70 μm subimmersed, angular, composed of brown textura angularis. Conidiophores arising from stromata aggregated in dense fascicles, cylindrical, 90–192 × 3–5 μm, 4–15-septate, straight to slightly curved, unbranched, medium brown, becoming paler toward the apex, smooth. Conidiogenous cells terminal, proliferating sympodially, 9–71 × 2.5–5 μm, pale brown, cylindrical, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, cylindrical to obclavate, straight to slightly curved, 36–97 × 3–5 μm, apex rounded to obtuse, base obconically truncate, 3–8-septate, guttulate, pale brown, smooth, 2.5–5 μm wide; hila neither thickened nor darkened, 1.5–2.5 μm diam.

Culture characteristics — Very slow-growing (12–14 mm diam after 20 d), raised with smooth, lobate margins, aerial mycelium sparse, white and greyish, reverse black, sterile.

Specimens examined. BRAZIL, São Paulo, Botanic Garden, on leaves of Richardsonia sp. (Rubiaceae), 4 Feb. 1901, A. Puttemans (holotype BPI 440387); Rio de Janeiro, Nova Friburgo, Mury, on leaves of Richardia brasiliensis, 9 June 2013, R.W. Barreto (epitype designated here VIC 42661, MBT202068, culture ex-epitype COAD 1568; isoepitype CBS H-22172, culture ex-isoepitype CPC 25248).

Notes — The epitype of P. richardsoniicola, designated here, is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same country as the type. Pseudocercospora richardsoniicola is phylogenetically closely related to P. basitruncata, and sister to P. plunkettii (Fig. 1, clade 2). Pseudocercospora basitruncata occurs on a distantly related host (Eucalyptus sp.) belonging to a different host family (Myrtaceae) and has a clearly distinct morphology – shorter conidiophores (12–60 μm) and narrower conidia (2.5–3.5 μm) (Crous 1998). For P. plunkettii see notes above. Pseudocercospora richardsoniicola is distinct from other species in the ITS phylogeny, and closely related to P. plunkettii and P. richardsoniicola in the tef1 and actA phylogenies.

Pseudocercospora rigidae Meir. Silva & O.L. Pereira, Mycotaxon 102: 261. 2007 — Fig. 18

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g018.jpg

Pseudocercospora rigidae (VIC 42726). a. Palicourea rigida with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. close-up of lesion with fruiting; e, f. fasciculate conidiophores; g–j. conidia. — Scale bars: e, f = 10 μm.

Leaf spots amphigenous, irregular or vein delimited, pale brown, surrounded by a dark brown to black border, confluent, covering large areas of the leaf surface, 2–15.5 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata well-developed, subepidermal, erumpent, dark brown, 16–27 × 19–53 μm, composed of brown textura globosa. Conidiophores amphigenous, fasciculate, arising from the subepidermal stromata, 21–85 × 3–5 μm, 3–9-septate, straight to geniculate-sinuous, rarely branched below, dark brown, smooth. Conidiogenous cells terminal or lateral, proliferation percurrently and sometimes sympodially, 12–23 × 3–4 μm, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, pale brown to brown, smooth, guttulate, obclavate-cylindrical, straight to slightly curved, 25–99 × 3–5 μm, apex obtuse to subacute, 2–2.5 μm wide, 0–7-septate; hila slightly thickened, slightly darkened not refractive, 1.5–2 μm diam.

Culture characteristics — Slow-growing (19–22 mm diam after 20 d), raised, corrugated with smooth, lobate margins, aerial mycelium velvety, olivaceous grey, reverse olivaceous black, sterile.

Specimens examined. BRAZIL, Minas Gerais, Carrancas, on leaves of Palicourea rigida (Rubiaceae), Mar. 2007, O.L. Pereira (holotype VIC 30472); Paraopeba, Floresta Nacional de Paraopeba (FLONA), on leaves of Palicourea rigida, 30 Mar. 2013, M. Silva (epitype designated here VIC 42726, MBT202069, culture ex-epitype COAD 1472; isoepitype CBS H-22150, culture ex-isoepitype CPC 25175).

Notes — The epitype of P. rigidae, designated here, is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same locality as the type. This study represents the first phylogenetic data available for this species, showing that it is basal to a clade containing P. catalpigena, P. pallida, P. plumeriifolii and P. rhapisicola (see morphological differences of these species in the above notes under P. plumeriifolii) (Fig. 1, clade 17). It is not possible to distinguish P. rigidae from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and it is closely related to P. zelkovae in the tef1 phylogeny.

Pseudocercospora sennae-multijugae Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB814905; Fig. 19

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g019.jpg

Pseudocercospora sennae-multijugae (VIC 42775). a. Senna multijuga with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion with fruiting; d. cross-section showing the internal mycelium; e. fasciculate conidiophores; f. conidiogenous cells; g–j. conidia. — Scale bars: d–j = 10 μm.

Etymology. Name derived from the plant host Senna multijuga.

Leaf spots amphigenous, grey-brown in the centre, surrounded by a dark brown to black margin, mostly in the border of leaves, irregular, 2–18 mm diam. Mycelium internal, subhyaline, consisting of septate, smooth hyphae, 2.5–3 μm diam wide. External mycelium subhyaline, consisting of septate, smooth hyphae, 2.5–4 μm diam. Stromata well-developed, substomatal, 25–67 μm diam, brown, composed of brown textura angularis. Conidiophores hypophyllous, sporodochial, arising from stroma, emerging through stomata, 8–14 × 2–4.5 μm, 0–2-septate, straight to sinuous, unbranched, medium brown to brown, smooth. Conidiogenous cells terminal, or conidiophores reduced to conidiogenous cells, 8–11 μm long, medium brown, subcylindrical, smooth, proliferating sympodially. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, olivaceous brown, finely guttulate, smooth, cylindrical to narrowly obclavate, straight to curved, 11–81 × 3–4 μm, apex obtuse, base obconically truncate, 2.5–4 μm wide, 2–7-septate; hila neither thickened nor darkened, 2–2.5 μm diam.

Culture characteristics — Slow-growing (18–20 mm diam after 20 d), raised, corrugated with irregular margins, aerial mycelium sparse, olivaceous grey, reverse green-black, sterile.

Specimen examined. BRAZIL, Goiás, Alto Paraíso de Goiás, Parque Nacional da Chapada dos Veadeiros, on leaves of Senna multijuga (Fabaceae), 23 Apr. 2013, M. Silva (holotype VIC 42775; culture ex-type COAD 1519, isotype CBS H-22158, culture ex-isotype CPC 25206).

Notes — Nine species of Pseudocercospora have previously been recorded on members of Senna, namely P. angustata, P. cassiae-alatae, P. cassiae-fistulae, P. cassiae-occidentalis, P. cassiae-siameae, P. nigricans, P. simulate, P. singaporensis and P. taichugensis (Farr & Rossman 2015). Two Pseudocercospora species known on Senna have a similar morphology to P. sennae-multijugae, namely P. nigricans, which occurs on different hosts on Fabaceae, and P. taichungensis reported on Senna atomataria and Cassia fistula (Farr & Rossman 2015). Pseudocercospora nigricans differs from P. sennae-multijugae by having well-developed stromata (25–67 μm diam) and branched, longer conidiophores (30–100 μm) (Brown & Morgan-Jones 1977), while P. taichungensis has longer and narrower conidiophores (10–25 × 1–3 μm) and shorter and narrower conidia (20–55 × 1.5–3 μm) (Hsieh & Goh 1990). Phylogenetically, P. sennae-multijugae clustered in the same clade with P. cercidis-chinensis, a species described on another member of the Fabaceae, Cersis chinensis (Fig. 1, clade 10). It is not possible to distinguish P. sennae-multijugae from numerous other Pseudocercospora spp. based solely on an ITS phylogeny, or from P. cercidis-chinensis, P. solani-pseudocapsicicola and P. pyracanthigena in the tef1 phylogeny. In the actA phylogeny it cannot be distinguished from P. acericola, P. cercidis-chinensis, P. fukuokaensis and P. mali. Morphologically, all species above differ from P. sennae-multijugae. Pseudocercospora cercidis-chinensis differs by having longer and narrower conidiophores (10–40 × 3–3.5 μm) (Shin & Braun 2000). Pseudocercospora pyracanthigena has narrower conidiophores (2–3 μm) and shorter conidia (30–45 μm) (Crous et al. 2013a), whereas P. acericola differs by having longer and wider conidiophores (10–65 × 4–5.5 μm) and longer and wider conidia (35–145 × 4–6 μm) (Chupp 1954). Pseudocercospora fukuokaensis has longer conidiophores (5–30 μm) and shorter and narrower conidia (30–70 × 2–3.5 μm) (Chupp 1954), while P. mali differs by having longer conidiophores (8–40 μm) and narrower conidia (1.5–3 μm) (Deighton 1976).

Pseudocercospora solani-pseudocapsicicola Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB814906; Fig. 20

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g020.jpg

Pseudocercospora solani-pseudocapsicicola (VIC 42807). a. Solanum pseudocapsicum with leaf spots; b. leaf spots on upper and lower leaf surface; c. close-up of lesion; d. close-up of lesion with fruiting; e, f. conidiophores emerging through stomata; g. conidiogenous cells; h–k. conidia. — Scale bars: e, f, h–k = 10 μm, g = 20 μm.

Etymology. Name derived from the plant host Solanum pseudocapsicum.

Leaf spots amphigenous, elliptical to irregular, scattered, with pale yellow areas on upper surface, 2–12 mm diam. Internal mycelium subhyaline, septate, branched, smooth, 3–5 μm diam. Stromata lacking. Conidiophores hypophyllous, in loose fascicles, arising from internal hyphae, through stomata, subcylindrical, 10–35 × 3–5 μm, 0–3-septate, straight to geniculate-sinuous, unbranched or rarely branched, pale olivaceous to pale brown, smooth. Conidiogenous cells terminal, unbranched, pale brown, subcylindrical, smooth, proliferating sympodially and percurrently, 10–27 × 3–4.5 μm. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, olivaceous to pale brown, smooth, obclavate-cylindrical, straight to curved, 42–128 × 2–3.5 μm, apex obtuse, base obconically truncate, 2–3 μm wide, 2–6-septate; hila not thickened, not darkened, 1–2.5 μm diam.

Culture characteristics — Very slow-growing (13–16 mm diam after 20 d), raised, with smooth to slightly irregularly lobate margins, aerial mycelium sparse, olivaceous grey, reverse iron-grey to green-black, sterile.

Specimen examined. BRAZIL, Minas Gerais, Viçosa, Sítio Criciúma, on leaves of Solanum pseudocapsicum (Solanaceae), 23 Jan. 2014, M. Silva (holotype VIC 42807, culture ex-type COAD 1974; isotype CBS H-22166, culture ex-isotype CPC 25229).

Notes — There are 21 species of Pseudocercospora known to occur on Solanaceae (Chupp 1954, Crous & Braun 2003). Only one species is described on Solanum pseudocapsicum, namely P. fasciculata described from Argentina (Deighton 1976). Pseudocercospora fasciculata is quite different from P. solani-pseudocapsicicola by having well-developed stroma, and longer and narrower conidiophores (80–110 × 2.5–3 μm). Two other species described on Solanaceae are morphologically more similar to P. solani-pseudocapsicicola, namely P. marcelinae described on Solanum micranthum in Argentina (Crous & Braun 2003) and P. venezuelae on Solanum argenteum in Venezuela and Brazil (Crous & Braun 2003). The former species differs from P. solani-pseudocapsicicola by having well-developed stromata, conidiophores which are shorter and narrower (5–25 × 2–4 μm) and shorter conidia (15–70 μm) (Chupp 1954), while P. venezuelae has well-developed stromata, conidiophores which are longer, arranged in dense fascicles (10–60 μm) and shorter conidia (2–4 μm) (Deighton 1976). Pseudocercospora solani-pseudocapsicicola grouped closely, but with poor support, with P. pyracanthigena (Fig. 1, clade 12), a species known to cause leaf spots on Pyracantha angustifolia (Rosaceae). Nevertheless, it is both morphologically and phylogenetically distinct from P. pyracanthigena. Pseudocercospora pyracanthigena is morphologically distinct from P. fasciculata in having shorter and narrower conidiophores (7–15 × 2–3 μm) and shorter conidia (30–45 μm) (Crous et al. 2013a). Deighton (1976) examined the original material of P. fasciculata and mentioned that “the type material is in very poor condition” and suggested that “further collections of this species are much to be desired”. An epitype therefore needs to be designated for this species. It is not possible to distinguish P. solani-pseudocapsicicola from numerous other Pseudocercospora spp. based solely on an ITS phylogeny, and it cannot be distinguished from P. cercidis-chinensis, P. sennae-multijugae and P. trinidadensis in the tef1 phylogeny. In the actA phylogeny it is closely related to P. pothomorphii (COAD 1450) and Pseudocercospora sp. (CPC 10645).

Pseudocercospora stizolobii (Syd. & P. Syd.) Deighton, Mycol. Pap. 140: 153. 1976 — Fig. 21

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g021.jpg

Pseudocercospora stizolobii (VIC 42791). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of lesion with fruiting; d, e. fasciculate conidiophores; f. conidiogenous cells; g–j. conidia. — Scale bars: d–j = 10 μm.

Basionym. Cercospora stizolobii Syd. & P. Syd., Ann. Mycol. 11: 270. 1913.

Descriptions & Illustrations — Chupp (1954: 335), Hsieh & Goh (1990: 204, f. 157).

Culture characteristics — Very slow-growing (16 mm diam after 20 d); colonies erumpent, surface folded, moderate aerial mycelium, smooth to slightly irregular lobate margins darker than the rest of the colony. Surface olivaceous grey; reverse olivaceous black.

Specimen examined. BRAZIL, Goiás, Alto Paraíso de Goiás, Parque Nacional da Chapada dos Veadeiros, on leaves of Mucuna aterrima (Fabaceae), 26 Apr. 2013, M. Silva (CBS H-22160, VIC 42791, COAD 1532, CPC 25217).

Notes — Although this species was previously reported from Brazil (Crous & Braun 2003), this study represents the first phylogenetic data for this taxon (Fig. 1, clade 7). Pseudocercospora stizolobii is distinct from other species in the tef1 and actA phylogenies, and slightly different from P. atromarginalis, P. chengtuensis and P. fuligena in the ITS phylogeny.

Pseudocercospora struthanthi U. Braun et al., Cryptog. Mycol. 23: 316. 2002 — Fig. 22

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g022.jpg

Pseudocercospora struthanthi (VIC 42766). a. Struthanthus flexicaulis with leaf spots; b. leaf spots on upper and lower leaf surface; c. fasciculate conidiophores; d–g. conidia. — Scale bars: c–g = 10 μm.

Leaf spots amphigenous, circular, 4–10 mm diam, dark brown, margin poorly defined, sometimes with the chlorotic halo. Internal mycelium indistinct. External mycelium absent. Stromata small or well-developed, 21–43 × 32–63 μm, subimmersed or erumpent, angular, brown, composed of brown textura angularis. Conidiophores amphigenous, predominantly hypophyllous, aggregated in dense fascicles, cylindrical to subcylindrical, 7.5–31 × 3–5.5 μm, 0–3-septate, straight, unbranched, brown, smooth. Conidiogenous cells terminal, 7.5–17× 3–5 μm brown, smooth, conidiophores usually reduced to conidiogenous cells. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, finely guttulate, pale brown to brown, smooth, obclavate to cylindrical, straight to curved, 41–83.5 × 3–4 μm, apex obtuse to subacute, base obconically truncate to truncate, 2.5–3 μm wide, 1–10-septate; hila unthickened, not darkened, 1–2 μm diam.

Culture characteristics — Slow-growing (20 mm diam after 20 d); colonies erumpent, surface folded with moderate aerial mycelium and smooth, lobate margins. Surface olivaceous grey surrounded by a pale olivaceous grey margin; reverse iron-grey.

Specimens examined. BRAZIL, Ceará, Fortaleza, on leaves of Struthanthus sp. (Loranthaceae), 20 June 2000, F. Freire (paratype HAL 1719); Distrito Federal, Brasília, Estação Ecológica de Águas Emendadas, on leaves of Struthanthus flexicaulis, 19 Apr. 2013, M. Silva (epitype designated here VIC 42766, MBT202070, culture ex-epitype COAD 1512; isoepitype CBS H-22157, culture ex-isoepitype CPC 25199).

Notes — The epitype of P. struthanthi designated here is morphologically similar to the holotype, particularly in morphology of conidiophores and conidia, and originates from the same country as the type. Pseudocercospora struthanthi clusters closely together with P. pipers (Fig. 1, clade 8). It is not possible to distinguish P. struthanthi from numerous other Pseudocercospora spp. based solely on an ITS or actA phylogeny, and it cannot be distinguished from P. aeschynomenicola, P. piperis and Pseudocercospora sp. CPC 10645 in the tef1 phylogeny.

Pseudocercospora tecomicola (J.M. Yen) U. Braun & Bagyan., Sydowia 51: 12. 1999 — Fig. 23

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g023.jpg

Pseudocercospora tecomicola (VIC 42687). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of lesion with fruiting; d. conidiophores in small fascicle; e–g. conidia. — Scale bars: d–g = 10 μm.

Basionym. Cercospora tecomicola J.M. Yen, Rev. Mycol. 196. 1967.

Cercoseptoria tecomicola (J.M. Yen) J.M. Yen, Gard. Bull. Singapore 33: 154. 1980.

Leaf spots amphigenous, irregular, brown, 2–10 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata almost lacking or 14–35 μm diam, subimmersed, globular, brown, composed of brown textura globosa. Conidiophores amphigenous, in small fascicles, mostly reduced to conidiogenous cells, emerging through stomata, cylindrical, 8–20 × 2–3.5 μm, 0–1-septate, straight to sinuous, unbranched, pale brown, smooth. Conidiogenous cells terminal, pale brown, cylindrical, smooth, proliferating sympodially. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, finely guttulate, pale brown, smooth, cylindrical to narrowly obclavate, straight to slightly curved, 21.5–63 × 2–4 μm, apex rounded to subacute, base truncate, 2–4 μm wide, 0–7-septate; hila neither thickened nor darkened, 1.5–2.5 μm diam.

Culture characteristics — Slow-growing (28 mm diam after 20 d); colonies circular, erumpent, surface velvety, with moderate aerial mycelium, smooth to slightly irregular margins. Surface olivaceous grey surrounded by pale olivaceous grey margin; reverse iron-grey.

Specimen examined. BRAZIL, Minas Gerais, Universidade Federal de Viçosa, on leaves of Tecoma stans (Bignoniaceae), 31 July 2013, R.W. Barreto (CBS H-22175, VIC 42687, COAD 1585, CPC 25260).

Notes — Three Pseudocercospora spp. are known to occur on species of the host genus Tecoma, viz. P. sordida on Tecoma stans, T. radicans and Tecoma sp., P. tecomicola on T. stans and P. tecomae-heterophyllae on T. heterophylla and T. undulata (Crous & Braun 2003, Farr & Rossman 2015). Pseudocercospora sordida has been previously described from Brazil on Tecoma sp. (Viégas 1945, Hanlin 1992, Crous & Braun 2003), but is morphologically and phylogenetically (Fig. 1, clade 5) quite distinct from P. tecomicola (Fig. 1, clade 6). The present Pseudocercospora collection closely matches the morphological features of P. tecomicola (Yen 1967, Bagyanarayana & Braun 1999) previously reported from Barbados and Singapore. This is the first report of P. tecomicola associated with T. stans in Brazil. It is not possible to distinguish P. tecomicola from several other Pseudocercospora spp. based solely on the ITS phylogeny, but it is distinct in the tef1 phylogeny. In the actA phylogeny it is closely related to P. nogalesii and P. wulffiae.

Pseudocercospora trinidadensis (F. Stevens & Solheim) Crous et al., Mycotaxon 72: 179. 1999 — Fig. 24

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g024.jpg

Pseudocercospora trinidadensis (VIC 42851). a. Close-up of lesion; b. close-up of leaf spot with fruiting; c. sporodochial conidiophores; d. conidiogenous cells; e–j. conidia. — Scale bars: c–j = 10 μm.

Basionym. Cercospora trinidadensis F. Stevens & Solheim, Mycologia 23: 376. 1931.

Leaf spots amphigenous, grey-brown in the centre, surrounded by a dark brown to black margin, irregular, 3–11 mm diam. Mycelium internal, subhyaline, consisting of septate, smooth hyphae, 2.5–4 μm diam. External mycelium absent. Stromata small substomatal, globular, 9–13 μm diam, composed of brown textura globosa. Conidiophores amphigenous, sporodochial, mostly reduced to conidiogenous cells, 10–22 × 3–5 μm, 0–2-septate, straight to sinuous, unbranched, pale to medium brown, smooth. Conidiogenous cells terminal, pale to medium brown, subcylindrical, smooth, proliferating sympodially, 7–15 × 3–5 μm. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, olivaceous, finely guttulate, smooth, cylindrical to narrowly obclavate, straight to slightly curved, 29–88 × 3–5 μm, apex obtuse, base obconically truncate, 3–5 μm wide, 0–14-septate; hila neither thickened nor darkened, 2–2.5 μm diam.

Culture characteristics — Slow-growing (26 mm diam after 20 d); colonies erumpent, surface velvety, with sparse aerial mycelium, smooth to slightly irregular margins, margin of colony darker than colony interior. Surface olivaceous grey; reverse olivaceous black.

Specimens examined. BRAZIL, Rio de Janeiro, Nova Friburgo, Fazenda Barreto II, on leaves of Croton urucurana (Euphorbiaceae), 1 June 2014, R.W. Barreto (CBS H-22174, VIC 42851, COAD 1756, CPC 26082).

Notes — Pseudocercospora trinidadensis was reported from Trinidad and Tobago on leaves of Croton gossypiifolius (Crous & Braun 2003). The morphology of our specimen is in agreement with the description by Crous et al. (1999), and is reported here for the first time on Croton urucurana and from Brazil. Based on the multigene phylogenetic analysis it is closely related to P. cercidis-chinensis and P. sennae-multijugae (Fig. 1, clade 10). It is not possible to distinguish P. trinidadensis from numerous other Pseudocercospora spp. based solely on the ITS phylogeny, and it could barely be distinguished from P. euphorbiacearum and P. pini-densiflorae in the tef1 phylogeny. No actA sequence of P. trinidadensis was available for comparison.

Pseudocercospora vassobiae Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813592; Fig. 25

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g025.jpg

Pseudocercospora vassobiae (VIC 42676). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of lesion with fruiting; d, e. conidiophores in a loose fascicle; f. conidiogenous cells; g–k. conidia. — Scale bars: d–k = 10 μm.

Etymology. Name derived from host genus Vassobia.

Leaf spots amphigenous, irregular, becoming vein-delimited, brown to red, 3–8 mm diam. Internal mycelium indistinct. External mycelium absent. Stromata absent. Conidiophores hypophyllous, single or in small fascicles, emerging through stomata, 20–65 × 3–4 μm, 1–5-septate, straight to slightly curved, unbranched, brown, smooth. Conidiogenous cells terminal, integrated, cylindrical, proliferating percurrently, 10–43 × 3–4 μm, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, finely guttulate, brown, smooth, cylindrical to obclavate, straight to curved, 27–108 × 3–5 μm, apex subacute to subobtuse, base obconically truncate, 2.5–4.5 μm wide, 2–10-septate; hila neither thickened nor darkened, 1–2.5 μm diam.

Culture characteristics — Slow-growing (17–20 mm diam after 20 d); raised, corrugated, aerial mycelium sparse, margins lobate, olivaceous grey, reverse olivaceous black, sterile.

Specimen examined. BRAZIL, Rio de Janeiro, Nova Friburgo, on leaves of Vassobia breviflora (Solanaceae), 9 June 2013, R.W. Barreto (holotype VIC 42676, culture ex-type COAD 1572; isotype CBS H-22173, culture ex-isotype CPC 25251).

Notes — No species of Pseudocercospora have previously been described on Vassobia breviflora. Pseudocercospora vassobiae is morphologically similar to P. solani-asperi and P. daturina. Pseudocercospora solani-asperi is distinct from P. vassobiae by having shorter and wider conidiophores (10–60 × 3–5 μm) and shorter and narrower conidia (30–80 × 3–4 μm) (Baker & Dale 1951, Deighton 1976) and P. daturina differs from P. vassobiae by having longer and wider conidiophores (30–80 × 4–6 μm) and longer conidia (51–123 μm) (Yen 1965, Deighton 1976). Phylogenetically, P. vassobiae clusters separate from other species of Pseudocercospora for which comparison of DNA sequence data is presently available (Fig. 1, clade 14). It is not possible to distinguish P. vassobiae from numerous other Pseudocercospora spp. based solely on the ITS or actA phylogenies. No tef1 sequence of P. vassobiae was available for comparison.

Pseudocercospora wulffiae Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813623; Fig. 26

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g026.jpg

Pseudocercospora wulffia (VIC 42810). a. Leaf spots on upper and lower leaf surface; b. close-up of lesion; c. close-up of lesion with fruiting; d. cross-section showing internal mycelium; e. conidiophore emerging through stomata; f. conidia. — Scale bars: d–f = 10 μm.

Etymology. Name derived from the plant host genus Wulffia, from which it was collected.

Leaf spots amphigenous, irregular, grey-brown surrounded by a dark brown margin, on lower surface medium brown, with poorly defined margin, 8–20 mm diam. Internal mycelium subhyaline, consisting of septate, branched, smooth, 3–4 μm diam hyphae. External mycelium absent. Stromata well-developed, 14–41 × 21–39 μm, immersed in the substomatal chamber, angular to irregular, medium brown, composed of brown textura angularis. Conidiophores hypophyllous, sporodochial, cylindrical, emerging through stomata, mostly reduced to conidiogenous cells, 14–21 × 2–3 μm, 0–2-septate, straight, unbranched, pale to medium brown, becoming paler toward the apex, smooth. Conidiogenous cells terminal, integrated, subcylindrical, proliferating percurrently, 8–21 × 2–3 μm, pale brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, cylindrical, apex rounded to subobtuse, straight to curved, 37.5–87 × 2–3.5 μm, base obconically truncate, 2.5–3 μm wide, 2–6-septate, pale brown, finely guttulate, smooth; hila unthickened, not darkened, 1.5–2.5 μm diam.

Culture characteristics — Slow-growing (22 mm diam after 20 d); colonies erumpent, surface folded with sparse aerial mycelium and smooth, lobate margins. Surface olivaceous grey with patches of pale olivaceous grey; reverse iron-grey to greenish black.

Specimen examined. BRAZIL, Minas Gerais, Lavras, on leaves of Wulffia stenoglossa (Asteraceae), 29 Jan. 2014, M. Silva (holotype VIC 42810, culture ex-type COAD 1976; isotype CBS H-22168, culture ex-isotype CPC 25232).

Notes — The description of Muller & Chupp (1936) of a new species of Cercospora (C. wulffiae) on Wulffia stenoglossa from Viçosa, Brazil, was invalid because it lacked a Latin diagnosis (Crous & Braun 2003). Currently, C. wulffiae is regarded as synonym of P. wedeliae (≡ Cercospora wedeliae), which occurs on different Wedelia spp. (Deighton 1976, Crous & Braun 2003). Although they have different host genera, “the morphological characteristics are nearly alike that they are considered identical” (Chupp 1954). We recollected the Pseudocercospora on Wulffia stenoglossa, and based on our phylogenetic data, we show that the species of Pseudocercospora described on Wulffia and Wedelia are different taxa. A sequence of the ITS region of P. wulffia (GenBank KT290150) possesses only 96 % similarity with the ITS sequence of P. wedeliae (GenBank KJ201940) (Kirschner & Liu 2014), confirming that they re-present different species. Also see notes under P. manihotii, to which it is phylogenetically almost identical (Fig. 1, clade 6). It is not possible to distinguish P. wulffiae from several other Pseudocercospora spp. based solely on an ITS phylogeny, and it cannot be distinguished from P. manihotii in the tef1 phylogeny. In the actA phylogeny it is closely related to P. nogalesii and P. tecomicola.

Pseudocercospora xylopiae Meir. Silva, R.W. Barreto & Crous, sp. nov. — MycoBank MB813622; Fig. 27

An external file that holds a picture, illustration, etc.
Object name is per-37-142-g027.jpg

Pseudocercospora xylopiae (VIC 42723). a, b. Xylopia aromatica with leaf spots; c. leaf spots on upper and lower leaf surface; d. close-up of lesion; e. close-up of lesion with fruiting; f, g. conidiophores in loose fascicles; h, i. conidiogenous cells; j–p. conidia. — Scale bars: f–p= 10 μm.

Etymology. Name derived from the plant host genus Xylopia.

Leaf spots amphigenous, circular to irregular, sparse, brown to red-brown, white in the centre, sometimes surrounded by a reddish chlorotic halo, 4–7 mm diam. Internal mycelium indistinct. External mycelium abundant, brown, septate, forming conidiophores. Stromata absent. Conidiophores hypophyllous, in loose fascicles, forming a dense network, climbing leaf trichomes, 5–7-septate, 15–187 × 3–5 μm, branched, brown, smooth. Conidiogenous cells terminal or intercalary, subcylindrical, proliferating sympodially, 8–20 × 2.5–4 μm, geniculate, brown, smooth. Conidiogenous loci inconspicuous, unthickened, not darkened. Conidia solitary, guttulate, pale brown, smooth, subcylindrical, straight to gently curved, 30–86.5 × 3–4.5 μm, apex obtuse, base truncate, 2.5–4 μm wide, 3–10-septate; hila unthickened, neither darkened nor refractive, 1.5–2.5 μm.

Culture characteristics — Slow-growing (16 mm diam after 20 d); colonies erumpent, surface velvety, convex, with smooth to slightly irregular margins. Surface olivaceous grey with olivaceous black border; reverse iron-grey to green-black.

Specimen examined. BRAZIL, Minas Gerais, Paraopeba, Floresta Nacional de Paraopeba (FLONA), on leaves of Xylopia aromatica (Annonaceae), 3 Jan. 2013, M. Silva (holotype VIC 42723, culture ex-type COAD 1469; isotype CBS H-22149, culture ex-isotype CPC 25173).

Notes Only one species of Pseudocercospora was known to occur on a member of Xylopia (Farr & Rossman 2015), namely P. aethiopicae on Xylopia aethiopica from Sierra Leone (Deighton 1976). Pseudocercospora aethiopicae clearly differs from P. xylopiae by having shorter and narrower conidiophores (10–40 × 2.5–4 μm), arranged in dense fascicles, and not forming on external mycelium, and having smaller conidia, 32–65 × 2.5–3 μm (Deighton 1976). Additionally, P. xylopiae does not correspond to any sequences available in GenBank at present, and is phylogenetically related to P. purpurea (Fig. 1, clade 5). Hence, it is described here as a new species. It is not possible to distinguish P. xylopiae from several other Pseudocercospora spp. based solely on an ITS phylogeny, but it is distinct in the tef1 and actA phylogenies.

DISCUSSION

This publication provides a multigene (ITS, actA and tef1) phylogenetic comparison of Pseudocercospora spp. collected from 15 host families occurring in Brazil. Currently, Pseudocercospora is recognised as genus name for the fungal holomorph, although its biology and morphological diversity are still under investigation (Braun et al. 2013, 2014, 2015, Crous et al. 2013a, Hora Júnior et al. 2014). Crous et al. (2013a) noted that significant ramifications pertaining to plant health and quarantine will only be resolved once critical taxa occurring in the Americas and Europe have been recollected from their original hosts and localities, isolated and epitypified, allowing for DNA sequence-based comparisons. This study is part of a broader project aimed at recollecting and providing molecular data for cercosporoid fungi occurring in Brazil, while also contemplating the description of newly collected species of cercosporoid fungi.

Several biomes in Brazil remain underexplored and entire plant families have never been investigated by mycologists. A recent example of the extent of the mycodiversity in Brazil awaiting discovery was provided by Guatimosim et al. (2016) who surveyed cercosporoid fungi on ferns in Brazil. These collections resulted in a significant increase in the known fern mycobiota in Brazil. Additionally, there is a complete lack of molecular information in public databases for the majority of Brazilian cercosporoid species.

The ITS barcode region (Schoch et al. 2012) was not able to differentiate many taxa at species level, resolving only 12 out of the 82 species included in the Bayesian analysis based only on the ITS alignment (data not shown, see TreeBASE). The lack of resolution of this region for Pseudocercospora was already commented on by Crous et al. (2013a) and Bakhshi et al. (2014), and is further confirmed here. The partial gene sequences of the protein-coding regions actA and tef1 were individually better (resolving each approximately half of all included species) for the identification of Pseudocercospora spp. from Brazil, as was also reported by Crous et al. (2013a) and observed for other cercosporoid genera, such as Cercospora (Groenewald et al. 2013, Bakhshi et al. 2015) and Ramularia (Videira et al. 2015). The combined phylogeny presented in Fig. 1 allows for better species discrimination than a phylogeny derived from any individual locus. Most species could be resolved, although the resolving power of the combined analysis failed for species in some clades, such as clades 8 and 9. For many of the examined species, any given locus alone is insufficient for species recognition, and requires the inclusion of at least one additional locus to resolve the species. The low resolution per individual locus also adds up in the combined alignment, ranging from low to no support values for clades containing closely related species (for example in clades 8, 9, 12 and 17). In the present study, only 11 species (P. angolensis, P. chamaecristae, P. exilis, P. fijiensis, P. guianensis, P. macrospora, P. planaltinensis, P. plunkettii, P. richardsoniicola, Pseudocercospora sp. CBS 113387 and P. udagawana) were supported as distinct by all three loci in the Bayesian phylogenies. Future work on identifying a more robust molecular marker for species discrimination in Pseudocercospora is therefore essential.

Fungi included in Pseudocercospora have been regarded as host-specific (Crous et al. 2013a, Bakhshi et al. 2014). However the same authors also reported species occurring on more than one host. There is a great need for studies involving inoculation experiments to address questions regarding host specificity of Pseudocercospora and pseudocercospora-like taxa. Furthermore, the general view of Pseudocercospora spp. being host-specific may change as molecular confirmation of species identity becomes available for more strains of a given species. The generation and public availability of phylogenetically informative gene regions of Pseudocercospora spp. is of great phytopathological importance for understanding the epidemiology of many important plant diseases. One among many examples is provided by a ‘pending enigma’, involving P. fijiensis (the aetiological agent of black Sigatoka of banana – a devastating disease of bananas and plantains). Gasparotto et al. (2005) reported this fungus as occurring on the ornamental plant Heliconia psittacorum, a member of a distinct plant family (Heliconiaceae) in Brazil. That study was based on symptomatology, fungus morphology and cross inoculations. However, the use of DNA data could lead to more conclusive evidence of the status of the fungus on H. psittacorum, which could have consequences for black Sigatoka management, including proper treatment and quarantine regulations.

The present study represents the first organized effort towards generating molecular data to support the taxonomy of Pseudocercospora spp. from Brazil. It yielded information for 27 taxa, representing only a small fraction of yet unknown species diversity in this and other genera of cercosporoid fungi. Twelve taxa found in this study represented novel species. Additionally, a further eight epitype specimens were designated, while three species were newly reported from Brazil. One of the purposes of this study was to recollect Brazilian cercosporoids described by pioneers of the discipline such as A.S. Muller and A.P. Viégas. Other cercosporoid fungi described by these authors were also recollected, and they will be treated in future publications. Many additional species still need to be recollected to enable a better understanding of what may be the largest known genus of cercosporoid fungi.

Acknowledgments

The authors thank Conselho Nacional de Desenvolvimento Científico e Tecnológico (CNPq), the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES) and Fundação de Amparo a Pesquisa do Estado de Minas Gerais (FAPEMIG) for financial support. The authors acknowledge the administration and scientific staff of Floresta Nacional de Paraopeba, Parque Nacional da Chapada dos Veadeiros and Estação Ecológica de Águas Emendadas for providing facilities and permits for the exploratory surveys of the mycodiversity in their protected areas. Olinto L. Pereira wishes to thank the Instituto Chico Mendes de Conservação da Biodiversidade (ICMBio) for permission No. 38963 and No. 37587.

REFERENCES

  • Arzanlou M, Abeln ECA, Kema GHJ, et al. 2007. Molecular diagnostics for the Sigatoka disease complex of banana. Phytopathology 97: 1112–1118. [Abstract] [Google Scholar]
  • Arzanlou M, Crous PW, Zwiers L-H. 2010. Evolutionary dynamics of mating-type loci of Mycosphaerella spp. occurring on banana. Eukaryotic Cell 9: 164–172. [Europe PMC free article] [Abstract] [Google Scholar]
  • Arzanlou M, Groenewald JZ, Fullerton RA, et al. 2008. Multiple gene genealogies and phenotypic characters differentiate several novel species of Mycosphaerella and related anamorphs on banana. Persoonia 20: 19–37. [Europe PMC free article] [Abstract] [Google Scholar]
  • Bagyanarayana G, Braun U. 1999. Phytopathogenic micromycetes from India (II). Sydowia 51: 1–19. [Google Scholar]
  • Baker RED, Dale WT. 1951. Fungi of Trinidad and Tobago. Mycological Papers 33: 1–123. [Google Scholar]
  • Bakhshi M, Arzanlou M, Babai-Ahari A, et al. 2014. Multi-gene analyses of Pseudocercospora spp. from Iran. Phytotaxa 184: 245–264. [Google Scholar]
  • Bakhshi M, Arzanlou M, Babai-Ahari A, et al. 2015. Application of the consolidated species concept to Cercospora spp. from Iran. Persoonia 34: 65–86. [Europe PMC free article] [Abstract] [Google Scholar]
  • Barreto RW, Evans HC. 1995. The mycobiota of the weed Mikania micrantha in southern Brazil with particular reference to fungal pathogens for biological control. Mycological Research 99: 343–352. [Google Scholar]
  • Batista AC, De Souza RG, Peres GEP. 1960. Alguns Cercospora estudados no IMUR. Publicações. Instituto de micologia da Universidade de Recife 262: 1–36. [Google Scholar]
  • Braun U. 1995. A monograph of Cercosporella, Ramularia and allied genera (Phytopathogenic Hyphomycetes). Vol. 1 IHW Verlag, Eching, Germany. [Google Scholar]
  • Braun U. 1999. Taxonomic notes on some species of the Cercospora complex (V). Schlechtendalia 2: 1–28. [Google Scholar]
  • Braun U. 2003. Miscellaneous notes on some cercosporoid hyphomycetes. Biblioteca Lichenologica 86: 79–98. [Google Scholar]
  • Braun U, Catañeda-Ruiz R. 1991. Cercospora and allied genera from Cuba (II). Cryptogamic Botany 2/3: 289–297. [Google Scholar]
  • Braun U, Crous PW, Kamal 2003. New species of Pseudocercospora, Pseudocercosporella, Ramularia and Stenella (cercosporoid hyphomycetes). Mycological Progress 2: 197–2008. [Google Scholar]
  • Braun U, Crous PW, Nakashima C. 2014. Cercosporoid fungi (Mycosphaerellaceae) 2. Species on monocots (Acoraceae to Xyridaceae, excluding Poaceae). IMA Fungus 5: 203–390. [Europe PMC free article] [Abstract] [Google Scholar]
  • Braun U, Crous PW, Nakashima C. 2015. Cercosporoid fungi (Mycosphaerellaceae) 3. Species on monocots (Poaceae, true grasses). IMA Fungus 6: 25–97. [Europe PMC free article] [Abstract] [Google Scholar]
  • Braun U, David J, Freire FCO. 1999. Some cercosporoid hyphomycetes from Brazil. Cryptogamie Mycologie 20: 95–106. [Google Scholar]
  • Braun U, Freire FCO. 2002. Some cercosporoid hyphomycetes from Brazil – II. Cryptogamie Mycologie 23: 295–328. [Google Scholar]
  • Braun U, Freire FCO. 2004. Some cercosporoid hyphomycetes from Brazil – III. Cryptogamie Mycologie 25: 221–244. [Google Scholar]
  • Braun U, Freire FCO. 2006. Some cercosporoid hyphomycetes from Brazil – IV. Cryptogamie Mycologie 27: 231–248. [Google Scholar]
  • Braun U, Mel’nik VA. 1997. Cercosporoid fungi from Russia and adjacent countries. Trudy Botanicheskogo Instituta Imeni V.L. Komarova, Rossijskaya Akademiya Nauk St. Petersburg 20: 1–130. [Google Scholar]
  • Braun U, Nakashima C, Crous PW. 2013. Cercosporoid fungi (Mycosphaerellaceae) 1. Species on other fungi, Pteridophyta and Gymnospermae. IMA Fungus 4: 265–345. [Europe PMC free article] [Abstract] [Google Scholar]
  • Brown LG, Morgan-Jones G. 1977. Notes on Hyphomycetes. XX. ‘Cercospora-complex’ fungi of Cassia and Psoralea. Mycotaxon 6: 261–276. [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]
  • Chupp C. 1954. A monograph of the fungus genus Cercospora. Published by the author, Ithaca, New York, USA. [Google Scholar]
  • Churchill ACL. 2011. Mycosphaerella fijiensis, the black leaf streak pathogen of banana: progress towards understanding pathogen biology and detection, disease development, and the challenges of control. Molecular Plant Pathology 12: 307–328. [Europe PMC free article] [Abstract] [Google Scholar]
  • Crous PW. 1998. Mycosphaerella spp. and their anamorphs associated with leaf spot diseases of Eucalyptus. Mycologia Memoir 21: 1–170. [Google Scholar]
  • Crous PW, Alfenas AC, Barreto RW. 1997. Cercosporoid fungi from Brazil. 1. Mycotaxon 64: 405–430. [Google Scholar]
  • Crous PW, Aptroot A, Kang JC, et al. 2000. The genus Mycosphaerella and its anamorphs. Studies in Mycology 45: 107–121. [Google Scholar]
  • Crous PW, Braun U. 2003. Mycosphaerella and its anamorphs: 1. Names published in Cercospora and Passalora. CBS Biodiversity Series 1: 1–571. CBS-KNAW Fungal Biodiversity Centre, Utrecht, The Netherlands. [Google Scholar]
  • Crous PW, Braun U, Alfenas AC. 1999. Cercosporoid fungi from Brazil. 3. Mycotaxon 72: 171–193. [Google Scholar]
  • Crous PW, Braun U, Hunter GC, et al. 2013a. Phylogenetic lineages in Pseudocercospora. Studies in Mycology 75: 37–114. [Europe PMC free article] [Abstract] [Google Scholar]
  • Crous PW, Câmara MPS. 1998. Cercosporoid fungi from Brazil. 2. Mycotaxon 68: 299–310. [Google Scholar]
  • Crous PW, Gams W, Stalpers JA, et al. 2004. MycoBank: an online initiative to launch mycology into the 21st century. Studies in Mycology 50: 19–22. [Google Scholar]
  • Crous PW, Hawksworth DL, Wingfield MJ. 2015. Identifying and naming plant-pathogenic fungi: past, present, and future. Annual Review of Phytopathology 53: 247–267. [Abstract] [Google Scholar]
  • Crous PW, Kang JC, Braun U. 2001. A phylogenetic redefinition of anamorph genera in Mycosphaerella based on ITS rDNA sequence and morphology. Mycologia 93: 1081–1101. [Google Scholar]
  • Crous PW, Liebenberg MM, Braun U, et al. 2006. Re-evaluating the taxonomic status of Phaeoisariopsis griseola, the causal agent of angular leaf spot of bean. Studies in Mycology 55: 163–173. [Europe PMC free article] [Abstract] [Google Scholar]
  • Crous PW, Quaedvlieg W, Sarpkaya K, et al. 2013b. Septoria-like pathogens causing leaf and fruit spot of pistachio. IMA Fungus 4: 187–199. [Europe PMC free article] [Abstract] [Google Scholar]
  • Crous PW, Verkley GJM, Groenewald JZ, et al. (eds). 2009. Fungal Biodiversity. CBS Laboratory Manual Series 1: 1–269. CBS-KNAW Fungal Biodiversity Centre, Utrecht, The Netherlands. [Google Scholar]
  • Deighton FC. 1976. Studies on Cercospora and allied genera. VI. Pseudocercospora Speg., Pantospora Cif. and Cercoseptoria Petr. Mycological Papers 140: 1–168. [Google Scholar]
  • Farr DF, Rossman AY. 2015. Fungal databases, systematic mycology and microbiology laboratory, ARS, USDA. Retrieved March 4, 2015, from http://nt.ars-grin.gov/fungaldatabases/. [Google Scholar]
  • Frank J, Crous PW, Groenewald JZ, et al. 2010. Microcyclospora and Microcyclosporella: novel genera accommodating epiphytic fungi causing sooty blotch on apple. Persoonia 24: 93–105. [Europe PMC free article] [Abstract] [Google Scholar]
  • Furlanetto C, Dianese JC. 1999. Some Pseudocercospora species and a new Prathigada species from the Brazilian cerrado. Mycological Research 103: 1203–1209. [Google Scholar]
  • Gasparotto L, Pereira JCR, Urben AF, et al. 2005. Heliconia psittacorum: hospedeira de Mycosphaerella fijiensis, agente causal da sigatoka-negra da bananeira. Fitopatologia Brasileira 30: 423–425. [Google Scholar]
  • Goh TK, Hsieh WH. 1989. Studies on Cercospora and allied genera of Taiwan (VI). Transactions of the Mycological Society of the Republic of China 4 (2–3): 1–23. [Google Scholar]
  • Groenewald JZ, Nakashima C, Nishikawa J, et al. 2013. Species concepts in Cercospora: spotting the weeds among the roses. Studies in Mycology 75: 115–170. [Europe PMC free article] [Abstract] [Google Scholar]
  • Guatimosim E, Schwartsburd PB, Barreto RW, et al. 2016. Novel fungi from an old niche: cercosporoid and related sexual morphs on ferns. Persoonia 37: 106–141. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hanlin RT. 1992. Index to genera and species of ascomycetes described by A.P. Viégas. Mycotaxon 43: 207–230. [Google Scholar]
  • Hawksworth DL, Crous PW, Redhead SA, et al. 2011. The Amsterdam Declaration on Fungal Nomenclature. IMA Fungus 2: 105–112. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hernández-Gutiérrez A, Braun U, Dianese JC. 2014. Cercosporoid hyphomycetes on malpighiaceous hosts from the Brazilian Cerrado: species of Pseudocercospora on hosts belonging to Byrsonima. Mycological Progress 13: 193–210. [Google Scholar]
  • Hernández-Gutiérrez A, Dianese JC. 2009. New cercosporoid fungi from the Brazilian Cerrado 2. Species on hosts of the subfamilies Caesalpinioideae, Faboideae and Mimosoideae (Leguminosae s. lat.). Mycotaxon 107: 1–24. [Google Scholar]
  • Hernández-Gutiérrez A, Dianese JC. 2014. Cercosporoid hyphomycetes on malpighiaceous hosts from the Brazilian Cerrado: New Passalora and Pseudocercospora species on hosts of the genus Banisteriopsis. Mycological Progress 13: 365–371. [Google Scholar]
  • Hora Júnior BT, Macedo DM, Barreto RW, et al. 2014. Erasing the past: a new identity for the Damoclean pathogen causing South American leaf blight of rubber. PloS ONE 9: e104750. [Europe PMC free article] [Abstract] [Google Scholar]
  • Hsieh WH, Goh TK. 1990. Cercospora and similar fungi from Taiwan. Maw Chang Book Company, Taipei, Taiwan. [Google Scholar]
  • Hunter GC, Wingfield BD, Crous PW, et al. 2006. A multi-gene phylogeny for species of Mycosphaerella occurring on Eucalyptus leaves. Studies in Mycology 55: 147–161. [Europe PMC free article] [Abstract] [Google Scholar]
  • Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Molecular Biology and Evolution 30: 772–780. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kearse M, Moir R, Wilson A, et al. 2012. Geneious basic: an integrated and extendable desktop software platform for the organization and analysis of sequence data. Bioinformatics 28: 1647–1649. [Europe PMC free article] [Abstract] [Google Scholar]
  • Kimati H, Amorim L, Rezende JAM, et al. 2005. Manual de Fitopatologia. Volume 2: Doenças das Plantas Cultivadas. 4th Edition Editora Agronômica Ceres Ltda; São Paulo, Brazil. [Google Scholar]
  • Kirschner R. 2014. A new species and new records of cercosporoid fungi from ornamental plants in Taiwan. Mycological Progress 13: 483–491. [Google Scholar]
  • Kirschner R, Liu L-C. 2014. Mycosphaerellaceous fungi and new species of Venustosynnema and Zasmidium on ferns and fern allies in Taiwan. Phytotaxa 176: 309–323. [Google Scholar]
  • Maddison WP, Maddison DR. 2011. Mesquite: a molecular system for evolutionary analysis. Version 2.75. http://mesquiteproject.org. [Google Scholar]
  • Minnis AM, Kennedy AH, Grenier DB, et al. 2011. Asperisporium and Pantospora (Mycosphaerellaceae): epitypifications and phylogenetic placement. Persoonia 27: 1–8. [Europe PMC free article] [Abstract] [Google Scholar]
  • Muller AS, Chupp C. 1934. Cercosporae de Minas Gerais. Arquivos do Instituto de Biologia Vegetal Rio de Janeiro 1: 213–220. [Google Scholar]
  • Muller AS, Chupp C. 1936. Uma segunda contribuição às Cercosporae de Minas Gerais. Arquivos do Instituto de Biologia Vegetal Rio de Janeiro 3: 91–97. [Google Scholar]
  • Nguanhom J, Cheewangkoon R, Groenewald JZ, et al. 2015. Taxonomy and phylogeny of Cercospora spp. from Northern Thailand. Phytotaxa 233: 27–48. [Google Scholar]
  • Nylander JAA. 2004. MrModeltest v. 2.2. Program distributed by the author. Evolutionary Biology Centre, Uppsala University. [Google Scholar]
  • O’Donnell K, Kistler HC, Cigelnik E, et al. 1998. Multiple evolutionary origins of the fungus causing Panama disease of banana: concordant evidence from nuclear and mitochondrial gene genealogies. Proceedings of the National Academy of Sciences of the United States of America 95: 2044–2049. [Europe PMC free article] [Abstract] [Google Scholar]
  • Parreira DF, Silva M, Pereira OL, et al. 2014. Cercosporoid hyphomycetes associated with Tibouchina herbaceae (Melastomataceae) in Brazil. Mycological Progress 13: 691–702. [Google Scholar]
  • Pereira OL, Barreto RW. 2005. The mycobiota of the weed Mitracarpus hirtus in Minas Gerais (Brazil) with particular reference to fungal pathogens for biological control. Australasian Plant Pathology 34: 41–50. [Google Scholar]
  • Pereira OL, Barreto RW. 2006. Pseudocercospora palicoureae sp. nov. associated with the toxic rubiaceous weed Palicourea marcgravii in Brazil, with observations on its mycobiota. Fungal Diversity 23: 243–253. [Google Scholar]
  • Pereira OL, Barreto RW. 2007. The mycobiota of the cactus weed Pereskia aculeate in Brazil, with comments on the life-cycle of Uromyces pereskiae. Fungal Diversity 25: 127–140. [Google Scholar]
  • Pereira OL, Barreto RW, Ellison CA, et al. 2003. Corynespora cassiicola f. sp. lantanae: a potential biocontrol agent for Lantana camara from Brazil. Biological Control 26: 21–31. [Google Scholar]
  • Quaedvlieg W, Binder M, Groenewald JZ, et al. 2014. Introducing the Consolidated Species Concept to resolve species in the Teratosphaeriaceae. Persoonia 33: 1–40. [Europe PMC free article] [Abstract] [Google Scholar]
  • Rannala B, Yang Z. 1996. Probability distribution of molecular evolutionary trees: a new method of phylogenetic inference. Journal of Molecular Evolution 43: 304–311. [Abstract] [Google Scholar]
  • Rayner RW. 1970. A mycological colour chart. CMI and British Mycological Society, Kew, Surrey, England. [Google Scholar]
  • Rehner SA, Samuels GJ. 1994. Taxonomy and phylogeny of Gliocladium analysed from nuclear large subunit ribosomal DNA sequences. Mycological Research 98: 625–634. [Google Scholar]
  • Rocha FB, Hanada RE, Albuquerque ST, et al. 2013. Pseudocercospora piperis associated with leaf spots on Piper aduncum in Brazil. Australasian Plant Disease Notes 8: 101–103. [Google Scholar]
  • Rocha FB, Soares DJ, Barreto RW. 2008. Pseudocercospora species on Piperaceae from Viçosa, Minas Gerais, Brazil. Mycological Progress 7: 249–252. [Google Scholar]
  • Rodriguez F, Oliver JF, Marin A, et al. 1990. The general stochastic model of nucleotide substitutions. Journal of Theoretical Biology 142: 485–501. [Abstract] [Google Scholar]
  • Ronquist F, Huelsenbeck JP. 2003. MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572–1574. [Abstract] [Google Scholar]
  • Schoch CL, Seifert KA, Huhndorf S, et al. 2012. Nuclear ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode marker for fungi. Proceedings of the National Academy of Sciences of the United States of America 109: 6241–6246. [Europe PMC free article] [Abstract] [Google Scholar]
  • Shin HD, Braun U. 2000. Notes on Korean Cercosporae and allied genera (III). Mycotaxon 74: 105–118. [Google Scholar]
  • Silva JL, Barreto RW, Pereira OL. 2008. Pseudocercospora cryptostegiae-madagascariensis sp. nov. on Cryptostegia madagascariensis, an exotic vine involved in major biological invasions in Northeast Brazil. Mycopathologia 165: 364–367. [Abstract] [Google Scholar]
  • Soares DJ, Barreto RW. 2008. Fungal pathogens of the invasive riparian weed Hedychium coronarium from Brazil and their potential for biological control. Fungal Diversity 28: 85–96. [Google Scholar]
  • Spegazzini C. 1910. Mycetes Argentinenses (Series V). Anales del Museo Nacional de Historia Natural, Buenos Aires 20: 329–467. [Google Scholar]
  • Stewart EL, Liu Z, Crous PW, et al. 1999. Phylogenetic relationships among some cercosporoid anamorphs of Mycosphaerella based on rDNA sequence analysis. Mycological Research 103: 1491–1499. [Google Scholar]
  • Tamura K, Stecher G, Peterson D, et al. 2013. MEGA6: Molecular Evolutionary Genetics Analysis version 6.0. Molecular Biology and Evolution 30: 2725–2729. [Europe PMC free article] [Abstract] [Google Scholar]
  • Videira SIR, Groenewald JZ, Kolecka A, et al. 2015. Eluciding the Ramularia eucalypti species complex. Persoonia 34: 50–64. [Europe PMC free article] [Abstract] [Google Scholar]
  • Viégas AP. 1945. Alguns fungos do Brasil – Cercosporae. Boletim de Sociedade Brasileira de Agronomia 8: 1–160. [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]
  • White TJ, Bruns T, Lee S, et al. 1990. Amplification and direct sequencing of fungal ribosomal RNA genes for phylogenetics. In: Innis MA, Gelfand DH, Sninsky JJ, et al. (eds), A guide to methods and applications: 315–322. Academic Press, New York. [Google Scholar]
  • Wingfield MJ, De Beer ZW, Slippers B, et al. 2012. One fungus, one name promotes progressive plant pathology. Molecular Plant Pathology 13: 604–613. [Europe PMC free article] [Abstract] [Google Scholar]
  • Yen JM. 1965. Etude sur les champignous parasites du sud-est Asiatigue. III. Deuxième note sur quelques nouvelles espèces de Cercospora de Singapour. Revue Mycologie 30: 166–204. [Google Scholar]
  • Yen JM. 1967. Etude sur les champignous parasites du sud-est Asiatigue. VII: Quatrième note sur quelques Cercospora et stenella de Singapour (Malaisie). Première note sur quelques nouvelles espèces de Cercospora de Singapour. Revue Mycologie 32: 177–202. [Google Scholar]
  • Yen JM. 1983. Studies on parasitic fungi from South East Asia, 49. Parasitic fungi from Malaysia 25: Semipseudocercospora gen. nov. Mycotaxon 17: 361 1983. Studies on parasitic fungi from South East Asia, 49. Parasitic fungi from Malaysia 25: Semipseudocercospora gen. nov. Mycotaxon 17: 361–363. [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

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/17784480
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/17784480

Article citations


Go to all (9) 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.