06.04.2013 Views

The Colletotrichum gloeosporioides species complex - CBS - KNAW

The Colletotrichum gloeosporioides species complex - CBS - KNAW

The Colletotrichum gloeosporioides species complex - CBS - KNAW

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Studies in Mycology<br />

available online at www.studiesinmycology.org<br />

<strong>The</strong> <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

B.S. Weir 1* , P.R. Johnston 1 , and U. Damm 2<br />

1 Landcare Research, Private Bag 92170 Auckland, New Zealand; 2 <strong>CBS</strong>-<strong>KNAW</strong> Fungal Biodiversity Centre, Uppsalalaan 8, 3584 CT Utrecht, <strong>The</strong> Netherlands<br />

*Correspondence: Bevan Weir, WeirB@LandcareResearch.co.nz<br />

Abstract: <strong>The</strong> limit of the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is defined genetically, based on a strongly supported clade within the <strong>Colletotrichum</strong> ITS gene tree.<br />

All taxa accepted within this clade are morphologically more or less typical of the broadly defined C. <strong>gloeosporioides</strong>, as it has been applied in the literature for the past 50 years.<br />

We accept 22 <strong>species</strong> plus one sub<strong>species</strong> within the C. <strong>gloeosporioides</strong> <strong>complex</strong>. <strong>The</strong>se include C. asianum, C. cordylinicola, C. fructicola, C. <strong>gloeosporioides</strong>, C. horii, C.<br />

kahawae subsp. kahawae, C. musae, C. nupharicola, C. psidii, C. siamense, C. theobromicola, C. tropicale, and C. xanthorrhoeae, along with the taxa described here as new,<br />

C. aenigma, C. aeschynomenes, C. alatae, C. alienum, C. aotearoa, C. clidemiae, C. kahawae subsp. ciggaro, C. salsolae, and C. ti, plus the nom. nov. C. queenslandicum<br />

(for C. <strong>gloeosporioides</strong> var. minus). All of the taxa are defined genetically on the basis of multi-gene phylogenies. Brief morphological descriptions are provided for <strong>species</strong><br />

where no modern description is available. Many of the <strong>species</strong> are unable to be reliably distinguished using ITS, the official barcoding gene for fungi. Particularly problematic<br />

are a set of <strong>species</strong> genetically close to C. musae and another set of <strong>species</strong> genetically close to C. kahawae, referred to here as the Musae clade and the Kahawae clade,<br />

respectively. Each clade contains several <strong>species</strong> that are phylogenetically well supported in multi-gene analyses, but within the clades branch lengths are short because of<br />

the small number of phylogenetically informative characters, and in a few cases individual gene trees are incongruent. Some single genes or combinations of genes, such as<br />

glyceraldehyde-3-phosphate dehydrogenase and glutamine synthetase, can be used to reliably distinguish most taxa and will need to be developed as secondary barcodes for<br />

<strong>species</strong> level identification, which is important because many of these fungi are of biosecurity significance. In addition to the accepted <strong>species</strong>, notes are provided for names<br />

where a possible close relationship with C. <strong>gloeosporioides</strong> sensu lato has been suggested in the recent literature, along with all subspecific taxa and formae speciales within<br />

C. <strong>gloeosporioides</strong> and its putative teleomorph Glomerella cingulata.<br />

Key words: anthracnose, Ascomycota, barcoding, <strong>Colletotrichum</strong> <strong>gloeosporioides</strong>, Glomerella cingulata, phylogeny, systematics.<br />

Taxonomic novelties: Name replacement - C. queenslandicum B. Weir & P.R. Johnst. New <strong>species</strong> - C. aenigma B. Weir & P.R. Johnst., C. aeschynomenes B. Weir & P.R.<br />

Johnst., C. alatae B. Weir & P.R. Johnst., C. alienum B. Weir & P.R. Johnst, C. aotearoa B. Weir & P.R. Johnst., C. clidemiae B. Weir & P.R. Johnst., C. salsolae B. Weir & P.R.<br />

Johnst., C. ti B. Weir & P.R. Johnst. New sub<strong>species</strong> - C. kahawae subsp. ciggaro B. Weir & P.R. Johnst. Typification: Epitypification - C. queenslandicum B. Weir & P.R.<br />

Johnst.<br />

Published online: 21 August 2012; doi:10.3114/sim0011. Hard copy: September 2012.<br />

INTRODUCTION<br />

<strong>The</strong> name <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> was first proposed<br />

in Penzig (1882), based on Vermicularia <strong>gloeosporioides</strong>, the<br />

type specimen of which was collected from Citrus in Italy. Much<br />

of the early literature used this name to refer to fungi associated<br />

with various diseases of Citrus, with other <strong>species</strong> established for<br />

morphologically similar fungi from other hosts. However, several<br />

early papers discussed the morphological similarity between many<br />

of the <strong>Colletotrichum</strong> spp. that had been described on the basis of<br />

host preference, and used inoculation tests to question whether or<br />

not the <strong>species</strong> were distinct. Some of these papers investigated<br />

in culture the link between the various <strong>Colletotrichum</strong> <strong>species</strong> and<br />

their sexual Glomerella state (e.g. Shear & Wood 1907, Ocfemia<br />

& Agati 1925). Authors such as Shear & Wood (1907, 1913) and<br />

Small (1926) concluded that many of the <strong>species</strong> described on the<br />

basis of host preference were in fact the same, rejecting apparent<br />

differences in host preference as a basis for taxonomic segregation.<br />

Small (1926) concluded that the names Glomerella cingulata and<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> should be used for the sexual and<br />

asexual morphs, respectively, of the many <strong>Colletotrichum</strong> spp.<br />

they regarded as conspecific. <strong>Colletotrichum</strong> <strong>gloeosporioides</strong><br />

was stated to be the earliest name with a proven link to what they<br />

Copyright <strong>CBS</strong>-<strong>KNAW</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherlands.<br />

StudieS in Mycology 73: 115–180.<br />

regarded as a biologically diverse G. cingulata. <strong>The</strong> studies of<br />

von Arx & Müller (1954) and von Arx (1957, 1970) taxonomically<br />

formalised this concept.<br />

<strong>The</strong> “von Arxian” taxonomic concept for <strong>Colletotrichum</strong><br />

saw large numbers of <strong>species</strong> synonymised with the names C.<br />

graminicola (for grass-inhabiting <strong>species</strong>) and C. <strong>gloeosporioides</strong><br />

(for non-grass inhabiting <strong>species</strong> with straight conidia). <strong>The</strong><br />

genetic and biological diversity encompassed by these names<br />

was so broad that they became of little practical use to plant<br />

pathologists, conveying no information about pathogenicity, host<br />

range, or other attributes. <strong>The</strong> von Arx & Müller (1954) and von<br />

Arx (1957) studies were not based on direct examination of type<br />

material of all <strong>species</strong> and some of the synonymy proposed<br />

in these papers has subsequently been found to be incorrect.<br />

Examples include the segregation of C. acutatum (Simmonds<br />

1965) and C. boninense (Moriwaki et al. 2003) from C.<br />

<strong>gloeosporioides</strong> sensu von Arx (1957). Other studies published<br />

elsewhere in this volume (Damm et al. 2012a, b) show that<br />

several <strong>species</strong> regarded as synonyms of C. <strong>gloeosporioides</strong><br />

by von Arx (1957) are members of the C. acutatum <strong>complex</strong><br />

(e.g. C. godetiae, Gloeosporium limetticola, G. lycopersici, and<br />

G. phormii) or the C. boninense <strong>complex</strong> (e.g. C. dracaenae).<br />

Recent molecular studies have resulted in a much better<br />

understanding of phylogenetic relationships amongst the<br />

You are free to share - to copy, distribute and transmit the work, under the following conditions:<br />

Attribution: You must attribute the work in the manner specified by the author or licensor (but not in any way that suggests that they endorse you or your use of the work).<br />

Non-commercial: You may not use this work for commercial purposes.<br />

No derivative works: You may not alter, transform, or build upon this work.<br />

For any reuse or distribution, you must make clear to others the license terms of this work, which can be found at http://creativecommons.org/licenses/by-nc-nd/3.0/legalcode. Any of the above conditions can be waived if you get<br />

permission from the copyright holder. Nothing in this license impairs or restricts the author’s moral rights.<br />

115


Weir et al.<br />

grass-inhabiting <strong>species</strong> of the C. graminicola group and the<br />

development of a more useful taxonomy for this group of fungi<br />

(e.g. Hsiang & Goodwin 2001, Du et al. 2005, and Crouch et<br />

al. 2006). This group is now recognised as comprising several<br />

host-specialised, genetically well characterised <strong>species</strong>, but a<br />

modern taxonomy for C. <strong>gloeosporioides</strong> has yet to be resolved.<br />

Von Arx (1970) and Sutton (1980) distinguished the C.<br />

<strong>gloeosporioides</strong> group using conidial shape and size. A few apparently<br />

host-specialised, C. <strong>gloeosporioides</strong>-like taxa were retained by these<br />

authors, but the basis of their identification was often difficult to<br />

understand. Prior to the availability of DNA sequence data, taxonomic<br />

concepts within <strong>Colletotrichum</strong> were based on features such as host<br />

<strong>species</strong>, substrate, conidial size and shape, shape of appressoria,<br />

growth rate in culture, colour of cultures, presence or absence of<br />

setae, whether or not the teleomorph develops, etc. Some studies<br />

have found characters such as these useful for distinguishing<br />

groups within C. <strong>gloeosporioides</strong> (e.g. Higgins 1926, Gorter 1956,<br />

Hindorf 1973, and Johnston & Jones 1997). However, problems<br />

arise because many of these morphological features change under<br />

different conditions of growth (dependent upon growth media,<br />

temperature, light regime, etc.), or can be lost or change with repeated<br />

subculturing. Host preference is poorly controlled — even good, welldefined<br />

pathogens causing a specific disease can be isolated by<br />

chance from other substrates (e.g. Johnston 2000). <strong>Colletotrichum</strong><br />

conidia will germinate on most surfaces, form an appressorium,<br />

remain attached to that surface as a viable propagule or perhaps as<br />

a minor, endophytic or latent infection, and grow out from there into<br />

senescing plant tissue or onto agar plates if given the opportunity.<br />

In addition, the same disease can be caused by genetically distinct<br />

sets of isolates, the shared pathogenicity presumably independently<br />

evolved, e.g. the bitter rot disease of apple is caused by members<br />

of both the C. acutatum and C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>es<br />

(Johnston et al. 2005).<br />

Sutton (1992) commented on C. <strong>gloeosporioides</strong> that “No<br />

progress in the systematics and identification of isolates belonging<br />

to this <strong>complex</strong> is likely to be made based on morphology alone”.<br />

A start was made towards a modern understanding of this name<br />

with the designation of an epitype specimen with a culture derived<br />

from it to stabilise the application of the name (Cannon et al. 2008).<br />

Based on ITS sequences, the ex-epitype isolate belongs in a<br />

strongly supported clade, distinct from other taxa that have been<br />

confused with C. <strong>gloeosporioides</strong> in the past, such as C. acutatum<br />

and C. boninense (e.g. Abang et al. 2002, Martinez-Culebras et<br />

al. 2003, Johnston et al. 2005, Chung et al. 2006, Farr et al. 2006,<br />

Than et al. 2008). However, biological and genetic relationships<br />

within the broad C. <strong>gloeosporioides</strong> clade remain confused and ITS<br />

sequences alone are insufficient to resolve them.<br />

In this study we define the limits of the C. <strong>gloeosporioides</strong><br />

<strong>species</strong> <strong>complex</strong> on the basis of ITS sequences, the <strong>species</strong> we<br />

accept within the <strong>complex</strong> forming a strongly supported clade<br />

in the ITS gene tree (fig. 1 in Cannon et al. 2012, this issue). In<br />

all cases the taxa we include in the C. <strong>gloeosporioides</strong> <strong>complex</strong><br />

would fit within the traditional morphological concept of the C.<br />

<strong>gloeosporioides</strong> group (e.g. von Arx 1970, Mordue 1971, and<br />

Sutton 1980). Commonly used <strong>species</strong> names within the C.<br />

<strong>gloeosporioides</strong> <strong>complex</strong> include C. fragariae, C. musae, and C.<br />

kahawae. Since the epitype paper (Cannon et al. 2008), several<br />

new C. <strong>gloeosporioides</strong>-like <strong>species</strong> have been described in<br />

regional studies, where multi-gene analyses have shown the new<br />

<strong>species</strong> to be phylogenetically distinct from the ex-epitype strain of<br />

C. <strong>gloeosporioides</strong> (e.g. Rojas et al. 2010, Phoulivong et al. 2011,<br />

and Wikee et al. 2011).<br />

116<br />

<strong>The</strong> regional nature of most of these studies, the often restricted<br />

genetic sampling across the diversity of C. <strong>gloeosporioides</strong> globally,<br />

and the minimal overlap between isolates treated and gene regions<br />

targeted in the various studies, means that the relationship between<br />

the newly described <strong>species</strong> is often poorly understood.<br />

While some authors have embraced a genetically highly<br />

restricted concept for C. <strong>gloeosporioides</strong>, many applied researchers<br />

continue to use the name in a broad, group-<strong>species</strong> concept (e.g.<br />

Bogo et al. 2012, Deng et al. 2012, Kenny et al. 2012, Parvin et<br />

al. 2012, and Zhang et al. 2012). In this paper we accept both<br />

concepts as useful and valid. When used in a broad sense, we<br />

refer to the taxon as the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> or C.<br />

<strong>gloeosporioides</strong> s. lat.<br />

This paper aims to clarify the genetic and taxonomic<br />

relationships within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> using<br />

a set of isolates that widely samples its genetic, biological and<br />

geographic diversity. Type specimens, or cultures derived from<br />

type specimens, have been examined wherever possible. Although<br />

we do not treat all of the names placed in synonymy with C.<br />

<strong>gloeosporioides</strong> or Glomerella cingulata by von Arx & Müller (1954)<br />

and von Arx (1957, 1970), we treat all names for which a possible<br />

close relationship with C. <strong>gloeosporioides</strong> has been suggested in<br />

the recent literature, along with all subspecific taxa and formae<br />

speciales within C. <strong>gloeosporioides</strong> and G. cingulata.<br />

ITS sequences, the official barcoding gene for fungi (Seifert<br />

2009, Schoch et al. 2012), do not reliably resolve relationships within<br />

the C. <strong>gloeosporioides</strong> <strong>complex</strong>. We define <strong>species</strong> in the <strong>complex</strong><br />

genetically rather than morphologically, on the basis of phylogenetic<br />

analyses of up to eight genes. Following Cannon et al. (2012, this<br />

issue) the generic name <strong>Colletotrichum</strong> is used as the preferred<br />

generic name for all <strong>species</strong> wherever possible throughout this<br />

paper, whether or not a Glomerella state has been observed for that<br />

fungus, and whether or not the Glomerella state has a formal name.<br />

MATERIALS AND METHODS<br />

Specimen isolation and selection<br />

An attempt was made to sample the genetic diversity across C.<br />

<strong>gloeosporioides</strong> as widely as possible, with isolates from diverse<br />

hosts from around the world selected for more intensive study. A<br />

BLAST search of GenBank using the ITS sequence of the epitype<br />

culture of C. <strong>gloeosporioides</strong> (Cannon et al. 2008) provided a<br />

coarse estimate for the genetic limit of the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> and ITS diversity across the <strong>complex</strong> was used to<br />

select a genetically diverse set of isolates. Voucher cultures were<br />

obtained from the research groups who deposited the GenBank<br />

records. To these were added isolates representing the known<br />

genetic and morphological diversity of C. <strong>gloeosporioides</strong> from<br />

New Zealand, isolated from rots of native and introduced fruits,<br />

from diseased exotic weeds, and as endophytes from leaves of<br />

native podocarps. Additional isolates representing ex-type and<br />

authentic cultures of as many named taxa and formae speciales<br />

within the C. <strong>gloeosporioides</strong> <strong>complex</strong> as possible were obtained<br />

from international culture collections. Approximately 400 isolates<br />

belonging to the C. <strong>gloeosporioides</strong> <strong>complex</strong> were obtained.<br />

GAPDH gene sequences were generated for all isolates as an initial<br />

measure of genetic diversity. A subset of 156 isolates, selected to<br />

represent the range of genetic, geographic, and host plant diversity,<br />

was used in this research (Table 1).


Most of the New Zealand isolates had been stored as conidial<br />

suspensions made from single conidium or ascospore cultures and<br />

then stored at -80 °C in a 5 % glycerol/water suspension. Additional<br />

isolates from New Zealand were obtained from the ICMP culture<br />

collection, where isolates are stored as lyophilised (freeze-dried)<br />

ampoules or in a metabolically inactive state in liquid nitrogen<br />

at -196 °C. <strong>The</strong> storage history of most of the isolates received<br />

from other research groups is not known. Table 1 lists the isolates<br />

studied. All those supplying cultures are acknowledged at the<br />

end of this manuscript, and additional details on each culture are<br />

available on the ICMP website (http://www.landcareresearch.co.nz/<br />

resources/collections/icmp).<br />

Culture collection and fungal herbarium (fungarium)<br />

abbreviations used herein are: <strong>CBS</strong> = Centraalbureau voor<br />

Schimmelcultures (Netherlands), ICMP = International Collection<br />

of Microorganisms from Plants, MFLU = Mae Fah Luang University<br />

Herbarium (Thailand) MFLUCC = Mae Fah Luang University<br />

Culture Collection (Thailand), GCREC = University of Florida,<br />

Gulf Coast Research and Education Centre (USA), HKUCC = <strong>The</strong><br />

University of Hong Kong Culture Collection (China), IMI = CABI<br />

Genetic Resource Collection (UK), MAFF = Ministry of Agriculture,<br />

Forestry and Fisheries (Japan), DAR = Plant Pathology Herbarium<br />

(Australia), NBRC = Biological Resource Center, National Institute<br />

of Technology and Evaluation (Japan), BCC = BIOTEC Culture<br />

Collection (Thailand), GZAAS = Guizhou Academy of Agricultural<br />

Sciences herbarium (China), MUCL = Belgian Co-ordinated<br />

Collections of Micro-organisms, (agro)industrial fungi & yeasts<br />

(Belgium), BRIP = Queensland Plant Pathology Herbarium<br />

(Australia), PDD = New Zealand Fungal and Plant Disease<br />

Collection (New Zealand), BPI = U.S. National Fungus Collections<br />

(USA), STE-U = Culture collection of the Department of Plant<br />

Pathology, University of Stellenbosch (South Africa), and MCA =<br />

M. Catherine Aime’s collection series, Louisiana State University<br />

(USA).<br />

DNA extraction, amplification, and sequencing<br />

Mycelium was collected from isolates grown on PDA agar, and<br />

manually comminuted with a micropestle in 420 μL of Quiagen<br />

DXT tissue digest buffer; 4.2 μL of proteinase K was added and<br />

incubated at 55 °C for 1 h. After a brief centrifugation 220 μL of<br />

the supernatant was placed in a Corbett X-tractorGene automated<br />

nucleic acid extraction robot. <strong>The</strong> resulting 100 μL of pure DNA in<br />

TE buffer was stored at -30 °C in 1.5 mL tubes until use.<br />

Gene sequences were obtained from eight nuclear gene regions,<br />

actin (ACT) [316 bp], calmodulin (CAL) [756 bp], chitin synthase<br />

(CHS-1) [229 bp], glyceraldehyde-3-phosphate dehydrogenase<br />

(GAPDH) [308 bp], the ribosomal internal transcribed spacer<br />

(ITS) [615 bp], glutamine synthetase (GS) [907 bp], manganesesuperoxide<br />

dismutase (SOD2) [376 bp], and β-tubulin 2 (TUB2)<br />

[716 bp].<br />

PCR Primers used during this study are shown in Table 2.<br />

<strong>The</strong> standard CAL primers (O’Donnell et al. 2000) gave poor or<br />

non-specific amplification for most isolates, thus new primers<br />

(CL1C, CL2C) were designed for <strong>Colletotrichum</strong> based on the C.<br />

graminicola M1.001 genome sequence. <strong>The</strong> standard GS primers<br />

(Stephenson et al. 1997) sequenced poorly for some isolates due<br />

to an approx. 9 bp homopolymer T run 71 bp in from the end of<br />

the GSF1 primer binding site. A new primer, GSF3, was designed<br />

41 bp downstream of this region to eliminate the homopolymer<br />

slippage error from sequencing. <strong>The</strong> reverse primer GSR2 was<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

designed in the same location as GSR1 with one nucleotide<br />

change. Both new GS primers were based on similarity with a<br />

C. theobromicola UQ62 sequence (GenBank L78067, as C.<br />

<strong>gloeosporioides</strong>).<br />

<strong>The</strong> PCRs were performed in an Applied Biosystems Veriti<br />

<strong>The</strong>rmal Cycler in a total volume of 25 μL. <strong>The</strong> PCR mixtures<br />

contained 15.8 μL of UV-sterilised ultra-filtered water, 2.5 μL of 10×<br />

PCR buffer (with 20 mM MgCl 2 ), 2.5 μL of dNTPs (each 20 μM), 1<br />

μL of each primer (10 μM), 1 μL of BSA, 1 μL of genomic DNA, and<br />

0.2 μL (1 U) of Roche FastStart Taq DNA Polymerase.<br />

<strong>The</strong> PCR conditions for ITS were 4 min at 95 °C, then 35<br />

cycles of 95 °C for 30 s, 52 °C for 30 s, 72 °C for 45 s, and then<br />

7 min at 72 °C. <strong>The</strong> annealing temperatures differed for the other<br />

genes, with the optimum for each; ACT: 58 °C, CAL: 59 °C, CHS-<br />

1: 58 °C, GAPDH: 60 °C, GS: 54 °C, SOD2: 54 °C, TUB2: 55<br />

°C. Some isolates required altered temperatures and occasionally<br />

gave multiple bands, which were excised separately from an<br />

electrophoresis gel and purified. PCR Products were purified on a<br />

Qiagen MinElute 96 UF PCR Purification Plate.<br />

DNA sequences were obtained in both directions on an Applied<br />

Biosystems 3130xl Avant Genetic analyzer using BigDye v. 3.1<br />

chemistry, electropherograms were analysed and assembled in<br />

Sequencher v. 4.10.1 (Gene Codes Corp.).<br />

Phylogenetic analyses<br />

Multiple sequence alignments of each gene were made with<br />

ClustalX v. 2.1 (Larkin et al. 2007), and manually adjusted where<br />

necessary with Geneious Pro v. 5.5.6 (Drummond et al. 2011).<br />

Bayesian inference (BI) was used to reconstruct most of<br />

the phylogenies using MrBayes v. 3.2.1 (Ronquist et al. 2012).<br />

Bayesian inference has significant advantages over other methods<br />

of analysis such as maximum likelihood and maximum parsimony<br />

(Archibald et al. 2003) and provides measures of clade support as<br />

posterior probabilities rather than random resampling bootstraps.<br />

jModelTest v. 0.1.1 (Posada 2008) was used to carry out statistical<br />

selection of best-fit models of nucleotide substitution using the<br />

corrected Akaike information criteria (AICc) (Table 3). Initial<br />

analyses showed that individual genes were broadly congruent,<br />

thus nucleotide alignments of all genes were concatenated using<br />

Geneious, and separate partitions created for each gene with<br />

their own model of nucleotide substitution. Analyses on the full<br />

data set were run twice for 5 x 10 7 generations, and twice for 2 x<br />

10 7 generations for the clade trees. Samples were taken from the<br />

posterior every 1000 generations. Convergence of all parameters<br />

was checked using the internal diagnostics of the standard<br />

deviation of split frequencies and performance scale reduction<br />

factors (PSRF), and then externally with Tracer v. 1.5 (Rambaut<br />

& Drummond 2007). On this basis the first 25 % of generations<br />

were discarded as burn-in.<br />

An initial BI analysis treated all 158 isolates using a concatenated<br />

alignment for five of the genes, ACT, CAL, CHS-1, GAPDH, and ITS.<br />

<strong>Colletotrichum</strong> boninense and C. hippeastri were used as outgroups.<br />

A second BI analysis, restricted to ex-type or authentic isolates<br />

of each of the accepted <strong>species</strong>, was based on a concatenated<br />

alignment of all eight genes. A third set of BI analyses treated<br />

focussed on taxa within the Musae clade and the Kahawae clade.<br />

For each clade, the ex-type or authentic isolates, together with 2–3<br />

additional selected isolates of each accepted taxon where available,<br />

were analysed using a concatenated alignment of all eight genes,<br />

with C. <strong>gloeosporioides</strong> used as the outgroup for both analyses.<br />

117


Weir et al.<br />

Table 1. A list of strains used in this study.<br />

Species Culture* Host Country GenBank accession number<br />

118<br />

ITS GAPDH CAL ACT CHS-1 GS SOD2 TUB2<br />

C. aenigma ICMP 18608* Persea americana Israel JX010244 JX010044 JX009683 JX009443 JX009774 JX010078 JX010311 JX010389<br />

ICMP 18686 Pyrus pyrifolia Japan JX010243 JX009913 JX009684 JX009519 JX009789 JX010079 JX010312 JX010390<br />

C. aeschynomenes ICMP 17673*, ATCC 201874 Aeschynomene virginica USA JX010176 JX009930 JX009721 JX009483 JX009799 JX010081 JX010314 JX010392<br />

C. alatae <strong>CBS</strong> 304.67*, ICMP 17919 Dioscorea alata India JX010190 JX009990 JX009738 JX009471 JX009837 JX010065 JX010305 JX010383<br />

ICMP 18122 Dioscorea alata Nigeria JX010191 JX010011 JX009739 JX009470 JX009846 JX010136 JX010371 JX010449<br />

C. alienum IMI 313842, ICMP 18691 Persea americana Australia JX010217 JX010018 JX009664 JX009580 JX009754 JX010074 JX010307 JX010385<br />

ICMP 18703 Persea americana New Zealand JX010252 JX010030 JX009656 JX009528 JX009885<br />

ICMP 12071* Malus domestica New Zealand JX010251 JX010028 JX009654 JX009572 JX009882 JX010101 JX010333 JX010411<br />

ICMP 17972 Diospyros kaki New Zealand JX010247 JX009944 JX009655 JX009497 JX009750<br />

ICMP 18704 Persea americana New Zealand JX010253 JX010045 JX009658 JX009456 JX009886<br />

ICMP 18621 Persea americana New Zealand JX010246 JX009959 JX009657 JX009552 JX009755 JX010075 JX010308 JX010386<br />

ICMP 12068 Malus domestica New Zealand JX010255 JX009925 JX009660 JX009492 JX009889<br />

ICMP 18725 Malus domestica New Zealand JX010254 JX009943 JX009659 JX009530 JX009887<br />

C. aotearoa ICMP 18532 Vitex lucens New Zealand JX010220 JX009906 JX009614 JX009544 JX009764 JX010108 JX010338 JX010421<br />

ICMP 18734 Agathis australis New Zealand JX010211 JX010004 JX009627 JX009569 JX009878<br />

ICMP 18528 Berberis glaucocarpa New Zealand JX010199 JX009977 JX009615 JX009527 JX009879<br />

ICMP 17324 Kunzea ericoides New Zealand JX010198 JX009991 JX009619 JX009538 JX009770 JX010109 JX010344 JX010418<br />

ICMP 18533 Prumnopitys ferruginea New Zealand JX010197 JX010026 JX009624 JX009522 JX009769 JX010110 JX010340 JX010416<br />

ICMP 18535 Dacrycarpus dacrydioides New Zealand JX010201 JX009968 JX009617 JX009545 JX009766 JX010107 JX010364 JX010423<br />

ICMP 18577 Coprosma sp. New Zealand JX010203 JX009978 JX009612 JX009567 JX009851 JX010111 JX010360 JX010417<br />

ICMP 18529 Acmena smithii New Zealand JX010222 JX009956 JX009618 JX009539 JX009883<br />

ICMP 18537* Coprosma sp. New Zealand JX010205 JX010005 JX009611 JX009564 JX009853 JX010113 JX010345 JX010420<br />

ICMP 18536 Coprosma sp. New Zealand JX010204 JX009907 JX009610 JX009577 JX009852<br />

ICMP 18748 Ligustrum lucidum New Zealand JX010209 JX009918 JX009613 JX009453 JX009858<br />

ICMP 17326 Podocarpus totara New Zealand JX010202 JX010049 JX009616 JX009578 JX009768 JX010106 JX010341 JX010422<br />

ICMP 18540 Geniostoma ligustrifolium New Zealand JX010207 JX010043 JX009622 JX009514 JX009855<br />

ICMP 18541 Coprosma sp. New Zealand JX010208 JX009960 JX009607 JX009513 JX009856<br />

ICMP 18742 Meryta sinclairii New Zealand JX010210 JX010025 JX009626 JX009477 JX009862<br />

ICMP 18740 Dysoxylum spectabile New Zealand JX010218 JX009988 JX009625 JX009517 JX009763 JX010135 JX010368 JX010446<br />

ICMP 18530 Vitex lucens New Zealand JX010268 JX009911 JX009623 JX009521 JX009884 JX010112 JX010339 JX010419<br />

ICMP 18735 Hedychium gardnerianum New Zealand JX010221 JX010023 JX009620 JX009500 JX009880 JX010115 JX010343 JX010424


Table 1. (Continued).<br />

Species Culture* Host Country GenBank accession number<br />

ITS GAPDH CAL ACT CHS-1 GS SOD2 TUB2<br />

ICMP 18736 Lonicera japonica New Zealand JX010200 JX009912 JX009608 JX009454 JX009894<br />

ICMP 18548 Coprosma sp. New Zealand JX010206 JX009961 JX009609 JX009854 JX009445 JX010114 JX010342 JX010425<br />

ICMP 18543 Melicytus ramiflorus New Zealand JX010156 JX009983 JX009621 JX009524 JX009859<br />

C. asianum IMI 313839, ICMP 18696 Mangifera indica Australia JX010192 JX009915 JX009723 JX009576 JX009753 JX010073 JX010306 JX010384<br />

www.studiesinmycology.org<br />

MAFF 306627, ICMP 18603 Mangifera indica Philippines JX010195 JX009938 JX009725 JX009579 JX009825<br />

HKUCC 10862, ICMP 18605 Mangifera indica Thailand JX010194 JX010021 JX009726 JX009465 JX009787<br />

ICMP 18580*, <strong>CBS</strong> 130418 Coffea arabica Thailand FJ972612 JX010053 FJ917506 JX009584 JX009867 JX010096 JX010328 JX010406<br />

<strong>CBS</strong> 124960, ICMP 18648 Mangifera indica Panama JX010193 JX010017 JX009724 JX009546 JX009871<br />

Japan JX010292 JX009905 JX009583 JX009827<br />

Crinum asiaticum var.<br />

sinicum<br />

C. boninense MAFF 305972*, ICMP 17904,<br />

<strong>CBS</strong> 123755<br />

C. clidemiae ICMP 18706 Vitis sp. USA JX010274 JX009909 JX009639 JX009476 JX009777 JX010128 JX010353 JX010439<br />

ICMP 18658* Clidemia hirta USA, Hawaii JX010265 JX009989 JX009645 JX009537 JX009877 JX010129 JX010356 JX010438<br />

C. cordylinicola MFLUCC 090551*, ICMP 18579 Cordyline fruticosa Thailand JX010226 JX009975 HM470238 HM470235 JX009864 JX010122 JX010361 JX010440<br />

C. fructicola ICMP 12568 Persea americana Australia JX010166 JX009946 JX009680 JX009529 JX009762<br />

ICMP 17787 Malus domestica Brazil JX010164 JX009958 JX009667 JX009439 JX009807<br />

ICMP 17788 Malus domestica Brazil JX010177 JX009949 JX009672 JX009458 JX009808<br />

IMI 345051, ICMP 17819 Fragaria × ananassa Canada JX010180 JX009997 JX009668 JX009469 JX009820<br />

ICMP 18613 Limonium sinuatum Israel JX010167 JX009998 JX009675 JX009491 JX009772 JX010077 JX010310 JX010388<br />

ICMP 18698 Limonium sp. Israel JX010168 JX010052 JX009677 JX009585 JX009773<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

ICMP 18667 Limonium sp. Israel JX010169 JX009951 JX009679 JX009464 JX009775<br />

ICMP 18615 Limonium sp. Israel JX010170 JX010016 JX009678 JX009511 JX009776<br />

ICMP 18610 Pyrus pyrifolia Japan JX010174 JX010034 JX009681 JX009526 JX009788<br />

ICMP 18120 Dioscorea alata Nigeria JX010182 JX010041 JX009670 JX009436 JX009844 JX010091 JX010323 JX010401<br />

<strong>CBS</strong> 125395, ICMP 18645 <strong>The</strong>obroma cacao Panama JX010172 JX009992 JX009666 JX009543 JX009873 JX010098 JX010330 JX010408<br />

ICMP 18581*, <strong>CBS</strong> 130416 Coffea arabica Thailand JX010165 JX010033 FJ917508 FJ907426 JX009866 JX010095 JX010327 JX010405<br />

ICMP 18727 Fragaria × ananassa USA JX010179 JX010035 JX009682 JX009565 JX009812 JX010083 JX010316 JX010394<br />

<strong>CBS</strong> 120005, ICMP 18609 Fragaria × ananassa USA JX010175 JX009926 JX009673 JX009534 JX009792<br />

ICMP 17789 Malus domestica USA JX010178 JX009914 JX009665 JX009451 JX009809<br />

ICMP 18125 Dioscorea alata Nigeria JX010183 JX010009 JX009669 JX009468 JX009847<br />

C. fructicola (syn. C. ignotum) <strong>CBS</strong> 125397(*), ICMP 18646 Tetragastris panamensis Panama JX010173 JX010032 JX009674 JX009581 JX009874 JX010099 JX010331 JX010409<br />

C. fructicola (syn. Glomerella cingulata var. <strong>CBS</strong> 238.49(*), ICMP 17921 Ficus edulis Germany JX010181 JX009923 JX009671 JX009495 JX009839 JX010090 JX010322 JX010400<br />

minor)<br />

119


Weir et al.<br />

Table 1. (Continued).<br />

Species Culture* Host Country GenBank accession number<br />

120<br />

ITS GAPDH CAL ACT CHS-1 GS SOD2 TUB2<br />

C. <strong>gloeosporioides</strong> DAR 76936, ICMP 18738 Carya illinoinensis Australia JX010151 JX009976 JX009730 JX009542 JX009797<br />

IMI 356878*, ICMP 17821,<br />

Citrus sinensis Italy JX010152 JX010056 JX009731 JX009531 JX009818 JX010085 JX010365 JX010445<br />

<strong>CBS</strong> 112999<br />

ICMP 12939 Citrus sp. New Zealand JX010149 JX009931 JX009728 JX009462 JX009747<br />

ICMP 12066 Ficus sp. New Zealand JX010158 JX009955 JX009734 JX009550 JX009888<br />

ICMP 18730 Citrus sp. New Zealand JX010157 JX009981 JX009737 JX009548 JX009861<br />

ICMP 12938 Citrus sinensis New Zealand JX010147 JX009935 JX009732 JX009560 JX009746<br />

ICMP 18694 Mangifera indica South Africa JX010155 JX009980 JX009729 JX009481 JX009796<br />

<strong>CBS</strong> 119204, ICMP 18678 Pueraria lobata USA JX010150 JX010013 JX009733 JX009502 JX009790<br />

ICMP 18695 Citrus sp. USA JX010153 JX009979 JX009735 JX009494 JX009779<br />

ICMP 18697 Vitis vinifera USA JX010154 JX009987 JX009736 JX009557 JX009780<br />

C. <strong>gloeosporioides</strong> (syn. Gloeosporium <strong>CBS</strong> 273.51(*), ICMP 19121 Citrus limon Italy JX010148 JX010054 JX009745 JX009558 JX009903<br />

pedemontanum)<br />

C. hippeastri <strong>CBS</strong> 241.78, ICMP 17920 Hippeastrum sp. Netherlands JX010293 JX009932 JX009740 JX009485 JX009838<br />

C. horii ICMP 12942 Diospyros kaki New Zealand GQ329687 GQ329685 JX009603 JX009533 JX009748 JX010072 JX010296 JX010375<br />

ICMP 12951 Diospyros kaki New Zealand GQ329689 GQ329683 JX009602 JX009466 JX009751<br />

NBRC 7478*, ICMP 10492 Diospyros kaki Japan GQ329690 GQ329681 JX009604 JX009438 JX009752 JX010137 JX010370 JX010450<br />

ICMP 17968 Diospyros kaki China JX010212 GQ329682 JX009605 JX009547 JX009811 JX010068 JX010300 JX010378<br />

MAFF 306429, ICMP 17970 Diospyros kaki Japan JX010213 GQ329686 JX009606 JX009467 JX009824<br />

C. kahawae subsp. ciggaro ICMP 18539* Olea europaea Australia JX010230 JX009966 JX009635 JX009523 JX009800 JX010132 JX010346 JX010434<br />

ICMP 18728 Miconia sp. Brazil JX010239 JX010048 JX009643 JX009525 JX009850<br />

ICMP 18741 Kunzea ericoides New Zealand JX010229 JX010039 JX009631 JX009472 JX009767<br />

ICMP 18534 Kunzea ericoides New Zealand JX010227 JX009904 JX009634 JX009473 JX009765 JX010116 JX010351 JX010427<br />

ICMP 18544 Toronia toru New Zealand JX010240 JX009920 JX009632 JX009430 JX009860<br />

ICMP 18531 Persea americana New Zealand JX009463 JX009999 JX009647 JX009463 JX009749<br />

ICMP 12952 Persea americana New Zealand JX010214 JX009971 JX009648 JX009431 JX009757 JX010126 JX010348 JX010426<br />

ICMP 12953 Persea americana New Zealand JX010215 JX009928 JX009646 JX009499 JX009758<br />

<strong>CBS</strong> 112984, ICMP 17932 Dryandra sp. South Africa JX010237 JX009973 JX009633 JX009434 JX009833<br />

IMI 359911, ICMP 17931,<br />

Dryas octopetala Switzerland JX010236 JX009965 JX009637 JX009475 JX009832 JX010121 JX010354 JX010428<br />

<strong>CBS</strong> 12988<br />

<strong>CBS</strong> 237.49(*), ICMP 17922 Hypericum perforatum Germany JX010238 JX010042 JX009636 JX009450 JX009840 JX010120 JX010355 JX010432<br />

C. kahawae subsp. ciggaro (syn. Glomerella<br />

cingulata var. migrans)


Table 1. (Continued).<br />

Species Culture* Host Country GenBank accession number<br />

ITS GAPDH CAL ACT CHS-1 GS SOD2 TUB2<br />

C. kahawae subsp. ciggaro (syn. Glomerella <strong>CBS</strong> 124.22(*), ICMP 19122 Vaccinium sp. USA JX010228 JX009950 JX009744 JX009536 JX009902 JX010134 JX010367 JX010433<br />

rufomaculans var. vaccinii)<br />

C. kahawae subsp. kahawae <strong>CBS</strong> 982.69, ICMP 17915 Coffea arabica Angola JX010234 JX010040 JX009638 JX009474 JX009829 JX010125 JX010352 JX010435<br />

IMI 361501, ICMP 17905 Coffea arabica Cameroon JX010232 JX010046 JX009644 JX009561 JX009816 JX010127 JX010349 JX010431<br />

www.studiesinmycology.org<br />

IMI 319418*, ICMP 17816 Coffea arabica Kenya JX010231 JX010012 JX009642 JX009452 JX009813 JX010130 JX010350 JX010444<br />

<strong>CBS</strong> 135.30, ICMP 17928 Coffea sp. Kenya JX010235 JX010037 JX009640 JX009554 JX009831<br />

IMI 301220, ICMP 17811 Coffea arabica Malawi JX010233 JX009970 JX009641 JX009555 JX009817 JX010131 JX010347 JX010430<br />

C. musae <strong>CBS</strong> 192.31, ICMP 17923 Musa sp. Indonesia JX010143 JX009929 JX009690 JX009587 JX009841<br />

IMI 52264, ICMP 17817 Musa sapientum Kenya JX010142 JX010015 JX009689 JX009432 JX009815 JX010084 JX010317 JX010395<br />

ICMP 12931 Musa sp. New Zealand JX010140 JX009995 JX009688 JX009442 JX009756<br />

(imported)<br />

ICMP 18600 Musa sp. Philippines JX010144 JX010038 JX009686 JX009556 JX009848<br />

ICMP 12930 Musa sp. New Zealand JX010141 JX009986 JX009685 JX009566 JX009881<br />

ICMP 18701 Musa sp. Philippines JX010145 JX010047 JX009687 JX009551 JX009849<br />

<strong>CBS</strong> 116870*, ICMP 19119 Musa sp. USA JX010146 JX010050 JX009742 JX009433 JX009896 JX010103 JX010335 HQ596280<br />

USA JX010189 JX009936 JX009661 JX009486 JX009834 JX010087 JX010319 JX010397<br />

C. nupharicola <strong>CBS</strong> 469.96, ICMP 17938 Nuphar lutea subsp.<br />

polysepala<br />

USA JX010187 JX009972 JX009663 JX009437 JX009835 JX010088 JX010320 JX010398<br />

<strong>CBS</strong> 470.96*, ICMP 18187 Nuphar lutea subsp.<br />

polysepala<br />

<strong>CBS</strong> 472.96, ICMP 17940 Nymphaea ordorata USA JX010188 JX010031 JX009662 JX009582 JX009836 JX010089 JX010321 JX010399<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

C. psidii <strong>CBS</strong> 145.29*, ICMP 19120 Psidium sp. Italy JX010219 JX009967 JX009743 JX009515 JX009901 JX010133 JX010366 JX010443<br />

C. queenslandicum ICMP 1778* Carica papaya Australia JX010276 JX009934 JX009691 JX009447 JX009899 JX010104 JX010336 JX010414<br />

ICMP 1780 Carica sp. Australia JX010186 JX010010 JX009693 JX009504 JX009900<br />

ICMP 12564 Persea americana Australia JX010184 JX009919 JX009692 JX009573 JX009759<br />

ICMP 18705 Coffea sp. Fiji JX010185 JX010036 JX009694 JX009490 JX009890 JX010102 JX010334 JX010412<br />

C. salsolae ICMP 19051* Salsola tragus Hungary JX010242 JX009916 JX009696 JX009562 JX009863 JX010093 JX010325 JX010403<br />

<strong>CBS</strong> 119296, ICMP 18693 Glycine max (inoculated) Hungary JX010241 JX009917 JX009695 JX009559 JX009791<br />

C. siamense ICMP 12567 Persea americana Australia JX010250 JX009940 JX009697 JX009541 JX009761 JX010076 JX010309 JX010387<br />

DAR 76934, ICMP 18574 Pistacia vera Australia JX010270 JX010002 JX009707 JX009535 JX009798 JX010080 JX010313 JX010391<br />

ICMP 12565 Persea americana Australia JX010249 JX009937 JX009698 JX009571 JX009760<br />

<strong>CBS</strong> 125379, ICMP 18643 Hymenocallis americana China JX010258 JX010060 GQ849451 GQ856776 GQ856729<br />

ICMP 18121 Dioscorea rotundata Nigeria JX010245 JX009942 JX009715 JX009460 JX009845 JX010092 JX010324 JX010402<br />

ICMP 18118 Commelina sp. Nigeria JX010163 JX009941 JX009701 JX009505 JX009843<br />

121


Weir et al.<br />

Table 1. (Continued).<br />

Species Culture* Host Country GenBank accession number<br />

122<br />

ITS GAPDH CAL ACT CHS-1 GS SOD2 TUB2<br />

ICMP 18117 Dioscorea rotundata Nigeria JX010266 JX009954 JX009700 JX009574 JX009842<br />

ICMP 18739 Carica papaya South Africa JX010161 JX009921 JX009716 JX009484 JX009794<br />

ICMP 18570 Persea americana South Africa JX010248 JX009969 JX009699 JX009510 JX009793<br />

ICMP 18569 Persea americana South Africa JX010262 JX009963 JX009711 JX009459 JX009795<br />

ICMP 18578*, <strong>CBS</strong> 130417 Coffea arabica Thailand JX010171 JX009924 FJ917505 FJ907423 JX009865 JX010094 JX010326 JX010404<br />

HKUCC 10884, ICMP 18575 Capsicum annuum Thailand JX010256 JX010059 JX009717 JX009455 JX009785<br />

HKUCC 10881, ICMP 18618 Capsicum annuum Thailand JX010257 JX009945 JX009718 JX009512 JX009786<br />

ICMP 18572 Vitis vinifera USA JX010160 JX010061 JX009705 JX009487 JX009783<br />

ICMP 18571 Fragaria × ananassa USA JX010159 JX009922 JX009710 JX009482 JX009782<br />

ICMP 18573 Vitis vinifera USA JX010271 JX009996 JX009712 JX009435 JX009784<br />

ICMP 17795 Malus domestica USA JX010162 JX010051 JX009703 JX009506 JX009805 JX010082 JX010315 JX010393<br />

ICMP 17791 Malus domestica USA JX010273 JX009933 JX009708 JX009508 JX009810<br />

ICMP 17797 Malus domestica USA JX010263 JX009984 JX009704 JX009461 JX009806<br />

ICMP 17785 Malus domestica USA JX010272 JX010051 JX009706 JX009446 JX009804<br />

C. siamense (syn. C. hymenocallidis) <strong>CBS</strong> 125378(*), ICMP 18642 Hymenocallis americana China JX010278 JX010019 JX009709 GQ856775 GQ856730 JX010100 JX010332 JX010410<br />

C. siamense (syn. C. jasmini-sambac) <strong>CBS</strong> 130420(*), ICMP 19118 Jasminum sambac Vietnam HM131511 HM131497 JX009713 HM131507 JX009895 JX010105 JX010337 JX010415<br />

Stylosanthes guianensis Australia JX010291 JX009948 JX009598 JX009498 JX009822 JX010067 JX010303 JX010381<br />

C. theobromicola MUCL 42295, ICMP 17958,<br />

<strong>CBS</strong> 124250<br />

ICMP 18566 Olea europaea Australia JX010282 JX009953 JX009593 JX009496 JX009801 JX010071 JX010297 JX010376<br />

ICMP 18565 Olea europaea Australia JX010283 JX010029 JX009594 JX009449 JX009802 JX010070 JX010298 JX010374<br />

ICMP 18567 Olea europaea Australia JX010287 JX009985 JX009599 JX009457 JX009803 JX010069 JX010299 JX010377<br />

ICMP 18576 Limonium sp. Israel JX010279 JX010022 JX009595 JX009532 JX009771<br />

ICMP 17895 Annona diversifolia Mexico JX010284 JX010057 JX009600 JX009568 JX009828 JX010066 JX010304 JX010382<br />

ICMP 15445 Acca sellowiana New Zealand JX010290 JX010027 JX009601 JX009509 JX009893<br />

<strong>CBS</strong> 125393, ICMP 18650 <strong>The</strong>obroma cacao Panama JX010280 JX009982 JX009590 JX009503 JX009872<br />

<strong>CBS</strong> 124945*, ICMP 18649 <strong>The</strong>obroma cacao Panama JX010294 JX010006 JX009591 JX009444 JX009869 JX010139 JX010372 JX010447<br />

ICMP 17099 Fragaria × ananassa USA JX010285 JX009957 JX009588 JX009493 JX009778<br />

ICMP 17100 Quercus sp. USA JX010281 JX009947 JX009596 JX009507 JX009781<br />

IMI 348152, ICMP 17814 Fragaria vesca USA JX010288 JX010003 JX009589 JX009448 JX009819 JX010062 JX010301 JX010379<br />

C. theobromicola (syn. C. fragariae) <strong>CBS</strong> 142.31(*), ICMP 17927 Fragaria × ananassa USA JX010286 JX010024 JX009592 JX009516 JX009830 JX010064 JX010295 JX010373<br />

C. theobromicola (syn. C. <strong>gloeosporioides</strong> f. MUCL 42294(*), ICMP 17957, Stylosanthes viscosa Australia JX010289 JX009962 JX009597 JX009575 JX009821 JX010063 JX010302 JX010380<br />

stylosanthis)<br />

<strong>CBS</strong> 124251


Table 1. (Continued).<br />

Species Culture* Host Country GenBank accession number<br />

ITS GAPDH CAL ACT CHS-1 GS SOD2 TUB2<br />

C. ti ICMP 5285 Cordyline australis New Zealand JX010267 JX009910 JX009650 JX009553 JX009897 JX010124 JX010363 JX010441<br />

ICMP 4832* Cordyline sp. New Zealand JX010269 JX009952 JX009649 JX009520 JX009898 JX010123 JX010362 JX010442<br />

C. tropicale MAFF 239933, ICMP 18672 Litchi chinensis Japan JX010275 JX010020 JX009722 JX009480 JX009826 JX010086 JX010318 JX010396<br />

<strong>CBS</strong> 124949*, ICMP 18653 <strong>The</strong>obroma cacao Panama JX010264 JX010007 JX009719 JX009489 JX009870 JX010097 JX010329 JX010407<br />

www.studiesinmycology.org<br />

<strong>CBS</strong> 124943, ICMP 18651 Annona muricata Panama JX010277 JX010014 JX009720 JX009570 JX009868<br />

Xanthorrhoea preissii Australia JX010261 JX009927 JX009653 JX009478 JX009823 JX010138 JX010369 JX010448<br />

C. xanthorrhoeae BRIP 45094*, ICMP 17903,<br />

<strong>CBS</strong> 127831<br />

IMI 350817a, ICMP 17820 Xanthorrhoea sp. Australia JX010260 JX010008 JX009652 JX009479 JX009814<br />

Glomerella cingulata “f.sp. camelliae” ICMP 10643 Camellia × williamsii UK JX010224 JX009908 JX009630 JX009540 JX009891 JX010119 JX010358 JX010436<br />

ICMP 18542 Camellia sasanqua USA JX010223 JX009994 JX009628 JX009488 JX009857 JX010118 JX010357 JX010429<br />

ICMP 10646 Camellia sasanqua USA JX010225 JX009993 JX009629 JX009563 JX009892 JX010117 JX010359 JX010437<br />

* = ex-type or authentic culture, (*) = ex-type or authentic culture of synonymised taxon. Sequences downloaded from GenBank, not generated as part of this project are in bold font. Collection abbreviations are listed in the methods.<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Several <strong>species</strong>-trees analyses were conducted using BEAST<br />

v. 1.7.1 (Drummond et al. 2012). Species-trees combine multi-gene<br />

and multiple isolate data to reconstruct the evolutionary history<br />

of hypothesised <strong>species</strong>, rather than individual isolates. BEAST<br />

does not use concatenation, but rather co-estimates the individual<br />

gene trees embedded inside the summary <strong>species</strong> tree. It also<br />

estimates the time since each <strong>species</strong> shared a common ancestor<br />

(divergence times). For these analyses the <strong>species</strong> tree ancestral<br />

reconstruction option was selected (Heled & Drummond 2010), the<br />

gene data partitioned as for BI and the substitution model for each<br />

gene was selected based on the models selected using jModelTest.<br />

<strong>The</strong> individual isolates were grouped into sets of <strong>species</strong> by setting<br />

<strong>species</strong> names as trait values. A strict clock was used for the<br />

GAPDH gene (as an all intronic sequence it was assumed to be<br />

accumulating mutations at a steady rate) and the other gene clock<br />

rates were estimated relative to GAPDH, using an uncorrelated<br />

lognormal relaxed clock. <strong>The</strong> <strong>species</strong> tree prior used for all genes<br />

was the Yule process, with the ploidy type set to nuclear autosomal.<br />

Uninformative priors were used for all parameters, and were<br />

allowed to auto optimise.<br />

<strong>The</strong> first <strong>species</strong>-tree analysis was conducted using the 158<br />

isolate, five gene dataset, with C. boninense and C. hippeastri as<br />

the outgroups. <strong>The</strong> MCMC chain was set to 1 × 10 8 generations<br />

for the <strong>species</strong> <strong>complex</strong> tree and samples were taken from the<br />

posterior every 1000 generations. <strong>The</strong> analysis was run twice<br />

independently. <strong>The</strong> effective sample size (ESS) and traces of<br />

all parameters and convergence of the two runs was checked<br />

with Tracer and a summary maximum clade credibility <strong>species</strong><br />

tree was built with TreeAnnotator v. 1.7.1 (Drummond et al.<br />

2012) using a 10 % burn-in and a posterior probability limit of<br />

0.5, setting the heights of each node in the tree to the mean<br />

height across the entire sample of trees for that clade. Separate<br />

analyses were conducted using all eight genes and the same<br />

restricted set of isolates chosen to represent taxa within the<br />

Musae clade and the Kahawae clade as were used for the BI<br />

analyses of the eight gene concatenated analyses outlined<br />

above. For each of the Musae and Kahawae clade analyses,<br />

the MCMC chain was set to 5 × 10 7 generations, but otherwise<br />

run as for the five gene dataset.<br />

To illustrate the potential limitations of ITS to discriminate <strong>species</strong><br />

within the C. <strong>gloeosporioides</strong> <strong>complex</strong>, an UPGMA tree was built<br />

of all 158 ITS sequences, using the Geneious tree builder tool. A<br />

UPGMA tree visually approximates a BLAST search, which is based<br />

on distances (and sequence length) rather than corrected nucleotide<br />

substitutions of more sophisticated, model-based analyses.<br />

Sequences derived in this study were lodged in GenBank (Table<br />

1), the concatenated alignment and trees in TreeBASE (www.<br />

treebase.org) study number S12535, and taxonomic novelties in<br />

MycoBank (Crous et al. 2004).<br />

Morphology<br />

Detailed morphological descriptions are provided only for those<br />

<strong>species</strong> with no recently published description. Few specimens<br />

were examined from infected host material; the descriptions<br />

provided are mostly from agar cultures. Cultures were grown on<br />

Difco PDA from single conidia, or from single hyphal tips for the few<br />

specimens where no conidia were formed, with culture diameter<br />

measured and appearance described after 10 d growth at 18–20 o C<br />

under mixed white and UV fluorescent tubes, 12 h light/12 h dark.<br />

Colour codes follow Kornerup & Wanscher (1963).<br />

123


Weir et al.<br />

Table 2. Primers used in this study, with sequences and sources.<br />

Gene Product name Primer Direction Sequence (5’–3’) Reference<br />

ACT Actin ACT-512F Foward ATG TGC AAG GCC GGT TTC GC Carbone & Kohn 1999<br />

ACT-783R Reverse TAC GAG TCC TTC TGG CCC AT Carbone & Kohn 1999<br />

CAL Calmodulin CL1 Foward GAR TWC AAG GAG GCC TTC TC O’Donnell et al. 2000<br />

CL2A Reverse TTT TTG CAT CAT GAG TTG GAC O’Donnell et al. 2000<br />

CL1C Foward GAA TTC AAG GAG GCC TTC TC This study<br />

CL2C Reverse CTT CTG CAT CAT GAG CTG GAC This study<br />

CHS-1 Chitin synthase CHS-79F Foward TGG GGC AAG GAT GCT TGG AAG AAG Carbone & Kohn 1999<br />

CHS-345R Reverse TGG AAG AAC CAT CTG TGA GAG TTG Carbone & Kohn 1999<br />

GAPDH Glyceraldehyde-3-phosphate dehydrogenase GDF Foward GCC GTC AAC GAC CCC TTC ATT GA Templeton et al. 1992<br />

GDR Reverse GGG TGG AGT CGT ACT TGA GCA TGT Templeton et al. 1992<br />

GS Glutamine synthetase GSF1 Foward ATG GCC GAG TAC ATC TGG Stephenson et al. 1997<br />

GSF3 Foward GCC GGT GGA GGA ACC GTC G This study<br />

GSR1 Reverse GAA CCG TCG AAG TTC CAG Stephenson et al. 1997<br />

GSR2 Reverse GAA CCG TCG AAG TTC CAC This study<br />

ITS Internal transcribed spacer ITS-1F Foward CTT GGT CAT TTA GAG GAA GTA A Gardes & Bruns 1993<br />

ITS-4 Reverse TCC TCC GCT TAT TGA TAT GC White et al. 1990<br />

SOD2 Manganese-superoxide dismutase SODglo2-F Foward CAG ATC ATG GAG CTG CAC CA Moriwaki & Tsukiboshi 2009<br />

SODglo2-R Reverse TAG TAC GCG TGC TCG GAC AT Moriwaki & Tsukiboshi 2009<br />

TUB2 β-Tubulin 2 T1 Foward AAC ATG CGT GAG ATT GTA AGT O’Donnell & Cigelnik 1997<br />

T2 Reverse TAG TGA CCC TTG GCC CAGT TG O’Donnell & Cigelnik 1997<br />

Bt2b Reverse ACC CTC AGT GTA GTG ACC CTT GGC Glass & Donaldson 1995<br />

Table 3. Nucleotide substitution models used in phylogenetic<br />

analyses.<br />

Gene All taxa Musae clade Kahawae clade<br />

ITS TrNef+G TrNef+G TrNef<br />

GAPDH HKY+G TPM1uf+G TrN<br />

CAL TIM1+G TIM1+G TrN+G<br />

ACT HKY+G TrN JC<br />

CHS-1 TrNef+G TrNef+G K80<br />

GS TIM2+G TIM3+G<br />

SOD2 HKY+G GTR+I+G<br />

TUB2 TrN+G HKY+G<br />

Conidia were measured and described using conidia taken from<br />

the conidial ooze on acervuli and mounted in lactic acid, at least<br />

24 conidia were measured for each isolate, range measurements<br />

are provided in the form (lower extreme–) 25 % quartile – 75 %<br />

quartile (–upper extreme), all ranges were rounded to the nearest<br />

0.5 µm. Cultures were examined periodically for the development<br />

of perithecia. Ascospores were measured and described from<br />

perithecia crushed in lactic acid.<br />

Appressoria were producing using a slide culture technique. A<br />

small square of agar was inoculated on one side with conidia and<br />

immediately covered with a sterile cover slip. After 14 d the cover<br />

slip was removed and placed in a drop of lactic acid on a glass<br />

slide.<br />

All morphological character measurements were analysed<br />

with the statistical programme “R” v. 2.14.0 (R Development Core<br />

Team 2011). <strong>The</strong> R package ggplot2 (Wickham 2009) was used for<br />

graphical plots. <strong>The</strong> box plots show the median, upper and lower<br />

124<br />

quartiles, and the ‘whisker’ extends to the outlying data, or to a<br />

maximum of 1.5× the interquartile range, individual outliers outside<br />

this range are shown as dots.<br />

Taxa treated in the taxonomic section<br />

Species, subspecific taxa, and formae speciales within the C.<br />

<strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> are treated alphabetically by<br />

epithet. <strong>The</strong> names of formae speciales are not governed by the<br />

International Code of Botanical Nomenclature (ICBN) (McNeill<br />

et al. 2006, Art. 4, Note 4), and are hence enclosed in quotation<br />

marks to indicate their invalid status. Other invalid names that<br />

are governed by the ICBN are also enclosed in quotation marks.<br />

Accepted names are marked with an asterisk (*). <strong>The</strong> breadth of<br />

the taxonomic names treated includes:<br />

all taxonomic names with DNA sequence data in GenBank<br />

that place them in the C. <strong>gloeosporioides</strong> <strong>complex</strong> as it has been<br />

defined here on the basis of the ITS gene tree. <strong>The</strong> sense that the<br />

names were used in GenBank may have been misapplied;<br />

names that have been used in the literature in recent years<br />

for which a possible relationship to C. <strong>gloeosporioides</strong> has been<br />

suggested;<br />

all subspecific taxa and formae speciales within C.<br />

<strong>gloeosporioides</strong> and Glomerella cingulata.<br />

We have not considered the full set of <strong>species</strong> in <strong>Colletotrichum</strong>,<br />

Gloeosporium and Glomerella that were placed in synonymy with<br />

C. <strong>gloeosporioides</strong> or Glomerella cingulata by von Arx & Müller<br />

(1954) or von Arx (1957, 1970).<br />

For each accepted <strong>species</strong>, comments are provided regarding<br />

the limitations of ITS, the official barcoding gene for fungi, to<br />

distinguish that <strong>species</strong> from others within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong>.


RESULTS<br />

Phylogenetics<br />

DNA sequences of five genes were obtained from all 158 isolates<br />

included in the study and concatenated to form a supermatrix of<br />

2294 bp. <strong>The</strong> gene boundaries in the alignment were: ACT: 1–316,<br />

CAL: 317–1072, CHS-1: 1073–1371, GAPDH: 1372–1679, ITS:<br />

1680–2294. A BI analysis of the concatenated dataset is presented<br />

in Fig.1. This tree is annotated with the <strong>species</strong> boundaries of the<br />

taxa that we accept in the C. <strong>gloeosporioides</strong> <strong>complex</strong>, and the<br />

clades representing these taxa formed the basis for investigating<br />

the morphological and biological diversity of our <strong>species</strong>. Ex-type<br />

and authentic isolates are highlighted in bold and labelled with the<br />

names under which they were originally described. <strong>The</strong> posterior<br />

probability (PP) support for the grouping of most <strong>species</strong> ranges<br />

from 1 to 0.96, however support for deeper nodes is often lower,<br />

e.g. 0.53 for the root of C. ti and C. aotearoa, indicating that the<br />

branching may be uncertain for the root of these <strong>species</strong>. Branch<br />

lengths and node PP are typically lower within a <strong>species</strong> than<br />

between <strong>species</strong>.<br />

<strong>The</strong> large number of taxa in Fig. 1 makes it difficult to visualise<br />

the interspecific genetic distance between the recognised <strong>species</strong>.<br />

<strong>The</strong> unrooted tree in Fig. 2 represents the results of a BI analysis<br />

based on a concatenation of all eight genes, but restricted to the<br />

ex-type or authentic cultures from each of the accepted taxa. <strong>The</strong><br />

analysis was done without out-group taxa and clearly shows two<br />

clusters of closely related <strong>species</strong> that we informally label the<br />

Musae clade, and the Kahawae clade.<br />

To better resolve relationships within the Musae and Kahawae<br />

clades a further set of BI analyses included eight genes and,<br />

wherever possible, several representative isolates of each of the<br />

accepted <strong>species</strong>. All eight gene sequences were concatenated<br />

to form a supermatrix for each clade. For the Musae clade of<br />

32 isolates the alignment was 4199 bp and the gene boundaries<br />

were: ACT: 1–292, TUB2: 293–1008, CAL: 1009–1746, CHS-1:<br />

1747–2045, GAPDH: 2046–2331, GS: 2332–3238, ITS: 3239–<br />

3823, SOD2: 3824–4199. For the Kahawae clade of 30 isolates<br />

the alignment was 4107 bp and the gene boundaries were: ACT:<br />

1–281, TUB2: 282–988, CAL: 989–1728, CHS-1: 1729–2027,<br />

GAPDH: 2028–2311, GS: 2312–3179, ITS: 3198–3733, SOD2:<br />

3734–4107. <strong>The</strong> additional genes sequenced provided additional<br />

support for our initial <strong>species</strong> delimitations with better resolution<br />

for some closely related <strong>species</strong>. For example, the highly<br />

pathogenic coffee berry isolates (referred to here as C. kahawae<br />

subsp. kahawae) were distinguished from other C. kahawae<br />

isolates.<br />

Analyses based on concatenated data sets can mask<br />

incongruence between individual gene trees. <strong>The</strong> low levels of<br />

support within some parts of the <strong>species</strong>-tree analysis (Fig. 3),<br />

in part reflects incongruence between gene trees. <strong>The</strong> levels of<br />

support for the Kahawae clade and for the Musae clade are strong<br />

(PP=1) but the <strong>species</strong> that we accept within these clades have<br />

lower levels of support than is shown between the other <strong>species</strong><br />

outside of the clades. <strong>The</strong> scale bar in Fig. 3 represents a time scale,<br />

calibrated at zero for the present day, and at 1 for the last common<br />

ancestor (LCA) of the C. <strong>gloeosporioides</strong> and C. boninense <strong>species</strong><br />

<strong>complex</strong>es. <strong>The</strong> separate <strong>species</strong>-tree analyses for the Musae and<br />

Kahawae clades provide a finer resolution of evolutionary history<br />

within each clade, the time scale based on the same calibration<br />

as Fig. 3.<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

<strong>The</strong> UPGMA-based ITS gene tree (Fig 6). shows that C.<br />

theobromicola, C. horii, C. <strong>gloeosporioides</strong>, G. cingulata “f. sp.<br />

camelliae”, C. asianum, C. musae, C. alatae, C. xanthorrhoeae all<br />

form monophyletic clades and may be distinguished with ITS, but<br />

many <strong>species</strong> are unable to be discriminated using this gene alone.<br />

Note that C. cordylinicola and C. psidii are represented by a single<br />

isolate, meaning that variation within ITS sequences across these<br />

<strong>species</strong> has not been tested.<br />

Morphology and biology<br />

Brief morphological descriptions, based on all specimens examined,<br />

are provided for only those <strong>species</strong> with no recently published<br />

description. Conidial sizes for all accepted <strong>species</strong> are summarised<br />

in Fig. 7. Within a <strong>species</strong>, conidial sizes are reasonably consistent<br />

across isolates, independent of geographic origin or host. However,<br />

differences between <strong>species</strong> are often slight and size ranges often<br />

overlap (Fig. 7). <strong>The</strong> shape of appressoria is generally consistent<br />

within a <strong>species</strong>, some being simple in outline, others <strong>complex</strong> and<br />

highly lobed.<br />

Several <strong>species</strong> are characterised in part by the development<br />

of perithecia in culture. <strong>The</strong>se include four <strong>species</strong> in the Musae<br />

clade (C. alienum, C. fructicola, C. queenslandicum, and C.<br />

salsolae) and three in the Kahawae clade (C. clidemiae, C.<br />

kahawae subsp. ciggaro, and C. ti). Ascospore size and shape<br />

can be a useful <strong>species</strong>-level diagnostic feature (Fig. 8). In most<br />

<strong>species</strong> the ascospores are strongly curved and typically tapering<br />

towards the ends, but in C. clidemiae and C. ti, they are more or<br />

less elliptic with broadly rounded ends and not, or only slightly,<br />

curved. Individual isolates within any of these <strong>species</strong> may lose the<br />

ability to form perithecia, perhaps associated with cultural changes<br />

during storage.<br />

Large, dark-walled stromatic structures are present in the<br />

cultures of some <strong>species</strong> not known to form perithecia. Often<br />

embedded in agar, less commonly on the surface or amongst<br />

the aerial mycelium, these structures differ from perithecia in<br />

comprising a compact tissue of tightly tangled hyphae rather than<br />

the pseudoparenchymatous, angular cells typical of perithecial<br />

walls. <strong>The</strong>y have a soft, leathery texture compared to the more<br />

brittle perithecia. <strong>The</strong>se stromatic structures sometimes develop<br />

a conidiogenous layer internally, and following the production<br />

of conidia they may split open irregularly, folding back to form a<br />

stromatic, acervulus-like structure. <strong>The</strong>se kind of structures are<br />

also formed by some <strong>species</strong> in the C. boninense <strong>species</strong> <strong>complex</strong><br />

(Damm et al. 2012b, this issue).<br />

<strong>The</strong> macroscopic appearance of the cultures is often highly<br />

divergent within a <strong>species</strong> (e.g. C. fructicola and C. theobromicola),<br />

in most cases probably reflecting different storage histories of the<br />

isolates examined. Prolonged storage, especially with repeated<br />

subculturing, results in staling of the cultures, the aerial mycelium<br />

often becoming dense and uniform in appearance and colour, and<br />

a loss of conidial and perithecial production, and variable in growth<br />

rate (Fig. 9). In some <strong>species</strong>, individual single ascospore or single<br />

conidial isolates show two markedly different cultural types, see<br />

notes under C. kahawae subsp. ciggaro.<br />

Some <strong>species</strong> appear to be host specialised, e.g. C. horii,<br />

C. kahawae subsp. kahawae, C. nupharicola, C. salsolae, C. ti,<br />

and C. xanthorrhoeae, but those most commonly isolated have<br />

broad host and geographic ranges, e.g. C. fructicola, C. kahawae<br />

subsp. ciggaro, C. siamense, and C. theobromicola. <strong>Colletotrichum</strong><br />

<strong>gloeosporioides</strong> s. str. is commonly isolated from Citrus in many<br />

125


Weir et al.<br />

126<br />

1<br />

1<br />

ICMP 17789 Malus USA<br />

ICMP 17819 Fragaria Canada<br />

ICMP 18609 Fragaria USA<br />

ICMP 18120 Dioscorea Nigeria<br />

1<br />

ICMP 17787 Malus Brazil<br />

ICMP 17788 Malus Brazil<br />

ICMP 18125 Dioscorea Nigeria<br />

ICMP 17921 Glomerella cingulata var. minor Ficus Germany<br />

ICMP 18645 <strong>The</strong>obroma Panama<br />

0.96 ICMP 18667 Limonium Israel<br />

ICMP 18613 Limonium Israel<br />

0.98<br />

ICMP 18615 Limonium Israel<br />

0.97 ICMP 18610 Pyrus Japan<br />

ICMP 18698 Limonium Israel<br />

1 1 ICMP 18727 Fragaria USA<br />

ICMP 18581 C. fructicola Coffea Thailand<br />

0.99 ICMP 12568 Persea Australia<br />

ICMP 18646 C. ignotum Tetragastris Panama<br />

1 ICMP 18187 C. nupharicola Nuphar USA<br />

1<br />

1 ICMP 17940 Nuphar USA<br />

ICMP 17938 Nuphar USA<br />

0.92 ICMP 12071 C. alienum Malus New Zealand<br />

ICMP 18691 Persea Australia<br />

ICMP 18704 Persea New Zealand<br />

ICMP 12068 Malus New Zealand<br />

1<br />

ICMP 17972 Diospyros New Zealand<br />

ICMP 18703 Persea New Zealand<br />

ICMP 18725 Malus New Zealand<br />

ICMP 18621 Persea New Zealand<br />

1 0.99<br />

1<br />

ICMP 17817 Musa Kenya<br />

ICMP 12931 Musa New Zealand (imported)<br />

ICMP 17923 Musa Indonesia<br />

0.95<br />

ICMP 18600 Musa Philippines<br />

1<br />

ICMP 12930 Musa New Zealand<br />

98<br />

ICMP 19119 C. musae Musa USA<br />

ICMP 18701 Musa Philippines<br />

1 ICMP 18608 C. aenigma Persea Israel<br />

ICMP 18686 Pyrus Japan<br />

1 ICMP 18739 Carica South Africa<br />

0.92 ICMP 18572 Vitis USA<br />

0.51 ICMP 18571 Fragaria USA<br />

1 ICMP 18118 Commelina Nigeria<br />

0.74 ICMP 18117 Dioscorea Nigeria<br />

0.68<br />

ICMP 18578 C. siamense Coffea Thailand<br />

0.98<br />

ICMP 17795 Malus USA<br />

1 ICMP 18570 Persea South Africa<br />

1 ICMP 12567 Persea Australia<br />

ICMP 12565 Persea Australia<br />

0.68<br />

0.53<br />

1 ICMP 18573 Vitis USA<br />

ICMP 17785 Malus USA<br />

ICMP 18643 Hymenocallis China<br />

ICMP 18642 C. hymenocallidis Hymenocallis China<br />

1 ICMP 17797 Malus USA<br />

ICMP 17791 Malus USA<br />

0.83<br />

1<br />

ICMP 18569 Persea South Africa<br />

ICMP 18574 Pistacia Australia<br />

1 ICMP 18575 Capsicum Thailand<br />

0.99 0.99ICMP<br />

18618 Capsicum Thailand<br />

ICMP 19118 C. jasmini-sambac Jasminum Vietnam<br />

1<br />

ICMP 18121 Dioscorea Nigeria<br />

ICMP 17673 C. <strong>gloeosporioides</strong> “f. sp. aeschynomenes” Aeschynomene USA<br />

1 ICMP 18653 C. tropicale <strong>The</strong>obroma Panama<br />

1 ICMP 18651 Annona Panama<br />

1<br />

ICMP 18672 Litchi Japan<br />

1 ICMP 1780 Carica Australia<br />

1<br />

1 ICMP 12564 Persea Australia<br />

ICMP 1778 C. <strong>gloeosporioides</strong> var. minus Carica Australia<br />

1 ICMP 18705 Coffea Fiji<br />

1 ICMP 18693 Glycine (inoculated) Hungary<br />

ICMP 19051 C. <strong>gloeosporioides</strong> “f. sp. salsolae” Salsola Hungary<br />

1<br />

1<br />

ICMP 18605 Mangifera Thailand<br />

ICMP 18648 Mangifera Panama<br />

ICMP 18696 Mangifera Australia<br />

ICMP 18603 Mangifera Philippines<br />

ICMP 18580 C. asianum Coffea Thailand<br />

0.88 ICMP 12939 Citrus New Zealand<br />

ICMP 18678 Pueraria USA<br />

0.76 ICMP 17821 C. <strong>gloeosporioides</strong> Citrus Italy<br />

0.78 ICMP 18738 Carya Australia<br />

ICMP 12938 Citrus New Zealand<br />

1<br />

ICMP 18694 Mangifera South Africa<br />

ICMP 18695 Citrus USA<br />

ICMP 18730 Citrus New Zealand<br />

0.99<br />

ICMP 12066 Ficus New Zealand<br />

ICMP 18697 Vitis USA<br />

ICMP 19121 Gloeosporium pedemontanum lemon drink Italy<br />

1 ICMP 18122 Dioscorea Nigeria<br />

ICMP 17919 C. <strong>gloeosporioides</strong> “f. alatae” Dioscorea India<br />

C. fructicola<br />

C. nupharicola<br />

C. alienum<br />

C. musae<br />

C. aenigma<br />

C. siamense<br />

C. aeschynomenes<br />

C. tropicale<br />

C. queenslandicum<br />

C. salsolae<br />

C. asianum<br />

C. <strong>gloeosporioides</strong><br />

(sensu stricto)<br />

C. alatae<br />

Fig. 1. A Bayesian inference phylogenetic tree of 156 isolates in the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>. <strong>The</strong> tree was built using concatenated sequences of the<br />

ACT, CAL, CHS-1, GAPDH, and ITS genes each with a separate model of DNA evolution. Bayesian posterior probability values ≥ 0.5 are shown above nodes. Culture accession<br />

numbers are listed along with host plant genus and country of origin. Ex-type and authentic cultures are emphasised in bold font, and include the taxonomic name as originally<br />

described. Species delimitations are indicated with grey boxes. <strong>Colletotrichum</strong> boninense and C. hippeastri isolates are used as outgroups. <strong>The</strong> scale bar indicates the number<br />

of expected changes per site.


Fig. 1. (Continued).<br />

parts of the world, but has been isolated from other hosts as well,<br />

such as Ficus, Mangifera, Pueraria, and Vitis. Not all of the <strong>species</strong><br />

with a broad host range are found everywhere, for example in New<br />

Zealand C. alienum is commonly associated with cultivated fruits,<br />

whereas <strong>species</strong> such as C. siamense and C. fructicola, common<br />

on these same cultivated fruits in other parts of the world, have not<br />

been reported from New Zealand.<br />

Taxonomy<br />

Based on results of the multigene concatenated BI phylogenies,<br />

we accept 22 <strong>species</strong> plus one sub<strong>species</strong> within the C.<br />

www.studiesinmycology.org<br />

1.0<br />

1<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

0.99<br />

ICMP 17814 Fragaria USA<br />

ICMP 17099 Fragaria USA<br />

ICMP 17927 C. fragariae Fragaria USA<br />

1<br />

1 ICMP 18649 C. theobromicola <strong>The</strong>obroma Panama<br />

ICMP 18650 <strong>The</strong>obroma Panama<br />

0.72 ICMP 17100 Quercus USA<br />

ICMP 18567 Olea Australia<br />

0.99<br />

0.61<br />

0.64<br />

ICMP 17958 Stylosanthes Australia<br />

ICMP 15445 Acca New Zealand<br />

ICMP 17895 Annona Mexico<br />

0.88<br />

1 ICMP 17957 C. <strong>gloeosporioides</strong> f. stylosanthis<br />

Stylosanthes Australia<br />

1 ICMP 18565 Olea Australia<br />

ICMP 18566 Olea Australia<br />

ICMP 18576 Limonium Israel<br />

1 ICMP 17903 C. xanthorrhoeae Xanthorrhoea Australia<br />

ICMP 17820 Xanthorrhoea Australia<br />

ICMP 10492 C. horii Diospryos Japan<br />

1<br />

ICMP 17970 Diospryos Japan<br />

0.85<br />

ICMP 17968 Diospryos China<br />

1<br />

ICMP 12951 Diospryos New Zealand<br />

ICMP 12942 Diospryos New Zealand<br />

C. horii<br />

1 ICMP 18543 Melicytus New Zealand<br />

0.93 ICMP 17324 Kunzea New Zealand<br />

ICMP 18533 Prumnopitys New Zealand<br />

ICMP 18528 Berberis New Zealand<br />

0.97 ICMP 18535 Dacrycarpus New Zealand<br />

0.68 ICMP 18734 Berberis New Zealand<br />

ICMP 18540 Geniostoma New Zealand<br />

0.99 ICMP 18532 Vitex New Zealand<br />

ICMP 17326 Podocarpus New Zealand<br />

0.54 ICMP 18577 Coprosma New Zealand<br />

0.66 ICMP 18536 Coprosma New Zealand<br />

0.82 ICMP 18541 Coprosma New Zealand<br />

ICMP 18537 C. aotearoa Coprosma New Zealand<br />

0.66<br />

ICMP 18530 Vitex New Zealand<br />

ICMP 18735 Hedychium New Zealand<br />

0.79 ICMP 18548 Coprosma New Zealand<br />

ICMP 18736 Lonicera New Zealand<br />

1 0.69<br />

ICMP 18748 Ligustrum New Zealand<br />

0.83<br />

0.69 ICMP 18742 Meryta New Zealand<br />

C. aotearoa<br />

0.53 ICMP 18740 Dysoxylum New Zealand<br />

ICMP 18529 Acmena New Zealand<br />

1 ICMP 4832 C. ti Cordyline New Zealand<br />

ICMP 5285 Cordyline New Zealand<br />

ICMP 12952 Persea New Zealand<br />

1<br />

ICMP 18531 Persea New Zealand<br />

ICMP 12953 Persea New Zealand<br />

C. ti<br />

0.96<br />

ICMP 17915 Coffea Angola<br />

ICMP 17905 Coffea Cameroon<br />

ICMP 18544 Toronia New Zealand<br />

ICMP 17932 Dryandra South Africa<br />

ICMP 17922 Glomerella cingulata var. migrans Hypericum Germany<br />

1<br />

ICMP 17816 C. kahawae subsp. kahawae Coffea Kenya<br />

ICMP 18741 Kunzea New Zealand<br />

C. kahawae<br />

0.75<br />

ICMP 18534 Kunzea New Zealand<br />

ICMP 17931 Dryas Switzerland<br />

ICMP 18539 C. kahawae subsp. ciggaro Olea Australia<br />

ICMP 18728 Miconia Brazil<br />

0.99 ICMP 19122 Glomerella rufomaculans var. vaccinii Vaccinium USA<br />

ICMP 17928 Coffea Kenya<br />

1<br />

ICMP 17811 Coffea Malawi<br />

0.97 ICMP 10643 Glomerella cingulata “f. sp. camelliae” Camellia UK<br />

1 ICMP 18542 Camellia USA<br />

0.99 ICMP 10646 Camellia USA<br />

1 ICMP 18706 Vitis USA<br />

ICMP 18658 C. <strong>gloeosporioides</strong> “f. sp. clidemiae”Clidemia USA<br />

ICMP 19120 C. psidii Psidium Italy<br />

ICMP 18579 C. cordylinicola Cordyline Thailand<br />

1<br />

ICMP 17904 C. boninense Crinum Japan<br />

ICMP 17920 C. hippeastri Hippeastrum Netherlands<br />

C. theobromicola<br />

C. xanthorrhoeae<br />

Glomerella cingulata “f. sp.camelliae”<br />

C. clidemiae<br />

C. psidii<br />

C. cordylinicola<br />

<strong>gloeosporioides</strong> <strong>complex</strong>. Isolates authentic for G. cingulata “f.<br />

sp. camelliae” form a genetically distinct group, but this is not<br />

formally named because of doubts over its relationship to C.<br />

camelliae. Based on DNA sequence comparisons, a few other<br />

isolates almost certainly represent additional unnamed <strong>species</strong>.<br />

We do not formally describe them because most are known from<br />

a single isolate, often stale, with little understanding of either their<br />

morphological or biological diversity. In the Musae clade these<br />

include ICMP 18614 and ICMP 18616, both from grape from<br />

Japan, and ICMP 18726 from pawpaw from the Cook Islands,<br />

and in the Kahawae clade ICMP 18699 from chestnut from Japan.<br />

<strong>The</strong>se isolates are not included in the phylogenies in this study,<br />

127


Weir et al.<br />

128<br />

C. alatae<br />

C. asianum<br />

C. fructicola<br />

C. nupharicola<br />

C. alienum<br />

C. horii<br />

C. aenigma<br />

C. musae<br />

C. aeschynomenes<br />

C. siamense<br />

Musae clade<br />

C. tropicale<br />

C. ti<br />

C. queenslandicum<br />

C. aotearoa<br />

C. salsolae<br />

C. cordylinicola<br />

C psidii<br />

C. clidemia<br />

C. <strong>gloeosporioides</strong><br />

Kahawae clade<br />

G. cingulata “f.sp. camelliae”<br />

C. kahawae subsp. ciggaro<br />

C. kahawae subsp. kahawae<br />

C. xanthorrhoeae<br />

0.0060<br />

C. theobromicola<br />

1.0 0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

1<br />

1<br />

0.30<br />

0.73<br />

0.97<br />

0.99<br />

0.30<br />

Fig. 2. An unrooted Bayesian inference phylogenetic tree<br />

of ex-type and authentic cultures of the 24 taxa within<br />

the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>,<br />

illustrating their relative genetic distances, as indicated by<br />

branch lengths. <strong>The</strong>re are two clusters within the <strong>species</strong><br />

<strong>complex</strong>, the ‘Musae clade’ and the ‘Kahawae clade’. <strong>The</strong><br />

tree was build using concatenated sequences of the ACT,<br />

TUB2, CAL, CHS-1, GAPDH, GS, ITS, and SOD2 genes<br />

each with a separate model of DNA evolution.<br />

0.91<br />

0.41<br />

0.37<br />

0.57<br />

0.95<br />

1<br />

0.37<br />

0.47<br />

0.45<br />

0.64<br />

0.98<br />

1<br />

0.53<br />

0.46<br />

0.45<br />

0.47<br />

C. fructicola<br />

C. nupharicola<br />

C. alienum<br />

C. aeschynomenes<br />

C. siamense<br />

C. aenigma<br />

C. tropicale<br />

C. musae<br />

C. salsolae<br />

C. queenslandicum<br />

C. asianum<br />

C. <strong>gloeosporioides</strong><br />

C. alatae<br />

G.c. “f. sp. camelliae”<br />

C. kahawae<br />

C. clidemia<br />

C. ti<br />

C. cordylinicola<br />

C. psidii<br />

C. aotearoa<br />

C. horii<br />

C. xanthorrhoeae<br />

C. theobromicola<br />

C. boninense<br />

C. hippeastri<br />

Fig. 3. A Bayesian inference <strong>species</strong>-tree of the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>. <strong>The</strong> tree was built by grouping all 158 isolates into <strong>species</strong> and simultaneously estimating<br />

the individual five gene trees (ACT, CAL, CHS-1, GAPDH, and ITS) and the summary <strong>species</strong> tree using BEAST. <strong>The</strong> scale is an uncalibrated clock set relative to the last<br />

common ancestor of the C. <strong>gloeosporioides</strong> and C. boninense <strong>species</strong> <strong>complex</strong>es.


1<br />

0.13<br />

A<br />

0.87<br />

B<br />

1<br />

0.94<br />

www.studiesinmycology.org<br />

1<br />

1<br />

1<br />

0.117<br />

0.99<br />

1<br />

1<br />

1<br />

1<br />

1<br />

0.104<br />

0.82<br />

1<br />

1<br />

1<br />

1<br />

1<br />

1<br />

1<br />

1<br />

ICMP 18574 Pistacia Australia<br />

0.0030<br />

0.97<br />

1<br />

ICMP 18613 Limonium Israel<br />

ICMP 18727 Fragaria USA<br />

ICMP 18581 C. fructicola Coffea Thailand<br />

ICMP 17921 Glomerella cingulata var. minor Ficus Germany<br />

ICMP 18645 <strong>The</strong>obroma Panama<br />

ICMP 17795 Malus USA<br />

ICMP 17940 Nuphar USA<br />

ICMP 18691 Persea Australia<br />

ICMP 19118 C. jasmini-sambac Jasminum Vietnam<br />

ICMP 12567 Persea Australia<br />

ICMP 18121 Dioscoria Nigeria<br />

ICMP 18653 C. tropicale <strong>The</strong>obroma Panama<br />

ICMP 18705 Coffea Fiji<br />

ICMP 19051 C.g. “f. sp. salsolae” Salsola Hungary<br />

ICMP 18696 Mangifera Australia<br />

ICMP 18187 C. nupharicola Nuphar USA<br />

ICMP 18621 Persea New Zealand<br />

ICMP 12071 C. alienum Malus New Zealand<br />

ICMP 18686 Pyrus Japan<br />

ICMP 18646 C. ignotum Tetragastris Panama<br />

ICMP 18120 Dioscorea Nigeria<br />

0.69<br />

ICMP 17938 Nuphar USA<br />

ICMP 18608 C. aenigma Persea Israel<br />

1<br />

0.091<br />

1<br />

0.95<br />

1<br />

0.99<br />

1<br />

ICMP 17817 Musa Kenya<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

ICMP 19119 C. musae Musa USA<br />

ICMP 18642 C. hymenocallidis Hymenocallis China<br />

ICMP 18578 C. siamense Coffee Thailand<br />

ICMP 17673 C.g. “f. sp. aeschynomenes”Aeschynomene USA<br />

ICMP 18672 Litchi Japan<br />

1<br />

ICMP 1778 C. <strong>gloeosporioides</strong> var. minus Carica Australia<br />

ICMP 18580 C. asianum Coffea Thailand<br />

1<br />

0.078<br />

0.44<br />

0.55<br />

0.065<br />

0.33<br />

0.47<br />

0.75<br />

0.052<br />

ICMP 17821 C. <strong>gloeosporioides</strong> Citrus Italy<br />

0.90<br />

0.039<br />

0.97<br />

0.97<br />

0.60<br />

0.026<br />

0.013<br />

C. fructicola<br />

C. nupharicola<br />

C. alienum<br />

C. aenigma<br />

C. musae<br />

C. siamense<br />

C. aeschynomenes<br />

C. tropicale<br />

C. queenslandicum<br />

C. salsolae<br />

C. asianum<br />

0.0<br />

C. fructicola<br />

C. alienum<br />

C. nupharicola<br />

C. aenigma<br />

C. musae<br />

C. aeschynomenes<br />

C. siamense<br />

C. tropicale<br />

C. salsolae<br />

C. queenslandicum<br />

C. asianum<br />

C. <strong>gloeosporioides</strong><br />

Fig. 4. A Bayesian inference phylogenetic tree of 32 selected isolates in the Musae clade of the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>. <strong>The</strong> tree was build using<br />

concatenated sequences of the ACT, TUB2, CAL, CHS-1, GAPDH, GS, ITS, and SOD2 genes each with a separate model of DNA evolution. Other details as per Fig.1. B. A<br />

<strong>species</strong>-tree constructed from the same data, the scale is an clock set relative to the last common ancestor of the Musae clade and C. <strong>gloeosporioides</strong> s. str., as calibrated in<br />

Fig. 3.<br />

but DNA sequences from these isolates have been accessioned<br />

into GenBank (ITS: JX009423–JX009428, GAPDH: JX009416–<br />

JX009422, ACT: JX009404–JX009407, CAL: JX009408–<br />

JX009411, CHS-1: JX009412–JX009415).<br />

Many of the <strong>species</strong> that we recognise fall into one of two<br />

clades, the informally named Musae clade and Kahawae clade (Fig.<br />

2). Each clade contains several <strong>species</strong> that are phylogenetically<br />

well supported in multi-gene analyses, but within the clades branch<br />

lengths are short because of the small number of phylogenetically<br />

informative characters. This is reflected in the low support values in<br />

the gene tree analyses for the <strong>species</strong> we accept within that clade<br />

(Figs 3, 4). Both the Musae and Kahawae clades contain ex-type<br />

or authentic cultures from several long accepted <strong>species</strong>. In this<br />

work we have made a pragmatic decision to minimise taxonomic<br />

disruption, so that monophyletic subclades within the Kahawae<br />

and Musae clades are accepted as <strong>species</strong> where they include<br />

129


Weir et al.<br />

ex-type or authentic cultures. <strong>The</strong> Musae clade thus includes C.<br />

fructicola, C. musae, C. nupharicola, C. siamense, and C. tropicale;<br />

and the Kahawae clade includes C. cordylinicola, C. psidii, and C.<br />

kahawae. Also belonging in the latter is Glomerella cingulata “f.<br />

sp. camelliae”. To provide a consistent taxonomic treatment of the<br />

subclades resolved within the Musae and Kahawae clades, several<br />

new <strong>species</strong> and one new sub<strong>species</strong> are proposed. In the Musae<br />

130<br />

1<br />

A<br />

B<br />

0.62<br />

0.94<br />

0.0040<br />

0.58<br />

1<br />

1<br />

0.99 ICMP 17905 Coffea Cameroon<br />

1 ICMP 17915 Coffea Angola<br />

ICMP 17811 Coffea Malawi<br />

1<br />

ICMP 17816 C. kahawae Coffea Kenya<br />

0.89<br />

ICMP 18534 Kunzea New Zealand<br />

0.95 ICMP 17931 Dryas Switzerland<br />

C. kahawae<br />

subsp.<br />

1 ICMP 17922 G. c. var. migrans Hypericum Germany<br />

1 0.89 ICMP 19122 G. rufomaculans var. vaccinii Vaccinium USA<br />

0.98<br />

ICMP 18539 C. kahawae subsp. ciggaro Olea Australia<br />

ICMP 12952 Persea New Zealand<br />

ICMP 10643 Glomerella cingulata “f. sp. camelliae” Camellia UK<br />

1 ICMP 18542 Camellia USA<br />

1<br />

ICMP 10646 Camellia USA<br />

ICMP 18706 Vitis USA<br />

ICMP 18658 C. <strong>gloeosporioides</strong> “f. sp.clidemia” Clidemia USA<br />

ICMP 19120 C. psidii Psidium Italy<br />

C. psidii<br />

0.99 ICMP 18548 Coprosma New Zealand<br />

0.99 ICMP 18537 C. aotearoa Coprosma New Zealand<br />

1<br />

1 ICMP 18735 Hedychium New Zealand<br />

ICMP 18740 Dysoxylum New Zealand<br />

0.98<br />

0.56<br />

ICMP 18530 Vitex New Zealand<br />

ICMP 18532 Vitex New Zealand<br />

ICMP 17324 Kunzea New Zealand<br />

1<br />

ICMP 18533 Prumnopitys New Zealand<br />

1<br />

ICMP 18577 Coprosma New Zealand<br />

1 ICMP 17326 Podocarpus New Zealand<br />

ICMP 18535 Dacrycarpus New Zealand<br />

1 ICMP 4832 C. ti Cordyline New Zealand<br />

ICMP 5285 Cordyline New Zealand<br />

ICMP 18579 C. cordylinicola Cordyline Thailand<br />

C. ti<br />

ICMP 17821 C. <strong>gloeosporioides</strong> Citrus Italy<br />

0.2 0.18 0.16 0.14 0.12 0.1<br />

0.08 0.06 0.04 0.02<br />

1<br />

0.17<br />

0.32<br />

0.39<br />

1<br />

0.91<br />

1<br />

C. kahawae<br />

subsp. kahawae<br />

ciggaro<br />

G. cingulata “f. sp. camelliae”<br />

C. clidemia<br />

C. aotearoa<br />

C. cordylinicola<br />

C. kahawae subsp. ciggaro<br />

C. kahawae subsp. kahawae<br />

G. c. “f. sp. camelliae”<br />

C. clidemia<br />

C. ti<br />

C. psidii<br />

C. cordylinicola<br />

C. aotearoa<br />

C. <strong>gloeosporioides</strong><br />

Fig. 5. A Bayesian inference phylogenetic tree of 30 selected isolates in the Kahawae clade of the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>. <strong>The</strong> tree was build using<br />

concatenated sequences of the ACT, TUB2, CAL, CHS-1, GAPDH, GS, ITS, and SOD2 genes each with a separate model of DNA evolution. Other details as per Fig.1. B. A<br />

<strong>species</strong>-tree constructed from the same data, the scale is a clock set relative to the last common ancestor of the Kahawae clade and C. <strong>gloeosporioides</strong> s. str., as calibrated<br />

in Fig. 3.<br />

clade these are C. aenigma, C. aeschynomenes, C. alienum,<br />

C. queenslandicum, and C. salsolae; in the Kahawae clade C.<br />

aotearoa, C. clidemiae, C. kahawae subsp. ciggaro, and C. ti. <strong>The</strong><br />

other accepted <strong>species</strong>, well resolved in all of the gene trees, are C.<br />

alatae, C. asianum, C. <strong>gloeosporioides</strong>, C. horii, C. theobromicola,<br />

and C. xanthorrhoeae.<br />

0.0


B<br />

A<br />

www.studiesinmycology.org<br />

0.3<br />

1bp<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

C.queenslandicum<br />

C.tropicale<br />

C.ti<br />

C.clidemiae<br />

C.siamense<br />

C.aotearoa<br />

C.siamense<br />

C.ti<br />

C.aotearoa<br />

C.aotearoa<br />

C.alienum<br />

C.siamense<br />

C.aenigma<br />

C.salsolae<br />

C.siamense<br />

C.kahawae<br />

C.alienum<br />

C.theobromicola<br />

C.theobromicola<br />

C.theobromicola<br />

C.horii<br />

C.aotearoa<br />

C.fructicola<br />

C.aeschynomenes<br />

C.fructicola<br />

C.queenslandicum<br />

C. nupharicola<br />

C.queenslandicum<br />

C.siamense<br />

C.siamense<br />

C.<strong>gloeosporioides</strong><br />

C.<strong>gloeosporioides</strong><br />

C.aotearoa<br />

C.nupharicola<br />

C.kahawae<br />

C.aotearoa<br />

C.kahawae<br />

C.aotearoa<br />

C.psidii<br />

G. c. “f. sp. camelliae”<br />

C.cordylinicola<br />

C.asianum<br />

C.musae<br />

C.musae<br />

C.alatae<br />

C.xanthorrhoeae<br />

C.hippeastri<br />

C.boninense<br />

Fig. 6. An UPGMA tree of ITS sequences from 156 isolates in the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>. Isolate names have been replaced with <strong>species</strong> present<br />

in each clade. Species that are in monophyletic clades are emphasised in bold font to indicate those for which ITS barcoding is likely to work well. B: A 50 % majority-rule<br />

consensus Bayesian inference tree of the same data, showing the collapse of structure when analysed with a more robust method.<br />

131


Weir et al.<br />

Conidial width variation<br />

Conidial length variation<br />

132<br />

8<br />

35<br />

7<br />

30<br />

6<br />

25<br />

20<br />

5<br />

width<br />

15<br />

4<br />

(µm)<br />

length (µm)<br />

10<br />

G. c.“f.sp. camelliae”<br />

C. xanthorrhoeae<br />

C. tropicale<br />

C. ti<br />

C. theobromicola<br />

C. siamense<br />

C. salsaloe<br />

C. queenslandicum<br />

C. nupharicola<br />

C. musae<br />

C. kahawae ssp. kahawae<br />

C. kahawae ssp. ciggaro<br />

C. horii<br />

C. gloeosporiodes<br />

C. fructicola<br />

C. clidemiae<br />

C. asianum<br />

C. aotearoa<br />

C. alienum<br />

C. alatae<br />

C. aeschynomenes<br />

C. aenigma<br />

G. c.“f.sp. camelliae”<br />

C. xanthorrhoeae<br />

C. tropicale<br />

C. ti<br />

C. theobromicola<br />

C. siamense<br />

C. salsaloe<br />

C. queenslandicum<br />

C. nupharicola<br />

C. musae<br />

C. kahawae ssp. kahawae<br />

C. kahawae ssp. ciggaro<br />

C. horii<br />

C. <strong>gloeosporioides</strong><br />

C. fructicola<br />

C. clidemiae<br />

C. asianum<br />

C. aotearoa<br />

C. alienum<br />

C. alatae<br />

C. aeschynomenes<br />

C. aenigma<br />

Fig. 7. Box plots showing the variation<br />

in length and width of conidia produced<br />

by the cultures examined in this study.<br />

<strong>The</strong> dashed lines show the mean length<br />

(16.74 µm) and width (5.1 µm) across the<br />

<strong>species</strong> <strong>complex</strong> (n = 1958).


Ascospore width variation<br />

Ascospore length variation<br />

6.5<br />

24<br />

6.0<br />

22<br />

www.studiesinmycology.org<br />

5.5<br />

20<br />

5.0<br />

18<br />

4.5<br />

16<br />

4.0<br />

width (µm)<br />

length (µm)<br />

14<br />

3.5<br />

C. ti<br />

C. fructicola<br />

C. clidemiae<br />

C. alienum<br />

C. aenigma<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

C. xanthorrhoeae<br />

C. ti<br />

C. kahawae ssp. ciggaro<br />

C. fructicola<br />

C. clidemiae<br />

C. alienum<br />

C. aenigma<br />

C. xanthorrhoeae<br />

C. kahawae ssp. ciggaro<br />

Fig. 8. Box plots showing the variation in<br />

length and width of ascospores produced<br />

by the cultures examined in this study. <strong>The</strong><br />

dashed lines show the mean length (17.46<br />

µm) and width (4.8 µm) across the <strong>species</strong><br />

<strong>complex</strong> (n = 452).<br />

133


Weir et al.<br />

134<br />

Culture size at10days<br />

80<br />

60<br />

40<br />

diameter (mm)<br />

20<br />

G. c.“f.sp. camelliae”<br />

C. xanthorrhoeae<br />

C. tropicale<br />

C. ti<br />

C. theobromicola<br />

C. siamense<br />

C. salsaloe<br />

C. queenslandicum<br />

C. nupharicola<br />

C. musae<br />

C. kahawae ssp. kahawae<br />

C. kahawae ssp. ciggaro<br />

C. <strong>gloeosporioides</strong><br />

C. fructicola<br />

C. cordylinicola<br />

C. clidemiae<br />

C. asianum<br />

C. aotearoa<br />

C. alienum<br />

C. alatae<br />

C. aeschynomene<br />

C. aenigma<br />

Fig. 9. A box plot of the diameter of cultures grown on PDA agar at 18 °C for 10 d. <strong>The</strong> dashed line shows the mean culture size (61.56 mm) across the <strong>species</strong> <strong>complex</strong> (n =<br />

719). Note that the data is skewed by some fast growing cultures that reached the agar plate diam (85 mm) in under 10 d.


* <strong>Colletotrichum</strong> aenigma B. Weir & P.R. Johnst., sp. nov.<br />

MycoBank MB563759. Fig. 10.<br />

Etymology: from the Latin aenigma, based on the enigmatic<br />

biological and geographic distribution of this <strong>species</strong>.<br />

Holotype: Israel, on Persea americana, coll. S. Freeman Avo-37-<br />

4B, PDD 102233; ex-holotype culture ICMP 18608.<br />

Colonies grown from single conidia on Difco PDA 30–35 mm<br />

diam after 10 d. Aerial mycelium sparse, cottony, white, surface<br />

of agar uniformly pale orange (7A5) towards centre, more<br />

or less colourless towards edge, conidia not associated with<br />

well differentiated acervuli and no masses of conidial ooze. In<br />

reverse pale orange towards centre. Conidiogenous cells arising<br />

haphazardly from dense, tangled hyphae across agar surface,<br />

short-cylindric with a poorly differentiated conidiogenous locus.<br />

Conidia often germinating soon after release, sometimes forming<br />

appressoria, so forming a thin, compact, layer of germinated,<br />

septate conidia, germ tubes, and appressoria across the central<br />

part of the colony surface. Conidia (12–)14–15(–16.5) × (5–)6–<br />

6.5(–7.5) µm (av. 14.5 × 6.1 µm, n = 53), cylindric with broadly<br />

rounded ends. Appressoria 6–10 µm diam, subglobose or with a<br />

few broad lobes.<br />

Geographic distribution and host range: known from only two<br />

collections, one from Pyrus pyrifolia from Japan, the other from<br />

Persea americana from Israel.<br />

Genetic identification: ITS sequences are insufficient to separate C.<br />

aenigma from C. alienum and some C. siamense isolates. <strong>The</strong>se<br />

taxa are best distinguished using TUB2 or GS.<br />

Notes: Although the biology of this <strong>species</strong> is more or less unknown,<br />

it has been found in two widely separate regions and is, therefore,<br />

likely to be found to be geographically widespread in the future.<br />

Genetically distinct within the Musae clade, this <strong>species</strong> has a<br />

distinctive appearance in culture with sparse, pale aerial mycelium<br />

and lacking differentiated acervuli.<br />

Other specimen examined: Japan, on Pyrus pyrifolia, coll. H. Ishii Nashi-10 (ICMP<br />

18686).<br />

* <strong>Colletotrichum</strong> aeschynomenes B. Weir & P.R. Johnst.,<br />

sp. nov. MycoBank MB563590. Fig. 11.<br />

= C. <strong>gloeosporioides</strong> “f. sp. aeschynomenes” (Daniel et al. 1973, as<br />

aeschynomene).<br />

Etymology: Based on C. <strong>gloeosporioides</strong> “f. sp. aeschynomenes”,<br />

referring to the host from which this <strong>species</strong> was originally<br />

described.<br />

Holotype: USA, Arkansas, on Aeschynomene virginica stem lesion,<br />

coll. D. TeBeest 3-1-3, PDD 101995; ex-type culture ICMP 17673<br />

= ATCC 201874.<br />

Colonies grown from single conidia on Difco PDA 25–35 mm diam<br />

after 10 d, aerial mycelium sparse, cottony, white, surface of colony<br />

with numerous acervuli, some with dark bases, with orange conidial<br />

ooze; in reverse more or less colourless apart from the dark<br />

acervuli and orange conidial masses showing through the agar.<br />

Conidia (14–)17–18.5(–20) × 4(–5) µm (av. 17.6 × 4.1 µm, n =<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

30), cylindric, straight, tapering slightly near both ends. Appressoria<br />

mostly elliptic to subfusoid, deeply lobed. Perithecia not seen.<br />

Geographic distribution and host range: Reported only from USA,<br />

pathogenic to Aeschynomeme.<br />

Genetic identification: ITS sequences do not distinguish<br />

C. aeschynomenes from C. fructicola. <strong>The</strong>se taxa are best<br />

distinguished using TUB2, GAPDH, or GS.<br />

Notes: <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. aeschynomenes”<br />

has been used to refer to isolates pathogenic to Aeschynomene<br />

virginica, later developed as the weed biocontrol agent Collego<br />

(references in Ditmore et al. 2008). It has also been reported from<br />

a range of other hosts (TeBeest 1988). Our analyses, based on a<br />

single, authentic strain of C. <strong>gloeosporioides</strong> “f. sp. aeschynomenes”<br />

(TeBeest 3.1.3, apparently the source of the single spore isolate<br />

originally used in the development of Collego, Ditmore et al. (2008))<br />

show it to be genetically distinct within the Musae clade of the C.<br />

<strong>gloeosporioides</strong> <strong>complex</strong>. Genetically close to the geographically<br />

and biologically diverse C. siamense, it differs morphologically from<br />

this <strong>species</strong> in having slightly longer and narrower conidia which<br />

taper slightly toward the ends, and in having larger, strongly lobed<br />

appressoria.<br />

An isolate deposited as C. <strong>gloeosporioides</strong> f. sp.<br />

aeschynomenes in <strong>CBS</strong> (<strong>CBS</strong> 796.72) by G.E. Templeton, one of<br />

the early C. <strong>gloeosporioides</strong> f. sp. aeschynomenes researchers<br />

(Daniel et al. 1973), is genetically distinct to TeBeest 3.1.3 and has<br />

been identified by Damm et al. (2012a, this issue) as C. godetiae, a<br />

member of the C. acutatum <strong>complex</strong>. <strong>The</strong> strain that we examined<br />

(Te Beest 3.1.3) matches genetically another strain often cited<br />

in the C. <strong>gloeosporioides</strong> f. sp. aeschynomenes literature (Clar-<br />

5a = ATCC 96723) (GenBank JX131331). It is possible that two<br />

distinct <strong>species</strong>, both highly pathogenic to Aeschynomene in<br />

Arkansas, have been confused. A survey of additional isolates of<br />

<strong>Colletotrichum</strong> highly virulent to Aeschynomene in Arkansas would<br />

clarify the interpretation of the past literature on this pathogen. For<br />

example, C. <strong>gloeosporioides</strong> “f. sp. aeschynomenes” was initially<br />

reported as specific to Aeschynomene virginica (Daniel et al.<br />

1973), while later studies reported isolates putatively of the same<br />

taxon, to have a wider host range (TeBeest 1988).<br />

Cisar et al. (1994) reported fertile ascospores from<br />

crosses between isolates identified as C. <strong>gloeosporioides</strong> “f.<br />

sp. aeschynomenes” and isolates of C. <strong>gloeosporioides</strong> “f. sp.<br />

jussiaeae”, a pathogen of Jussiaea decurrens. <strong>The</strong> position of<br />

C. <strong>gloeosporioides</strong> “f. sp. jussiaeae” within our phylogeny is not<br />

known, but these taxa could prove useful for better understanding<br />

of the biological differences between phylogenetically defined<br />

<strong>species</strong> of <strong>Colletotrichum</strong>.<br />

Specimen examined: USA, Arkansas, on Aeschynomene virginica stem lesion, coll.<br />

D. TeBeest 3.1.3 (ICMP 17673 = ATCC 201874).<br />

* <strong>Colletotrichum</strong> alatae B. Weir & P.R. Johnst., sp. nov.<br />

MycoBank MB563747. Fig. 12.<br />

= <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. alatae” R.D. Singh, Prasad & R.L. Mathur,<br />

Indian Phytopathol. 19: 69. 1966. [nom. inval., no Latin description, no type<br />

designated].<br />

Etymology: Based on the invalid name C. <strong>gloeosporioides</strong> “f. alatae”<br />

(Singh et al. 1966), referring to Dioscorea alata, the scientific name<br />

for yam.<br />

135


Weir et al.<br />

Fig. 10. <strong>Colletotrichum</strong> aenigma. A, C, D, E, F. ICMP 18608 – ex-holotype culture. B. ICMP 18616. A–B. Cultures on PDA, 10 d growth from single conidia, from above and<br />

below. C–D. Conidia. E–F. Appressoria. Scale bar C = 20 µm. Scale bar of C applies to C–F.<br />

Holotype: India, Rajasthan, Udaipur, on Dioscorea alata leaves and<br />

stems, coll. K.L. Kothari & J. Abramham, 1959, <strong>CBS</strong> H-6939; extype<br />

culture and putatively authentic isolate of C. <strong>gloeosporioides</strong> f.<br />

alatae <strong>CBS</strong> 304.67 = ICMP 17919.<br />

136<br />

Colonies grown from single conidia on Difco PDA 30–40 mm diam<br />

after 10 d. Ex-holotype culture looks “stale”, with low, felted, dense,<br />

pale grey aerial mycelium, orange agar surface showing through<br />

near the margin, scattered dark based acervuli with orange conidial


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 11. <strong>Colletotrichum</strong> aeschynomenes. ICMP 17673 – ex-holotype culture. A–C. Appressoria. D. Conidiogenous cells. E. Conidia. F. Cultures on PDA, 10 d growth from single<br />

conidia, from above and below. Scale bar of A = 20 µm. Scale bar of A applies to A–E.<br />

masses near centre; in reverse deep pinkish orange with patches of<br />

grey pigment near centre. ICMP 18122 with aerial mycelium sparse,<br />

colony surface with numerous discrete, dark-based acervuli with<br />

bright orange conidial ooze, margin of colony feathery; in reverse<br />

irregular sectors with pale grey pigment within the grey, otherwise<br />

colourless apart from the colour of the acervuli and conidial<br />

masses. Conidia (14.5−)18–19.5(−23.5) × (4.5−)5−5.5(−6.5) µm<br />

(av. 18.9 × 5.2 µm, n = 40), cylindric, straight, ends rounded, a few<br />

tapering towards the basal end. Appressoria mostly simple, elliptic<br />

to fusoid in shape, sometime developing broad, irregular lobes,<br />

about 7–13.5 × 5–10.5 µm. Perithecia not seen.<br />

137


Weir et al.<br />

Fig. 12. <strong>Colletotrichum</strong> alatae. ICMP 18122. A. Cultures on PDA, 10 d growth from single conidia, from above and below. B–C. Appressoria. D. Conidiogenous cells and conidia.<br />

E. Conidia. F. Setae. Scale bars B, F = 20 µm. Scale bar of B applies to B–E.<br />

138


Geographic distribution and host range: Known only from yam<br />

(Dioscorea alata), from Nigeria, Barbardos, India, Guadeloupe.<br />

Genetic identification: ITS sequences distinguish C. alatae from all<br />

other taxa.<br />

Notes: Anthracnose diseases of yam are found throughout the<br />

regions where the host is grown (e.g. Winch et al. 1984, Prasad &<br />

Singh 1960, Singh et al. 1966, Abang et al. 2002, 2003). Isolates<br />

from diseased yam leaves are morphologically (Winch et al. 1984)<br />

and genetically (Abang et al. 2002) diverse. Both of these authors<br />

used a broad <strong>species</strong> concept, grouping all isolates sourced from<br />

yam under the single name C. <strong>gloeosporioides</strong>. In this paper we<br />

accept part of that diversity to represent a distinct <strong>species</strong>, newly<br />

described here as C. alatae. <strong>The</strong> type specimen of C. alatae<br />

matches the SGG (slow growing grey) group of Abang et al. (2002),<br />

the group that these authors found to be more pathogenic to yam<br />

than the other morphological and genetic groups they recognised<br />

within C. <strong>gloeosporioides</strong>. In addition to the Nigerian isolates of<br />

Abang et al. (2002), isolates from yam from Barbados (isolates<br />

SAS8 and SAS9 from Sreenivasaprasad et al. 1996), Guadeloupe<br />

(GenBank accession GQ495617) and India (<strong>CBS</strong> 304.67 and<br />

GenBank accession FJ940734) belong in this clade, while no<br />

isolates from other hosts have been found.<br />

Other isolates from yam that we sequenced included some<br />

representing the Abang et al. (2002) FGS group (Abang Cg22 =<br />

ICMP 18120, Abang Cg13 = ICMP 18125, Abang CgS6 = ICMP<br />

18117, Abang CgS2 = ICMP 18121), a group distinguished from<br />

the highly pathogenic SGG isolates by faster growth in culture and<br />

shorter conidia (Abang et al. 2002). Two of these isolates (ICMP<br />

18120, 18125) genetically match C. fructicola, the others match C.<br />

siamense.<br />

Several names have been applied to <strong>Colletotrichum</strong><br />

specimens from anthracnose of yam stems and leaves, including<br />

Gloeosporium pestis Massee, G. “dioscoreae” Sawada (nom.<br />

inval.; no Latin diagnosis), <strong>Colletotrichum</strong> dioscoreae Av.-Saccá<br />

1917, and C. dioscoreae Tehon 1933. In addition, Gloeosporium<br />

bomplandii Speg. was described from a host doubtfully<br />

identified as Dioscorea. Because of the broad genetic diversity<br />

of <strong>Colletotrichum</strong> spp. associated with diseased yam, the lack<br />

of cultures from any of these early type specimens, and the<br />

uncertainty to which part of the yam-associated diversity they<br />

correspond, we have chosen not to use these names for our<br />

newly recognised, yam-specialised pathogen. Whether the postharvest<br />

tuber rot referred to as dead skin disease of yam (Abang<br />

et al. 2003, Green & Simmons 1994) is caused by the same<br />

<strong>Colletotrichum</strong> population as associated with diseased foliage is<br />

not known.<br />

Other specimen examined: Nigeria, Kpite, on Dioscorea alata leaf, coll. M.M. Abang<br />

Cg25, 2001 (ICMP 18122).<br />

* <strong>Colletotrichum</strong> alienum B. Weir & P.R. Johnst., sp. nov.<br />

MycoBank MB563591. Figs 13, 14.<br />

Etymology: Based on the biology of this <strong>species</strong>, confined to exotic<br />

hosts and presumed to be a recent introduction to Australasia.<br />

Holotype: New Zealand, Auckland, Kumeu research orchard,<br />

Malus domestica fruit rot, coll. P.R. Johnston C824, 14 Aug. 1987,<br />

PDD 101996; ex-type culture ICMP 12071.<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Colonies grown from single conidia on Difco PDA 85 mm diam after<br />

10 d. Colonies often with distinct sectors; some with cottony, grey<br />

aerial mycelium with numerous dark-based acervuli and orange<br />

conidial ooze visible through the mycelium; others with dense,<br />

cottony to felted mycelium, fewer acervuli and these hidden by<br />

the dense mycelium. In reverse, irregular dark grey patches and<br />

sectors masking the pale orange coloured pigmentation. ICMP<br />

18691 looks “stale” with slow growth, dense, pale aerial mycelium<br />

and sparse conidial production and no perithecia. Conidia (12.5–)<br />

15.5–17.5(–22) × (3–)5–5.5(–6) µm (av. 16.5 × 5.0 µm, n = 70),<br />

cylindric with broadly rounded ends. Appressoria mostly simple,<br />

globose to short-cylindric, a few with broad, irregular lobes; ICMP<br />

18691 has mostly lobed appressoria. Perithecia forming in most<br />

cultures after about 10 d, dark-walled, globose with short, narrow<br />

ostiolar neck. Ascospores (14.5–)17–19.5(–22) × 4–5(–6) µm (av.<br />

18.1 × 4.6 µm, n = 55), cylindric, curved, tapering slightly to each<br />

end.<br />

Geographic distribution and host range: Known only from Australia<br />

and New Zealand, common on a wide range of introduced fruit<br />

crops.<br />

Genetic identification: ITS sequences do not separate C. alienum<br />

from some C. siamense isolates. <strong>The</strong>se taxa are best distinguished<br />

using CAL or GS.<br />

Notes: Common on commercial fruit crops, this fungus was referred<br />

to as C. <strong>gloeosporioides</strong> Group A by Johnston & Jones (1997) and<br />

Johnston et al. (2005).<br />

Other specimens examined: Australia, New South Wales, Murwillumbah, on Persea<br />

americana (DAR 37820 = IMI 313842 = ICMP 18691). New Zealand, Auckland,<br />

Oratia, Shaw Rd, on Malus domestica fruit rot, coll. P.R. Johnston C938.5, 14 Apr.<br />

1988 (ICMP 18725); Bay of Plenty, Katikati, on Diospyros kaki ripe fruit rot, coll. M.A.<br />

Manning, Jun. 1989 (ICMP 17972); Bay of Plenty, Te Puke, on Persea americana<br />

ripe fruit rot, coll. W.F.T. Hartill, 2 Feb. 1988 (ICMP 18704); Bay of Plenty, Te Puna,<br />

on Persea americana ripe fruit rot, coll. W.F.T. Hartill, 25 Jan. 1988 (ICMP 18703);<br />

Bay of Plenty, on Persea americana ripe fruit rot, coll. W.F.T. Hartill, Feb. 1991<br />

(ICMP 18621); Waikato, Hamilton, on Malus domestica fruit rot, coll. G.I. Robertson,<br />

May 1988 (ICMP 12068).<br />

* <strong>Colletotrichum</strong> aotearoa B. Weir & P.R. Johnst., sp. nov.<br />

MycoBank MB800213. Figs 15, 16.<br />

Etymology: Based on the Maori name for New Zealand; most<br />

isolates from native New Zealand plants belong here.<br />

Holotype: New Zealand, Auckland, Glen Innes, Auckland University<br />

campus, on Coprosma sp. incubated berries, coll. B. Weir C1282.4,<br />

30 Apr 2009, PDD 101076; ex-type culture ICMP 18537.<br />

Colonies grown from single conidia on Difco PDA 70–85 mm diam<br />

after 10 d, several isolates with restricted growth, 50–55 mm diam<br />

with an irregularly scalloped margin. Aerial mycelium cottony to<br />

dense cottony, tufted near centre, grey to dark grey, scattered,<br />

small, dark-based acervuli and large, globose, stromatic structures<br />

partially embedded in agar, these sometimes splitting apart and<br />

forming conidia. In reverse typically with pinkish-orange pigments,<br />

variable in intensity, in some isolates this colour partially hidden<br />

by more or less concentric bands of dark grey pigment. Conidia<br />

(12–)16–17.5(–21.5) × (4.5–)5–5.5(–6.5) µm (av. 16.9 × 5.2 µm,<br />

n = 216), cylindric, straight, apex broadly rounded, often tapering<br />

slightly towards subtruncate base, 0-septate, hyaline. Appressoria<br />

139


Weir et al.<br />

Fig. 13. <strong>Colletotrichum</strong> alienum. A, E, F. ICMP 12071 – ex-holotype culture. B. ICMP 18703. C–D. ICMP 12068. G–I. ICMP 18691 (ex DAR 37820). A–B. Appressoria. C–D. Asci<br />

and ascospores. E. Conidia. F. Conidiogenous cells. G. Appressoria. H. Conidia. I. Conidiogenous cells. Scale bar D = 20 µm. Scale bar of D applies to A–I.<br />

140


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 14. <strong>Colletotrichum</strong> alienum. A. ICMP 12071 – ex-holotype culture. B. ICMP 12068. C. ICMP 18691 (ex DAR 37820). A–C. Cultures on PDA, 10 days growth from single<br />

conidia, from above and below.<br />

variable in shape, simple to broadly lobed, sometimes in groups,<br />

sometimes intercalary, about 7–17 × 4–9.5 µm. Perithecia not seen<br />

in culture.<br />

Geographic distribution and host range: Confirmed only from New<br />

Zealand, but GenBank records suggest C. aotearoa also occurs in<br />

China (see below). In New Zealand this <strong>species</strong> is common on a<br />

taxonomically diverse set of native plants, as both a fruit rot and a<br />

leaf endophyte, and has also been isolated from leaves of several<br />

<strong>species</strong> of naturalised weeds.<br />

Genetic identification: ITS sequences do not separate C. aotearoa<br />

from several taxa in the Kahawae and Musae clades. This <strong>species</strong><br />

can be distinguished using several other genes, including TUB2,<br />

CAL, GS, and GAPDH.<br />

Notes: All isolates in the C. <strong>gloeosporioides</strong> <strong>complex</strong> from New<br />

Zealand native plants studied here belong in the Kahawae<br />

clade, and most of these are C. aotearoa; a small number of leaf<br />

endophyte isolates from New Zealand native trees are C. kahawae<br />

subsp. ciggaro. <strong>The</strong> C. aotearoa isolates have been isolated as<br />

endophytes from symptomless leaves as well as from rotting fruit<br />

from native trees. Morphologically indistinguishable from isolates<br />

of C. kahawae subsp. ciggaro, this <strong>species</strong> is distinguished<br />

genetically with all genes sampled, except ITS. <strong>The</strong> GAPDH gene<br />

tree splits C. aotearoa into two well supported clades, but these do<br />

not correlate to any other features, either geographic or biological.<br />

Isolates associated with distinctive and common leaf spots on<br />

Meryta sinclairii, first recorded by Beever (1984), belong in this<br />

<strong>species</strong>. Whether isolates of C. aotearoa from other hosts are able<br />

to cause the same disease on Meryta is not known.<br />

Also in C. aotearoa are a range of isolates from weeds that<br />

have become naturalised in New Zealand. We assume that C.<br />

aotearoa is a New Zealand native <strong>species</strong>. It has a broad host<br />

range amongst native plants and has apparently jumped host to<br />

some weeds. It has never been found associated with cultivated<br />

plants or as a rot of cultivated fruit.<br />

<strong>Colletotrichum</strong> aotearoa may also occur in China. ITS<br />

sequences from isolates from Boehmeria from China (GenBank<br />

records GQ120479 and GQ120480) from Wang et al. (2010)<br />

match exactly a set of C. aotearoa isolates. ITS between-<strong>species</strong><br />

differences within the C. <strong>gloeosporioides</strong> <strong>complex</strong> are very small,<br />

so this match needs confirming with additional genes. C. aotearoa<br />

was referred to as Undescribed Group 2 by Silva et al. (2012b).<br />

Other specimens examined: New Zealand, Auckland, Freemans Bay, on Vitex<br />

lucens fruit, coll. P.R. Johnston C1252.1, 26 Aug. 2007 (ICMP 18532; PDD 92930).<br />

on Berberis sp. leaf spot, coll. N. Waipara C69 (ICMP 18734); Auckland, Mangere,<br />

on Berberis glaucocarpa leaf spot, coll. N. Waipara C7, Jun. 2007 (ICMP 18528);<br />

Auckland, Waitakere Ranges, on Kunzea ericoides leaf endophyte, coll. S. Joshee<br />

7Kun3.5, Jan. 2004 (ICMP 17324); Auckland, Waitakere Ranges, on Prumnopitys<br />

ferruginea leaf endophyte, coll. S. Joshee 8Mb5.1, Jan. 2004 (ICMP 18533); Auckland,<br />

Waitakere Ranges, on Dacrycarpus dacrydioides leaf endophyte, coll. S. Joshee<br />

5K5.9, Jan. 2004 (ICMP 18535); Auckland, St Johns, Auckland University campus,<br />

on Coprosma sp. incubated berries, coll. B. Weir C1282.1, 30 Apr. 2009 (ICMP<br />

18577); Auckland, Mt Albert, on Acmena smithii lesions fruit, coll. P.R. Johnston C847,<br />

9 Sep. 1987 (ICMP 18529); Auckland, Glen Innes, Auckland University campus, on<br />

Coprosma sp. incubated berries, coll. B. Weir C1282.3, 30 Apr. 2009 (ICMP 18536);<br />

Auckland, Orakei, on Ligustrum lucidum leaf spot, coll. C. Winks & D. Than M136.3<br />

(ICMP 18748); Auckland, Waitakere Ranges, on Podocarpus totara leaf endophyte,<br />

coll. S. Joshee 3T5.6, Jan. 2004 (ICMP 17326); Auckland, Waitakere Ranges, Huia,<br />

on Geniostoma ligustrifolium leaf endophyte, coll. S. Bellgard M128, 8 Jul. 2010 (ICMP<br />

18540); Auckland, Waitakere Ranges, Huia, on Coprosma sp. rotten berry, coll. S.<br />

Bellgard M130-2, 8 Jul. 2010 (ICMP 18541); Auckland, Waiheke Island, Palm Beach,<br />

on Meryta sinclairii leaf spot, coll. P.R. Johnston C1310.1, 21 Mar. 2010 (PDD 99186;<br />

ICMP 18742); Auckland, Tiritiri Island, on Dysoxylum spectabile fruit rot, coll. P.R.<br />

141


Weir et al.<br />

Fig. 15. <strong>Colletotrichum</strong> aotearoa. A. ICMP 17324. B. ICMP 18529. C. ICMP 18548. D. 18532. E. ICMP 18540. A–C. Appressoria. D. Conidiogenous cells. E. Conidia. Scale bar<br />

A = 20 µm. Scale bar of A applies to A–E.<br />

Johnston C1220, 12 Feb. 1997 (PDD 67042; ICMP 18740); Northland, Whangaruru,<br />

on Vitex lucens fruit rot, coll. P.R. Johnston C880.1, L. Brako, P. Berry, 28 Jan. 1988<br />

(PDD 48408; ICMP 18530); on Berberis sp. leaf spot, coll. N. Waipara C77 (ICMP<br />

18735), on Lonicera japonica leaf spot, coll. N. Waipara J3 (ICMP 18736); Wellington,<br />

Waikanae, on Coprosma sp. leaf, coll. B. Weir C1285, 14 May 2009 (ICMP 18548);<br />

Auckland, Wenderholm Regional Park, on Melicytus ramiflorus leaf endophyte, coll.<br />

G.C. Carroll MELRA, 16 Sep. 2009 (ICMP 18543).<br />

* <strong>Colletotrichum</strong> asianum Prihastuti, L. Cai & K.D. Hyde,<br />

Fungal Diversity 39: 96. 2009. Fig. 17.<br />

Prihastuti et al. (2009) provide a description of this <strong>species</strong>.<br />

Geographic distribution and host range: Known on Mangifera<br />

indica from Australia, Colombia, Japan, Panama, Philippines, and<br />

Thailand; also reported on Coffea arabica from Thailand.<br />

142<br />

Genetic identification: <strong>Colletotrichum</strong> asianum is distinguished from<br />

all other taxa using any of the genes tested, including ITS.<br />

Notes: Although the type specimen is from coffee, this fungus is isolated<br />

commonly from mango (Mangifera indica) (e.g. Morphological Group<br />

1 from Than et al. 2008; IMI 313839 from Australia; MAFF 306627<br />

from Japan). Isolates referred to <strong>Colletotrichum</strong> indet. sp. 1 by Rojas<br />

et al. (2010), also associated with mango fruit rots, again match C.<br />

asianum. Based on ITS sequences, isolates Man-63 and Man-69<br />

cited by Afanador-Kafuri et al. (2003) from mango from Colombia, are<br />

probably also C. asianum. Several papers have reported genetically<br />

uniform populations of C. <strong>gloeosporioides</strong> associated with M. indica<br />

around the world (e.g. Hodson et al. 1993, Alahakoon et al. 1994,<br />

Sanders & Korsten 2003) and these perhaps also represent C.<br />

asianum, although DNA sequences are not available to confirm this.


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 16. <strong>Colletotrichum</strong> aotearoa. A. ICMP 18537 – ex-holotype culture. B. ICMP 18548. C. ICMP 18532. D. ICMP 18740. E. ICMP 18533. F. ICMP 18530. A–F. Cultures on<br />

PDA, 10 d growth from single conidia, from above and below.<br />

Three earlier <strong>species</strong>, originally described from leaves rather<br />

than fruit of Mangifera, may provide earlier names for C. asianum<br />

but type material for these <strong>species</strong> has not been examined in this<br />

study; C. mangiferae Kelkar, Gloeosporium mangiferae Henn. 1898,<br />

and G. mangiferae Racib. 1900. As with most substrates, several<br />

different <strong>species</strong> of <strong>Colletotrichum</strong> often occur on the same host.<br />

143


Weir et al.<br />

For example, Damm et al. (2012a, b, this issue) report members<br />

of the C. acutatum and C. boninense <strong>species</strong> <strong>complex</strong>es, C.<br />

simmondsii, C. fioriniae, and C. karstii, from mango from Australia.<br />

Isolates from Capsicum reported by Than et al. (2008) as C.<br />

<strong>gloeosporioides</strong> Morphological Group 2 (e.g. isolates Ku4 = ICMP<br />

144<br />

Fig. 17. <strong>Colletotrichum</strong> asianum. A. ICMP 18648 (ex<br />

<strong>CBS</strong> 124960). B. ICMP 18580 (ex MFLU 090234). C.<br />

ICMP 18603 (ex MAFF 306627). D. ICMP 18604 (ex<br />

HKUCC 18602). E. ICMP 18696 (ex IMI 313839). A–E.<br />

Cultures on PDA, 10 d growth from single conidia, from<br />

above and below.<br />

18575 and Ku8 = ICMP 18618), were referred to as C. asianum by<br />

Hyde et al. (2009), however they are genetically distinct from C.<br />

asianum and belong to C. siamense based on our analyses.<br />

<strong>The</strong> C. asianum protologue designates the holotype as MFLU<br />

090234, and the culture derived from the holotype as “BCC” with


no strain number. <strong>The</strong> ex-holotype culture is listed as BDP-I4 in the<br />

Prihastuti et al. (2009) Table 1, but this number is not mentioned<br />

in the description. Culture BDP-I4 was obtained from the authors<br />

(Prihastuti et al. 2009) for this study.<br />

Specimens examined: Australia, New South Wales, Sextonville, on Mangifera<br />

indica, 1987 (IMI 313839 = ICMP 18696). Philippines, on Mangifera indica (MAFF<br />

306627 = ICMP 18603). Thailand, Chiang Mai, on Mangifera indica fruit, coll. P.P.<br />

Than M3 (HKUCC 10862 = ICMP 18605); Chiang Mai, on Mangifera indica fruit,<br />

coll. P.P. Than M4 (HKUCC 10863 = ICMP 18604); Mae Lod Village, Mae Taeng<br />

District, Chiang Mai, on Coffea arabica berries, coll. H. Prihastuti BPD-I4, 16 Jan.<br />

2008 (ex-holotype culture of C. asianum from specimen MFLU 090234 = ICMP<br />

18580 = <strong>CBS</strong> 130418). Panama, Gamboa, on Mangifera indica fruit rot, coll. S. Van<br />

Bael GJS 08-144, Jul 2008 (<strong>CBS</strong> 124960 = ICMP 18648).<br />

<strong>Colletotrichum</strong> boehmeriae Sawada, Hakubutsu Gakkwai<br />

Kwaihô (Trans. Nat. Hist. Soc. Formosa) 17: 2. 1914.<br />

Notes: Sawada (1922) provided an English translation of his original<br />

description. This <strong>species</strong> was described as a stem pathogen of<br />

Boehmeria nivea, and remains in use in this sense (e.g. Li & Ma<br />

1993). Wang et al. (2010) cite several GenBank accessions from<br />

isolates they identify as C. <strong>gloeosporioides</strong> that cause severe<br />

disease of Boehmeria. Based on a comparison of the GenBank data<br />

with our ITS gene tree, these and other isolates from the same host<br />

deposited by the same authors (GQ120479–GQ120499), appear<br />

to represent three different taxa within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> — C. <strong>gloeosporioides</strong> s. str., C. aotearoa, and C. fructicola.<br />

Isolates representative of all three taxa are reportedly pathogenic<br />

on Boehmeria (Wang et al. 2010). <strong>The</strong> genetic relationship of these<br />

fungi needs to be confirmed using additional genes.<br />

<strong>Colletotrichum</strong> camelliae Massee, Bull. Misc. Inform. Kew.<br />

1899: 91. 1899.<br />

Notes: <strong>Colletotrichum</strong> camelliae was described by Massee (in<br />

Willis 1899) from the living leaves of tea (Camellia sinensis) from<br />

Sri Lanka. It was placed in synonymy with C. <strong>gloeosporioides</strong> by<br />

von Arx (1957). Although not listed by Hyde et al. (2009), the name<br />

is widely used in the trade and semi-popular literature as the causal<br />

agent of the brown blight disease of tea (e.g. Sosa de Castro et al.<br />

2001, Muraleedharan & Baby 2007).<br />

We have been unable to sample <strong>Colletotrichum</strong> isolates<br />

from tea with typical brown blight symptoms. <strong>The</strong>re are four<br />

GenBank accessions of <strong>Colletotrichum</strong> from tea, two from China<br />

(EU732732, FJ515007), one from Japan (AB218993), and<br />

another from Iran (AB548281), referred variously to C. camelliae,<br />

C. crassipes and C. <strong>gloeosporioides</strong>. Although ITS sequences<br />

only are available for these geographically widespread isolates,<br />

the DNA sequence of the Iranian isolate appears to match C.<br />

<strong>gloeosporioides</strong> s. str., while those from the other three isolates<br />

are all very similar to each other. <strong>The</strong> ITS sequence from these<br />

isolates matches that of <strong>CBS</strong> 232.79, from tea shoots from Java<br />

(GenBank JX009429). GAPDH and ITS sequences from <strong>CBS</strong><br />

232.79 (GenBank JX009417, JX009429) place this isolate in C.<br />

fructicola. Note that <strong>CBS</strong> 571.88, isolated from tea from China<br />

and deposited as Glomerella cingulata, is a <strong>Colletotrichum</strong> sp.<br />

outside C. <strong>gloeosporioides</strong> s. lat., based on ITS sequences<br />

(GenBank JX009424).<br />

We tested the pathogenicity of <strong>CBS</strong> 232.79 and isolates of G.<br />

cingulata “f. sp. camelliae” (see below) using detached tea leaves<br />

and found that only the G. cingulata “f. sp. camelliae” isolates were<br />

strong pathogens (unpubl. data).<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

<strong>The</strong> genetic relationship between the pathogen of ornamental<br />

Camellia (here referred to G. cingulata “f. sp. camelliae”), isolates<br />

from tea with DNA sequence data in GenBank, and isolates<br />

associated with brown blight symptoms of tea remain unresolved.<br />

Additional isolates with known pathogenicity, collected from typical<br />

brown blight symptoms from the field, are required to determine<br />

whether or not there are two distinct pathogens of Camellia, one of<br />

tea, the other of ornamental varieties.<br />

Other <strong>Colletotrichum</strong> <strong>species</strong> reported from tea include C.<br />

“theae-sinensis”, an invalid recombination of Gloeosporium theaesinensis<br />

I. Miyake, proposed by Yamamoto (1960). Moriwaki and<br />

Sato (2009) summarised the taxonomic history of this name and<br />

transferred G. theae-sinensis to Discula on the basis of DNA<br />

sequences. Sphaerella camelliae Cooke and its recombination<br />

Laestadia camelliae (Cooke) Berl. & Voglino were listed by von Arx<br />

& Müller (1954) as synonyms of Glomerella cingulata. This <strong>species</strong><br />

is now accepted as Guignardia camelliae (Cooke) E.J. Butler ex<br />

Petch and is regarded as the causal agent of copper blight disease<br />

of tea (Spaulding 1958).<br />

Thang (2008) placed C. camelliae in synonymy with C.<br />

coccodes, presumably on the basis of the Species Fungorum<br />

synonymy (www.<strong>species</strong>fungorum.org, website viewed 6 Oct<br />

2010). Thang (2008) questioned the synonymy, noting differences<br />

between the descriptions of the two <strong>species</strong> provided by Massee<br />

(in Willis 1899) and Sutton (1980) respectively.<br />

<strong>Colletotrichum</strong> caricae F. Stevens & J.G. Hall, Z.<br />

Pflanzenkrankh., 19: 68. 1909.<br />

Notes: Placed in synonymy with C. <strong>gloeosporioides</strong> by von Arx<br />

(1957), C. caricae was listed as a separate <strong>species</strong> by Sutton (1992).<br />

It was described from fruits and leaves of Ficus carica from the USA<br />

(Stevens & Hall 1909) but is poorly understood both morphologically<br />

and biologically. Its genetic relationship to and within the C.<br />

<strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>, and to other Ficus-associated<br />

<strong>species</strong> such as <strong>Colletotrichum</strong> ficus Koord. and Glomerella cingulata<br />

var. minor (here placed in synonymy with C. fructicola) is unknown.<br />

Glomerella cingulata (Stonem.) Spauld. & H. Schrenk,<br />

Science, n.s. 17: 751. 1903.<br />

Basionym: Gnomoniopsis cingulata Stonem., Bot. Gaz. 26: 101.<br />

1898.<br />

= Gloeosporium cingulatum G.F. Atk., Bull. Cornell Univ. Agric. Exp. Sta. 49:<br />

306. 1892. [fide Stoneman 1898]<br />

Notes: Stoneman (1898) described Glomerella cingulata from<br />

diseased stems of Ligustrum vulgare from the USA and reported the<br />

development of perithecia in cultures initiated from conidia of what<br />

she considered its asexual morph, Gloeosporium cingulatum. <strong>The</strong>re<br />

are recent reports of anthracnose diseases of Ligustrum (e.g. Alfieri<br />

et al. 1984, Vajna & Bagyinka 2002) but the relationship of isolates<br />

causing this disease to the C. <strong>gloeosporioides</strong> <strong>complex</strong> is not known.<br />

Glomerella cingulata is often linked taxonomically to the<br />

anamorph <strong>Colletotrichum</strong> <strong>gloeosporioides</strong>, and the name has in the<br />

past been applied in an equally broad sense to C. <strong>gloeosporioides</strong><br />

s. lat. (e.g. Small 1926, von Arx & Müller 1954). However, it is<br />

unlikely that the type specimen of G. cingulata represents the<br />

same <strong>species</strong> as C. <strong>gloeosporioides</strong> s. str. (see notes under C.<br />

<strong>gloeosporioides</strong>). <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> s. str. is not<br />

known to form perithecia in culture, and there are no isolates of<br />

C. <strong>gloeosporioides</strong> s. str. known to us that are associated with a<br />

Glomerella state on diseased stems of Ligustrum, An isolate of C.<br />

145


Weir et al.<br />

Fig. 18. Glomerella cingulata “f. sp. camelliae”. A, C, D. ICMP 10643. B, E. ICMP 10646. A–B. Appressoria. C. Conidiogenous cells. D–E. Conidia. Scale bar A = 20 µm. Scale<br />

bar of A applies to A–E.<br />

aotearoa (ICMP 18748) was isolated from Ligustrum lucidum in<br />

New Zealand, but it was not associated with a stem lesion and no<br />

C. aotearoa isolates were observed forming perithecia.<br />

Glomerella cingulata var. brevispora Wollenw., Z.<br />

Parasitenk. (Berlin) 14: 260. 1949.<br />

Notes: Described from fruit rots from Germany, this name has not<br />

been used since. No cultures are available and its relationship to<br />

and within the C. <strong>gloeosporioides</strong> <strong>complex</strong> is not known.<br />

146<br />

* Glomerella cingulata “f. sp. camelliae” (Dickens & Cook<br />

1989). Figs 18, 19.<br />

Notes: Dickens & Cook (1989) proposed the name Glomerella<br />

cingulata “f. sp. camelliae” for isolates morphologically typical of<br />

C. <strong>gloeosporioides</strong> s. lat. that were highly pathogenic to leaves and<br />

shoots of ornamental Camellia saluenensis hybrids, causing the<br />

disease Camellia twig blight. <strong>The</strong>se authors reported the fungus<br />

from plants imported into the UK from New Zealand and noted that<br />

a similar disease had been reported from plants grown in the UK,


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 19. Glomerella cingulata “f. sp. camelliae”. A. ICMP 18542. B. ICMP 10643. C. ICMP 10646. A–C. Cultures on PDA, 10 d growth from single conidia, from above and below.<br />

USA, Australia, France, and Italy. <strong>The</strong> disease has been reported<br />

from Camellia japonica, C. reticulata, and C. sasanqua. Although<br />

isolated in the UK from plants imported from New Zealand, this<br />

pathogen has not yet been found on Camellia plants growing in<br />

New Zealand.<br />

We have sequenced authentic isolates cited by Dickens &<br />

Cook (1989) as well as isolates pathogenic to Camellia saluenensis<br />

collected from the USA. <strong>The</strong>y are similar to each other genetically<br />

as well as biologically and morphologically. ITS sequences alone<br />

distinguish G. cingulata “f. sp. camelliae” from all other taxa in the<br />

C. <strong>gloeosporioides</strong> <strong>complex</strong>, and there is good genetic evidence to<br />

consider these isolates to be representative of a distinct <strong>species</strong><br />

within the C. kahawae clade. A new <strong>species</strong> is not proposed here<br />

because the relationship between the G. cingulata “f. sp. camelliae”<br />

isolates and C. camelliae, the fungus causing brown blight of tea,<br />

remains uncertain.<br />

Dickens & Cook (1989) also reported two C. acutatum strains<br />

from ornamental Camellia <strong>species</strong> that were avirulent in tests<br />

with detached Camellia cv. Donation leaves. Strain IMI 351261,<br />

deposited 1992 in IMI by R. Cook, is likely to be one of them. This<br />

strain was confirmed as belonging to the C. acutatum <strong>species</strong><br />

<strong>complex</strong> and identified as C. lupini, which causes lupin anthracnose<br />

and is occasionally found on other hosts (Damm et al. 2012a, this<br />

issue). Another strain from Camellia reticulata from China belongs<br />

to C. fioriniae, also a <strong>species</strong> in the C. acutatum <strong>complex</strong>, while<br />

a strain from New Zealand (ICMP 10338) is C. boninense s. str.<br />

(Damm et al. 2012a, b, this issue).<br />

See notes under C. camelliae.<br />

Specimens examined: UK, plants imported from New Zealand, on Camellia ×<br />

williamsii, coll. Dickens & Cook 82/437, 1982 (authentic culture of Glomerella<br />

cingulata “f. sp. camelliae” – ICMP 10643; dried culture PDD 56488). USA,<br />

Mississippi, on Camellia sasanqua twig blight, coll. W.E. Copes CG02g, May 2002<br />

(ICMP 18542); South Carolina, on Camellia sp., coll. G. Laundon 20369, 1 Jan.<br />

1982 (ICMP 10646).<br />

Glomerella cingulata var. crassispora Wollenw., Z.<br />

Parasitenk. (Berlin) 14: 260. 1949.<br />

Notes: Described from Coffea arabica from a glasshouse in<br />

Germany, this name has not been used since. No cultures are<br />

available and its relationship to and within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> is not known.<br />

Glomerella cingulata “f. sp. manihotis” (Chevaugeon<br />

1956)<br />

Notes: See notes under <strong>Colletotrichum</strong> manihotis.<br />

Glomerella cingulata var. minor Wollenw., Z. Parasitenk.<br />

(Berlin) 14: 261. 1949.<br />

= Gloeosporium elasticae Cooke & Massee, Grevillea 18: 74. 1890. [fide<br />

Wollenweber & Hochapfel 1949]<br />

Notes: Placed here in synonymy with C. fructicola.<br />

Glomerella cingulata var. minor was described from Ficus from<br />

Germany, but Wollenweber & Hochapfel (1949) noted that the<br />

same fungus occurred also on other hosts in Europe, Africa, and<br />

America, including Malus and Coffea. Genetically the ex-holotype<br />

culture of G. cingulata var. minor (<strong>CBS</strong> 238.49) matches the type<br />

specimen of C. fructicola, although the culture itself appears to be<br />

stale, with slow growth and an irregularly scalloped margin (see<br />

147


Weir et al.<br />

images under C. fructicola). Wollenweber & Hochapfel (1949) used<br />

the name Gloeosporium elasticae Cooke & Massee for the conidial<br />

state of G. cingulata var. minor, the type specimens for both names<br />

being from Ficus.<br />

See also notes under C. queenslandicum.<br />

Specimen examined: Germany, Berlin-Dahlem, from Ficus edulis leaf spot, May<br />

1936 (ex-holotype culture of G. cingulata var. minor – <strong>CBS</strong> 238.49 = ICMP 17921).<br />

Glomerella cingulata var. migrans Wollenw., Z. Parasitenk.<br />

(Berlin) 14: 262. 1949.<br />

Notes: Placed here in synonymy with C. kahawae subsp. ciggaro,<br />

see notes under this <strong>species</strong>.<br />

Specimen examined: Germany, Berlin-Dahlem, on stem of Hypericum perforatum,<br />

Jun. 1937 (ex-holotype culture of Glomerella cingulata var. migrans – <strong>CBS</strong> 237.49<br />

= ICMP 17922).<br />

Glomerella cingulata “var. orbiculare” Jenkins & Winstead,<br />

Phytopathology 52: 15. 1962.<br />

Notes: Listed in Index Fungorum, this name was mentioned<br />

in an abstract, but is invalid (no Latin description) and never<br />

formally published. It was being used to refer to the teleomorph of<br />

<strong>Colletotrichum</strong> orbiculare, not part of the C. <strong>gloeosporioides</strong> <strong>complex</strong><br />

(Cannon et al. 2012, this issue). Glomerella lagenaria (Pass.)<br />

F. Stevens, a recombination of the anamorphic name Fusarium<br />

lagenarium Pass., has also been used to refer to this teleomorph.<br />

Correll et al. (1993) comment on the pathogenicity of cucurbitassociated<br />

strains that form a Glomerella state in culture, suggesting<br />

a degree of confusion around the application of these names.<br />

Glomerella cingulata “f. sp. phaseoli” (Kimati & Galli 1970).<br />

Notes: Both G. cingulata “f. sp. phaseoli” (e.g. Castro et al. 2006)<br />

and Glomerella lindemuthiana (e.g. Rodríguez-Guerra et al. 2005,<br />

as G. lindemuthianum) have been used for the teleomorph of<br />

<strong>Colletotrichum</strong> lindemuthianum in the recent literature, the two<br />

names placed in synonymy by Sutton (1992). This fungus is not part<br />

of the C. <strong>gloeosporioides</strong> <strong>complex</strong> (Cannon et al. 2012, this issue).<br />

Glomerella cingulata var. sorghicola Saccas, Agron. Trop.<br />

(Maracay). 9: 171. 1954.<br />

Notes: Not a member of the C. <strong>gloeosporioides</strong> <strong>complex</strong>. Sutton<br />

(1992) suggested using this name to refer to the teleomorph of<br />

<strong>Colletotrichum</strong> sublineola, although Crouch et al. (2006) note that<br />

C. sublineola has no known teleomorph.<br />

* <strong>Colletotrichum</strong> clidemiae B. Weir & P.R. Johnst. sp. nov.<br />

MycoBank MB563592. Figs 20, 21.<br />

= <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. clidemiae” (Trujillo et al. 1986).<br />

Etymology: Based on the host reportedly susceptible to this <strong>species</strong>.<br />

Holotype: USA, Hawai’i, Aiea, on Clidemia hirta leaf spot, coll. S.A.<br />

Ferreira & K. Pitz, 14 May 2010, PDD 101997; ex-type culture<br />

ICMP 18658.<br />

Colonies grown from single conidia on Difco PDA 25 mm diam after<br />

10 d, aerial mycelium grey, cottony, sparse, surface of colony with<br />

148<br />

numerous small, dark-based acervuli with deep orange conidial<br />

ooze and scattered setae, in reverse more or less colourless except<br />

for the acervuli and masses of conidial ooze showing through. After<br />

18 d numerous globose, pale walled protoperithecia developing<br />

near centre of colony. Conidia (16−)18−20(−26.5) × (4.5−)5.5−6<br />

µm (av. 19.3 × 5.5 µm, n = 48), broad-cylindric, ends broadly<br />

rounded, longer conidia sometimes tapering slightly towards the<br />

base. Appressoria variable in shape, some simple, subglobose,<br />

but often with a small number of broad, irregular lobes. Perithecia<br />

mature after about 21 d, dark-walled, about 200–250 µm diam with<br />

short ostiolar neck, perithecial wall of 3–4 layers of angular cells<br />

10–15 µm diam with walls thin, pale brown to brown. Asci 8-spored<br />

60–67 × 10–14 µm. Ascospores (14–)15.5–19(–21.5) × 4.5–5.5(–<br />

6.5) µm (av. 17.2 × 5.0 µm, n = 46), oblong-elliptic, tapering to<br />

rounded ends, widest point toward one end, in side view flat on one<br />

side, rarely curved and if so, then slightly.<br />

Geographic distribution and host range: First reported from<br />

Clidemia, native to Panama, and subsequently introduced to<br />

Hawai’i as a pathogen of that host. Genetically matching isolates<br />

occur on native Vitis and Quercus spp. in Florida (see notes below).<br />

Genetic identification: ITS sequences do not separate C. clidemiae<br />

from C. aotearoa. <strong>The</strong> two <strong>species</strong> are best distinguished using<br />

ACT, GAPDH, or GS.<br />

Notes: Isolates referred to C. <strong>gloeosporioides</strong> “f. sp. clidemiae” by<br />

Trujillo et al. (1986) were highly pathogenic to Clidemia, but not to<br />

the other <strong>species</strong> of Melastomataceae tested. No voucher cultures<br />

of the original isolates collected from Panama were kept, but<br />

recent specimens isolated from naturalised Clidemia hirta plants in<br />

Hawai’i with typical disease symptoms are genetically uniform and<br />

distinct within the Kahawae clade. Phylogenetic, biological, and<br />

morphological evidence support this fungus being described as a<br />

new <strong>species</strong> within the C. <strong>gloeosporioides</strong> <strong>complex</strong>.<br />

A fungus isolated from a Vitis sp. in Florida and referred to as<br />

“Glomerella cingulata native host” by MacKenzie et al. (2007), is<br />

genetically close to our isolates from Clidemia and is here referred<br />

to the same <strong>species</strong>. Data in MacKenzie et al. (2007) shows the<br />

same fungus occurs on both Vitis and Quercus in Florida. Micromorphologically<br />

the isolates from Clidemia and from Vitis that<br />

we examined are similar with respect to the size and shape of<br />

appressoria, conidia, and ascospores. <strong>The</strong>y are distinct in cultural<br />

appearance, the cultures of the Vitis-associated fungus having<br />

aerial mycelium darker and more dense, and a faster growth rate.<br />

Similar variation in cultural appearance is present in several of the<br />

phylogenetically defined <strong>species</strong> that we recognise. Whether or<br />

not the Clidemia-associated isolates are biologically distinct from<br />

the Vitis- and Quercus-associated isolates from Florida requires<br />

pathogenicity tests to determine.<br />

Other specimens examined: USA, Florida, Sarasota, on Vitis sp. leaf, coll. S.<br />

MacKenzie SS-Grape-12, 2002 (ICMP 18706); Hawai’i, Aiea, on Clidemia hirta leaf<br />

spot, coll. S.A. Ferreira & K. Pitz, 14 May 2010 (ICMP 18659, ICMP 18660, ICMP<br />

18661, ICMP 18662, ICMP 18663).<br />

<strong>Colletotrichum</strong> coffeanum F. Noak, Z. Pflanzenkrankh. 11:<br />

202. 1901.<br />

Notes: Waller et al. (1993) discussed the use of the names<br />

<strong>Colletotrichum</strong> coffeanum and Gloeosporium coffeanum Delacr.<br />

and the geographic and biological differences between these


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 20. <strong>Colletotrichum</strong> clidemiae. A, B, E. ICMP 18658 – ex-holotype culture. C, D. ICMP 18706. A, D. Appressoria. B, C. Asci and ascospores. E. Conidia. Scale bar C = 20<br />

µm. Scale bar of C applies to A–E.<br />

<strong>species</strong> and the pathogen of coffee berries, C. kahawae. Both C.<br />

coffeanum and G. coffeanum were described from leaves of coffee,<br />

the two <strong>species</strong> distinguished by Noak (1901) by the presence<br />

or absence of setae in the acervuli. <strong>The</strong>re is a wide range of C.<br />

<strong>gloeosporioides</strong>-like <strong>species</strong> on coffee plants (see Waller et al.<br />

1993 and notes under C. kahawae) and the relationships of C.<br />

coffeanum and G. coffeanum within the C. <strong>gloeosporioides</strong> <strong>species</strong><br />

<strong>complex</strong> remain uncertain.<br />

149


Weir et al.<br />

Fig. 21. <strong>Colletotrichum</strong> clidemiae. A. ICMP 18658 – ex-holotype culture. ICMP<br />

18706. Cultures on PDA, 10 d growth from single conidia, from above and below.<br />

<strong>Colletotrichum</strong> cordylines Pollacci, Atti Ist. Bot. Univ.<br />

Pavia, Serie 2, 5: 44. 1899.<br />

Notes: Described from leaves of Cordyline indivisa from a botanical<br />

garden in Italy, the genetic and biological status of this <strong>species</strong> is not<br />

known. Two Cordyline-associated <strong>species</strong> are accepted in this study,<br />

C. cordylinicola from Thailand and the newly described C. ti from New<br />

Zealand. <strong>The</strong> original description of C. cordylines is brief (Pollacci<br />

1899) but it specifically mentions setae more than 100 µm long.<br />

<strong>Colletotrichum</strong> cordylinicola is described as lacking setae (Phoulivong<br />

et al. 2011) and in C. ti they are rare and when present much less than<br />

100 µm long. <strong>The</strong> phylogenetic significance of this apparent difference<br />

and confirmation that these names represent different fungi requires<br />

DNA sequences to be generated from type material of C. cordylines.<br />

* <strong>Colletotrichum</strong> cordylinicola Phoulivong, L. Cai & K.D.<br />

Hyde, Mycotaxon 114: 251. 2011 [“2010”]. Fig. 22.<br />

Phoulivong et al. (2011) provide a description.<br />

Geographic distribution and host range: Known only from Cordyline<br />

from Thailand and Eugenia from Laos.<br />

Genetic identification: ITS sequences separate C. cordylinicola<br />

from all other <strong>species</strong>.<br />

Notes: Phoulivong et al. (2011) report C. cordylinicola from<br />

Cordyline (Agavaceae) and Eugenia (Myrtaceae). <strong>The</strong>y noted<br />

that the isolate from Eugenia was not pathogenic to Cordyline and<br />

vice versa, and they also showed that the specimens from the<br />

two hosts are genetically somewhat distinct, although forming a<br />

sister relationship amongst the taxa included in their analysis. <strong>The</strong><br />

calmodulin gene tree generated from our sequence data together<br />

with the sequences provided by Phoulivong et al. (2011) (GenBank<br />

accession HM470236) supports placing the isolates from Eugenia<br />

and from Cordyline in the same <strong>species</strong> (unpubl. data).<br />

<strong>Colletotrichum</strong> cordylinicola is genetically distinct from a<br />

<strong>species</strong> associated with Cordyline leaf spots from New Zealand,<br />

described here as C. ti. See also notes under C. cordylines.<br />

150<br />

Specimen examined: Thailand, Chiang Mai, Sam Sai District, Maejo Village, on<br />

Cordyline fruticosa, coll. S. Phoulivong, 15 Mar. 2009 (ex-holotype culture –<br />

MFLUCC 090551 = ICMP 18579). Note that the ex-holotype culture was mistakenly<br />

cited as MFUCC 090551 by Phoulivong et al. (2011).<br />

<strong>Colletotrichum</strong> crassipes (Speg.) Arx, Verh. Kon. Ned.<br />

Akad. Wetensch., Afd. Natuurk., Sect. 2, 51(3): 77. 1957.<br />

Basionym: Gloeosporium crassipes Speg., Rivista Vitic. Enol. 2:<br />

405. 1878.<br />

Notes: Several isolates identified as <strong>Colletotrichum</strong> crassipes<br />

that have sequences accessioned to GenBank belong in C.<br />

<strong>gloeosporioides</strong> s. lat. GenBank accessions identified as C.<br />

crassipes that have a publically available culture include C.<br />

kahawae subsp. ciggaro (STE-U 5302 = <strong>CBS</strong> 112988 – AY376529,<br />

AY376577, FN557348, FN557538, and FN599821; STE-U 4445<br />

= <strong>CBS</strong> 112984 – AY376530, AY376578, – FN557347, FN557537,<br />

and FN599820), along with several other <strong>species</strong> outside of<br />

the C. <strong>gloeosporioides</strong> <strong>complex</strong> (<strong>CBS</strong> 169.59 = IMI 309371 –<br />

AJ536230, FN557344, and FN599817; <strong>CBS</strong> 159.75 – FN557345<br />

and FN599818; <strong>CBS</strong> 109355 – FN557346 and FN599819).<br />

Those with no isolates in a public collection include C. kahawae<br />

subsp. ciggaro (CORCS3 cited in Yang et al. (2011), HM584410,<br />

HM582002, HM585412), C. fructicola (strain 080912009 Jining,<br />

unpubl. data, FJ515007), and a possibly undescribed <strong>species</strong><br />

within the Kahawae clade (strain SYJM02, unpubl. data,<br />

JF923835). Originally described from the berries of Vitis vinifera<br />

from Italy (Spegazzini 1878), the identity of C. crassipes remains<br />

unresolved. <strong>The</strong>re is confusion regarding its morphology. Von Arx<br />

(1970) uses the name C. crassipes for fungi in which setae are<br />

rare, conidia are 22–34 × 6–8 µm (more or less matching the<br />

original description), and the lobed appressoria are distinctively<br />

globose in shape. Sutton (1980) uses a different morphological<br />

concept – setae common (according to Sutton these are rare in the<br />

otherwise morphologically similar C. musae), conidia 10–15 × 4.5–<br />

6.5 µm (Sutton’s concept of C. <strong>gloeosporioides</strong> is characterised by<br />

narrower conidia), and the appressoria deeply lobed. <strong>The</strong> conidial<br />

width cited for C. <strong>gloeosporioides</strong> by Sutton (1980), 3–4.5 µm, is<br />

narrower than we have found for all the taxa we accept within C.<br />

<strong>gloeosporioides</strong> s. lat., whereas his C. crassipes measurement of<br />

4.5–6.5 µm matches many of the taxa we recognise. Several of<br />

these taxa also have deeply lobed appressoria.<br />

<strong>Colletotrichum</strong> dracaenae Allesch., Rabenhorst’s<br />

Kryptogamen-Flora von Deutschland, Oesterreich und der<br />

Schweiz, Ed. 2, 1(7): 560. 1902.<br />

Notes: Farr et al. (2006) examined the type specimen of this<br />

<strong>species</strong> and concluded it was a member of C. <strong>gloeosporioides</strong> s.<br />

lat., based on conidial size and shape. Genetic data is not available<br />

to confirm this. See also discussion under C. petchii in Damm et al.<br />

(2012b, this issue)<br />

<strong>Colletotrichum</strong> fragariae A.N. Brooks, Phytopathology 21:<br />

113. 1931.<br />

Notes: Placed here in synonymy with C. theobromicola. See notes<br />

and additional specimens examined under C. theobromicola.<br />

<strong>The</strong> name C. fragariae was originally applied to isolates<br />

associated with a disease of strawberry (Fragaria × ananassa)<br />

runners (stolons) and petioles in Florida (Brooks 1931). Although<br />

the name was placed in synonymy with C. <strong>gloeosporioides</strong> by


von Arx (1957), it has continued to be used in the literature for<br />

strawberry-associated <strong>Colletotrichum</strong> isolates. It was accepted as<br />

distinct by Sutton (1992), although he noted confusion surrounding<br />

application of the name. Designation of one of Brook’s cultures<br />

(<strong>CBS</strong> 142.31 = IMI 346325) as the epitype of C. fragariae by Buddie<br />

et al. (1999) has allowed a modern, genetic basis for this name to<br />

be fixed. <strong>The</strong> ex-epitype culture of C. fragariae sits in a strongly<br />

supported clade containing isolates from a wide range of hosts<br />

from many parts of the world, including the ex-epitype culture of C.<br />

theobromicola, an earlier name for C. fragariae in the sense that we<br />

accept these <strong>species</strong> in this paper.<br />

<strong>The</strong>re are several <strong>species</strong> from the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> which inhabit diseased strawberry plants, and as shown<br />

by MacKenzie et al. (2007, 2008) isolates that genetically match<br />

the epitype of C. fragariae have a wide host range. Despite its<br />

name MacKenzie et al. (2007, 2008) regarded this fungus as<br />

simply one of a group of several <strong>species</strong> sometimes found on<br />

strawberry. Our study confirms that members of the C. fragariae/<br />

theobromicola clade occur throughout the world on a wide range of<br />

hosts. Within the diversity of the C. fragariae/theobromicola clade,<br />

there is a subclade consisting of the C. fragariae epitype and two<br />

contemporary ex-strawberry isolates from the USA (Fig. 1), further<br />

work will be needed to establish if the strawberry stolon disease is<br />

restricted to this subclade. Despite regular surveys this disease has<br />

not been found on strawberries in New Zealand.<br />

Xie et al. (2010b) provides a good example of the confusion<br />

that continues to surround the application of <strong>Colletotrichum</strong> names<br />

to isolates from strawberry. <strong>The</strong>se authors noted that putative C.<br />

<strong>gloeosporioides</strong> and C. fragariae isolates were difficult to distinguish<br />

using ITS sequences, the only sequences that they generated.<br />

Xie et al. (2010b) found 4 groups of isolates, each with a slightly<br />

different ITS sequence, two of those groups they considered to<br />

be C. fragariae and two to be C. <strong>gloeosporioides</strong>. To classify their<br />

isolates as either C. fragariae of C. <strong>gloeosporioides</strong> they used a<br />

restriction enzyme method based on Martinez-Culebras et al.<br />

(2000). Incorporating their ITS sequences into our ITS alignment,<br />

one of their groups genetically matches C. tropicale, one matches<br />

C. <strong>gloeosporioides</strong> s. str., one matches C. fructicola, and one<br />

matches C. siamense. <strong>The</strong>se relationships are based on ITS<br />

sequences only — the genetic differences between some of these<br />

<strong>species</strong> are small and are indicative only of possible relationships.<br />

However, it is clear that none of the Xie et al. (2010b) sequences<br />

match those of the epitype of C. fragariae. <strong>The</strong>re are also several<br />

<strong>species</strong> within the C. acutatum <strong>species</strong> <strong>complex</strong> associated with<br />

Fragaria (Damm et al. 2012, this issue).<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 22. <strong>Colletotrichum</strong> cordylinicola. ICMP 18579 (ex<br />

MFLUCC 090551 – ex-holotype culture). A. Cultures on<br />

PDA, 10 d growth from single conidia, from above and<br />

below.<br />

Specimen examined: USA, Florida, on Fragaria × ananassa, coll. A.N. Brooks, 1931<br />

(ex-epitype culture – <strong>CBS</strong> 142.31 = ICMP 17927).<br />

* <strong>Colletotrichum</strong> fructicola Prihastuti, L. Cai & K.D. Hyde,<br />

Fungal Diversity 39: 158. 2009. Fig. 23.<br />

= <strong>Colletotrichum</strong> ignotum E.I. Rojas, S. A. Rehner & Samuels, Mycologia 102:<br />

1331. 2010.<br />

= Glomerella cingulata var. minor Wollenw., Z. Parasitenk. (Berlin) 14: 261.<br />

1949.<br />

Prihastuti et al. (2009) and Rojas et al. (2010) provide descriptions.<br />

Geographic distribution and host range: Originally reported<br />

from coffee berries from Thailand (as C. fructicola) and as a<br />

leaf endophyte from several plants in Central America (as C.<br />

ignotum), isolates that we accept as C. fructicola are biologically<br />

and geographically diverse. Known from Coffea from Thailand,<br />

Pyrus pyrifolia from Japan, Limonium from Israel, Malus domestica<br />

and Fragaria × ananassa from the USA, Persea americana from<br />

Australia, Ficus from Germany, Malus domestica from Brazil,<br />

Dioscorea from Nigeria, and <strong>The</strong>obroma and Tetragastris from<br />

Panama.<br />

Genetic identification: ITS sequences do not separate C. fructicola<br />

from C. aeschynomenes and some C. siamense isolates. <strong>The</strong>se<br />

taxa are best distinguished using GS or SOD2.<br />

Notes: Rojas et al. (2010) noted the occurrence of two distinct<br />

haplotype subgroups (A4-3 and A5-4) within their concept of C.<br />

ignotum. Our genetic analyses resolve the two clades representative<br />

of these two subgroups. However, together they are monophyletic<br />

within the Musae clade of the C. <strong>gloeosporioides</strong> <strong>complex</strong>, and we<br />

retain them here as a single <strong>species</strong>. Both clades include isolates<br />

from a wide range of hosts from many countries, and both are<br />

similar in morphology and cultural appearance. <strong>The</strong> types of both<br />

C. fructicola and C. ignotum are in the same haplotype subgroup.<br />

<strong>The</strong> C. fructicola protologue designates the holotype as MFLU<br />

090228, but the culture derived from holotype as “BCC” with no<br />

specimen number. <strong>The</strong> ex-holotype culture is listed as BDP-I16 in<br />

Table 1 of Prihastuti et al. (2009) but this number is not mentioned<br />

in the description. Culture BDP-I16 was obtained from the authors<br />

(Prihastuti et al. 2009) for this study and deposited as ICMP 18581.<br />

See also notes under G. cingulata var. minor.<br />

Specimens examined: Australia, Queensland, Bli-Bli, on Persea americana fruit<br />

rot, coll. L. Coates 24154 (ICMP 12568). Brazil, Rio Grande do Sul State, on Malus<br />

domestica leaf, coll. T. Sutton BR 8 2001, Jan. 2001 (ICMP 17787); Santa Catarina<br />

State, on Malus domestica leaf, coll. T. Sutton BR 21 2001, Jan. 2001 (ICMP 17788).<br />

151


Weir et al.<br />

Fig. 23. <strong>Colletotrichum</strong> fructicola. A. ICMP 12568. B. ICMP 18615. C. ICMP 18581 (ex MFLU 090228 – ex-holotype culture of C. fructicola). D. ICMP 18610. E. ICMP 18646<br />

(ex <strong>CBS</strong> 125379 – ex-holotype culture of C. ignotum). F. ICMP 17921 (ex <strong>CBS</strong> 238.49 – ex-holotype culture of G. cingulata var. minor). A–F. Cultures on PDA, 10 d growth from<br />

single conidia, from above and below.<br />

Canada, Ontario, on Fragaria × ananassa, Jan. 1991 (IMI 345051 = ICMP 17819).<br />

Germany, Berlin-Dahlem Botanical Garden, on Ficus edulis leaf spot, (ex-holotype<br />

culture of Glomerella cingulata var. minor – <strong>CBS</strong> 238.49 = ICMP 17921). Indonesia,<br />

152<br />

Java, Bandung, Pangheotan Estate, on Camellia sinensis shoots, coll. H. Semangun,<br />

Apr. 1979 (<strong>CBS</strong> 232.79 = ICMP 18656). Israel, on Limonium sinuatum leaf lesion, coll.<br />

S. Freeman L32 (cited in Moriwaki et al. 2006) (ICMP 18613); on Limonium sp. leaf


lesion, coll. S. Freeman L50 (cited in Maymon et al. 2006) (ICMP 18698); on Limonium<br />

sp. leaf lesion, coll. S. Freeman Cg2 (cited in Maymon et al. 2006) (ICMP 18667);<br />

on Limonium sinuatum, coll. S. Freeman L11 (cited in Maymon et al. 2006) (ICMP<br />

18615). Japan, Saga, on Pyrus pyrifolia, coll. H. Ishii sA02-5-1 (cited in Chung et al.<br />

2006) (ICMP 18610). Nigeria, Ibadan, on Dioscorea alata leaf spot, M. Abang Cg13<br />

(cited in Abang et al. 2002) (ICMP 18125); Ilesha, Dioscorea rotundata leaf spots, coll.<br />

M. Abang Cg22 (cited in Abang et al. 2002) (ICMP 18120). Panama, Barro Colorado<br />

Monument, Tetragastris panamensis leaf endophyte, coll. E.I. Rojas E886, 1 Jun. 2004<br />

(ex-holotype culture of C. ignotum − <strong>CBS</strong> 125397 = ICMP 18646); <strong>The</strong>obroma cacao<br />

leaf endophyte, coll. E. Rojas E183 (<strong>CBS</strong> 125395 = ICMP 18645). Thailand, Chiang<br />

Mai, Pa Daeng Village, on Coffea arabica berry, coll. H. Prihastuti BPD-I16, 12 Dec.<br />

2007 (ex-holotype culture of C. fructicola, from specimen MFLU 090228 – ICMP<br />

18581 = <strong>CBS</strong> 130416). USA, on Fragaria × ananassa crown, F. Louws 9, (ICMP<br />

18727); Florida, on Fragaria × ananassa, coll. F.A. Ueckes FAU552 (<strong>CBS</strong> 120005 =<br />

BPI 747977 = ICMP 18609); North Carolina, Lincoln County, on Malus domestica fruit,<br />

coll. T. Sutton CROTTS 13 2001, Jan. 2001 (ICMP 17789).<br />

* <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> (Penz.) Penz. & Sacc.,<br />

Atti Reale Ist. Veneto Sci. Lett. Arti., Serie 6, 2: 670. 1884.<br />

Fig. 24.<br />

Basionym: Vermicularia <strong>gloeosporioides</strong> Penz., Michelia 2: 450.<br />

1882.<br />

= Gloeosporium pedemontanum Pupillo, Ann. Sperim. Agrar. n.s. 6: 57. 1952.<br />

Cannon et al. (2008) provide a description of the <strong>species</strong>.<br />

Geographic distribution and host range: Most isolates of C.<br />

<strong>gloeosporioides</strong> are associated with Citrus, and in many parts of<br />

the world this fungus is common on Citrus, but it also occurs on<br />

other hosts including Ficus, Mangifera, Pueraria, and Vitis. <strong>The</strong><br />

isolate reported as a pathogen of paper mulberry (Broussonetia<br />

papyrifera) by Yan et al. (2011) matches C. <strong>gloeosporioides</strong> s. str.<br />

genetically.<br />

Genetic identification: ITS separates C. <strong>gloeosporioides</strong> from all<br />

other <strong>species</strong>.<br />

Notes: <strong>The</strong> name <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> is currently in<br />

common use in two senses, one a genetically and biologically<br />

broad sense more or less following von Arx (1957, 1970) and<br />

Sutton (1992), including the whole <strong>species</strong> <strong>complex</strong>, the other a<br />

strict sense, encompassing only those specimens genetically<br />

matching the epitype selected for this name by Cannon et al.<br />

(2008). Depending on the context, use of the name in either<br />

sense can be useful. When used in a broad sense in this paper,<br />

it is referred to as the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> or C.<br />

<strong>gloeosporioides</strong> s. lat.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> is often linked taxonomically to<br />

the teleomorph Glomerella cingulata, see notes under G. cingulata.<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Specimens examined: Australia, New South Wales, Tamworth, on Carya<br />

illinoinensis (DAR 76936; ICMP 18738). Italy, Calabria, on Citrus sinensis,<br />

(ex-epitype culture of C. <strong>gloeosporioides</strong> − IMI 356878 = <strong>CBS</strong> 112999 = ICMP<br />

17821); on Citrus limon juice, coll. G. Goidánich, 1951 (ex-holotype culture of<br />

Gloeosporium pedemontanum – <strong>CBS</strong> 273.51 = ICMP 19121). New Zealand,<br />

Auckland, Sandringham, on Citrus sp. fruit, coll. P.R. Johnston C1014.6, 2 May<br />

1988 (ICMP 12939); Auckland, Sandringham, on Ficus sp. fruit, coll. P.R. Johnston<br />

C945.2, 9 May 1988 (ICMP 12066); Auckland, on Citrus sp. fruit, coll. G. Caroll, Feb<br />

2010 (ICMP 18730); Northland, Kerikeri, Kapiro Rd, on Citrus sinensis fruit, coll. P.R.<br />

Johnston C1009.2, 10 Aug. 1988 (ICMP 12938). South Africa, on Mangifera indica,<br />

coll. L. Korsten Cg68 (ICMP 18694). USA, Georgia, on Pueraria lobata (AR2799 =<br />

<strong>CBS</strong> 119204 = BPI 871837 = ICMP 18678); Florida, on Citrus sp. leaf lesion, coll. N.<br />

Peres SRL-FTP-9 (ICMP 18695); Florida, on Vitis vinifera leaf lesion, coll. N. Peres<br />

LAGrape8 (ICMP 18697).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. aeschynomenes”<br />

(Daniel et al. 1973, as aeschynomene).<br />

Notes: See <strong>Colletotrichum</strong> aeschynomenes.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. alatae” R.D. Singh,<br />

Prasad & R.L. Mathur, Indian Phytopathol. 19: 69. 1966.<br />

[nom. inval., no Latin description, no type designated]<br />

Notes: See <strong>Colletotrichum</strong> alatae.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. aleuritis Saccas &<br />

Drouillon [as “aleuritidis”], Agron. Trop. (Nogent-sur-Marne)<br />

6: 249. 1951.<br />

≡ Glomerella cingulata var. aleuritis Saccas & Drouillon [as “aleuritidis”],<br />

Agron. Trop. (Nogent-sur-Marne) 6: 251. 1951.<br />

Notes: Originally described from Aleurites fordii and A. montaba<br />

from French Equatorial Africa, these names have not been used<br />

since being described and the genetic relationship of this fungus<br />

to and within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is unknown.<br />

Although the original publications have not been seen, both names<br />

were tagged as invalid in the Index of Fungi 2: 53, 57 (1952).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. clidemiae” (Trujillo<br />

et al. 1986).<br />

Notes: See <strong>Colletotrichum</strong> clidemiae.<br />

Fig. 24. <strong>Colletotrichum</strong> <strong>gloeosporioides</strong>. A. ICMP 17821<br />

(ex IMI 356878 – ex-epitype culture). A. Cultures on PDA,<br />

10 d growth from single conidia, from above and below.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. cucurbitae”<br />

(Menten et al. 1980).<br />

153


Weir et al.<br />

Notes: First described from cucumber, this fungus is widely<br />

regarded as a synonym of C. orbiculare in the plant pathology<br />

literature (e.g. Snowdon 1991, da Silva et al. 2011).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. cuscutae” (Zhang<br />

1985).<br />

Notes: A strain identified by this name was developed as a<br />

mycoherbicide against dodder (Cuscuta chinensis in China (Zhang<br />

1985). This strain referred to as “Lu Bao No.1” is apparently<br />

included in the study of Guerber et al. (2003) as strain 783 and<br />

belongs to the C. acutatum <strong>species</strong> <strong>complex</strong>. Other strains from<br />

dodder in the USA included in the same study were revealed to be<br />

C. fioriniae, while a strain from Dominica was found to represent a<br />

new <strong>species</strong>, both belonging to the C. acutatum <strong>species</strong> <strong>complex</strong><br />

as well (Damm et al. 2012, this issue).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. gomphrenae Perera,<br />

Revista Fac. Agron. Univ. Nac. La Plata 41: 12. 1965.<br />

Notes: Originally described from Gomphrena globosa, the name has<br />

not been used since it was described and its genetic relationship<br />

to and within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is unknown.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. hederae Pass., Atti<br />

Reale Accad. Italia, Rendiconti., Serie 4, 6: 469. 1889.<br />

Notes: <strong>The</strong> original description of this Hedera-inhabiting <strong>species</strong>,<br />

with fusiform, straight to curved conidia suggests that it is a<br />

synonym of the Hedera pathogen C. trichellum.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. heveae (Petch) Saccas,<br />

Agron. Trop. (Nogent-sur-Marne) 14: 430. 1959.<br />

Basionym: <strong>Colletotrichum</strong> heveae Petch, Ann. Roy. Bot. Gard.<br />

Peradeniya 3(1): 8. 1906.<br />

Notes: Originally described from the leaves of seedlings of Hevea<br />

brasiliensis from Sri Lanka, this fungus was described with very<br />

broad conidia, 18–24 × 7.5–8 µm. Carpenter & Stevenson (1954)<br />

considered this, and several other <strong>Colletotrichum</strong>, Gloeosporium<br />

and Glomerella <strong>species</strong> described from rubber, to be synonyms of<br />

C. <strong>gloeosporioides</strong>. <strong>The</strong> genetic relationship of these <strong>species</strong> to<br />

and within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is unknown. See<br />

also notes in Damm et al. (2012b, this issue) under <strong>Colletotrichum</strong><br />

annelatum.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. hyperici” (Harris<br />

1993).<br />

Notes: This name was first used by Harris (1993) for strains of<br />

C. <strong>gloeosporioides</strong> pathogenic to Hypericum perforatum. Earlier<br />

studies by Hildebrand & Jensen (1991) had found the Hypericum<br />

pathogen to be pathogenic also on several other plants. <strong>The</strong><br />

genetic relationship of the Hypericum pathogen to and within the<br />

C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is unknown. Note that the exholotype<br />

culture of G. cingulata var. migrans, a variety here placed<br />

in synonymy with C. kahawae subsp. ciggaro, was also isolated<br />

from Hypericum.<br />

154<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. jussiaeae” (Boyette<br />

et al. 1979).<br />

Notes: Strains identified as C. <strong>gloeosporioides</strong> “f. sp. jussiaeae” are<br />

highly pathogenic, specialised pathogens of Jussiaea decurrens<br />

(Boyette et al. 1979). <strong>The</strong> genetic relationship of this taxon to and<br />

within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>, or to <strong>Colletotrichum</strong><br />

jussiaeae Earle, is unknown. Isolates pathogenic to Jussiaea have<br />

a similar conidial germination self-inhibitor profile to another isolate<br />

identified as C. fragariae (Tsurushima et al. 1995). <strong>The</strong> authentic<br />

isolate of C. <strong>gloeosporioides</strong> “f. sp. jussiaeae” deposited as ATCC<br />

52634, is not included in this study.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. malvae” (Makowski<br />

& Mortensen 1989).<br />

Notes: Strains identified as C. <strong>gloeosporioides</strong> “f. sp. malvae” were<br />

registered as a bioherbicide against round leafed mallow in Canada<br />

(Makowski & Mortensen 1989). <strong>The</strong> fungus was subsequently<br />

recognised as belonging to the C. orbiculare <strong>species</strong> <strong>complex</strong><br />

(Bailey et al. 1996).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. manihotis”<br />

(Chevaugeon 1956).<br />

Notes: See <strong>Colletotrichum</strong> manihotis.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. melongenae Fournet,<br />

Ann. Mus. Civico Storia Nat. Genova 5: 13. 1973.<br />

Notes: In addition to C. <strong>gloeosporioides</strong> f. melongenae, the names<br />

C. <strong>gloeosporioides</strong> “f. sp. melongenae”, C. melongenae Av.-<br />

Saccá 1917, and C. melongenae Lobik 1928 have been used to<br />

refer to fungi associated with anthracnose diseases of Solanum<br />

melongena (e.g. Sherf & McNab 1986, Kaan 1973). Other names<br />

used for isolates from the same host have included Gloeosporium<br />

melongenae Ellis & Halst. 1891 and G. melongenae Sacc. 1916.<br />

<strong>The</strong> genetic relationships of these eggplant-associated taxa to and<br />

within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> remain unknown.<br />

Solanum melongena associated <strong>species</strong> are known also from the<br />

C. boninense <strong>species</strong> <strong>complex</strong> (Damm et al. 2012b, this issue).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. miconiae” (Killgore<br />

et al. 1999).<br />

Notes: Killgore et al. (1999) reported that the isolates they<br />

recognised as C. <strong>gloeosporioides</strong> “f. sp. miconiae” were highly<br />

specialised pathogens of Miconia calvescens, unable to infect<br />

the closely related Clidemia hirta. <strong>The</strong> original voucher cultures<br />

are no longer available (pers. comm., Robert Barreto). Recently<br />

collected isolates from Miconia from the type locality in Brazil have<br />

proved to be genetically diverse across the C. <strong>gloeosporioides</strong><br />

<strong>species</strong> <strong>complex</strong>, with isolates in both the Kahawae and Musae<br />

clades (unpubl. data). For now the genetic position of this pathogen<br />

remains unresolved.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. minus Simmonds,<br />

Queensland J. Agric. Anim. Sci. 25: 178A. 1968.<br />

Notes: See <strong>Colletotrichum</strong> queenslandicum.


<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. nectrioidea Gonz.<br />

Frag., Bol. Soc. Brot., 2: 52. 1924.<br />

Notes: Originally described from Citrus aurantium from Portugal,<br />

the name has not been used since it was described and its genetic<br />

relationship to and within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

is unknown.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. ortheziidae”<br />

(Marcelino et al. 2008).<br />

Notes: Marcelino et al. (2008) clearly show that the Orthezia<br />

praelonga pathogen belongs in the C. acutatum <strong>species</strong> <strong>complex</strong>,<br />

despite referring to the fungus only as C. <strong>gloeosporioides</strong> “f. sp.<br />

ortheziidae”. See also notes under C. nymphaeae in Damm et al.<br />

(2012a, this issue).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. pilosae” (Singh<br />

1974).<br />

Notes: First described from leaves of Bidens pilosa, this name has<br />

not been used since it was described and its genetic relationship<br />

to and within the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is unknown.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. stylosanthis Munaut,<br />

Mycol. Res. 106: 591. 2002.<br />

Notes: Placed here in synonymy with C. theobromicola; see notes<br />

under C. theobromicola.<br />

Irwin & Cameron (1978) and Munaut et al. (2002) described<br />

different diseases of Stylosanthes associated with Type A and Type<br />

B isolates of C. <strong>gloeosporioides</strong> f. stylosanthis, the two groups of<br />

isolates distinguished morphologically by growth rate in culture and<br />

by conidial morphology. Compared with Type A, the Type B isolates<br />

had a slower growth rate on PDA, and conidia more variable in size<br />

and shape (Irwin & Cameron 1978). <strong>The</strong>y were also distinguished<br />

genetically using RFLP and similar methods (e.g. Munaut et al.<br />

1998, 2002). Munaut et al. (2002) used ITS1 sequences to show<br />

the C. <strong>gloeosporioides</strong> f. stylosanthis to be related to an isolate they<br />

identified as C. fragariae. We regard C. fragariae to be a synonym<br />

of C. theobromicola, with putatively authentic Type A (HM335, C.<br />

<strong>gloeosporioides</strong> f. stylosanthis “f. sp. guianensis”) and Type B (HM<br />

336, C. <strong>gloeosporioides</strong> f. stylosanthis “f. sp. stylosanthis”) isolates<br />

both also belonging to this <strong>species</strong>. From the ITS1 sequence data<br />

available, isolates regarded as typical of Type A (RAPD cluster I)<br />

and of Type B (RAPD cluster II) by Munaut et al. (1998) all belong<br />

in C. theobromicola in the sense that we are using the name; their<br />

RAPD cluster III isolate could be C. tropicale, and their RAPD<br />

cluster IV isolates are probably C. fructicola.<br />

<strong>The</strong> cultures of C. <strong>gloeosporioides</strong> f. stylosanthis that we used<br />

were originally studied by Irwin & Cameron (1978), and selected<br />

as the “types” of “f. sp. guianensis” and “f. sp. stylosanthis” by<br />

Munaut et al. (2002). Both isolates have a ‘stale’ growth form, no<br />

longer forming conidia in culture and with aerial mycelium closely<br />

appressed to the agar surface, resulting in an almost slimy colony<br />

surface. Both isolates had a slow growth rate, similar to that<br />

reported for Type B isolates by Irwin & Cameron (1978). Genetically<br />

both isolates were identical for all the genes we sequenced. This<br />

identity should be checked against additional isolates, especially<br />

some matching Type A sensu Irwin & Cameron (1978) with respect<br />

to both pathogenicity and growth form.<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Sherriff et al. (1994), using ITS2 and partial 28S rDNA<br />

sequences, found isolates they considered to represent C.<br />

<strong>gloeosporioides</strong> f. stylosanthis Type A and Type B respectively to<br />

be genetically distinct. However, their ITS2 sequences show that<br />

the putative Type B isolate in their study was in fact a member of<br />

the C. boninense <strong>species</strong> <strong>complex</strong>.<br />

Specimens examined: Australia, Queensland, Townsville, on Stylosanthes viscosa,<br />

coll. J.A.G. Irwin 21365 (HM335), 1976 (ex-holotype culture of C. <strong>gloeosporioides</strong><br />

f. stylosanthis – MUCL 42294 = ICMP 17957 = <strong>CBS</strong> 124251); Samford, on<br />

Stylosanthes guianensis, coll. J.A.G. Irwin 21398 (HM336), 1979 (MUCL 42295 =<br />

ICMP 17958 = <strong>CBS</strong> 124250).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. stylosanthis “f. sp.<br />

guianensis” (Munaut et al. 2002)<br />

≡ <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. guianensis” (Vinijsanum et al.<br />

1987).<br />

Notes: See notes and specimens examined under C.<br />

<strong>gloeosporioides</strong> f. stylosanthis.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. stylosanthis “f. sp.<br />

stylosanthis” (Munaut et al. 2002).<br />

Notes: See notes and specimens examined under C.<br />

<strong>gloeosporioides</strong> f. stylosanthis.<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. uredinicola”<br />

(Singh 1975).<br />

Notes: Described from uredinia and telia of Ravenelia sessilis<br />

on pods of Albizia lebbek, this name has not been used since it<br />

was described and its genetic relationship to and within the C.<br />

<strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong> is unknown.<br />

<strong>Colletotrichum</strong> gossypii Southw., J. Mycol. 6: 100. 1891.<br />

= Glomerella gossypii Edgerton, Mycologia 1: 119. 1909.<br />

Notes: This <strong>species</strong> was originally described from the USA and was<br />

reported to cause disease symptoms on all parts of cotton plants,<br />

but especially the bolls (Southworth 1891, Edgerton 1909). Isolates<br />

identified as C. gossypii by Shear & Wood (1907) were reportedly<br />

associated with a Glomerella state in culture, and Edgerton (1909)<br />

described Glomerella gossypii from diseased, mature cotton plants<br />

in the USA Edgerton (1909) discussed differences in ascospore<br />

shape between G. gossypii and fruit-rotting isolates of G.<br />

cingulata, with G. gossypii having elliptic, not curved ascospores.<br />

Von Arx (1957) considered C. gossypii to be a synonym of C.<br />

<strong>gloeosporioides</strong> and von Arx & Müller (1954) regarded G. gossypii<br />

to be a synonym of G. cingulata.<br />

Modern authors have recognised two pathogens of cotton, C.<br />

gossypii and C. gossypii var. cephalosporioides. <strong>Colletotrichum</strong><br />

gossypii is reportedly the cause of cotton anthracnose, a damping-off<br />

disease of cotton seedlings, and C. gossypii var. cephalosporioides<br />

the cause of ramulosis, a disease causing abnormal branching<br />

of mature plants (Bailey et al. 1996, Silva-Mann et al. 2005). In<br />

a study based on ITS2 sequences, Bailey et al. (1996) found C.<br />

gossypii and C. gossypii var. cephalosporioides to be genetically<br />

distinct but with both belonging to the C. <strong>gloeosporioides</strong> <strong>species</strong><br />

<strong>complex</strong>. Silva-Mann et al. (2005) also distinguished the two<br />

taxa genetically, based on an AFLP analysis. <strong>The</strong> only DNA<br />

sequences available for isolates identified as C. gossypii and C.<br />

155


Weir et al.<br />

gossypii var. cephalosporioides are ITS2 and the D2 region of<br />

the rDNA LSU, neither of which resolves their relationships within<br />

the C. <strong>gloeosporioides</strong> <strong>complex</strong>. Whether the seedling pathogen<br />

regarded by Silva-Mann et al. (2005) and Bailey et al. (1996) to be<br />

C. gossypii represents the <strong>species</strong> first described from cotton in<br />

the USA is not known. <strong>The</strong> genetic relationship of these apparently<br />

biologically specialised fungi requires additional sequences to be<br />

generated from authentic isolates with known pathogenicity.<br />

<strong>Colletotrichum</strong> gossypii var. cephalosporioides A.S.<br />

Costa, Bragantia 6: 5. 1946.<br />

≡ <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “var. cephalosporioides” (A.S. Costa)<br />

Follin & Mangano, Coton et fibres tropicales 37: 209. 1983. [comb. inval.,<br />

no full reference to basionym]<br />

Notes: See notes under <strong>Colletotrichum</strong> gossypii.<br />

* <strong>Colletotrichum</strong> horii B. Weir & P.R. Johnst., Mycotaxon<br />

111: 21. 2010.<br />

Weir & Johnston (2010) and Xie et al. (2010a) provide descriptions.<br />

Geographic distribution and host range: Associated with fruit and<br />

stem disease of Diospyros kaki from China, Japan, and New<br />

Zealand. Xie et al. (2010a) noted minor symptoms on inoculated<br />

fruit of Capsicum annuum, Musa acuminata, and Cucurbita pepo,<br />

but noted that the fungus had never been associated with disease<br />

symptoms on these hosts from the field.<br />

Genetic identification: ITS distinguishes C. horii from all other<br />

<strong>species</strong>.<br />

Specimens examined: See Weir & Johnston (2010).<br />

<strong>Colletotrichum</strong> hymenocallidis Yan L. Yang, Zuo Y. Liu,<br />

K.D. Hyde & L. Cai, Fungal Diversity 39: 138. 2009.<br />

Notes: Placed here in synonymy with <strong>Colletotrichum</strong> siamense.<br />

See notes and additional specimens examined under C. siamense.<br />

Yang et al. (2009) reported this <strong>species</strong> as a leaf pathogen of<br />

Hymenocallis americana. <strong>The</strong>y distinguished C. hymenocallidis<br />

from C. siamense, also described from Hymenocallis, primarily<br />

on the basis of a multi-gene phylogeny and differences in colony<br />

colour. Although gene selection was appropriate for resolving<br />

genetic relationships within the C. <strong>gloeosporioides</strong> group, Yang<br />

et al. (2009) included only five isolates of the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> in their phylogeny. Based on this isolate selection, the<br />

C. hymenocallidis isolates were genetically distinct from the<br />

C. siamense isolates. However, in our analysis, in which the C.<br />

siamense/C. hymenocallidis group is represented by 30 isolates<br />

from a wide range of hosts from all over the world, authentic isolates<br />

of the two <strong>species</strong> fall within a monophyletic clade that cannot be<br />

further subdivided phylogenetically.<br />

<strong>The</strong> Latin part of the C. hymenocallidis protologue designates<br />

a culture (“Holotypus: Cultura (CSSN2)”) as the holotype but<br />

the English citation of the type specimen corrects this apparent<br />

mistake, citing CSSN2 as an ex-holotype culture, with the herbarium<br />

specimen GZAAS 080001 as the holotype.<br />

Specimen examined: China, Guangxi, Nanning, on Hymenocallis americana leaf<br />

spot, coll. Y.L. Yang GZAAS 080001, 19 Jun 2008 (ex-holotype culture of C.<br />

hymenocallidis − <strong>CBS</strong> 125378 = ICMP 18642).<br />

156<br />

<strong>Colletotrichum</strong> ignotum E.I Rojas, S.A. Rehner & Samuels,<br />

Mycologia 102: 1331. 2010.<br />

Notes: Placed here in synonymy with <strong>Colletotrichum</strong> fructicola. See<br />

notes and additional specimens examined under C. fructicola.<br />

Specimen examined: Panama: Barro Colorado Monument, Tetragastris panamensis<br />

leaf endophyte, coll. E.I. Rojas E886, 1 Jun 2004 (ex-holotype culture of C.<br />

ignotum − <strong>CBS</strong> 125397 = ICMP 18646).<br />

<strong>Colletotrichum</strong> jasmini-sambac Wikee, K.D. Hyde, L. Cai<br />

& McKenzie, Fungal Diversity 46: 174. 2011.<br />

Notes: Placed here in synonymy with <strong>Colletotrichum</strong> siamense<br />

based on the ITS, GAPDH, CAL, TUB2, and ACT gene sequences<br />

from the ex-holotype culture, deposited in GenBank by Wikee et<br />

al. (2011).<br />

Wikee et al. (2011) discussed similarities between C. jasminisambac,<br />

C. siamense and C. hymenocallidis, three <strong>species</strong><br />

genetically close in their phylogenetic analysis. <strong>The</strong> broader range<br />

of isolates representing C. siamense in our analysis shows that<br />

these <strong>species</strong> form a single, monophyletic clade that cannot be<br />

sensibly subdivided (see notes under C. siamense).<br />

Specimen examined: Vietnam, Cu Chi District, Trung An Ward, on living leaves of<br />

Jasminum sambac, Jan. 2009, coll. Hoa Nguyen Thi LLTA–01 (ex-holotype culture<br />

of C. jasmini-sambac – <strong>CBS</strong> 130420 = ICMP 19118).<br />

* <strong>Colletotrichum</strong> kahawae J.M. Waller & Bridge subsp.<br />

kahawae, Mycol. Res. 97: 993. 1993. Fig. 25.<br />

Waller et al. (1993) provide a description.<br />

Geographic distribution and host range: Known only from Coffea<br />

from Africa.<br />

Genetic identification: ACT, CAL, CHS-1, GAPDH, TUB2, SOD2,<br />

and ITS sequences are the same as those from C. kahawae<br />

subsp. ciggaro. <strong>The</strong> two sub<strong>species</strong> can be distinguished by GS<br />

sequences; C. kahawae subsp. kahawae has a 22 bp deletion and<br />

a single C to T transition. Collectively, the two sub<strong>species</strong> can be<br />

distinguished from all other <strong>species</strong> using ITS sequences alone.<br />

Notes: <strong>Colletotrichum</strong> kahawae was proposed by Waller et al. (1993)<br />

as a name to refer specifically to <strong>Colletotrichum</strong> isolates causing<br />

Coffee Berry Disease (CBD), to taxonomically distinguish these<br />

disease-causing isolates from the several other <strong>Colletotrichum</strong><br />

spp. that can be isolated from coffee plants, including C. coffeanum<br />

(see notes under C. coffeanum). <strong>Colletotrichum</strong> kahawae is an<br />

apparently clonal population (Varzea et al. 2002), widespread on<br />

coffee in Africa, and with a distinctive growth form and biology<br />

(Waller et al. 1993).<br />

In this paper C. kahawae sensu Waller et al. (1993) is reduced<br />

to sub<strong>species</strong>. Based on ACT, CAL, CHS-1, GAPDH, TUB2, SOD2,<br />

and ITS gene sequences the coffee berry pathogen cannot be<br />

distinguished from isolates from a wide range of other hosts that<br />

are not pathogenic to coffee. Those other isolates are referred to<br />

here as C. kahawae subsp. ciggaro. We retain a distinct taxonomic<br />

label for the coffee berry pathogen to reflect its biosecurity<br />

importance. In addition to its biology, C. kahawae subsp. kahawae<br />

can be distinguished metabolically, and genetically using GS gene<br />

sequences. Waller et al. (1993) used a metabolic test, an inability


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 25. <strong>Colletotrichum</strong> kahawae subsp. kahawae. A, E. ICMP 17905 (ex IMI 361501). B–C. ICMP 17816 (ex IMI 319418 – ex-holotype culture). C. ICMP 17915 (ex <strong>CBS</strong><br />

982.69). A–B. Appressoria. C. Conidia. D–E. Cultures on PDA, 10 d growth from single conidia, from above and below. Scale bar C = 20 µm. Scale bar of C applies to A–C.<br />

157


Weir et al.<br />

to utilise either citrate or tartrate as a sole carbon source, to help<br />

characterise isolates as C. kahawae. None of our C. kahawae<br />

subsp. kahawae isolates were able to utilise either citrate or tartrate,<br />

whereas all of the C. kahawae subsp. ciggaro isolates were able<br />

to utilise one or both of these carbon sources (Weir & Johnston<br />

2009). All of the C. kahawae subsp. kahawae isolates share a 22<br />

bp deletion in the glutamine synthetase gene, lacking in the C.<br />

kahawae subsp. ciggaro isolates. Note that one of the isolates<br />

metabolically and genetically typical C. kahawae subsp. kahawae<br />

(<strong>CBS</strong> 982.69) was reported by Gielink & Vermeulen (1983) to be<br />

non-pathogenic to coffee, but we have not independently checked<br />

this result.<br />

<strong>The</strong> isolates we accept as C. kahawae subsp. kahawae show<br />

two cultural types, one matching the description of Waller et al.<br />

(1993), slow growing, darkly pigmented cultures with conidia<br />

developing mostly in the aerial mycelium. <strong>The</strong> second cultural type<br />

grew even more slowly, had little or no pigmentation within the agar,<br />

and the colony surface was covered with numerous acervuli and<br />

orange conidial masses. Metabolically and genetically both cultural<br />

types were the same, and pathogenicity tests showed that the nonpigmented<br />

isolates caused CBD (unpubl. data, D. Silva, Centro de<br />

Investigação das Ferrugens do Cafeeiro). Rodriguez et al. (1991)<br />

reported further variation in cultural appearance amongst CBD<br />

causing isolates.<br />

Waller et al. 1993 stated that C. kahawae was not known to form<br />

ascospores. However, Gielink & Vermeulen (1983) observed the<br />

production of perithecia on coffee berries that had been inoculated<br />

with CBD-causing isolates, many months after inoculation and death<br />

of the berries. At least one of the isolates that they cited with this<br />

biology, <strong>CBS</strong> 135.30, has the GS sequence typical of C. kahawae<br />

subsp. kahawae. Vermeulen et al. (1984) grew cultures from the<br />

perithecia that developed on the previously inoculated berries, and<br />

found that none were pathogenic to coffee. It is possible that the<br />

perithecia developing on inoculated berries reported by Gielink &<br />

Vermeulen (1983) were from other <strong>Colletotrichum</strong> spp. present on<br />

the berries before they were inoculated, and represented <strong>species</strong><br />

distinct from C. kahawae subsp. kahawae. A similar situation has<br />

been noted with some of our inoculations, where <strong>species</strong> present<br />

on tissues prior to inoculation, either endophytic or latent, started<br />

to sporulate on the dead tissue following inoculation (unpubl. data,<br />

B.S. Weir).<br />

Based on ITS sequences, most of the accessions in GenBank<br />

identified as C. kahawae and isolated from coffee, match our<br />

concept of C. kahawae subsp. kahawae. <strong>The</strong>re are two exceptions,<br />

AF534468 (from Malawi) and AY376540 (STE-U 5295 = IMI 319424<br />

= <strong>CBS</strong> 112985, from Kenya). <strong>The</strong> Kenyan isolate was cited as C.<br />

kahawae in Lubbe et al. (2004). Based on the ITS sequences, and<br />

the TUB2 sequence from isolate STE-U 5295 (AY376588), these<br />

isolates represent C. siamense.<br />

Specimens examined: Angola, Ganada, on Coffea arabica berry, coll. J.N.M. Pedro<br />

16/65, 2 Jun. 1965 (IMI 310524 = <strong>CBS</strong> 982.69 = ICMP 17915). Cameroon, on<br />

Coffea arabica (IMI 361501 = ICMP 17905). Kenya, Ruiru, Kakuzi Estate, on Coffea<br />

arabica young shoots, coll. D.M. Masaba 22/87, 29 Jan. 1987 (ex-holotype culture<br />

of C. kahawae – IMI 319418 = ICMP 17816); on Coffea sp., coll. E.C. Edwards,<br />

May 1930 (<strong>CBS</strong> 135.30 = ICMP 17982). Malawi, on Coffea arabica (IMI 301220 =<br />

ICMP 17811).<br />

* <strong>Colletotrichum</strong> kahawae subsp. ciggaro B. Weir & P.R.<br />

Johnst., subsp. nov. MycoBank MB563758. Figs 26, 27.<br />

= Glomerella cingulata var. migrans Wollenw., Z. Parasitenk. (Berlin) 14: 262.<br />

1949.<br />

= Glomerella rufomaculans var. vaccinii Shear, Bull. Torrey Bot. Club. 34: 314.<br />

1907.<br />

158<br />

Etymology: Based on the title of the Jim Jarmusch movie “Coffee<br />

and Cigarettes”, referring to the close genetic relationship between<br />

C. kahawae subsp. ciggaro and the coffee pathogen C. kahawae<br />

subsp. kahawae; ciggaro is Portuguese for cigarette.<br />

Holotype: Australia, on Olea europaea, coll. V. Sergeeva UWS124,<br />

1989, PDD 102232; ex-type culture ICMP 18539.<br />

Colonies grown from single conidia on Difco PDA 75–85 mm diam<br />

after 10 d for most isolates, the ex-holotype culture of G. cingulata<br />

var. migrans 48–49 mm diam. Aerial mycelium cottony, grey, dense,<br />

or in some isolates with dark stromatic masses and associated<br />

orange conidial ooze showing through mycelium from agar<br />

surface; in reverse agar with pinkish-orange pigments (6B4–7B4),<br />

irregular scattered black spots, and variable levels of development<br />

of overlying dark grey to green-grey pigments (4C2–5D4), these<br />

sometimes in discrete sectors. See notes below about a divergent<br />

growth form single ascospore cultures from perithecia in culture.<br />

Conidia form on dark-based acervuli, (12–)16–19.5(–29) × (4.5–)<br />

5(–8) µm (av. 17.8 × 5.1 µm, n = 214), cylindric, straight, apex<br />

rounded, often tapering slightly towards the base. Appressoria<br />

typically cylindric to fusoid in shape, deeply lobed. Perithecia<br />

numerous, forming tightly packed clumps, individual perithecia<br />

globose, small, about 250 µm diam, with a short ostiolar neck. Asci<br />

55–100 × 10–12 µm, 8–spored. Ascospores (13.5–)17.5–20(–24)<br />

× (4–)4.5–5(–6.5) µm (av. 18.8 × 4.8 µm, n = 121), gently curved,<br />

tapering to quite narrow, rounded ends, widest point usually<br />

towards one end of the spore.<br />

Geographic distribution and host range: Known from Australia,<br />

Germany, New Zealand, and South Africa. Both host and geographic<br />

range of the isolates we accept in C. kahawae subsp. ciggaro<br />

are broad. Genetic identification: ACT, CAL, CHS-1, GAPDH,<br />

TUB2, SOD2, and ITS sequences match those from C. kahawae<br />

subsp. kahawae. <strong>The</strong> two sub<strong>species</strong> can be distinguished by GS<br />

sequences. Collectively, the two sub<strong>species</strong> can be distinguished<br />

from all other <strong>species</strong> using ITS sequences alone.<br />

Notes: <strong>The</strong> authentic isolate of G. cingulata var. migrans (<strong>CBS</strong> 237.49)<br />

differed from all other isolates we accept in C. kahawae subsp.<br />

ciggaro by its slower growth rate. Wollenweber & Hochapfel (1949)<br />

distinguished Glomerella cingulata var. migrans from G. cingulata var.<br />

cingulata on the basis of pathogenicity (G. cingulata var. migrans was<br />

pathogenic to Hypericum and not to apple) and because of its slightly<br />

longer ascospores and shorter conidia — ascospores average 21 ×<br />

4.2 µm versus 18 × 4.6 µm, conidia average 14 × 5.2 µm versus 18 ×<br />

5 µm (Wollenweber & Hochapfel 1949). We were unable to produce<br />

ascospores from <strong>CBS</strong> 237.49, the conidia were similar in size to that<br />

reported by Wollenweber & Hochapfel (1949), averaging 16.6 × 5.3<br />

µm. However, the average ascospore and conidial lengths of our C.<br />

kahawae subsp. ciggaro isolates varied across the range cited by<br />

Wollenweber & Hochapfel (1949) for both G. cingulata var. cingulata<br />

and G. cingulata var. migrans, the average ascospore length from<br />

individual isolates ranging from 16.6 to 20 µm, the average conidial<br />

length ranging from 14.9 to 21.2 µm.<br />

Glomerella rufomaculans var. vaccinii was described by Shear<br />

(1907) for a fungus isolated from cranberry that was morphologically<br />

identical to isolates from apple and other hosts but which appeared<br />

to be biologically distinct (Shear 1907, Shear & Wood 1913). A<br />

putatively authentic isolate of this <strong>species</strong>, deposited by Shear<br />

in <strong>CBS</strong> in 1922, matches C. kahawae subsp. ciggaro genetically.<br />

Polashock et al. (2009) discussed the diversity of <strong>Colletotrichum</strong>


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 26. <strong>Colletotrichum</strong> kahawae subsp. ciggaro. A. ICMP 12952. B, D. ICMP 17932 (ex <strong>CBS</strong> 112984). E, H. ICMP 17931 (ex IMI 359911). C, F. ICMP 18539 – ex-holotype<br />

culture. G. ICMP 18531. A–B. Asci and ascospores. C–D. Appressoria. E. Setae. F–H. Conidia. Scale bars A, E = 20 µm. Scale bar of A applies to A–D, F–H.<br />

spp. associated with North American cranberry fruit rots, reporting a<br />

close match between their isolates and C. kahawae. Incorporation<br />

of their ITS sequences into our alignment confirms this. Whether or<br />

not there is a genetically distinct, cranberry specialised taxon within<br />

C. kahawae requires additional genes to be sequenced from the<br />

cranberry-associated isolates.<br />

<strong>Colletotrichum</strong> kahawae subsp. ciggaro was referred to as C.<br />

<strong>gloeosporioides</strong> Group B by Johnston & Jones (1997) and Johnston<br />

159


Weir et al.<br />

Fig. 27. <strong>Colletotrichum</strong> kahawae subsp. ciggaro. A. ICMP 12953. B. ICMP 18534. C. ICMP 17922 (ex <strong>CBS</strong> 237.49 – ex-holotype culture of Glomerella cingulata var. migrans).<br />

D. ICMP 17932 (ex <strong>CBS</strong> 112984). E. ICMP 18539 – ex-holotype culture of C. kahawae subsp. ciggaro. F. ICMP 12952 – single ascospore cultures from single conidial isolate.<br />

Cultures on PDA, 10 d growth from single conidia, from above and below.<br />

et al. (2005), and as Undescribed Group 1 by Silva et al. (2012b).<br />

Single ascospore isolates derived from perithecia forming in<br />

single conidial cultures of the avocado-associated isolates of C.<br />

160<br />

kahawae subsp. ciggaro from New Zealand showed two highly<br />

divergent growth forms (Fig. 27F). One typical of the “wild type”<br />

(cottony, grey to dark grey aerial mycelium with dark-based acervuli


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 28. <strong>Colletotrichum</strong> musae. A. ICMP 12930. B. ICMP 18600. C. ICMP 17817 (ex IMI 52264). Cultures on PDA, 10 d growth from single conidia, from above and below.<br />

and orange conidial masses visible through the mycelium, in<br />

reverse with pinkish-orange pigmentation, in places this masked by<br />

irregular patches or sectors of dark grey pigmentation), the other<br />

more or less lacking aerial mycelium, the surface of the colony<br />

covered with small, pale-based acervuli with bright orange conidial<br />

ooze, in reverse bright orange from the conidial ooze. Although<br />

common from single ascospores, the bright, conidial cultural type<br />

is rarely formed by isolates from nature (unpubl. data). Similar<br />

dimorphic cultural types have been observed also from single<br />

ascospore isolates from a member of the C. boninense <strong>complex</strong>,<br />

C. constrictum (unpubl. data, P.R. Johnston).<br />

Other specimens examined: Brazil, on leaves of Miconia sp., coll. R. Barreto<br />

RWB1054, 2009 (ICMP 18728). Germany, Berlin-Dahlem, on stem of Hypericum<br />

perforatum, Jun. 1937 (ex-holotype culture of Glomerella cingulata var. migrans<br />

– <strong>CBS</strong> 237.49 = ICMP 17922). New Zealand, Auckland, Waitakere Ranges, on<br />

leaves of Kunzea ericoides, coll. S. Joshee 5Kun3.10 (ICMP 18741); Auckland,<br />

Waitakere Ranges, on leaves of K. ericoides, coll. S. Joshee 7Kun5.2 (ICMP 18534);<br />

Auckland, Waitakere Ranges, on leaves of Toronia toru, coll. G. Carroll TOROTO3<br />

(ICMP 18544); Te Puke, on Persea americana fruit rot, coll. W.F.T. Hartill, 19 Jan.<br />

1989 (ICMP 18531); Te Puke, on P. americana fruit rot, coll. W.F.T. Hartill, 8 Feb.<br />

1988 (ICMP 12952); Te Puke, on P. americana fruit rot, coll. W.F.T. Hartill, 28 Sep.<br />

1991 (ICMP 12953). South Africa, Madeira, on Dryandra sp., coll. J.E. Taylor, 1<br />

Apr. 2001 (<strong>CBS</strong> 112984, as C. crassipes = ICMP 17932). Switzerland, on Dryas<br />

octopetala, coll. P. Cannon (IMI 359911 = <strong>CBS</strong> 12988 = ICMP 17931). USA, on<br />

Vaccinium macrocarpum leaves, coll. C.L. Shear, Apr. 1922 (authentic culture of<br />

G. rufomaculans var. vaccinii – <strong>CBS</strong> 124.22 = ICMP 19122).<br />

<strong>Colletotrichum</strong> manihotis Henn., Hedwigia 43: 94. 1904.<br />

Notes: Anthracnose is an important disease of cassava (e.g.<br />

Chevaugeon 1956, Makambila 1994, Fokunang et al. 2000,<br />

Owolade et al. 2008), variously referred to <strong>Colletotrichum</strong> manihotis,<br />

Gloeosporium manihotis Henn., Glomerella manihotis (Sacc.)<br />

Petr., Glomerella cingulata “f. sp. manihotis”, or C. <strong>gloeosporioides</strong><br />

“f. sp. manihotis”. <strong>The</strong> original descriptions of both C. manihotis<br />

and Gloeosporium manihotis are of <strong>species</strong> with short, broad<br />

conidia (8–15 × 4–6 µm), and Chevaugeon (1956) regarded all of<br />

these cassava-associated fungi as con-specific. However, based<br />

on Fokunang et al. (2000), a morphologically highly diverse set of<br />

<strong>Colletotrichum</strong> isolates are associated with diseased plants. <strong>The</strong>re<br />

are three GenBank accessions of <strong>Colletotrichum</strong> from cassava,<br />

all from China, and although only ITS sequences are available for<br />

these isolates, they appear to represent a single, distinct <strong>species</strong><br />

within the C. <strong>gloeosporioides</strong> <strong>complex</strong>. How these Chinese isolates<br />

relate to cassava-associated isolates from other parts of the world<br />

is not known.<br />

* <strong>Colletotrichum</strong> musae (Berk. & M.A. Curtis) Arx, Verh.<br />

Kon. Ned. Akad. Wetensch., Afd. Natuurk., Sect. 2 51(3):<br />

107. 1957. Fig. 28.<br />

Basionym: Myxosporium musae Berk. & M.A. Curtis, Grevillea 3:<br />

13. 1874.<br />

Su et al. (2011) provide a description.<br />

Geographic distribution and host range: Found in association with<br />

fruit lesions of Musa spp. in many regions.<br />

Genetic identification: ITS sequences separate C. musae from all<br />

other <strong>species</strong>.<br />

Notes: <strong>Colletotrichum</strong> musae was originally described from North<br />

Carolina (Berkeley 1874), and the name was recently epitypified<br />

by Su et al. (2011) on the basis of a specimen collected in Florida<br />

161


Weir et al.<br />

(ex-epitype culture <strong>CBS</strong> 116870). Su et al. (2011) cite several<br />

strains from Thailand that match their concept of C. musae, and<br />

isolates from anthracnose symptoms on banana fruit from several<br />

parts of the world are the same based on our study. <strong>The</strong>se<br />

isolates form a well-supported clade within the C. <strong>gloeosporioides</strong><br />

<strong>species</strong> <strong>complex</strong>, show low levels of genetic differentiation, and<br />

based on ITS sequences are consistent with C. musae sensu<br />

Sreenivasaprasad et al. (1996), Nirenberg et al. (2002) and Shenoy<br />

et al. (2007). <strong>The</strong> morphology in culture agrees with the description<br />

of Sutton & Waterston (1970).<br />

We have not seen a Glomerella state in culture and none was<br />

mentioned by Su et al. (2011). However, Rodriguez & Owen (1992)<br />

reported rare production of perithecia from crosses between two<br />

of 14 isolates identified as C. musae. It is not known whether the<br />

isolates studied by Rodriguez & Owen (1992) match our concept of<br />

C. musae genetically, but it is possible that this <strong>species</strong> behaves in a<br />

similar way to some <strong>species</strong> in the C. acutatum <strong>complex</strong>, where the<br />

sexual morph can be generated in culture under suitable conditions<br />

(Guerber & Correll 2001). <strong>The</strong> name “Glomerella musae”, used by<br />

Rodriguez & Owen (1992) and Krauss et al. (2001), has never been<br />

validly published.<br />

More than one <strong>species</strong> of <strong>Colletotrichum</strong> has been found<br />

in association with rotting banana fruit. From isolates with well<br />

characterised sequence data these include a <strong>species</strong> belonging to<br />

C. acutatum s. lat. (Sherriff et al. 1994, Johnston & Jones 1997) that<br />

is described as C. paxtonii (Damm et al. 2012a, this issue), and C.<br />

karstii that belongs to the C. boninense <strong>species</strong> <strong>complex</strong> (Damm et<br />

al. 2012b, this issue). <strong>The</strong> latter forms a sexual stage in culture and<br />

is known from Musa in South America and Australia, as well as from<br />

many other hosts worldwide, often as an endophyte. Species in the C.<br />

boninense <strong>species</strong> <strong>complex</strong> have been previously confused with C.<br />

<strong>gloeosporioides</strong> s. lat. Greene (1967) referred isolates pathogenic to<br />

banana that were not associated with a teleomorph to C. musae, and<br />

a second non-pathogenic <strong>species</strong> that formed fertile ascospores, to<br />

C. <strong>gloeosporioides</strong>. Whether Glomerella musarum Petch, described<br />

from leaves of banana and cited as the teleomorph of C. musae by<br />

Sutton (1992) and Hyde et al. (2009), is a synonym of C. musae in<br />

the sense we use the name here is not known, but seems unlikely<br />

given the rare production of perithecia by this <strong>species</strong>.<br />

Specimens examined: Indonesia, on Musa sp., coll. G. von Becze, Jan. 1931 (<strong>CBS</strong><br />

192.31 = ICMP 17923). Kenya, on Musa sapientum, coll. R.M. Nattrass 1850, 1<br />

Jan. 1953 (IMI 52264 = ICMP 17817). New Zealand, Auckland (imported fruit), on<br />

Musa sp., coll. P.R. Johnston C1197.1, 24 May 1991 (ICMP 12931; PDD 59100);<br />

Auckland (fruit imported from the Phillipines), on Musa sp., coll. S. Bellgard, 5 May<br />

2009 (ICMP 18600); Auckland, Mt Albert Research Centre, Musa sp. spots on green<br />

fruit, coll. P.R. Johnston C809.2, 12 Aug. 1987 (ICMP 12930; PDD 46160); Auckland<br />

(fruit imported from the Phillipines), on Musa sp., coll. B. Weir, 17 May 2009 (ICMP<br />

18701; PDD 97438). USA, Florida, on Musa sp., coll. M. Arzanlou A-1 (ex-epitype<br />

culture of C. musae – <strong>CBS</strong> 116870 = ICMP 19119).<br />

Glomerella musarum Petch, Ann. Roy. Bot. Gard.<br />

Peradeniya 6(3): 223. 1917.<br />

Notes: See notes under C. musae.<br />

* <strong>Colletotrichum</strong> nupharicola D.A. Johnson, Carris & J.D.<br />

Rogers, Mycol. Res. 101: 647. 1997. Fig. 29.<br />

Johnson et al. (1997) provide a description.<br />

Geographic distribution and host range: Known only from the USA,<br />

on the aquatic plants Nuphar and Nymphaea spp.<br />

162<br />

Genetic identification: One of the two ITS haplotypes of C.<br />

nupharicola is identical with C. queenslandicum. All other<br />

genes distinguish this <strong>species</strong> well from other <strong>species</strong> in the C.<br />

<strong>gloeosporioides</strong> <strong>complex</strong>.<br />

Notes: Sequence data from the ex-holotype culture of C. nupharicola<br />

places it within the C. <strong>gloeosporioides</strong> <strong>complex</strong>, genetically close to<br />

C. fructicola and C. alienum in the Musae clade. This apparently<br />

host-specific <strong>species</strong> and has a distinctive, slow growth in culture<br />

and massive conidia (Johnson et al. 1997).<br />

Johnson et al. (1997) compare C. nupharicola with another<br />

water plant pathogen, C. nymphaeae, that is epitypified and shown<br />

to belong to the C. acutatum <strong>species</strong> <strong>complex</strong> by Damm et al.<br />

(2012a, this issue).<br />

Specimens examined: USA, Washington, King Co., on Nuphar lutea subsp.<br />

polysepala, coll. D.A. Johnson A-7, Oct. 1993 (<strong>CBS</strong> 469.96 = ICMP 17938);<br />

Washington, Yakima Co., on N. lutea subsp. polysepala, coll. D.A. Johnson A-2,<br />

Oct. 1993 (ex-holotype culture − <strong>CBS</strong> 470.96 = ICMP 17939); Rhode Island, on<br />

Nymphaea ordorata, coll. R.D. Goos RDG-291, 1979 (<strong>CBS</strong> 472.96 = ICMP 18187).<br />

Gloeosporium pedemontanum Pupillo, Ann. Sperim. Agrar.<br />

n.s. 6: 57. 1952.<br />

Notes: Placed here in synonymy with C. <strong>gloeosporioides</strong>. See<br />

notes under C. <strong>gloeosporioides</strong>.<br />

Specimen examined: Italy, on Citrus limon juice, coll. G. Goidánich, 1951 (exholotype<br />

culture of G. pedemontanum – <strong>CBS</strong> 273.51 = ICMP 19121).<br />

* <strong>Colletotrichum</strong> psidii Curzi, Atti dell’Istituto Botanico<br />

dell’Università di Pavia, ser. 3, 3: 207. 1927. Fig. 30.<br />

Colonies grown from single conidia on Difco PDA 58–63 mm diam<br />

after 10 d, aerial mycelium dense, cottony to felted, uniform in<br />

height, white to off-white; in reverse uniformly pale creamy yellow<br />

(2A2–2A3) or in some cultures becoming dull greyish yellow (2D2–<br />

2E2) towards the centre. No conidiogenous cells or conidia seen.<br />

Geographic distribution and host range: Known from a single<br />

isolate, from Psidium from Italy.<br />

Genetic identification: Although known from only one isolate, ITS<br />

sequences separate C. psidii from all other taxa.<br />

Notes: A putatively authentic isolate of this <strong>species</strong>, deposited<br />

in <strong>CBS</strong> by Curzi shortly after publication of C. psidii, represents<br />

a genetically distinct <strong>species</strong> within the Kahawae clade. <strong>The</strong> only<br />

available culture is stale, no longer forming conidia. Curzi (1927)<br />

describes the conidia as 12–15 × 3.5–4.5 µm, cylindric with<br />

rounded ends, straight or rarely slightly curved.<br />

Anthracnose diseases have been noted for Psidium spp.<br />

(guava) from several tropical regions of the world (e.g. MacCaughey<br />

1917, Venkatakrishniah 1952, Liu 1972, Misra 2004). It is likely that<br />

several <strong>Colletotrichum</strong> spp. are associated with guava fruit rots.<br />

Whether the fungus described by Curzi from an Italian botanical<br />

garden represents one of the <strong>species</strong> causing a guava disease in<br />

the tropics is not known. All other members of the Kahawae clade<br />

are predominantly tropical, so perhaps this fungus was introduced<br />

to Italy along with its host plant. Misra (2004) uses C. psidii to refer<br />

to a <strong>Colletotrichum</strong> <strong>species</strong> with curved conidia.<br />

One other <strong>species</strong> has been described from this host,<br />

Glomerella psidii (basionym Gloeosporium psidii), the relationship


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 29. <strong>Colletotrichum</strong> nupharicola. A. ICMP 17939 (ex <strong>CBS</strong> 470.96 – ex-holotype culture). B. ICMP 17938 (ex <strong>CBS</strong> 469.96). C. ICMP 18187 (ex <strong>CBS</strong> 472.96). Cultures on<br />

PDA, 10 d growth from single conidia, from above and below.<br />

Fig. 30. <strong>Colletotrichum</strong> psidii (ICMP 19120,<br />

ex <strong>CBS</strong> 145.29 – authentic culture).<br />

Cultures on PDA, 10 d growth from single<br />

hyphal tips, from above and below.<br />

163


Weir et al.<br />

of this <strong>species</strong> to C. psidii remains unknown. A new <strong>species</strong> on<br />

Psidium guajava, C. guajavae, belonging to the C. acutatum<br />

<strong>species</strong> <strong>complex</strong>, is described elsewhere in this volume (Damm<br />

et al. 2012a).<br />

Specimen examined: Italy, Rome, on Psidium sp., coll. M. Curzi (authentic culture<br />

of C. psidii – <strong>CBS</strong> 145.29 = ICMP 19120).<br />

Glomerella psidii (Delacr.) J. Sheld., Bull. West Virginia<br />

Agric. Exp. Sta. 104: 311. 1906.<br />

Basionym: Gloeosporium psidii Delacr., Bull. Soc. Mycol. France.<br />

19: 144. 1903.<br />

Notes: Sheldon (1906) produced perithecia in culture from<br />

isolates he considered typical of Gloeosporium psidii and on this<br />

basis recombined the <strong>species</strong> described by Delacroix (1903) in<br />

Glomerella. <strong>The</strong> relationship of G. psidii to <strong>Colletotrichum</strong> psidii,<br />

also described from guava, is not known. See notes under C. psidii.<br />

* <strong>Colletotrichum</strong> queenslandicum B. Weir & P.R. Johnst.,<br />

nom. nov. et stat. nov. MycoBank MB563593. Fig. 31.<br />

Basionym: <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. minus Simmonds,<br />

Queensland J. Agric. Anim. Sci. 25: 178A. 1968. [as var. minor]<br />

Etymology: based on the region from which the type specimen of<br />

this <strong>species</strong> was collected.<br />

Holotype: Australia, Queensland, Ormiston, on Carica papaya, coll.<br />

J.H. Simmonds, Oct. 1965, IMI 117612.<br />

Epitype: Australia, Queensland, Brisbane, on Carica papaya, coll.<br />

J.H. Simmonds 11663C, Sep. 1965, epitype here designated PDD<br />

28797; ex-epitype culture ICMP 1778.<br />

Colonies grown from single conidia on Difco PDA 62–74 mm diam<br />

after 10 d, aerial mycelium either dense, cottony, uniform, grey,<br />

or with aerial mycelium lacking, towards centre of colony with<br />

numerous, small acervuli with dark bases and orange conidial<br />

ooze; in reverse cultures with copious aerial mycelium uniformly<br />

dark grey (1F2), those with little aerial mycelium having a pinkish<br />

brown (8B4) pigment within the agar, the dark bases of the acervuli<br />

and the colour of the conidial ooze visible through the agar. Conidia<br />

(12–)14.5–16.5(–21.5) × (3.5–)4.5–5(–6) µm (av. 15.5 × 4.8 µm, n<br />

= 96), cylindric, straight, sometimes slightly constricted near centre,<br />

ends broadly rounded. Appressoria about 6–12 µm diam., globose<br />

to short-cylindric, rarely lobed. Perithecia not seen.<br />

Geographic distribution and host range: Known from Carica papaya<br />

and Persea americana from Queensland, Australia, and from<br />

Coffea berries from Fiji. Simmonds (1965) reported from Australia<br />

what he considered to be the same fungus also from Mangifera<br />

indica, Malus sylvestris, and “many other hosts”.<br />

Genetic identification: ITS sequences do not separate C.<br />

queenslandicum from some C. fructicola, some C. siamense, and<br />

some C. tropicale isolates. It is best distinguished from these taxa<br />

using TUB2, GAPDH, or GS.<br />

Notes: <strong>The</strong> ex-type cultures cited by Simmonds (1968) are no<br />

longer in storage at BRIP in Queensland (R. Shivas, pers. comm.)<br />

and presumably lost. However, we do have two cultures identified<br />

as C. <strong>gloeosporioides</strong> var. minus by Simmonds and isolated from<br />

164<br />

the same host from the same locality as the holotype (Simmonds<br />

isolates 16633C and 1647A2), that had been sent to Joan Dingley<br />

in 1965 and subsequently stored in the ICMP culture collection.<br />

<strong>The</strong> culture selected here as epitype (Simmonds 11663C = ICMP<br />

1778) matches the Simmonds (1965) description of this fungus as<br />

having “an abundance of aerial mycelium in culture”. Our conidial<br />

measurements from ICMP 1778 and 1780 are broader than those<br />

given by Simmonds (1965), but he does note that “Confusion can<br />

occur between narrower strains of C. <strong>gloeosporioides</strong> and broader<br />

strains of C. <strong>gloeosporioides</strong> var. minus …”. Simmonds (1965) also<br />

notes that perithecia may rarely be seen in cultures of some isolates.<br />

<strong>The</strong> isolates accepted here as C. queenslandicum are<br />

genetically distinct within the Musae clade of C. <strong>gloeosporioides</strong> s.<br />

lat. <strong>Colletotrichum</strong> minus Zimm. (1901) requires that we propose a<br />

nom. nov. for this fungus at <strong>species</strong> rank.<br />

Simmonds (1965) considered C. <strong>gloeosporioides</strong> var. minus<br />

to be the conidial state of Glomerella cingulata var. minor<br />

Wollenw. Wollenweber & Hochapfel (1949) used the name<br />

Gloeosporium elasticae Cooke & Massee for the conidial state<br />

of G. cingulata var. minor, the type specimens for both names<br />

being from Ficus. Simmonds (1965) noted that it was not possible<br />

to transfer G. elasticae to <strong>Colletotrichum</strong> because <strong>Colletotrichum</strong><br />

elasticae had already been published for a different fungus.<br />

However, rather than proposing a nom. nov. for Gloeosporium<br />

elasticae, he described C. <strong>gloeosporioides</strong> var. minus as a new<br />

variety, with a different type specimen. Glomerella cingulata var.<br />

minor is genetically distinct from the specimen Simmonds chose<br />

as the type of C. <strong>gloeosporioides</strong> var. minus, see notes under G.<br />

cingulata var. minor.<br />

Other specimens examined: Australia, Queensland, Brisbane, on Carica sp.,<br />

coll. J.H. Simmonds 16347A2 (ICMP 1780, dried culture stored as PDD 28797);<br />

Queensland, Home Hill, on Persea americana, coll. L. Coates 22516, Feb. 1983<br />

(ICMP 12564). Fiji, on Coffea sp. berry, coll. R. Gounder, Apr. 1988 (ICMP 18705).<br />

Glomerella rufomaculans var. vaccinii Shear, Bull. Torrey<br />

Bot. Club. 34: 314. 1907.<br />

Notes: Placed here in synonymy with <strong>Colletotrichum</strong> kahawae<br />

subsp. ciggaro. See notes under C. kahawae subsp. ciggaro. Note<br />

that Saccardo & Trotter (1913) place Shear’s variety in Glomerella<br />

fructigena (Clint.) Sacc., a rarely used <strong>species</strong> name, placed in<br />

synonymy with G. cingulata by von Arx & Müller (1954).<br />

Specimen examined: USA, on Vaccinium macrocarpum leaves, coll. C.L. Shear,<br />

Apr. 1922 (authentic isolate of G. rufomaculans var. vaccinii – <strong>CBS</strong> 124.22 = ICMP<br />

19122).<br />

* <strong>Colletotrichum</strong> salsolae B. Weir & P.R. Johnst., sp. nov.<br />

MycoBank MB563589. Fig. 32.<br />

= <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> “f. sp. salsolae” (Berner et al. 2009).<br />

Etymology: Based on C. <strong>gloeosporioides</strong> “f. sp. salsolae”, referring<br />

to the host from which this fungus was originally collected.<br />

Holotype: Hungary, on Salsola tragus, coll. D. Berner [specimen<br />

from plants inoculated with strain 96-067, originally collected I.<br />

Schwarczinger & L. Vajna on Salsola tragus from Bugac, near<br />

Kiskunsag National Park, 1996], BPI 878740; ex-holotype culture<br />

ICMP 19051.<br />

Colonies grown from single conidia on Difco PDA 38–42 mm diam<br />

after 10 d, aerial mycelium sparse, cottony, pale grey, surface of


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 31. <strong>Colletotrichum</strong> queenslandicum. A, C, E. ICMP 1778 – ex-epitype culture. B, F. ICMP 1780. D, G. ICMP 12564. H. ICMP 18705. A–B. Appressoria. C–D. Conidia. E–H.<br />

Cultures on PDA, 10 d growth from single conidia, from above and below. Scale bar A = 20 µm. Scale bar of A applies to A–D.<br />

colony dark, a more or less continuous layer of acervulus-like<br />

structure with deep orange brown conidial masses and numerous<br />

setae; in reverse dark purplish-black near centre of colony, dark<br />

olivaceous near the margin. Conidia (10–)14–16.5(–20.5) × (4.5–)<br />

5.5–6(–7.5) µm (av. 15.3 × 5.8 µm, n = 24), highly variable in size<br />

and shape, subglobose to long-cylindric, apex usually broadly<br />

165


Weir et al.<br />

Fig. 32. <strong>Colletotrichum</strong> salsolae. A, C–H. ICMP 19051 – ex-holotype culture. B. BPI 878740 – holotype. A. Cultures on PDA, 10 d growth from single conidia, from above and<br />

below. B. Lesion on stem, dried type specimen. C. Conidiogenous cells. D–E. Conidia. F–G. Appressoria. Scale bars B = 1 mm, C = 20 µm. Scale bar of C applies to C–G.<br />

rounded, small truncate scar at base. Conidiogenous cells 13–18<br />

× 4–6.5 µm, cylindric to flask-shaped, tapering at apex to narrow,<br />

166<br />

phialidic conidiogenous locus, wall at base often encrusted with<br />

dark brown material. Appressoria sparsely developed, cylindric to


elliptic, simple; many putatively partially developed appressoria,<br />

similar in shape to those with dark and thick walls and also with<br />

an appressorial pore, but the wall remains thin and only slightly<br />

pigmented. Perithecia not seen.<br />

Geographic distribution and host range: Known from throughout the<br />

geographic range of Salsola tragus (Berner et al. 2009), reported in<br />

nature only from Salsola spp.<br />

Genetic identification: ITS sequences of C. salsolae are very close<br />

to C. alienum and some C. siamense isolates. <strong>The</strong>se <strong>species</strong> can<br />

be distinguished using TUB2 or GAPDH.<br />

Notes: Isolates of C. <strong>gloeosporioides</strong> pathogenic to Salsola<br />

tragus were reported by Schwarczinger et al. (1998) and referred<br />

to as C. <strong>gloeosporioides</strong> “f. sp. salsolae” by Berner et al. (2009).<br />

Although mildly pathogenic to a wide range of hosts in glasshouse<br />

pathogenicity tests, this fungus causes severe disease only on<br />

Salsola spp. with the exception of S. orientalis, S. soda, and S.<br />

vermiculata (Berner et al. 2009).<br />

<strong>Colletotrichum</strong> salsolae belongs to the Musae clade, and<br />

although genetically close to several other <strong>species</strong>, it is biologically<br />

and morphologically distinctive.<br />

Other specimen examined: Hungary, additional isolate of strain selected as the<br />

holotype, recovered from inoculated Glycine max plants (MCA 2498 = <strong>CBS</strong> 119296<br />

= ICMP 18693).<br />

* <strong>Colletotrichum</strong> siamense Prihastuti, L. Cai & K.D. Hyde,<br />

Fungal Diversity 39: 98. 2009. Fig. 33.<br />

= <strong>Colletotrichum</strong> jasmini-sambac Wikee, K.D. Hyde, L. Cai & McKenzie, Fungal<br />

Diversity 46: 174. 2011.<br />

= <strong>Colletotrichum</strong> hymenocallidis Yan L. Yang, Zuo Y. Liu, K.D. Hyde & L. Cai,<br />

Fungal Diversity 39: 138. 2009.<br />

Descriptions of this <strong>species</strong> are provided by Prihastuti et al. (2009),<br />

Wikee et al. (2011), and Yang et al. (2009).<br />

Geographic distribution and host range: <strong>Colletotrichum</strong> siamense<br />

was originally described from coffee from Thailand, but our concept<br />

of this <strong>species</strong> is biologically and geographically diverse, found on<br />

many hosts across several tropical and subtropical regions.<br />

Genetic identification: ITS sequences do not reliably separate C.<br />

siamense from C. alienum, C. fructicola, or C. tropicale. <strong>The</strong>se<br />

<strong>species</strong> are best distinguished using CAL or TUB2.<br />

Notes: Yang et al. (2009) and Wikee et al. (2011) discussed genetic<br />

and morphological differences between C. siamense, C. jasminisambac,<br />

and C. hymenocallidis. However, both studies used a<br />

limited set of isolates within the C. <strong>gloeosporioides</strong> <strong>complex</strong>, making<br />

interpretation of the genetic differences difficult. <strong>The</strong> morphological<br />

differences they described are commonly seen as within-<strong>species</strong><br />

variation in other <strong>Colletotrichum</strong> spp. In our analysis, C. siamense is<br />

represented by 30 isolates from a wide range of hosts from several<br />

tropical regions, and forms a monophyletic clade that cannot be<br />

further subdivided genetically. Variation in cultural appearance is<br />

broad but in part this probably reflects the different conditions under<br />

which the isolates had been stored. Shape and size of appressoria,<br />

and the characteristically small conidia are similar in all isolates.<br />

Based on matching translation elongation factor (TEF) and<br />

TUB2 sequences, isolates referred by Rojas et al. (2010) to<br />

<strong>Colletotrichum</strong> sp. indet. 2 also represent C. siamense. Note that<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

TEF data was excluded from our phylogenetic analyses because<br />

the TEF gene tree was often incongruent with the trees from the<br />

other genes that we sequenced. For example, compare our isolate<br />

ICMP 17797 (GenBank GU174571) with isolates Rojas et al. (2010)<br />

cite as <strong>Colletotrichum</strong> sp. indet. 2, V1H1_1 (GenBank GU994297)<br />

and 7767 (GenBank GU994298).<br />

<strong>The</strong> C. siamense protologue designates the holotype as MFLU<br />

090230, but the culture derived from holotype as “BCC” with no<br />

specimen number. <strong>The</strong> ex-holotype culture is listed as BDP-I2 in<br />

Table 1 of Prihastuti et al. (2009) but not in the description of the<br />

<strong>species</strong>. Strain BDP-I2 was obtained from the authors (Prihastuti et<br />

al. 2009) for this study and deposited as ICMP 18578.<br />

Specimens examined: Australia, New South Wales, Murwillumbah, on Persea<br />

americana fruit rot, coll. L. Coates 23695, 1 Apr. 1990 (ICMP 12567); New South<br />

Wales, Muswellbrook, on Pistacia vera (DAR 76934 = ICMP 18574); Queensland,<br />

Mt Tamborine, on Persea americana fruit rot, coll. L. Coates T10-1, 1 Sep. 1993<br />

(ICMP 12565). China, Guangxi, Nanning, on Hymenocallis americana leaf spot,<br />

coll. Y.L. Yang CSSN2, 19 Jun. 2008 (ex-holotype culture of C. hymenocallidis<br />

− <strong>CBS</strong> 125378 = ICMP 18642); Guangxi Province, Nanning, on H. americana leaf,<br />

coll. Y.L. Yang CSSN3 (<strong>CBS</strong> 125379 = ICMP 18643). Nigeria, Ibadan, on Dioscorea<br />

rotundata seed, coll. M. Abang CgS2 (ICMP 18121); Ibadan, on D. rotundata seed,<br />

coll. M. Abang CgS6 (ICMP 18117); Ibadan, on Commelina sp. leaf, coll. M. Abang<br />

Cg29 (ICMP 18118). South Africa, on Carica papaya fruit, coll. L. Korsten PMS<br />

1 (ICMP 18739). on Persea americana, coll. L. Korsten Cg227 (ICMP 18570); on<br />

Persea americana, coll. L. Korsten Cg231 (ICMP 18569). Thailand, Chiang Mai,<br />

Mae Lod Village, on Coffea arabica berries, coll. H. Prihastuti BPD-I2, 12 Dec.<br />

2007 (ex-holotype culture of C. siamense – MFLU 090230 = ICMP 18578).<br />

Kanchanaburi, on Capsicum annuum, P.P. Than Ku4 (HKUCC 10884 = ICMP<br />

18575); Nakhonpathon, on C. annuum, coll. P.P. Than Ku8 (HKUCC 10881 = ICMP<br />

18618). USA, Florida, on Vitis vinifera leaf, coll. N. Peres ssgrape 10 (ICMP 18572);<br />

Florida, on Fragaria × ananassa crown, coll. N. Peres strawberry 6 (ICMP 18571);<br />

Florida, on V. vinifera leaf, coll. N. Peres DI-grape-6 (ICMP 18573); North Carolina,<br />

Wilkes County, on Malus domestica fruit, coll. T. Sutton LD Cg12 2001 (ICMP<br />

17795); North Carolina, Johnston County, on M. domestica fruit, coll. T. Sutton GD<br />

8 2002 (ICMP 17791); North Carolina, Johnston County, on M. domestica fruit, coll.<br />

T. Sutton GD 7 2002 (ICMP 17797); Alabama, on M. domestica fruit, coll. T. Sutton<br />

AL 1 2001 (ICMP 17785). Vietnam, Cu Chi District, Trung An Ward, on living leaves<br />

of Jasminium sambac, Jan. 2009, coll. Hoa Nguyen Thi LLTA–01 (ex-holotype<br />

culture of C. jasmini-sambac – <strong>CBS</strong> 130420 = ICMP 19118).<br />

* <strong>Colletotrichum</strong> theobromicola Delacr., Bull. Soc. Mycol.<br />

France. 31: 191. 1905. Fig. 34.<br />

= <strong>Colletotrichum</strong> fragariae A.N. Brooks, Phytopathology 21: 113. 1931.<br />

= <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. stylosanthis Munaut, Mycol. Res. 106: 591.<br />

2002.<br />

= <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. stylosanthis “f. sp. stylosanthis” (Munaut<br />

et al. 2002).<br />

= <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. stylosanthis “f. sp. guianensis” (Munaut et<br />

al. 2002).<br />

A modern description of this <strong>species</strong> is provided by Rojas et al.<br />

(2010).<br />

Geographic distribution and host range: Broadly distributed in<br />

tropical and subtropical regions on a wide range of hosts.<br />

Genetic identification: ITS sequences distinguish C. theobromicola<br />

from all other <strong>species</strong>.<br />

Notes: <strong>The</strong> ex-epitype culture of <strong>Colletotrichum</strong> fragariae, the exneotype<br />

culture of C. theobromicola, and the ex-holotype culture of<br />

C. <strong>gloeosporioides</strong> f. stylosanthis, selected by Buddie et al. (1999),<br />

Rojas et al. (2010), and Munaut et al. (2002) respectively, belong in<br />

a clade that we accept genetically as a single <strong>species</strong>. Also in this<br />

clade are authentic isolates of C. <strong>gloeosporioides</strong> f. stylosanthis<br />

“f. sp. stylosanthis” and C. <strong>gloeosporioides</strong> f. stylosanthis “f. sp.<br />

guianensis” (but see notes under C. <strong>gloeosporioides</strong> f. stylosanthis).<br />

167


Weir et al.<br />

Fig. 33. <strong>Colletotrichum</strong> siamense. A. ICMP 18642 (ex <strong>CBS</strong> 125378 – ex-holotype culture of C. hymenocallidis). B. ICMP 18578 (ex MFLU 090230 – ex-holotype culture of C.<br />

siamense). C. ICMP 12565. D. ICMP 18574 (ex DAR 76934). E. ICMP 18618 (ex HKUCC 10881). F. ICMP 18121. Cultures on PDA, 10 d growth from single conidia, from<br />

above and below.<br />

<strong>Colletotrichum</strong> theobromicola as accepted here contains<br />

several putatively specialised pathogens, including the pathogen of<br />

strawberry runners described by Brooks (1931) as C. fragariae, and<br />

168<br />

the pathogens of Stylosanthes referred to as C. <strong>gloeosporioides</strong> f.<br />

stylosanthis (Munaut et al. 2002). Future studies may show that<br />

the <strong>species</strong> should be segregated based on their pathogenicity


www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Fig. 34. <strong>Colletotrichum</strong> theobromicola. A. ICMP 17957 (ex MUCL 42294 – ex-holotype culture of C. <strong>gloeosporioides</strong> f. stylosanthis). B. ICMP 17927 (ex <strong>CBS</strong> 142.31 – ex-epitype<br />

culture of C. fragariae). C. ICMP 17958 (ex MUCL 42295). D. ICMP 17895. E. ICMP 18567. F. ICMP 18566. Cultures on PDA, 10 d growth from single conidia, from above and<br />

below.<br />

to specific hosts. See also notes under C. fragariae and C.<br />

<strong>gloeosporioides</strong> f. stylosanthis.<br />

Munaut et al. (2002) distinguished C. <strong>gloeosporioides</strong> f.<br />

stylosanthis from isolates they considered to represent C.<br />

169


Weir et al.<br />

Fig. 35. <strong>Colletotrichum</strong> ti. A. PDD 24881 – holotype. B. PDD 30206. C, D, F, H. ICMP 4832 – ex-holotype culture. E, G. ICMP 19444. A–B. Lesions on dried herbarium<br />

specimens. C–E. Appressoria. F. Conidia. G. Conidiogenous cells. H. Ascospores. Scale bars A = 1 mm, C = 20 µm. Scale bar of A applies to A–B, scale bar of C applies to C–H.<br />

<strong>gloeosporioides</strong> f. <strong>gloeosporioides</strong> because of 2 additional C’s at<br />

positions 93 and 94 in the ITS1 region, giving a string of 7 C’s at<br />

this position. This characteristic feature of the ITS-1 is found also in<br />

the ex-neotype isolate of C. theobromicola, the ex-epitype isolate<br />

of C. fragariae and all other isolates of C. theobromicola, although<br />

a few isolates have 3 additional C’s rather than 2. None of the<br />

170<br />

other isolates that we sampled from the C. <strong>gloeosporioides</strong> <strong>species</strong><br />

<strong>complex</strong> have this characteristic string of C’s.<br />

Rojas et al. (2010) provide a description for their concept of<br />

C. theobromicola, MacKenzie et al. (2008) for C. fragariae, and<br />

Irwin & Cameron (1978) for C. <strong>gloeosporioides</strong> f. stylosanthis “f. sp.<br />

stylosanthis” (as C. <strong>gloeosporioides</strong> Type A) and C. <strong>gloeosporioides</strong>


f. stylosanthis “f. sp. guianensis” (as C. <strong>gloeosporioides</strong> Type B).<br />

In cultural appearance the isolates we accept in this <strong>species</strong> are<br />

variable, from the very dark ex-neotype isolate of C. theobromicola<br />

to the slow-growing, pale coloured C. <strong>gloeosporioides</strong> f. stylosanthis<br />

“f. sp. guianensis”. None of the isolates that we examined formed<br />

perithecia in culture. All had conidia tapering slightly towards<br />

each end, this more pronounced towards the base, matching<br />

the description of C. fragariae by Gunnell & Gubler (1992), who<br />

regarded the conidial shape as distinctive for the <strong>species</strong>. Some<br />

of the isolates studied by Gunnell & Gubler (1992) were included<br />

in the study of MacKenzie et al. (2008), their genetic concept of C.<br />

fragariae matching ours.<br />

See also notes under C. fragariae and C. <strong>gloeosporioides</strong> f.<br />

stylosanthis.<br />

Specimens examined: Australia, Queensland, Townsville, on Stylosanthes viscosa,<br />

coll. J.A.G. Irwin 21365 (HM335), 1976 (ex-holotype culture of C. <strong>gloeosporioides</strong> f.<br />

stylosanthis − MUCL 42294 = ICMP 17957); Samford, on Stylosanthes guianensis,<br />

coll. J.A.G. Irwin 21398 (HM336), 1979 (MUCL 42295 = ICMP 17958); New South<br />

Wales, Olea europaea fruit, coll. V. Sergeeva UWS 128, 21 Apr. 2008 (ICMP 18566);<br />

New South Wales, O. europaea fruit, coll. V. Sergeeva UWS 130, 21 Apr. 2008<br />

(ICMP 18565); New South Wales, O. europaea fruit, coll. V. Sergeeva UWS 98, 8<br />

Apr. 2008 (ICMP 18567). Israel, on Limonium sp. leaf lesion, coll. S. Freeman P1<br />

(cited in Maymon et al. 2006) (ICMP 18576). Mexico, on Annona diversifolia, coll.<br />

R. Villanueva-Aroe Gro-7, Jul. 2003 (ICMP 17895). New Zealand, Kerikeri, on Acca<br />

sellowiana, coll. M.A. Manning MM317, 1 Feb. 2004 (ICMP 15445). Panama, Chiriqui<br />

Province, San Vicente, on <strong>The</strong>obroma cacao pod lesion, coll. E.J. Rojas ER08-9, Jan.<br />

2008 (<strong>CBS</strong> 125393 = ICMP 18650); Chiriqui Province, Escobal, on T. cacao leaf spot,<br />

coll. E.J. Rojas GJS 08-50, Jan. 2008 (ex-neotype culture of C. theobromicola − <strong>CBS</strong><br />

124945 = ICMP 18649). USA, Florida, Dover, Plant City, on Fragaria × ananassa, coll.<br />

S. MacKenzie 326-1, 1988 (ICMP 17099); Florida, Lake Alfred, on Quercus sp. leaf,<br />

coll. S. MacKenzie LA-oak-13, 2002 (ICMP 17100); Louisiana, on F. vesca, 1985 (IMI<br />

348152 = ICMP 17814); Florida, on F. × ananassa, coll. A.N. Brooks, 1931 (ex-epitype<br />

culture of C. fragariae − <strong>CBS</strong> 142.31 = ICMP 17927).<br />

* <strong>Colletotrichum</strong> ti B. Weir & P.R. Johnst., sp. nov.<br />

MycoBank MB563594. Figs 35, 36.<br />

Etymology: Based on the Maori name for Cordyline australis, tī.<br />

Holotype: New Zealand, Taupo, on Cordyline sp., coll. J.M. Dingley<br />

65187, Sep. 1965, PDD 24881; ex-holotype culture ICMP 4832.<br />

Leaf spots oblong to elliptic in shape, up to about 1 × 2 mm, sometimes<br />

coalescing when close together on a leaf, pale grey and necrotic in<br />

the centre with a reddish margin; acervuli numerous, base pale to<br />

dark grey, with scattered, dark brown setae about 50–80 µm long.<br />

Perithecia not seen on infected leaves. Freshly isolated colonies on<br />

Difco PDA 50–55 mm diam after 10 d, margin slightly irregular and<br />

feathery, aerial mycelium lacking from ex-holotype culture, when<br />

present fine, cottony, pale grey, surface of colony dark towards the<br />

centre, pale pinkish orange (7A6) towards margin, conidia forming<br />

over all parts of culture, mostly not associated with well differentiated<br />

acervuli, setae not observed; in reverse purple (12E3) near centre,<br />

orange outside, sometimes with concentric rings of grey pigment.<br />

Conidiogenous cells cylindric, mostly 15–25 × 3.5–4.5 µm, towards<br />

centre of colony arranged in closely packed palisade, towards margin<br />

the conidiophores with a much looser structure, irregularly branched,<br />

conidiogenous loci at apex and often also at septa. Conidia (11.5–<br />

)14–17.5(–23.5) × (4–)5–5.5(–7.5) µm (av. 16 × 5.2 µm, n = 53),<br />

cylindric, ends broadly rounded, sometimes tapering towards basal<br />

end. Appressoria often narrow-cylindric, often tapering towards<br />

apex, sometimes irregularly lobed. Perithecia developing in small<br />

numbers in culture after about 4 wk, solitary, scattered across plate,<br />

dark-walled, globose with well-developed, tapering ostiolar neck.<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Asci (60–)65–75(–78) × (10–) 11(–12) µm (av. 69.6 × 11 µm, n = 5),<br />

cylindric to subfusoid, 8–spored. Ascospores (14.5–)15.5–16.5(–19)<br />

× (4.5–)5–5.5(–6) µm (av. 15.9 × 5.2 µm, n=18), broad-cylindric,<br />

ends broadly rounded, not tapering to the ends, in side view mostly<br />

flat on one side, often slightly curved.<br />

Geographic distribution and host range: Known only from Cordyline<br />

spp. from New Zealand.<br />

Genetic identification: ITS sequences do not distinguish C. ti from<br />

C. aotearoa. <strong>The</strong> two <strong>species</strong> can be distinguished using TUB2 or<br />

GAPDH.<br />

Notes: A member of the Kahawae clade, this fungus causes a leaf<br />

spot of Cordyline spp. in New Zealand. It is genetically distinct<br />

from C. cordylinicola, described from Cordyline fruticosa from<br />

Thailand. Based on the published description of C. cordylinicola<br />

(Phoulivong et al. 2011) the two fungi are morphologically similar.<br />

Inoculation tests using culture ICMP 5285 when freshly isolated<br />

(J.M. Dingley, unpublished data), showed it to be pathogenic to<br />

Cordyline australis, forming spots on leaves 2 wk after inoculation,<br />

but causing no symptoms on apple, even after wounding.<br />

Although only four of the specimens examined have been<br />

compared genetically, all of the cited specimens examined match<br />

in terms of associated symptoms and conidial size and shape.<br />

A specimen from Cordyline banksii (PDD 78360) has narrower<br />

conidia, forms perithecia on the infected leaves, and perhaps<br />

represents a different <strong>species</strong>. Specimens accepted here as C. ti<br />

were referred to Glomerella cingulata by Laundon (1972).<br />

<strong>The</strong> appearance in culture varies between isolates. <strong>The</strong> J.M.<br />

Dingley cultures, first isolated in the mid-1960’s, have dense, felted<br />

aerial mycelium and limited conidial production; one has a much<br />

slower growth rate than the more recent collections.<br />

Other specimens examined: New Zealand, Auckland, on Cordyline australis, coll.<br />

J.M. Dingley 6653, Mar. 1966 (PDD 30206; ICMP 5285); Taranaki, New Plymouth,<br />

Duncan and Davies Nursery, on C. australis × C. banksii leaf spots, coll. G.F.<br />

Laundon LEV 3343, 26 May 1969 (PDD 50634); Taranaki, New Plymouth, Duncan<br />

and Davies Nursery, on C australis × C. banksii leaf spots, coll. G.F. Laundon, 26<br />

May 1969 (PDD 26775); Waikato, Cambridge, Anton Nursery, on C. australis leaf<br />

spots, coll. L.A. Houghton, 23 Jul. 1992 (PDD 61219; ICMP 19444).<br />

* <strong>Colletotrichum</strong> tropicale E.I. Rojas, S.A. Rehner &<br />

Samuels, Mycologia 102: 1331. 2010. Fig. 37.<br />

Rojas et al. (2010) provide a description.<br />

Geographic distribution and host range: Rojas et al. (2010) noted<br />

that C. tropicale has been isolated from a wide range of hosts<br />

in forests in tropical America, from rotting fruit as well as leaf<br />

endophytes. We include also an isolate from tropical Japan, from<br />

Litchi chinensis leaves.<br />

Genetic identification: ITS sequences do not separate C. tropicale<br />

from some C. siamense or some C. queenslandicum isolates.<br />

<strong>Colletotrichum</strong> tropicale is best distinguished using TUB2, CHS-1,<br />

GS, or SOD2.<br />

Notes: <strong>Colletotrichum</strong> tropicale is genetically close to C. siamense<br />

and the two <strong>species</strong> share a number of morphological features;<br />

slow growth in culture, short and broad conidia with broadly<br />

rounded ends and often slightly constricted near the centre, and<br />

simple appressoria.<br />

171


Weir et al.<br />

Fig. 36. <strong>Colletotrichum</strong> ti. A. ICMP 19444. B. ICMP 4832 – ex-holotype culture. C. ICMP 5285. Cultures on PDA, 10 d growth from single conidia, from above and below.<br />

Fig. 37. <strong>Colletotrichum</strong> tropicale. A. ICMP 18653 (ex <strong>CBS</strong> 124949 – ex-holotype culture). B. ICMP 18651 (ex <strong>CBS</strong> 124943). C. ICMP 18672 (ex MAFF 239933). Cultures on<br />

PDA, 10 d growth from single conidia, from above and below.<br />

172


Specimens examined: Japan, Okinawa, on Litchi chinensis leaf (MAFF 239933 =<br />

ICMP 18672). Panama, Barro Colardo Monument, on <strong>The</strong>obroma cacao leaf, coll.<br />

E.I. Rojas, L.C. Mejía, Z. Maynard 5101, 2008 (ex-holotype culture – <strong>CBS</strong> 124949<br />

= ICMP 18653); Escobal, Chiriqui, on Annona muricata fruit rot, coll. E.I. Rojas GJS<br />

08-42 (<strong>CBS</strong> 124943 = ICMP 18651).<br />

* <strong>Colletotrichum</strong> xanthorrhoeae R.G. Shivas, Bathgate &<br />

Podger, Mycol. Res. 102: 280. 1998. Fig. 38.<br />

Shivas et al. (1998) provide a description. One of the isolates<br />

we examined (ICMP 17820) formed fertile perithecia in culture,<br />

a feature not mentioned in the original description. Perithecia<br />

are dark-walled, globose with a prominent, narrow neck, wall<br />

comprising several layers of pseudoparenchymatous cells 8–15<br />

µm diam, with several layers of densely packed hyphae outside<br />

this. Asci 75–100 × 10–12 µm, 8-spored. Ascospores (17–)18.5–<br />

20(–22) × (5–)5.5–6 µm (av. 19.4 × 5.6 µm, n = 24), more or less<br />

elliptic, tapering to narrow, rounded ends, in side view flattened on<br />

one side, but generally not curved.<br />

Genetic identification: ITS sequences distinguish C. xanthorrhoeae<br />

from all other <strong>species</strong>.<br />

Notes: This pathogen of Xanthorrhoea has a distinctive morphology,<br />

with a very slow growth rate in culture and large conidia which taper<br />

towards the basal end. <strong>The</strong> ascospore shape is distinct to that of<br />

most taxa within the C. <strong>gloeosporioides</strong> group, which typically have<br />

bent or curved ascospores.<br />

Specimens examined: Australia, Western Australia, Melville, on Xanthorrhoea<br />

preissii leaf spots, coll. F.D. Podger, Jan. 1994 (ex-holotype culture − BRIP 45094<br />

= ICMP 17903 = <strong>CBS</strong> 127831); Queensland, Cunningham’s Gap, Main Ranges<br />

National Park, on Xanthorrhoea sp. leaf spot (IMI 350817a = ICMP 17820).<br />

DISCUSSION<br />

<strong>The</strong> <strong>species</strong> that we accept in the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong><br />

<strong>species</strong> <strong>complex</strong> together form a strongly supported clade in the<br />

<strong>Colletotrichum</strong> ITS gene tree (fig. 1 in Cannon et al. 2012, this issue).<br />

All <strong>species</strong> are micro-morphologically typical of C. <strong>gloeosporioides</strong><br />

sensu von Arx (1970) and Sutton (1992). However, morphology<br />

alone cannot unequivocally place an isolate in this <strong>complex</strong>,<br />

making the ITS particularly important for identification at the<br />

<strong>species</strong> <strong>complex</strong> level in <strong>Colletotrichum</strong>. For example, members of<br />

the C. boninense <strong>species</strong> <strong>complex</strong> (Damm et al. 2012b, this issue)<br />

and C. cliviae (Yang et al. 2009) are micro-morphologically similar<br />

to <strong>species</strong> in the C. <strong>gloeosporioides</strong> <strong>complex</strong> but genetically distinct<br />

(Cannon et al. 2012, this issue). <strong>The</strong> utility of ITS sequences is<br />

enhanced by their strong representation in GenBank, but this can<br />

also be a problem. Nilsson et al. (2006) summarised the frequency<br />

of inaccurately annotated data in GenBank. <strong>The</strong> diversity of<br />

taxonomic concepts around the name C. <strong>gloeosporioides</strong> makes<br />

this a particular problem. This is illustrated by the phylogeny<br />

presented by Hyde et al. (2010), based on GenBank accessions<br />

of ITS sequences identified as C. <strong>gloeosporioides</strong> and Glomerella<br />

cingulata, that shows the taxa represented belong to many <strong>species</strong><br />

in different <strong>Colletotrichum</strong> <strong>species</strong> <strong>complex</strong>es. See notes under C.<br />

boehmeriae, C. crassipes, and C. kahawae subsp. kahawae for<br />

specific examples of misidentified GenBank accessions.<br />

<strong>The</strong> <strong>species</strong> we accept are based on a phylogenetic <strong>species</strong><br />

concept, all <strong>species</strong> forming strongly supported, monophyletic<br />

clades within our multigene phylogenies. However, not all terminal<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

clades are recognised as named <strong>species</strong>. In most cases any well<br />

supported, within-<strong>species</strong> phylogenetic structure evident in the<br />

multi-gene phylogeny is not resolved consistently in all gene trees.<br />

This lack of congruence between gene trees is a signal that the<br />

diversity being sampled is below the <strong>species</strong> level, according to<br />

the logic of the genealogical concordance phylogenetic <strong>species</strong><br />

recognition (GCPSR) concept (Taylor et al. 2000). Although the<br />

concatenation of gene sequences is a convenient way to present<br />

multigene data, it masks discordance between individual gene<br />

phylogenies. An alternative method, using a <strong>species</strong>-tree approach<br />

(Figs 3, 4B, 5B) combines multi-gene data from multiple isolates<br />

hypothesised to represent a single <strong>species</strong>, so that the evolutionary<br />

history of the <strong>species</strong> rather than that of individual isolates is<br />

estimated. Fig. 3, shows the results of such an analysis for the C.<br />

<strong>gloeosporioides</strong> <strong>complex</strong>, Figs 4B and 5B show relationships within<br />

the Musae and Kahawae clades respectively, at an expanded<br />

scale. Posterior probabilities for some of the speciation events<br />

are low, particularly within the Musae and Kahawae clades. This<br />

may be because although the <strong>species</strong>-trees algorithms account<br />

for incomplete lineage sorting (Heled & Drummond 2010, Chung<br />

& Ané 2011), most do not compensate for horizontal gene transfer,<br />

reassortment, or introgression. Hybridisation could also result<br />

in discordant gene phylogenies. Hybrids are known in the C.<br />

acutatum <strong>complex</strong>, e.g. Glomerella acutata, a hybrid formed by<br />

crossing C. acutatum and C. fioriniae strains in the laboratory,<br />

and a putative hybrid strain between the same two <strong>species</strong> that<br />

had been collected from terminal crook disease on Pinus in New<br />

Zealand, where both <strong>species</strong> occur in nature (Damm et al. 2012a,<br />

this issue). Hybrids also form in the C. <strong>gloeosporioides</strong> <strong>complex</strong>,<br />

e.g. the Carya and Aeschynomene populations discussed by Cisar<br />

et al. (1994), more or less genetically equivalent to our <strong>species</strong><br />

within the C. <strong>gloeosporioides</strong> <strong>complex</strong>.<br />

Our taxonomic conclusions are based, of necessity, on the<br />

limited set of genes sampled. Potentially more powerful genes,<br />

such as ApMAT and Apn25L (Silva et al. 2012a) may provide<br />

finer resolution within the <strong>species</strong>-level clades that we recognise.<br />

However, even with these potentially more informative genes, the<br />

low levels of genetic divergence across the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> may always provide a technical challenge (Silva et al.<br />

2012a). <strong>The</strong> low level of diversity within this <strong>species</strong> <strong>complex</strong> is<br />

reflected by the branch lengths in fig. 2, Cannon et al. (2012, this<br />

volume), and is especially true across the Musae clade, where<br />

average pairwise identity between all isolates treated in our 5<br />

gene alignment is 98.6 %. Pairwise identity between isolates of<br />

C. siamense and C. theobromicola, two <strong>species</strong> showing strong<br />

within-<strong>species</strong> phylogenetic structure, are 99.4 % and 99.6%<br />

respectively. This suggests that the <strong>species</strong> recognised within the<br />

C. <strong>gloeosporioides</strong> <strong>complex</strong> are very recently evolved and Silva et<br />

al. (2012b) provide data supporting this. <strong>The</strong>ir hypothesis of recent<br />

evolution of host-specialised <strong>Colletotrichum</strong> populations from more<br />

generalist fungi was also invoked in relation to the C. acutatum<br />

<strong>complex</strong> by Lardner et al. (1999) using the “episodic selection”<br />

framework of Brasier (1995).<br />

Several of the <strong>species</strong> we accept contain isolates with divergent<br />

lifestyles, for example C. aotearoa, C. clidemiae, C. kahawae, and<br />

C. theobromicola. Each of these <strong>species</strong> includes isolates capable<br />

of causing specific diseases. In the case of C. kahawae, recent<br />

pathogenicity tests have shown that only some isolates are able<br />

to cause coffee berry disease (Silva et al. 2012a, Silva & Weir,<br />

unpubl. data) and that these isolates can be distinguished using GS<br />

sequences (this study), Apn25L and MAT1-2-1 (Silva et al. 2012b).<br />

Because of the well understood pathogenicity of isolates within<br />

173


Weir et al.<br />

Fig. 38. <strong>Colletotrichum</strong> xanthorrhoeae. ICMP 17903 (ex BRIP 45094 – ex-holotype culture). A. Cultures on PDA, 10 d growth from single conidia, from above and below. B.<br />

Culture on PDA at 4 wk showing sectoring with variation in pigmentation and growth form. C–D. Asci and ascospores. E. Perithecial wall in squash mount. Scale bar C = 20<br />

µm. Scale bar of C applies to C–E.<br />

174


C. kahawae, the biosecurity importance of coffee berry disease,<br />

and the ability to distinguish the disease-causing isolates using<br />

carefully selected genetic markers, we recognise the diseasecausing<br />

isolates taxonomically at the subspecific level. Future study<br />

of the comparative pathogenicity of isolates within C. aotearoa, C.<br />

clidemiae, and C. theobromicola may reveal genetically distinct,<br />

host-specialised pathogenic populations within these <strong>species</strong> that<br />

future workers may also choose to recognise taxonomically.<br />

<strong>The</strong> classification we accept here is deliberately taxonomically<br />

conservative, minimising nomenclatural changes. This reflects<br />

continuing uncertainty about sensible <strong>species</strong> limits within the<br />

C. <strong>gloeosporioides</strong> <strong>complex</strong> that relate to low levels of genetic<br />

divergence across the <strong>complex</strong>, gene selection, isolate selection,<br />

and a lack of understanding of the mechanisms driving <strong>species</strong><br />

and population divergence amongst these fungi. For example,<br />

the two haplotype subgroups of C. fructicola are not distinguished<br />

taxonomically because collectively they form a monophyletic clade,<br />

both subgroups include sets of isolates with similar geographic and<br />

host diversity, and there is no practical need to distinguish them<br />

taxonomically.<br />

Molecular tools are increasingly being used for day-to-day<br />

identification by biosecurity officers and plant pathology researchers,<br />

providing a need for both a taxonomy that closely reflects groups that<br />

are resolved genetically, as well as simple and reliable protocols for<br />

identifying those taxa. <strong>The</strong> internal transcribed spacer region (ITS)<br />

has been proposed as the official fungal barcoding gene (Schoch<br />

et al. 2012). Although ITS is useful at the <strong>species</strong> <strong>complex</strong> level, it<br />

does a poor job of resolving <strong>species</strong> within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong>, resolving only 10 of 22 accepted <strong>species</strong>. This reflects<br />

the low number of base changes in the ITS region across the C.<br />

<strong>gloeosporioides</strong> <strong>complex</strong>; <strong>species</strong> often distinguished by only one<br />

or two base changes. In some cases, chance variation in the ITS<br />

sequence within or between <strong>species</strong> means that some <strong>species</strong><br />

cannot be distinguished (Fig. 6). Examples of taxa with identical<br />

ITS sequences include C. clidemiae, C. tropicale, C. ti and some<br />

C. siamense isolates; C. fructicola and some C. siamense isolates;<br />

and C. alienum, C. aenigma and some C. siamense isolates.<br />

Protein-coding genes and their introns often have more<br />

variation than ITS, and the need for secondary barcodes based<br />

on these kinds of genes has been discussed in relation to some<br />

groups of fungi (Fitzpatrick et al. 2006, Aguileta et al. 2008, Weir<br />

& Johnston 2011). Ideally, one of the seven protein coding genes<br />

that were used in this study could be proposed as a secondary<br />

barcode to obtain an accurate identification of <strong>species</strong> within the<br />

C. <strong>gloeosporioides</strong> <strong>complex</strong>. A preliminary analysis of the genes<br />

performance as barcodes was conducted as part of Cai et al.<br />

(2009) with GAPDH, CAL, and ACT performing well, but CHS-<br />

1, ITS, and TEF (EF1α) poorly. However, the analysis (Cai et al.<br />

2009) included only five <strong>species</strong> within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong>, the Musae and Kahawae clades being treated at the level<br />

of <strong>species</strong>. With the final classification presented here, none of<br />

the genes we analysed provides an effective barcode ont its own<br />

across the entire <strong>complex</strong>. Of the single genes, TUB2, GS, and<br />

GAPDH are amongst the most effective at distinguishing <strong>species</strong>.<br />

However, C. clidemiae is polyphyletic in the TUB2 gene tree and<br />

GS sequences are needed to distinguish C. fructicola and C.<br />

alienum. With GS, C. aotearoa, C. kahawae subsp. ciggaro, and C.<br />

siamense are paraphyletic. GAPDH is the easiest of all the genes<br />

tested to amplify and sequence, however when using this gene GS<br />

sequences are needed to distinguish C. fructicola from C. alienum<br />

and C. aeschynomenes from C. siamense, and C. tropicale is<br />

paraphyletic. In the <strong>species</strong> descriptions we provide notes on which<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

genes are the best for genetic identifications, and in Table 4 these<br />

are summarised for all <strong>species</strong> and genes. For <strong>species</strong> represented<br />

by a single or only a few isolates the <strong>species</strong> boundaries may not<br />

be accurate, we recommend two protein-coding genes in addition<br />

to ITS for sequence-based identifications. A meta-analysis of<br />

DNA barcodes across the whole genus will be required to find the<br />

combination of genes that are effective for all <strong>species</strong> of the genus<br />

that distinguish all <strong>Colletotrichum</strong> <strong>species</strong>.<br />

Several studies have shown that cultural morphology can be<br />

useful for grouping isolates when they are sampled at a local or<br />

regional level (e.g. Johnston & Jones 1997, Prihastuti et al. 2009).<br />

However, our experience is that such groups often break down<br />

when the geographic sample within a clade is extended to a global<br />

scale. Many of the <strong>species</strong> we accept have few or no diagnostic<br />

morphological or cultural features that can be consistently and<br />

reliably used to identify them. Our morphological examinations<br />

were confined to cultures on Difco PDA agar plates, and we will<br />

have missed any features that develop solely in association with<br />

plant material. In addition, the cultures we used have been sourced<br />

from different labs and collections from around the world, many<br />

with no information on storage history. Storage history and method<br />

has a major impact on the appearance of <strong>Colletotrichum</strong> in culture.<br />

Cultures can become “stale” during storage, losing the ability to<br />

produce pigments, the aerial mycelium often becoming very dense<br />

and felted, and losing the ability to form well-differentiated acervuli,<br />

conidia, or perithecia. In some clades, even freshly isolated cultures<br />

are highly variable, forming distinct sectors with differences in the<br />

production of pigment, aerial mycelium, acervuli, and conidia.<br />

Some isolates form two very different cultural types from single<br />

conidia or ascospores derived from colonies themselves started<br />

from single ascospores. Figure 27F shows single ascospore<br />

cultures from an isolate of C. kahawae subsp. ciggaro. One has<br />

the typical appearance of cultures of this fungus isolated from the<br />

field. <strong>The</strong> other, with a uniform, dense layer of conidia across the<br />

colony surface without well differentiated acervuli and more or less<br />

no aerial mycelium, is common from single ascospore isolates in<br />

culture, but rarely found in cultures isolated directly from the field.<br />

This kind of variation, and that revealed from sectoring during<br />

colony growth, makes morphological variation difficult to interpret<br />

for accurate identification.<br />

Many of the <strong>species</strong> recognised in this work remain poorly<br />

understood in terms of their pathogenicity and host preference.<br />

This in part reflects a lack of certainty about the biological<br />

relationship between the fungi and the plants from which they<br />

were isolated. Species that are pathogenic on one host can<br />

also be isolated from others following opportunistic colonisation<br />

of senescing tissue, such as the C. salicis example discussed<br />

by Johnston (2000, as Glomerella miyabeana). <strong>The</strong> multiple<br />

<strong>Colletotrichum</strong> spp. associated with a single host are likely to<br />

have a variety of life styles: primary pathogens of healthy tissue,<br />

<strong>species</strong> with the ability to invade and cause minor disease<br />

when the host plant is under stress, <strong>species</strong> that develop latent<br />

infections and fruit only following senescence of the host tissue<br />

or ripening of host fruit and endophytic <strong>species</strong> that sporulate<br />

only following host tissue death. <strong>The</strong> combination of this range of<br />

distinct life styles, the fact that several <strong>Colletotrichum</strong> spp. may<br />

become established on a single host, and the ability of most of<br />

these <strong>species</strong> to also establish on a range of other hosts, has<br />

been a large part of the confusion surrounding <strong>species</strong> limits<br />

within <strong>Colletotrichum</strong>.<br />

In some cases, apparently clear differences in pathogenicity<br />

of isolates in the C. <strong>gloeosporioides</strong> <strong>complex</strong> are not reflected<br />

175


Weir et al.<br />

Table 4. Performance of individual genes at resolving <strong>species</strong> within the <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>. Y – <strong>species</strong><br />

distinguished from all others. N – <strong>species</strong> not distinguished from all others. N* – distinguishes at the sub<strong>species</strong> level.<br />

Species ITS GAPDH CAL TUB2 ACT CHS-1 GS SOD<br />

C. fructicola N N Y N N Y Y Y<br />

C. nupharicola N Y Y Y Y Y Y Y<br />

C. alienum N N Y N N Y Y Y<br />

C. musae Y Y Y Y Y Y Y Y<br />

C. aenigma N Y Y Y N Y Y Y<br />

C. siamense N N Y Y N N N N<br />

C. aeschynomenes N N N Y N Y Y Y<br />

C. tropicale N N N Y Y N Y Y<br />

C. queenslandicum N Y Y Y N N Y N<br />

C. salsolae N Y Y Y Y Y N Y<br />

C. asianum Y Y Y Y Y Y Y Y<br />

C. <strong>gloeosporioides</strong> Y Y Y Y Y Y Y Y<br />

C. alatae Y Y Y Y Y Y Y Y<br />

C. theobromicola Y Y Y Y Y Y Y Y<br />

C. xanthorrhoeae Y Y Y Y Y Y Y Y<br />

C. horii Y Y Y Y Y Y Y Y<br />

C. aotearoa N N Y Y N Y Y N<br />

C. ti N Y Y Y N Y Y Y<br />

C. kahawae N Y N Y Y N N* N<br />

G. cingulata “f. sp. camelliae” Y N Y Y Y N Y Y<br />

C. clidemiae N N N N Y Y Y N<br />

C. psidii Y Y Y Y N Y Y Y<br />

C. cordylinicola Y Y Y Y N Y Y Y<br />

genetically. For example, the fungi referred to as C. <strong>gloeosporioides</strong><br />

f. stylosanthis “f. sp. guianensis” and C. <strong>gloeosporioides</strong> f.<br />

stylosanthis “f. sp. stylosanthis”, are reportedly associated with two<br />

distinct diseases of Stylosanthes (Irwin & Cameron 1978; Munaut<br />

et al. 2002), but both taxa genetically match C. theobromicola and<br />

are here placed in synonymy with C. theobromicola. It is possible<br />

that screening additional genes across a set of isolates from<br />

Stylosanthes with known pathogenicity will reveal one or more<br />

genes that generate a phylogeny that correlates with pathogenicity.<br />

This is the case with another specialised pathogen, C. kahawae.<br />

Originally described as a pathogen of green coffee berries, almost<br />

genetically identical isolates have subsequently been found on a<br />

wide range of hosts (see notes under C. kahawae). <strong>The</strong> isolates<br />

from other hosts are not pathogenic to coffee berries (Silva et al.<br />

2012b). <strong>The</strong> difference in pathogenicity correlates with a genetic<br />

difference in the GS gene, and we taxonomically recognise this<br />

biologically specialised population at the sub<strong>species</strong> level. A similar<br />

approach could potentially be taken for other biologically distinct<br />

populations within a genetically strongly supported <strong>species</strong>.<br />

Despite the epitypification of C. <strong>gloeosporioides</strong> in 2008, web<br />

search hits on the name C. <strong>gloeosporioides</strong> from papers published<br />

over the past 12 mo show that many authors will continue to use<br />

the name in the sense of the C. <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong>,<br />

presumably regarding this level of identification as sufficient for their<br />

research. All of the isolates that we accept in the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> share the string 5’–GGGCGGGT–3’ about 139–142 bases<br />

after the ITS1F primer binding site. Based on a comparison with<br />

GenBank data, this string appears to be specific to isolates that<br />

we would accept as members of the C. <strong>gloeosporioides</strong> <strong>complex</strong>.<br />

176<br />

Several authors have developed PCR-based, rapid identification<br />

tools for distinguishing members of the C. <strong>gloeosporioides</strong> <strong>complex</strong><br />

from members of the C. acutatum <strong>species</strong> <strong>complex</strong>. This has been<br />

prompted because some members of the C. acutatum <strong>complex</strong><br />

have conidia without the acute ends characteristic of this <strong>species</strong> as<br />

described by Simmonds (1965), and have at times been confused<br />

with C. <strong>gloeosporioides</strong> (Damm et al. 2012, this issue). Primers<br />

reportedly specific to C. <strong>gloeosporioides</strong> include the CgInt primer for<br />

ITS (Mills et al. 1992). In our data set this primer sequence is found<br />

in C. <strong>gloeosporioides</strong> s. str., C. fructicola, and C. siamense but all<br />

of the other taxa that we recognise within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> have one or more bases not matching the CgInt primer.<br />

<strong>The</strong> practical impact of these differences will depend in part on<br />

the position of the mismatch and stringency of the PCR reaction.<br />

Talhinas et al. (2005) discussed the TBCG primer for β-tubulin,<br />

and this is found within all of our taxa within the C. <strong>gloeosporioides</strong><br />

group except C. musae and C. asianum. Liu et al. (2011) describe<br />

characteristic RFLP bands from glutamine synthetase using the<br />

restriction enzyme Pst1. Based on our sequences, this method will<br />

generate the characteristic C. <strong>gloeosporioides</strong> bands reported by Liu<br />

et al. (2011) for C. aenigma, C. alienum, C. aotearoa, C. asianum,<br />

C. clidemiae, C. cordylinicola, C. fructicola, C. <strong>gloeosporioides</strong> s.<br />

str., C. horii, C. queenslandicum, C. salsolae, C. siamense, C. ti,<br />

and C. tropicale. Different banding patterns will be produced by C.<br />

aeschynomenes (band sizes 253, 316, 388), the two C. kahawae<br />

subspp. (band sizes 112, 388, 457), G. cingulata “f. sp. camelliae”<br />

(band sizes 51, 112, 337, 457), and C. musae (band sizes 388,<br />

552), but none match the bands reported for C. acutatum by these<br />

authors.


Comparison of our data with gene sequences reported as<br />

C. <strong>gloeosporioides</strong> in recent papers allows most to be placed<br />

with confidence in one of the <strong>species</strong> that we accept. <strong>The</strong>re are<br />

exceptions, such as the pecan-associated isolates from Liu et al.<br />

(2011), and the pistachio-associated isolates reported by Yang et<br />

al. (2011), both of which appear to represent undescribed <strong>species</strong><br />

within the C. <strong>gloeosporioides</strong> <strong>complex</strong>. Clearly, more <strong>species</strong> remain<br />

to be described within the C. <strong>gloeosporioides</strong> <strong>complex</strong>. In addition,<br />

taxonomic issues still to be resolved amongst the <strong>species</strong> discussed<br />

in this paper include the relationship between G. cingulata “f. sp.<br />

camelliae” and C. camelliae, the identity of the cotton pathogens<br />

referred to C. gossypii, the identity of the cassava pathogens referred<br />

to C. manihotis, the relationship between C. aeschynomenes and C.<br />

<strong>gloeosporioides</strong> “f. sp. jussiaeae”, whether the various yam diseases<br />

discussed in the literature are all caused by C. alatae, and whether<br />

the isolates of C. aotearoa from Meryta leaf spots form a biologically<br />

distinct population. A more general question relates to better<br />

understanding the frequency of hybrids within the C. <strong>gloeosporioides</strong><br />

<strong>complex</strong> and the impact of this on the interpretation of the phylogenies<br />

within the <strong>complex</strong>. <strong>The</strong> impact of hybridisation on the evolution of<br />

disease specialised populations has barely been explored.<br />

ACKNOWLEDGEMENTS<br />

This study was funded by AGMARDT (<strong>The</strong> Agricultural and Marketing Research<br />

and Development Trust) grant no. 892, the New Zealand Ministry of Science and<br />

Innovation through backbone funding of the “Defining New Zealand’s Land Biota”<br />

programme, and the New Zealand Tertiary Education Commission through the<br />

“Molecular diagnostics: capitalising on a million DNA barcodes” project. Shaun<br />

Pennycook, Landcare Research, provided valuable nomenclatural advice, Duckchul<br />

Park and Paula Wilkie, Landcare Research, provided valuable technical assistance.<br />

Diogo Silva, Centro de Investigação das Ferrugens do Cafeeiro, Portugal, carried<br />

out the C. kahawae pathogenicity tests. David Cleland, ATCC, provided DNA<br />

sequences for ATCC 96723. This study has been possible only through the provision<br />

of cultures by the DAR, ICMP, <strong>CBS</strong>, CABI, and MAFF culture collections, as well<br />

as by many individual researchers — Matthew Abang, Deutsche Sammlung von<br />

Mikroorganismen und Zellkulturen, Braunschweig, Germany; Robert Barreto,<br />

Universidade Federal de Viçosa, Brazil; George Carroll, University of Oregon,<br />

USA; Dana Berner, Foreign Disease-Weed Science Research Unit, USDA-ARS, Ft.<br />

Detrick, USA; Lindy Coates, Brad McNeil and Roger Shivas, Queensland Primary<br />

Industries and Fisheries, Indooroopilly, Queensland, Australia; W.E. Copes, USDA-<br />

ARS Small Fruit Research Unit, Poplarville, Massachusetts, USA; Kerry Everett<br />

and Mike Manning, Plant and Food Research, Auckland, New Zealand; David Farr,<br />

Gary Samuels and Steve Rehner, Systematic Mycology & Microbiology Laboratory<br />

USDA, Beltsville, USA; Steve Ferriera and K. Pitz, University of Hawaii, USA; Stan<br />

Freeman, Dept. of Plant Pathology and Weed Research, <strong>The</strong> Volcani Center, Israel;<br />

Hideo Ishii, National Institute for Agro-Environmental Sciences, Tsukuba, Japan;<br />

Sucheta Joshee and Nick Waipara, Landcare Research, Auckland, New Zealand;<br />

Lisa Korsten, University of Pretoria, South Africa; Lei Cai, State Key Laboratory of<br />

Mycology, Institute of Microbiology, Chinese Academy of Sciences, Beijing, China;<br />

Frank Louws and Turner Sutton, North Carolina State University, USA; Stephen<br />

Mackenzie and Natalia Peres, University of Florida, USA; Nelson Massola, Escola<br />

Superior de Agricultura “Luiz de Queiroz”, Brazil; Jean-Yves Meyer, Department of<br />

Research Ministère d’Education, de l’Enseignement Supérieur et de la Recherche<br />

Gouvernement de Polynésie française, French Polynesia; François Munaut,<br />

Université Catholique de Louvain, Belgium; Vera Sergeeva, University of Western<br />

Sydney, New South Wales, Australia; Amir Sharon, Tel Aviv University, Israel; Po<br />

Po Than, Chinese Academy of Forestry, Kunming, China; Ramon Villanueva-Aroe,<br />

University of València, Spain; Jing-Ze Zhang, Zhejiang University, Hangzhou, China.<br />

REFERENCES<br />

Aa HA van der (1978). A leaf spot of Nymphaea alba in the Netherlands. Netherlands<br />

Journal of Plant Pathology 84: 109–115.<br />

Abang MM, Winter S, Green KR, Hoffmann P, Mignouna HD, Wolf GA (2002).<br />

Molecular identification of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> causing anthracnose<br />

of yam in Nigeria. Plant Pathology 51: 63–71.<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Abang MM, Winter S, Mignouna HD, Green KR, Asiedu R (2003). Molecular<br />

taxonomic, epidemiological and population genetic approaches to understanding<br />

yam anthracnose disease. African Journal of Biotechnology 2: 486–496.<br />

Afanador-Kafuri L, Minz D, Maymon M, Freeman S (2003). Characterisation of<br />

<strong>Colletotrichum</strong> isolates from tamarillo, passiflora, and mango in Colombia and<br />

identification of a unique <strong>species</strong> from the genus. Phytopathology 93: 579–587.<br />

Aguileta G, Marthey S, Chiapello H, Lebrun MH, Rodolphe F, et al. (2008). Assessing<br />

the Performance of Single-Copy Genes for Recovering Robust Phylogenies<br />

Systematic Biology 57: 613–627.<br />

Alahakoon PW, Brown AE, Sreenivasaprasad S (1994). Genetic characterisation<br />

of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> isolates obtained from mango. International<br />

Journal of Pest Management 40: 225–229.<br />

Alfieri Jr SA, Langdon KR, Wehlburg C, Kimbrough JW (1984). Index of Plant<br />

Diseases in Florida (Revised). Florida Department of Agriculture and Consumer<br />

Services, Division of Plant Industries Bulletin 11: 1–389.<br />

Archibald JK, Mort ME, Crawford DJ (2003). Bayesian inference of phylogeny: a<br />

non-technical primer. Taxon 52: 187–191<br />

Arx JA von (1957). Die Arten der Gattung <strong>Colletotrichum</strong> Cda. Phytopathologische<br />

Zeitschrift 29: 413–468.<br />

Arx JA von (1970). A revision of the fungi classified as Gloeosporium. Bibliotheca<br />

Mycologica 24: 1–203.<br />

Arx JA von, Müller E (1954). Die Gattung der amerosporen Pyrenomyceten.<br />

Beiträge zur Kryptogamenflora der Schweiz 11(1): 1–434.<br />

Bailey JA, Nash C, Morgan LW, O’Connell RJ, TeBeest DO (1996). Molecular<br />

taxonomy of <strong>Colletotrichum</strong> <strong>species</strong> causing anthracnose on the Malvaceae.<br />

Phytopathology 86: 1076–1083.<br />

Beever RE (1984). Observations on puka (Meryta sinclairii) on the Chickens Islands,<br />

with an assessment of the hypothesis that it was transferred there by the Maori.<br />

Tane 30: 77–92.<br />

Berkeley MJ (1874). Notices of North American fungi. Grevillea 3: 1–17.<br />

Berner DK, Bruckart WL, Cavin CA, Michael JL, Carter ML, Luster DG (2009).<br />

Best linear unbiased prediction of host-range of the facultative parasite<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp. salsolae, a potential biological control<br />

agent of Russian thistle. Biological Control 51: 158–168.<br />

Bogo A, Casa RT, Rufato L, Gonçalves MJ (2012). <strong>The</strong> effect of hail protection nets<br />

on Glomerella leaf spot in ‘royal Gala’ apple. Crop Protection 31: 40–44<br />

Boyette CD, Templeton GE, Smith JR (1979). Control of waterprimrose (Jussiaea<br />

decurrens) and northern jointvetch (Aeschynomene virginica) with fungal<br />

pathogens. Weed Science 27: 497–501.<br />

Brasier CM (1995). Episodic selection as a force in fungal microevolution, with<br />

special reference to clonal speciation and hybrid introgression. Canadian<br />

Journal of Botany 73 (Supplement 1): S1213–S1221.<br />

Brooks AN (1931). Anthracnose of strawberry caused by <strong>Colletotrichum</strong> fragariae n.<br />

sp. Phytopathology 21: 739–744.<br />

Buddie AG, Martínez-Culebras P, Bridge PD, García MD, Querol A, et al. (1999).<br />

Molecular characterisation of <strong>Colletotrichum</strong> strains derived from strawberry.<br />

Mycological Research 103: 385–394.<br />

Cai L, Hyde KD, Taylor PWJ, Weir BS, Waller J, et al. (2009). A polyphasic approach<br />

for studying <strong>Colletotrichum</strong>. Fungal Diversity 39: 183–204.<br />

Cannon PF, Buddie AG, Bridge PD (2008). <strong>The</strong> typification of <strong>Colletotrichum</strong><br />

<strong>gloeosporioides</strong>. Mycotaxon 104: 189–204.<br />

Cannon PF, Damm U, Johnston PR, Weir BS (2012). <strong>Colletotrichum</strong> – current status<br />

and future directions. Studies in Mycology 73: 181–213.<br />

Carbone I, Kohn LM (1999). A method for designing primer sets for speciation<br />

studies in filamentous ascomycetes. Mycologia 91: 553–556.<br />

Carpenter JB, Stevenson JA (1954). A secondary leaf spot of the Hevea rubber tree<br />

caused by Glomerella cingulata. Plant Disease Reporter 38: 494–499.<br />

Castro RA, Mendes-Costa MC, Souza EA (2006). Dimorfismo de ascósporos em<br />

Glomerella cingulata f. sp. phaseoli. Fitopatologia Brasileira 31: 598–600.<br />

Chevaugeon J (1956). Les maladies cryptogamiques du manioc en Afrique<br />

Occidentale. Encyclopédie Mycologique 28: 1–205.<br />

Chung WS, Ishii H, Nishimura K, Fukaya M, Yano K, Kajitani Y (2006). Fungicide<br />

sensitivity and phylogenetic relationship of anthracnose fungi isolated from<br />

various fruit crops in Japan. Plant Disease 90: 506–512.<br />

Chung Y, Ané C (2011). Comparing Two Bayesian Methods for Gene Tree/Species<br />

Tree Reconstruction: Simulations with Incomplete Lineage Sorting and<br />

Horizontal Gene Transfer. Systematic Biology 60: 261–275<br />

Cisar CR, Spiegel FW, TeBeest DO, Trout C (1994). Evidence for mating between<br />

isolates of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> with different host specificities.<br />

Current Genetics 25: 330–335.<br />

Correll JC, Rhoads DD, Guerber JC (1993). Examination of mitochondrial DNA<br />

restriction fragment length polymophisms, DNA fingerprints, and randomly<br />

amplified polymorphic DNA of <strong>Colletotrichum</strong> orbiculare. Phytopathology 83:<br />

1199–1204.<br />

Crouch JA, Clarke BB, Hillman BI (2006). Unravelling evolutionary relationships<br />

amongst divergent lineages of <strong>Colletotrichum</strong> causing anthracnose disease in<br />

turfgrass and corn. Phytopathology 96: 46–60.<br />

177


Weir et al.<br />

Crous PW, Gams W, Stalpers JA, Robert V, Stegehuis G (2004). MycoBank: an<br />

online initiative to launch mycology into the 21st century. Studies in Mycology<br />

50: 19–22.<br />

Curzi M (1927). De novis Eumycetibus. Atti dell’Istituto Botanico della Università e<br />

Laboratorio Crittogamico di Pavia 3 (3): 203–208.<br />

Damm U, Cannon PF, Woudenberg JHC, Crous PW (2012a). <strong>The</strong> <strong>Colletotrichum</strong><br />

acutatum <strong>species</strong> <strong>complex</strong>. Studies in Mycology 73: 37–113.<br />

Damm U, Cannon PF, Woudenberg JHC, Johnston PR, Weir BS, et al. (2012b).<br />

<strong>The</strong> <strong>Colletotrichum</strong> boninense <strong>species</strong> <strong>complex</strong>. Studies in Mycology 73: 1–36.<br />

Daniel JT, Templeton GE, Smith RJ, Fox WT (1973). Biological control of northern<br />

jointvetch in rice with an endemic fungal disease. Weed Science 21: 303–<br />

307.<br />

Delacroix G (1903). De la tavelure goyaves produite par le Gloeosporium psidii nov.<br />

sp. G. Del. Bulletin de la Société Mycologique de France 19: 143–145.<br />

Deng Y, Zhang M, Luo H (2012). Identification and antimicrobial activity of two<br />

alkaloids from traditional Chinese medicinal plant Tinospora capillipes.<br />

Industrial Crops and Products 37: 298–302.<br />

Dickens JSW, Cook RTA (1989). Glomerella cingulata on Camellia. Plant Pathology<br />

38: 75–85.<br />

Ditmore M, Moore JW, TeBeest DO (2008). Interactions of two selected field isolates<br />

of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp. aeschynomene on Aeschynomene<br />

virginica. Biological Control 47: 298–308.<br />

Drummond AJ, Ashton B, Buxton S, Cheung M, Cooper A, et al. (2011). Geneious<br />

v5.4, Available from http://www.geneious.com/<br />

Drummond AJ, Suchard MA, Xie D, Rambaut A (2012). Bayesian phylogenetics with<br />

BEAUti and the BEAST 1.7. Molecular Biology and Evolution 29: 1969–1973.<br />

Du M, Schardl CL, Nuckles EM, Vaillancourt L (2005). Using mating-type gene<br />

sequences for improved phylogenetic resolution of <strong>Colletotrichum</strong> <strong>species</strong><br />

<strong>complex</strong>es. Mycologia 97: 641–658.<br />

Edgerton CW (1909). <strong>The</strong> perfect stage of the cotton anthracnose. Mycologia 1:<br />

115–120.<br />

Farr DF, Aime MC, Rossman AY, Palm ME (2006). Species of <strong>Colletotrichum</strong> on<br />

Agavaceae. Mycological Research 110: 1395–1408.<br />

Fitzpatrick DA, Logue ME, Stajich JE, Butler G (2006). A fungal phylogeny based<br />

on 42 complete genomes derived from supertree and combined gene analysis.<br />

BMC Evolutionary Biology 6: 99.<br />

Fokunang CN, Dixon AGO, Akem CN, Ikotun T (2000). Cutural, morphological<br />

and pathogenic variability in <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp. manihotis<br />

isolates from cassava (Manihot esculenta) in Nigeria. Pakistan Journal of<br />

Biological Sciences 3: 542–546.<br />

Gardes M, Bruns TD (1993). ITS primers with enhanced specificity for basidiomycetes<br />

– application to the identification of mycorrhizae and rusts. Molecular Ecology<br />

2: 113–118.<br />

Glass NL, Donaldson GC (1995). Development of primer sets designed for use with<br />

the PCR to amplify conserved genes from filamentous ascomycetes. Applied<br />

and Environmental Microbiology. 61: 1323–1330.<br />

Gielink AJ, Vermeulen H (1983). <strong>The</strong> ability of American and African <strong>Colletotrichum</strong><br />

isolates to cause coffee berry disease symptoms and the association of some<br />

isolates with Glomerella cingulata. Netherlands Journal of Plant Pathology 89:<br />

188–190.<br />

Gorter GJMA (1956). Anthracnose fungi of olives. Nature 178: 1129–1130.<br />

Green KR, Simons SA (1994). ‘Dead skin’ on yams (Dioscorea alata) caused by<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong>. Plant Pathology 43: 1062–1065.<br />

Greene GL (1967). Effect of 2-dehydroxy-D-glocose on the respiration of tropical<br />

anthracnose fungi. Physiologia Plantarum 20: 580–586.<br />

Guerber JC, Correll JC (2001). Characterisation of Glomerella acutata, the<br />

teleomorph of <strong>Colletotrichum</strong> acutatum. Mycologia 93: 216–229.<br />

Guerber JC, Liu B, Correll, JC, Johnston PR (2003). Characterization of diversity in<br />

<strong>Colletotrichum</strong> acutatum sensu lato by sequence analysis of two gene introns,<br />

mtDNA and intron RFLPs, and mating compatibility. Mycologia 95: 872–895.<br />

Gunnell PS, Gubler WD (1992). Taxonomy and morphology of <strong>Colletotrichum</strong><br />

<strong>species</strong> pathogenic to strawberry. Mycologia 84: 157–165.<br />

Harris P (1993). Effects, constraints and the future of weed biocontrol. Agriculture,<br />

Ecosystems and Environment 46: 289–303.<br />

Heled J, Drummond AJ (2010). Bayesian Inference of Species Trees from Multilocus<br />

Data. Molecular Biology and Evolution 27: 570–580.<br />

Higgins BB (1926). Anthracnose of pepper (Capsicum anuum L.). Phytopathology<br />

16: 333–345.<br />

Hildebrand PD, Jensen KIN (1991). Potential for the biological control of St<br />

Johns’s-wort (Hypericum perforatum) with an endemic strain of <strong>Colletotrichum</strong><br />

<strong>gloeosporioides</strong>. Canadian Journal of Plant Pathology 13: 60–70.<br />

Hindorf H (1973). <strong>Colletotrichum</strong>-Population auf Coffea arabica L. in Kenia II.<br />

Qualitative und quantitative Unterschiede in der <strong>Colletotrichum</strong>-Population.<br />

Phytopathologische Zeitschrift 77: 216–234<br />

Hosdson A, Mills PR, Brown AE (1993). Ribosomal and mitochondrial DNA<br />

polymorphisms in <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> isolated from tropical fruits.<br />

Mycological Research 97: 329–335.<br />

178<br />

Hsiang T, Goodwin PH (2001). Ribosomal DNA sequence comparisons of<br />

<strong>Colletotrichum</strong> graminicola from turfgrasses and other hosts. European Journal<br />

of Plant Pathology 107: 593–599.<br />

Hyde KD, Cai L, Cannon PF, Crouch JA, Crous PW, et al. (2009). <strong>Colletotrichum</strong> –<br />

names in current use. Fungal Diversity 39: 147–182.<br />

Hyde KD, Chomnunti P, Crous PW, Groenewald JZ, Damm U, et al. (2010). A case<br />

for re-inventory of Australia’s plant pathogens. Persoonia 25: 50–60.<br />

Irwin JAG, Cameron DF (1978). Two diseases in Stylosanthes caused by<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> in Australia, and pathogenic specialisation<br />

within one of the causal organisms. Australian Journal of Agricultural Research<br />

29: 3015–317.<br />

Johnson DA, Carris LM, Rogers JD (1997). Morphological and molecular<br />

characterization of <strong>Colletotrichum</strong> nymphaeae and C. nupharicola sp.<br />

nov. on water-lilies (Nymphaea and Nuphar). Mycological Research 101:<br />

641–649.<br />

Johnston PR (2000). <strong>The</strong> importance of phylogeny in understanding host<br />

relationships within <strong>Colletotrichum</strong>. In: <strong>Colletotrichum</strong>: host specificity,<br />

pathogenicity, and host-pathogen interactions (Prusky D, Dickman MB,<br />

Freeman S eds). APS Press, St Paul, Minnesota: 21–28.<br />

Johnston PR, Jones D (1997). Relationships among <strong>Colletotrichum</strong> isolates from<br />

fruit rots assessed using rDNA sequences. Mycologia 89: 420–430.<br />

Johnston PR, Pennycook SR, Manning MA (2005). Taxonomy of fruit-rotting fungal<br />

pathogens: what’s really out there? New Zealand Plant Protection 58: 42–46.<br />

Kaan F (1973). A study on inheritance of eggplant (Solanum melongena L.) resistance<br />

to fruit anthracnose (<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp. melongenae Penzig<br />

Fournet). Annales De l’Amelioration des Plantes 23:127–151.<br />

Kenny MK, Galea VJ, Price TV (2012). Germination and growth of <strong>Colletotrichum</strong><br />

acutatum and <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> isolates from coffee in Papua<br />

New Guinea and their pathogenicity. Australasian Plant Pathology (online,<br />

doi:10.1007/s13313-012-0117-7).<br />

Killgore EM, Sugiyama LS, Barreto RW, Gardner DE (1999). Evaluation of<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> for Biological Control of Miconia calvescens in<br />

Hawaii. Plant Disease 83: 964.<br />

Kornerup A, Wanscher JH (1963). Methuen handbook of colour. Methuen and Co.<br />

London.<br />

Krauss U, Matthews P, Bidwell R, Hocart M, Anthony F (2001). Strain discrimination<br />

by fungal antagonists of <strong>Colletotrichum</strong> musae: implications for biocontrol of<br />

crown rot of banana. Mycological Research 105: 67–76.<br />

Lardner R, Johnston PR, Plummer KM, Pearson MN (1999). Morphological and<br />

molecular analysis of <strong>Colletotrichum</strong> acutatum sensu lato. Mycological<br />

Research 103: 275–285.<br />

Larkin MA, Blackshields G, Brown NP, Chenna R, McGettigan PA, et al. (2007).<br />

Clustal W and Clustal X v. 2.0. Bioinformatics 23: 2947–2948.<br />

Laundon GF (1972) [1971]. Records of fungal plant diseases in New Zealand – 2.<br />

New Zealand Journal of Botany 9: 610–624.<br />

Li R, Ma H (1993). Studies in the occurrence and control of ramie anthracnose. Acta<br />

Phytophylacica Sinica 20: 83–89. [in Chinese]<br />

Liu LJ (1972). Identification and occurrence of perfect stage and cultural and<br />

morpholoical variants of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> from guava in Puerto<br />

Rico. Journal of Agriculture of the University of Puerto Rico 56: 171–180.<br />

Liu B, Louws FJ, Sutton TB, Correll JC (2011). A rapid qualitative molecular method<br />

for the identification of <strong>Colletotrichum</strong> acutatum and C. <strong>gloeosporioides</strong>.<br />

European Journal of Plant Pathology 132: 593–607.<br />

Lubbe CM, Denman S, Cannon PF, Groenewald JZ, Lamprecht SC, Crous PW<br />

(2004). Characterisation of <strong>Colletotrichum</strong> <strong>species</strong> associated with Proteaceae.<br />

Mycologia 96: 1268–1279.<br />

MacCaughey V (1917). <strong>The</strong> guavas of the Hawaiian Islands. Bulletin of the Torrey<br />

Botanical Club 44: 513–524.<br />

Mackenzie SJ, Seijo TE, Legard DE, Timmer LW, Peres NA (2007). Selection for<br />

pathogenicity to strawberry in populations of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong><br />

from native plants. Phytopathology 97: 1130–1140.<br />

Mackenzie SJ, Mertely JC, Seijo TE, Peres NA (2008). <strong>Colletotrichum</strong> fragariae is a<br />

pathogen on hosts other than strawberry. Plant Disease 92: 1432–1438.<br />

McNeill J, Barrie FR, Burdet HM, Demoulin V, Hawksworth DL, et al. (2006).<br />

International Code of Botanical Nomenclature (Vienna Code). Adopted by the<br />

Seventeenth International Botanical Congress, Vienna, Austria, July 2005.<br />

Regnum Vegetabile 146. 568 p.<br />

Makambila C (1994). <strong>The</strong> fungal diseases of cassava in the Republic of Congo,<br />

Central Africa. African Crop Science Journal 2: 511–517.<br />

Makowski RMD, Mortensen K (1989). <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp.<br />

malvae as a bioherbicide for round-leafed mallow (Malva pusilla): conditions<br />

for successful control in the field. In: Proceedings of the 7 th International<br />

Symposium for the Biological Control of Weeds. (Delfosse ES ed.). Istituto<br />

Sperimentale per la Patalogia Vegetale, Rome: 513–522.<br />

Marcelino J, Giordano R, Gouli S, Gouli V, Parker BL, et al. (2008). <strong>Colletotrichum</strong><br />

acutatum var. fioriniae (teleomorph: Glomerella acutata var. fioriniae var. nov.)<br />

infection of a scale insect. Mycologia 100: 353–374.


Martínez-Culebras PV, Barrio E, García MD, Querol A (2000). Identification of<br />

<strong>Colletotrichum</strong> <strong>species</strong> responsible for anthracnose of strawberry based on<br />

the internal transcribed spacers of the ribosomal region. FEMS Microbiology<br />

Letters 189: 97–101.<br />

Martinez-Culebras PV, Querol A, Suarez-Fernandez MB, Garcia-Lopez MD,<br />

Barrio E (2003). Phylogenetic relationships among <strong>Colletotrichum</strong> pathogens<br />

of strawberry and design of PCR primers for their identification. Journal of<br />

Phytopathology 151: 135–143.<br />

Maymon M, Zveibil A, Pivonia S, Minz D, Freeman S (2006). Identification<br />

and characterisation of benomyl-resistant and sensitive- populations of<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> from statice (Limonium spp.). Phytopathology<br />

96: 542–548.<br />

Menten JOM, Kimati H, Costa CP (1980). Reação de populações de pepino<br />

(Cucumis sativus L.) a <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp. cucurbitae (Berk.<br />

et Mont.) N. Comb. Summa Phytopathologica 5: 127–133.<br />

Mills PR, Sreenivasaprasad S, Brown AE (1992). Detection and differentiation<br />

of <strong>Colletotrichum</strong> gl Maymon oeosporioides isolates using PCR. FEMS<br />

Microbiological Letters 98: 137–144.<br />

Misra AK (2004). Guava diseases – their symptoms, causes and management.<br />

In: Diseases of fruits and vegetables. Diagnosis and management Volume 2.<br />

(Naqvi SAMH. ed.). Kluwer Academic Publishers, Dordrecht: 81–119.<br />

Mordue JEM (1971). Glomerella cingulata. CMI Desriptions of Plant Pathogenic<br />

Fungi No. 315. Commonwealth Mycological Institute, Kew.<br />

Moriwaki J, Sato T, Tsukiboshi T (2003). Morphological and molecular characterisation<br />

of <strong>Colletotrichum</strong> boninense sp. nov. from Japan. Mycoscience 44: 47–53.<br />

Moriwaki J, Sato T (2009). A new combination for the causal agent of tea<br />

anthracnose: Discula theae-sinensis (I. Miyake) Moriwaki & Toy. Sato, comb.<br />

nov. Journal of General Plant Pathology 75: 359–361.<br />

Moriwaki J, Tsukiboshi T (2009). <strong>Colletotrichum</strong> echinochloae, a new <strong>species</strong> on<br />

Japanese barnyard millet (Echinochloa utilis). Mycoscience 50: 273–280.<br />

Munaut F, Hamaide N, Maraite H (2002). Genomic and pathogenic diversity in<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> from wild native Mexican Stylosanthes spp.,<br />

and taxonomic implications. Mycological Research 106: 579–593.<br />

Munaut F, Hamaide N, Vander Stappen J, Maraite H (1998). Genetic relationships<br />

among isolates of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> from Stylosanthes spp. in<br />

Africa and Australia using RAPD and ribosomal DNA markers. Plant Pathology<br />

47: 641–648.<br />

Muraleedharan N, Baby UI (2007). Tea disease: ecology and control. In: Encyclopedia<br />

of Pest Management Vol. 2 (Pimentel D, ed.) [electronic resource]. CRC Press,<br />

Boca Raton: 668–671.<br />

Nilsson RH, Ryberg M, Kristiansson E, Abarenkov K, Larsson KH, Kõljalg U (2006).<br />

Taxonomic reliability of DNA sequences in public sequence databases: a fungal<br />

perspective. PLoS ONE 1(1): e59.<br />

Nirenberg HI, Feiler U, Hagedorn G (2002). Description of <strong>Colletotrichum</strong> lupini<br />

comb. nov. in modern terms. Mycologia 94: 307–320.<br />

Noak F (1901). Die Krankheiten des Kaffeebaumes in Brasilien. Zeitschrift für<br />

Pflanzenkrankheiten 11: 196–203.<br />

Ocfemia, GO, Agati JA (1925). <strong>The</strong> cause of anthracnose of avocado, mango and<br />

upo in the Philippine Islands. Philippine Agriculturalist 14: 199–216.<br />

O’Donnell K, Cigelnik E (1997). Two divergent intragenomic rDNA ITS2 types within<br />

a monophyletic lineage of the fungus Fusarium are nonorthologous. MoIecular<br />

Phylogenetics and Evolution 7:103–116.<br />

O’Donnell K, Nirenberg HI, Aoki T, Cigelnik E (2000). A Multigene phylogeny of the<br />

Gibberella fujikuroi <strong>species</strong> <strong>complex</strong>: Detection of additional phylogenetically<br />

distinct <strong>species</strong>. Mycoscience 41: 61–78.<br />

Owolade OF, Dixon AGO, Alabi BS, Akande SR, Olakojo SA (2008). A combining<br />

ability analysis of cassava Manihot esculenta Crantz genotypes to anthracnose<br />

disease. Electronic Journal of Environmental, Agricultural and Food Chemistry<br />

7: 2959–2968.<br />

Parvin S, Lee OR, Sathiyaraj G, Khorolragchaa A, Kim YJ, et al. (2012).<br />

Interrelationship between calmodulin (CaM) and H2O2 in abscisic acid-induced<br />

antioxidant defense in the seedlings of Panax ginseng. Molecular Biology<br />

Reports 39: 7327–7338.<br />

Penzig AGO (1882). Fungi agrumicoli. Contribuzione allo studio dei funghi parassiti<br />

degli agrumi. Michelia 2: 385–508.<br />

Phoulivong S, Cai L, Parinn N, Chen H, Abd-Elsalam K, et al. (2011). A new <strong>species</strong><br />

of <strong>Colletotrichum</strong> from Cordyline fruticosa and Eugenia javanica causing<br />

anthracnose disease. Mycotaxon 114: 247–257.<br />

Polashock JJ, Caruso FL, Oudemans PV, McManus PS, Crouch JA (2009). <strong>The</strong><br />

North American cranberry fruit rot fungal community: a systematic overview<br />

using morphological and phylogenetic affinities. Plant Pathology 58: 1116–<br />

1127.<br />

Pollacci G (1899). Contribuzione alla micologia ligustica. Atti dell’ Istituto Botanico<br />

dell’ Università di Pavia, Serie 2 5: 29–46.<br />

Posada D. (2008). jModelTest: Phylogenetic Model Averaging. Molecular Biology<br />

and Evolution 25: 1253–1256<br />

Prasad N, Singh RD (1960). Anthracnose disease of Dioscorea alata L. (Yam, Eng.,<br />

www.studiesinmycology.org<br />

<strong>The</strong> ColletotriChum <strong>gloeosporioides</strong> <strong>species</strong> <strong>complex</strong><br />

Ratalu, Hind.). Current Science 29: 66–67.<br />

Prihastuti H, Cai L, Chen H, McKenzie EHC, Hyde KD (2009). Characterisation<br />

of <strong>Colletotrichum</strong> <strong>species</strong> associated with coffee berries in northern Thailand.<br />

Fungal Diversity 39: 89–109.<br />

R Development Core Team (2011). R: A language and environment for statistical<br />

computing. R Foundation for Statistical Computing, Vienna, Austria.<br />

Rambaut A, Drummond AJ (2007). Tracer v. 1.4, Available from http://beast.bio.<br />

ed.ac.uk/Tracer<br />

Rodrigues CJ, Varzea VMP, Hindorf H, Medeiros EF (1991). Strains of <strong>Colletotrichum</strong><br />

coffeanum Noack causing coffee berry disease in Angola and Malawi with<br />

characteristics different to the Kenya strain. Journal of Phytopathology 131:<br />

205–209.<br />

Rodriguez RJ, Owen JL (1992). Isolation of Glomerella musae [teleomorph<br />

of <strong>Colletotrichum</strong> musae (Berk. & Curt.) Arx] and segregation analysis of<br />

ascospore progeny. Experimental Mycology 16: 291–301.<br />

Rodríguez-Guerra R, Ramírez-Rueda M, Cabral-Enciso M, García-Serrano M, Lira-<br />

Maldonado Z, et al. (2005). Heterothallic mating observed between Mexican<br />

isolates of Glomerella lindemuthiana. Mycologia 97: 793–803.<br />

Rojas EI, Rehner SA, Samuels GJ, Van Bael SA, Herre EA, et al. (2010).<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> s.l. associated with <strong>The</strong>obroma cacao and<br />

other plants in Panamá: multilocus phylogenies distinguish host-associated<br />

pathogens from asymptomatic endophytes. Mycologia 102: 1318–1338.<br />

Ronquist F, Teslenko M, van der Mark P, Ayres D, Darling A, et al. (2012). MrBayes v.<br />

3.2: Efficient Bayesian phylogenetic inference and model choice across a large<br />

model space. Systematic Biology 61: 539–542.<br />

Saccardo PA, Trotter A (1913). Supplementum Universale Pars. IX. Ascomycetae-<br />

Deuteromycetae. Sylloge Fungorum 22: 1–1612.<br />

Sanders GM, Korsten L (2003). Comparison of cross inoculation potential of South<br />

African avocado and mango isolates of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong>.<br />

Microbiological Research 158: 143–150.<br />

Sawada T (1922). New Japanese fungi notes and translations – XI. Mycologia 14:<br />

81–89.<br />

Schoch CL, Seifert KA, Huhndorf S, Robert V, Spouge JL, et al. (2012). Nuclear<br />

ribosomal internal transcribed spacer (ITS) region as a universal DNA barcode<br />

marker for Fungi. Proceedings of the National Academy of Sciences 109:<br />

6241–6246.<br />

Schwarczinger I, Vajna L, Bruckart WL (1998). First report of <strong>Colletotrichum</strong><br />

<strong>gloeosporioides</strong> on Russian thistle. Plant Disease 82: 1405.<br />

Seifert KA (2009). Progress towards DNA barcoding of fungi. Molecular Ecology<br />

Resources 9 (Suppl. 1): 83–89.<br />

Shear CL (1907). New <strong>species</strong> of fungi. Bulletin of the Torrey Botanical Club 34:<br />

305–317.<br />

Shear CL, Wood AK (1907). Ascogenous forms of Gloeosporium and <strong>Colletotrichum</strong>.<br />

Botanical Gazette 43: 259–266.<br />

Shear CL, Wood AK (1913). Studies of fungus parasites belonging to the genus<br />

Glomerella. USDA Bureau of Plant Industry Bulletin 252: 1–110.<br />

Sheldon JL (1906). <strong>The</strong> ripe rot or mummy disease of guavas. Bulletin of the West<br />

Virginia University Experimental Station 104: 299–315.<br />

Shenoy BD, Jeewon R, Lam WH, Bhat DJ, Than PP, et al. (2007). Morpho-molecular<br />

characterisation and epitypification of <strong>Colletotrichum</strong> capsici (Glomerellaceae,<br />

Sordariomycetes), the causative agent of anthracnose on chilli. Fungal<br />

Diversity 27: 197–211.<br />

Sherf AF, MacNab AA (1986). Vegetable diseases and their control. John Wiley and<br />

Sons. USA.<br />

Sherriff C, Whelan MJ, Arnold GM, Lafay JF, Brygoo Y, Bailey JA (1994). Ribosomal<br />

DNA sequence analysis reveals new <strong>species</strong> groupings in the genus<br />

<strong>Colletotrichum</strong>. Experimental Mycology 18: 121–138.<br />

Shivas RG, Bathgate J, Podger FD (1998). <strong>Colletotrichum</strong> xanthorrhoeae sp.<br />

nov. on Xanthorrhoea in Western Australia. Mycological Research 102:<br />

280–282.<br />

Silva DN, Talhinas P, Várzea V, Cai L, Paulo OS, Batista D (2012a). Application<br />

of the Apn2/MAT locus to improve the systematics of the <strong>Colletotrichum</strong><br />

<strong>gloeosporioides</strong> <strong>complex</strong>: an example from coffee (Coffea spp.) hosts.<br />

Mycologia 104: 396–409.<br />

Silva DN, Talhinhas P, Cai L, Manuel L, Gichuru EK, et al. (2012b). Host-jump<br />

drives rapid and recent ecological speciation of the emergent fungal pathogen<br />

<strong>Colletotrichum</strong> kahawae. Molecular Ecology 21: 2655–2670.<br />

Silva VN da, Guzzo SD, Mantovanello-Lucon CM, Harakava R (2011). Promoção de<br />

crescimento e inducção de resistência à antracnose por Trichoderma spp. em<br />

pepineiro. Pesquisa Agropecuária Brasileira 46: 1609–1618.<br />

Silva-Mann R, Carvalho Vieira MGG, Machado JC, Bernandino Filho JR, Salgado<br />

KCC, Stevens MR (2005). AFLP markers differentiate isolates of <strong>Colletotrichum</strong><br />

gossypii from C. gossypii var. cephalosporioides. Fitopatologia Brasileira 30:<br />

169–172.<br />

Simmonds JH (1965). A study of the <strong>species</strong> of <strong>Colletotrichum</strong> causing ripe fruit<br />

rots in Queensland. Queensland Journal of Agricultural and Animal Sciences<br />

22: 437–459.<br />

179


Weir et al.<br />

Simmonds JH (1968). Type specimens of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> var. minor<br />

and <strong>Colletotrichum</strong> acutatum. Queensland Journal of Agricultural and Animal<br />

Sciences 25: 178A.<br />

Singh UP (1974). A new forma specialis of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> Penz.<br />

from India. Beihefte zur Nova Hedwigia 47: 451–452.<br />

Singh UP (1975). <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> Penzig - A mycoparasite on<br />

Ravenelia sessilis Berkeley. Current Science 44: 623–624.<br />

Singh RD, Prasad N, Mathur RL (1966). On the taxonomy of the fungus causing<br />

anthracnose of Dioscorea alata L. Indian Phytopathology 19: 65–71.<br />

Small W (1926). On the occurrences of a <strong>species</strong> of <strong>Colletotrichum</strong>. Transactions of<br />

the British Mycological Society 11: 112–137.<br />

Snowdon AL 1991. A colour atlas of post-harvest diseases and disorders of fruit and<br />

vegetables. Volume 2. Vegetables. Wolfe Scientific.<br />

Sosa de Castro NT, Cabrera de Alvarez MG, Alvarez RE (2001). Primera<br />

información de <strong>Colletotrichum</strong> camelliae como patógeno de Camellia japonica,<br />

en Corrientes. Retrieved 6 Oct 2010 from www1.unne.edu.ar/cyt/2001/5-<br />

Agrarias/A-056.pdf.<br />

Southworth EA (1891). Antracnose of cotton. Journal of Mycology 6: 100–105.<br />

Spaulding P (1958). Diseases of foreign trees growing in the United States, an<br />

annotated list. USDA Agriculture Handbook 139.<br />

Spegazzini CL (1878). Ampelomiceti Italici, ossia enumerazione, diagnosi e storia<br />

dei principali parassiti della vite. Rivista Viticoltura ed Enologia 2: 405–411.<br />

Sreenivasaprasad S, Mills PR, Meehan B M, Brown AE (1996). Phylogeny and<br />

systematics of 18 <strong>Colletotrichum</strong> <strong>species</strong> based on ribosomal DNA spacer<br />

sequences. Genome 39: 499–512.<br />

Stephenson SA, Green JR, Manners JM, Maclean DJ (1997). Cloning and<br />

characterisation of glutamine synthetase from <strong>Colletotrichum</strong> <strong>gloeosporioides</strong><br />

and demonstration of elevated expression during pathogenesis on Stylosanthes<br />

guianensis. Current Genetics 31: 447–454.<br />

Stevens FL, Hall JG (1909). Eine neue Feigen-Anthraknose (Colletotrichose).<br />

Zeitschrift für Pflanzenkrankheiten 19: 65–68.<br />

Stoneman B (1898). A comparative study of the development of some anthracnoses.<br />

Botanical Gazette 26: 69–120.<br />

Su YY, Noireung P, Liu F, Hyde KD, Moslem MA, et al. (2011). Epitypification<br />

of <strong>Colletotrichum</strong> musae, the causative agent of banana anthracnose.<br />

Mycoscience 52: 376–382.<br />

Sutton BC, Waterston JM (1970). <strong>Colletotrichum</strong> musae. CMI Descriptions of Plant<br />

Pathogenic Fungi and Bacteria No. 222.<br />

Sutton BC (1980). <strong>The</strong> Coelomycetes. Commonwealth Mycological Institite. Kew.<br />

Sutton BC (1992). <strong>The</strong> genus Glomerella and its anamorph <strong>Colletotrichum</strong>. In:<br />

<strong>Colletotrichum</strong> Biology, Pathology and Control. (Bailey JA, Jeger MJ, eds).<br />

CAB International, Wallingford: 1–26.<br />

Talhinas P, Sreenivasaprasad S, Neves-Martins J, Oliviera H (2005). Molecular and<br />

phenotypic analyses reveal association of diverse <strong>Colletotrichum</strong> acutatum<br />

groups and a low level of C. <strong>gloeosporioides</strong> with olive anthracnose. Applied<br />

and Environmental Microbiology 71: 2987–2998.<br />

Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, et al. (2000). Phylogenetic<br />

<strong>species</strong> recognition and <strong>species</strong> concepts in fungi. Fungal Genetics and<br />

Biology 31: 21–32.<br />

Taylor JW, Berbee ML (2006). Dating divergences in the Fungal Tree of Life: review<br />

and new analyses. Mycologia 98: 838–849.<br />

TeBeest DO (1988). Additions to host range of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp.<br />

aeschynomene. Plant Disease 72: 16–18.<br />

Templeton MD, Rikkerink EHA, Solon SL, Crowhurst RN. (1992). Cloning and<br />

molecular characterization of the glyceraldehyde-3-phosphate dehydrogenaseencoding<br />

gene and cDNA from the plant pathogenic fungus Glomerella<br />

cingulata. Gene 122: 225–230.<br />

Than PP, Jeewon R, Hyde KD, Pongsupasamit S, Mongkolporn O, Taylor PWJ<br />

(2008). Characterisation and pathogenicity of <strong>Colletotrichum</strong> <strong>species</strong><br />

asscoaited with anthracnose on chilli (Capsicum spp.) in Thailand. Plant<br />

Pathology 57: 562–572.<br />

Thang MM (2008). Biodiversity survey of coelomycetes in Burma. Australasian<br />

Mycologist 27: 74–110.<br />

Trujillo EE, Latterell FM, Rossi AE (1986). <strong>Colletotrichum</strong> <strong>gloeosporioides</strong>, a<br />

possible biological control agent for Clidemia hirta in Hawaiian forests. Plant<br />

Disease 70: 974–976.<br />

Tsurushima T, Ueno T, Fukami H, Irie H, Inoue M (1995). Germination of selfinhibitors<br />

from <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> f. sp. jussiaea. Molecular Plant-<br />

Microbe Interactions 8: 652–657.<br />

180<br />

Vajna L, Bagyinka T (2002). Occurrence of privet anthracnose in Hungary caused by<br />

Glomerella cingulata. Plant Disease 51: 403.<br />

Varzea VMP, Rodrigues Junior CJ, Lewis BG (2002). Distinguishing characteristics<br />

and vegetative compatibility of <strong>Colletotrichum</strong> kahawae in comparison with<br />

other related <strong>species</strong> from coffee. Plant Pathology 51: 202–207.<br />

Venkatakrishniah NS (1952). Glomerella psidii (Del.) Sheld. and Pestalotia psidii<br />

Pat. associated with a cankerous disease of guava. Proceedings of the Indian<br />

Academy of Plant Sciences 36: 129–134.<br />

Vermeulen H, De Ferrante B, Gielink AJ (1984). Genetic and morphological diversity<br />

of mono-spore isolates of Glomerella cingulata associated with coffee berry<br />

disease. Netherlands Journal of Plant Pathology 90: 213–223.<br />

Vinijsanun T, Irwin JAG, Cameron DF (1987). Host range of three strains of<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> from tropical pasture legumes, and comparative<br />

histological studies of interactions between Type B disease-producing strains<br />

and Stylosanthes scabra (non-host) and S. guianensis (host). Australian<br />

Journal of Botany 35: 665–677.<br />

Waller JM, Bridge PD, Black R, Hakiza G (1993). Characterisation of the coffee berry<br />

disease pathogen, <strong>Colletotrichum</strong> kahawae sp. nov. Mycological Research 97:<br />

989–994.<br />

Wang XX, Wang B, Liu JL, Chen J, Cui XP, et al. (2010). First reports of anthracnose<br />

caused by <strong>Colletotrichum</strong> gloesporioides on ramie in China. Plant Disease 94:<br />

1508.<br />

Weir BS, Johnston PR (2009). Defining and delimiting genetic <strong>species</strong> in<br />

<strong>Colletotrichum</strong>. Asian Mycologyical Congress 2009. Taichung Taiwan<br />

November 15–19. pg O-128.<br />

Weir BS, Johnston PR (2010). Characterisation and neotypification of Gloeosporium<br />

kaki Hori as <strong>Colletotrichum</strong> horii nom. nov. Mycotaxon 111: 209–219.<br />

Weir BS, Johnston PR (2011). Towards a secondary fungal barcode for<br />

<strong>Colletotrichum</strong> [Sordariomycetes, Ascomycota]. Forth international barcode of<br />

life conference, Adelaide Australia, 28 November – 3 December 2011.<br />

White TJ, Bruns T, Lee S, Taylor JW. (1990). Amplification and direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: A Guide to<br />

Methods and Applications. (Innis MA, Gelfand DH, Sninsky JJ, White TJ eds).<br />

Academic Press: New York: 315–322.<br />

Wickham H (2009). ggplot2: elegant graphics for data analysis. Springer New York.<br />

Wikee S, Cai L, Pairin N, McKenzie EHC, Su YY, et al. (2011). <strong>Colletotrichum</strong><br />

<strong>species</strong> from Jasmine (Jasminum sambac). Fungal Diversity 46: 171–182.<br />

Willis JC (1899). DCLII - Tea and coffee diseases. Bulletin of Miscellaneous<br />

Information Royal Botanical Gardens Kew 1899: 89–94.<br />

Winch JE, Newhook FJ, Jackson GVH, Cole JS (1984). Studies of <strong>Colletotrichum</strong><br />

<strong>gloeosporioides</strong> disease on yam, Dioscorea alata, in Solomon Islands. Plant<br />

Pathology 33: 467–477.<br />

Wollenweber HW, Hochapfel H (1949). Beiträge zur Kenntnis parasitärer und<br />

saprophytischer Pilze. VI. Vermicularia, <strong>Colletotrichum</strong>, Gloeosporium,<br />

Glomerella und ihre Beziehung zur Fruchtfäule. Zeitschrift für Parasitenkunde<br />

14: 181–268.<br />

Xie L, Zhang JZ, Cai L, Hyde KD (2010a). Biology of <strong>Colletotrichum</strong> horii, the causal<br />

agent of persimmon anthracnose. Mycology 1: 242–253.<br />

Xie L, Zhang JZ, Wan Y, Hu DW (2010b). Identification of <strong>Colletotrichum</strong> spp.<br />

isolated from strawberry in Zhejiang Province and Shanghai City, China.<br />

Journal of Zhejiang University Science B (Biomedicine & Biotechnology) 11:<br />

61–70.<br />

Yamamoto W (1960). A revision of the genus and <strong>species</strong> names of anthracnose<br />

fungi in Japan. Plant Protection 14: 49–52. [in Japanese]<br />

Yan J, Wu PS, Du HZ, Zhang QE (2011). First report of black spot caused by<br />

<strong>Colletotrichum</strong> <strong>gloeosporioides</strong> on paper mulberry in China. Plant Disease 95:<br />

880.<br />

Yang SY, Su SC, Liu T, Fan G, Wang J (2011). First report of anthracnose caused<br />

by <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> on pistachio (Pistacia vera) in China. Plant<br />

Disease 95: 1314.<br />

Yang YL, Liu ZY, Cai L, Hyde KD, Yu ZN, McKenzie EHC (2009). <strong>Colletotrichum</strong><br />

anthracnose of Amaryllidaceae. Fungal Diversity 39: 123–146.<br />

Zhang TY (1985). A forma specialis of <strong>Colletotrichum</strong> <strong>gloeosporioides</strong> on Cuscuta<br />

sp. Acta Mycologica Sinica 4: 223–239. [In Chinese]<br />

Zhang X, Yang G, Yang J, Shu Q (2012). Physiological mechanism of resistance to<br />

anthracnose of different Camellia varieties. African Journal of Biotechnology<br />

11: 2026–2031.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!