Europe PMC

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

Abstract 


A comprehensive phylogenetic reassessment of the ascomycete genus Cosmospora (Hypocreales, Nectriaceae) is undertaken using fresh isolates and historical strains, sequences of two protein encoding genes, the second largest subunit of RNA polymerase II (rpb2), and a new phylogenetic marker, the larger subunit of ATP citrate lyase (acl1). The result is an extensive revision of taxonomic concepts, typification, and nomenclatural details of many anamorph- and teleomorph-typified genera of the Nectriaceae, most notably Cosmospora and Fusarium. The combined phylogenetic analysis shows that the present concept of Fusarium is not monophyletic and that the genus divides into two large groups, one basal in the family, the other terminal, separated by a large group of species classified in genera such as Calonectria, Neonectria, and Volutella. All accepted genera received high statistical support in the phylogenetic analyses. Preliminary polythetic morphological descriptions are presented for each genus, providing details of perithecia, micro- and/or macro-conidial synanamorphs, cultural characters, and ecological traits. Eight species are included in our restricted concept of Cosmospora, two of which have previously documented teleomorphs and all of which have Acremonium-like microconidial anamorphs. A key is provided to the three anamorphic species recognised in Atractium, which is removed from synonymy with Fusarium and epitypified for two macroconidial synnematous species and one sporodochial species associated with waterlogged wood. Dialonectria is recognised as distinct from Cosmospora and two species with teleomorph, macroconidia and microconidia are accepted, including the new species D. ullevolea. Seven species, one with a known teleomorph, are classified in Fusicolla, formerly considered a synonym of Fusarium including members of the F. aquaeductuum and F. merismoides species complex, with several former varieties raised to species rank. Originally a section of Nectria, Macroconia is raised to generic rank for five species, all producing a teleomorph and macroconidial anamorph. A new species of the Verticillium-like anamorphic genus Mariannaea is described as M. samuelsii. Microcera is recognised as distinct from Fusarium and a key is included for four macroconidial species, that are usually parasites of scale insects, two of them with teleomorphs. The four accepted species of Stylonectria each produce a teleomorph and micro- and macroconidial synanamorphs. The Volutella species sampled fall into three clades. Pseudonectria is accepted for a perithecial and sporodochial species that occurs on Buxus. Volutella s. str. also includes perithecial and/or sporodochial species and is revised to include a synnematous species formerly included in Stilbella. The third Volutella-like clade remains unnamed. All fungi in this paper are named using a single name system that gives priority to the oldest generic names and species epithets, irrespective of whether they are originally based on anamorph or teleomorph structures. The rationale behind this is discussed.

Free full text 


Logo of simycolStudies in MycologyAbout the JournalInstructions to AuthorsEditorial BoardWebshopSubscribe
Stud Mycol. 2011; 68: 79–113.
PMCID: PMC3065986
PMID: 21523190

An overview of the taxonomy, phylogeny, and typification of nectriaceous fungi in Cosmospora, Acremonium, Fusarium, Stilbella, and Volutella

Abstract

A comprehensive phylogenetic reassessment of the ascomycete genus Cosmospora (Hypocreales, Nectriaceae) is undertaken using fresh isolates and historical strains, sequences of two protein encoding genes, the second largest subunit of RNA polymerase II (rpb2), and a new phylogenetic marker, the larger subunit of ATP citrate lyase (acl1). The result is an extensive revision of taxonomic concepts, typification, and nomenclatural details of many anamorph- and teleomorph-typified genera of the Nectriaceae, most notably Cosmospora and Fusarium. The combined phylogenetic analysis shows that the present concept of Fusarium is not monophyletic and that the genus divides into two large groups, one basal in the family, the other terminal, separated by a large group of species classified in genera such as Calonectria, Neonectria, and Volutella. All accepted genera received high statistical support in the phylogenetic analyses. Preliminary polythetic morphological descriptions are presented for each genus, providing details of perithecia, micro- and/or macro-conidial synanamorphs, cultural characters, and ecological traits. Eight species are included in our restricted concept of Cosmospora, two of which have previously documented teleomorphs and all of which have Acremonium-like microconidial anamorphs. A key is provided to the three anamorphic species recognised in Atractium, which is removed from synonymy with Fusarium and epitypified for two macroconidial synnematous species and one sporodochial species associated with waterlogged wood. Dialonectria is recognised as distinct from Cosmospora and two species with teleomorph, macroconidia and microconidia are accepted, including the new species D. ullevolea. Seven species, one with a known teleomorph, are classified in Fusicolla, formerly considered a synonym of Fusarium including members of the F. aquaeductuum and F. merismoides species complex, with several former varieties raised to species rank. Originally a section of Nectria, Macroconia is raised to generic rank for five species, all producing a teleomorph and macroconidial anamorph. A new species of the Verticillium-like anamorphic genus Mariannaea is described as M. samuelsii. Microcera is recognised as distinct from Fusarium and a key is included for four macroconidial species, that are usually parasites of scale insects, two of them with teleomorphs. The four accepted species of Stylonectria each produce a teleomorph and micro- and macroconidial synanamorphs. The Volutella species sampled fall into three clades. Pseudonectria is accepted for a perithecial and sporodochial species that occurs on Buxus. Volutella s. str. also includes perithecial and/or sporodochial species and is revised to include a synnematous species formerly included in Stilbella. The third Volutella-like clade remains unnamed. All fungi in this paper are named using a single name system that gives priority to the oldest generic names and species epithets, irrespective of whether they are originally based on anamorph or teleomorph structures. The rationale behind this is discussed.

Keywords: Article 59, Buxus, codon model, holomorph concept, unitary nomenclature, synnematous hyphomycetes

INTRODUCTION

This paper focuses on phylogenetic and taxonomic reassessment of the prevailing concept of the ascomycete genus Cosmospora (Nectriaceae, Hypocreales) (Samuels et al. 1991, Rossman et al. 1999). This genus has been assumed to be polyphyletic because of its anamorphic and biological diversity, a fact recently reinforced by phylogenetic studies on a limited sampling of species (Zhang & Zhuang 2006, Luo & Zhuang 2008, Samuels et al. 2009). The majority of described Cosmospora species have Acremonium-like or Fusarium-like anamorphs, but hyphomycetous anamorphs classified in Chaetopsina, Cylindrocladiella, Gliocladiopsis, Mariannaea, Penicillifer, Stilbella, Verticillium, and Volutella have also been associated with the genus (Samuels et al. 1991 as Nectria subgenus Dialonectria, Rossman et al. 1999). The prevailing concept of Cosmospora is unified by the teleomorph, which tends to be relatively nondescript, with usually solitary, astromatic, smooth, thin-walled perithecia, often orange or reddish, and changing to dark red in KOH, and 1-septate ascospores in a cylindrical ascus with a simple apex of refractive apical ring; for convenience we will refer to this concept as Cosmospora sensu Rossman.

Before DNA-based phylogenetic studies significantly influenced fungal taxonomy, anamorph taxonomy in the Hypocreales had shifted away from classical form-taxa towards a practice that correlated teleomorphic and anamorphic generic concepts (Samuels & Seifert 1987). Preceding the segregation of Nectria sensu Booth into many teleomorph genera in three families, Rossman (1993) suggested the delimitation of each teleomorph genus with one anamorph genus, the so-called 1:1 genus concept. Taxonomic equivalency between linked teleomorph and anamorph genera was proposed for several groups of the Bionectriaceae, Nectriaceae, and Hypocreaceae. Within the Cosmospora complex, for example, this rationale was used in the corresponding generic concepts for Nectricladiella (teleomorph), with Cylindrocladiella (anamorph) (Schoch et al. 2000), and Chaetopsinectria (teleomorph) with Chaetopsina (anamorph) (Luo & Zhuang 2010).

Booth's broad concept of Nectria dominated for 30 years; he recognised “groups” of species including the Episphaeria group (Booth 1959). This group, with additional species, was revised first as Nectria subgenus Dialonectria by Samuels et al. (1991), and then elevated to generic rank as Cosmospora (Rossman et al. 1999). The latter is typified by C. coccinea (= Nectria cosmariospora, not Nectria coccinea, which is a different fungus), which Saccardo (1883) listed as the only member of Nectria subgenus Cosmospora. Cosmospora coccinea produces orange, solitary, superficial perithecia and verrucose, brownish ascospores; its anamorph is Verticillium olivaceum (Gams 1971).

The relationship of the prevailing concept of Cosmospora with the generic concept of the economically important anamorph genus Fusarium is significant. In the present taxonomic system, about 20 Fusarium species or varieties are linked to Cosmospora sensu Rossman (Gräfenhan et al. 2008). There has been a reluctance to apply the 1:1 genus concept or strict monophyly to the present generic concept of Fusarium, which exhibits a striking lack of correlation with teleomorph/holomorph generic concepts in the Nectriaceae. Species with teleomorphs classified in other orders of ascomycetes were excluded from Fusarium some time ago, namely Microdochium nivale (Xylariales, Samuels & Hallett 1983) and Plectosporium tabacinum (Glomerellales, Palm et al. 1995). As now delimited, Fusarium is still linked to six teleomorph genera in the Nectriaceae, i.e. Albonectria, Cosmospora, Cyanonectria, Gibberella (the teleomorph genus associated with the type species of Fusarium), and Haematonectria, with some species remaining in Nectria sensu Booth. Members of a seventh genus, Neocosmospora, fall into the Fusarium solani/Haematonectria clade (O'Donnell et al. 2008), but no Fusarium-like macroconidia are produced by these species.

Throughout the modern history of Fusarium, taxonomists have consistently recognised the distinctiveness of several groups of species first considered as discrete taxonomic sections by Wollenweber (1931). Most species of sections Eupionnotes, Macroconia, Pseudomicrocera, and Arachnites produce characteristic colonies in vitro, growing slower and producing less aerial mycelium than species of other sections (Gerlach & Nirenberg 1982), often with spreading orange, macroconidial slime known as pionnotes. As shown for most taxonomic sections of Fusarium, sections Eupionnotes and Macroconia are polyphyletic (O'Donnell 1993, Torzilli et al. 2002, Schroers et al. 2009). Some of the morphological characters used to define these sections, including macroconidial shape and colony characters in vitro, are plesiomorphic and shared by distantly related species. For Acremonium-like anamorphs, a similar or even more complex pattern of plesiomorphy is known; preliminary revisions to that generic concept are presented by Summerbell et al. (2011).

Although there have been discussions of narrowing the generic concept of Fusarium at specialist symposia, arguments have not been presented in print nor have nomenclatural changes been proposed. The prevailing concept of Fusarium is essentially that of Wollenweber (1931) and Wollenweber & Reinking (1935) with the exclusion of some species; for convenience we refer to this concept as Fusarium sensu Wollenweber. The need to reevaluate more than 20 anamorph generic names considered synonyms of Fusarium has caused some hesitancy in modifying this concept; these type studies are initiated here. Previous studies provided inconclusive phylogenetic evidence to demonstrate the distinctiveness of the Gibberella and Cosmospora clades, but sampled inadequately from other anamorph and teleomorph genera in the Nectriaceae (O'Donnell 1993, Zhang & Zhuang 2006, Luo & Zhuang 2008, Samuels et al. 2009). We sampled more broadly here, including 93 species originally assigned to about 11 teleomorph and 13 anamorph genera.

Our phylogenetic analysis, combined with morphological and ecological considerations, suggests the recognition of about 13 well supported lineages within Cosmospora sensu Rossman that can be recognised at the generic level. Fusarium sensu Wollenweber splits into two major groups, which we will refer to as the “terminal Fusarium clade” centred on Gibberella, and a collection of lineages in the basal part of the Nectriaceae that we will refer to as the “basal Fusarium-like clades”. In the latter, we resurrect the genera Dialonectria, Fusicolla, Microcera, and Stylonectria for species and varieties of the former Fusarium sections Arachnites, Eupionnotes, Macroconia, Pseudomicrocera, and Submicrocera, Acremonium section Nectroidea, and several fungicolous, entomogenous, and soil-borne species classified in Cosmospora sensu Rossman. Cosmospora s. str. is redelimited as a morphologically and phylogenetically restricted genus including only species with anamorphs originally ascribed to Acremonium or Verticillum. We raise Nectria sect. Macroconia to generic rank for a small group of species with large Fusarium-like macroconidia and minute perithecia. We epitypify the classical hyphomycete genus Atractium, sometimes listed as a synonym of Fusarium, and consider two other anamorph genera associated with Cosmospora, namely Mariannaea and Volutella.

The result is a revision of the Cosmospora sensu Rossman clade into segregate genera that should provide phylogenetic clarity to subsequent monographic revisions and facilitate the description of new species in appropriate genera. The basal Fusarium-like clades, for the most part the slow growing pionnotal species formerly associated with Cosmospora sensu Rossman, are distributed in seven monophyletic genera, six of them provided with pre-existing generic names. Another paper concerns genera of the terminal Fusarium clade, including the former Nectria desmazieri, with teleomorphs that morphologically are somewhat Cosmospora-like (Schroers et al. 2011).

In common with the papers by Schroers et al. (2011) and Summerbell et al. (2011), we adopt a single-name nomenclature, employing the oldest available generic name in combination with the oldest available species epithet, irrespective of whether these names could be interpreted as teleomorphic or anamorphic. In some cases these cross-morph combinations violate Article 59. In our opinion, the International Code of Botanical Nomenclature (ICBN, McNeill et al. 2006) should be like any legal code and be governed by its own basic principles. This is analogous to a constitution; when laws within a legal structure are found to be unconstitutional, they are rejected. Art. 59 violates Principle III of the ICBN, that the correct name is based on priority of publication. We give precedence to the Principles rather than the contradictory article and essentially reject Art. 59 as unconstitutional. According to Art. 59, when a valid and legitimate name is transferred into a genus that does not match its karyological type, i.e. an anamorph epithet is moved into a teleomorph genus or visa versa, the name can be considered superfluous or incorrect or contrary to Art. 59.1, but the resulting binomial is still valid and legitimate. By this interpretation, combination of a valid, legitimate anamorph-typified epithet to a teleomorph-typified generic name or a valid teleomorph-typified epithet to an anamorph-typified generic name, results in a binomial that is incorrect for the holomorph. Incorrect names may become correct later (cf. Art. 52.3) provided they have a valid/legitimate basionym and the part of the Code (i.e. Art. 59) that makes the names incorrect is changed. According to the title of Chapter VI of the Code, Art. 59 only applies to pleiomorphic fungi, i.e. species where both the teleomorph and anamorph(s) are known. In this interpretation, names for monomorphic species resulting from the transfer of anamorph epithets into teleomorph-typified genera or visa versa would be correct, valid, and legitimate. In this paper, we explicitly state which names may be “incorrect” according to this interpretation of the present Code. However, we hope that the growing support for single name nomenclature that was evident at the International Mycological Congresses in 2002, 2006, and 2010 will discourage anyone from attempting to “correct” them.

MATERIALS AND METHODS

Fungal isolates and herbarium specimens

Ninety-three taxa of Nectriaceae were included in the phylogenetic study with Acremonium lichenicola selected as outgroup (Table 1) based on prior analyses (Gräfenhan et al. 2008). Morphological observations of colonies and anamorph characters are based on strains grown on potato-dextrose agar (PDA; Difco), cornmeal agar (CMA; Acumedia, Lansing, Michigan) and synthetic low nutrient agar (SNA; Nirenberg 1976) in the laboratory at room temperature (about 22–25 °C) under ambient light conditions. Measurements for some structures are presented as a range of one standard deviation above and below the calculated mean, with extreme observed values given in parentheses, and the number of measured structures noted. Colour codes refer to Kornerup & Wanscher (1978). Herbarium abbreviations are from Holmgren et al. (1990). Abbreviations of culture collections follow the World Federation of Culture Collections code (wdcm.nig.ac.jp/wfcc).

Table 1.

Taxa used in molecular phylogenetic analysis.

Unitary names used in phylogeniesTeleomorph name (most recent)Anamorph name (most recent)StrainOther No.Collector/DepositorCountrySubstratumGenBank Accession No.
rpb2acl1ITSLSU
Acremonium lichenicola W. Gams Acremonium lichenicola W. Gams CBS 425.66* K.W. Gams Germany Betula sp., old leaf HQ897724 HQ897861
Acremonium macroclavatum Ts. Watan. Acremonium macroclavatum Ts. Watan. CBS 123922* MAFF 238162 T. Watanabe Japan Soil HQ897740 HQ897876 HQ897806
Acremonium tsugae W. Gams Acremonium tsugae W. Gams CBS 788.69* J.E. Bier Canada Tsuga heterophylla HQ897728 HQ897865
“Albonectria” albida (Rossman) Guu & Y.M. Ju Albonectria albida (Rossman) Guu & Y.M. Ju BBA 67603* ATCC 44543; BBA 65209; C.T.R. 71-110 C. T. Rogerson Jamaica Bark of woody stem HQ897738 HQ897874 HQ897804
Albonectria albosuccinea (Pat.) Rossman & Samuels Albonectria albosuccinea (Pat.) Rossman & Samuels BBA 64502* ATCC 44544; C.T.R. 71-188; NRRL 20459 C.T. Rogerson Venezuela Wood HQ897699 HQ897837 HQ897788 U34554
Albonectria rigidiuscula (Berk. & Broome) Rossman & Samuels Albonectria rigidiuscula (Berk. & Broome) Rossman & Samuels Fusarium decemcellulare Brick CBS 122570 BPI 863840; G.J.S. 01-170 G.J. Samuels Cameroon Bark HQ897760 HQ897896 HQ897815
“Albonectria” verrucosa (Pat.) Rossman & Samuels Albonectria verrucosa (Pat.) Rossman & Samuels CBS 102163 ATCC 208923; BBA 64786; G.J.S. 84-426 G.J. Samuels Venezuela Recently cut bamboo HQ897784 HQ897920
Atractium crassum (Wollenw.) Seifert & Gräfenhan Fusarium merismoides var. crassum Wollenw. CBS 180.31* NRRL 20894 H.W. Wollenweber Germany Water tap HQ897722 HQ897859 U88110
Atractium stilbaster Link Stilbella fusca (Sacc.) Seifert DAOM 215627 K.A. Seifert Canada / Quebec Cut stump HQ897748 HQ897884 HQ843769
Chaetopsina penicillata Samuels Chaetopsinectria chaetopsinae-penicillatae (Samuels) J. Luo & W.Y. Zhuang Chaetopsina penicillata Samuels CBS 608.92* ATCC 56205; G.J.S. 77-21 G.J. Samuels New Zealand Beilschmiedia tawa, bark HQ897709 HQ897847 HQ897798
Cosmospora arxii (W. Gams) Gräfenhan & Schroers Acremonium arxii W. Gams CBS 748.69* K.W. Gams Germany Hypoxylon sp. HQ897725 HQ897862
Cosmospora butyri (J.F.H. Beyma) Gräfenhan, Seifert & Schroers Tilachlidium butyri J.F.H. Beyma CBS 301.38* MUCL 9950 Knudson Denmark Butter HQ897729 HQ897866
Cosmospora coccinea Rabenh. Cosmospora coccinea Rabenh. Verticillium olivaceum W. Gams CBS 341.70 K.W. Gams Germany Hymenium of Inonotus nodulosus on Fagus sylvatica HQ897777 HQ897913 HQ897827
Cosmospora cymosa (W. Gams) Gräfenhan & Seifert Acremonium cymosum W. Gams CBS 762.69* K.W. Gams Germany Inonotus radiatus, decaying fruiting body HQ897778 HQ897914 HQ897828
Cosmospora khandalensis (Thirum. & Sukapure) Gräfenhan & Seifert Cephalosporium khandalense Thirum. & Sukapure CBS 356.65* ATCC 16091; IMI 112790; MUCL 7974 M.J. Thirumalachar India Bambusa sp., decaying stem HQ897723 HQ897860
Cosmospora lavitskiae (Zhdanova) Gräfenhan & Seifert Gliomastix lavitskiae Zhdanova CBS 530.68* ATCC 18666; IMI 133984 L.A. Beljakova Ukraine Plant debris on surface soil HQ897726 HQ897863
“Cosmospora” stegonsporii Rossman, D.F. Farr & Akulov Cosmospora stegonsporii Rossman, D.F. Farr & Akulov CBS 122305* A.R. 4385; BPI 878274 A.Y. Akulov Ukraine Stegonsporium pyriforme on bark HQ897733 HQ897869
Cosmospora cf. viridescens (C. Booth) Gräfenhan & Seifert Nectria cf. viridescens C. Booth CBS 102433 M. Reblova Czech Republic Tilia sp., dead tree HQ897712 HQ897850
Cosmospora sp. CBS 213.70 K.W. Gams Poland Fomitopsis pinicola HQ897727 HQ897864
Cyanonectria buxi (Fuckel) Schroers, Gräfenhan & Seifert Gibberella buxi (Fuckel) G. Winter Fusarium buxicola Sacc. BBA 64985 M.E. Noordeloos Netherlands Buxus sempervirens HQ897746 HQ897882 HQ897809
Cyanonectria cyanostoma (Sacc. & Flageolet) Samuels & Chaverri Cyanonectria cyanostoma (Sacc. & Flageolet) Samuels & Chaverri BBA 70964* BPI 748307; CBS 101734; G.J.S. 98-127 G.J. Samuels & F. Candoussau France Buxus sempervirens, bark HQ897759 HQ897895 HQ897814 FJ474076
Cylindrocladium sp. CBS 125514 K.A.S. 1732 K.A. Seifert New Zealand Soil under Leptospermum scoparium HQ897735 HQ897871 HQ897801
Cylindrodendrum sp. DAOM 226786 K.A.S. 872 K.A. Seifert Australia / New South Wales Rotten wood HQ897750 HQ897886 HQ843773
Dialonectria cf. episphaeria (Tode : Fr.) Cooke Cosmospora cf. episphaeria (Tode : Fr.) Rossman & Samuels CBS 125494 DAOM 235830; T.G. 2006-11 T. Gräfenhan Canada / Ontario Old ascomycete stromata on deciduous tree HQ897756 HQ897892 HQ897811
Dialonectria ullevolea Seifert & Gräfenhan Fusarium aquaeductuum var. medium Wollenw. CBS 125493 DAOM 235827; T.G. 2007-56 T. Gräfenhan USA / Pennsylvania Ascomycete on Fagus americana HQ897782 HQ897918
“Fusarium” cavispermum Corda Fusarium cavispermum Corda BBA 64137 CBS 184.77; NRRL 20837; NRRL 22279 T. Nilsson Sweden Untreated pine pole HQ897762 HQ897898
“Fusarium” ciliatum (Alb. & Schw.) Link Fusarium ciliatum (Alb. & Schw.) Link BBA 62172 ATCC 16068; ATCC 24137; CBS 191.65; CBS H-12687; IMI 112499; NRRL 20431 H. Richter Germany On Fagus sylvatica HQ897764 HQ897900 HQ897818 AF228349
“Fusarium” dimerum Penz. Fusarium dimerum Penz. CBS 254.50 NRRL 36384 Mack Netherlands Man, sputum HQ897695 EU926279
“Fusarium” domesticum (Fr.) Bachm. Fusarium domesticum (Fr.) Bachm. CBS 116517 NRRL 29976 K. O'Donnell Switzerland Cheese HQ897694 EU926219
Fusarium graminearum Schwabe Gibberella zeae (Schwein.) Petch Fusarium graminearum Schwabe NRRL 31084 PH-1 USA / Michigan Zea mays FGSG026590 FGSG060390
“Fusarium” lunatum (Ellis & Everh.) Arx Fusarium lunatum (Ellis & Everh.) Arx BBA 63199 CBS 632.76; NRRL 20690; NRRL 37067 W. Gerlach Germany Gymnocalcium damsii HQ897766 HQ897902 HQ897819
“Fusarium” melanochlorum (Casp.) Sacc. Fusarium melanochlorum (Casp.) Sacc. CBS 202.65 ATCC 16069; B 700014030; BBA 62248; NRRL 36353 H. Richter Austria Branch canker on Fagus sylvatica HQ897769 HQ897905 AF228353
“Fusarium” merismoides var. chlamydosporale Wollenw. Fusarium merismoides var. chlamydosporale Wollenw. CBS 179.31* NRRL 20839 H.W. Wollenweber USA / Wisconsin Ostrya virginiana HQ897721 U88109
“Fusarium” nematophilum Nirenberg & G. Hagedorn Fusarium nematophilum Nirenberg & G. Hagedorn BBA 70838 A. Westphal USA / California Beta vulgaris / Heterodera schachtii HQ897693 HQ897834 HQ897786
Fusarium sambucinum Fuckel Gibberella pulicaris (Fr.) Sacc. Fusarium sambucinum Fuckel BBA 70569 H.I. Nirenberg Germany Humulus lupulus HQ897751 HQ897887
Fusarium sublunatum Reinking Fusarium sublunatum Reinking BBA 62431*CBS 189.34; NRRL 13384; NRRL 20840 O.A. Reinking Costa Rica Soil of banana plantation HQ897780 HQ897916 HQ897830
Fusarium verticillioides (Sacc.) Nirenberg Gibberella moniliformis Wineland Fusarium verticillioides (Sacc.) Nirenberg NRRL 20956 FGSC 7600; FRC M-3125 USA / California Zea mays FVEG092860 FVEG046670
“Fusarium” sp. DAOM 235648 BBA 62195; CBS 119875; K.A.S. 2872; MRC 1652 R. Schneider Germany Solanum lycopersicum HQ897698 HQ897836 HQ897787
Fusicolla acetilerea (Tubaki C. Booth & T. Harada) Gräfenhan & Seifert Fusarium merismoides var. acetilereum Tubaki C. Booth & T. Harada BBA 63789* IMI 181488; NRRL 20827 Miyoshi Japan Polluted soil HQ897701 HQ897839 HQ897790 U88108
Fusicolla aquaeductuum (Radlk. & Rabenh.) Gräfenhan, Seifert & Schroers Fusarium aquaeductuum var. aquaeductuum (Radlk. & Rabenh.) Lagerh. BBA 63669 CBS 734.79; NRRL 20686 W. Gerlach Germany Drinking water HQ897742 HQ897878
Fusicolla betae (Desm.) Bonord. Fusarium betae (Desm.) Sacc. BBA 64317* C. Bauers Germany On young plants of Triticum aestivum HQ897781 HQ897917
Fusicolla epistroma (Höhn.) Gräfenhan & Seifert Fusarium epistroma (Höhn.) C. Booth BBA 62201* IMI 85601; NRRL 20439 W.G. Bramley UK Ascomycete on Betula sp. HQ897765 HQ897901 AF228352
Fusicolla matuoi (Hosoya & Tubaki) Gräfenhan & Seifert Cosmospora matuoi Hosoya & Tubaki Fusarium matuoi Hosoya & Tubaki CBS 581.78 ATCC 18694; MAFF 238445: NRRL 20427 T. Matsuo Japan Albizzia julibrissin HQ897720 HQ897858
Fusicolla violacea Gräfenhan & Seifert Fusarium merismoides var. violaceum W. Gerlach, nom. inval.CBS 634.76* BBA 62461; NRRL 20896 D. Ershad Iran Quadraspidiotus perniciosus on living on branch of Prunus domestica HQ897696 U88112
Geejayessia atrofusca (Schw.) Schroers & Gräfenhan Nectria atrofusca (Schwein.) Ellis & Everh. Fusarium staphyleae Samuels & Rogerson CBS 125482 DAOM 238118; T.G. 2006-01 T. Gräfenhan Canada / Ontario Staphylea trifolia, twigs HQ897775 HQ897911 HQ897825
Geejayessia celtidicola Gräfenhan & Schroers CBS 125481 DAOM 238129; T.G. 2006-29 T. Gräfenhan Canada / Ontario Celtis occidentalis, twigs HQ897772 HQ897908 HQ897822
Geejayessia cicatricum (Berk.) Schroers Nectria cicatricum (Berk.) Tul. & C. Tul. CBS 125550CBS H-20375; H.J.S. 1374 H.-J. Schroers & M. Žerjav Slovenia Buxus sempervirens, twigs HQ897697 HQ897835
Geejayessia desmazieri (Becc. & De Not.) Schroers, Gräfenhan & Seifert Nectria desmazieri Becc. & De Not. Fusarium fuckelii Sacc. CBS 313.34 NRRL 20474 E.W. Mason UK Buxus sempervirens, dead twig HQ897703 HQ897841 HQ897792 U88125
Geejayessia zealandica (Cooke) Schroers Cosmospora zealandica (Cooke) Samuels & Nirenberg Fusarium zealandicum Nirenberg & Samuels BBA 65034 BPI 802575; CBS 101913; G.J.S. 86-509 G.J. Samuels New Zealand Plagianthus, timber HQ897745 HQ897881 HQ897808
Haematonectria illudens (Berk.) Samuels & Nirenberg Haematonectria illudens (Berk.) Samuels & Nirenberg Fusarium illudens C. Booth BBA 67606 G.J.S. 82-98; NRRL 22090 G.J. Samuels New Zealand Beilschmiedia tawa HQ897692 HQ897833 AF178393 AF178362
Haematonectria ipomoeae (Halst.) Samuels & Nirenberg Haematonectria ipomoeae (Halst.) Samuels & Nirenberg Fusarium striatum Sherb. BBA 64379 NRRL 22147 H.I. Nirenberg Germany Passiflora edulis HQ897753 HQ897889
Heliscus lugdunensis Sacc. & Therry Nectria lugdunensis J. Webster Heliscus lugdunensis Sacc. & Therry CBS 125485 DAOM 235831; T.G. 2008-07 T. Gräfenhan USA / Arizona Populus fremontii, twigs in stream HQ897731 HQ897867
Heliscus submersus H.J. Huds. Heliscus submersus H.J. Huds. CBS 394.62* H.J. Hudson UK HQ897707 HQ897845 HQ897796
Macroconia leptosphaeriae (Niessl) Gräfenhan & Schroers Cosmospora leptosphaeriae (Niessl) Rossman & Samuels ? Fusarium sphaeriae var. majus Wollenw. CBS 100001CBS H-6030 L. Rommelaars Netherlands On Leptosphaeria on dead stem of Urtica dioica HQ897755 HQ897891 HQ897810
Macroconia papilionacearum (Seaver) Gräfenhan & Seifert Cosmospora papilionacearum (Seaver) Rossman & Samuels ?Fusarium gigas Speg. CBS 125495 DAOM 238119; T.G. 2007-03 T. Gräfenhan USA / Florida Black ascomycete on Fabaceae HQ897776 HQ897912 HQ897826
Macroconia sp. CBS 125496 T.G. 2008-08 T. Gräfenhan USA / Arizona Quercus sp., branch in stream of water HQ897732 HQ897868
Mariannaea elegans (Corda) Samson ? Nectria mariannaeae Samuels & Seifert Mariannaea elegans (Corda) Samson DAOM 226709 K.A.S. 948 K.A. Seifert Canada / Ontario Betula sp., wood HQ897747 HQ897883 HQ843768
Mariannaea samuelsii Seifert & Bissett DAOM 235814*CBS 125515; K.A.S. 1307 J. Bissett Guatemala Soil under Podocarpus HQ897752 HQ897888 HQ843767 HQ843766
Microcera coccophila Desm. Fusarium coccophilum (Desm.) Wollenw. & Reinking CBS 310.34 NRRL 13962 H.W. Wollenweber Italy Scale insect on Laurus nobilis HQ897705 HQ897843 HQ897794
Microcera diploa (Berk. & M.A. Curtis) Gräfenhan & Seifert Cosmospora diploa (Berk. & M.A. Curtis) Rossman & Samuels Fusarium coccidicola Henn. BBA 62173 CBS 735.79; NRRL 13966 W. Gerlach Iran Quadraspidiotus perniciosus on living on branch of Prunus domestica HQ897763 HQ897899 HQ897817
Microcera larvarum (Fuckel) Gräfenhan, Seifert & Schroers Fusarium larvarum Fuckel CBS 169.30 NRRL 22102 H.W. Wollenweber Japan Aphids on Pyrus communis HQ897717 HQ897855
Microcera rubra Gräfenhan & Seifert Fusarium larvarum var. rubrum W. Gerlach, nom. inval. BBA 62460*CBS 638.76; NRRL 20475; NRRL 22111; NRRL 22170 W. Gerlach Iran Quadraspidiotus perniciosus on living on branch of Prunus domestica HQ897767 HQ897903 HQ897820
Nalanthamala diospyri (Crand.) Schroers &. M.J. Wingfield Nalanthamala diospyri (Crand.) Schroers &. M.J. Wingfield CBS 429.89 ATCC 22206 B.S. Crandall USA / Mississippi Diospyros virginiana HQ897718 HQ897856
“Nectria” cinereopapillata Henn. & E. Nyman Nectria cinereopapillata Henn. & E. Nyman CBS 264.36 H.W. Wollenweber Sierra Leone Cassia sieberiana HQ897710 HQ897848 HQ897799
“Nectria” diminuta Berk. Cosmospora diminuta (Berk.) Rossman & Samuels CBS 114636 BPI 864173; G.J.S. 00-181 G.J. Samuels USA / North Carolina Quercus virginiana, dead tree HQ897758 HQ897894 HQ897813
“Nectria” cf. flavoviridis (Fuckel) Wollenw. Nectria flavoviridis (Fuckel) Wollenw. BBA 65542 G.J. Samuels USA / New York On fungus on decorticated wood HQ897702 HQ897840 HQ897791
“Nectria” magnoiiae M.L. Lohman & Hepting Nectria magnoliae M.L. Lohman & Hepting CBS 380.50* BPI 552527 G.H. Hepting USA / North Carolina Liriodendron tulipifera HQ897713 HQ897851
Nectria miltina (Mont.) Mont. Nectria miltina (Mont.) Mont. CBS 125499 T.G. 2008-02 T. Gräfenhan USA / Arizona Yucca elata HQ897730
Nectria nigrescens Cooke Nectria nigrescens Cooke CBS 125500 DAOM 235832; T.G. 2006-18 T. Gräfenhan Canada / Ontario Acer sp., twig HQ897757 HQ897893 HQ897812
Nectria pseudotrichia Berk. & M.A. Curtis Nectria pseudotrichia Berk. & M.A. Curtis Tubercularia lateritia (Berk.) Seifert DAOM 235820 T.G. 2007-41 T. Gräfenhan USA / Florida Dead herbaceous plant HQ897706 HQ897844 HQ897795
“Nectria” rishbethii C. Booth Cosmospora rishbethii (C. Booth) Rossman & Samuels CBS 496.67* IMI 070248b; MUCL 4133 J. Rishbeth UK Pinus sylvestris, stump HQ897714 HQ897852
“Nectria” rubropeziza Wollenw. Nectria rubropeziza Wollenw. CBS 234.31* H.W. Wollenweber USA / Maryland Tree trunk HQ897708 HQ897846 HQ897797
“Nectria” setofusariae Samuels & Nirenberg Nectria setofusariae Samuels & Nirenberg Fusarium setosum Nirenberg & Samuels CBS 635.92 A.R. 3333; BBA 65063; BPI 1113176; G.J.S. 88-12 A.Y. Rossman French Guiana Bark of recently cut tree HQ897704 HQ897842 HQ897793
“Nectria” ventricosa C. Booth Nectria ventricosa C. Booth Fusarium ventricosum Appel & Wollenw. BBA 62452 CBS 748.79; NRRL 20846; NRRL 22113 K.H. Domsch Germany Wheat field soil HQ897761 HQ897897 HQ897816 L36613
“Nectria” ventricosa C. Booth Nectria ventricosa C. Booth Fusarium ventricosum Appel & Wollenw. CBS 430.91 NRRL 25729 U. Kuchenbäcker Germany Robinia pseudoacacia, twig HQ897771 HQ897907
Nectria sp. CBS 125498 T.G. 2006-33 T. Gräfenhan Canada / Ontario Abies balsamea HQ897737 HQ897873 HQ897803
Neocosmospora vasinfecta E.F. Sm. Neocosmospora vasinfecta E.F. Sm. NRRL 22166 ATCC 62199 L.M. Carris USA / Illinois Heterodera glycines EU329497 DQ094319 DQ236361
Neonectria coccinea (Pers.) Rossman & Samuels Neonectria coccinea (Pers.) Rossman & Samuels Cylindrocarpon candidum (Link) Wollenw. CBS 125484 DAOM 235835; T.G. 2007-17 T. Gräfenhan Germany Fagus sylvatica HQ897785 HQ897921 HQ897832
Neonectria ditissima (Tul. & C. Tul.) Samuels & Rossman Neonectria ditissima (Tul. & C. Tul.) Samuels & Rossman Cylindrocarpon heteronema (Berk. & Broome) Wollenw. CBS 125486 DAOM 235836; T.G. 2006-21 T. Gräfenhan Canada / Ontario Fagus americana, branch HQ897774 HQ897910 HQ897824
Neonectria fuckeliana (C. Booth) Castl. & Rossman Neonectria fuckeliana (C. Booth) Castl. & Rossman CBS 239.29* IMI 039700 H.W. Wollenweber UK Picea sitchensis, bark HQ897711 HQ897849
Pseudonectria buxi (DC.) Seifert, Gräfenhan & Schroers Pseudonectria rousseliana (Mont.) Wollenw. Volutella buxi (DC.) Berk. CBS 125483 T.G. 2007-69A K.W. Gams Spain Buxus sempervirens, leaves HQ897719 HQ897857 HQ897800
“Pseudonectria” pachysandricola B.O. Dodge Pseudonectria pachysandricola B.O. Dodge Volutella pachysandricola B.O. Dodge DAOM 195309 E.J. Mathers USA / Florida Pachysandra sp., nursery stock HQ897743 HQ897879 HQ897807
Pseudonectria sp. BBA 71336 H.I. Nirenberg Germany Buxus sempervirens, leaves HQ897741 HQ897877
Stylonectria cf. applanata Höhn. Nectria applanata var. succinea Höhn. CBS 125489 T.G. 2008-24 T. Gräfenhan Canada / Ontario Ascomycete on Betula sp. HQ897739 HQ897875 HQ897805
Stylonectria carpini Gräfenhan Nectria applanata Fuckel DAOM 235819 W.J. 3013 H. Voglmayr Austria On Melanconis spodiaea on Carpinus betulus HQ897773 HQ897909 HQ897823
Stylonectria purtonii (Grev.) Gräfenhan Cosmospora purtonii (Grev.) Rossman & Samuels DAOM 235818 T.G. 2007-30 T. Gräfenhan Germany On small branches of Picea abies HQ897783 HQ897919 HQ897831
Stylonectria wegeliniana (Rehm) Gräfenhan, Voglmayr & Jaklitsch Cosmospora wegeliniana (Rehm) Rossman & Samuels CBS 125490 WU 29855 H. Voglmayr Austria Stromata of Hapalycystis bicaudata on Ulmus glabra HQ897754 HQ897890
Stylonectria sp. CBS 125491 T.G. 2007-21 T. Gräfenhan Germany Ascomycete on Carpinus / Ulmus ? HQ897779 HQ897915 HQ897829
Thelonectria discophora (Mont.) P. Chaverri & C. Salgado Neonectria discophora (Mont.) Mantiri & Samuels Cylindrocarpon ianthothele var. majus Wollenw. CBS 125487 DAOM 235837; T.G. 2007-34 T. Gräfenhan Germany Aesculus hippocastanum HQ897700 HQ897838 HQ897789
Thelonectria lucida (C. Booth) P. Chaverri & C. Salgado Cylindrocarpon lucidum C. Booth DAOM 226723 K.A.S. 1007 K.A. Seifert Canada / British Columbia Pseudotsuga menziesii, root HQ897734 HQ897870
Volutella ciliata (Alb. & Schwein.) Fr. Volutella ciliata (Alb. & Schwein.) Fr. DAOM 226718 K.A.S. 972 J.A. Traquair Canada / Ontario Agricultural soil HQ897736 HQ897872 HQ897802
Volutella citrinella (Cooke & Massee) Seifert Cosmospora stilbellae (Samuels & Seifert) Rossman & Samuels Stilbella aciculosa (Ellis & Everh.) Seifert DAOM 226720 K.A.S. 978 R.J. Bandoni & A.A. Bandoni Canada / British Columbia Solanum tuberosum, debris HQ897770 HQ897906 HQ897821 HQ843771
Volutella consors (Ellis & Everh.) Seifert, Gräfenhan & Schroers Cosmospora consors (Ellis & Everh.) Rossman & Samuels Volutella minima Höhn. CBS 328.77 C.T.R. 72-347 C.T. Rogerson USA / North Carolina Magnolia fraseri, old inflorescence HQ897716 HQ897854
Volutella consors (Ellis & Everh.) Seifert, Gräfenhan & Schroers Cosmospora consors (Ellis & Everh.) Rossman & Samuels Volutella minima Höhn. CBS 139.79 G.H. Boerema Netherlands Decaying orchid bulb HQ897715 HQ897853
— A.R. = Amy Y. Rossman personal collection; ATCC = American Type Culture Collections, Manassas, Virginia, USA; B = Mycological Herbarium at the Botanical Museum, Berlin, Germany; BBA = Julius Kühn-Institute, Institute for Epidemiology and Pathogen Diagnostics, Berlin & Braunschweig, Germany; BPI = U.S. National Fungus Collections, USDA, ARS, Beltsville, Maryland, USA; CBS = Centraalbureau voor Schimmelcultures, Utrecht, The Netherlands; C.T.R. = Clark T. Rogerson personal collection; DAOM = Canadian National Mycological Herbarium and Culture Collection, AAFC, Ottawa, Ontario, Canada; FGSC = Fungal Genetics Stock Center, School of Biological Sciences, University of Missouri, Kansas City, Missouri, USA; FRC = Fusarium Research Center, Department of Plant Pathology, Penn State University, University Park, Pennsylvania, USA; G.J.S. = Gary J. Samuels personal collection; H.J.S. = Hans-Josef Schroers personal collection; IMI = CABI Bioservices, Egham, Surrey, UK; K.A.S. = Keith A. Seifert personal collection; MAFF = Microbial Culture Collection, National Institute of Agrobiological Sciences, Tsukuba, Japan; MRC = Microbial Culture Collection, South African Medical Research Council, Tygerberg, South Africa; MUCL = (Agro)Industrial Fungi & Yeasts Collection, Université Catholique de Louvain, Louvain-la-Neuve, Belgium; NRRL = ARS Culture Collection, USDA, NCAUR, Peoria, Illinois, USA; T.G. = Tom Gräfenhan personal collection; W.J. = Walter Jaklitsch personal collection; WU = Herbarium, Department of Plant Systematics and Evolution, Faculty of Life Sciences, University Vienna, Austria.
*— Type or other authentic material.
— GenBank accession numbers beginning with HQ were newly generated. All other sequences were obtained from GenBank.
0— Locus number in the Fusarium genome database (http://www.broad.mit.edu/annotation/fungi/fusarium)

DNA extractions, PCR and DNA sequencing

DNA extractions were performed using UltraClean Microbial DNA Isolation Kits (MO BIO Laboratories Inc., Carlsbad, California) from mycelium scraped from colonies grown on PDA using a sterile scalpel. DNA concentration and quality were determined by Nanodrop ND-1000 spectrometer (Thermo Scientific, Wilmington, Delaware) and preparations were diluted to 1–5 ng/μL of DNA template.

The second largest subunit of the RNA polymerase II (rpb2) was amplified following the protocol of de Cock & Lévesque (2004) using the primer combinations 5F2/7cR and 7cF/11aR (O'Donnell et al. 2007) in a total reaction volume of 20 μL. PCR products of the larger subunit of the ATP citrate lyase (acl1, Nowrousian et al. 2000) was amplified using the newly designed primers acl1-230up (5'-AGC CCG ATC AGC TCA TCA AG-3') and acl1-1220low (5'-CCT GGC AGC AAG ATC VAG GAA GT-3') in a total reaction volume of 20 μL following the same protocol. PCR reactions were placed in an Eppendorf thermal cycler (Westbury, New York) and processed with the following temperature profile for the rbp2 regions: 3 min at 95 °C (initial denaturation), 5 cycles 45 s at 95 °C (denaturation), 45 s at 60 °C (annealing), 2 min at 72 °C (extension), followed by 5 cycles with annealing at 58 °C, followed by 30 cycles with annealing at 54 °C, with a final extension 8 min at 72 °C. The temperature profile for the acl1 region was as follows: 3 min at 95 °C, 5 cycles 45 s at 95 °C, 45 s at 64 °C, 2 min at 72 °C, followed by 5 cycles with annealing at 62 °C, followed by 30 cycles with annealing at 56 °C, with a final extension 8 min at 72 °C. For forward and reverse strands, sequencing reactions were performed directly without cleaning PCR amplicons, using a BigDye sequencing kit (Applied Biosystems, Foster City, California) on an ABI3130 DNA Analyzer (Applied Biosystems). The following profile was used for the sequencing reactions: 95 °C for 3 min, then for 40 cycles at 95 °C for 30 s, 50 °C for 15 s, 60 °C for 2 min. Contig sequences were assembled using Sequencher v. 4.9 (Gene Codes Corporation, Ann Arbor, Michigan) and aligned manually using BioEdit 7 (Hall 1999). Protein coding DNA sequences were aligned along the reading frame of the corresponding amino acid sequence and divided into 3 partitions, rpb2 region 5–7, rpb2 region 7–11, and acl1. Intergenic spacer regions and introns of the rpb2 and acl1 gene sequences could not be reliably aligned and were excluded from the final alignment. Additional ITS sequences were generated for some of the species mentioned below using the methods described by Nguyen & Seifert (2008).

All DNA sequences generated in this study are deposited in GenBank (accession numbers listed in Table 1 and in the Taxonomy part as barcodes). We have designated some of these as “DNA barcodes” when they represent type, authentic, or thoroughly validated strains.

Phylogenetic analyses

The combined and partitioned data set of the protein encoding regions of rpb2 and acl1 was used to search for the best maximum likelihood (ML) tree employing the GARLI v. 1 software (Zwickl 2006) implemented by the CIPRES project at the San Diego Supercomputer Center (www.phylo.org). The best-fit substitution model under the Akaike information criterion (Akaike 1974) was determined by using Modeltest v. 3.7 (Posada & Crandall 1998) and PAUP v. 4.0b10 (Swofford 2003). The GTR + I + G nucleotide substitution model was selected, which assumes an estimated proportion of invariant sites and 8 gamma-distributed rate categories to account for rate heterogeneity across sites. 100 independent ML heuristic phylogenetic analyses were performed using a starting tree generated by stepwise-addition (attachmentspertaxon = 2) and 10 000 generations without topology improvement parameter.

To correct for positive and divergent selection in molecular evolution of protein encoding DNA sequences, ML analyses were performed with GARLI using a codon substitution model that considers the ratio of nonsynonymous (dN) to synonymous (dS) rates of nucleotide substitution (dN/dS = ω). The GTR-like substitution model was selected with F3×4 codon frequencies (observed frequency at each codon position) and dN/dS values and proportions falling in three discrete categories ω1 < ω2 < ω3 (M3 model with site classes K = 3, Yang et al. 2000). Ten independent ML heuristic phylogenetic analyses were performed using a starting tree generated by stepwise-addition (attachmentspertaxon = 2) and 10 000 generations without improving the topology parameter.

Non-parametric bootstrapping of 1 000 ML pseudo-replicates of the data was used to assess clade support with GARLI. Because of the extended time necessary for ML bootstrap analysis under the M3 codon model, the measure of clade support was calculated using the parameters of the GTR + I + G nucleotide model given above. ML bootstrap probabilities (ML-BP) for the splits were mapped onto the best phylogenetic tree inferred under the M3 codon substitution model using SumTrees of the DendroPy v. 3.7 phylogenetic computing library (Sukumaran & Holder 2010).

Bayesian posterior probabilities (PP) were obtained from the combined and partitioned rpb2/acl1 data set using MrBayes v. 3.1.2 (Ronquist & Huelsenbeck 2003) implemented by the CIPRES project (see above). The GTR + I + G substitution model was selected assuming an estimated proportion of invariant sites and 4 gamma-distributed rate categories to account for rate heterogeneity across sites. Two independent Markov chain Monte Carlo analysis (MCMC) runs each with 4 chains were performed simultaneously. The analysis was run for 10 000 000 generations, sampling every 1 000 generations for a total of 10 001 trees. The first 1 500 000 generations were discarded as burn-in. Each of the two independent MCMC runs yielded 8 501 trees from each partition. The resulting six tree files (total 51 006 trees) were used to calculate PPs. These posterior probabilities were mapped onto the best phylogenetic tree using SumTrees of the DendroPy package.

Heuristic searches for the most parsimonious (MP) trees using PAUP v. 4.0b10 (Swofford 2003) were based on 1 026 parsimony informative, unordered and equally weighted characters; gaps were treated as missing data. Starting trees were obtained via 100 stepwise, random addition sequences. Other settings included auto-increase for MAXTREES, the tree-bisection-reconnection branch-swapping algorithm, the MULTREES option, and assigning any possible character state to an internal node with STEPMATRIX. MP bootstrap probabilities (MP-BP) were assessed by 1 000 heuristic pseudoreplicates using the same settings as above but with 20 stepwise, random addition sequences. By using SumTrees of the DendroPy package, the MP-BP support for the splits were mapped onto the best phylogenetic tree.

RESULTS

Sequence alignment

The combined and partitioned data set of two protein encoding genes for 93 taxa and outgroup consisted of 2 250 bp, translating to 750 amino acids. The rpb2 sequences (1 764 bp) had two coding regions (rpb2 5–7 and rpb2 7–11) with an intergenic spacer, which was removed from the final alignment. The acl1 amplicon comprised a coding region of 420 bp and a single intron of 200–500 bp, which was also removed.

Phylogenetic analyses

One hundred independent ML analyses under the GTR + I + G nucleotide substitution model of the combined and partitioned data set (rpb2 5–7 with 488 parsimony-informative characters, rpb2 7–11 with 387 parsimony-informative characters, and acl1 with 206 parsimony-informative characters) resulted in a single best ML tree with –lnL = –57,309.9782 (not shown). The parameters for the GTR + I + G model of the rpb2 5–7 partition were as follows: Estimated base frequencies; A = 0.2098, C = 0.2885, G = 0.2691, T = 0.2326; substitution rates AC = 2.104, AG = 6.386, AT = 2.011, CG = 0.767, CT = 9.725, GT = 1.000; proportion of invariable sites I = 0.3861; gamma distribution shape parameter α = 0.8858. The parameters for the GTR + I + G model of the rpb2 7–11 partition were as follows: Estimated base frequencies; A = 0.2033, C = 0.3050, G = 0.2538, T = 0.2379; substitution rates AC = 1.680, AG = 7.167, AT = 2.089, CG = 0.914, CT = 10.966, GT = 1.000; proportion of invariable sites I = 0.5253; gamma distribution shape parameter α = 0.8815. The parameters for the GTR + I + G model of the acl1 partition were as follows: Estimated base frequencies; A = 0.1774, C = 0.3655, G = 0.2369, T = 0.2202; substitution rates AC = 0.982, AG = 2.844, AT = 0.638, CG = 0.839, CT = 7.876, GT = 1.000; proportion of invariable sites I = 0.4834; gamma distribution shape parameter α = 0.9192.

Ten independent ML analyses under the codon substitution model (M3 with K = 3) of the combined and partitioned data set (rpb2 5–7 with 294 parsimony-informative characters, rpb2 7–11 with 292 parsimony-informative characters, and acl1 with 145 parsimony-informative characters) resulted in a single best ML tree with –lnL = –54,991.4885 (Fig. 1). The parameters for the M3 codon model of the rpb2 5–7 partition were as follows: 61 empirical codon frequencies (F3×4 method); substitution rates AC = 1.234, AG = 2.380, AT = 1.222, CG = 0.743, CT = 2.758, GT = 1.000; and three estimated nonsynonymous rate categories ω1 = 0.0020 with p1 = 0.6471, ω2 = 0.0726 with p2 = 0.2452, ω3 = 0.3214 with p3 = 0.1077. The parameters for the M3 codon model of the rpb2 7–11 partition were as follows: 61 empirical codon frequencies (F3×4 method); substitution rates AC = 1.023, AG = 2.820, AT = 1.177, CG = 0.933, CT = 2.489, GT = 1.000; and three estimated non-synonymous rate categories ω1 = 0.0020 with p1 = 0.8918, ω2 = 0.0925 with p2 = 0.0985, ω3 = 0.5436 with p3 = 0.0097. The parameters for the M3 codon model of the acl1 partition were as follows: 61 empirical codon frequencies (F3×4 method); substitution rates AC = 1.863, AG = 3.515, AT = 1.290, CG = 1.264, CT = 3.346, GT = 1.000; and three estimated non-synonymous rate categories ω1 = 0.0031 with p1 = 0.8025, ω2 = 0.1007 with p2 = 0.1211, ω3 = 0.4420 with p3 = 0.0763. These dN/dS ratios (ω < 1) verify a significant departure from neutrality (ω ≈ 1) of the rpb2 and acl1 data partitions implying natural selection against changes of amino acids in the encoding genes studied.

An external file that holds a picture, illustration, etc.
Object name is 79fig1.jpg

(p. 89). Maximum likelihood (ML) tree under the M3 codon model inferred from combined rpb2 + acl1 gene sequence data set. Negative-log likelihood (-lnL) of the ML tree is –54,991.4885. Branches with ML-BP and MP-BP values of > 75 % and PP scores > 0.95 are in bold. Internodes that are supported with individual values of ML-BP or MP-BP > 75 % or PP scores > 0.95, respectively, are drawn in bold and grey. Symbols following strain numbers indicate different morphs known for the species: ○ = microconidial state, [open diamond] = Fusarium-like macroconidial state, [open triangle] = teleomorph. Vertical bars in red indicate members of Cosmospora sensu Rossman et al. (1999), yellow bars taxa of the basal Fusarium-like clade, and a dark grey bar species of the terminal Fusarium clade, respectively.

In comparison, the best ML tree for the M3 codon model received a significantly better negative-log likelihood score than the best ML tree under the GTR + I + G nucleotide substitution model. The topology of the phylograms did not differ for the clades studied. Only some basal lineages such as “Nectriadiminuta, “N.rubropeziza, and “Pseudonectriapachysandricola grouped differently using different substitution models, probably a result of long branch attraction.

Similarily, heuristic searches of the parsimony analysis yielded a single most parsimonious tree (not shown), which did not have a significantly different topology than that of the ML analyses. The MP tree was 14 023 steps with a consistency index (CI) of 0.152, a retention index (RI) of 0.492, a rescaled CI (RC) of 0.075, and a homoplasy index (HI) of 0.848.

1 000 ML pseudoreplicates, two independent MCMC analyses, and 1 000 heuristic bootstrap replicates of the combined and partitioned data set conducted with GARLI, MrBayes and PAUP, respectively, yielded majority consensus trees with highly concordant topologies (not shown) similar to that of the best ML tree generated for the M3 codon model. Internodes with significant clade support are drawn in thicker lines on the best ML tree topology (Fig. 1). Nodes were considered strongly supported when ML bootstrap proportions (ML-BP) is ≥ 75 %, Bayesian posterior probabilities (PP) is ≥ 0.95, and MP bootstrap proportions (MP-BP) is ≥ 75 % (Lutzoni et al. 2004).

Polyphyly of Cosmospora sensu Rossman

In the best ML tree (Fig. 1), species formerly placed in Cosmospora sensu Rossman fall into several major clades. The first major clade includes Volutella with four strains of three species, V. ciliata, V. citrinella (“Cosmospora” stilbellae) and V. consors (“C.” consors), in a strongly supported clade. Chaetopsina penicillata (= Chaetopsinectria or “Cosmospora” chaetopsinae-penicillatae) is a well supported sister species of Volutella, confirming the close phylogenetic relationship of Chaetopsina and Volutella (Zhang & Zhuang 2006, Luo & Zhuang 2010). Although not strongly supported, the Volutella/Chaetopsina group is the sister clade to a diverse fungal clade consisting of species of Calonectria, Cylindrodendrum, Heliscus, Mariannaea, and Neonectria.

The second major clade includes species formerly classified as Nectria applanata, Cosmospora purtonii, and C. wegeliniana. This clade is strongly supported and comprises species having ascomata with perithecial walls mainly consisting of two regions, and which are probably host-specific. These species are transferred to Stylonectria in the taxonomic section below.

The third and largest clade includes several subclades including the type species of Cosmospora, C. coccinea, and species with Fusarium-, Acremonium- and Verticillium-like anamorphs, which are classified in Cosmospora sensu stricto, Dialonectria, Fusicolla, Macroconia, and Microcera below. Cosmospora coccinea forms a strongly supported clade with other well-known species of the genus with Acremonium-like anamorphs, such as C. butyri, C. cymosa, and C. viridescens. This clade contains a group of species with similar microconidial anamorphs and a fairly constant ecological niche, delineating the new generic concept of Cosmospora s. str. Basal to Cosmospora is the strongly supported Dialonectria clade, which contains D. episphaeria and a new species, D. ullevolea. With “Nectria” rishbethii as a sister species, this subclade is delimited from another strongly supported subclade with species of Macroconia and Microcera, and Fusicolla matuoi. Macroconia and Microcera are sister clades, and include species such as Macroconia papilionacearum and Mac. leptosphaeriae as well as Microcera coccophila, Mic. diploa, and Mic. larvarum. These subclades, together with a few “residual” species classified in Fusarium such as “F.” cavispermum, “F.” ciliatum, “F.” melanochlorum, and “F.” merismoides var. chlamydosporale, are all phylogenetically distinct from the terminal Fusarium clade discussed below.

The terminal Fusarium clade contains a group of fungi with Cosmospora-like teleomorphs, of which only “Nectria” zealandica was formally combined in Cosmospora (Nirenberg & Samuels 2000). The terminal clade includes “Nectria” desmazieri and “N.” atrofusca, and is dealt with in more detail by Schroers et al. (2011).

As a singleton, “Nectria” diminuta does not group with any of the clades mentioned above. In all analyses under various substitution models (data not shown), “N.” diminuta fell neither in the terminal Fusarium clade nor the basal Fusarium-like clade nor any of the Cosmospora sensu Rossman groups (Fig. 1). This positional artifact may be caused by long-branch attraction or a paucity of parsimony-informative characters for the basal taxa in the combined DNA sequence data set.

Polyphyly of Fusarium sensu Wollenweber

The genus Fusarium is taxonomically linked to the teleomorph genus Gibberella, because they share the same species as type, F. sambucinum and G. pulicaris. In nature, Gibberella teleomorphs occur less frequently than their Fusarium anamorphs (Rossman et al. 1999). In the ML tree (Fig. 1), the Gibberella clade, representing Fusarium in the strict sense and including the type species in addition to F. graminearum, F. sublunatum, and F. verticillioides, is strongly supported. In Fig. 1 and Schroers et al. (2011), Gibberella is the sister clade to Cyanonectria. The terminal Fusarium clade in Fig. 1, including species with teleomorphs described in Albonectria, Cyanonectria, Gibberella, Haematonectria, and Neocosmospora, did not receive a statistically significant support similar to that obtained in other phylogenetic analyses (Schroers et al. 2009). The basal lineage of the terminal Fusarium clade is represented by the “Nectria” ventricosa species complex. Within the terminal Fusarium group, members of Albonectria and the Haematonectria/Neocosmospora species complex as well as the species pair “Albonectriaalbida and “Fusariumnematophilum always formed strongly supported groups.

The basal Fusarium-like clade, with numerous members formerly classified in Fusarium sections Arachnites, Eupionnotes, Macroconia, Pseudomicrocera, and Submicrocera, is phylogenetically and phenotypically distinct from the terminal Fusarium clade mentioned above. The basal clade splits into several subclades similar to what is described above for Cosmospora sensu Rossman. Therefore we have given these groups genus rank in the taxonomy part below.

Another genus of Fusarium-like species is represented by Atractium. Atractium crassum (“Fusarium” merismoides var. crassum) did not fall within the basal or terminal Fusarium clades. Together with Atractium stilbaster, it forms a strongly supported sister lineage to a group of fungi including species of Chaetopsina, Pseudonectria, and Volutella.

Polyphyly of Volutella sensu lato

As mentioned above, Volutella and Chaetopsina form a well supported lineage that is distinct from Cosmospora s. str. and the basal Fusarium-like clade. The type of the genus Pseudonectria, P. buxi, together with another similar species (BBA 71336), form a strongly supported sister group to the Volutella/Chaetopsina lineage. Chaetopsina separates Pseudonectria from species of Volutella s. str. In contrast to the above-mentioned clades, “Pseudonectria” pachysandricola and “Nectria” rubropeziza comprise a fairly well supported clade that branches off near the root of the tree and that separates the basal from the terminal Fusarium clade (Fig. 1). Thus, `“P.” pachysandricola is only distantly related to the type species of Pseudonectria and the Volutella s. str. group.

DISCUSSION

In revising the taxa associated with Cosmospora sensu Rossman, we focused on both teleomorph and anamorph phenotypes and ecological parameters guided by molecular phylogenetics. Resolving the taxonomy and nomenclature of Cosmospora requires resolving the phylogenetic relationships of many species presently included in Fusarium sensu Wollenweber. Previously published phylogenies of Fusarium, e.g. Summerbell & Schroers (2002), O'Donnell et al. (2010), sampled sparingly from teleomorphs of the Nectriaceae associated with other anamorph genera. It is clear from the analysis presented here in Fig. 1 and elsewhere in this volume by Chaverri et al. (2011), that as presently defined, Fusarium is not monophyletic. The basal Fusarium-like lineages and terminal Fusarium clade are separated by other genera that represent large genetic and taxonomic diversity. Although the sampling of species outside of the core Fusarium clade exceeds that of previous studies, this is still a relatively small subsample of these other genera. For example, Cylindrocladium, represented by one species here, includes about 50 known species, and the Cylindrocarpon clade including the teleomorph genera Ilyonectria, Neonectria, Rugonectria, and Thelonectria, and the anamorph genus Campylocarpon (see Chaverri et al. 2011), has at least 70 species. Volutella, discussed below, is probably similarly speciose, although no comprehensive revision exists. The hyphomycete genera Cylindrodendrum, Heliscus, and Mariannaea and many Acremonium-like species also occur in this clade.

In our analyses based on two genes including a standard barcode marker for Fusarium, rpb2, and a new phylogenetic marker, acl1, statistical support is weak for the backbone of the phylogenetic tree. Similar problems exist with published nuclear ribosomal large subunit trees, e.g. Summerbell & Schroers (2002), Zhang & Zhuang (2006), and Luo & Zhuang (2008). In the five gene analysis by Chaverri et al. (2011), the statistical support for the backbone of the Nectriaceae is stronger, and the few members sampled in the basal Fusarium-like clade and terminal Fusarium clade both form well-supported, distinct monophyletic groups. It would be preferable if the bootstrap and probability support for the relative arrangment of these clades were stronger, but in a polyphasic treatment, this is only one kind of evidence. Although molecular analyses do not strongly support our conclusion that the basal and terminal clades of Fusarium are phylogenetically distinct, there are also no data to support the taxonomic hypothesis that Fusarium sensu Wollenweber is monophyletic. Thus, neither monophyly nor the 1:1 teleomorph:anamorph genus argument supports the classical concept of Fusarium. We are confident that additional DNA sequencing data will add support to our conclusion that these major clades diverged long ago. Our decision results in a monophyletic concept of Fusarium s. str., although the terminal Fusarium clade retains some problematic groups that will require further consideration (cf. Schroers et al. 2011). Additional sampling of outlying Fusarium-like species will undoubtedly lead to the recognition of other genera.

The Hypocreales is an anamorph rich order, with the majority of holomorphic species having at least one anamorph, and with many apparently solely anamorphic species. One of the main character suites of the Nectriaceae are sporodochial anamorphs with slimy macroconidia produced from phialides, which are broadly distributed in the family and probably represent the plesiomorphic condition. The three best known macroconidial groups were placed in the classical genera Fusarium sensu Wollenweber, Cylindrocarpon, and Cylindrocladium, the latter now treated by its teleomorph generic name, Calonectria (Lombard et al. 2010). Often, macroconidial anamorphs are accompanied by microconidial, Acremonium-like synanamorphs, with small ameroconidia produced from phialides and enveloped in slime. These are probably also plesiomorphic in the family and homologous to similar “microconidial” anamorphs in other families of the order. In some lineages, macroconidia seem to have disappeared, while in other lineages, microconidia seem to have disappeared. Verticillate anamorphs occur in some clades, in particular Chaetopsina and Mariannaea, presumably derived from Acremonium-like progenitors. In addition to micro- and macroconidia, mesoconidia have been described in a few species of Fusarium (Pascoe 1990) as intermediate between micro- and macroconidia, but dry and produced from holoblastic conidiogenous cells, while megaconidia were described by Crous & Seifert (1998) in a few species of Calonectria, significantly larger than macroconidia and produced only under some cultural conditions.

Fusarium-like conidia occur in several orders of Ascomycota (Seifert 2001). In the Nectriaceae, the phylogenetic distribution of this character is disjunct. Because the phylogenetic backbone of the family is weakly supported in most analyses including ours, there are two possible interpretations for the distribution of the Fusarium-like conidium. If the Fusarium-like conidium is plesiomorphic in the Nectriaceae, then the cylindrical macroconidia of Calonectria and Neonectria were derived from it, and the taxa delimited by the ancestral Fusarium-like conidium have become paraphyletic. Alternatively, but perhaps less probable, the Fusarium-like conidium has evolved several times in the family, and the taxon delimited by this character is polyphyletic.

A practical problem with dividing Fusarium is the existence of 22 generic names sometimes considered synonyms (Table 2). These names must be considered in any division of the genus, which means that the identities of their type species in modern terms must be understood. Many of the synonyms come from the work of Wollenweber, whose herbarium studies are largely documented in his series Fusarium autographice delineata (Wollenweber 1916). Unfortunately, Wollenweber did not rigorously employ a type concept that conforms with today's standards, and we have discovered that many of his interpretations cannot be verified. The status of some of the 22 synonyms can be evaluated on the basis of existing knowledge and we examined type specimens of relevant genera for this study (Table 2); the precise status of a few of these genera remains uncertain. We focused on older generic synonyms, seriously considering Atractium (1809), Microcera (1848), Pionnotes (1849), and Fusicolla (1851).

Table 2.

Anamorphic genera reported as synonyms of Fusarium and interpretation of their type species according to present knowledge.

Generic nameType speciesSynonymy proposed byIdentity of type speciesPresent status
Fusisporium Link 1809 F. aurantiacum Link 1809 : Fr. Wollenweber (1916) F. graminum Corda or F. sporotrichioides Sherb. = Fusarium, Gams & Nirenberg 1989
Atractium Link 1809 A. stilbaster Link 1809 Wollenweber & Reinking (1935) A. stilbaster Link Distinct genus in Nectriaceae, this paper
Selenosporium Corda 1837 S. tubercularioides Corda 1837 ≡ Fusarium tubercularioides (Corda) Sacc. 1886 Lindau (1910), Wollenweber & Reinking (1935) F. avenaceum (Corda) Sacc. or F. lateritium Nees = Fusarium, Holubová-Jechová et al. 1994
Microcera Desm. 1848 M. coccophila Desm. 1848 Wollenweber & Reinking (1935) M. coccophila Desm. Distinct genus in Nectriaceae, this paper
Pionnotes Fr. 1849 P. capitata (Schw.) Fr. 1849 ≡ Fusarium capitatum Schw. 1832 Wollenweber & Reinking (1935) Dacrymyces sp. (PH!) = Dacrymyces, Seifert et al. in prep.
Fusicolla Bonord. 1851 F. betae (Desm. : Fr.) Bonord. 1851 ≡ Fusisporium betae Desm. 1830 : Fr. Wollenweber (1916), Wollenweber & Reinking (1935) Fusicolla betae (Desm.) Bonord. Distinct genus in Nectriaceae, this paper
Sporotrichella P. Karst. 1887 S. rosea P. Karst. 1887 Wollenweber & Reinking (1935) F. sporotrichioides Sherb. = Fusarium
Lachnidium Giard 1891 L. acridiorum Giard 1891 Saccardo (1901), Wollenweber & Reinking (1935) F. solani complex = Fusarium
Discocolla Prill. & Delacr. 1894 D. pirina Prill. & Delacr. 1894 Wollenweber & Reinking (1935) F. lactis Pirotta & Riboni = Fusarium
Septorella Allesch. 1897 S. salaciae Allesch. 1897 Höhnel (1912) Unknown Status uncertain
Trichofusarium Bubák 1906 T. rusci Bubák 1906 ≡ Fusarium roseum var. rusci Sacc. 1886 Wollenweber & Reinking (1935), Sutton (1986) Pycnofusarium rusci D. Hawksw. & Punith. Considered distinct by Schroers (pers. comm.)
Ustilaginoidella Essed 1911 U. musaeperda Essed 1911 Brandes (1919) F. oxysporum complex = Fusarium
Stagonostroma Died. 1914 S. dulcamarae (Pass.) Died. 1914 ≡ Stagonospora dulcamarae Pass. 1890 Sutton (1977) Unknown Status uncertain
Fusariopsis Horta 1919 F. derrienii Horta 1919 Dodge (1935) Unknown Unknown
DiscofusariumPetch 1921D. tasmaniense (McAlpine) Petch 1921Microcera tasmanica McAlpine 1904 ≡ Fusarium tasmanicum (McAlpine) Rossman 1983 Rossman (1983) “Fusarium” anamorph of “Nectriacoccidophaga (Petch) Rossman 1983 Unknown
PseudomicroceraPetch 1921P. henningsii (Koord.) Petch 1921Aschersonia henningsii Koord. 1907 Wollenweber & Reinking (1935) Microcera diploa = Microcera, this paper
Fusidomus Grove 1929 Not designated Sutton (1977) Unknown Status uncertain
Infrafungus Cif. 1951 I. micropus (Sacc.) Cif. 1951 ≡ Fusarium micropus Sacc. 1921 Wollenweber & Reinking (1935) Fusarium lateritium complex = Fusarium
Euricoa Bat. & H. Maia 1955 E. dominguesii Bat. & H. Maia 1955 Summerbell & Schroers (2002) F. solani complex
Hyaloflorea Bat. & H. Maia 1955 H. ramosa Bat. & H. Maia 1955 W. Gams (pers. comm.) F. solani complex = Fusarium
Pseudofusarium Matsush. 1971 P. fusarioideum Matsush. 1971 = Pseudofusarium semitectum (Berk. & Rav.) Matsush. 1975 Pascoe (1990) F. semitectum auct. Fusarium, Matsushima 1980
Pycnofusarium Punith. 1973 P. rusci D. Hawksw. & Punith. 1973 Sutton (1986) Pycnofusarium rusci D. Hawksw. & Punith. = Trichofusarium, Schroers (pers. comm.)

We considered two scenarios to resolve the para/polyphyly of Fusarium. The first was to adopt broad generic concepts and to maintain the two main lineages as genera, i.e. the terminal lineage including the type species of Fusarium, and the basal Fusarium-like lineage that includes most of the species attributed to Cosmospora sensu Rossman. The perithecial walls of the species of these two clades have clearly different micromorphology. Cultures generally differ in colony morphology and growth rates, produce different metabolites, and the species have different ecological preferences, especially host specificity. However, this separation was unsatisfactory because these two large clades themselves lacked convincing statistical support, and the amount of morphological diversity incorporated in both of these large clades was huge, rendering the resulting taxonomy meaningless from a practical point of view. In particular, the generic name Cosmospora would be supplanted by the oldest available name Microcera, resulting in a genus incorporating many large, phylogenetically well-supported clades, some of which are sufficiently well-defined ecologically and morphologically to be recognised as distinct genera on their own. In this broad concept of Microcera, anamorphs with Fusarium-like macroconidia would still not be monophyletic, because of the existence of a large clade of microconidial, Acremonium-like anamorphs that is terminal within this basal clade.

The second option was to adopt the genera as well-supported, ecologically or morphologically distinct clades within the basal lineage. Although this results in more genera, the concepts are more homogenous and the system is practical. We followed this second approach, and the details of the generic names adopted are included in the Taxonomy section below. Fortunately, we were able to assign existing generic or subgeneric taxa to most of the clades. Cosmospora is retained for the clade with Acremonium-like microconidial anamorphs, and Microcera is reintroduced in something similar to its nineteenth century delimitation, as a genus of insect pathogens producing striking, flame-like conidiomata, usually on scale insects. Despite the number of genera segregated, this revision keeps the core of common, economically important Fusarium species intact. Of the species included in the popular Nelson et al. (1983) system and its more speciose successor (Leslie et al. 2006), only the F. aquaeductuum and F. merismoides species complexes are removed to Fusicolla. The more difficult decision concerning the generic fate of the Fusarium solani species complex remains to be decided.

Both Gams & Nirenberg (1989) and Seifert (2001) emphasised the importance of delimiting genera using polythetic concepts, i.e. concepts based on the occurrence of variable sets of shared characters with no single character considered essential for inclusion. Although we provide preliminary descriptions below, the development of robust polythetic diagnoses for the genera remains a work in progress. This is just the beginning of a taxonomic reevaluation of Fusarium and morphologically similar genera that, with increased sampling and more genomic analysis, will result in the recognition and definition of additional segregate genera. This revision provides a foundation for the discovery and phylogenetic classification of a large amount of presently unrecognised diversity representing both holomorphic and anamorphic species.

It is unfortunate that our decision to attempt to implement a single name nomenclature to these fungi coincides with what may be equally a controversial decision to split Fusarium. In general, Fusarium workers have had little interest in teleomorphs and most will have no reluctance to abandon a dual nomenclature of little relevance to them. Because teleomorphs are rarely seen in culture, except for that of F. graminearum, they are considered the domain of taxonomic specialists and their nomenclatural primacy is an historical annoyance. The introduction of single scientific names for polythetically characterised holomorphs and the recognition of a single nomenclaturally valid name for all taxonomic ranks seem inevitable steps towards the stabilisation of fungal taxonomy (Rossman & Samuels 2005). We encourage mycologists to accept our proposed nomenclature as a sincere attempt to provide a functional single-name system that respects the principles of the ICBN and refrain from attempting to perpetuate a dual nomenclatural system where it is unlikely to be used by most scientists working on the practical aspects of these fungi.

TAXONOMY

In this section, we consider the classification, nomenclature, and typification of the species examined in our phylogenetic studies and implement the taxonomic conclusions discussed above. Where possible, we have examined holotype specimens, other authentic material, and/or ex-type cultures, as well as material conforming to the concepts of Wollenweber. When feasible, we designate lectotype or epitype specimens to stabilise species concepts and provide living material for further studies. Many species are pleomorphic having a teleomorph, a macroconidial, Fusarium-like anamorph, and a microconidial or Acremonium-like anamorph, or any combination of these. The morphs recorded for each species are indicated on Fig. 1. The species are not redescribed here. In some cases, species concepts applied by various authors deviate from the strict concept of the species as typified. Therefore, we refer only to descriptions and illustrations already published that represent the species indicated by the typification.

Atractium Link: Fr., Mag. Ges. naturf. Freunde, Berlin 3: 10 (tab. I, fig. 11), 1809: Fries, Syst. Mycol. 1: xli, 1821.

Type species: Atractium stilbaster Link 1809.

Emended generic diagnosis

Teleomorph unknown. Conidiophores aggregated into sporodochia or synnemata, nonstromatic; in culture, sometimes becoming pionnotal. When produced synnemata determinate, pale brown, composed of a stipe of parallel hyphae and a divergent capitulum of conidiophores giving rise to a slimy conidial mass; differentiated marginal hyphae absent. Conidiophore branching once or twice monochasial, 2-level verticillate, monoverticillate or irregularly biverticillate. Conidiogenous cells monophialidic, hyaline, subulate, with conspicuous periclinal thickening. Conidial masses yellow to orange. Conidia (0–)1–5-septate, clavate, obovoid or gently curved, rarely ellipsoidal, with a rounded apical cell, and somewhat conical basal cell, lacking a differentiated foot. Chlamydospores produced in culture by some species. Cultures growing relatively slowly, usually less than 30 mm diam in 14 d, with little aerial mycelium.

One of the commonly cited synonyms of Fusarium is the name Atractium, described immediately following and on the same page as its more famous cousin. The original diagnosis for Atractium and its type species A. stilbaster reads:

Atractium. Stroma elongatum, capitatum. Sporidia fusiformia, non septata, capitulo instrata. Stroma stilbiforme, sporidia eadem quae Fusidiorum. Contextus stromatis, uti videtur, tenue floccosus, floccis parallelis. Capitulum sub microscopio composito, aqua adfusa, in sporidia fere diffluit. Unica species, nondum descripta. A. stilbaster, stipite cylindrico, capitulo globoso, utroque glabro lutescente. In truncis fagorum caesorum occurrit, vix ulta ½ lin. longa, fugax, stipite facili evanescente et capitulo in sporidia diffluente. Rarius invenit am. Ditmar. Iconem v. fig. 11.”

The protologue includes a drawing by Ditmar (reproduced here as Fig. 2B), which shows what could either be a capitate, synnematous fungus, similar to Stilbella or possibly a myxomycete with a ruptured sporangium as seen in species of Trichia and many other genera, growing on a stump of Fagus. Link was confused about the septation of conidia of A. stilbaster. The protologues for both Fusarium and Atractium explicitly state, “Sporidia fusiformia, nonseptata...”. Link (1816) added two more species to Atractium that Nees (1817) transferred to Fusarium without explanation. Link (1825) adjusted his observation and reported septate conidia in A. stilbaster, transferring it to Fusarium, and implicitely modifying his original species concept, and thus the generic concept of Atractium, to include species with septate conidia. These reinterpretations led subsequent authors, such as Berkeley, Fuckel, and Saccardo, whose systematic philosophy would not allow synnematous species to be included in the sporodochial genus Fusarium, to place synnematous Fusarium-like species in Atractium. In the 19th century, the prevailing concept of Atractium evolved to represent pale or colourful synnematous fungi with slimy conidial masses, usually with falcate, septate conidia. Tulasne & Tulasne (1861, 1865) noted the similarity of Atractium and Microcera (reintroduced below), and Petch (1921) commented on the modification of Link's original concept to include species with septate conidia. The species added to Atractium were often associated with the teleomorph genus Sphaerostilbe, the species of which were revised by Seifert (1985a).

An external file that holds a picture, illustration, etc.
Object name is 79fig2.jpg

Atractium species. A. Atractium crassum, as illustrated in the protologue by Wollenweber (1930). B. Atractium stilbaster, original drawing by Ditmar accompanying the protologue of Atractium, designated here as lectotype for A. stilbaster. C. Atractium crassum, digital photographs of living conidiomata (left) and a conidium (right) from a collection made in Ontario, Canada (K.A.S. 809).

Following the work of Wollenweber & Reinking (1935), who equated A. stilbaster with Fusarium aquaeductuum var. medium (now Dialonectria ullevolea, see below), Atractium was usually listed as a synonym of Fusarium. The proposed synonymy is curious because this species does not produce synnemata, the dominant feature of Link's drawing of A. stilbaster. There is no reason to follow Wollenweber & Reinking's interpretation and no evidence that Wollenweber, in his work for either Die Fusarien or Fusarium autographice deliniata, saw authentic material of A. stilbaster.

We were unable to locate authentic material of A. stilbaster, the original species of Atractium, from the herbaria of Link (B), Persoon (L) or Fries (UPS, UPS-Fries). The drawing with the protologue must be regarded as the lectotype; it shows what we interpret as a capitate, synnematous fungus (Fig. 2). The confusion over whether or not the conidia were septate, described above, is instructive in the interpretation of the identity of this fungus. To fix the application of the name, an epitype specimen should be designated of a synnematous fungus occurring on wood of Fagus in Germany. Seifert (1985a) provided a description and illustration of a fungus he called Stilbella fusca, a common, synnematous fungus on water-saturated, decayed wood, including trunks of Fagus, in northern Europe including Germany. It is the most frequently collected species attributed to the pre-1985 concept of Didymostilbe, and was often reported as D. eichleriana. This species produces slimy, obovate to obclavate conidia that are usually curved, from long phialides on branched conidiophores. The present concept includes specimens with predominantly aseptate conidia, but most specimens have only 1-septate conidia (Seifert 1985a). This species thus matches both Link's original concept and his subsequent revised concept of A. stilbaster in all salient details, especially noting that other authors included it in Atractium. A culture of this fungus isolated from bark in Germany, CBS 410.67, is thus selected as the epitype for A. stilbaster, applying Atractium for this clade identified in Fig. 1.

The three species of Atractium accepted here are all associated with water in some way. Atractium stilbaster and A. holubovae (not known in culture) are associated with water saturated decaying wood, and A. crassum was isolated twice from drinking water in Germany.

In our phylogenetic analysis (Fig. 1), two species (A. stilbaster and A. crassum) form a well-supported monophyletic clade in the Nectriaceae. The clade is also basal to Chaetospina, Pseudonectria, and Volutella as discussed below.

We did not attempt a systematic reevaluation of the 24 species attributed to Atractium, but a summary of present knowledge is presented in Table 3.

Table 3.

Species attributed to Atractium and their current status. Basic nomenclatural data from Index Fungorum (www.indexfungorum.org).

Species, authority and year of publicationStatusReference
A. aurantiacum (Corda) Bonord. 1851 Unknown
A. brunaudiana Sacc. 1883 Unknown
A. candiduli Sacc. 1883 = Cylindrocarpon candidulum (Sacc.) Wollenw.
A. ciliatumLink 1816 Basionym of “Fusarium” ciliatum (Link) Link This paper
A. cristatum Demelius 1923 Unknown
A. cronartioides Speg. 1883 Unknown
A. flammeolumHöhn. 1915Nomen dubiumSeifert 1985a
A. flammeum Berk. & Ravenel 1854 = Microcera coccophila Desm. This paper
A. flavoviride Sacc. 1883 Synonym of A. stilbasterSeifert 1985a
A. fuscum Sacc. 1883 Synonym of A. stilbasterSeifert 1985a
A. gelatinosum (Pers.) Sacc. 1886 No type in L, nomen dubiumSeifert 1985a
A. indicum Chona & Munjal 1956 Unknown
A. lusitanicum Sousa da Câmara & Luz 1941 Unknown
A. micropus (Pers.) Sacc. 1886 No type in L, nomen dubiumSeifert 1985a
A. olivaceum Kunze & J.C. Schmidt 1817 No type in B, nomen dubiumSeifert 1985a
A. pallens Nees 1818 Type in B examined, is a coelomycete This paper
A. pallidum Bonord. 1851 Unknown
A. pallidum Berk. & M.A. Curtis 1868 Unknown
A. pulvinatumLink 1816 Type in B examined, not an Atractium This paper
A. rigidum Bonord. 1864 Unknown
A. stilbaster Link 1809 Accepted species This paper
A. therryanum Sacc. 1879 Anamorph of Dermea morthieri (Fuckel) Nannf. Groves 1946
A. trematis Hansf. 1944 Unknown
A. tubericola Sacc. & Peglion. 1902 Unknown

No teleomorphs are conclusively known for this genus, and there are no other published names that could be applied to this clade. Seifert (1985a) discussed the association of A. stilbaster with “Nectria” flavoviridis and Sphaerostilbe fusca, concluding that the reported association of this teleomorph and anamorph was probably coincidental. Our reexamination of the type material suggests that the KOH–perithecia on the specimen are more likely to represent the teleomorph of a species of Fusicolla, the macroconidia of which also occur on the specimen, rather than the teleomorph of A. stilbaster.

KEY TO ACCEPTED ATRACTIUM SPECIES

  • 1. Conidia mostly (0–)1–3 septate; synnematous conidiomata produced..................................................................................................... 2
    1. Conidia mostly 3(–5) septate; synnemata not produced........................................................................................................... A. crassum

  • 2. Conidia 37–49 × 4–5.5 μm; phialides 30–54 × 1.5–2.5 μm.................................................................................................... A. holubovae
    2. Conidia 15–25 × 2–4.5 μm, phialides 20–40 × 1.5–2.5 μm...................................................................................................... A. stilbaster

Accepted species

Atractium stilbaster Link 1809, Mag. Ges. naturf. Freunde, Berlin 3: 10.
Basionym: Fusarium stilbaster (Link) Link in Willdenow, Sp. pl., Edn 4 6(2): 106. 1825 (1824).

  • = Atractium fuscum Sacc., Syll. Fung. 2: 514. 1883.

  • ≡ Stilbella fusca (Sacc.) Seifert, Stud. Mycol. 35: 77. 1985.

See Seifert (1985a, as Stilbella fusca) for other synonyms.

Typification: Illustration published in Mag. Ges. naturf. Freunde, Berlin 3 as tab. I, fig. 11, lectotype designated here, reproduced here as Fig. 2B. epitype of A. stilbaster designated here: Germany, Bayrischer Wald, Rachelseewand, on bark, Jul. 1967, W. Gams, CBS 410.67.

Other material examined: See Seifert (1985a). Canada, Quebec, Gatineau Park, Lac Bourgeous, on cut end of stump, Jul. 1992, K.A. Seifert, DAOM 215627.

Notes: Seifert (1985a) provided illustrations and a complete description of this species. The variability in conidium dimensions and septation reported by Seifert (1985a) may indicate the existence of several closely related but possibly morphologically diagnosable species.

Atractium crassum (Wollenw.) Seifert & Gräfenhan, comb. & stat. nov. MycoBank MB519420.
Basionym: Fusarium merismoides var. crassum Wollenw., Fus. autogr. del. 3: 857. 1930. (The publication of the same species in Zeitschrift für Parasitenkunde 3(3): 308. 1931 was apparently after the cited 1930 publication).

Typification: Germany, Berlin, isolated from drinking water, 1928, H.W. Wollenweber 3119, lectotype designated here, CBS. ex-type cultures CBS 180.31 = NRRL 20894. GenBank barcodes: HQ897722 (rpb2), HQ897859 (acl1).

Notes: This species was described and illustrated by Wollenweber (1930, reproduced here as Fig. 2A), Wollenweber & Reinking (1935), and Gerlach & Nirenberg (1982). The strains described by the latter authors are now degenerated, and the following details come from their description. Fresh cultures grow slowly, 15–30 mm diam after 10 d on PDA, and sometimes produce Coremium-like structures. The macroconidia are gently curved with a rounded to somewhat conical basal cell and a rounded apical cell; there is no foot to the basal cell. They are mostly 3–5-septate; 3-septate conidia average 52 × 5 μm (ranging 37–60 × 4.5–5.5), 4–5-septate 60 × 5.5 μm (50–65 × 5-6), 1–2 septate 31 × 4.5 μm (25–37 × 3–6). Chlamydospores are terminal, intercalary or in conidia, round, 7–12 μm diam.

A second culture, BBA 62257, was illustrated by Gerlach (1972) and Gerlach & Nirenberg (1982) but is no longer available. A dried culture kept in the CBS herbarium is designated as lectotype above, because it is the only known original material. Wollenweber's published illustration of the type strain (Fig. 2A) represents the macroconidia of his taxon well. Epitypification must await the isolation of a fresh culture and specimen that can demonstrate the salient morphological features more completely than the existing cultures.

This species developed in damp chambers on small twigs collected from cold, running river water in Ontario, Canada, but the cultures were not preserved and the fungus cannot be relocated on the original specimen. Attempts to recollect and reisolate the fungus from the same locality were unsuccessful. The conidiomata on the natural substrate were glistening white and flame-shaped; the bundles of parallel macroconidia give the appearance of minute synnemata (Fig. 2C). However, little conidiomatal tissue is actually produced, and the phialides arise from a typical, Fusarium-like sporodochium of interwoven but not stromatic hyphae and conidiogenous cells.

Atractium holubovae (Seifert, S.J. Stanley & K.D. Hyde) Seifert, comb. nov. MycoBank MB519421.
Basionym: Stilbella holubovae Seifert, S.J. Stanley & K.D. Hyde, Sydowia 47: 258, 1995.

Typification: Philippines, Negros Occiental, Bario Caliban, Caliban River, on submerged wood, Dec. 1994, K.D. Hyde & E. Arimas, holotype DAOM 214961.

Notes: This species was described and illustrated by Seifert et al. (1995) in the absence of pure cultures and is transferred here on the basis of its morphological similarity with A. stilbaster. It is known from the holotype and two subsequent records on submerged wood collected from streams in Asia (Sivichai et al. 2002, Fryar et al. 2004).

Cosmospora Rabenh., Hedwigia 2: 59. 1862.

Type species: Cosmospora coccinea Rabenh. 1862.

Stroma inconspicous or absent. Perithecia scattered to gregarious, pyriform with an acute or apical papilla, collapsing cupulate or pinched when dry, orange red or bright red, turning dark red in KOH+, smooth walled, usually 150–450 μm high. Asci cylindrical to narrowly clavate, with an apical ring, 8 uniseriate or partly biseriate ascospores. Ascospores initially hyaline but becoming yellow brown to reddish brown, 1-septate, becoming tuberculate when mature. Conidiophores Acremonium-like, either lateral phialides on somatic hyphae, or with one or two layers of monochasial branching, or verticillate, hyaline. Phialides monophialidic, cylindrical to subulate, hyaline. Microconidia ellipsoidal, oblong or clavate or slightly allantoid, aseptate, hyaline, in slimy heads. Macroconidia absent. Chlamydospores usually not seen, but produced on some media.

Colonies on PDA slow growing, 15–25 mm diam in 14 d at room temperature, surface powdery, felt-like, floccose, cottony, white, pale pink, ochre to olivaceous green, sporulation usually abundant, arising directly from agar surface or from sometimes abundant aerial mycelium.

Habitat: On fruiting bodies and stromata of other fungi, e.g. Fomitopsis, Hypoxylon, Inonotus, Stereum, often isolated from soil.

Notes: About 65 species have been attributed to Cosmospora sensu Rossman. This concept is relatively broad, encompassing a great deal of anamorphic variability, although the teleomorph morphology is relatively conserved, with small, orange or reddish KOH+ perithecia with thin walls, cylindrical asci with or without an apical ring, and eight, uniseriate, 1-septate ascospores; stroma development is usually limited. Our phylogenetic analyses (Fig. 1) identify several distinct lineages within the prevailing concept of Cosmospora. New teleomorph genera have already been proposed for some lineages, namely Nectricladiella (a synonym of the anamorphically typified genus Cylindrocladiella) and Chaetopsinectria (a synonym of the anamorphically typified genus Chaetopsina). In general, well-supported clades correlate with anamorph types, although Fusarium-like anamorphs are found in several lineages.

Here, we propose a more restricted concept for Cosmospora, limiting it to the clade of species surrounding the type, C. coccinea, which have only microconidial, Acremonium-like anamorphs and tend to occur on other fungi. Other microconidial genera recognised are Mariannaea and Volutella. The clades with Fusarium-like anamorphs are reclassified below in the reintroduced genera Dialonectria, Fusicolla, and Microcera, with Macroconia elevated to generic rank from its previous sectional rank in Nectria. A small residue of species remains in Cosmospora sensu Rossman that are not redisposed here.

Although several of the new combinations propose the transfer of an anamorph typified name to a teleomorphically typified genus, as explained in the Introduction, the results are correct, legitimate, and valid for those species that are not pleomorphic, i.e. those that lack a teleomorph and are outside Art. 59 of the ICBN.

Accepted species

Cosmospora coccinea Rabenh., Hedwigia 2: 59. 1862 [non Nectria coccinea (Pers.) Fr. 1849].

  • = Verticillium olivaceum W. Gams, Cephalosporium-artige Schimmelpilze, p. 129. 1971.

Typification: Germany, near Laubach, on rotting pores of a polypore, Solms, Fungi europaei no. 459, lectotype BPI designated by Rossman et al. 1999.

Other material examined: Germany, Bayrischer Wald, Arberseewand, on hymenium of Inonotus nodulosus on Fagus sylvatica, Aug. 1967, W. Gams 680, CBS 341.70 = VKM F-2863; Kr. Plön, near Dobersdorf, on hymenium of Inonotus radiatus on Alnus, Oct. 1965, W. Gams 1104, CBS 343.70; Eifel, Geeser Wald near Gerolstein, on Inonotus radiatus, Sep. 1970, W. Gams, CBS 841.70; Eifel, Geeser Wald near Gerolstein, on Inonotus radiatus, Sep. 1970, W. Gams, CBS 983.70 = VKM F-2862; Neubrandenburg, Kleppelshager Forst near Friedland, on Inonotus radiatus, Oct. 1978, P. Hübsch H78/40, CBS 704.79; Bayern, on Inonotus nodulosus, dead crust, on fallen branch of Fagus sylvatica, 1993, T.R. Lohmeyer & R. Boesmiller 93/62, A.R. 2741 = BPI 802729 = CBS 114050; Nordrhein-Westfalen, Detmold, Krebsteich, on Inonotus nodulosus on Fagus sylvatica, Apr. 2007, T. Gräfenhan 2007-37, DAOM 235821.

Notes: For descriptions, illustrations, and additional taxonomic synonyms of the microconidial anamorph, see Gams (1971); the teleomorph is briefly described by Rossman et al. (1999).

Cosmospora arxii (W. Gams) Gräfenhan & Schroers, comb. nov. MycoBank MB519422.
Basionym: Acremonium arxii W. Gams, Cephalosporium-artige Schimmelpilze, p. 123. 1971.

Typification: Germany, Niedersachsen, near Wilhelmshaven, Neuenburger Urwald, on Hypoxylon sp., May 1965, W. Gams, holotype CBS H-6635, ex-type culture CBS 748.69 GenBank barcodes: HQ897725 (rpb2), HQ897862 (acl1).

Other material examined: Germany, Nordrhein-Westfalen, Kamen, Heerener Holz, on Hypoxylon on Fagus, Apr. 2007, T. Gräfenhan 2007-22, DAOM 235822; Nordrhein-Westfalen, Detmold, Externsteine, on Hypoxylon on Fagus sylvatica, Apr. 2007, T. Gräfenhan 2007-28, DAOM 235823; Nordrhein-Westfalen, Detmold, Donoper Teich, on Hypoxylon on Fagus sylvatica, Apr. 2007, T. Gräfenhan 2007-29, DAOM 235824 & T.G. 2007-33, DAOM 235825; USA, Pennsylvania, near Salt Springs State Park, on Hypoxylon on Acer, May 2007, T. Gräfenhan 2007-55, DAOM 235826.

Notes: The teleomorph of Cosmospora arxii is commonly found on Hypoxylon spp. on Fagus in North America and Europe, but has not been described yet; its morphology is similar to that of C. viridescens. For a description, illustrations, and discussion of the microconidial anamorph, see Gams (1971) and notes under C. berkeleyana below.

Cosmospora berkeleyana (P. Karst.) Gräfenhan, Seifert & Schroers, comb. nov. MycoBank MB519423.
Basionym: Verticillium berkeleyanum P. Karst., Meddeland. Soc. Fauna Fl. Fenn. 18: 64. 1891.

  • Acremonium berkeleyanum (P. Karst.) W. Gams, Netherlands J. Pl. Pathol. 88: 76. 1982.

Typification: Finland, near Mustiala, on Stereum hirsutum on Betula, Oct. 1890, P.A. Karsten 2310, holotype H.

Notes: For a description and discussion of this microconidial species, see Karsten (1891) and Gams & Zaayen (1982).

Although some have considered the teleomorph to be the heterotypic Hypomyces berkeleyanus Plowr. & Cooke (≡ Sphaerostilbella berkeleyana (Plowr. & Cooke) Samuels & Candoussau), our observations complicate the situation considerably. Because our phylogenetic results suggest that this is a species complex, the proposed synonyms applied to the teleomorph-anamorph connections for Cosmospora berkeleyana need to be re-evaluated (Fig. 1). These synonyms include Acremonium butyri, Cephalosporium khandalense, Gliomastix lavitskiae, Nectria vilior, and N. viridescens (Gams 1971, Samuels et al. 1990, Rossman et al. 1999). In our phylogenetic analysis, all of these putative synonyms can be interpreted as distinct species of Cosmospora.

Cosmospora berkeleyana, C. vilior, and C. viridescens have often been considered synonymous, but this now seems unlikely and each name must be re-evaluated. Samuels et al. (1990, 1991) studied and discussed the type material of C. vilior on a valsaceous stroma from Brazil. Because no fresh material from subtropical South America is available, we are unable to reinterpret Samuels' concept in phylogenetic terms. Cosmospora viridescens was described from a fungal host on Salix in Europe and thus may have distinct host relationships and geographical distribution. Possible morphological distinctions between these two teleomorphs are discussed below under C. viridescens.

Gams & Zaayen (1982) studied a recent specimen and culture identified as Acremonium berkeleyanum, which was unavailable for our study (The Netherlands, Oostelijk Flevoland, Abbert-bos, perceel O66, on Stereum hirsutum, July 1981, W. Gams, CBS 501.81). A similar fungus producing perithecia and the characteristic greenish Acremonium-like anamorph on basidiocarps of S. hirsutum on Alnus rubra is common in British Columbia, Canada (Seifert, unpubl. data).

Until species limits can be more clearly established, we prefer not to epitypify C. berkeleyana or C. vilior. The diversity of substrates and broad geographic distribution recorded for C. berkeleyana (Gams 1971, www.cbs.knaw.nl/databases) suggest that additional phylogenetic species await discovery in this complex.

Cosmospora butyri (J.F.H. Beyma) Gräfenhan, Seifert & Schroers, comb. nov. MycoBank MB519428.
Basionym: Tilachlidium butyri J.F.H. Beyma, Zentralbl. Bakteriol., 2 Abt. 99: 388. 1938.

  • Acremonium butyri (J.F.H. Beyma) W. Gams, Cephalosporium-artige Schimmelpilze, p. 126. 1971.

Typification: Denmark, Copenhagen, butter, Knudsen, holotype CBS H-6601, ex-type cultures CBS 301.38 = MUCL 9950. GenBank barcodes: HQ897729 (rpb2), HQ897866 (acl1).

Notes: No teleomorph is known, but see notes under C. berkeleyana above. This microconidial species is described, illustrated, and discussed by van Beyma (1938) and Gams (1971). As noted by Summerbell et al. (2011), there may be more than one fungus preserved as CBS 301.38; we have not examined the holotype specimen.

Cosmospora cymosa (W. Gams) Gräfenhan & Seifert, comb. nov. MycoBank MB519429.
Basionym: Acremonium cymosum W. Gams, Cephalosporium-artige Schimmelpilze, p. 131. 1971.

Typification: Germany, Schleswig-Holstein, Kr.Rendsburg, Enkendorfer Gehölz, on decaying Inonotus radiatus, Oct. 1965, W. Gams, lectotype designated here CBS H-5054, isotype CBS H-6603, ex-type culture CBS 762.69. GenBank barcodes: HQ897778 (rpb2), HQ897914 (acl1).

Other material examined: Germany, Kr.Plön, Dobersdorfer Wald, on Inonotus radiatus on Alnus glutinosa, June 1965, W. Gams 512A, CBS H-8146, CBS 258.70.

Notes: For description and illustrations of this microconidial anamorphic species, see Gams (1971). No teleomorph is known.

Cosmospora khandalensis (Thirum. & Sukapure) Gräfenhan & Seifert, comb. nov. MycoBank MB519430.
Basionym: Cephalosporium khandalense Thirum. & Sukapure, Mycologia 58: 359. 1966.

Typification: India, Maharashtra, Khandala, on decaying stem of Bambusa, Aug. 1964, M.J. Thirumalachar, holotype HACC 148, isotype CBS H-15076, ex-type cultures ATCC 16091 = CBS 356.65 = IMI 112790 = MUCL 7974. GenBank barcodes: HQ897723 (rpb2), HQ897860 (acl1).

Notes: The microconidial anamorph of this species as typified here is described and illustrated by Sukapure & Thirumalachar (1966) and discussed by Gams (1971). See notes above under C. berkeleyana.

Cosmospora lavitskiae (Zhdanova) Gräfenhan & Seifert, comb. nov. MycoBank MB519431.
Basionym: Gliomastix lavitskiae Zhdanova, Mikrobiol. Zhurn. 28: 37. 1966.

Typification: Ukraine, Poltawa region, on plant debris from rhizosphere soil of Zea mays, July 1961, holotype D.K. Zabolotny Institute of Microbiology and Virology of the National Academy of Sciences of Ukraine, ex-type cultures ATCC 18666 = CBS 530.68 = IMI 133984 = VKM F-1324. GenBank barcodes: HQ897726 (rpb2), HQ897863 (acl1).

Notes: The microconidial anamorph of the species is described and illustrated by Zhdanova (1966) and discussed by Gams (1971). No teleomorph is known. See notes above under C. berkeleyana.

Cosmospora viridescens (C. Booth) Gräfenhan & Seifert, comb. nov. MycoBank MB519432.
Basionym: Nectria viridescens C. Booth, Mycol. Papers 73: 89. 1959.

Typification: UK, England, Yorkshire, Sawley Woods, on black pyrenomycete on branches of Salix, Apr. 1954, C. Booth, holotype IMI 56736, isotype DAOM 83074.

Notes: The microconidial anamorph and teleomorph of this species as typified are described, illustrated, and discussed by Booth (1959) and Gams (1971).

Cosmospora viridescens is morphologically similar to C. vilior, but the latter has tuberculate ascospores, compared to the spinulose ascospores of C. viridescens (Samuels et al. 1990). Both species have Acremonium-like anamorphs with green colonies, and their perithecia occur on black, valsaceous stromata. Ascospore isolates made from perithecia collected on stromata of Hypoxylon and Ustulina in temperate areas often yield green colonies similar to C. viridescens, but are probably different from the tropical or subtropical species identified as C. vilior. Furthermore, differences in substrate specificity and geographic distribution support the distinction of C. viridescens from the other Cosmospora species mentioned above.

Cosmospora viridescens cannot be correlated with any described Acremonium species, nor can any of the described Acremonium species in this complex be unequivocally connected to any of the described teleomorphic species. Of the species in this complex with names based on anamorphic types, only C. arxii unequivocally has a known teleomorph, but it has apparently never been named.

Dialonectria (Sacc.) Cooke, Grevillea 12: 109. 1884. MycoBank MB1491.

Type species: Dialonectria episphaeria (Tode: Fr.) Cooke 1884 as D. sanguinea.

Stroma inconspicous or absent. Perithecia scattered and solitary or in small groups, pyriform with a short acute or round apical papilla, collapsing cupulate or pinched when dry, orange red to carmine red, turning dark red in KOH+, smooth-walled, usually < 200 μm high. Asci cylindrical to narrowly clavate, with an apical ring, 8 uniseriate ascospores. Ascospores hyaline to pale brown, 1-septate, smooth or becoming tuberculate when mature. Conidiophores initially as lateral phialides on somatic hyphae, sometimes verticillate, hyaline. Phialides monophialidic, subulate to subclavate, hyaline. Microconidia ellipsoidal to clavate, aseptate, hyaline, abundant. Macroconidia, if present, subcylindrical, moderately curved, slightly narrowing toward each end, apical cell often slightly hooked with a more or less pointed tip, basal cell not or scarcely pedicellate, predominantly 3–5-septate, hyaline, mostly thin-walled. Chlamydospores not observed.

Colonies on PDA slow growing, 25–50 mm diam in 14 d at room temperature, surface smooth, white to orange, aerial mycelium sparse, often becoming pionnotal, i.e. with abundant sporulation occuring in slimy masses over colony surface, often without discrete sporodochia.

Habitat: Mostly growing on stromata of other ascomycetes on deciduous trees.

Notes: Dialonectria was introduced first as a subgenus of Nectria and was revised in that context by Samuels et al. (1991), with a delimitation that more or less correlated with what the same authors later assigned to Cosmospora sensu Rossman. With the more restricted delimitation of Cosmospora adopted above, we also propose a restricted concept of Dialonectria around its type species, D. episphaeria. Most of the ~45 other species ascribed to Dialonectria by various authors have been reassigned or synonymised with other species by students of Nectria over the past 30 years.

Several phylogenetically distinct lineages are known within the D. episphaeria complex, one of which is described as a new species below.

Accepted species

Dialonectria episphaeria (Tode: Fr.) Cooke as D. sanguinea, Grevillea 12: 110. 1884.
Basionym: Sphaeria episphaeria Tode: Fr., Tode, Fungi Mecklenb. Sel. 2: 21. 1791: Fries, Syst. Mycol. 2: 454. 1823.

  • Nectria episphaeria (Tode: Fr.) Fr., Summa Veg. Scand. 2: 388. 1846.

  • Cucurbitaria episphaeria (Tode: Fr.) O. Kuntze, Rev. Gen. Plant. 3: 461. 1898.

  • Fusarium episphaeria (Tode) W.C. Snyder & H.N. Hansen, Amer. J. Bot. 32: 662. 1945.

  • Cosmospora episphaeria (Tode: Fr.) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 121. 1999.

Typification: Origin unknown, lectotype designated by Booth (1959) in L 0112704, Herb. Lugd. Bat. 910267659 ex Herb. Persoon, isotype TNS.

Notes: For description, illustrations, and discussion of the teleomorph, see Booth (1959). The anamorph produces micro- and macroconidia and is described by Gerlach & Nirenberg (1982) and Nelson et al. (1983).

The morphological species Dialonectria episphaeria splits into at least five phylogenetic lineages, which share similar phenotypic traits (Gräfenhan et al. 2008). There is presently no fresh, well-characterised material on Diatrype on Crataegus from northern Germany suitable for epitypification. The anamorph of D. episphaeria was often reported as or referred to as Fusarium aquaeductuum var. medium, e.g. Gerlach & Nirenberg 1982, but we consider this to represent a different phylogenetic species that is described below as a new species.

Dialonectria ullevolea Seifert & Gräfenhan, sp. nov. MycoBank MB519433. Fig. 3A–J.

An external file that holds a picture, illustration, etc.
Object name is 79fig3.jpg

A–J. Dialonectria ullevolea, ex-type strain (BBA 64549). A. Micro- and macroconidia formed on CMA after 18 d. B. Colony surface on PDA after 1 mo. C. Microconidia formed on CMA after 18 d. D–E. Phialides bearing microconidia on agar surface (D) and submerged (E) on CMA after 14 d. F. Cylindrical asci with obliquely uniseriate ascospores. G. Pyriform perithecia in culture on CMA after 50 d. H. Cells at surface of perithecial wall mounted in water. I. Perithecial apex mounted in water. J. Ascospores in optical section mounted in water. Scale bars: C, D, E, J = 10 μm; A, F, H, I = 20 μm; G = 200 μm.

  • = Fusarium aquaeductuum var. medium Wollenw., Fus. autogr. del., no. 844. 1930.

Etymology: K.A.S. recalls impassioned discussion on the topic of dividing Fusarium with P. Crous, K. O'Donnell, M. Stadler, and B. Summerell during the 7th International Mycological Congress in Oslo, Norway, August 2002; this is commemorated with Dialonectria ullevolea, named for the Ullevol pub, where this discussion occurred.

Coloniae in agaro CMA perithecia fertilia, aurantiaco-rubra vel rubra formantes; perithecia pyriformia, papilla brevi praedita, Dialonectriae episphaeriae similia, ascosporis dilute brunneis, bicellularibus, (8.7–)9.7–11(–12.5) × (3.7–)4–4.5(–4.8) μm. Conidiophora primum phialides simplices ex hyphis orientes, deinde irregulariter ramosa, nonnumquam verticillata. Monophialides subulatae vel subclavatae, 8–20 × 1.5–2.3 μm. Conidia copiosa in pionnote conidiophororum aggregatorum vel in conidiophoris singulis, tenuitunicata, hyalina: microconidia ellipsoidea vel clavata, unicellularia, (3–)3.5–5(–6.5) × 1–1.5(–1.7) μm, fere copiosa; macroconidia plerumque 3–5-septata, 1-septata: 10–25 × 1.5–2 μm, 3-septata: (20–)30–42(–48) × (1.8-)2–2.5(–2.7) μm, 4–5-septata: (30–)37–43.5(–50) × (1.8-)2–2.5(–2.7) μm, 6–7-septata: 40–48(–52) × (2–)2.3–2.7 μm, subcylindrica, modice curvata, utrinque paulo angustata, sursum saepe paulo uncinata et plus minusve acutata; ad basim vix an non pedicellata. Coloniae in agaro PDA lente crescentes, 25–30 mm diam. post 14 dies, dilute aurantiae vel griseo-aurantiae. Mycelium aerium absens vel appressum, pionnotes aurantia iuxta coloniam mediam. Corpora sclerotialia absentia.

On CMA, the type culture forms fertile, orange red to bright red perithecia, pyriform each with a short apical papilla, morphologically similar to Dialonectria episphaeria as described by Booth (1959); ascospores pale brown, 1-septate, (8.7–)9.7–11(–12.5) × (3.7–) 4–4.5(–4.8) μm (n = 50).

Colonies slow-growing on PDA, 25–30 mm diam in 14 d at room temperature. Surface light orange (5A5) to greyish orange (5B5) in colony centre, whitish at margin, margin smooth to broadly lobed. Reverse similar in colour but less bright with a slightly yellowish tinge (6A4 to 6B5). Aerial mycelium sparse or occasionally with floccose spots, lacking or appressed at margin. Sporulation in orange pionnotal masses, first observed near colony centre. Sclerotial bodies not observed.

In culture on CMA: Conidiophores initially unbranched, with phialides arising laterally from hyphae, later irregularly or occasionally verticillately branched. Phialides monophialidic, subulate to subclavate, 8–20 × 1.5–2.3 μm, hyaline. Conidia produced abundantly in pionnotes of aggregated conidiophores or on single conidiophores, delicate, hyaline. Microconidia ellipsoidal to clavate, aseptate, (3–)3.5–5 (–6.5) × 1–1.5(–1.7) μm (n = 30), hyaline, abundant. Macroconidia subcylindrical, moderately curved, slightly narrowing toward each end, apical cell often slightly hooked with a more or less pointed tip; basal cell not or scarcely pedicellate, predominantly 3–5-septate, 1-septate: 10–25 × 1.5–2 μm (n = 5), 3-septate: (20–)30–42(–48) × (1.8-)2–2.5(–2.7) μm (n = 40), 4–5-septate: (30–)37–43.5(–50) × (1.8-)2–2.5(–2.7) μm (n = 30), 6–7-septate: 40–48(–52) × (2–)2.3–2.7 μm (n = 25). Chlamydospores not observed.

Typification: Netherlands, Baarn, Groeneveld, perithecia on branch of Fagus sylvatica, July 1984, K.A. Seifert 357, holotype CBS H-3565, ex-type cultures BBA 64549 = CBS 512.84 = NRRL 20688. GenBank barcodes: HQ897749 (rpb2), HQ897885 (acl1).

Other material examined: USA, Pennsylvania, near Salt Springs State Park, on pyrenomycete stroma on Fagus, May 2007, T. Gräfenhan 2007-56, DAOM 235827; Canada, Quebec, Mayo, Forêt la Blanche, on pyrenomycete stroma on deciduous tree, Oct. 2007, T. Gräfenhan 2007-72, DAOM 235828.

Notes: To preserve the taxonomic concept of F. aquaeductuum var. medium sensu Wollenweber (1930), we typify Dialonectria ullevolea with an isolate from Fagus sylvatica collected in The Netherlands. The species produces a teleomorph and both microconidial and macroconidial synanamorphs; it seems to be pan-temperate and has been collected in Europe and North America.

Fusicolla Bonord., Handbuch der allgemeinen Mykologie p. 150. 1851.

Type species: Fusicolla betae (Desm.) Bonord. 1851.

Stroma erumpent from host with hyphae forming a slimy, pale orange sheet over the substratum, with perithecia fully or partially immersed. Perithecia scattered to gregarious, or in small groups, globose to pyriform with a short acute or disk-like papilla, pinched when dry, yellow, pale buff to orange, KOH–, smooth walled, usually 100–200 μm high. Asci cylindrical to narrowly clavate, with an apical ring, 8 uniseriate ascospores. Ascospores hyaline to pale brown, 1-septate, smooth or becoming slightly verrucose when mature. Conidiophores initially as lateral phialides on somatic hyphae, sometimes monochasial, verticillate or penicillate, hyaline. Phialides monophialidic, cylindrical to subulate, hyaline. Microconidia sparse or absent, ellipsoidal to allantoid, aseptate, hyaline. Macroconidia falcate, more or less straight, or moderately to clearly curved, slightly narrowing toward each end, apical cell often hooked with a more or less pointed tip, basal cell slightly pedicellate, predominantly 1–3-septate, or 3–5-septate, in one species up to 10-septate, hyaline, mostly thin-walled. Chlamydospores absent, sparse, or abundant, when present globose, single, in pairs or chains, sometimes in macroconidia.

Colonies on PDA slow growing, 30–55 mm diam in 14 d at room temperature, surface smooth, whitish to pale brown, pink or orange, sometimes with violet or reddish-brown tones, often entirely pionnotal; aerial mycelium sparse or abundant, turf-like, felt-like, or coremioid if with violet or reddish-brown tones.

Habitat: On soil or plant matter in contact with soil, on woody material, slime flux of trees, sometimes on stromata of other fungi, in flowing water including drinking water and sewage.

Notes: Fusicolla has generally been considered a synonym of Fusarium (see notes under F. betae below), but is adopted here for elements of the F. aquaeductuum and F. merismoides species complexes. Some of the varieties attributed to those two species by other authors are raised to species rank. The application of the name Fusarium merismoides var. chlamydosporale remains uncertain at this time, while F. merismoides var. crassum is transferred to Atractium above.

Eight other species were described in Fusicolla before the genus was synonymised with Fusarium by Wollenweber (1916, see below), six of them by Karsten. We have not seen the type specimens of any of these species, which have apparently not been revised since their original descriptions.

Accepted species

Fusicolla betae (Desm.) Bonord., Handbuch der allgemeinen Mykologie p. 150. 1851. Fig. 4.
Basionym: Fusisporium betae Desm., Ann. Sci. Nat., Bot., Sér. 1, 19: 436. 1830.

An external file that holds a picture, illustration, etc.
Object name is 79fig4.jpg

Fusicolla betae, lectotype (K). A. Sporodochium. B. Conidia and phialides. Scale bar in B = 10 μm.

  • Fusarium betae (Desm.) Sacc., Michelia 2: 132. 1880.

  • Pionnotes betae (Desm.) Sacc., Syll. Fung. 4: 726. 1886.

  • Pionnotes rhizophila var. betae (Desm.) De Wild. & Durieu, Prodr. Fl. Belg. 2: 367. 1898.

Typification: France, on tuber of Beta vulgaris, spring 1826, Desmazières, lectotype designated here K(M) 167520, Plantes Cryptogames du Nord de la France, no. 305; epitype designated here: Germany, Schleswig-Holstein, Kiel, on young plants of Triticum aestivum, Jan. 1983, C. Bauers, preserved culture BBA 64317. GenBank barcodes: HQ897781 (rpb2), HQ897917 (acl1).

Other material identified: Germany, northern Germany, rotting potato tuber, E. Langerfeld DE 8, FRC E-0114 = MRC 2196 = NRRL 47186. Turkey, roots of Papaver, 2007, G. Turhan, T.G. 2007-70. UK, on Beta vulgaris, IMI 105043 = NRRL 22133.

Notes: Morphologically, Fusicolla betae closely resembles other members of the Fusicolla merismoides species complex, and critical taxonomic reevaluation of this complex is required to develop reliable species concepts.

There has been confusion over the identity of this species with two independent concepts in the literature. Wollenweber (1916, no. 99, 100) probably studied type material of Fusisporium betae, but later listed the species as synonym of Fusarium merismoides irrespective of precedence of the older species epithet (Wollenweber & Reinking 1935). Following this, the genus Fusicolla was usually listed as a synonym of Fusarium, e.g. Carmichael et al. 1980. Alternatively, Chupp (1954, p. 111) cited Fusarium betae and “Fusidium betae Desm.” (probably a lapsus for Fusisporium) as synonyms of Cercospora beticola. He cited only the type of C. beticola and types of other Cercospora names synonymised with C. beticola; types of the Fusarium/Fusidium names were not cited. We conjecture that he proposed the synonymy based on the identity of the host and a general congruence in conidial size and septation. Crous & Braun (2003) followed the latter synonymy including Fusisporium betae as a synonym of Cercospora apii s. lat.; they also did not see type material (U. Braun, pers. comm.). Our studies of the lectotype designated above confirm that Desmaziéres' fungus produces sporodochia, phialides, and Fusarium-like conidia identical to those of the epitype selected above.

Fusicolla acetilerea (Tubaki, C. Booth & T. Harada) Gräfenhan & Seifert, comb. et stat. nov. MycoBank MB519434.
Basionym: Fusarium merismoides var. acetilereum Tubaki, C. Booth & T. Harada, Trans. Brit. Mycol. Soc. 66: 355. 1976.

Typification: Japan, Osaka, near Osaka University, soil, 1973, T. Miyoshi, holotype IFO 30040, ex-type cultures IMI181488 = BBA 63789 = NRRL 20827. GenBank barcodes: HQ897701 (rpb2), HQ897839 (acl1).

Other material identified: Australia, soil, FRC E-0052 = NRRL 13261, FRC E-0120 = NRRL 47187, FRC E-0121 = NRRL 47188, ICMP 10485 = NRRL 39744, IMI 175962 = NRRL 22137. Philippines, Nueva Vizcaya, FRC E-0164 = NRRL 47201. south Africa, soil, FRC E-0130 = NRRL 47191, FRC E-0136 = NRRL 47193, FRC E-0205 = NRRL 47210, FRC E-0226 = NRRL 47215, FRC E-0229 = NRRL 47844, FRC E-0257 = NRRL 47222, FRC E-0265 = NRRL 47224, FRC E-0287 = NRRL 47231, FRC E-0288 = NRRL 47232. Zambia, soil, FRC E-0208 = NRRL 47212.

Notes: This species produces both macroconidia and microconidia. The holotype is described, illustrated, and discussed by Tubaki et al. (1976) and Gerlach & Nirenberg (1982).

Fusicolla aquaeductuum (Radlk. & Rabenh.) Gräfenhan, Seifert & Schroers, comb. nov. MycoBank MB519435.
Basionym: Selenosporium aquaeductuum Radlk. & Rabenh., Kunst-Gewerbe-Blatt 49: 10. 1863.

  • Fusarium aquaeductuum (Radlk. & Rabenh.) Lagerh., Centralbl. Bakteriol. Parasitenk. 9: 655. 1891.

Typification: Germany, Bayern, München, water fountain near Gasteigberg, Nov. 1862, L. Radlkofer, lectotype designated here B 700014034. A permanent slide prepared by Radlkofer and sent to Wollenweber is selected here as the lectotype of Selenosporium aquaeductuum; it is the only known authentic material. epitype designated here: Germany, Berlin-Dahlem, Julius-Kühn-Institute (formerly BBA), isol. ex plugged water tap in BBA, May 1985, H.I. Nirenberg, ex-type cultures BBA 64559 = CBS 837.85 = NRRL 20865 = NRRL 37595. GenBank barcodes: HQ897744 (rpb2), HQ897880 (acl1).

Other material examined: Germany, Berlin, drinking water, 1974, W. Gerlach, BBA 63669 = CBS 734.79 = NRRL 20686; The Netherlands, Baarn, rubber tubing, 1953, A.L. van Beverwijk, CBS H-12677, CBS 268.53 = NRRL 22115.

Notes: No teleomorph is known for this species. For a description, illustrations, and discussion of the microconidial and macroconidial synanamorphs of this species as epitypified here, see Gerlach & Nirenberg (1982).

In Radlkofer (1863), two figures illustrate Selenosporium aquaeductuum, one showing 1–2(–4)-septate conidia borne on phialides. Wollenweber (1916) studied a permanent slide originally prepared by Radlkofer and drew the fungus with 1-septate and 3–4-septate conidia. On the herbarium sheet with that slide, Wollenweber noted the presence of two Fusarium species, F. aquaeductuum with 1-septate conidia, 18–22 × 1.5–2 μm and F. biasolettianum with 3-septate conidia, 30–55 × 2–2.5 μm. Based on similarities of the phenotype and substrate preferences, we classify Fusarium aquaeductuum in Fusicolla.

Wollenweber & Reinking (1935) included Microcera brachyspora Sacc. & Scalia as a synonym of F. aquaeductuum, but this should be confirmed with type studies.

Wollenweber (1931) linked Fusarium aquaeductuum var. aquaeductuum to “Nectria” episphaeria var. coronata (syn. “Nectria” purtonii, see below); subsequently this anamorph-teleomorph connection was accepted by Booth (1959), Gerlach and Nirenberg (1982), Samuels et al. (1991), and Rossman et al. (1999). According to our phylogenetic results, “Nectria” purtonii is not a member of Fusicolla but belongs to Stylonectria. The reported anamorph-teleomorph connection could not be confirmed here.

Fusicolla epistroma (Höhn.) Gräfenhan & Seifert, comb. nov. MycoBank MB519436.
Basionym: Dendrodochium epistroma Höhn., Sitzungsber. Kaiserl. Akad. Wiss., Wien, Math.-Naturwiss. Kl., Abt. 1, 118: 424. 1909.

  • Fusarium epistroma (Höhn.) C. Booth as F. epistromum, The Genus Fusarium p. 66. 1971.

Typification: Germany, Brandenburg, “Schmidt's Grund” near Tamsel, on old stromata of Diatrypella favacea on branches of Betula, Nov. 1906, P. Vogel, Sydow's Mycotheca germanica 648 Hymenula epistroma, lectotype B 700014042 designated here, isotypes FH 00286649, K, S F40143. epitype designated here: UK, England, Yorkshire, Aberford & Gundale, on Diatrypella on Betula, Apr. 1961, C. Booth, IMI 85601, ex-type cultures ATCC 24369 = BBA 62201 = NRRL 20461 = NRRL 20439. GenBank barcodes: HQ897765 (rpb2), HQ897901 (acl1).

Other material examined: Germany, Triglitz, 1907, O. Jaap, herb. von Höhnel 3087, FH 00286650.

Notes: For descriptions, illustrations, and discussion of the micro- and macroconidial synanamorphs of this species, see Booth (1971) and Gerlach & Nirenberg (1982).

An anamorph-teleomorph connection of F. epistromum with Nectria (“Cosmospora“) magnusiana was suggested by Höhnel (1909) and later followed by Jaap (1910), Booth (1959), Gerlach & Nirenberg (1982), and Samuels et al. (1991). Höhnel (1909) based his assumption on the observation that both fungi occurred on the same host fungus, Diatrypella favacea. However, he did not collect or observe the teleomorph together with his Dendrodochium epistroma. Wollenweber (1924, No. 539) studied a specimen of N. magnusiana collected by Jaap (Fungi selecti exs. 418) and questioned the link with Höhnel's anamorphic fungus. Booth's (1959) report of the anamorph-teleomorph connection included a drawing of the anamorph that lacks attribution to a specimen, but looks much like Wollenweber's Fusaria autographice delineata no. 539. The conidiophores and conidia are similar, having subulate phialides and non-septate, oblong to allantoid conidia. We compared Rehm's type material (S F84956, B 700014041) to the description given by Samuels et al. (1991) based on Jaap's exsiccati. In contrast to the latter, the KOH–ascomatal wall of the type specimen appears slightly verrucose and the colour is dark orange-brown with an obtuse apex and an ostiolar area that becomes almost black. Mature ascospores of Nectria magnusiana measure (12–)13–14.5(–15.5) × (5.5–)5.8–6.5(–6.8) μm and are significantly wider than those of the Jaap exsiccata studied by Samuels et al. (1991). The type material of N. magnusiana is reminiscent of Neonectria or Nectria s.str. An anamorph was associated with the same stroma from which perithecia developed. Its buff-coloured hymenium bears oblong-ellipsoidal microconidia conidia, 3.5–8 × 1–2 μm. These microconidia match those observed in two authentic collections of Dendrodochium epistroma (Sydow's Mycotheca Germanica 648 and Jaap's Fungi Selecti Exsiccati 349). Booth (1959) and Samuels et al. (1991) concluded that D. epistroma is the anamorph of N. magnusiana, both being host specific to Diatrypella favacea. Only a few Fusarium-like macroconidia were found on the type material of Dendrodochium epistroma, but macroconidia were lacking on the hymenium of the type collection of N. magnusiana. Interestingly, in culture the ex-type isolate of Fusicolla epistroma produces predominantly 3-septate conidia, rarely microconidia. From this, it remains unclear whether the associated anamorph on the type material of N. magnusiana is Fusicolla epistroma. Therefore, we decided to designate the epitype for F. epistroma based on Booth's material and not to consider the older species name Nectria magnusiana for this species.

Fusicolla matuoi (Hosoya & Tubaki) Gräfenhan & Seifert, comb. nov. MycoBank MB519437.
Basionym: Fusarium matuoi Hosoya & Tubaki, Mycoscience 45: 264. 2004.

  • Cosmospora matuoi Hosoya & Tubaki, Mycoscience 45: 262. 2004.

  • [= Fusarium splendens Matuo & Takah. Kobay., nom. nud., Trans. Mycol. Soc. Japan 2(4): 13. 1960].

Typification: Japan, Honshu, Yamagata Pref., Mamurogawa-machi, Mogami-gun, on Albizia julibrissin, Oct. 1958, T. Kobayashi, holotype TNS F-11127, ex-type culture MAFF 410976.

Other material examined: Iran, Prov. Gilan, near Bandarepahlavi, on rotting stalk of Zea mays, Oct. 1968, D. Ershad, BBA 62154 = FRC E-0089 = NRRL 47180. Japan, on Albizia julibrissin, Oct. 1959, T. Kobayashi, ATCC 18694 = CBS 581.78 = MAFF 238445 = NRRL 20427.

Notes: For a description, illustrations, and discussion of the teleomorph and micro- and macroconidial synanamorphs of this species, see Hosoya & Tubaki (2004).

Fusicolla merismoides (Corda) Gräfenhan, Seifert & Schroers, comb. nov. MycoBank MB519438.
Basionym: Fusarium merismoides Corda, Icon. Fung. 2: 4. 1838.

Typification: Czech Republic, Prague, on very wet shards of a plant pot, winter 1836, Corda, holotype PRM 155493.

Notes: Fusicolla merismoides is morphologically well characterised and has been widely accepted as a distinctive species (Wollenweber 1931, Booth 1971, Gerlach & Nirenberg 1982, Nelson et al. 1983, Leslie et al. 2006, Domsch et al. 2007); these authors provide descriptions, illustrations, and discussion of the macroconidial anamorph of this species. The morphological species concept was established by Wollenweber & Reinking (1935), who synonymised numerous taxa with Fusarium merismoides var. merismoides. Unlike F. betae, which is mainly known from roots and tubers of plants, F. merismoides is commonly isolated from soils, polluted water, slime fluxes of trees, rotting plant material, and many other substrates. Gräfenhan et al. (2008) discovered several phylogenetic lineages in the F. merismoides morphological species, including some ascospore isolates; the same conclusion can be drawn from publically available sequences attached to this name. We studied Corda's type material deposited in PRM and could not come to a satisfying conclusion on the selection of an appropriate epitype based solely on the macroconidial characteristics. Moreover, after examination of authentic material of Fusarium biasolettianum (PRM 155487), we could not confirm the reported synomymy with Fusicolla merismoides (Wollenweber & Reinking 1935). Macroconidia of Fusarium biasolettianum have almost an pointed and slightly hooked apical cell and a pedicellate basal cell (Fig. 5) that rather resemble macroconidium characteristics of Fusarium s. str. species. Rossman et al. (1999) mentioned Chrysogluten biasolettianum nom. rej., but there is no nomenclatural connection between this teleomorphic fungus and F. biasolettianum; the coincidental epithets indicate only that they were named in honour of the Italian botanist B. Biasoletto.

An external file that holds a picture, illustration, etc.
Object name is 79fig5.jpg

Fusarium biasolettianum, authentic material (PRM 155487). Macroconidia. Scale bar = 10 μm.

Most of the varieties within F. merismoides are distinct species, either within Fusicolla or in sister genera.

Fusicolla violacea Gräfenhan & Seifert, sp. nov. MycoBank MB519439.

  • = Fusarium merismoides var. violaceum W. Gerlach, Phytopathol. Z. 90: 34. 1977. nom. inval. Art. 37.

Latin description in Gerlach, Phytopath. Z. 90: 34-35. 1977 under the name “Fusarium merismoides var. violaceum“.

Typification: Iran, Prov. Gilan, near Rasht, on Quadraspidiotus perniciosus (San José insect) scaleon dying twig of Prunus domestica, Nov. 1968, W. Klett, holotype CBS 634.76, permanently cryopreserved culture, ex-type cultures BBA 62461 = NRRL 20896. GenBank barcodes: HQ897696 (rpb2).

Notes: For descriptions, illustrations, and discussion of the micro- and macroconidial synanamorphs of this species, see Gerlach (1977) and Gerlach & Nirenberg (1982).

The taxon was not validly published because the author did not designate a holotype, instead listing one living strain with accession numbers in two culture collections as “Cultura typica”.

Macroconia (Wollenw.) Gräfenhan, Seifert & Schroers, gen. et stat. nov. MycoBank MB519441.
Basionym: Nectria sect. Macroconia Wollenw., Angew. Bot. 8: 179. 1926. MycoBank MB519440.

Type species: Nectria leptosphaeriae Niessl in Krieger 1886, here recognised as Macroconia leptosphaeriae (Niessl) Gräfenhan & Schroers.

Stroma inconspicous or absent. Perithecia solitary, subglobose with or without a small apical papilla, collapsing cupulate when dry, orange to carmine red, KOH+ dark red to violet, sometimes with hyphal hairs arising from outer wall, usually 100–250 μm high. Asci cylindrical to narrowly clavate, with a simple apex, 8 uniseriate to partially biseriate ascospores. Ascospores yellowish, 1-septate, smooth or becoming striate when mature. Conidiophores initially as lateral phialides on somatic hyphae, later monochasial to verticillate, hyaline. Phialides monophialidic, cylindrical to subulate, hyaline. Microconidia absent or very rare, when present ellipsoidal to allantoid, hyaline. Macroconidia robust, subcylindrical to moderately curved, apical cell conical or hooked, basal cell mostly conspicuously pedicellate, 3–7(–14)-septate, hyaline, mostly thick-walled. Chlamydospores absent or rare, when present globose, single, in pairs, or in chains in hyphae.

Colonies on PDA slow- or very slow-growing, 7–10 or ~ 45 mm diam in 14 d at room temperature, whitish to orange or reddish brown; aerial mycelium abundant, with discrete pink, orange or reddish brown sporodochia or small pionnotes.

Habitat: Mostly growing on stromata of other ascomycetes on herbaceous plants or deciduous trees.

Notes: Based on the section name originally in Nectria (Wollenweber 1926), but also used as a “Gruppe” in Fusarium (Wollenweber & Reinking 1935), we raise Macroconia to generic rank here for five species with large Fusarium-like macroconidia and minute perithecia.

Accepted species

Macroconia leptosphaeriae (Niessl) Gräfenhan & Schroers, comb. nov. MycoBank MB519442.
Basionym: Nectria leptosphaeriae Niessl in Krieger, Fungi Saxonici Exsiccati. Die Pilze Sachsen's 4: No. 165. 1886.

  • Cucurbitaria leptosphaeriae (Niessl) O. Kuntze, Rev. Gen. Plant. 3: 461. 1898.

  • Hypomyces leptosphaeriae (Niessl) Wollenw., Fus. autogr. del., Edn 1: No. 57. 1916.

  • Lasionectria leptosphaeriae (Niessl) Petch, Trans. Brit. Mycol. Soc. 21: 267. 1938.

  • Cosmospora leptosphaeriae (Niessl) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 122. 1999.

  • ? = Fusarium sphaeriae var. majus Wollenw., Fus. autogr. del. No. 859. 1930.

Typification: Germany, Sachsen, Königstein Fortress, church yard, on Leptosphaeria doliolum on stems of Urtica dioica, Sept. & Oct. 1885, W. Krieger, Krieger's Fungi saxonici 165, lectotype designated here K(M) 165805, isotype B, BPI, K.

Other material examined: Canada, Ontario, Ottawa, Britannia, near Mud Lake, on Leptosphaeria on dead stem of Urtica dioica, July 2008, T. Gräfenhan 2008-15, DAOM 235833. Italy, Latio, ancient Etruscan village Corviano near Bomarzo, on Leptosphaeria on dead stem of Urtica dioica, Aug. 2008, T. Gräfenhan 2008-19, DAOM 235834. The Netherlands, Tilburg, on Leptosphaeria on dead stem of Urtica dioica, L. Rommelaars, CBS 100001, CBS-H 6030.

Notes: For description and illustration of the macroconidial anamorph and teleomorph of this species, see Weese (1916), Wollenweber (1916, No. 57; 1926; 1930, No. 859), Booth (1959, 1971), and Samuels et al. (1991).

The distinction between Macroconia leptosphaeriae and M. sphaeriae is based on the size of ascospores and conidia in the type collections. According to Wollenweber (1926), the ascospores of the type material of M. leptosphaeriae are smaller (14–18 × 5–5.5 μm) than those of M. sphaeriae (19–25 × 5.8–6.5 μm). These observations were partly confirmed by Samuels et al. (1991), who discussed the history and synonymy of the species. Five-septate conidia of M. leptosphaeriae measure 74–105 × 5–7 μm, whereas 5-septate conidia of M. sphaeriae are 45–73 × 4.5–5.5 μm (Wollenweber 1926). Further morphological studies of fresh collections from Leptosphaeria on Urtica are needed to confirm these species boundaries in these two species of Macroconia. Also, the occurrence of cellular hairs or sterile appendages on ascomatal walls needs to be reviewed critically. Therefore, we refrain from designating epitype material for M. leptopshaeriae here.

Macroconia cupularis(J. Luo & W.Y. Zhuang) Gräfenhan & Seifert, comb. nov. MycoBank MB519443.
Basionym: Cosmospora cupularis J. Luo & W.Y. Zhuang, Fungal Diversity 31: 88. 2008.

Typification: China, Zhejiang, Hangzhou, Taihuyuan, 500 m alt., on fruitbodies of a black ascomycete (Stylodothis sp.) on twigs of an unidentified tree, Sep. 2005, J. Luo and W.Y. Li 6790-2, holotype HMAS 97514, ex-type culture HMAS 173240. GenBank barcodes: EF121864 (ITS), EF121870 (28S rDNA).

Notes: For description, illustrations, and discussion of the teleomorph and macroconidial anamorph of this species, see Luo & Zhuang (2008). Its inclusion in Macroconia is inferred from the morphology and sequences provided in the protologue, although we did not include the species in our own analysis.

Macroconia gigas (J. Luo & W.Y. Zhuang) Gräfenhan & Seifert, comb. nov. MycoBank MB519444.
Basionym: Cosmospora gigas J. Luo & W.Y. Zhuang, Fungal Diversity 31: 85. 2008 non Fusarium gigas Speg., Anales Soc. Ci. Argent. 22: 221. 1886.

Typification: Taiwan, Nantou, Huisun Forestry Farm, 700 m alt., on rotten stem of bamboo associated with other fungi, Aug. 2005, W.Y. Zhuang 6598, holotype HMAS 99592, ex-type culture HMAS 173239; paratype ibid., W.Y. Zhuang, 6595, HMAS 97513. GenBank barcodes: EF121863 (ITS), EF121869 (28S rDNA).

Notes: For description, illustrations, and discussion of this teleomorph and macroconidial anamorph of this species, see Luo & Zhuang (2008). Its inclusion in Macroconia is inferred from the morphology and sequences provided in the protologue, although we did not include the species in our own analysis.

Macroconia papilionacearum (Seaver) Gräfenhan & Seifert, comb. nov. MycoBank MB519445.
Basionym: Nectria papilionacearum Seaver, Mycologia 1: 62. 1909.

  • Cosmospora papilionacearum (Seaver) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 124. 1999.

  • ? = Fusarium gigas Speg., Anales Soc. Ci. Argent. 22: 221. 1886.

Typification: USA, Missouri, Lebanon, on living Lespedeza with Parodiella perisporioides, Jul. 1887, Kellerman 1003, lectotype NY designated by Samuels et al. 1991.

Other material examined: USA, Florida, Tampa, near Hillsborough River State Park, on pyrenomycete on Fabaceae, Dec. 2006, T. Gräfenhan 2007-03, CBS 125495 = DAOM 238119.

Notes: For a description, illustrations, and discussion of the teleomorph, see Samuels et al. (1991). Our material collected in Florida closely resembles the description of M. papilionacearum given by Samuels et al. (1991), except for the smooth ascospores; the specimen from Florida has striate ascospores. In culture, the macroconidial anamorph of the Florida collection corresponded with the sketchy descriptions of Fusarium gigas (Wollenweber 1916, Wollenweber & Reinking 1935, Booth 1971, Gerlach & Nirenberg 1982). We found no anamorphic structures during our examination of the type material of Fusarium gigas (Paraguay, Arroyo-Guazu, on sterile pyrenomycete on culm of Bambusaceae, Jan. 1882, B. Balansa, Pl. du Paraguay 3471, Spegazinni's Fungi Guaranitici 426, B 700014033, B 700014032, PAD). The synonymy of M. papilionacearum with the macroconidial anamorph represented by the name “Fusariumgigas should be confirmed using fresh South American material.

Macroconia sphaeriae (Fuckel) Gräfenhan & Schroers, comb. nov. MycoBank MB519446.
Basionym: Fusarium sphaeriae Fuckel, Jahrb. Nassauischen Vereins Naturk. 23–24: 370. 1870.

  • ? = Nectria leptosphaeriae var. macrospora Wollenw., Angew. Bot. 8: 187. 1926.

Typification: Germany, Hessen, Rheingau, Reichartshausen near Oestrich-Winkel, on Leptosphaeria (Sphaeria) dioica on Urtica dioica, in spring, L. Fuckel, Fuckel Fungi Rhenani 212, lectotype designated here G 00111017, isotypes B, DAOM 126601 = Herb. Barbey-Boissier 2634.

Notes: The macroconidial anamorph and the teleomorph of this species is described, illustrated, and discussed by Wollenweber (1916, No. 58; 1926). The proposed new combination moves an anamorphically typified epithet into a teleomorphically typified genus, resulting in a valid, legitimate but technically incorrect name under the present Art. 59.

Macroconia sphaeriae can be distinguished from M. leptosphaeriae by its larger ascospores and smaller conidia (Wollenweber 1926; see M. leptosphaeriae above). The lectotype material in G had a few ascomata, but the two perithecia studied contained neither asci nor ascospores. The isotype material lacked teleomorph structures. We follow Wollenweber's (1926) conclusion and treat the two as separate species.

Mariannaea G. Arnaud ex Samson, Stud. Mycol. 6: 74. 1974.

Type species: mariannaea elegans (Corda) Samson 1974.

Stroma absent or inconspicuous. Perithecia solitary, globose with a flat apex, not collapsing or collapsing by lateral pinching when dry, pale yellow, orange or brown, KOH–, smooth or finely roughened, 250–350 μm high. Asci cylindrical to narrowly clavate, with a sometimes inconspicuous apical ring, 8 uniseriate or apically biseriate ascospores. Ascospores hyaline, 1-septate, smooth to spinulose when mature. Conidiophores verticillate to penicilliate, hyaline, with conidiogenous cells arising directly from the stipe or from whorls of metulae on lower parts of the stipe, the stipe hyaline or yellowish brown at the base, often roughened at the base. Phialides monophialidic, flask shaped, hyaline, usually with obvious periclinal thickening and inconspicuous collarettes. Conidia aseptate, hyaline, in imbricate chains that eventually collapse to form slimy heads. Chlamydospores produced by some species.

Notes: Mariannaea is a common hyphomycete genus in soil and on woody substrates, and includes mononematous species with verticillate conidiophores, phialidic conidiogenous cells, and often imbricate chains of aseptate conidia. The genus was validly published by Samson (1974) and his concept is accepted for this anamorph typified genus, with the addition of teleomorph characters above. Although the conidia are small, the conidiophores and conidia are not comparable to microconidia of the Fusarium complex, and the genus is included here because of the similarity of its teleomorph to the Cosmospora complex. In common with many of the teleomorph-anamorph connections discussed in this paper, the exact identities of the relevant morphs are imprecise. A teleomorph of a fungus similar to M. elegans was described from specimens collected in Jamaica and Venezuela as “Nectria” mariannaea by Samuels & Seifert (1991). Although it is Cosmospora-like, the name was not transferred by Rossman et al. (1999) and remains misclassified in Nectria. As discussed below, it seems unlikely that “N.” mariannaea is the teleomorph of M. elegans s. str., and we are unable to infer its identity with any other of the named anamorphic species. An LSU sequence for the ex-type culture of N. mariannaeae was deposited in GenBank (AY554242) by Schroers et al. (2005); the LSU of the ex-type of M. samuelsii (HQ843766) differs by 5 substitutions from N. mariannaeae, and 3 substitutions from M. aquaticola. Thus, given the limited amount of variation in the ITS and LSU normally seen in the Nectriaceae, the phylogenetic data suggest that M. aquaticola, N. mariannaeae and M. samuelsii represent different species. We elect not to describe a new genus for N. mariannaeae, preferring to use the older Mariannaea as a holomorphic genus. Transferring it to Mariannaea would create a tautonym (Art. 23.4), thus, we have elected to leave this name in limbo until its taxonomic status can be more thoroughly evaluated.

Some of the species described in Mariannaea do not belong to the Nectriaceae, but to the Cordycipitaceae (Liang 1991, Liu et al. 2002). A phylogenetic analysis of internal transcribed spacer sequences of nectriacious Mariannaea species was provided by Li et al. (2009) and suggests the existence of four species, including the type, M. elegans, a variety distinguished from the type that seems to be distinct at the species level, i.e. M. aquaticola, M. camptospora, and M. elegans var. punicea. To this we add a fifth species, M. samuelsii described below.

Mariannaea samuelsii Seifert & Bissett, sp. nov. MycoBank MB519447. Fig. 6.

An external file that holds a picture, illustration, etc.
Object name is 79fig6.jpg

Mariannaea samuelsii, ex-type strain. A, B. Obverse and reverse of 14 d old colony on oatmeal agar. C, D. Conidiophores showing verticillate branching. E. Imbricate conidial chains. F. Conidia. Scale bars = 10 μm.

Coloniae in agaro malto et peptono confecto post 7 dies 21 mm diam, aureo-brunneae vel brunneolo-aurantiae; in agaro farina avenae confecto 28–29 mm diam, sub luce aurantio-griseae, obscuritate griseo-aurantiae. Conidiophora 100–200 μm longa, stipite 2–3.5 μm lato, bis vel ter verticillata, verticillos terminales (2–)3–5 phialidum, in verticillis subterminalibus 25–35 μm distantibus 1–3 phialides ferentia; raro phialides singulae circa 20 μm longae ex hyphis repentibus orientes. Phialides 12–30 μm longae, in parte latissima 2–3.5 μm latae, subulatae, in summo periclinaliter inspissatae, collari inconspicuo cylindrico praeditae. Conidia 3.5–7.5 × 2.5–3.5 μm, late fusiformia vel ellipsoidea, symmetrica, sed saepe asymmetrice ex apertura conidiogena protrusa, hyalina, levia, in catenis imbricatis saepe collabentibus adhaerentia. Holotypus DAOM 235814 (cultura dessicata).

On Blakeslee's MEA: Conidiophores arising from the agar surface, from aerial hyphae or fascicles, mostly 100–200 μm long, the axis 2–3.5 μm wide, branching 2–3 level verticillate, with a terminal whorl of (2–)3–5 phialides, and 1–2 lower nodes of 1–3 phialides spaced 25–35 μm apart, sometimes with a basal branch that repeats the pattern of 1–2 levels of verticillate branching, rarely with phialides single and terminal on an intercalary cell about 20 μm long. Phialides 12–30 μm long, 2–3.5 μm wide at broadest part (19.8 +/- 0.9 × 2.9 +/- 0.06, n = 25), subulate, sometimes with base slightly swollen, often longest in basal whorls, periclinal thickening obvious with phase contrast, collarette inconspicuous, about 1 × 1 μm, cylindrical. Conidia 3.5–7.5 × 2.5–3.5 μm (6.0 +/- 0.2 × 3.1 +/- 0.06, n = 25), broadly fusiform or ellipsoidal, L/B ratio about 2–2.5, symmetrical but often sitting asymmetrically on conidiogenous aperture, hyaline, smooth-walled, in imbricate chains that quickly collapse into hyaline, slimy heads. Chlamydospores rarely produced, globose to ellipsoidal, hyaline, ~5–10 × 3–5 μm, in chains of up to five cells.

Colonies on Blakeslee's MEA after 7 d about 21 mm diam, golden brown to brownish orange (5–6D6) in centre, fading towards entire margin, planar, with sparsely lanose aerial mycelium and fascicles, reverse concolourous; sporulation more intense on MEA in presence of 12:12 h fluorescent light:continuous darkness, agar surface mealy. On OA 28–29 mm diam, orange gray (5B2) in light, and grayish orange (5D2) in dark, fading towards entire, thin margin, with moderately dense lanose aerial mycelium and fascicles, reverse concolourous.

Typification: Guatemala, Zacapa Prov., San Lorenzo Mt., isolated from soil under Podocarpus sp., surface litter and humus horizons, containing roots, 0–2 cm, 12 Jul. 1986, John Bissett, herb. DAOM 235814, ex-type culture CBS 125515. GenBank barcodes: HQ843766 (28S rDNA), HQ843767 (ITS), HQ897752 (rpb2), HQ897888 (acl1).

Notes: Mariannaea samuelsii is morphologically similar to M. elegans, the type of the genus (Samson 1974), and the recently described M. aquaticola (Li et al. 2009) in producing verticillate conidiophores and imbricate chains of fusiform conidia. The conidiophores of M. aquaticola and M. samuelsii are generally less elaborately branched than those of M. elegans, and lack basal roughening. The size ranges of the conidia of these three species overlap, with conidia of M. samuelsii (3.5–7.5 × 2.5–3.5 μm) intermediate in length between the shorter conidia of M. elegans (4–6 × 1.5–2.5 μm) and the longer conidia of M. aquaticola (5–10 × 2–4.5 μm). Mariannaea elegans produces chlamydospores, which have not been seen in M. aquaticola and are rarely and sparsely produced in M. samuelsii.

Mariannaea samuelsii differs by four base-pair substitutions (two in the ITS1, two in the ITS2) from M. aquaticola, its sister species.

Microcera Desm., Ann. Sci. Nat., Bot., sér. 3, 10: 359. 1848.

  • = Pseudomicrocera Petch, Trans. Brit. Mycol. Soc. 7: 164. 1921.

Type species: microcera coccophila Desm. 1848.

Stroma and/or white byssus covering host. Perithecia solitary or in groups, globose, with a blunt papilla, collapsing cupulate or pinched when dry, orange to dark red, KOH+ dark red or violet, finely roughened, 200–400 μm high. Asci cylindrical to narrowly clavate, with an apical ring, 8 uniseriate ascospores. Ascospores hyaline to pale yellow-brown, 1(–3)-septate, smooth or becoming tuberculate when mature. Conidiophores initially as lateral phialides on somatic hyphae, later monochasial, verticillate to penicilliate, hyaline, usually forming discrete sporodochia or synnemata on the host. Phialides monophialidic, cylindrical to subulate to subclavate, hyaline. Microconidia absent. Macroconidia pale, orange, pink or bright red in mass, subcylindrical, moderately curved, or conspicuously curved, apical cell often slightly or conspicuously hooked, basal cell scarcely to conspicuously pedicellate, mostly (0–)3–5-septate, but up to 12 septate in one species, hyaline, mostly thick-walled. Chlamydospores not observed.

Colonies on PDA slow growing, 18–35 mm diam in 14 d at room temperature, surface smooth, felt-like or floccose, whitish to bright orange-red, sometimes with violet or vinaceous tones; aerial mycelium sparse or appressed, sporulation occurring in sporodochia or sometimes in slimy masses (pionnotes).

Habitat: Mostly parasites of scale insects, also reported on aphids, adelgids, and sometimes isolated as saprobes from soil or plant debris.

Notes: Along with Atractium discussed above, Microcera was a generic name used for synnematous Fusarium-like fungi, but in this case mostly parasites of scale insects. Our phylogenetic analysis confirms the significance of this ecological association, and the genus is here redefined to include additional non-synnematous species associated with scale insects, some of which are sometimes also found on other substrates. Until the 1920's, the generic name Microcera was widely used for entomogenous species with slender, falcate conidia (McAlpine 1899, 1904; Parkin 1906; Trabut 1907; Miyabe & Sawada 1913; Petch 1921). The original concept of Microcera included one species, M. coccophila, based on two collections made by Roberge near Caen, France. Desmazières did not mention perithecia on these specimens, but from the conidial shape he inferred a close relationship with Fusarium. Tulasne & Tulasne (1861, 1865) studied these and additional specimens from the type and other locations. They redescribed the species as a holomorph as Sphaerostilbe flammea, but concluded that Desmazières' Microcera was a “Stilbum” with long, curved, Fusarium-like macroconidia. Petch (1921) revised this group of entomogenous species and studied the type material of M. coccophila, finding perithecia on well-developed stromata associated with the synnemata of the anamorph. Mature perithecia were red with ascospores measuring 12–18 × 5–7 μm (Petch 1921).

The taxonomic synonymy of Microcera with Fusarium followed the work of Wollenweber. Wollenweber (1916) first classified F. ciliatum in Microcera, based on his study of two herbarium specimens originally identified as Fusarium pallens (Wollenweber 1916; 1st edition, No. 435, 436). Later, Wollenweber & Reinking (1935) discarded Microcera and placed its species in Fusarium. In his first monographic revision of Fusarium, Wollenweber (1931) did not consider M. coccophila, but subsequently revised his generic concept profoundly (Wollenweber & Reinking 1935). Then, M. coccophila, along with species described in other genera such as Atractium, Discofusarium, Fusidium, Fusisporium, Fusoma, Microcera, Pionnotes, Pseudomicrocera, and Selenosporium were placed in Fusarium. Of these, only the type species of Pseudomicrocera (Ps. henningsii) would now be considered a member of the Microcera clade. After Wollenweber's work, Microcera was included as a synonym in major revisions of Fusarium, e.g. Booth (1971), Gerlach & Nirenberg (1982), Nelson et al. (1983), and Leslie et al. (2006).

Twenty species were included in Microcera by various authors, and the present status of most species is known (Table 4). We presently accept four species, which can be keyed out as follows.

Table 4.

Species attributed to Microcera and their current status. Basic nomenclatural data from Index Fungorum (www.indexfungorum.org).

Species, authority and year of publicationStatusReference
M. acuminata (Ellis & Everh.) Höhn. 1919 = Fusarium acuminatum Wollenweber & Reinking 1935
M. aurantiicolaPetch 1921 = M. larvarum This paper
M. brachyspora Sacc. & Scalia 1904 ? = Fusicolla aquaeductuum Wollenweber & Reinking 1935
M. ciliata (Link) Wollenw. 1916 = “Fusarium” ciliatum, status unclear
M. clavariella Speg. 1886 = Cladosterigma fusispora Pat. Seifert 1985b
M. coccidophthoraPetch 1921 = Fusarium tasmanicum (McAlpine) Rossman 1983Rossman 1983
M. coccophila Desm. 1848 Accepted species This paper
M. curta Sacc. 1909 = M. larvarum This paper
M. erumpens Ellis & Everh. 1894 Unknown
M. fujikuroiMiyabe & Sawada 1913 = M. diploa This paper
M. henningsii (Koord.) Petch 1914 = M. diploa This paper
M. massariae Sacc. 1886 = “Fusariumciliatum, see above Wollenweber & Reinking 1935
M. merrillii Syd. 1914 = M. diploa This paper
M. mytilaspidisMcAlpine 1904 = Fusarium lateritium var. longumWollenweber & Reinking 1935
M. orthospora Syd. 1924 = Mycogloea orthospora (Syd.) R. McNabb ex Dingley 1989 Dingley 1989
M. parlatoriae Trab. 1907 = M. larvarum This paper
M. pluriseptata Cooke & Massee 1888 = M. coccophila This paper
M. rectispora Cooke & Massee = Tetracrium rectisporum (Cooke & Massee) Petch 1921Petch 1921
M. tasmanicaMcAlpine 1904 = Fusarium tasmanicum (McAlpine) Rossman 1983Rossman 1983
M. tonduzii Pat. 1912 = M. larvarum This paper

KEY TO SPECIES OF MICROCERA

  • 1. Macroconidia straight to slightly curved, up to 140 μm long, up to 12 septate........................................................................ M. coccophila
    1. Macroconidia distinctly curved, usually less than 120 μm long, mostly 3–5 septate................................................................................... 2

  • 2. Macroconidia slender, 40–120 μm long.......................................................................................................................................... M. diploa
    2. Macroconidia usually less than 40 μm long................................................................................................................................................. 3

  • 3. Agar colonies with red pigments...................................................................................................................................................... M. rubra
    3. Agar colonies lacking red pigments............................................................................................................................................ M. larvarum

Accepted species

Microcera coccophila Desm., Ann. Sci. Nat., Bot., Sér. 3, 10: 359. 1848. Fig. 7A, B.
Basionym: Tubercularia coccophila (Desm.) Bonord., Abh. Geb. Mykol., p. 96. 1864.

An external file that holds a picture, illustration, etc.
Object name is 79fig7.jpg

Two Microcera species. A, B. Microcera coccophila. A. Habit, with conical red perithecia on a stroma growing over scale insect and flame-like synnema emerging from the top. B. Macroconidia. C, D. M. larvarum. C. Flame-like conidiomata on scale insect. D. Conidia. Scale bars = 10 μm.

  • Fusarium coccophilum (Desm.) Wollenw. & Reinking, Die Fusarien, p. 34. 1935.

  • Fusarium episphaeria f. coccophilum (Desm.) W.C. Snyder & H.N. Hansen, Amer. J. Bot. 32: 662. 1945.

  • = Microcera pluriseptata Cooke & Massee in Cooke, Grevillea 17: 43. 1888.

Typification: France, Normandy, near Caen, on Eulecanium tiliae (nut scale) on living and young trunks of Salix and Fraxinus excelsior, Feb. 1847, M. Roberge, lectotype designated here K (M) 165807, Plantes Cryptogames de France, Ed. II, Ser. I, No. 1350, isotypes P, K (M) 165806, Plantes Cryptogames de France Ed. I, Ser. I, No. 1750.

Additional material examined: Japan, Saitama, Hiki-gun, Ogawa-machi, on scale insect on Broussonetia kazinoki × B. papryifera, Jul. 1993, G. Okada.

Notes: The macroconidial anamorph and the teleomorph of this species as lectotypified here is described and discussed in detail by Petch (1921). For description, illustrations, and further taxonomic synonyms of the anamorph, see Gerlach & Nirenberg (1982).

There has been confusion about synonymies and anamorph-teleomorph connections between this fungus, M. diploa, and M. larvarum. Petch (1921) synonymised the anamorphic name Atractium flammeum Berk. & Ravenel with Microcera coccophila, arguing that Sphaerostilbe flammea Tul. & C. Tul. represented the holomorph of M. coccophila and that Sphaerostilbe coccophila Tul. & C. Tul. was actually a different species, M. larvarum (as “Nectria” aurantiicola). He cited two Desmazières exsiccati of M. coccophila, namely Plantes Cryptogames de France, Ed. I, Ser. I, No. 1750 and ibid. Ed. II, Ser. I, No. 1350. Our reexamination of the latter confirms Petch's observation that mature perithecia have 1-septate ascospores, 12–18 × 5–7 μm, associated with the anamorph. “Nectriaflammea reportedly has larger ascospores (Dingley 1951, 15–24 × 6–10 μm; Booth 1971, 1981b, 16–20 × 7.5–10 μm). The anamorph-teleomorph connection of Microcera coccophila with “Nectriaflammea needs to be critically reevaluated.

Gräfenhan et al. (2008) noted the occurrence of several phylogenetic species among anamorph and teleomorph collections that are morphologically similar to M. coccophila, M. diploa, and M. larvarum.

Microcera diploa (Berk. & M.A. Curtis) Gräfenhan & Seifert, comb. nov. MycoBank MB519448.
Basionym: Nectria diploa Berk. & M.A. Curtis, J. Linn. Soc., Bot. 10: 378. 1869.

  • Cucurbitaria diploa (Berk. & M.A. Curtis) O. Kuntze, Rev. Gen. Plant. 3: 461. 1898.

  • Creonectria diploa (Berk. & M.A. Curtis) Seaver, Mycologia 1: 190. 1909.

  • Calonectria diploa (Berk. & M.A. Curtis) Wollenw., Angew. Bot. 8: 193. 1926.

  • Cosmospora diploa (Berk. & M.A. Curtis) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 121. 1999.

  • = Fusarium coccidicola Henn. [as “coccideicola”], Bot. Jahrb. Syst. 34: 57. 1904.

  • = Aschersonia henningsii Koord., Bot. Untersuch. Java p. 213. 1907.

  • Microcera henningsii (Koord.) Petch, Ann. Roy. Bot. Gard. Peradeniya 5: 533. 1914.

  • Pseudomicrocera henningsii (Koord.) Petch, Trans. Brit. Mycol. Soc. 7: 164. 1921.

  • = Microcera fujikuroi Miyabe & Sawada, J. Coll. Agric. Tohoku Imp. Univ. 5: 83. 1913.

  • = Microcera merrillii Syd., Ann. Mycol. 12: 576. 1914.

Typification: Cuba, on individual scale insects on bark, C. Wright 606 ex Herb. Berk., Fungi Cubensis Wrightiana 767, lectotype K designated by Booth 1971, isotypes FH 00286651, FH 00286652, NYS.

Notes: The holotype of this species is consistent with the descriptions of the teleomorph by Booth (1971) and Rossman (1983). The macroconidial anamorph is described by Booth (1971), Gerlach & Nirenberg (1982), and Rossman (1983). As explained in the introduction, under the present Art. 59, the proposed new combination results in a technically incorrect but valid and legitimate name.

Microcera diploa is an entomogenous species reported from many tropical and subtropical regions (Booth 1971, Rossman 1983), commonly found on various scale insects sitting on several plant species. Booth (1971) studied the type collection and reported pustules of perithecia on a stroma associated with the anamorph. From our observations of the same material, it is clear that the stromata developed over individual scale insects. In agreement with Rossman (1983), we follow Booth's decision to interpret the Cuban specimen as the type of Nectria diploa. Several Fusarium species were synonymised with M. diploa, namely F. derridis, F. juruanum, and F. pentaclethrae, which were described only from herbaceous material (Wollenweber & Reinking 1935). We studied Hennings' material (F. derridis = B 700014017; F. juruanum = B 700014035, B 700014036; F. pentaclethrae = B 700014037), and none seem to be insect-associated. Therefore, we reject these synonymies, except for F. coccidicola as listed above.

Microcera larvarum (Fuckel) Gräfenhan, Seifert & Schroers, comb. nov. MycoBank MB519449. Fig. 7C, D.
Basionym: Fusarium larvarum Fuckel, Jahrb. Nassauischen Vereins Naturk. 23–24: 369. 1870.

  • = Microcera parlatoriaeTrab., Bull. Agric. Algérie Tunisie 13: 33. 1907.

  • = Microcera curta Sacc., Ann. Mycol. 7: 437. 1909.

  • = Microcera tonduzii Pat., Bull. Soc. Mycol. France 28: 142. 1912.

  • = Microcera aurantiicola Petch, Trans. Brit. Mycol. Soc. 7: 163. 1921.

Typification: Germany, Hessen, Rheingau, near Oestrich-Winkel, on larva cuticles of insects on apple trees, in spring, L. Fuckel, lectotype designated here G 00111015 epitype designated here: Iran, Prov. Gilan, near Rasht, on Quadraspidiotus perniciosus (San José insect) scale on Prunus domestica, Oct. 1968, W. Gerlach & D. Ershad, epitype BBA, ex-type cultures BBA 62239 = CBS 738.79 = MUCL 19033 = NRRL 20473. GenBank barcodes: HQ897768 (rpb2), HQ897904 (acl1).

Notes: For descriptions, illustrations, and further taxonomic synonyms of the teleomorph and macroconidial anamorph of this species, see Petch (1921), Wollenweber (1931), Booth (1971, 1981a, c), and Gerlach & Nirenberg (1982).

Our phylogenetic analysis and that of Bills et al. (2009) clearly indicate that the two varieties of M. larvarum segregated by Gerlach (1977) warrant species rank; M. larvarum var. rubrum is recognised as a distinct species below. Bills et al. (2009) studied parnafungin production by species of this complex, and their data suggest that perhaps two additional phylogenetic species may exist in this group.

The synonymy of Microcera larvarum with “Nectriaaurantiicola cited by Booth (1971, 1981a), Gerlach & Nirenberg (1982), and Rossman et al. (1999) should be critically reviewed.

Microcera rubra Gräfenhan & Seifert, sp. nov. MycoBank MB519450.

  • = Fusarium larvarum var. rubrum W. Gerlach, Phytopath. Z. 90: 38. 1977. nom. inval. Art. 37.

Latin description in Gerlach, Phytopath. Z. 90: 38. 1977 under the name “Fusarium” larvarum var. rubrum.

Typification: Iran, Prov. Gilan, near Rasht, on Quadraspidiotus perniciosus (San José insect) scale on Prunus domestica, Oct. 1968, W. Gerlach & D. Ershad, holotype CBS H-714, ex-type cultures BBA 62460 = CBS 638.76 = NRRL 20475 = NRRL 22111 = NRRL 22170. GenBank barcodes: HQ897767 (rpb2), HQ897903 (acl1).

Notes: For descriptions, illustrations, and discussion of this macroconidial species, see Gerlach (1977) and Gerlach & Nirenberg (1982); for phylogenetic relationships, see Bills et al. (2009).

The taxon was not validly published because the author did not designate a holotype, instead listing one living strain with accession numbers in two culture collections as “Cultura typica”.

Pseudonectria Seaver, Mycologia 1: 48. 1909.

Type species: Pseudonectria rousseliana (Mont.) Clements & Shear 1931, here recognised as P. buxi (DC.) Seifert, Gräfenhan & Schroers.

Notes: Pseudonectria as presently circumscribed is not monophyletic (Fig. 1), with two species branching out in separate clades in the Nectriaceae. The type species of Pseudonectria, together with an undescribed taxon, forms a sister clade to Atractium. The second species, “Pseudonectriapachysandricola together with “Nectriadiminuta and “N.rubropeziza, falls between the terminal and basal Fusarium-like clade. Therefore, only one species is presently recognised in this genus, with the teleomorph typifying the oldest available generic name Pseudonectria 1909, and the anamorph representing the type of the later generic name Chaetodochium 1932. There is presently no acceptable generic name for “Pseudonectriapachysandricola, which is well described and illustrated by Dodge (1944) and Rossman et al. (1993).

The anamorphs of Pseudonectria are fairly well understood pathogens on the Buxaceae (Bezerra 1963, Rossman et al. 1993), but these species are usually cited under their anamorph names, i.e.Volutellabuxi and “V.” pachysandricola. Because these species do not share common morphological characters with Volutella s. str. (see below) and are phylogenetically distinct, these anamorph names should not be used. The phylogenetic relationship of a biologically and morphologically similar species described from Ruscus aculeatus, “V.” rusci, remains unresolved.

Pseudonectria buxi (DC.) Seifert, Gräfenhan & Schroers, comb. nov. MycoBank MB519451.
Basionym: Tubercularia buxi DC., Flore française, Edn. 3 (Paris) 6: 110. 1815.

  • Chaetostroma buxi (DC.) Corda, Icon. Fung. 2: 30. 1838.

  • Volutella buxi (DC.) Berk., Outl. Brit. Fungi p. 340. 1860.

  • Chaetodochium buxi (DC.) Höhn., Mitt. bot. Inst. tech. Hochsch. Wien 9: 45. 1932.

  • = Pseudonectria rousseliana (Mont.) Clements & Shear, Genera of Fungi p. 280. 1931.

  • Nectria rousseliana Mont. in Castagne, Cat. P1. Marseille Suppl. p. 44. 1851. For additional obligate synonyms, see Rossman et al. 1993.

Notes: Bezerra (1963) and Rossman et al. (1993) redescribed and illustrated both the anamorph and teleomorph of P. buxi, a common pathogen of Buxus sempervirens. The conidia of the anamorph tend toward fusiform, a shape not seen in species of Volutella s. str., and the sporodochia tend to be broadly attached to the substratum. These are subtle characters, and at present we cannot suggest robust morphological characters to unequivocally distinguish the anamorphs of Pseudonectria from Volutella. However, the teleomorphs are rather different, with the perithecia of Volutella being red and those of Pseudonectria being green.

Because this fungus has a known teleomorph and anamorph, Art. 59 applies, and our transfer of an anamorphically typified epithet to a teleomorphically typified generic name is technically incorrect according to the present ICBN, but it is valid and legitimate.

StylonectriaHöhn., Sitzungsber. Kaiserl. Akad. Wiss., Wien, Math.-Naturwiss. Kl., Abt. 1, 124: 52. 1915.

Type species: Stylonectria applanata Höhn. 1915.

Stroma thin, whitish or yellow, hyphal or subiculum-like. Perithecia gregarious in groups of up to 20, subglobose, pyriform to subcylindrical, with a rounded or broad, circular, flat disc on a venter-like neck, sometimes laterally collapsing when dry, pale yellow, orange-red, orange-brown, or pale to dark red, KOH+ dark red to purple, yellow in lactic acid, smooth, usually shiny, slightly iridescent, 150–250(–350) μm high. Perithecial wall consisting of two regions: inner region of hyaline, thin-walled, compressed, elongate cells; outer region of distinct, isodiametric to oblong, angular or globose, thick-walled cells. Asci cylindrical to clavate, apex simple or with a ring, with 8 uniseriate, biseriate or irregularly disposed ascospores. Ascospores hyaline or yellow to pale brown, 1-septate, cylindrical to allantoid or ellipsoidal, smooth or tuberculate, generally thick-walled. Conidiophores initially formed mostly as unbranched phialides on somatic hyphae, occasionally loosely branched, sometimes forming small sporodochia. Phialides monophialidic, almost cylindrical to subcylindrical, often with a distinct collarette. Microconidia sparsely produced, allantoid to lunulate, slightly to strongly curved, aseptate, in slimy heads. Macroconidia orange in mass, subcylindrical or moderately to strongly curved, falcate, mostly 0–1-septate, apex narrower than base, apical cell blunt or hooked, basal cell not or scarcely pedicellate. Chlamydospores not observed.

In culture on PDA slow- to very slow-growing, 10–30 mm diam in 14 d at room temperature, surface white, later becoming off-white to pale or bright orange, occasionally with orange sporodochia; aerial mycelium mostly lacking, if present, sparse and appressed margin smooth to broadly lobed

Habitat: Restricted to stromata of ascomycetes, mainly in the Diaporthales.

Notes: Stylonectria was described by Höhnel (1915) as an anamorph genus with the type and only species, S. applanata, for which the teleomorph was considered to be “Nectria” applanata var. succinea. Booth (1959) presented convincing evidence that Höhnel (1915) actually was dealing with a teleomorphic fungus, which was further explained by Rossman et al. (1999). Species of Stylonectria are considered to be host specific, probably to the fungal host, which itself may be host specific to the plant.

Accepted species

Stylonectria applanata Höhn., Sitzungsber. Kaiserl. Akad. Wiss., Wien, Math.-Naturwiss. Kl., Abt. 1, 124: 52. 1915.

  • = Nectria applanata var. succinea Höhn., Sitzungsber. Kaiserl. Akad. Wiss., Wien, Math.-Naturwiss. Kl., Abt. 1, 124: 51. 1915.

Typification: Austria, Niederösterreich, near Sonntagsberg, on stromata of Melogramma bulliardii on dead twigs of Corylus avellana, Aug. 1914, P. Strasser, lectotype designated here FH 00286663.

Notes: For descriptions and discussion of the teleomorph, microconidial anamorph, and macroconidial synanamorph of this species, see von Höhnel (1915) and Weese (1916).

Von Höhnel (1915) distinguished “Nectriaapplanata var. succinea from “N.applanata var. applanata based on the pale yellow colour of the translucent perithecia. Otherwise, the two varieties were described with identical macro- and microscopic characters. Because host specificity is an important character for distinguishing species of Stylonectria (cf. Gräfenhan 2009), we recognise S. applanata as a distinct species from S. carpini, described below, i.e. Nectria applanata var. applanata.

Stylonectria carpini Gräfenhan, nom. nov. MycoBank MB519452.

  • Nectria applanata Fuckel, Jahrb. Nassauischen Vereins Naturk. 25–26: 310. 1871 (1872).

  • Cucurbitaria applanata (Fuckel) O. Kuntze, Rev. Gen. Plant. 3: 460. 1898.

  • Dialonectria applanata (Fuckel) Petch, Trans. Brit. Mycol. Soc. 25: 170. 1941.

Etymology: The species epithet is derived from the plant host genus Carpinus.

Typification: Germany, Hessen, Rheingau, Aepfelbach im Oestricherwald, on black pyrenomycete on decaying, corticated branches of Carpinus betulus, L. Fuckel, Fuckel Fungi Rhenani 2356, lectotype designated here G 00111018, isotypes B 700014054, FH 00286648, K, DAOM 119800 = Herb. Barbey-Boissier 862.

Other material examined: Austria, Niederösterreich, Gießhübl, Wasserspreng, Talgrund, (Finsterer Gang), MTB 7863/1, on Melanconis spodiaea on Carpinus betulus, Aug. 2006, H. Voglmayr W.J. 3013, DAOM 235819. Germany, Schleswig-Holstein, near Stegelkamp, Naturwaldzelle Endern, on pyrenomycete on Carpinus betulus, Aug. 2008, T. Gräfenhan 2008-17, DAOM 235829.

Notes: This species produces both a micro- and a macroconidial synanamorph in addition to a teleomorph. Our examination of Höhnel's type material of Stylonectria applanata (FH 00286663) and that of Fuckel's “Nectriaapplanata (G 00111018) suggests the two species are not conspecific, but both are species of Stylonectria; the latter is therefore renamed here.

The distribution of Stylonectria carpini corresponds to the distribution of Carpinus betulinus in Europe. In North America, a different species of Stylonectria occurs on a black pyrenomycete on the congeneric native host, Carpinus caroliniana, and has a microconidial anamorph in culture and a distinctly different teleomorph. Collections made from a pyrenomycete on Betula are morphologically similar to S. carpini but phylogenetically distinct.

Stylonectria purtonii (Grev.) Gräfenhan, comb. nov. MycoBank MB519453.
Basionym: Sphaeria purtonii Grev., Scot. Crypt. Fl. 6: 23. 1828.

  • Nectria purtonii (Grev.) Berk., Outl. Brit. Fung. p. 394. 1860.

  • Dialonectria purtonii (Grev.) Cooke, Grevillea 12: 110. 1884.

  • Cucurbitaria purtonii (Grev.) O. Kuntze, Rev. Gen. Plant. 3: 461. 1898.

  • Cosmospora purtonii (Grev.) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 124. 1999.

Typification: UK, Scotland, Edinburgh, Rosslyn Woods, on black pyrenomycete on small branches of coniferous tree, 1820, Greville, lectotype E designated by Booth 1958.

Other material examined: France, Provence, St. Remy, on old stromata of pyrenomycete on Coronilla emerus, Oct. 1974, W. Gams, culture CBS 717.74. Germany, Nordrhein-Westfalen, Detmold, Externsteine, on small branches of felled trees of Picea abies, Apr. 2007, T. Gräfenhan 2007-30, DAOM 235818.

Notes: For descriptions, illustrations, and further taxonomic synonyms of the teleomorph as well as microconidial and macroconidial synanamorphs of this species, see Booth (1959) and Samuels (1976).

Stylonectria wegeliniana (Rehm) Gräfenhan, Voglmayr & Jaklitsch, comb. nov. MycoBank MB519454.
Basionym: Nectria episphaeria var. wegeliniana Rehm, Hewigia 30: 260. 1891.

  • Dialonectria wegeliniana (Rehm) Petch, Trans. Brit. Mycol. Soc. 21: 266. 1938 as D. wegeliana.

  • Cosmospora wegeliniana (Rehm) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 131. 1999.

Typification: switzerland, Heimiswyl bridge near Bern, on Hapalocystis bicaudata (= Pseudovalsa berkeleyi) on dry branches of Ulmus, Oct. 1887, Wegelin, Rehm Ascomyceten 1045, lectotype designated here S F86597, isotypeNY.

Other material examined: Austria, Niederösterreich, Distr. Mödling, Comm. Hinterbrühl, Wassergspreng, Finsterer Gang west of Gießhübl, margin of a forest road, elev. 400 m, map grid 7863/3, on Hapalocystis bicaudata on corticated dead branches of Ulmus glabra attached to the living tree, May 2009, H. Voglmayr, WU 29855, culture CBS 125490.

Notes: This species produces microconidia and macroconidia in culture; the teleomorph was only found in nature. For a description, illustrations, and discussion of the species, see Weese (1916).

Volutella Tode 1790: Fr. 1832. Fungi Mecklenb. Sel. 1: 28. 1790: Syst. Mycol. 3: 458, 466 1832, nom. cons. [non Volutella Forsk. 1775 (Lauraceae)]

Type species: Volutella ciliata (Alb. & Schw.: Fr.) Fr. 1832, typus cons.

Perithecia nonstromatic, pyriform, collapsing by lateral pinching or not collapsing when dry, brownish orange to brownish red, yellow in 100 % lactic acid, darkest around papilla, hyphal hairs covering surface, hyaline, thick walled. Perithecial wall 15–25 um wide, with two intergrading layers of angular cells; cells next to centrum thin walled; cells of layer region thick walled. Asci narrowly clavate to broadly cylindrical, apex with or without refractive ring, eight-spored. Ascospores fusiform or biconic, equally or unequally 2-celled, smooth or finely roughed, hyaline, white in mass, obliquely uniseriate or partially biseriate near base, completely filling each ascus. Conidiophores aggregated into sporodochia or synnemata, with an inconspicuous basal stroma; unbranched, hyaline setae around margin of conidiomata. Synnemata, when produced, determinate, pale, composed of a stipe of parallel hyphae and a divergent capitulum of conidiophores giving rise to a slimy conidial mass; differentiated marginal hyphae absent. Conidiophore branching once or twice monochasial, 2-level verticillate, monoverticillate or irregularly biverticillate. Conidiogenous cells monophialidic, hyaline, subulate, usually with conspicuous periclinal thickening. Conidial masses slimy, white, yellow, orange or pink. Conidia aseptate, hyaline, ellipsoidal, ovate or oblong. Chlamydospores produced in culture by some species. Verticillium-like synanamorph present in some species: Conidiophores hyaline, with 2 or more whorls of conidiogenous cells; phialides and conidia with similar characters to those described for the conidiomata. Agar cultures growing relatively slowly, usually less than 30 mm diam in 14 d, with little aerial mycelium.

Notes: Volutella is a classical hyphomycete genus that has received little study, despite the common occurrence and broad distribution of its species. The genus is typified by V. ciliata, which has sporodochial conidiomata with conspicuous hyaline, thick-walled, unbranched, spine-like setae, phialidic conidiogenous cells arising from more or less penicillately branched conidiophores, and ameroconidia accumulating in a profuse, colourful slime. Domsch et al. (2007) provided a general overview of the type and a few other soil-borne species of the genus. In anticipation of a more comprehensive revision of Volutella, the inclusion of one synnematous species in this genus is discussed here.

Volutella s. str. should be restricted to the clade that includes the type species, V. ciliata, V. consors (referred to as V. minima by Domsch et al. 2007), and the synnematous V. citrinella. The teleomorphs associated with Volutella provide clues to its polyphyly. “Cosmosporaconsors was reported as the teleomorph of V. ciliata by Samuels (1977, as Nectria consors); the identity of the anamorph was later changed to V. minima by Domsch et al. (2007). This species differs from V. ciliata primarily by its cylindrical conidia. Volutella citrinella, considered at more length below, has a similar teleomorph, “Nectriastilbellae. Neither teleomorph genus is appropriate, with Cosmospora now restricted to species with Acremonium-like anamorphs, discussed above, and Nectria is restricted to species with Tubercularia anamorphs (Hirooka et al. 2011). We have elected not to describe a new teleomorph genus for this clade, preferring to refer to these fungi by the oldest available generic name Volutella. As noted by Summerbell et al. (2011) in their discussion of Trichothecium, replacing a classic and well known generic name with a virtually unknown teleomorphically typified generic name would be taxonomically capricious. The other two holomorphic species with anamorphs attributed to Volutella are species presently classified in Pseudonectria (see above), which produce setose perithecia and aseptate ascospores, rather different than the smooth- or rough-walled perithecia and 1-septate ascospores of V. citrinella and V. consors.

The synnematous fungus V. citrinella was formerly known as Stilbella aciculosa (Seifert 1985a) but is more appropriately classified in Volutella. There have been scattered comments in the literature about synnematous species of Volutella, including the comment by Domsch et al. (2007) that some strains or species are “short stipitate”. Thus, the inclusion of synnematous species only subtly alters the existing generic concept. Although there was scant mention of Volutella in the monograph of the synnematous genus Stilbella by Seifert (1985a), it was included in the key to Stilbella-like genera because of these observations by other authors.

Few of the approximately 120 described species of Volutella have been revised, and most species were seldom reported after their original descriptions. A preliminary survey of type specimens accessioned in K by Seifert (unpublished) suggests that many of the described species represent Colletotrichum, Sarcopodium, and other anamorphic genera. Comparatively few species that conform to the modern concept were uncovered. However, given the morphological variation we have seen in unidentifed specimens and cultures, we suggest Volutella s. str. will ultimately include many more species.

Accepted species

Volutella ciliata (Alb. & Schwein.) Fr., Syst. Mycol. 3: 467. 1832.
Basionym: Tubercularia ciliata Alb. & Schwein., Consp. fung. p. 68. 1805.

Typification: We were unable to locate authentic material of T. ciliata; the sole specimen in the Schweinitz herbarium (PH) dates to a later publication (Schweinitz 1822). Because this name is formally conserved, careful attention must be paid to appropriate typification, and we chose not to propose a neotype or epitype here.

Volutella consors (Ellis & Everh.) Seifert, Gräfenhan & Schroers, comb. nov. MycoBank MB519455.
Basionym: Dialonectria consors Ellis & Everh., J. Mycol. 4(12): 122. 1888.

  • Nectria consors (Ellis & Everh.) Seaver, Mycologia 1: 61. 1909.

  • Nectriella consors (Ellis & Everh.) Sacc., Syll. fung. 9: 941. 1891.

  • Cosmospora consors (Ellis & Everh.) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 119. 1999.

  • ? = Volutella comata Ellis, Bull. Torrey Bot. Club 9: 20. 1892.

  • ? = Volutella minima Höhn., Sber. Akad. Wiss. Wien, Math.-naturw. Kl., Abt. 1, 118: 1543. 1909.

Typification: USA: Louisiana, St. Martinsville, Sep. 1888, Langlois 1485. holotype NY (examined by Samuels 1977).

Material examined: Volutella comata. USA, New Jersey, Newfield, on fallen petioles on Robinia, June 1881, Ellis North American Fungi no. 811. Isotypes DAOM, K.

Notes: Volutella consors predates the commonly used name for this morphological species, V. minima and the newly synonymised V. comata. As noted in the Introduction, the transfer of a teleomorph typified name into an anamorph genus creates a technically incorrect name that is nevertheless valid and legitimate.

Several morphological variants of this species exist including specimens with reddish brown sporodochial tissues and white conidial masses as in the isotypes of V. comata or white stipes and bright yellow conidial masses as in several specimens from India in CBS-H and IMI 205174, as Stilbella sp. In addition, some living strains have Verticillium-like synanamorphs as noted but not illustrated by Matsushima (1975) and visible in the strain CBS 552.89. This is probably a species complex, and the synonymies with V. minima and V. comata should be reevaluated in future studies.

Volutella citrinella(Cooke & Massee) Seifert, comb. nov. MycoBank MB519456. Fig. 8.
Basionym: Stilbum citrinellum Cooke & Massee, Grevillea 16: 81. 1887.

An external file that holds a picture, illustration, etc.
Object name is 79fig8.jpg

Volutella citrinella, colony and microscopic characters. A, B. Colony on oatmeal agar showing typical purple pigment and yellowish slime of the synnemata. C, D. Determinate synnemata developed in culture. E. Conidiophores. F. Conidia. G, H, I. Seta-like marginal hypha in culture (DAOM 226716, 165570). Scale bars = 10 μm.

  • = Stilbum aciculosum Ellis & Everhart, J. Mycol. 1: 153. 1885.

  • Stilbella aciculosa (Ellis & Everhart) Seifert, Stud. Mycol. 27: 44. 1985 non Volutella aciculosa (Ellis & Harkn.) Sacc., Syll. fung. 4: 687. 1886.

  • = Nectria stilbellae Samuels & Seifert, Sydowia 43: 250. 1991.

  • Cosmospora stilbellae (Samuels & Seifert) Rossman & Samuels in Rossman, Samuels, Rogerson & Lowen, Stud. Mycol. 42: 125. 1999.

For other synonyms, see Seifert (1985a) under Stilbella aciculosa.

Notes: The holomorph was described and illustrated by Samuels & Seifert (1991). Seifert (1985a) noted that the hyphae of the synnema stipes of this species sometimes become slightly thick-walled, and, if they diverge from the synnema, may appear somewhat seta-like. With the sister relationship of V. citrinella to V. ciliata revealed by the phylogenetic analysis, the taxonomic significance of this morphological observation becomes clear. Examination of three cultures of this fungus and reexamination of a slide of the holotype of Stilbum aciculosum revealed thickened hyphae with nearly occluded lumina in all of them. These hyphae (Fig. 7F–H) are 1.5–3 μm wide with cell walls thickened up to 1 μm at the base, thinning towards the acute apex. They are common on specimens from nature. In culture, they are less frequent sometimes giving the synnemata a slightly hirsute appearance, but they generally do not penetrate into the capitulum.

In addition to the distributional records provided by Seifert (1985a), specimens have since been examined originating in Grenada, New Zealand, and South Africa.

Acknowledgments

This study would have been impossible without the tremendous contributions of Gary J. Samuels. His habit of sharing collections selflessly and his holistic approach to describing fungi make him a role model for both beginning and experienced mycologists. He could have been easily a co-author of this study, but then we could not have written this acknowledgement for him!

We thank Kerry O'Donnell for providing unpublished DNA sequences (rpb2) for many Fusicolla and Microcera strains deposited with NRRL. We enjoyed numerous fruitful discussions with Scott Redhead on interpretations of Article 59 of the ICBN but assume full responsibility for the nomenclatural decisions made here. We thank the curators of the herbaria B, DAOM, G, FH, K, L, PAD, PH, PRM, S, and UPS for the loan of type and other material or for searching for relevant material. We are especially indebted to Harrie Sipman at the Botanical Museum in Berlin for providing information on fungal specimens deposited by H. Richter and H.W. Wollenweber.

We appreciate the continued support of and consultation with our mentor Prof. W. Gams, who prepared the Latin diagnosis for M. samuelsii, while proclaiming his horror at the nomenclatural mutations proposed here by his devoted former students.

We are grateful for DNA sequencing by Gerry Louis-Seize, and for the support from the Canadian Collection of Fungal Cultures. Last but not least, we thank W. Jaklitsch and H. Voglmayr for permission to study their specimens of Stylonectria and the examination of type material of S. wegeliniana.

Notes

Taxonomic novelties: New genus: Macroconia (Wollenw.) Gräfenhan, Seifert & Schroers. New species: Dialonectria ullevolea Seifert & Gräfenhan, Fusicolla violacea Gräfenhan & Seifert, Mariannaea samuelsiiSeifert & Bissett, Microcera rubra Gräfenhan & Seifert. New combinations: Atractium holubovae (Seifert, S.J. Stanley & K.D. Hyde) Seifert, Atractium crassum (Wollenw.) Seifert & Gräfenhan, Cosmospora arxii (W. Gams) Gräfenhan & Schroers, Cosmospora berkeleyana (P. Karst.) Gräfenhan, Seifert & Schroers, Cosmospora butyri (J.F.H, Beyma) Gräfenhan, Seifert & Schroers, Cosmospora cymosa (W. Gams) Gräfenhan & Seifert, Cosmospora khandalensis (Thirum. & Sukapure) Gräfenhan & Seifert, Cosmospora lavitskiae (Zhdanova) Gräfenhan & Seifert, Cosmospora viridescens (C. Booth) Gräfenhan & Seifert, Fusicolla acetilerea(Tubaki, C. Booth & T. Harada) Gräfenhan & Seifert, Fusicolla aquaeductuum (Radlk. & Rabenh.) Gräfenhan, Seifert & Schroers, Fusicolla epistroma (Höhn.) Gräfenhan & Seifert, Fusicolla matuoi (Hosoya & Tubaki) Gräfenhan & Seifert, Fusicolla merismoides (Corda) Gräfenhan, Seifert & Schroers, Macroconia cupularis (J. Luo & W.Y. Zhuang) Gräfenhan & Seifert, Macroconia gigas (J. Luo & W.Y. Zhuang) Gräfenhan & Seifert, Macroconia leptosphaeriae (Niessl) Gräfenhan & Schroers, Macroconia papilionacearum (Seaver) Gräfenhan & Seifert, Macroconia sphaeriae (Fuckel) Gräfenhan & Schroers, Microcera diploa (Berk. & M.A. Curtis) Gräfenhan & Seifert, Microcera larvarum (Fuckel) Gräfenhan, Seifert & Schroers, Pseudonectria buxi (DC.) Seifert, Gräfenhan & Schroers, Stylonectria purtonii (Grev.) Gräfenhan, Stylonectria wegeliniana (Rehm) Gräfenhan, Voglmayr & Jaklitsch, Volutella citrinella (Ellis & Everh.) Seifert, Volutella consors (Ellis & Everh.) Seifert, Gräfenhan & Schroers. New name: Stylonectria carpini Gräfenhan.

References

  • Akaike H (1974). A new look at the statistical model identification. IEEE Transactions on Automatic Control 19: 716–723. [Google Scholar]
  • Beyma van FH (1938). Beschreibung einiger neuer Pilzarten aus dem “Centraalbureau voor Schimmelcultures” Baarn (Holland). Zentralblatt für Bakteriologie, Parasitenkunde und Infektionskrankheiten, 2. Abt. 99: 381–394. [in German] [Google Scholar]
  • Bezerra JL (1963). Studies on Pseudonectria rousseliana. Acta Botanica Neerlandica 12: 58–63. [Google Scholar]
  • Bills GF, Platas G, Overy DP, Collado J, Fillola A, Jiménez MR, Martín J, del Val AG, Vicente F, Tormo JR, Peláez F, Calati K, Harris G, Parish C, Xu D, Roemer T (2009). Discovery of the parnafungins, antifungal metabolites that inhibit mRNA polyadenylation, from the Fusarium larvarum complex and other hypocrealean fungi. Mycologia 101: 449–472. [Abstract] [Google Scholar]
  • Booth C (1959). Studies of Pyrenomycetes IV. Nectria (part I). Mycological Papers 73: 1–115. [Google Scholar]
  • Booth C (1971). The genus Fusarium. 1–234. International Mycological Institute, Kew.
  • Booth C (1981a). Nectria aurantiicola. CMI Descriptions of pathogenic fungi and bacteria 714: 1–2. [Google Scholar]
  • Booth C (1981b). Nectria flammea. CMI Descriptions of pathogenic fungi and bacteria 715: 1–2. [Google Scholar]
  • Booth C (1981c). Perfect states (teleomorphs) of Fusarium species. 446–452. In: Fusarium: Disease, Biology, and Taxonomy (Nelson PE, Toussoun TA, Cook RJ, eds.). The Pennsylvania State University Press, University Park and London.
  • Brandes EW (1919). Banana wilt. Phytopathology 9: 339–389. [Google Scholar]
  • Carmichael JW, Kendrick WB, Conners IL, Sigler L (1980). Genera of Hyphomycetes. i–x, 1–386. Edmonton, Canada; University of Alberta Press.
  • Chaverri P, Salgado C, Hirooka Y, Rossman AY, Samuels GJ (2011). Delimitation of Neonectria and Cylindrocarpon (Nectriaceae, Hypocreales, Ascomycota) and related genera with Cylindrocarpon-like anamorphs. Studies in Mycology 68: 57–78 (this issue). [Europe PMC free article] [Abstract] [Google Scholar]
  • Chupp C (1954). A monograph of the fungus genus Cercospora. 1–667. Ithaca, NY; Chupp.
  • Crous PW, Braun U (2003). Mycospherella and its anamorphs.1. Names published in Cercospora and Passalora. CBS Biodiversity Series 1: 1–571. [Google Scholar]
  • Crous PW, Braun U, Schubert K, Groenewald JZ (2007). Delimiting Cladosporium from mophologically similar genera. Studies in Mycology 58: 33–56. [Europe PMC free article] [Abstract] [Google Scholar]
  • Crous PW, Seifert KA (1998). Megaconidia as an additional taxonomic character in Cylindrocladium, with a note on Cylindrocladiopsis. Fungal Diversity 1: 53–64. [Google Scholar]
  • Cock AWAM de, Lévesque A (2004). New species of Pythium and Phytophthora. Studies in Mycology 50: 481–488. [Google Scholar]
  • Dingley JM (1951). The Hypocreales of New Zealand. II. The genus Nectria. Journal of the Royal Society of New Zealand 79: 177–202. [Google Scholar]
  • Dingley JM (1989). Reappraisal of Microcera othospora and Myxosporium hoheriae. Memoirs of the New York Botanical Garden 49: 206–209. [Google Scholar]
  • Dodge BO (1944). A new Pseudonectria on Pachysandra. Mycologia 36: 532–537. [Google Scholar]
  • Dodge CW (1935). Medical Mycology. C. V. Mosby Co., St. Louis.
  • Domsch KH, Gams W, Anderson T (2007). Compendium of soil fungi. 2nd ed. 1–672. IHW-Verlag, Eching, Germany.
  • Fryar SC, Booth W, Davies J, Hodgkiss JI, Hyde KD (2004). Distribution of fungi on wood in the Tutong River, Brunei. Fungal Diversity 17: 17–38. [Google Scholar]
  • Gams W (1971). Cephalosporium-artige Schimmelpilze (Hyphomycetes). 1–262. G. Fischer, Stuttgart.
  • Gams W, Nirenberg HI (1989). A contribution to the generic definition of Fusarium. Mycotaxon 35: 407–416. [Google Scholar]
  • Gams W, Zaayen van A (1982). Contribution to the taxonomy and pathogenicity of fungicolous Verticillium species. I. Taxonomy. Netherlands Journal of Plant Pathology 88: 57–78. [Google Scholar]
  • Gerlach W (1972). Fusarien aus Trinkwasserleitungen. Annales Agriculturae Fenniae 11: 298–302. [Google Scholar]
  • Gerlach W (1977). Drei neue Varietäten von Fusarium merismoides, F. larvarum und F. chlamydosporum. Phytopathologische Zeitschrift 90: 31–42. [Google Scholar]
  • Gerlach W, Nirenberg HI (1982). The genus Fusarium — A pictorial atlas. Mitteilungen der Biologischen Bundesanstalt für Land- und Forstwirtschaft 209: 1–406. [Google Scholar]
  • Gräfenhan T (2009). Contributions to the taxonomy of the ascomycete genus Cosmospora and its anamorphs. Zeitschrift für Mykologie 75: 175–188. [in German] [Google Scholar]
  • Gräfenhan T, Nirenberg HI, Seifert KA (2008). The hidden diversity of Cosmospora and its Fusarium anamorphs. Journal of Plant Pathology 90 (3, Suppl.): S3.18. [Google Scholar]
  • Groves JW (1946). The North American species of Dermea. Mycologia 38: 352–431. [Google Scholar]
  • Hall TA (1999). BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41: 95–98. [Google Scholar]
  • Hirooka Y, Rossman AY, Chaverri P (2011). Morphological and phylogenetic revision of the Nectria cinnabarina species complex. Studies in Mycology 68: 35–56 (this issue). [Europe PMC free article] [Abstract] [Google Scholar]
  • Höhnel F von (1909). Fragmente zur Mykologie: VI. Mitteilung (Nr. 182 bis 288). Sitzungsberichte der mathematisch-naturwissenschaftlichen Klasse der Kaiserlichen Akademie der Wissenschaften, Wien 118 275–452. [Google Scholar]
  • Höhnel F von (1912). Fragmente zur Mykologie (XIV. Mitteilung, Nr. 719 bis 792). Sitzungsberichte der mathematisch-naturwissenschaftlichen Klasse der Kaiserlichen Akademie der Wissenschaften, Wien 121: 339–784. [Google Scholar]
  • Höhnel F von (1915). Fragmente zur Mykologie (XVII. Mitteilung, Nr. 876 bis 943). Sitzungsberichte der mathematisch-naturwissenschaftlichen Klasse der Kaiserlichen Akademie der Wissenschaften, Wien 124: 49-159. [Google Scholar]
  • Holmgren PK, Holmgren NH, Barnett LC (1990). Index herbariorum. Part I: The herbaria of the world. 8th edition. New York Botanical Garden, New York.
  • Holubová-Jechová V, Gams W, Nirenberg HI (1994). Revisiones Generum Obscurorum Hyphomycetum: a revision of the Selenosporium species described by A.C.J. Corda. Sydowia 46: 247–256. [Google Scholar]
  • Hosoya T, Tubaki K (2004). Fusarium matuoi sp. nov. and its teleomorph Cosmospora matuoi sp. nov. Mycoscience 45: 261–270. [Google Scholar]
  • Jaap O (1910). Verzeichnis der bei Triglitz in der Prignitz beobachteten Ascomyceten. Verhandlungen des Botanischen Vereins der Provinz Brandenburg 52: 109–150. [Google Scholar]
  • Karsten PA (1891). Symbolae ad mycologicam fennicam. Pars XXX. Meddelanden af Societas pro Fauna et Flora Fennica 18: 61–74. [Google Scholar]
  • Kornerup A, Wanscher JH (1978). Methuen handbook of colour, 3rd edition. 1–252. Eyre Methuen, London.
  • Leslie JF, Summerell BA, Bullock S (2006). The Fusarium laboratory manual. 1–388. Blackwell Publishing, Ames, Oxford, and Victoria.
  • Li C, Kurniawati E, Hyde KD (2009) Morphological and molecular characterization of Mariannaea aquaticola sp. nov. collected from freshwater habitats. Mycological Progress 9: 337–343. [Google Scholar]
  • Liang Z-Q (1991). Determination and identification of anamorph of Cordyceps pruinosa. Acta Mycologica Sinica 10: 72–80. [Google Scholar]
  • Lindau G (1910). Dr. L. Rabenhorst's Kryptogamen-Flora von Deutschland, Oesterreich und der Schweiz. Zweite Auflage. Erster Band: Die Pilze Deutschlands, Österreichs und der Schweiz. X. Abteilung: Fungi imperfecti: Hyphomycetes (zweite Hälfte), Dematiaceae (Phaeophragmiae bis Phaeostaurosporae), Stilbaceae, Tuberculariaceae, sowie Nachträge, Nährpflanzenverzeichnis und Register. Leipzig.
  • Link JHF (1816). Observationes in ordines plantarum naturales. Dissertatio secunda. Magazin. Gesellschaft Naturforschender Freunde zu Berlin 7: 25–45. [Google Scholar]
  • Link JHF (1825). Caroli a Linne Species plantarum: exhibentes plantas rite cognitas, ad genera relatas, cum differentiis specificis, nominibus trivialibus, synonymis selectis, locis natalibus, secundum systema sexuale digestas 6(2): 106. Impensis G. C. Nauk, Berolini. [Google Scholar]
  • Liu Z-Y, Liang Z-Q, Liu A-Y, Yao Y-J, Hyde KD, Yu Z-N (2002). Molecular evidence for teleomorph–anamorph connections in Cordyceps based on ITS-5.8S rDNA sequences. Mycological Research 106: 1100–1108. [Google Scholar]
  • Lombard L, Crous PW, Wingfield BD, Wingfield MJ (2010). Systematics of Calonectria: a genus of root, shoot and foliar pathogens. Studies in Mycology 66: 1–71. [Europe PMC free article] [Abstract] [Google Scholar]
  • Luo J, Zhuang W-Y (2008). Two new species of Cosmospora (Nectriaceae, Hypocreales) from China. Fungal Diversity 31: 83–93. [Google Scholar]
  • Luo J, Zhuang W-Y (2010). Chaetopsinectria (Nectriaceae, Hypocreales), a new genus with Chaetopsina anamorphs. Mycologia 102: 976–984. [Abstract] [Google Scholar]
  • Lutzoni F, Kauff F, Cox CJ, McLaughlin D, Celio G, et al. (2004). Assembling the fungal tree of life: progress classification and evolution of subcellular traits. American Journal of Botany 91: 1446–1480. [Abstract] [Google Scholar]
  • Matsushima T (1975). Icones Microfungorum a Matsushima lectorum. 1–209 Published by the author, Kobe, Japan.
  • Matsushima T (1980). Saprophytic microfungi from Taiwan. Part 1. Hyphomycetes. Matsushima Mycological Memoirs 1: 1–82. [Google Scholar]
  • McAlpine D (1899). Fungus diseases of citrus trees in Australia, and their treatment. 1–132. Australia, Melbourne; Agriculture Department of Victoria.
  • McAlpine D (1904). Two new fungi parasitic on scale insects. Journal of the Department of Agriculture, Victoria 2: 646–648. [Google Scholar]
  • McNeill J, Barrie FR, Burdet HM, Demoulin V, Hawksworth DL, Marhold K, Nicolson DH, Prado J, Silva PC, Skog JE, Wiersema JH, Turland NJ (2006). International Code of Botanical Nomenclature (Vienna Code). Regnum Vegetabile 146. A.R.G. Gantner Verlag KG.
  • Miyabe K, Sawada K (1913). On fungi parasitic on scale-insects found in Formosa. Journal of the College of Agriculture, Tohoku Imperial University 5: 73–90. [Google Scholar]
  • Nees von Esenbeck, CDG (1817). System der Pilze und Schwämme. 1–334. Würzburg, Germany.
  • Nelson PE, Toussoun TA, Marasas WFO (1983). Fusarium species: An illustrated manual for identification. 1–193. Pennsylvania State University Press, University Park, Pennsylvania.
  • Nguyen HDT, Seifert KA (2008). Description and DNA barcoding of three new species of Leohumicola from South Africa and the United States. Persoonia 21: 57–69. [Europe PMC free article] [Abstract] [Google Scholar]
  • Nirenberg HI (1976). Untersuchungen über die morphologische und biologische Differenzierung in der Fusarium-Sektion Liseola. Mitteilungen aus der Biologischen Bundesanstalt für Land- und Forstwirtschaft 169: I–v + 1–117. [Google Scholar]
  • Nirenberg HI, Samuels GJ (2000). Nectria and Fusarium. II. Cosmospora zealandica comb. nov. and its anamorph, Fusarium zealandicum sp. nov. Canadian Journal of Botany 78: 1482–1487. [Google Scholar]
  • Nowrousian M, Kück U, Loser K, Weltring KM (2000). The fungal acl1 and acl2 genes encode two polypeptides with homology to the N- and C-terminal parts of the animal ATP citrate lyase polypeptide. Current Genetics 37: 189–193. [Abstract] [Google Scholar]
  • O'Donnell K (1993). Fusarium and its near relatives. 224–233. In: The fungal holomorph: mitotic, meiotic and pleomorphic speciation in fungal systematic (Reynolds DR, Taylor JW, eds.). CAB International: Wallingford, UK.
  • O'Donnell K, Sarver BA, Brandt M, Chang DC, Noble-Wang J, Park BJ, Sutton DA, Benjamin L, Lindsley M, Padhye A, Geiser DM, Ward TJ (2007). Phylogenetic diversity and microsphere array-based genotyping of human pathogenic Fusaria, including isolates from the multistate contact lens-associated U.S. keratitis outbreaks of 2005 and 2006. Journal of Clinical Microbiology 45: 2235–2248. [Europe PMC free article] [Abstract] [Google Scholar]
  • O'Donnell K, Sutton DA, Fothergill A, McCarthy D, Rinaldi MG, Brandt ME, Zhang N, Geiser DM (2008). Molecular phylogenetic diversity, multilocus haplotype nomenclature, and in vitro antifungal resistance within the Fusarium solani species complex. Journal of Clinical Microbiology 46: 2477–2490. [Europe PMC free article] [Abstract] [Google Scholar]
  • O'Donnell K, Sutton DA, Rinaldi MG, Gueidan C, Crous PW, Geiser DM (2010). A novel multi-locus sequence typing scheme reveals high genetic diversity of human pathogenic members of the Fusarium incarnatum-F. equiseti and F. chlamydosporum species complexes within the U.S. Journal of Clinical Microbiology 47: 3851–3861. [Europe PMC free article] [Abstract] [Google Scholar]
  • Palm ME, Gams W, Nirenberg HI (1995). Plectosporium, a new genus for Fusarium tabacinum, the anamorph of Plectosphaerella cucumerina. Mycologia 87: 397–406. [Google Scholar]
  • Parkin J (1906). Fungi parasitic upon scale-insects (Coccidae and Aleurodidae): a general account with special reference to Ceylon forms. Annals of the Royal Botanic Gardens, Peradeniya 3: 11–82. [Google Scholar]
  • Pascoe IG (1990). Fusarium morphology. I. Identification and characterisation of a third conidial type, the mesoconidium. Mycotaxon 37: 121–160. [Google Scholar]
  • Petch T (1921). Studies in entomogenous fungi. Transactions of the British Mycological Society 7: 89–167. [Google Scholar]
  • Posada D, Crandall KA (1998). Modeltest: testing the model of DNA substitution. Bioinformatics 14: 817–818. [Abstract] [Google Scholar]
  • Radlkofer L (1863). Ueber die Verunreinigung eines der Münchener Trinkwasser. Kunst- und Gewerbe-Blatt des polytechnischen Vereins für das Königreich Bayern 49: 1–17. [Google Scholar]
  • Ronquist F, Huelsenbeck JP (2003). MRBAYES 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572–1574. [Abstract] [Google Scholar]
  • Rossman AY (1983). The phragmosporous species of Nectria and related genera. Mycological Papers 150: 1–164. [Google Scholar]
  • Rossman AY (1993). Holomorphic hypocrealean fungi: Nectria sensu stricto and teleomorphs of Fusarium. 149–160. In: The fungal holomorph: mitotic, meiotic and pleomorphic speciation in fungal systematic (Reynolds DR, Taylor JW, eds.). CAB International: Wallingford, UK.
  • Rossman AY, Samuels GJ (2005). Towards a single scientific name for species of fungi. Inoculum 56: 1–3. [Google Scholar]
  • Rossman AY, Samuels GJ, Lowen R (1993). Leuconectria clusiae gen. nov. and its anamorph Gliocephalotrichium bulbilium with notes on Pseudonectria. Mycologia 85: 685–704. [Google Scholar]
  • Rossman AY, Samuels GJ, Rogerson CT, Lowen R (1999). Genera of Bionectriaceae, Hypocreaceae and Nectriaceae (Hypocreales, Ascomycetes). Studies in Mycology 42: 1–248. [Google Scholar]
  • Saccardo PA (1883). Sylloge fungorum omnium hucusque cognitorum 2: 1–815 + i–lxix + 1–77. [Google Scholar]
  • Samson RA (1974). Paecilomyces and some allied hyphomycetes. Studies in Mycology 6: 1–119. [Google Scholar]
  • Samuels GJ (1976). A revision of the fungi formerly classified as Nectria subgenus Hyphonectria. Memoirs of the New York Botanical Garden 26(3): 1–126. [Google Scholar]
  • Samuels GJ (1977). Nectria consors and its Volutella conidial state. Mycologia 69: 255–262. [Google Scholar]
  • Samuels GJ, Doi Y, Rogerson CT (1990). Hypocreales. Memoirs of the New York Botanical Garden 59: 6–108. [Google Scholar]
  • Samuels GJ, Hallett IC (1983). Microdochium stoveri and Monographella stoveri, new combinations for Fusarium stoveri and Micronectriella stoveri. Transactions of the British Mycological Society 81: 473–483. [Google Scholar]
  • Samuels GJ, Lu B-S, Chaverri P, Candoussau F, Fournier J, Rossman AY (2009). Cyanonectria, a new genus for Nectria cyanostoma and its Fusarium anamorph. Mycological Progress 8: 49–58. [Google Scholar]
  • Samuels GJ, Rossman AY, Lowen R, Rogerson CT (1991). A synopsis of Nectria subgen. Dialonectria. Mycological Papers 164: 1–48. [Google Scholar]
  • Samuels GJ, Seifert KA (1987). Taxonomic implications of variation among hypocrealean anamorphs. 29–56. In: Pleomorphic fungi: the diversity and its taxonomic implication (Sugiyama J, ed.). Kodansha, Tokyo and Elsevier, Amsterdam.
  • Samuels GJ, Seifert KA (1991). Two new species of Nectria with Stilbella and Mariannaea anamorphs. Sydowia 43: 249–263. [Google Scholar]
  • Schoch C, Crous PW, Wingfield MJ, Wingfield BD (2000). Phylogeny of Calonectria and selected hypocrealean genera with cylindrical macroconidia. Studies in Mycology 45: 45–62. [Google Scholar]
  • Schroers H-J, Gräfenhan T, Nirenberg HI, Seifert KA (2011). A revision of Cyanonectria and Geejayessia gen. nov. and related species with Fusarium-like anamorphs. Studies in Mycology 68: 115–138 (this issue). [Europe PMC free article] [Abstract] [Google Scholar]
  • Schroers H-J, Geldenhuis MM, Wingfield MJ, Schoeman MH, Yen Y-F, Shen W-C, Wingfield BD (2005). Classification of the guava wilt fungus Myxosporium psidii, the palm pathogen Gliocladium vermoesenii and the persimmon wilt fungus Acremonium diospyri in Nalanthamala. Mycologia 97: 375–395. [Abstract] [Google Scholar]
  • Schroers H-J, Lamprecht SC, O'Donnell K, Kammeyer PL, Johnson S, Sutton DA, Rinaldi MG, Summerbell RC (2009). Taxonomy and phylogeny of the Fusarium dimerum species group. Mycologia 101: 44–70. [Abstract] [Google Scholar]
  • Schweinitz LD de (1822). Synopsis fungorum Carolinae superioris. Johann Ambrosis Barth, Leipzip (reprinted by J. Carmer, Vaduz, 1976).
  • Seifert KA (1985a). A monograph of Stilbella and allied hyphomycetes. Studies in Mycology 27: 1–235. [Google Scholar]
  • Seifert KA (1985b). Notes on some apocryphal genera of synnematal Hyphomycetes. Transactions of the British Mycological Society 85: 123–133. [Google Scholar]
  • Seifert KA (2001). Fusarium and anamorph generic concepts. 15–28. In: Summerell BA, Leslie JF, Backhouse D, Bryden WL, Burgess LW (eds): Fusarium: Paul E. Nelson Memorial Symposium. American Phytopathological Society, St. Paul, Minnesota.
  • Seifert KA, Samuels GJ (2000). How should we look at anamorphs? Studies in Mycology 45: 5–18. [Google Scholar]
  • Seifert KA, Stanley SJ, Hyde KD (1995). Stilbella holubovae, a new synnematous hyphomycete species on driftwood from the Philippines and South Africa. Sydowia 47: 258–262. [Google Scholar]
  • Sivichai S, Jones EBG, Hywel-Jones N (2002). Fungal colonisation of wood in a freshwater stream at Tad Ta Phu, Khao Yai National Park, Thailand. Fungal Diversity 10: 113–129. [Google Scholar]
  • Sukapure RS, Thirumalachar MJ (1966). Conspectus of species of Cephalosporium with particular reference to Indian species. Mycologia 58: 351–361. [Abstract] [Google Scholar]
  • Sukumaran J, Holder MT (2010). DendroPy: a Python library for phylogenetic computing. Bioinformatics 26: 1569–1571. [Abstract] [Google Scholar]
  • Summerbell RC, Gueidan C, Schroers H-J, Hoog GS de, Starink M, Iperen A van, Arocha Rosete Y, Guarro J, Scott JA (2011). Acremonium phylogenetic overview and revision of Gliomastix, Sarocladium, and Trichothecium. Studies in Mycology 68: 139–162 (this issue). [Europe PMC free article] [Abstract] [Google Scholar]
  • Summerbell RC, Schroers H-J (2002). Analysis of phylogenetic relationship of Cylindrocarpon lichenicola and Acremonium falciforme to the Fusarium solani species complex and a review of similarities in the spectrum of opportunistic infections caused by these fungi. Journal of Clinical Microbiology 40: 2866–2875. [Europe PMC free article] [Abstract] [Google Scholar]
  • Sutton BC (1977). Coelomycetes VI. Nomenclature of generic names proposed for coelomycetes. Mycological Papers 141: 1–253. [Google Scholar]
  • Sutton BC (1986). Improvizations on conidial themes. Transactions of the British Mycological Society 86: 1–38. [Google Scholar]
  • Swofford DL (2003). PAUP*. Phylogenetic Analysis Using Parsimony (*and Other Methods). Version 4. Sinauer Associates, Sunderland, Massachusetts.
  • Torzilli AP, Balakrishna S, O'Donnell K, Lawrey J (2002). The degradative activity of a lichenicolous Fusarium sp. compared to entomogenous species. Mycological Research 106: 1204–1210. [Google Scholar]
  • Trabut L (1907). Un champignon parasite de la cochenille des orangers. Bulletin agricole de l'Algerie et de la Tunisie 13: 32–108. [Google Scholar]
  • Tubaki K, Booth C, Harada T (1976). A new variety of Fusarium merismoides. Transactions of the British Mycological Society 66: 355–356. [Google Scholar]
  • Tulasne ELR, Tulasne C (1861). Selecta Fungorum Carpologia 1: 129–130. [Google Scholar]
  • Tulasne ELR, Tulasne C (1865). Selecta Fungorum Carpologia 3: 104–106. [Google Scholar]
  • Weese J (1916). Beiträge zur Kenntnis der Hypocreaceen. Sitzungsberichte der Kaiserlichen Akademie der Wissenschaften. Mathematisch-Naturwissenschaftliche Klasse, Abt. 1, 125: 465–575. [Google Scholar]
  • Wollenweber HW (19301924 1916). Fusaria Autographice Delineata. Collectio Specierum ex Herbariis Variis Selectarum et ab Auctore Lectarum Cultarumque Synonymis et Excludendis Additis quas Determinavit, in Sectiones Digessit, Comparavit cum Hypocreaceis Analogis Praemissis ad Methodi Naturalis Normas et Culturae Purae Experientiam H.W. Wollenweber 1: i, 1–509; 2: i, 510–659; 3: i, 660–1100. Germany, Berlin. [Google Scholar]
  • Wollenweber HW (1926). Pyrenomyceten-Studien. II. Angewandte Botanik 8: 168–212. [Google Scholar]
  • Wollenweber HW (1931). Fusarium-Monographie. Fungi parasitici et saprophytici. Zeitschrift für Parasitenkunde 3: 269–516. [Google Scholar]
  • Wollenweber HW, Reinking OA (1935). Die Fusarien: ihre Beschreibung, Schadwirkung und Bekämpfung. 1–355. Germany, Berlin.
  • Yang ZH, Nielsen R, Goldman N, Pedersen A-MK (2000). Codon-substitution models for heterogeneous selection pressure at amino acid sites. Genetics 155: 431–449. [Europe PMC free article] [Abstract] [Google Scholar]
  • Zhang X-M, Zhuang W-Y (2006). Phylogeny of some genera in the Nectriaceae (Hypocreales, Acomycetes) inferred from 28S nrDNA partial sequences. Mycosystema 25: 15–22. [Google Scholar]
  • Zhdanova NN (1966). Rare and new species of Dematiaceae isolated from maize rhizosphere of various climatic belts of the Ukrainian SSR. Mikrobiologicheskii Zhurnal, Kiev 28: 36–40. [in Russian] [Abstract] [Google Scholar]
  • Zwickl DJ (2006). Genetic algorithm approaches for the phylogenetic analysis of large biological sequence data sets under the maximum likelihood criterion. Ph.D. dissertation, The University of Texas at Austin.

Articles from Studies in Mycology are provided here courtesy of Westerdijk Fungal Biodiversity Institute

Citations & impact 


Impact metrics

Jump to Citations
Jump to Data

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/3660436
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/3660436

Article citations


Go to all (100) article citations

Other citations

Data 


Data behind the article

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

Similar Articles 


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