ArticlePDF Available

Epitypification of Fusarium oxysporum – clearing the taxonomic chaos

Authors:

Abstract and Figures

Fusarium oxysporum is the most economically important and commonly encountered species of Fusarium. This soil-borne fungus is known to harbour both pathogenic (plant, animal and human) and non-pathogenic strains. However, in its current concept F. oxysporum is a species complex consisting of numerous cryptic species. Identification and naming these cryptic species is complicated by multiple subspecific classification systems and the lack of living ex-type material to serve as basic reference point for phylogenetic inference. Therefore, to advance and stabilise the taxonomic position of F. oxysporum as a species and allow naming of the multiple cryptic species recognised in this species complex, an epitype is designated for F. oxysporum. Using multi-locus phylogenetic inference and subtle morphological differences with the newly established epitype of F. oxysporum as reference point, 15 cryptic taxa are resolved in this study and described as species.
Content may be subject to copyright.
© 2018-2019 Naturalis Biodiversity Center & Westerdijk Fungal Biodiversity Institute
You are free to share - to copy, distribute and transmit the work, under the following conditions:
Attribution: You must attribute the work in the manner specified by the author or licensor (but not in any way that suggests that they endorse you or your use of the work).
Non-commercial: You may not use this work for commercial purposes.
No derivative works: You may not alter, transform, or build upon this work.
For any reuse or distribution, you must make clear to others the license terms of this work, which can be found at http://creativecommons.org/licenses/by-nc-nd/3.0/legalcode. Any of the above conditions can be
waived if you get permission from the copyright holder. Nothing in this license impairs or restricts the author’s moral rights.
Persoonia 43, 2019: 1– 47 ISSN (Online) 1878-9080
www.ingentaconnect.com/content/nhn/pimj https://doi.org/10.3767/persoonia.2019.43.01
RESEARCH ARTICLE
INTRODUCTION
Fusarium oxysporum is the most economically important and
commonly encountered species of Fusarium. This soil-borne
asexual fungus is known to harbour both pathogenic (plant,
animal and human) and non-pathogenic strains (Leslie & Sum-
merell 2006) and is also ranked fifth on a list of top 10 fungal
pathogens based on scientific and economic importance (Dean
et al. 2012, Geiser et al. 2013). Historically, F. oxysporum has
been defined by the asexual phenotype as no sexual morph
has yet been discovered, even though several studies have
indicated the possible presence of a cryptic sexual cycle (Arie
et al. 2000, Yun et al. 2000, Aoki et al. 2014, Gordon 2017).
This is further supported by phylogenetic studies that place
F. oxysporum within the Gibberella Clade (Baayen et al. 2000,
O’Donnell et al. 2009, 2013). These studies also showed that
F. oxysporum displays a complicated phylogenetic substruc-
ture, indicative of multiple cryptic species within F. oxysporum
(Gordon & Martyn 1997, Laurence et al. 2014). As with other
Fusarium species complexes, the F. oxysporum species com-
plex (FOSC) has suffered from multiple taxonomic/ classification
systems applied in the past.
Diederich F.L. von Schlechtendal first introduced F. oxysporum
in 1824, isolated from a rotten potato tuber (Solanum tubero-
sum) collected in Berlin, Germany. Wollenweber (1913) placed
F. oxysporum within the section Elegans along with eight other
Fusarium species and numerous varieties and forms based
on similarity of the micro- and macroconidial morphology and
dimensions. Snyder & Hansen (1940) later consolidated and
reduced all species within the section Elegans into F. oxysporum
and designated 25 special forms (formae speciales) within this
species. These special forms were further expanded on by Gor-
don (1965) to 66, most of which are still used in literature today.
The use of special forms or formae speciales as subspecific
rank in F. oxysporum classification has become common prac-
tice due to the broad morphological delineation of this species
(Leslie & Summerell 2006). This informal subspecific rank is
defined based on the plant pathogenicity of the particular F. oxy-
sporum strain and excludes both clinical and non-pathogenic
strains (Armstrong & Armstrong 1981, Gordon & Martyn 1997,
Kistler 1997, Baayen et al. 2000, Leslie & Summerell 2006).
Therefore, F. oxysporum strains attacking the same plant host
are generally considered to belong to the same special form.
Although this homologous trait has led to erroneous assump-
tions considering a specific special form to be phylogenetically
monophyletic, several studies (O’Donnell et al. 1998, 2004,
2009, O’Donnell & Cigelnik 1999, Baayen et al. 2000, Lievens
et al. 2009b, Van Dam et al. 2016) have highlighted the para-
and polyphyletic relationships within several F. oxysporum
special forms, e.g., F. oxysporum f. sp. batatas, F. oxysporum
f. sp. cubense and F. oxysporum f. sp. vasinfectum. Addition-
ally, several F. oxysporum special forms are able to infect and
cause disease in more than one (sometimes unrelated) plant
hosts, whereas others are highly specialised to a specific plant
host (Armstrong & Armstrong 1981, Gordon & Martyn 1997,
Kistler 1997, Baayen et al. 2000, Leslie & Summerell 2006,
Fourie et al. 2011).
Naming F. oxysporum special forms are not subject to the Inter-
national Code of Nomenclature for algae, fungi, and plants (ICN;
McNeill et al. 2012, Thurland et al. 2018), and therefore no
diagnosis (in Latin and/or English), nor the deposit of type ma-
terial in a recognised repository is required. This decision was
made due to the difficulty in accepting special forms within the
Code, even though these strains are of great importance to plant
pathologists and breeders (Deighton et al. 1962, Gordon 1965,
Armstrong & Armstrong 1981). Several studies on F. oxysporum
indicate that between 70 to over 150 special forms are known
in F. oxysporum (Booth 1971, Armstrong & Armstrong 1981,
Kistler 1997, Baayen et al. 2000, Leslie & Summerell 2006,
Lievens et al. 2008, O’Donnell et al. 2009, Fourie et al. 2011,
Epitypification of Fusarium oxysporum
clearing the taxonomic chaos
L. Lombard1, M. Sandoval-Denis1,2, S.C. Lamprecht3, P.W. Crous1,2,4
1 Westerdijk Fungal Biodiversity Institute, Uppsalalaan 8, 3584 CT Utrecht,
The Netherlands;
corresponding author e-mail: l.lombard@westerdijkinstitute.nl.
2 Faculty of Natural and Agricultural Sciences, Department of Plant Sciences,
University of the Free State, P.O. Box 339, Bloemfontein 9300, South Africa.
3 ARC-Plant Health and Protection, Private Bag X5017, Stellenbosch, 7599,
Western Cape, South Africa.
4 Wageningen University and Research Centre (WUR), Laboratory of Phyto-
pathology, Droevendaalsesteeg 1, 6708 PB Wageningen, The Netherlands.
Key words
cryptic species
diversity
human and plant pathogens
species complex
subspecific classification
Abstract Fusarium oxysporum is the most economically important and commonly encountered species of Fusa-
rium. This soil-borne fungus is known to harbour both pathogenic (plant, animal and human) and non-pathogenic
strains. However, in its current concept F. oxysporum is a species complex consisting of numerous cryptic species.
Identification and naming these cryptic species is complicated by multiple subspecific classification systems and the
lack of living ex-type material to serve as basic reference point for phylogenetic inference. Therefore, to advance
and stabilise the taxonomic position of F. oxysporum as a species and allow naming of the multiple cryptic species
recognised in this species complex, an epitype is designated for F. oxysporum. Using multi-locus phylogenetic
inference and subtle morphological differences with the newly established epitype of F. oxysporum as reference
point, 15 cryptic taxa are resolved in this study and described as species.
Article info Received: 20 June 2018; Accepted: 19 October 2018; Published: 18 December 2018.
2Persoonia – Volume 43, 2019
Laurence et al. 2014, Gordon 2017). At present Index Fungo-
rum (http:// www.indexfungorum.org/) lists 124 special forms in
F. oxysporum, whereas MycoBank (http: //www.mycobank.org/)
list 127 special forms. Further careful scrutiny of literature re-
vealed that 144 special forms have been named until February
2018 (Table 1). Although the special forms concept of Snyder
& Hansen (1940) is still applied today, additional subspecific
classification systems for special forms of F. oxysporum have
also been introduced, which include haplotypes, races and
vegetative compatibility groups (VCGs).
The haplotype subspecific classification system was introduced
by Chang et al. (2006) and later expanded upon by O’Donnell
et al. (2008, 2009) to include strains from both the FOSC
and Neocosmospora (formerly the F. solani (FSSC) species
complex). This classification system is based on unique multi-
locus genotypes within the species complex, aimed to resolve
communication problems among public health and agricultural
scientists (O’Donnell et al. 2008). Chang et al. (2006) proposed
a standardised haplotype nomenclature system that depict
the species complex, species and genotype. O’Donnell et al.
(2009) was able to identify 256 unique two-locus haplotypes
from 850 isolates representing 68 special forms of F. oxysporum
as well as environmental and clinical strains. However, this
classification system is not in common use as a reference, and
a continuously updated database is required.
One of the most important subspecific ranks applied to special
forms of F. oxysporum are physiological pathotypes or races.
This classification system is of great importance to plant breed-
ers, especially for resistance breeding. Traditionally, race demar-
cation is based on cultivar specificity linked to specific resistance
genes of the plant host cultivar (Armstrong & Armstrong 1981,
Kistler 1997, Baayen et al. 2000, Roebroeck 2000, Fourie et al.
2011, Epstein et al. 2017). However, race designation has been
inconsistent in the past (Gerlagh & Blok 1988, Correll 1991,
Kistler 1997, Fourie et al. 2011) with several different nomen-
clatural systems being applied (Gabe 1975, Risser et al. 1976,
Armstrong & Armstrong 1981) to further cause confusion (Kistler
1997). With advances in molecular technology, identification
of races has been simplified using sequence-characterised
amplified region (SCAR) primers (Lievens et al 2008, Epstein
et al. 2017, Gilardi et al. 2017). However, time consuming and
laborious pathogenicity tests are still needed to identify new
emerging races and to test whether newly developed plant
cultivars are resistant to known races (Epstein et al. 2017,
Gilardi et al. 2017).
The use of vegetative compatibility (also known as heterokar-
yon compatibility) has formed an integral part of subspecific
classification of F. oxysporum special forms and non-pathogenic
strains. Formation of a stable heterokaryon between two auxo-
trophic nutritional mutants is regulated by several vic or het
incompatibility loci (Correll 1991, Leslie 1993) indicating that
the strains are homogenic at these loci (Correll 1991) and con-
sidered to be part of the same VCG. Therefore, classification
using vegetative compatibility is based on genetic similarity at
specific loci and not pathogenicity, providing a crude marker
for population genetic studies (Correll 1991, Gordon & Martyn
1997, Leslie 1993, Leslie & Summerell 2006). Puhalla (1985),
utilizing nit mutants, was the first to identify VCGs in F. oxyspo-
rum and characterised 16 VCGs in a collection of 21 F. oxyspo-
rum strains. The numbering system applied by Puhalla (1985),
which is still used today, consists of a three-digit numerical
code indicating the special form followed by digit(s) indicating
the VCG (Katan 1999, Katan & Di Primo 1999). Conventional
VCG characterisation is a relatively objective, time consuming
and laborious assay only indicating genetic similarity and not
genetic difference (Kistler 1997). Therefore, several PCR-based
detection methods have been developed to identify economi-
cally important VCGs as diagnostic tool (Fernandez et al. 1998,
Pasquali et al. 2004a, c, Lievens et al. 2008), e.g., F. oxysporum
f. sp. cubense TR4 VCG01213 (Dita et al. 2010).
Until recently, limited knowledge on the genetic premise for host
specificity in F. oxysporum was available (Gordon & Martyn
1997, Kistler 1997, Baayen et al. 2000). However, the discovery
of a lineage-specific chromosome (or transposable/effector/
accessory chromosome) in F. oxysporum f. sp. lycopersici by
Ma et al. (2010), in which the host specific virulence genes lie
(Van der Does et al. 2008, Takken & Rep 2010, Ma et al. 2013),
has provided a new view into the evolution of pathogenicity in
F. oxysporum. In vitro transfer of these accessory chromosomes
into non-pathogenic F. oxysporum strains has converted the
latter strains into host-specific pathogens, providing evidence
that host-specific pathogenicity could be acquired through
horizontal transfer of accessory chromosomes (Takken & Rep
2010, Ma et al. 2010, 2013, Van Dam et al. 2016, Van Dam &
Rep 2017). Therefore, the special form name can be linked to
the accessory chromosome whereas race demarcation can be
linked to the specific virulence genes carried on these acces-
sory chromosomes.
The genetic and functional mechanisms of the infection process
in plants of various special forms of F. oxysporum has been
well documented (Di Pietro et al. 2003, Ma et al. 2013, Upasani
et al. 2016, Gordon 2017). However, these same mechanisms
are still poorly understood in human and animal infections
(O’Donnell et al. 2004, Guarro 2013, Van Diepeningen et al.
2015). Fusarium oxysporum has been linked to fungal keratitis
(Hemo et al. 1989, Chang et al. 2006) and dermatitis (Guarro
& Gene 1995, Romano et al. 1998, Ninet et al. 2005, Cutuli et
al. 2015, Van Diepeningen et al. 2015), and has been isolated
from contaminated hospital water systems (Steinberg et al.
2015, Edel-Hermann et al. 2016) and medical equipment (Bar-
ton et al. 2016, Carlesse et al. 2017) posing a serious threat
to immunocompromised patients. Several recent reports also
indicate that F. oxysporum is able to infect immunocompetent
patients (Jiang et al. 2016, Khetan et al. 2018). In general,
fusariosis is difficult to treat as Fusarium species display a
remarkable resistance to antifungal agents (Guarro 2013, Al-
Hatmi et al. 2018). However, some antimycotics are known to
be effective against F. oxysporum related fusariosis (Al-Hatmi
et al. 2018). Recently, both mycotoxins beauvericin and fusaric
acid, produced by F. oxysporum strains that can infect tomato,
have been shown to be important virulence determinants to
infect immunosuppressed mice (López-Berges et al. 2013,
López-Díaz et al. 2018).
Strains of F. oxysporum are known to produce a cocktail of
polyketide secondary metabolites, some with unknown function
and toxicities (Marasas et al. 1984, Mirocha et al. 1989, Bell
et al. 2003, Desjardins 2006, Manici et al. 2017). Some of the
better-known toxins produced by F. oxysporum include beau-
vericin (Marasas et al. 1984, Logrieco et al. 1998, López-Berges
et al. 2013), fusaric acid (Marasas et al. 1984, López-Díaz et
al. 2018) and fumonisins (Rheeder et al. 2002) to name a few.
Mycotoxicological studies on F. oxysporum has thus far only
focused on a strain to strain basis and therefore no link has
yet been established between special form and/or race and
mycotoxin production capabilities.
In light of the complicated and sometimes confusing classifi-
cation systems applied to F. oxysporum taxonomy and nomen-
clature, the question has risen whether F. oxysporum truly
represent a species (Kistler 1997). Given that F. oxysporum is
a common, widespread, soil-borne fungus, with a global distri-
bution and high economic importance, this question requires
urgent attention. Therefore, to advance and stabilize the taxo-
nomic and nomenclatural position of F. oxysporum and allow
3
L. Lombard et al.: Epitypification of Fusarium oxysporum
adzukicola Kitazawa & Yanagita 1984, 1989 Summerell et al. 2010 Katan & Di Primo 1999
aechmeae Sauthoff & Gerlach 1957, 1958 Fusarium bulbigenum f. aechmeae Gordon 1965, Armstrong & Gherbawy 1999, O’Donnell et al. 2009
Sauthoff & Gerlach, Gratenwelt 57: Armstrong 1968, 1981, Booth 1971,
390. 1957 Summerell et al. 2010
albedinis Sergent & Beguet 1921, Killian & Maire 1930, Cylindrophora albedinis Kill. & Maire, Gordon 1965, Armstrong & Tantaoui et al. 1996, Kistler Tantaoui & Boisson 1991, Tantaoui & Fernandez 1993
Malençon 1934, Louvet & Toutain 1981 Bull. Soc. Hist. Nat. Afrique N. 21: Armstrong 1968, 1981, Booth 1971, et al. 1998, Katan 1999 Tantaoui et al. 1996, Fernandez et al. 1994, 1998,
89 –101. 1930 Summerell et al. 2010 Skovgaard et al. 2001, Mbofung et al. 2007, Lievens
Fusarium albedinis (Kill. & Maire) et al. 2008, O’Donnell et al. 2009, Elliott et al. 2010,
Malençon, Compt. Rend. Acad. Sci. 198: Mirtalebi & Banihashemi 2014
1259–1261. 1930
Fusarium oxysporum var. albedinis
(Kill. & Maire) Malençon, Rev. Mycol.
(Paris) 15: 45– 60. 1950
aleuritis Suelong 1981 Suelong 1981
allii Matuo et al. 1979 Yoo et al. 1993, Katan & Di O’Donnell et al. 2009
Primo 1999
amaranthi Chen & Swart 2001 Summerell et al. 2010 Chen & Swart 2001 Chen & Swart 2001
anethi Janson 1951, Gordon 1965 Gordon 1965, Armstrong &
Armstrong 1968, 1981, Booth 1971,
Summerell et al. 2010
anoectochili Huang et al. 2014 Huang et al. 2014 Huang et al. 2014 Huang et al. 2014
apii Snyder & Hansen 1940 Fusarium apii P.E. Nelson & Sherb., Snyder & Hansen 1940, Gordon Schneider & Norelli 1981, Puhalla Puhalla 1984a, b, Correll Wang et al. 2001, O’Donnell et al. 2009,
Tech. Bull. Mich. Agric. Exp. Sta. 155: 1965, Armstrong & Armstrong 1984a, b, Epstein et al. 2017 et al. 1986, 1987, Toth & Chakrabarti et al. 2011, Epstein et al. 2017
42. 1937 1968, 1981, Booth 1971, Lacy 1991, Kistler et al.
Fusarium oxysporum f. apii (P.E. Summerell et al. 2010 1998, Katan 1999
Nelson & Sherb.) W.C. Snyder & H.N.
Hansen, Amer. J. Bot. 27: 66. 1940
Fusarium bulbigenum var. apii (P.E.
Nelson & Sherb.) Raillo, Fungi of the
genus Fusarium: 250. 1950
Fusarium apii var. pallidum P.E. Nelson &
Sherb., Tech. Bull. Mich. Agric. Exp.
Sta. 155: 42. 1937
arctii Matuo et al. 1975 Summerell et al. 2010 O’Donnell et al. 2009
asparagi Cohen 1946 Gordon 1965, Armstrong & Blok & Bollen 1997, Elmer Baayen et al. 2000, Mbofung et al. 2007,
Armstrong 1968, 1981, Booth 1971, & Stephens 1989, Yoo et al. O’Donnell et al. 2009, Poli et al. 2012, Mirtalebi &
Summerell et al. 2010 1993, Kistler et al. 1998, Banihashemi 2014
Katan 1999, Katan & Di
Primo 1999
basilica Dzidzariya 1968, Armstrong & Armstrong 1981 Fusarium oxysporum var. basilicum Armstrong & Armstrong 1968, 1981, Elmer et al. 1994, Kistler Chiocchetti et al. 1999, 2001, Pasquali et al. 2006,
Dzidzariya, Pishch. Prom. SSR: Summerell et al. 2010 et al. 1998, Katan 1999, Lievens et al. 2008, O’Donnell et al. 2009
129–140. 1968 Katan & Di Primo 1999
batatas Wollenweber 1914, 1931 Fusarium batatas Wollenw., J. Agric. Snyder & Hansen 1940, Armstrong & Armstrong 1958b, 1968, Katan 1999, Katan & Di O’ Donnell et al. 1998, Kim et al. 2001, Mbofung
Res. 2: 268. 1914 Gordon 1965, Armstrong & Booth 1971 Primo 1999 et al. 2007, Lievens et al. 2009b, O’Donnell et al.
Fusarium bulbigenum var. batatas Armstrong 1968, 1981, Booth 1971, 2009, Pinaria et al. 2015
(Wollenw.) Wollenw., Z. Parasitenk. Summerell et al. 2010
(Berlin) 3: 414. 1931
Fusarium oxysporum f. batatas
(Wollenw.) W.C. Snyder & H.N.
Hansen, Amer. J. Bot. 27: 66. 1940
benincasae Gerlagh & Ester 1985 Gerlagh & Blok 1988
betae Stewart 1931 Fusarium conglutinans var. betae Snyder & Hansen 1940, Gordon Armstrong & Armstrong 1976 Harveson & Rush 1997, Cramer et al. 2003, Nitschke et al. 2009, O’Donnell
D. Stewart, Phytopathology 9: 59. 1931 1965, Armstrong & Armstrong 1968, Kistler et al. 1998, et al. 2009, Hill et al. 2011, Covey et al. 2014
Fusarium orthoceras var. betae 1981, Booth 1971, Summerell Webb et al. 2013
Table 1 List of known special forms of Fusarium oxysporum.
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
4Persoonia – Volume 43, 2019
betae (cont.) (D. Stewart) Padwick, Indian J. Agric. et al. 2010
Sci. 10: 282. 1940
Fusarium oxysporum f. betae
(D. Stewart) W.C. Snyder & H.N.
Hansen, Amer. J. Bot. 27: 66. 1940
Fusarium oxysporum var. orthoceras
(Appel & Wollenw.) Bilaǐ, The Fusaria:
282. 1955
bouvardiae Marziano et al. 1987 O’Donnell et al. 2009
brassica Williams et al. 2016 Williams et al. 2016
callistephi Beach 1918 Fusarium conglutinans var. callistephi Snyder & Hansen 1940, Gordon Armstrong & Armstrong 1971 Mbofung et al. 2007, O’Donnell et al. 2009, Poli et al.
Beach, Rep. Michigan Acad. Sci. 1965, Armstrong & Armstrong 1968, 2012
29: 297. 1918 1981, Booth 1971, Summerell
Fusarium orthoceras var. callistephi et al. 2010
(Beach) Padwick, Indian J. Agric. Sci.
10: 283. 1940
Fusarium oxysporum f. callistephi
(Beach) W.C. Snyder & H.N. Hansen,
Amer. J. Bot. 27: 66. 1940
Fusarium conglutinans var. majus
Wollenw., Fusaria Autographica
Delineata 3: 981. 1930
canariensis Mercier & Louvet 1973, Feather et al. 1979 Summerell et al. 2010 Katan 1999, Pyler et al. Pyler et al 2000, Gunn & Summerell 2002, Mbofung
2000, Gunn & Summerell et al. 2007, Lievens et al. 2009b, Elliott et al. 2010,
2002 Laurence et al. 2015, Pinaria et al. 2015
cannabis Noviello & Snyder 1962 Gordon 1965, Armstrong & Armstrong O’Donnell et al. 2009
1968, 1981 Booth 1971
capsici Black et al. 1993
carthami Klisiewicz & Houston 1963 Gordon 1965, Armstrong & Arm- Klisiewicz & Thomas 1970a, b, Shende et al. 2015
strong 1968, 1981, Booth 1971, Klisiewicz 1975
Summerell et al. 2010
cassiae Armstrong 1954, Gordon 1965 Gordon 1965, Armstrong & Arm- O’Donnell et al. 2009
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
cattleyae Foster 1955 Gordon 1965, Armstrong & Arm- Baayen & Kleijn 1989 O’Donnell et al. 2009
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
cepae Hanzawa 1914 Fusarium cepae Hanzawa, Snyder & Hansen 1940, Gordon Molnár et al. 1990, Yoo et al. Gherbawy 1999, Mbofung et al. 2007, Galván et al.
Mykol. Zentbl. 5: 5. 1914 1965, Armstrong & Armstrong 1968, 1993, Katan & Di Primo 2008, O’Donnell et al. 2009, Bayraktar et al. 2010,
Fusarium oxysporum f. cepae 1981, Booth 1971, Summerell 1999, Swift et al. 2002, Lin et al. 2010, Southwood et al. 2012, Mirtalebi &
(Hanzawa) W.C. Snyder & H.N. et al. 2010 Widodo et al. 2008, Banihashemi 2014, Taylor et al. 2016
Hansen, Amer. J. Bot. 27: 66. 1940 Bayraktar et al. 2010,
Fusarium oxysporum var. cepae Southwood et al. 2012
(Hanzawa) Raillo, Fungi of the genus
Fusarium: 253. 1950
chrysanthemi Armstrong et al. 1970 Armstrong & Armstrong 1968, Huang et al. 1992, Troisi et al. 2013 Puhalla 1985, Correll et al. Kim et al. 2001, Pasquali et al. 2003, 2004a, b, c,
1981, Booth 1971, 1987, Kistler et al. 1998, Bogale et al. 2007, Lievens et al. 2008, O’Donnell
Summerell et al. 2010 Katan 1999, Pasquali et al. et al. 2009, Li et al. 2010, Lin et al. 2010, Troisi et al.
2004c 2010, 2013
ciceris Padwick 1940, Erwin 1958, Matuo & Sato 1962 Fusarium orthoceras var. ciceri Armstrong & Armstrong 1968, Haware & Nene 1982, Barve et al. Kistler et al. 1998 Kelly et al. 1994, 1998, García- Pedrasjas et al. 1999,
Padwick, Indian J. Agr. Sci. 10: 1981, Booth 1971 2001, Jiménez-Gasco et al. 2001, Barve et al. 2001, Jiménez-Gasco et al. 2001, 2002,
241–284. 1940 2004a, b, Jiménez-Gasco & 2004a, b, Jiménez-Gasco & Jiménez-Díaz 2003,
Fusarium lateritium f. ciceri (Padwick) Jiménez-Díaz 2003, Sharma et al. 2004, 2014, 2016, Honnareddy & Dubey
Erwin, Phytopathology 48: 500. 1958 Sharma et al. 2004, Honnareddy & 2006, Bayraktar et al. 2008, Dubey & Singh 2008,
Dubey 2006, Gurjar et al. 2009, Gurjar et al. 2009, Dubey et al. 2012, Demers et al.
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
5
L. Lombard et al.: Epitypification of Fusarium oxysporum
ciceris (cont.) Dubey et al. 2012, Demers et al. 2014, 2014, Ghosh et al. 2015, Upasani et al. 2016, Williams
Upasani et al. 2016 et al. 2016
cichorii Poli et al. 2012 Poli et al. 2012
citri Timmer et al. 1979, Timmer 1982 Hannachi et al. 2015
coffeae Alvarez 1945, Wellman 1954 Fusarium bulbigenum var. coffeae Gordon 1965, Armstrong & Armstrong
Álv. García, J. Agric. Univ. Puerto 1968, 1981, Booth 1971, Summerell
Rico 29: 8. 1945 et al. 2010
colocasiae Nishimura & Kudo 1994 Hirano & Arie 2009, Poli et al. 2013
conglutinans Wollenweber 1913, Padwick 1940 Fusarium conglutinans Wollenw., Snyder & Hansen 1940, Gordon Ramirez-Villupadua et al. 1985, Puhalla 1985, Bosland & Bosland & Williams 1987, Kistler et al. 1987, Kistler
Phytopathology 3 (1): 30. 1913 1965, Armstrong & Armstrong 1968, Armstrong & Armstrong 1952, Williams 1987, Correll et al. & Benny 1989, Crowhurst et al. 1995, Gherbawy
Fusarium orthoceras var. conglutinans 1981, Booth 1971, Summerell 1953, 1966 1987, Correll 1991, Kistler 1999, Kim et al. 2001, Bogale et al. 2007, Hirano &
(Wollenw.) Padwick, Indian J. Agric. et al. 2010 et al. 1998, Katan 1999, Arie 2009, O’Donnell et al. 2009, Srinivasan et al.
Sci. 10: 282. 1940 Katan & Di Primo 1999 2010, Poli et al. 2012, Covey et al. 2014, Zang et al.
Fusarium oxysporum f. conglutinans 2014, Hansen et al. 2015, Kashiwa et al. 2016, Li et
(Wollenw.) W.C. Snyder & H.N. Hansen, al. 2015, 2016, Taylor et al. 2016, Van Dam & Rep
Amer. J.Bot. 27: 66. 1940 2017
coriandrii Booth 1971, Armstrong & Armstrong 1981 Armstrong & Armstrong 1968, 1981,
Booth 1971, Summerell et al. 2010
crassulae Ortu et al. 2013 Ortu et al. 2013
croci Boerema & Hamers 1989 Roebroeck 2000 Roebroeck 2000 Roebroeck 2000, Palmero et al. 2014
crotalariae Kulkarni 1934, Gupta 1974 Fusarium vasinfectum var. crotalariae Armstrong & Armstrong 1968, 1981
Kulk., Indian J. Agric. Sci 4: 994. 1934
Fusarium udum f.sp. crotalariae (Kulk.)
Subram., The genus Fusarium: 114. 1971
cubense Smith 1910, Brandes 1919 Fusarium cubense E.F. Sm., Science, Snyder & Hansen 1940, Gordon See review by Fourie et al. 2011 See review by Fourie et al. See review by Fourie et al. 2011, Ploetz 2015 and
N.S. 31: 755. 1910 1965, Armstrong & Armstrong and Ploetz 2015 2011 and Ploetz 2015, Lin & Shen 2017, Mostert et al. 2017, Aguayo et al.
Fusarium cubense var. inodoratum E.W. 1968, 1981, Booth 1971, Mostert et al. 2017 2017, Van Dam & Rep 2017, Czislowski et al. 2017
Brandes, Phytopathology 9: 374. 1919 Summerell et al. 2010
Fusarium oxysporum var. cubense
(E.F. Sm.) Wollenw., Die Fusarien,
ihre Beschreibung, Schadwirkung und
Bekämpfung: 119. 1935
Fusarium oxysporum f. cubense
(E.F. Sm.) W.C. Snyder & H.N. Hansen,
Amer. J. Bot. 27: 66. 1940
cucumerinum Owen 1956 Gordon 1965, Armstrong & Arm- Armstrong & Armstrong 1978b, Ahn et al. 1998, Kistler et al. Namiki et al. 1994, Vakalounakis & Fragkiadakis 1999,
strong 1968, 1981, Booth 1971, Armstrong et al. 1978, Gerlagh & 1998, Katan 1999, Kim et al. 2001, Skovgaard et al. 2001, Wang et al.
Summerell et al. 2010 Blok 1988 Katan & Di Primo 1999, 2001, Vakalounakis et al. 2004, Lievens et al. 2007,
Vakalounakis & Fragkiadakis 2008, Hirano & Arie 2009, O’Donnell et al. 2009, Lin
1999, Vakalounakis et al. et al. 2010, Poli et al. 2013, Scarlett et al. 2013,
2004 Mirtalebi & Banihashemi 2014, Bertoldo et al. 2015
cucurbitacearum Gerlagh & Blok 1988 Gerlagh & Blok 1988 Bogale et al. 2007, O’Donnell et al. 2009, Bennett
et al. 2013
cumini Patel et al. 1957 Summerell et al. 2010 Talaviya et al. 2014, Nawade et al. 2017
cyclaminis Gerlach 1954 Gordon 1965, Armstrong & Arm- Woudt et al. 1995, Kistler Woudt et al. 1995, Gherbawy 1999, Kim et al. 2001,
strong 1968, 1981, Booth 1971, et al. 1998, Katan 1999, O’Donnell et al. 2009, Lecomte et al. 2016
Summerell et al. 2010 Lori et al. 2012
dahliae Summerell et al. 2010 Summerell et al. 2010
delphinii Laskaris 1949 Gordon 1965, Armstrong & Arm- Kondo et al. 2013
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
dianthi Snyder & Hansen 1940 Fusarium dianthi Prill. & Delacr., Compt. Snyder & Hansen 1940, Gordon Hood & Stewart 1957, Garibaldi Puhalla 1985, Correll et al. Manicom et al. 1990, Manicom & Baayen 1993,
Rend. Acad. Sci.: 744–745. 1899 1965, Armstrong & Armstrong 1968, 1975, 1977, 1983, Baayen et al. 1988, 1987, Hadar et al. 1989, Manulis et al. 1994, Crowhurst et al. 1995, Baayen
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
6Persoonia – Volume 43, 2019
dianthi (cont.) Fusarium oxysporum f. dianthi (Prill. & 1981, Booth 1971, Summerell Aloi & Baayen 1993, Summerell et al. Molnár et al. 1990, Manicom et al. 1997, 2000, Gherbawy 1999, Kim et al. 2001,
Delacr.) W.C. Snyder & H.N. Hansen, et al. 2010 2010 et al. 1990, Aloi & Baayen Skovgaard et al. 2001, Bogale et al. 2007, Lievens
Amer. J. Bot. 27: 66. 1940 1993, Baayen et al. 1997, et al. 2008, Hirano & Arie 2009, O’Donnell et al. 2009,
Fusarium oxysporum f. sp. barbati Kistler et al. 1998, Katan Poli et al. 2013, Bertoldo et al. 2015, Pinaria et al.
W.C. Snyder, Phytopathology 31: 1056. 1941 1999, Katan & Di Primo 2015, Koyyappurath et al. 2016, Taylor et al. 2016
Fusarium oxysporum var. dianthi (Prill. & 1999
Delacr.) Raillo, Fungi of the genus Fusarium:
255. 1950
dioscoreae Wellman 1972
echeveriae Ortu et al. 2015a Ortu et al. 2015a
elaeagni Armstrong & Armstrong 1968 Fusarium oxysporum var. orthoceras Armstrong & Armstrong 1968, 1981,
(Appel & Wollenw.) Bilaǐ, The Fusaria: Booth 1971, Summerell et al. 2010
282. 1955
elaeidis Gordon 1965 Gordon 1965, Booth 1971, See Flood 2006 for prior See Flood 2006 for prior publications; Bogale et al.
Armstrong & Armstrong 1981, publications 2007, O’Donnell et al. 2009, Elliott et al. 2010
Summerell et al. 2010
erucae Chatterjee & Rai 1974
erythroxyli Sands et al. 1997 Summerell et al. 2010 Sands et al. 1997, Kistler Sands et al. 1997, Lievens et al. 2009b, O’Donnell
et al. 1998, Katan 1999, et al. 2009
Katan & Di Primo 1999
eucalypti Arya & Jain 1962 Gordon 1965, Armstrong & Arm-
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
eustomae Raabe 1985a Bertoldo et al. 2015
fabae Yu & Fang 1948 Gordon 1965, Armstrong & Arm- Mbofung et al. 2007, O’Donnell et al. 2009, Srinivasan
strong 1968, 1981, Booth 1971 et al. 2010, Mirtalebi & Banihashemi 2014
fatshederae Triolo & Lorenzini 1983 O’Donnell et al. 2009
foli see Hirooka et al. 2008 Hirooka et al. 2008
fragariae Winks & Williams 1965 Armstrong & Armstrong 1968, 1981, Katan & Di Primo 1999, Kim et al. 2001, Nagarajan et al. 2004, 2006, Hirano
Booth 1971, Summerell et al. 2010 Nagarajan et al. 2006 & Arie 2009, O’Donnell et al. 2009, Chakrabarti et
al. 2011, Fang et al. 2013, Poli et al. 2013, Suga et
al. 2013, Bertoldo et al. 2015, Czislowski et al. 2017,
Henry et al. 2017
freesia Taylor et al. 2016
garlic Matuo et al. 1986 Yoo et al. 1993, Katan &
Di Primo 1999
gerberae Von Arx 1952, Gordon 1965 Gordon 1965, Armstrong & Armstrong 1968,
1981, Booth 1971, Summerell et al. 2010
gladioli Massey 1926, Snyder & Hansen Fusarium oxysporum var. gladioli Snyder & Hansen 1940, Gordon Roebroeck & Mes 1992, Mes Molnár et al. 1990, Mes et Mes et al. 1994, Crowhurst et al. 1995, Baayen et al.
1940, Buxton 1955 Massey, Phytopathology 16: 511. 1926 1965, Armstrong & Armstrong et al. 1994, De Haan et al. 2000 al. 1994, Kistler et al. 1998, 2000, De Haan et al. 2000, Kim et al. 2001, Bogale
Fusarium oxysporum f. gladioli 1968, 1981, Booth 1971, Summerell Katan 1999, Katan & Di et al. 2007, O’Donnell et al. 2009, Elliott et al. 2010,
(Massey) W.C. Snyder & H.N. Hansen, et al. 2010 Primo 1999, Di Primo et al. Lin et al. 2010, Pinaria et al. 2015, Van Dam & Rep
Amer. J. Bot. 27: 66. 1940 2002 2017
Fusarium orthoceras var. gladioli
L. McCulloch, Phytopathology 34:
280. 1944
glycines Armstrong & Armstrong 1965 Armstrong & Armstrong 1968, 1981, Lievens et al. 2009b, O’Donnell et al. 2009, Pinaria
Booth 1971, Summerell et al. 2010 et al. 2015, Koyyappurath et al. 2016
hebes Raabe 1985b Gordon 1965, Armstrong & Armstrong
1968, 1981, Booth 1971, Summerell
et al. 2010
heliconiae Waite 1963 (see Ploetz 2006)
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
7
L. Lombard et al.: Epitypification of Fusarium oxysporum
heliotropae Netzer & Weintal 1987 Mbofung et al. 2007, O’Donnell et al. 2009
herbemontis Gordon 1965 Fusarium oxysporum var. herbemontis Gordon 1965, Armstrong & Armstrong
Tochetto, Revta Agron., Porto Alegre: 1968, 1981, Booth 1971, Summerell
82– 89. 1954 et al. 2010
iridiacearum Roebroeck 2000 Roebroeck 2000 Roebroeck 2000 Roebroeck 2000
koae Gardner 1980 Shiraishi et al. 2012 O’Donnell et al. 2009, Shiraishi et al. 2012
laciniati Pandotra et al. 1971 Summerell et al. 2010
lactucae Matuo & Motohashi 1967, Summerell et al. 2010 Fujinaga et al. 2001, 2003, 2005, Kistler et al. 1998, Katan Fujinaga et al. 2005, 2014, Shimazu et al. 2005,
Hubbard & Gerik 1993 2014, Yamauchi et al. 2001, 2004, 1999, Ogiso et al. 2002, Mbofung et al. 2007, Pasquali et al. 2007, 2008,
Ogiso et al. 2002, Shimazu et al. Yamauchi et al. 2004, Lievens et al. 2008, Hirano & Arie 2009, O’Donnell
2005, Pasquali et al. 2007, 2008, Pasquali et al. 2005, 2008, et al. 2009, Lin et al. 2010, 2014, Mbofung & Pryor
Lin et al. 2014, Gilardi et al. 2017 Pintore et al. 2017 2010, Poli et al. 2012, 2013, Mirtalebi & Banihas-
hemi 2014, Bertoldo et al. 2015, Gilardi et al. 2017
lagenariae Matuo & Yamamoto 1967 Armstrong & Armstrong 1968, Armstrong & Armstrong 1978b Katan & Di Primo 1999 Okuda et al. 1998, Kim et al. 2001, Galván et al.
1981, Booth 1971, Summerell 2008, Hirano & Arie 2009, O’Donnell et al. 2009,
et al. 2010 Poli et al. 2013
lathyri Bhide & Uppal 1948 Fusarium oxysporum var. lathyri Gordon 1965,Armstrong & Arm-
V.P. Bhide & Uppal, Phytopathology 38: strong 1968, 1981, Booth 1971,
560– 567. 1948 Summerell et al. 2010
lentis Vasudeva & Srinivasan 1952 Fusarium orthoceras var. lentis Gordon 1965, Armstrong & Arm- Pouralibaba et al. 2016, 2017 Belabid & Fortas 2002 Belabid et al. 2004, O’Donnell et al. 2009, Taheri et
Vasudeva & Sriniv., Indian Phytopathol. strong 1968, 1981, Booth 1971, al. 2010, Datta et al. 2011, Mohammadi et al. 2011,
5: 28. 1953 Summerell et al. 2010 Rafique et al. 2015, Al-Husien et al. 2017, Nourollahi
& Madahjalai 2017
lilii Imle 1942 Gordon 1965, Armstrong & Arm- Löffler & Rumine 1991, Baayen et al. 1998, 2000, Kim et al. 2001, Skovgaard
strong 1968, 1981, Booth 1971, Baayen et al. 1998, Kistler et al. 2001, Wang et al. 2001, O’Donnell et al. 2009,
1981, Summerell et al. 2010 et al. 1998, Katan 1999, Lin et al. 2010, Baysal et al. 2013, Van Dam & Rep
Katan & Di Primo 1999 2017
lini Bolley 1901 Fusarium lini Bolley, Proc. Ann. Snyder & Hansen 1940, Gordon Katan & Di Primo 1999, Baayen et al. 2000, Bogale et al. 2007, O’Donnell
Meeting Soc. Prom. Agr. Sci. 22: 42. 1965, Armstrong & Armstrong 1968, Baayen et al. 2000 et al. 2009, Pinaria et al. 2015, Taylor et al. 2016
1901 1981, Booth 1971, Summerell
Fusarium oxysporum f. lini (Bolley) et al. 2010
W.C. Snyder & H.N. Hansen, Amer.
J. Bot. 27: 66. 1940
loti Bergstrom & Kalb 1995 Wunsch et al. 2009 Galván et al. 2008, O’Donnell et al. 2009, Wunsch
et al. 2009
luffae Kawai et al. 1958 Summerell et al. 2010 Armstrong & Armstrong 1978b Kim et al. 1993, Wang et al. 2001, Lin et al. 2010
lupini Snyder & Hansen 1940 Snyder & Hansen 1940, Gordon Richter 1941, Armstrong & Armstrong Kistler et al. 1998, Katan Bogale et al. 2007, O’Donnell et al. 2009
1965, Armstrong & Armstrong 1968, 1964, Rataj-Guranowska et al. 1984 1999, Katan & Di Primo
1981, Booth 1971, Summerell 1999
et al. 2010
lycopersici Wollenweber 1913 Fusarium oxysporum subsp. lycopersici Snyder & Hansen 1940, Gordon Alexander & Tucker 1945, Gerdemann Puhalla 1985, Correll et al. Elias & Schneider 1992; Elias et al. 1993, Crowhurst
Sacc., Syll. Fung. 4: 705. 1886 1965, Armstrong & Armstrong 1968, & Finley 1951, Gabe 1975, Elias & 1987, Hadar et al. 1989, et al. 1995, Marlatt et al. 1996, Mes et al. 1998,
Fusarium lycopersici Bruschi, Rc. 1981, Booth 1971, Summerell Schneider 1992, Elias et al. 1993, Molnár et al. 1990, Correll Gherbawy 1999, Kim et al. 2001, Bao et al. 2002, Cai
Accad. Naz. Lincei: 298. 1912 et al. 2010 Marlatt et al. 1996, Mes et al. 1998, 1991, Elias & Schneider et al. 2003, Hirano & Arie 2006, 2009, Bogale et al.
Fusarium lycopersici (Sacc.) Wollenw., Cai et al. 2003, Hirano & Arie 2006, 1991, 1992, Marlatt et al. 2007, Mbofung et al. 2007, Lievens et al. 2009a, b,
Phytopathology 3 (1): 29. 1913 Lievens et al. 2009a 1996, Kistler et al. 1998, O’Donnell et al. 2009, Elliott et al. 2010, Inami et al.
Fusarium oxysporum f. lycopersici Mes et al. 1998, Katan 2010, Ma et al. 2010, See review by Takken & Rep
(Sacc.) W.C. Snyder & H.N. Hansen, 1999, Katan & Di Primo 2010, Chakrabarti et al. 2011, Poli et al. 2012, 2013,
Amer. J. Bot. 27: 66. 1940 1999, Cai et al. 2003 Thatcher et al. 2012, Baysal et al. 2013, Bennett et
al. 2013, Covey et al. 2014, Gawehns et al. 2014,
Mirtalebi & Banihashemi 2014, Bertoldo et al. 2015,
Hansen et al. 2015, Nirmaladevi et al. 2016, Taylor
et al. 2016, Williams et al. 2016, Bilju et al. 2017, Van
Dam & Rep 2017, Jelinski et al. 2017
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
8Persoonia – Volume 43, 2019
magnoliae Lin & Chen 1994
matthiolae Baker 1948 Booth 1971, Summerell et al. 2010 Correll 1991, Kistler et al. Kistler et al. 1987, Mbofung et al. 2007, O’Donnell
1998, Katan 1999 et al. 2009, Srinivasan et al. 2010, Poli et al. 2012
medicaginis Weimer 1928 Fusarium oxysporum var. medicaginis Snyder & Hansen 1940, Gordon Puhalla 1985, Correll et al. Mbofung et al. 2007, O’Donnell et al. 2009, Srinivasan
Weimer, J. Agric. Res. 37: 425. 1928 1965, Armstrong & Armstrong 1987, Molnár et al. 1990, et al. 2010, Poli et al. 2012, Mirtalebi & Banihashemi
Fusarium oxysporum f. medicaginis 1968, 1981, Booth 1971, Kistler et al. 1998, Katan 2014, Thatcher et al. 2016, Williams et al. 2016,
(Weimer) W.C. Snyder & H.N. Hansen, Summerell et al. 2010 1999 Czislowski et al. 2017
Amer. J. Bot. 27: 66. 1940
melongenae Matuo & Ishigami 1958 Gordon 1965, Armstrong & Hadar et al. 1989, Kistler Crowhurst et al. 1995, Kim et al. 2001, Hirano & Arie
Armstrong 1968, Booth 1971, et al. 1998, Katan 1999, 2009, O’Donnell et al. 2009, Altinok & Can 2010,
1981, Summerell et al. 2010 Katan & Di Primo 1999, Baysal et al. 2010, Bennett et al. 2013, Poli et al.
Altinok & Can 2010, Altinok 2013, Bertoldo et al. 2015, Dong et al. 2017
2013, Altinok et al. 2013
melonis Leach & Currence 1938, Fusarium bulbigenum var. niveum Snyder & Hansen 1940, Gordon Risser & Mas 1965, Risser et al. Correll et al. 1987, Jacobson & Gordon 1990b, Kim et al. 1993, 2001,
Snyder & Hansen 1940 Leach & Curr., Minnisota Agric. Exp. 1965, Armstrong & Armstrong 1976, Armstrong & Armstrong 1978b, Jacobson & Gordon 1988, Crowhurst et al. 1995, Namiki et al. 1998, 2001,
Sta. Tech. Bull. 129: 1–32. 1938 1968, 1981, Booth 1971, Gerlagh & Blok 1988, Katan et al. 1994, 1990a, Hadar et al. 1989, Gherbawy 1999, Skovgaard et al. 2001, Mbofung et
Summerell et al. 2010 Luongo et al. 2014, Mirtalebi & Correll 1991, Katan et al. al. 2007, Hirano & Arie 2009, Lievens et al. 2009b,
Banihashemi 2014, Sebastiani et 1994, Kistler et al. 1998, O’Donnell et al. 2009, Lin et al. 2010, Bennett et al.
al. 2017 Katan 1999, Katan & Di 2013, Poli et al. 2013, Covey et al. 2014, Gawehns
Prim o 199 9, Mir talebi & et al. 2014, Luongo et al. 2014, Ma et al. 2014, Mirtalebi
Banihashemi 2014 & Banihashemi 2014, Bertoldo et al. 2015, Hansen
et al. 2015, Pinaria et al. 2015, Schmidt et al. 2016,
Taylor et al. 2016, Williams et al. 2016, Van Dam &
Rep 2017, Sebastiani et al. 2017
meniscoideum (var.) Bugnicourt 1939 Gerlach & Nirenberg 1982 O’Donnell et al. 2009
momordicae Sun & Huang 1983 Skovgaard et al. 2001, O’Donnell et al. 2009, Lin et
al. 2010, Bennett et al. 2013, Chen et al. 2015
mori Pastrana et al. 2017 Pastrana et al. 2017 Pastrana et al. 2017
narcissi Wollenweber & Reinking 1935, Snyder & Hansen 1940, Gordon Linfield 1993, Crowhurst et al. 1995, O’Donnell et al.
Snyder & Hansen 1940 1965, Armstrong & Armstrong 2009, Taylor et al. 2016, Van Dam & Rep 2017
1968, 1981, Booth 1971,
Summerell et al. 2010
nelumbicola Gordon 1965 Fusarium bulbigenum var. nelumbicola Snyder & Hansen 1940, Gordon
Y. Nisik. & Kyoto Watan., Ber. Ohara 1965, Armstrong & Armstrong
Inst. Landw. Biol. Okayama Univ.: 3. 1968, 1981, Booth 1971,
1953 Summerell et al. 2010
nicotianae Johnson 1921 Fusarium oxysporum var. nicotianae Booth 1971, Summerell et al. 2010 Bogale et al. 2007, O’Donnell et al. 2009
J. Johnson, J. Agric. Res. 20: 525. 1921
niveum Wollenweber & Reinking 1935 Fusarium niveum E.F. Sm., Bull. Snyder & Hansen 1940, Gordon Reid 1958, Crall 1963, Netzer 1976, Puhalla 1985, Correll et al. Kim et al. 1993, 2001, Crowhurst et al. 1995, Zhang
U.S.D.A. 1894 1965, Armstrong & Armstrong Armstrong & Armstrong 1978b, 1987, Hadar et al. 1989, et al. 2005, Bogale et al. 2007, Hirano & Arie 2009,
Fusarium bulbigenum var. niveum 1968, 1981, Booth 1971, Martyn 1987, Gerlagh & Blok 1988, Larkin et al. 1988, 1990, O’Donnell et al. 2009, Lin et al. 2010, Chakrabarti et
(E.F. Sm.) Wollenw., Die Fusarien: Summerell et al. 2010 Martyn & Bruton 1989, Larkin et al. Correll 1991, Kistler et al. al. 2011, Poli et al. 2013, Gawehns et al. 2014, Mirtalebi
117. 1935 1990, Zhou et al. 2010 1998, Katan 1999, Katan & Banihashemi 2014, Bertoldo et al. 2015, Ren et al.
Fusarium oxysporum f. niveum & Di Primo 1999 2015, Van Dam & Rep 2017, Czislowski et al. 2017
(E.F. Sm.) W.C. Snyder & H.N. Hansen,
Amer. J. Bot. 27: 66. 1940
opuntiarum Gordon 1965 Fusarium oxysporum var. opuntiarum Gordon 1965, Armstrong & Katan & Di Primo 1999 Baayen et al. 2000, Mbofung et al. 2007, O’Donnell
Pettinari, Annali Sper. Agr.: 1419. 1951 Armstrong 1968, 1981, et al. 2009, Ortu et al. 2013, Pinaria et al. 2015,
Booth 1971, Summerell et al. 2010 Koyyappurath et al. 2016, Bertetti et al. 2017
orthoceras Bilaǐ 1955
oxysporum (var.) Von Schlechtendahl 1824 Gerlach & Nirenberg 1982
palmarum Elliott et al. 2010 O’Donnell et al. 2009, Elliott et al. 2010, 2017,
Giesbrecht et al. 2013
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
9
L. Lombard et al.: Epitypification of Fusarium oxysporum
papaveris Ortu et al. 2015b Summerell et al. 2010 Katan 1999 Bertetti et al. 2014, Ortu et al. 2015b
passiflorae Gordon 1965 Gordon 1965, Armstrong & Arm- Gherbawy 1999, Bogale et al. 2007, Lievens et al.
strong 1968, 1981, Booth 1971, 2009b, O’Donnell et al. 2009, Chakrabarti et al. 2011,
Summerell et al. 2010 Dos Santos Silva et al. 2013, Gawehns et al. 2014,
Pinaria et al. 2015, Koyyappurath et al. 2016, Czislowski
et al. 2017
perillae Kim et al. 2002
perniciosum Toole 1941 Fusarium perniciosum Hepting, Gordon 1965, Armstrong & Arm- Toole 1952 Crowhurst et al. 1995, Bogale et al. 2007, Mbofung
Circ. U.S.D.A.: 7. 1939 strong 1968, 1981, Booth 1971, et al. 2007, Lievens et al. 2009b, O’Donnell et al.
Fusarium oxysporum f. perniciosum Summerell et al. 2010 2009, Elliott et al. 2010, Bennett et al. 2013, Pinaria
(Hepting) Toole, Phytopathology 31: et al. 2015
599. 1941
Fusarium vasinfectum var. perniciosum
(Hepting) Carrera, Monatsh. Landw.:
483. 1955
phaseoli Kendrick & Snyder 1942b Gordon 1965, Armstrong & Arm- Ribeiro 1977, Ribeiro & Hagedorn Woo et al. 1996, Kistler Woo et al. 1996, Cramer et al. 2003, Zanotti et al.
strong 1968, 1981, Booth 1971, 1979, Salgado & Schwartz 1993, et al. 1998, Katan 1999, 2006, Alves-Santos et al. 2002b, Bogale et al. 2007,
Summerell et al. 2010 Woo et al. 1996, Alves-Santos et al. Katan & Di Primo 1999, Mbofung et al. 2007, Hirano & Arie 2009, O’Donnell
2002a, Cramer et al. 2003, Henrique Alves-Santos et al. 2002a et al. 2009, De Vega-Bartol et al. 2011, Baysal et al.
et al. 2015 2013, Poli et al. 2013, Mirtalebi & Banihashemi 2014,
Da Silva et al. 2014, Bertoldo et al. 2015, De Sousa
et al. 2015
phormii Wager 1947 Gordon 1965, Armstrong & Arm-
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
pini Hartig 1892, Snyder & Hansen 1940 Fusisporium aurantiacum Link, Mag. O’Donnell et al. 2009
Ges. Naturf. Freunde Berlin 3: 19. 1809
Fusoma pini Hartig, Forstl.-Naturwiss.
Z. 1: 432– 436. 1892
Fusarium blasticola Rostr., Gartner-
Tidende 1895: 122. 1895
Fusarium oxysporum f. pini (Hartig) W.C.
Snyder & H.N. Hansen, Amer. J. Bot. 27: 66. 1940
Fusarium oxysporum f. sp. blasticola Bilaǐ,
Fusarii: 281. 1955
pisi Van Hall 1903, Snyder & Hansen 1940 Fusarium vasinfectum var. pisi C.J.J. Snyder & Hansen 1940, Gordon Snyder & Walker 1935, Snyder & Puhalla 1985, Correll et al. Coddington et al. 1987, Kistler et al. 1991, Whitehead
Hall, Ber. Deutsch. Bot. Ges. 21: 4. 1903 1965, Armstrong & Armstrong Hansen 1940, Schreuder 1951, 1987, Correll 1991, White- et al. 1992, Grajal-Martin et al. 1993, Gherbawy 1999,
Fusarium orthoceras var. pisi Linford, 1968, 1981, Booth 1971, Bolton et al. 1966, Armstrong & head et al. 1992, Kistler Skovgaard et al. 2001, O’Donnell et al. 2009,
Res. Bull. Agric. Exp. Stn Univ. Wis.: Summerell et al. 2010 Armstrong 1974, Kraft & Haglund et al. 1998, Katan 1999, Chakrabarti et al. 2011, Covey et al. 2014, Mirtalebi
11. 1928 1978, Haglund & Kraft 1979, Katan & Di Primo 1999, & Banihashemi 2014, Hansen et al. 2015, Taylor et
Fusarium oxysporum f. 8 W.C. Snyder, Coddington et al. 1987, Whitehead al. 2016, Williams et al. 2016, Van Dam & Rep 2017
Zentralbl. Bakteriol., 2. Abt.: 374. 1935 et al. 1992, Grajal-Martin et al. 1993
Fusarium oxysporum var. pisi (C.J.J. Hall)
Raillo, Fungi of the genus Fusarium: 254.
1950
Fusarium oxysporum var. orthoceras
(Appel & Wollenw.) Bilaǐ, Fusarii: 282. 1955
psidii Prasad et al. 1952 Gordon 1965, Armstrong & Arm- Gupta 2012, Mishra et al. 2013a, b, c, 2014
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
pyracanthae McRitchie 1973, Armstrong & Armstrong 1981 Armstrong & Armstrong 1968, 1981,
Summerell et al. 2010
querci Gordon 1965 Gordon 1965, Armstrong & Arm-
strong 1968, 1981, Booth 1971,
Summerell et al. 2010
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
10 Persoonia – Volume 43, 2019
quitoense Ochoa et al. 2004
radicis-capsici Lomas-Cano et al. 2014, 2016 Lomas-Cano et al. 2014
radicis-cucumerinum Vakalounakis 1996 Summerell et al. 2010 Katan 1999, Katan & Di Vakalounakis & Fragkiadakis 1999, Vakalounakis
Primo 1999, Vakalounakis et al. 2004, 2005, Lievens et al. 2007, Van Dam
& Fragkiadakis 1999, & Rep 2017
Vakalounakis et al. 2004,
2005, Tok & Kurt 2010
radicis-lupini Weimer 1944 Gordon 1965, Booth 1971,
Summerell et al. 2010
radicis-lycopersici Jarvis & Shoemaker 1978 Summerell et al. 2010 Puhalla 1985, Correll et al. Kim et al. 2001, Skovgaard et al. 2001, Balmas et al.
1987, Katan et al. 1991, 2005, Hirano & Arie 2006, 2009, Bogale et al. 2007,
Kistler et al. 1998, Katan Hibar et al. 2007, O’Donnell et al. 2009, Huang et al.
1999, Katan & Di Primo 2013, Poli et al. 2013, Covey et al. 2014, Mirtalebi &
1999, Rosewich et al. Banihashemi 2014, Bertoldo et al. 2015, Taylor
1999, Di Primo et al. 2001, et al. 2016
Balmas et al. 2005, Huang
et al. 2013
radicis-vanillae Koyyappurath et al. 2016 Koyyappurath et al. 2016
ranunculi Garibaldi & Gullino 1985
rapae Enya et al. 2008 Enya et al. 2008 Enya et al. 2008
raphani Kendrick & Snyder 1942a Gordon 1965, Armstrong & Bosland & Williams 1987, Kistler & Benny 1989, Kistler et al. 1991, Kim et al.
Armstrong 1968, 1981, Booth Kistler et al. 1998, Katan 2001, Bogale et al. 2007, Hirano & Arie 2009, O’Donnell
1971, Summerell et al. 2010 1999, Katan & Di Primo et al. 2009, Lin et al. 2010, Srinivasan et al. 2010,
1999 Poli et al. 2012, 2013, Covey et al. 2014, Bertoldo
et al. 2015, Taylor et al. 2016, Van Dam & Rep
2017, Kim et al. 2017
rauvolfiae Janardhanan et al. 1964 Gordon 1965, Armstrong & O’Donnell et al. 2009
Armstrong 1968, 1981, Booth
1971, Summerell et al. 2010
rhois Snyder et al. 1949 Gordon 1965, Armstrong & Mbofung et al. 2007
Armstrong 1968, 1981, Booth
1971, Summerell et al. 2010
ricini Gordon 1965 Fusarium orthoceras var. ricini Gordon 1965, Armstrong & Prasad et al. 2008, Reddy et al. 2012
Wollenw., Biologico 6: 148. 1940 Armstrong 1968, 1981, Booth
1971, Summerell et al. 2010
samaneae Wellman 1972
sansevieriae Gupta et al. 1982
sedi Raabe 1960 Gordon 1965, Armstrong &
Armstrong 1968, 1981, Booth
1971, Summerell et al. 2010
sesami Gordon 1965, Booth 1971 Fusarium vasinfectum var. sesami Gordon 1965, Armstrong & Basirnia & Banihashemi O’Donnell et al. 2009, Li et al. 2012,
Zaprom., Pflanzenschutz-Vers. Armstrong 1968, 1981, Booth 2005 Bennett et al. 2013
Sta. Taschkent: 36 pp. 1926 1971, Summerell et al. 2010
sesbaniae Gordon 1965, Booth 1971 Gordon 1965, Armstrong &
Armstrong 1968, 1981, Booth
1971, Summerell et al. 2010
spinaciae Hungerford 1923 Fusarium spinaciae Sherb., Snyder & Hansen 1940, Gordon Armstrong & Armstrong 1976 Kistler et al. 1998, Katan Baayen et al. 2000, Kim et al. 2001, Skovgaard et al.
Phytopathology 13: 209. 1923 1965, Armstrong & Armstrong 1999, Katan & Di Primo 2001, Kawabe et al. 2007, Mbofung et al. 2007, Hirano
Fusarium oxysporum f. spinaciae 1968, 1981, Booth 1971, 1999, Takehara et al. & Arie 2009, O’Donnell et al. 2009, Poli et al. 2012,
(Sherb.) W.C. Snyder & H.N. Hansen, Summerell et al. 2010 2003 2013, Bennett et al. 2013, Okubara et al. 2013,
Amer. J. Bot. 27: 66. 1940 Covey et al. 2014, Mirtalebi & Banihashemi
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
11
L. Lombard et al.: Epitypification of Fusarium oxysporum
spinaciae (cont.) Fusarium redolens f. spinaciae 2014, Bertoldo et al. 2015
(Sherb.) Subram., Hyphomycetes:
an account of Indian species,
except Cercosporae: 690. 1971
stachydis Gordon 1965 Gordon 1965, Armstrong &
Armstrong 1968, 1981, Booth
1971, Summerell et al. 2010
strigae Elzein & Kroschel 2006 Elzein et al. 2008, Zimmermann et al. 2015, 2016
tabernaemontanae Pande & Rao 1990
tanaceti Hirooka et al. 2008 Hirooka et al. 2008
tracheiphilum Wollenweber 1931, Snyder & Hansen 1940 Fusarium tracheiphilum E.F. Sm. 1899 Snyder & Hansen 1940, Gordon Armstrong & Armstrong 1950, 1980, Correll et al. 1987, Kistler Gherbawy 1999, Bao et al. 2002, Hirano & Arie 2009,
Fusarium bulbigenum var. tracheiphilum 1965, Armstrong & Armstrong Hare 1953, Swanson & Van Gundy et al. 1998, Katan 1999, O’Donnell et al. 2009, Lin et al. 2010, Troisi et al.
(E.F. Sm.) Wollenw., Z. Parasitenk. 1968, 1981, Booth 1971, 1985, Smith et al. 1999 Katan & Di Primo 1999, 2010, Bennett et al. 2013, Poli et al. 2013, Bertoldo
(Berlin) 3: 413. 1931 Summerell et al. 2010 Bao et al. 2002 et al. 2015, Koyyappurath et al. 2016
Fusarium oxysporum f. tracheiphilum
(E.F. Sm.) W.C. Snyder & H.N. Hansen,
Amer. J. Bot. 27: 66. 1940
trifolii Bilaǐ 1955 Fusarium trifolii Jacz., Jb. Pfl. krankh. Gordon 1965, Armstrong &
Russl. VII-VIII, Abt. 6. 1917 Armstrong 1968, 1981, Booth
Fusarium oxysporum var. trifolii (Jacz.) 1971, Summerell et al. 2010
Raillo, Fungi of the genus Fusarium: 255.
1950
tuberosi Snyder & Hansen 1940 Fusarium oxysporum var. solani Snyder & Hansen 1940, Gordon Molnár et al. 1990, Venter Gherbawy 1999, Lievens et al. 2009a,
Raillo, Fungi of the genus Fusarium: 1965, Armstrong & Armstrong et al. 1992, Kistler et al. O’Donnell et al. 2009
254. 1950 1968, 1981, Booth 1971, 1998, Katan 1999
Fusarium oxysporum var. solani Summerell et al. 2010
(Raillo) Bilaǐ, Fusarii: 281. 1955
tulipae Snyder & Hansen 1940 Gordon 1965, Armstrong & Katan 1999, Katan & Gherbawy 1999, Baayen et al. 2000, Kim et al. 2001,
Armstrong 1968, 1981, Booth Di Primo 1999 Skovgaard et al. 2001, Hirano & Arie 2009, O’Donnell
1971, Summerell et al. 2010 et al. 2009, Poli et al. 2013, Mirtalebi & Banihashemi
2014, Bertoldo et al. 2015, Pinaria et al. 2015, Swett
& Uchida 2015, Van Dam & Rep 2017
vanillae Tucker 1927 Fusarium batatas var. vanillae Gordon 1965, Armstrong & Katan & Di Primo 1999 O’Donnell et al. 2009, Chakrabarti et al. 2011,
Tucker, J. Agric. Res. 44: 1121. 1927 Armstrong 1968, 1981, Booth Adame-García et al. 2015, Pinaria et al. 2015,
1971, Summerell et al. 2010 Koyyappurath et al. 2016
vasconcella Ochoa et al. 2004
vasinfectum Atkinson 1892 Fusarium vasinfectum G.F. Atk., Snyder & Hansen 1940, Gordon Armstrong & Armstrong 1958a, 1960, Puhalla 1985, Correll et al. Assigbetse et al. 1994, Fernandez et al. 1994,
Bulletin of the Alabama Agricultural 1965, Armstrong & Armstrong 1978a, Ibrahim 1966, Kappelman 1987, Katan & Katan 1988, Crowhurst et al. 1995, Moricca et al. 1998, Skovgaard
Experiment Station: 28. 1892 1968, 1981, Booth 1971, 1983, Chen et al. 1985, Assigbetse Hadar et al. 1989, Correll et al. 2001, Smith et al. 2001, Abd-Elsalam et al. 2002,
Fusarium oxysporum f. vasinfectum Summerell et al. 2010 et al. 1994, Fernandez et al. 1994, 1991, Fernandez et al. 2004, 2006, Abo et al. 2005, Kim et al. 2005, 2017,
(G.F. Atk.) W.C. Snyder & H.N. Hansen, Nirenberg et al. 1994, Skovgaard et 1994, Davis et al. 1996, McFadden et al. 2006, Wang et al. 2006, 2010,
Amer. J. Bot. 27: 66. 1940 al. 2001, Kim et al. 2005, Holmes et al. Kistler et al. 1998, Katan Mbofung et al. 2007, Zambounis et al. 2007, Bennett
2009, Guo et al. 2015 1999, Katan & Di Primo et al. 2008, 2013, Holmes et al. 2009, O’Donnell et
1999, Abo et al. 2005, al. 2009, Elliot et al. 2010, Chakrabarti et al. 2011,
Wang et al. 2010 Egamberdiev et al. 2013, 2014, Da Silva et al. 2014,
Covey et al. 2014, Doan et al. 2014, Cianchetta et al.
2015, Guo et al. 2015, Pinaria et al. 2015, Crutcher
et al. 2016, Taylor et al. 2016, Van Dam & Rep 2017,
Ortiz et al. 2017
voandzeiae Armstrong et al. 1975 Armstrong & Armstrong 1981 O’Donnell et al. 2009
zingiberi Trujillo 1963 Pappalardo et al. 2009 Katan & Di Primo 1999 Crowhurst et al. 1995, O’Donnell et al. 2009,
Pappalardo et al. 2009, Chakrabarti et al. 2011,
Gupta et al. 2014, Czislowski et al. 2017
Table 1 (cont.)
formae speciales Description Synonym(s) Listed Race(s) VCG(s) Molecular studies
12 Persoonia – Volume 43, 2019
naming of the multiple cryptic species recognised in this spe-
cies complex, Fusarium isolates were collected from the type
locality in Berlin, Germany, and the type substrate, Solanum
tuberosum. Using molecular phylogenetic and morphological
tools, an epitype is designated for F. oxysporum in the present
study based on these collections.
MATERIALS AND METHODS
Isolates
Tubers of S. tuberosum (potato), displaying symptoms of dry
rot, were collected from several vegetable gardens in Berlin,
Germany. Potato tubers were placed individually in paper
bags, stored at 4 °C until transported to the laboratory for
further processing. After surface-sterilisation of the potato tu-
bers using a 10 % (v/v) sodium hypochlorite solution, pieces
of symptomatic tissue were removed from the leading edges
of the rot lesions and plated onto 2 % (w/v) potato dextrose
agar (PDA) amended with 100 µg/mL penicillin and 100 µg/
mL streptomycin, and peptone pentachloronitrobenzene agar
(PCNB; Nash & Snyder 1962) and incubated at 25 °C in the
dark. Axenic cultures were prepared on PDA from characteristic
Fusarium colonies. Additional strains, previously identified as
F. oxysporum, were obtained from the culture collection (CBS)
of the Westerdijk Fungal Biodiversity Institute (WFBI), Utrecht,
the Netherlands, and the working collection of Pedro W. Crous
(CPC) housed at WFBI (Table 2).
DNA isolation, PCR and sequencing
Total genomic DNA was extracted from isolates grown for 7 d on
PDA at 24 °C using a 12/12 h photoperiod using the Wizard®
Genomic DNA purification Kit (Promega Corporation, Madison,
WI, USA), according to the manufacturer’s instructions. Partial
gene sequences were determined for the β-tubulin (tub2),
calmodulin (cmdA), the intergenic spacer region of the rDNA
(IGS), RNA polymerase II second largest subunit (rpb2) and
translation elongation factor 1-alpha (tef1), using PCR protocols
described elsewhere (O’Donnell et al. 1998, 2007, 2009, 2010,
Lombard et al. 2015). Primer pairs T1/CYLTUB1R (O’Donnell
& Cigelnik 1997, Crous et al. 2004) for tub2, Cal228F/CAL2Rd
(Carbone & Kohn 1999, Groenewald et al. 2013) for cmdA,
iNL11/iCNS1 and the internal sequencing primers NLa/CNSa
(O’Donnell et al. 2009) for IGS, 5f2/7cr (Liu et al. 1999, Sung
et al. 2007) for rpb2, and EF1/EF2 (O’Donnell et al. 1998) for
tef1, were used for amplifications of the respective gene re-
gions. Integrity of the sequences was ensured by sequencing
the amplicons in both directions using the same primer pairs
as were used for amplification. Consensus sequences for
each locus were assembled in MEGA v. 7 (Kumar et al. 2016),
with the exception of the IGS locus, which was assembled in
Geneious R11 (Kearse et al. 2012). All sequences generated
in this study were deposited in GenBank (Table 1).
Phylogenetic analyses
Sequences of the individual loci were aligned using MAFFT
v. 7.110 (Katoh et al. 2017) and manually corrected where
necessary. The individual gene datasets were assessed for
incongruency prior to concatenation using a 70 % reciprocal
bootstrap criterion (Mason-Gamer & Kellogg 1996). Three in-
dependent phylogenetic algorithms, Maximum Parsimony (MP),
Maximum Likelihood (ML) and Bayesian inference (BI), were
employed for phylogenetic analyses. Phylogenetic analyses
were conducted for the individual loci and then as a multilocus
sequence dataset that included the cmdA, rpb2, tef1 and tub2
sequences.
For BI and ML, the best evolutionary models for each locus
were determined using MrModeltest (Nylander 2004) and in-
corporated into the analyses. MrBayes v. 3.2.1 (Ronquist &
Huelsenbeck 2003) was used for BI to generate phylogenetic
trees under optimal criteria for each locus. A Markov Chain
Monte Carlo (MCMC) algorithm of four chains was initiated in
parallel from a random tree topology with the heating parameter
set at 0.3. The MCMC analysis lasted until the average standard
deviation of split frequencies was below 0.01 with trees saved
every 1 000 generations. The first 25 % of saved trees were
discarded as the ‘burn-in’ phase and posterior probabilities (PP)
were determined from the remaining trees.
The ML analyses were performed using RAxML v. 8.2.9 (ran-
domised accelerated (sic) maximum likelihood for high per-
formance computing; Stamatakis 2014) through the CIPRES
website (http://www.phylo.org) to obtain another measure of
branch support. The robustness of the analysis was evalu-
ated by bootstrap support (BS) with the number of bootstrap
replicates automatically determined by the software. For MP,
analyses were done using PAUP (Phylogenetic Analysis Using
Parsimony, v. 4.0b10; Swofford 2003) with phylogenetic rela-
tionships estimated by heuristic searches with 1 000 random
addition sequences. Tree-bisection-reconnection was used,
with branch swapping option set on ‘best trees’ only. All charac-
ters were weighted equally and alignment gaps treated as fifth
state. Measures calculated for parsimony included tree length
(TL), consistency index (CI), retention index (RI) and rescaled
consistence index (RC). Bootstrap (BS) analyses (Hillis &
Bull 1993) were based on 1 000 replications. Alignments and
phylogenetic trees derived from this study were uploaded to
TreeBASE (www.treebase.org).
Genealogical concordance phylogenetic species
recognition (GCPSR)
In order to establish the recombination levels between the newly
proposed species in this study and their closest phylogenetic
relatives, pairwise homoplasy index (PHI) analyses were done
on the respective concatenated multilocus datasets (Bruen et al.
2006). The analyses were conducted as described by Quaed-
vlieg et al. (2014) using SplitsTree v. 4.14.4 (Huson & Bryant
2006). Therefore, a PHI value below 0.05 (fW < 0.05) would
indicate the presence of significant recombination in the dataset.
Split graphs were constructed for visualization of the relation-
ships between closely related species.
Morphological characterisation
All isolates were characterised following the protocols described
by Leslie & Summerell (2006) using potato dextrose agar (PDA;
recipe in Crous et al. 2009), synthetic nutrient-poor agar (SNA;
Nirenberg 1976) and carnation leaf agar (CLA; Fisher et al.
1982). Colony morphology, pigmentation, odour and growth
rates were evaluated on PDA after 3 and 7 d at 24 °C with
a 12/12 h cool fluorescent light/dark cycle as described by
Sandoval-Denis et al. (2018) and using the colour charts of
Rayner (1970). Micromorphological characters were examined
using water as mounting medium on a Zeiss Axioskop 2 plus
with Differential Interference Contrast (DIC) optics and a Nikon
AZ100 stereomicroscope both fitted with Nikon DS-Ri2 high
definition colour digital cameras to photo-document fungal
structures. Measurements were taken using the Nikon software
NIS-elements D v. 4.50 and the 95 % confidence levels were
determined for the conidial measurements with extremes given
in parentheses. For all other fungal structures examined, only the
extremes are presented. To facilitate the comparison of relevant
micro- and macroconidial features, composite photo plates were
assembled from separate photographs using PhotoShop CSS.
13
L. Lombard et al.: Epitypification of Fusarium oxysporum
Fusarium callistephi CBS 187.53T Callistephus chinensis callistephi The Netherlands MH484693 MH484784 MH484875 MH484966 MH485057
CBS 115423 Agathosma betulina South Africa MH484723 MH484814 MH484905 MH484996 MH485087
F. carminascens CBS 144739 = CPC 25792 Zea mays South Africa MH484752 MH484843 MH484934 MH485025 MH485116
CBS 144740 = CPC 25793 Z. mays South Africa MH484753 MH484844 MH484935 MH485026 MH485117
CBS 144741 = CPC 25795 Z. mays South Africa MH484754 MH484845 MH484936 MH485027 MH485118
CBS 144738 = CPC 25800T Z. mays South Africa MH484755 MH484846 MH484937 MH485028 MH485119
F. contaminatum CBS 111552 Pasteurized fruit juice The Netherlands MH484718 MH484809 MH484900 MH484991 MH485082
CBS 114899T Pasteurized chocolate milk Germany MH484719 MH484810 MH484901 MH484992 MH485083
CBS 117461 Tetra pack with milky nutrition The Netherlands MH484729 MH484820 MH484911 MH485002 MH485093
F. cugenangense CBS 620.72 = DSM 11271 = NRRL 36520 Crocus sp. gladioli Germany MH484697 MH484788 MH484879 MH484970 MH485061
CBS 130304 = BBA 69050 = NRRL 25433 Gossypium barbadense vasinfectum China MH484739 MH484830 MH484921 MH485012 MH485103
CBS 130308 = ATCC 26225 = NRRL 25387 Human toe nail New Zealand MH484738 MH484829 MH484920 MH485011 MH485102
CBS 131393 Vicia faba Australia MH484746 MH484837 MH484928 MH485019 MH485110
F. curvatum CBS 247.61 = BBA 8398 = DSM 62308 = NRRL 22545 Matthiola incana matthiolae Germany MH484694 MH484785 MH484876 MH484967 MH485058
CBS 238.94 = NRRL 26422 = PD 94/184T Beaucarnia sp. meniscoideum The Netherlands MH484711 MH484802 MH484893 MH484984 MH485075
CBS 141.95 = NRRL 36251 = PD 94/1518 Hedera helix The Netherlands MH484712 MH484803 MH484894 MH484985 MH485076
F. duoseptatum CBS 102026 = NRRL 36115 Musa sapientum cv. Pisang ambon cubense Malaysia MH484714 MH484805 MH484896 MH484987 MH485078
F. elaeidis CBS 217.49 = NRRL 36358 Elaeis sp. elaeidis Zaire MH484688 MH484779 MH484870 MH484961 MH485052
CBS 218.49 = NRRL 36359 Elaeis sp. elaeidis Zaire MH484689 MH484780 MH484871 MH484962 MH485053
CBS 255.52 = NRRL 36386 Elaeis guineensis elaeidis Unknown MH484692 MH484783 MH484874 MH484965 MH485056
F. fabacearum CBS 144742 = CPC 25801 Z. mays South Africa MH484756 MH484847 MH484938 MH485029 MH485120
CBS 144743 = CPC 25802T Glycine max South Africa MH484757 MH484848 MH484939 MH485030 MH485121
CBS 144744 = CPC 25803 G. max South Africa MH484758 MH484849 MH484940 MH485031 MH485122
F. foetens CBS 120665 Nicotiana tabacum Iran MH484736 MH484827 MH484918 MH485009 MH485100
F. glycines CBS 176.33 = NRRL 36286 Linum usitatissium lini Unknown MH484686 MH484777 MH484868 MH484959 MH485050
CBS 214.49 = NRRL 36356 Unknown Argentina MH484687 MH484778 MH484869 MH484960 MH485051
CBS 200.89 Ocimum basilicum basilici Italy MH484706 MH484797 MH484888 MH484979 MH485070
CBS 144745 = CPC 25804 G. max South Africa MH484759 MH484850 MH484941 MH485032 MH485123
CBS 144746 = CPC 25808T G. max South Africa MH484760 MH484851 MH484942 MH485033 MH485124
F. gossypinum CBS 116611 Gossypium hirsutum vasinfectum Ivory Coast MH484725 MH484816 MH484907 MH484998 MH485089
CBS 116612 G. hirsutum vasinfectum Ivory Coast MH484726 MH484817 MH484908 MH484999 MH485090
CBS 116613T G. hirsutum vasinfectum Ivory Coast MH484727 MH484818 MH484909 MH485000 MH485091
F. hoodiae CBS 132474T Hoodia gordonii hoodiae South Africa MH484747 MH484838 MH484929 MH485020 MH485111
CBS 132476 H. gordonii hoodiae South Africa MH484748 MH484839 MH484930 MH485021 MH485112
CBS 132477 H. gordonii hoodiae South Africa MH484749 MH484840 MH484931 MH485022 MH485113
F. languescens CBS 645.78 = NRRL 36531T Solanum lycopersicum lycopersici Morocco MH484698 MH484789 MH484880 MH484971 MH485062
CBS 646.78 = NRRL 36532 S. lycopersicum lycopersici Morocco MH484699 MH484790 MH484881 MH484972 MH485063
CBS 413.90 = ATCC 66046 = NRRL 36465 S. lycopersicum lycopersici Israel MH484708 MH484799 MH484890 MH484981 MH485072
CBS 300.91 = NRRL 36416 S. lycopersicum lycopersici The Netherlands MH484709 MH484800 MH484891 MH484982 MH485073
CBS 302.91 = NRRL 36419 S. lycopersicum lycopersici The Netherlands MH484710 MH484801 MH484892 MH484983 MH485074
CBS 872.95 = NRRL 36570 S. lycopersicum radicis-lycopersici Unknown MH484713 MH484804 MH484895 MH484986 MH485077
CBS 119796 = MRC 8437 Z. mays South Africa MH484735 MH484826 MH484917 MH485008 MH485099
F. libertatis CBS 144748 = CPC 25782 Aspalathus sp. South Africa MH484750 MH484841 MH484932 MH485023 MH485114
CBS 144747 = CPC 25788 Aspalathus sp. South Africa MH484751 MH484842 MH484933 MH485024 MH485115
CBS 144749 = CPC 28465T Rock surface South Africa MH484762 MH484853 MH484944 MH485035 MH485126
F. nirenbergiae CBS 129.24 Secale cereale Unknown MH484682 MH484773 MH484864 MH484955 MH485046
CBS 149.25 = NRRL 36261 Musa sp. cubense Unknown MH484683 MH484774 MH484865 MH484956 MH485047
Species Culture accession1 Host/ substrate Special form Origin GenBank accession
cmdA IGS rpb2 tef1 tub2
Table 2 Details of Fusarium strains included in the phylogenetic analyses.
14 Persoonia – Volume 43, 2019
F. nirenbergiae (cont.) CBS 181.32 = NRRL 36303 S. tuberosum USA MH484685 MH484776 MH484867 MH484958 MH485049
CBS 758.68 = NRRL 36546 S. lycopersicum lycopersici The Netherlands MH484695 MH484786 MH484877 MH484968 MH485059
CBS 744.79 = BBA 62355 = NRRL 22549 Passiflora edulis passiflorae Brazil MH484700 MH484791 MH484882 MH484973 MH485064
CBS 127.81 = BBA 63924 = NRRL 36229 Chrysanthemum sp. chrysanthemi USA MH484701 MH484792 MH484883 MH484974 MH485065
CBS 129.81 = BBA 63926 = NRRL 22539 Chrysanthemum sp. chrysanthemi USA MH484703 MH484794 MH484885 MH484976 MH485067
CBS 196.87 = NRRL 26219 Bouvardia longiflora bouvardiae Italy MH484704 MH484795 MH484886 MH484977 MH485068
CBS 840.88T Dianthus caryophyllus dianthi The Netherlands MH484705 MH484796 MH484887 MH484978 MH485069
CBS 115416 = CPC 5307 Agathosma betulina South Africa MH484720 MH484811 MH484902 MH484993 MH485084
CBS 115417 = CPC 5306 A. betulina South Africa MH484721 MH484812 MH484903 MH484994 MH485085
CBS 115419 = CPC 5308 A. betulina South Africa MH484722 MH484813 MH484904 MH484995 MH485086
CBS 115424 = CPC 5312 A. betulina South Africa MH484724 MH484815 MH484906 MH484997 MH485088
CBS 123062 = GJS 91-17 Tulip roots USA MH484737 MH484828 MH484919 MH485010 MH485101
CBS 130300 = NRRL 26368 Amputated human toe USA MH484743 MH484834 MH484925 MH485016 MH485107
CBS 130301 = NRRL 26374 Human leg ulcer USA MH484744 MH484835 MH484926 MH485017 MH485108
CBS 130303 S. lycopersicum radicis-lycopersici USA MH484741 MH484832 MH484923 MH485014 MH485105
CPC 30807 South Africa MH484768 MH484859 MH484950 MH485041 MH485132
F. odoratissimum CBS 794.70 = BBA 11103 = NRRL 22550 Albizzia julibrissin perniciosum Iran MH484696 MH484787 MH484878 MH484969 MH485060
CBS 102030 M. sapientum cv. Pisang mas cubense Malaysia MH484716 MH484807 MH484898 MH484989 MH485080
CBS 130310 = NRRL 25603 Musa sp. cubense Australia MH484740 MH484831 MH484922 MH485013 MH485104
F. oxysporum CBS 221.49 = IHEM 4508 = NRRL 22546 Camellia sinensis medicaginis South East Asia MH484690 MH484781 MH484872 MH484963 MH485054
CBS 144134ET S. tuberosum Germany MH484771 MH484862 MH484953 MH485044 MH485135
CBS 144135 S. tuberosum Germany MH484772 MH484863 MH484954 MH485045 MH485136
CPC 25822 Protea sp. South Africa MH484761 MH484852 MH484943 MH485034 MH485125
F. pharetrum CBS 144750 = CPC 30822 Aliodendron dichotomum South Africa MH484769 MH484860 MH484951 MH485042 MH485133
CBS 144751 = CPC 30824T A. dichotomum South Africa MH484770 MH484861 MH484952 MH485043 MH485134
F. trachichlamydosporum CBS 102028 = NRRL 36117 M. sapientum cv. Pisang awak legor cubense Malaysia MH484715 MH484806 MH484897 MH484988 MH485079
F. triseptatum CBS 258.50 = NRRL 36389T Ipomoea batatas batatas USA MH484691 MH484782 MH484873 MH484964 MH485055
CBS 116619 G. hirsutum vasinfectum Ivory Coast MH484728 MH484819 MH484910 MH485001 MH485092
CBS 119665 Sago starch Papua New Guinea MH484734 MH484825 MH484916 MH485007 MH485098
CBS 130302 = NRRL 26360 = FRC 755 Human eye USA MH484742 MH484833 MH484924 MH485015 MH485106
F. udum CBS 177.31 Digitaria eriantha South Africa MH484684 MH484775 MH484866 MH484957 MH485048
F. veterinarium CBS 109898 = NRRL 36153T Shark peritoneum The Netherlands MH484717 MH484808 MH484899 MH484990 MH485081
CBS 117787 Swab sample near filling apparatus The Netherlands MH484730 MH484821 MH484912 MH485003 MH485094
CBS 117790 Swab sample near filling apparatus The Netherlands MH484731 MH484822 MH484913 MH485004 MH485095
CBS 117791 Pasteurized milk-based product The Netherlands MH484732 MH484823 MH484914 MH485005 MH485096
CBS 117792 Pasteurized milk-based product The Netherlands MH484733 MH484824 MH484915 MH485006 MH485097
NRRL 54984 Mouse mucosa USA MH484763 MH484854 MH484945 MH485036 MH485127
NRRL 54996 Little blue penguin foot USA MH484764 MH484855 MH484946 MH485037 MH485128
NRRL 62542 Unknown animal faeces USA MH484765 MH484856 MH484947 MH485038 MH485129
NRRL 62545 Endoscope of veterinary clinic USA MH484766 MH484857 MH484948 MH485039 MH485130
NRRL 62547 Canine stomach USA MH484767 MH484858 MH484949 MH485040 MH485131
Fusarium sp. CBS 128.81 = BBA 63925 = NRRL 36233 Chrysanthemum sp. chrysanthemi USA MH484702 MH484793 MH484884 MH484975 MH485066
CBS 680.89 = NRRL 26221 Cucumis sativus cucurbitacearum The Netherlands MH484707 MH484798 MH484889 MH484980 MH485071
CBS 130323 Human nail Australia MH484745 MH484836 MH484927 MH485018 MH485109
1 ATCC: American Type Culture Collection, USA; BBA: Biologische Bundesanstalt für Land- und Forstwirtschaft, Berlin-Dahlem, Germany; CBS: Westerdijk Fungal Biodiverity Institute (WIFB), Utrecht, The Netherlands; CPC: Collection of P.W. Crous; DSM: Deutsche Sammlung von
Mikroorganismen und Zellkulturen GmbH, Braunschweig, Germany; FRC: Fusarium Research Center, Penn State University, Pennsylvania; GJS: Collection of Gary J. Samuels; IHEM: Institute of Hygiene and Epidemiology-Mycology Laboratory, Brussels, Belgium; MRC: National
Research Institute for Nutritional Diseases, Tygerberg, South Africa; NRRL: Agricultural Research Service Culture Collection, USA; PD: Collection of the Dutch National Plant Protection Organization, Wageningen, The Netherlands. T Ex-type culture; ETEpitype.
Species Culture accession1 Host/ substrate Special form Origin GenBank accession
cmdA IGS rpb2 tef1 tub2
Table 2 (cont.)
15
L. Lombard et al.: Epitypification of Fusarium oxysporum
RESULTS
Isolates
A total of 23 fusarium-like isolates were obtained from the
symptomatic tissues of the potato tubers. Of these, six isolates
displayed typical F. oxysporum-like phenotypes, of which two
(CBS 144134 and CBS 144135) were selected for further study.
Phylogenetic analyses
Approximately 500–650 bases were determined for cmdA,
tef1 and tub2, 880 bases for rpb2 and 2 650 bases for IGS.
Sequence comparisons of the IGS, rpb2 and tef1 gene regions
generated in this study, against those in the Fusarium-ID (http://
isolate.fusariumdb.org/blast.php) and Fusarium-MLST (http://
www.westerdijkinstitute.nl/fusarium/) databases revealed that
all isolates included in this study belonged to the FOSC. The
congruency analysis revealed no conflict between the cmdA,
rpb2, tef1 and tub2 sequence datasets and were therefore
combined. However, the IGS sequence dataset revealed major
conflict with several included taxa resolving into single lineages
due to the large number of ambiguous regions in this gene
region. Therefore, the IGS sequences were excluded from
further analyses.
For the BI and ML analyses, a K80 model for cmdA, an HKY+
G+I model for rpb2, an HKY+G for tef1 and SYM+I+G model for
tub2 were selected and incorporated into the analyses. The ML
tree topology confirmed the tree topologies obtained from the BI
and MP analyses, and therefore, only the ML tree is presented.
The combined four loci sequence dataset included 89 ingroup
taxa with F. foetens (CBS 120665) and F. udum (CBS 177.31)
as outgroup taxa. The dataset consisted of 2 679 characters
including gaps. Of these characters, 2 291 were constant, 211
parsimony-uninformative and 177 parsimony-informative. The
BI lasted for 1.2 M generations, and the consensus tree and
posterior probabilities (PP) were calculated from 8 814 trees
left after 2 937 were discarded as the ‘burn-in’ phase. The MP
analysis yielded 1 000 trees (TL = 574; CI = 0.747; RI = 0.858;
RC = 0.641) and a single best ML tree with -InL = 7353.014512
(Fig. 1).
In the phylogenetic tree (Fig. 1) the ingroup taxa resolved into
eight clades (I–VIII). Of these, Clades I, II, IV and VI represent
single well- (ML & MP-BS ≥ 75– 95 %; PP ≥ 0.95–0.98) to
highly (ML & MP-BS ≥ 96 %; PP ≥ 0.99–1.0) supported clades,
whereas Clades III, V, VII and VIII displayed substantial sub-
structure. Clade III included eight well- to highly supported sub-
clades as well as two single lineages. Sequence comparisons of
the rpb2 and tef1 sequences with those generated by Maryani
et al. (2019) revealed that both single lineages represented
F. duo septatum (CBS 102026) and F. tradichlamydosporum
(CBS 102028), respectively. Similarly, the subclade that include
isolates CBS 620.72, CBS 130304, CBS 130308 and CBS
131393 represent F. cugenangense. Both Clades V and VIII
resolved two subclades in each, and Clade VII included three
subclades. The phylogenetic relationships between Clades
I–VIII and their underlying subclades are further discussed in
the notes in the Taxonomy section.
The PHI tests revealed that no evidence of recombination
(fW = 0.43; Fig. 2a) was detected between each Clade (I–VIII)
and their underlining subclades. Similarly, the genealogical
exclusivity of the subclades in Clades III (fW = 0.43; Fig. 2b)
and VII (fW = 1.0; Fig. 2d), as well as between Clades IV–VIII
(fW = 0.06; Fig. 2c) was also confirmed. The basal subclade in
Clade VIII (fW = 0.031; Fig. 2c), however, showed significant
evidence for recombination among all isolates included.
Taxonomy
In this section we provide a new (emended) description of
F. oxysporum and designate an epitype for this species. The
following species are also recognised as new within the FOSC,
based on phylogenetic inference and morphological compari-
sons. Isolates CBS 128.81, CBS 680.89 and CBS 130323 in
Clade III are not treated further as these were sterile.
Fusarium callistephi L. Lombard & Crous, sp. nov. — Myco-
Bank MB826833; Fig. 3
Etymology. Name refers to the plant genus Callistephus from which this
fungus was isolated.
Typus. NetherlaNds, Oostenbrink, from Callistephus chinensis, 28 Feb.
1953, collector unknown (holotype CBS H-23608 designated here, culture
ex-type CBS 187.53).
Conidiophores carried on the aerial mycelium 60–110 µm
tall, unbranched or sparingly branched, bearing terminal or
intercalarily monophialides, often reduced to single phialides;
aerial phialides subulate to subcylindrical, smooth- and thin-
walled, 2–23 × 3 4 µm, periclinal thickening inconspicuous
or absent; aerial conidia forming small false heads on the tips
of the phialides, hyaline, ellipsoidal to falcate, smooth- and
thin-walled, 0–1-septate; 0-septate conidia: (6–)7–11(–14) ×
2–3 µm (av. 9 × 3 µm); 1-septate conidia: (13–)14–18 (–20) ×
2–4 µm (av. 16 × 3 µm). Sporodochia pale luteous to pale rosy
vinaceous, formed abundantly on carnation leaves. Conidio-
phores in sporodochia verticillately branched and densely
packed, consisting of a short, smooth- and thin-walled stipe,
4–7 × 2–4 µm, bearing apical whorls of 2–3 monophialides or
rarely as single lateral monophialides; sporodochial phialides
subulate to subcylindrical, 9–13 × 3– 4 µm, smooth- and thin-
walled, sometimes showing a reduced and flared collarette.
Sporodochial conidia falcate, curved dorsiventrally with almost
parallel sides tapering slightly towards both ends, with a blunt
to papillate, curved apical cell and a blunt to foot-like basal cell,
3–4(–5)-septate, hyaline, smooth- and thin-walled; 3-septate
conidia: (28–)33 –39(–40) × 3–5 µm (av. 36 × 4 µm); 4-septate
conidia: (30–)35 –41(– 42) × 3 5 µm (av. 38 × 4 µm); 5-septate
conidia: 36–44 (–47) × 4– 5 µm (av. 40 × 5 µm). Chlamydo-
spores not observed.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 2.9–4.2 mm /d at 24 °C. Colony surface
white to pale vinaceous, floccose with abundant aerial myce-
lium; colony margins irregular, lobate, serrate or filiform. Odour
absent. Reverse colourless, lacking diffusible pigment. On SNA,
hyphae hyaline, smooth-walled, lacking chlamydospores, aerial
mycelium sparse with moderate sporulation on the medium
surface. On CLA, aerial mycelium sparse with abundant pale
luteous to pale rosy vinaceous sporodochia forming on the
carnation leaves.
Additional material examined. south africa, Western Cape Province,
Piketberg, from Agathosma betulina, 2001, K. Lubbe, CBS 115423 = CPC
5311.
Notes — Fusarium callistephi formed a highly-supported
subclade in Clade III, closely related to F. cugenangense,
F. elaeidis and the untreated Fusarium clade. This species (co-
nidia 3–4 (–5)-septate) can be distinguished from F. cugen ang-
ense (conidia 3–6-septate; Maryani et al. 2019) and F. elaeidis
((1–)3 5-septate) based on septation of their macroconidia.
Additionally, F. cugenangense produces up to 3-septate micro-
conidia, a feature not seen in either F. callistephi or F. elaeidis.
Fusarium elaeidis readily formed polyphialidic conidiogenous
cells on the aerial mycelium, not seen in F. callistephi.
16 Persoonia – Volume 43, 2019
CBS 132477 (hoodiae)
CBS 144739
CBS 116619 (vasinfectum)
CBS 144743
CBS 128.81 (chrysanthemi)
CBS 221.49 (medicaginis)
CBS 187.53 (callistephi)
CBS 132476 (hoodiae)
Fusarium udum CBS 177.31
CPC 25822
CBS 102028 (cubense)
CBS 144742
CBS 144747
CBS 131393
CBS 130302
CBS 130323
CBS 102030 (cubense)
CBS 214.49 (dianthi)
CBS 680.89 (cucurbitacearum)
CBS 144134
CBS 255.52 (elaeidis)
CBS 115423
CBS 102026 (cubense)
CBS 144746
CBS 116611 (vasinfectum)
CBS 218.49 (elaeidis)
CBS 144749
CBS 144135
CBS 176.33 (lini)
CBS 144738
CBS 258.50 (batatas)
CBS 116613 (vasinfectum)
CBS 130308
CBS 217.49 (elaeidis)
CBS 144745
CBS 620.72 (gladioli)
CBS 119665
CBS 144741
CBS 130310 (cubense)
CBS 144740
CBS 130304 (vasinfectum)
CBS 132474 (hoodiae)
CBS 794.70 (perniciosum)
CBS 144748
CBS 116612 (vasinfectum)
Fusarium foetens CBS 120665
CBS 144744
100/50/1.0
80/85/1.0
-/-/0.99
-/-/0.99
97/96/1.0
78/66/0.95
-/-/0.99
99/78/0.99
-/-/1.0
66/-/-
89/66/1.0
84/71/1.0
100/50/1.0
-/-/0.96
95/96/1.0
100/100/1.0
96/96/1.0
79/-/1.0
78/68/1.0
63/61/-
100/99/1.0
52/54/-
84/-/1.0
77/73/1.0
99/98/1.0
100/99/1.0
x3
x3
x10
63/-/-
x3
Fusarium odoratissimum
Fusarium libertatis
Fusarium tradichlamydosporum
Fusarium duoseptatum
Fusarium fabacearum
Fusarium carminascens
Fusarium glycines
Fusarium gossypinum
Fusarium sp.
Fusarium callistephi
Fusarium elaeidis
Fusarium cugenangense
Fusarium hoodiae
Fusarium triseptatum
Fusarium oxysporum
I
II
III
IV
V
Fig. 1 The ML consensus tree inferred from the combined cmdA, rpb2, tef1 and tub2 sequence alignment. Thickened branches indicate branches present
in the ML, MP and Bayesian consensus trees. Support values (ML & MP bootstrap and posterior probability values) are indicated at the branches. The scale
bar indicates 0.02 expected changes per site. Clade numbers are provided on the right of the tree and these are used for reference in the treatment of the
species. The tree is rooted to F. foetens (CBS 120665) and F. udum (CBS 177.31). Epi- and ex-type strains are indicated in bold.
17
L. Lombard et al.: Epitypification of Fusarium oxysporum
0.02
CBS 646.78 (lycopersici)
CBS 129.24
CBS 117790
CBS 302.91 (lycopersici)
CBS 758.68 (lycopersici)
CBS 117461
CBS 141.95
CBS 114899
CBS 238.94 (meniscoideum)
CBS 123062
CBS 144751
CBS 130303 (radicis-lycopersici)
CBS 115416
CBS 840.88 (dianthi)
CBS 117787
CBS 127.81 (chrysanthemi)
CBS 300.91 (lycopersici)
CBS 115424
CBS 130300
NRRL 62542
CBS 109898
CBS 144750
NRRL 62545
NRRL 62547
CBS 645.78 (lycopersici)
CPC 30807
CBS 117791
CBS 247.61 (matthiolae)
CBS 181.32
CBS 129.81 (chrysanthemi)
CBS 111552
CBS 130301
CBS 119796
CBS 872.95 (radicis-lycopersici)
CBS 196.87 (bouvardiae)
CBS 413.90 (lycopersici)
CBS 115419
CBS 149.25 (cubense)
CBS 115417
NRRL 54996
NRRL 54984
CBS 117792
CBS 744.79 (passiflorae)
100/98/1.0
87/86/1.0
93/81/1.0
72/52/-
73/67/0.99
99/94/1.0
93/75/0.98
98/86/1.0
84/71/0.95
Fusarium languescens
Fusarium pharetrum
Fusarium contaminatum
Fusarium veterinarium
Fusarium curvatum
Fusarium nirenbergiae
VI
VII
VIII
Fig. 1 (cont.)
18 Persoonia – Volume 43, 2019
Fig. 2 Splitgraphs showing the results of the pairwise homoplasy index (PHI) test of newly described taxa using both LogDet transformation and splits decomposition. PHI test results (fW) < 0.05 indicate significant recombination
within the dataset. a. Representatives of all phylogenetic species resolved in this study; b. phylogenetic species in Clade III; c. phylogenetic species in Clades IV– VIII; d. phylogenetic species in Clade VII; e. isolates representing
F. nirenbergiae.
Fusarium udum CBS 177.31
Fusarium foetens CBS 120665
CBS 646.78
CBS 645.78
CBS 144751
CBS 144750
CBS 111552
CBS 117787
CBS 141.95
CBS 238.94
CBS 840.88
CBS 115416
CBS 144135
CBS 144134
CBS 144749
CBS 144747
Fusarium tradichlamydosporum CBS 102028
Fusarium duoseptatum CBS 102026
CBS 144746
CBS 200.89
CBS 144739
CBS 144738
CBS 144743
CBS 144744
CBS 116611
CBS 116612
CBS 115423
CBS 187.53
CBS 131393
CBS 128.81
CBS 680.89
CBS 258.50
CBS 116619
CBS 114899
CBS 109898
CBS 132474
CBS 132476
CBS 218.49
CBS 217.49
CBS 130304
Fusarium triseptatum
Fusarium sp.
Fusarium cugenangense
Fusarium elaeidis
Fusarium callistephi
Fusarium gossypinum
Fusarium fabacearum
Fusarium carminascens
Fusarium glycines
Fusarium libertatis
Fusarium hoodiae
Fusarium oxysporum
Fusarium nirenbergiae
Fusarium curvatum
Fusarium veterinarium
Fusarium phraretrum
Fusarium contaminatum
Fusarium languescens
CBS 217.49
CBS 187.53
CBS 115423
CBS 116612
CBS 116611
CBS 144738
CBS 144643
CBS 200.89
CBS144746
CBS 102028 Fusarium tradichlamydosporum
CBS 102026 Fusarium duoseptatum
CBS 680.89
CBS 128.81
CBS 131393
CBS 218.49
CBS 144739
CBS 144744
CBS 130304
0.001
Fusarium sp.
Fusarium cugenangense
Fusarium elaeidis
Fusarium callistephi
Fusarium gossypinum
Fusarium carminascens
Fusarium fabacearum
Fusarium glycines
CBS 258.50
CBS 144134
CBS 144135
CBS 132476
CBS 144751
CBS 144750
CBS 141.95
CBS 238.94
CBS 840.88
CBS 115416
CBS 116619
CBS 111552
CBS 645.78
CBS 132474
Fusarium oxysporum
Fusarium triseptatum
Fusarium nirenbergii
Fusarium curvatum
CBS 117787
CBS 109898
Fusarium veterinatum
Fusarium hoodiae
CBS 646.78
Fusarium languescens
Fusarium phraretrum
CBS 114899
Fusarium contaminatum
CBS 109898
CBS 144750
CBS 144751
CBS 111552
Fusarium phraretrum
CBS 114899
Fusarium contaminatum
CBS 117787
Fusarium veterinarium
CBS 129.24
CBS 115424
CPC 30807
CBS 130301
CBS 149.25
CBS 181.32
CBS 123062
CBS 115416
CBS 115417
CBS 758.68
CBS 840.88
CBS 130300
CBS 744.79
CBS 129.81
CBS 130303
CBS 196.87
CBS 115419
CBS 127.81
1.0E-4
W= 0.031
e
1.0E-4
W= 1.0
d
0,001
W= 0.43
a
W= 0.43
b
0.001
W= 0.06
c
19
L. Lombard et al.: Epitypification of Fusarium oxysporum
Fusarium carminascens L. Lombard, Crous & Lampr., sp.
nov. — MycoBank MB826835; Fig. 4
Etymology. Name refers to the almost carmine exudates this fungus
produces in its aerial mycelium when grown on PDA.
Typus. south africa, KwaZulu-Natal Province, from Zea mays, 2008,
S.C. Lamprecht (holotype CBS H-23609 designated here, culture ex-type
CBS 144738 = CPC 25800).
Conidiophores carried on the aerial mycelium 35 55 µm tall,
unbranched or sparingly branched, bearing terminal or interca-
larily phialides, often reduced to single phialides; aerial phialides
mono- and polyphialidic, subulate to subcylindrical, smooth- and
thin-walled, 8–18 × 3–4 µm, periclinal thickening inconspicu-
ous or absent; aerial conidia forming small false heads on the
tips of the phialides, hyaline, ellipsoidal to falcate, smooth- and
thin-walled, 0–1-septate; 0-septate conidia: (5–)7–11(–12) ×
2–3(–4) µm (av. 9 × 3 µm); 1-septate conidia: (12–)13–15(–18)
× 2– 4 µm (av. 14 × 3 µm). Sporodochia bright orange, formed
abundantly on carnation leaves. Conidiophores in sporodochia
verticillately branched and densely packed, consisting of a
short, smooth- and thin-walled stipe, 4–9 × 2–4 µm, bearing
apical whorls of 2–3 monophialides or rarely as single lateral
monophialides; sporodochial phialides subulate to subcylin-
drical, 5–13 × 2 4 µm, smooth- and thin-walled, sometimes
showing a reduced and flared collarette. Sporodochial conidia
falcate, curved dorsiventrally with almost parallel sides tapering
slightly towards both ends, with a blunt to papillate, curved api-
cal cell and a blunt to foot-like basal cell, (2–)3 –4(–5)-septate,
hyaline, smooth- and thin-walled; 2-septate conidia: 16–19 ×
3–4 µm (av. 18 × 3 µm); 3-septate conidia: (21–)26–36(–40) ×
3–5 µm (av. 31 × 4 µm); 4-septate conidia: (31–)33–43(–44) ×
4–5 µm (av. 38 × 4 µm); 5-septate conidia: 45– 51 × 4 µm (av.
48 × 4 µm). Chlamydospores globose to subglobose, formed
terminally, 48 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.0 mm/d at 24 °C. Colony surface
vinaceous purple to livid purple, floccose with abundant aerial
mycelium which produce an almost carmine exudate; colony
margins irregular, lobate, serrate or filiform. Odour absent.
Reverse dark livid to livid purple, lacking diffusible pigment.
On SNA, hyphae hyaline, smooth-walled, with abundant chla-
mydospores, aerial mycelium sparse with abundant sporulation
on the medium surface. On CLA, aerial mycelium sparse with
abundant bright orange sporodochia forming on the carnation
leaves.
Additional materials examined. south africa, KwaZulu-Natal Province,
from Zea mays, 2008, S.C. Lamprecht, CBS 144739 = CPC 25792, CBS
144740 = CPC 25793, CBS 144741 = CPC 25795.
Notes — Fusarium carminascens formed a well-supported
subclade in Clade III, closely related to F. fabacearum and
F. glycines. This species produced an almost carmine coloured
exudate in its aerial mycelium, a feature not observed in any of
the other strains studied here. Furthermore, F. carminascens
produces polyphialidic conidiogenous cells on its aerial myce-
lium, not observed in F. fabacearum or F. glycines.
Fig. 3 Fusarium callistephi (ex-type culture CBS 187.53). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c. conidiophores on surface of carnation leaf; d. sporodochia on carnation leaves; e– i. conidiophores and phialides on aerial my-
celium; j–k. sporodochia and sporodochial conidiophores; l. aerial conidia (microconidia); m. sporodochial conidia (macroconidia). — Scale bars: e– m = 10 µm.
20 Persoonia – Volume 43, 2019
Fusarium contaminatum L. Lombard & Crous, sp. nov. — Myco-
Bank MB826836; Fig. 5
Etymology. Name refers to the fact that this fungus was isolated from
contaminated food products.
Typus. GermaNy, Schluchtern, from pasteurized chocolate milk, Apr. 2004,
J. Houbraken (holotype CBS H-23610 designated here, culture ex-type CBS
114899).
Conidiophores carried on the aerial mycelium 15 85 µm tall,
unbranched or branched, bearing a single terminal or a whorl
of 2–4 monophialides or intercalarily monophialides, often
reduced to single phia lides; aerial phialides subulate to sub-
cylindrical, smooth- and thin-walled, 7–22 × 25 µm, periclinal
thickening inconspicuous or absent; aerial conidia forming small
false heads on the tips of the phialides, hyaline, ellipsoidal
to falcate, smooth- and thin-walled, 0–1-septate; 0-septate
conidia: 5–9 (–11) × 2–4 µm (av. 7 × 3 µm); 1-septate conidia:
(9–)10–14 (–17) × 2–4 µm (av. 12 × 3 µm). Sporodochia bright
orange, formed sparsely on carnation leaves. Conidiophores in
sporodochia verticillately branched and densely packed, con-
sisting of a short, smooth- and thin-walled stipe, 7–13 × 4 µm,
bearing apical whorls of 2–3 monophialides or rarely as single
lateral monophialides; sporodochial phialides subulate to sub-
cylindrical, 4–9 × 2–3 µm, smooth- and thin-walled, sometimes
Fig. 4 Fusarium carminascens (ex-type culture CBS 144738). a b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e– f. sporodochia on carnation leaves; g– j. conidiophores and phialides on
aerial mycelium; g– h. monophialides; i–j. polyphialides; k –l. chlamydospores; m– p. sporodochia and sporodochial conidiophores; o–p. phialides of sporo-
dochial conidiophores; q. aerial conidia (microconidia); r. sporodochial conidia (macroconidia). — Scale bars: g– r = 10 µm.
21
L. Lombard et al.: Epitypification of Fusarium oxysporum
showing a reduced and flared collarette. Sporodochial conidia
falcate, curved dorsiventrally with almost parallel sides tapering
slightly towards both ends, with a blunt to papillate, curved api-
cal cell and a blunt to foot-like basal cell, (2–)3-septate, hyaline,
smooth- and thin-walled; 2-septate conidia: (14–)15–17 × 3–4
µm (av. 16 × 3 µm); 3-septate conidia: (18 –)20– 26(–28) × 3–5
µm (av. 23 × 4 µm). Chlamydospores not observed.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.5 mm/d at 24 °C. Colony surface
white to pale vinaceous, floccose with abundant aerial myce-
lium; colony margins irregular, lobate, serrate or filiform. Odour
absent. Reverse rosy vinaceous, lacking diffusible pigment. On
SNA, hyphae hyaline, smooth-walled, lacking chlamydospores,
aerial mycelium sparse with abundant sporulation on the me-
dium surface. On CLA, aerial mycelium sparse with abundant
orange sporodochia forming on the carnation leaves.
Additional materials examined. NetherlaNds, from pasteurized fruit juice,
date and collector unknown, CBS 111552; from tetra pack with milky nutrition,
2005, collector unknown, CBS 117461.
Notes — Fusarium contaminatum formed a highly-supported
subclade in Clade VII, closely related to F. pharetrum and
F. veterinarium. This species produces small, 2 3-septate
macroconidia, whereas F. pharetrum produces much larger,
3(– 4)-septate macroconidia and F. veterinarium produces
slightly smaller, 1–(2–)3-septate macroconidia. None of these
three species produced any chlamydospores on SNA.
Fusarium curvatum L. Lombard & Crous, sp. nov. — Myco-
Bank MB826837; Fig. 6
Etymology. Name refers to the strongly curved sporodochial conidia pro-
duced by this fungus.
Typus. NetherlaNds, from Beaucarnia sp., 1994, J.W. Veenbaas-Rijks (holo-
type CBS H-23611 designated here, culture ex-type CBS 238.94 = NRRL
26422 = PD 94/184).
Conidiophores carried on the aerial mycelium 25 56 µm tall,
unbranched or sparingly branched, bearing terminal or interca-
Fig. 5 Fusarium contaminatum (ex-type culture CBS 114899). a– b. Colony on PDA; a. Surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c d. conidiophores on surface of carnation leaf; e–f. sporodochia on carnation leaves; g– k. conidiophores and phialides
on aerial mycelium; l. false head carried on phialide on aerial mycelium; m– p. sporodochia and sporodochial conidiophores; q. aerial conidia (microconidia);
r. sporodochial conidia (macroconidia). — Scale bars: g– l, q– r = 10 µm; m–p = 20 µm.
22 Persoonia – Volume 43, 2019
larily phialides, often reduced to single phialides or as phialidic
pegs; aerial phialides mono- and polyphialidic, subulate to sub-
cylindrical, smooth- and thin-walled, 3–30 × 2–5 µm, periclinal
thickening inconspicuous or absent; aerial conidia forming small
false heads on the tips of the phialides, hyaline, ellipsoidal
to falcate, smooth- and thin-walled, 0–1-septate; 0-septate
conidia: (4–)5 9(–11) × 2–4 µm (av. 7 × 3 µm); 1-septate
conidia: (10–)11–13 × 2 4 µm (av. 12 × 3 µm). Sporodochia
orange, formed abundantly on carnation leaves. Conidiophores
in sporodochia verticillately branched and densely packed,
consisting of a short, smooth- and thin-walled stipe, 8–10 × 2–4
µm, bearing apical whorls of 2–3 monophialides or rarely as
single lateral monophialides; sporodochial phialides subulate to
subcylindrical, 8–22 × 2– 4 µm, smooth- and thin-walled, some-
times showing a reduced and flared collarette. Sporodochial
conidia falcate, strongly curved or curved dorsiventrally with
almost parallel sides tapering slightly towards both ends, with
a blunt to papillate, curved apical cell and a blunt to foot-like
basal cell, (2–)3 5-septate, hyaline, smooth- and thin-walled;
2-septate conidia: (15–)16– 22(– 23) × 3 4 µm (av. 19 × 3 µm);
Fig. 6 Fusarium curvatum (ex-type culture CBS 238.94). a–b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e–f. sporodochia on carnation leaves; g i. conidiophores, monophialides and
polyphialides (arrows) on aerial mycelium; j. phialidic pegs on aerial mycelium; k– o. sporodochia and sporodochial conidiophores; p. aerial conidia (microco-
nidia); q. sporodochial conidia (macroconidia). — Scale bars: g– i, n = 20 µm; j, o–q = 10 µm, k– m = 50 µm.
23
L. Lombard et al.: Epitypification of Fusarium oxysporum
3-septate conidia: (18–)27– 39(–41) × 3– 5 µm (av. 33 × 4 µm);
4-septate conidia: (34–)37– 43(–46) × 3– 5 µm (av. 40 × 4 µm);
5-septate conidia: (30–)38 –46(–51) × 3–5 µm (av. 42 × 4 µm).
Chlamydospores not observed.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.5 mm/d at 24 °C. Colony surface
pale vinaceous to rosy vinaceous, floccose with abundant aerial
mycelium; colony margins irregular, lobate, serrate or filiform.
Odour absent. Reverse pale vinaceous, lacking diffusible pig-
ment. On SNA, hyphae hyaline, smooth-walled, lacking chlamy-
do spores, aerial mycelium sparse with abundant sporulation
on the medium surface. On CLA, aerial mycelium sparse with
abundant orange sporodochia forming on the carnation leaves.
Additional materials examined. GermaNy, Berlin-Dahlem, from Matthiola
incana, Feb. 1957, W. Gerlach, CBS 247.61 = BBA 8398 = DSM 62308 =
NRRL 22545. – NetherlaNds, from Hedera helix, 1994, J.W. Veenbaas-Rijks,
CBS 141.95 = NRRL 36251 = PD 94/1518.
Notes — Fusarium curvatum formed a highly-supported sub-
clade in Clade VIII, closely related to F. nirenbergiae. This
species produces strongly curved 3-septate macroconidia and
aerial polyphialidic conidiogenous cells, distinguishing it from
F. nirenbergiae. Additionally, F. curvatum failed to produce any
chlamydospores on SNA, whereas F. nirenbergiae produced
abundant chlamydospores.
Fusarium elaeidis L. Lombard & Crous, sp. nov. — MycoBank
MB826838; Fig. 7
Etymology. Name refers to the host plant genus Elaeis, from which this
fungus was first isolated.
Typus. Zaire, from Elaeis sp., 1949, T. Gogoi (holotype CBS H-23612
designated here, culture ex-type CBS 217.49 = NRRL 36358).
Conidiophores carried on the aerial mycelium 25 65 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily phialides, often reduced to single phialides or as phia-
lidic pegs; aerial phialides mono- and polyphialidic, subulate
to subcylindrical, smooth- and thin-walled, 3–14 × 3– 4 µm,
periclinal thickening inconspicuous or absent; aerial conidia
forming small false heads on the tips of the phialides, hyaline,
ellipsoidal to falcate, smooth- and thin-walled, 0–1-septate;
0-septate conidia: 6–10(–13) × 23 µm (av. 8 × 3 µm); 1-sep-
tate conidia: (9–)11–15(–17) × 2–4(–5) µm (av. 13 × 3 µm).
Sporodochia pale rosy vinaceous to orange, formed abundantly
on carnation leaves. Conidiophores in sporodochia verticillately
Fig. 7 Fusarium elaeidis (ex-type culture CBS 217.49). a –b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c d. conidiophores on surface of carnation leaf; e–f. sporodochia on carnation leaves; g. false head carried on a phialidic peg
on aerial mycelium; h. phialidic peg; i– j. conidiophores and phialides on aerial mycelium; j. polyphialide; k –l. sporodochia and sporodochial conidiophores;
m. aerial conidia (microconidia); n. sporodochial conidia (macroconidia). — Scale bars: g– n = 10 µm.
24 Persoonia – Volume 43, 2019
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 3–9 × 2 3 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 8–12 × 2–4
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a blunt
to foot-like basal cell, (1–)3 5-septate, hyaline, smooth- and
thin-walled; 1-septate conidia: (14–)15– 25(– 32) × 2 4 µm (av.
20 × 3 µm); 2-septate conidia: (17–)19 25 × 3 4 µm (av. 22
× 4 µm); 3-septate conidia: (22–) 3040(– 46) × (2–)34 µm
(av. 35 × 4 µm); 4-septate conidia: (34 –)36– 40(–43) × 3 – 5 µm
(av. 38 × 4 µm); 5-septate conidia: (36 –)37–43(–50) × 3 5 µm
(av. 40 × 4 µm). Chlamydospores not observed.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 2.6–3.4 mm /d at 24 °C. Colony surface
pale rosy vinaceous grey, floccose with abundant aerial my-
celium; colony margins irregular, lobate, serrate or filiform.
Odour absent. Reverse pale rosy vinaceous, lacking diffusible
pigment. On SNA, hyphae hyaline, smooth-walled, lacking chla-
mydospores, aerial mycelium sparse with abundant sporulation
on the medium surface. On CLA, aerial mycelium sparse with
abundant pale rosy vinaceous to orange sporodochia forming
on the carnation leaves.
Additional materials examined. Zaire, from Elaeis sp., 1949, T. Gogoi,
CBS 218.49 = NRRL 36359. – uNkNowN locality, from Elaeis guineensis,
1952, J. Fraselle, CBS 255.52 = NRRL 36386.
Notes — Fusarium elaeidis formed a highly-supported sub-
clade in Clade III, closely related to F. callistephi, F. cugenan-
gense and the untreated Fusarium clade. See notes under
F. callistephi for distinguishing morphological features.
Fusarium fabacearum L. Lombard, Crous & Lampr., sp. nov.
— MycoBank MB826839; Fig. 8
Etymology. Name refers to the plant family, Fabaceae, which includes
the plant host Glycine max from which this fungus was first isolated.
Typus. south africa, North West Province, from Glycine max, 2010, S.C.
Lamprecht (holotype CBS H-23613 designated here, culture ex-type CBS
144743 = CPC 25802).
Conidiophores carried on the aerial mycelium 25 50 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
11–15 × 3– 4 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (4–) 5– 9(–13) × 23 µm (av. 7 ×
Fig. 8 Fusarium fabacearum (ex-type culture CBS 144743). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c. conidiophores on surface of carnation leaf; d. sporodochia on carnation leaves; e. false head carried on a phialide on
aerial mycelium; f– h. conidiophores and phialides on aerial mycelium; i–k. sporodochia and sporodochial conidiophores; l. chlamydospore; m. aerial conidia
(microconidia); n. sporodochial conidia (macroconidia). — Scale bars: e– h, k–n = 10 µm; i– j = 50 µm.
25
L. Lombard et al.: Epitypification of Fusarium oxysporum
3 µm); 1-septate conidia: (12–)13 –15(–16) × 3– 4 µm (av. 14 ×
3 µm). Sporodochia pale luteous to orange, formed abundantly
on carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 4–7 × 3 µm, bearing apical whorls of 2 3
monophialides or rarely as single lateral monophialides; sporo-
dochial phialides subulate to subcylindrical, 7–10 × 2 4 µm,
smooth- and thin-walled, sometimes showing a reduced and
flared collarette. Sporodochial conidia falcate, curved dorsiven-
trally with almost parallel sides tapering slightly towards both
ends, with a blunt to papillate, curved apical cell and a blunt to
foot-like basal cell, (1–)3 –4 (–5)-septate, hyaline, smooth- and
thin-walled; 1-septate conidia: (15–)16– 24(– 25) × 3 4 µm (av.
20 × 3 µm); 3-septate conidia: (24 –)27–33(–36) × (2–)3– 5 µm
(av. 30 × 4 µm); 4-septate conidia: (32 –)33– 37(– 40) × 3 – 5 µm
(av. 35 × 4 µm); 5-septate conidia: (35–)3844 × 34 µm (av.
41 × 4 µm). Chlamydospores globose to subglobose, formed
terminally, 5–8 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.0–4.4 mm /d at 24 °C. Colony surface
pale vinaceous grey to vinaceous grey, floccose with abundant
aerial mycelium; colony margins irregular, lobate, serrate or
filiform. Odour absent. Reverse pale vinaceous grey, lacking
diffusible pigment. On SNA, hyphae hyaline, smooth-walled,
with abundant chlamydospores, aerial mycelium sparse with
abundant sporulation on the medium surface. On CLA, aerial
mycelium sparse with abundant pale luteous to orange sporo-
dochia forming on the carnation leaves.
Additional materials examined. south africa, North West Province, from
Glycine max, 2010, S.C. Lamprecht, CBS 144744 = CPC 25803; from Zea
mays, 2008, C.M. Bezuidenhout, CBS 144742 = CPC 25801.
Notes — Fusarium fabacearum formed a highly-supported
subclade in Clade III, closely related to F. carminascens and
F. glycines. See notes under F. carminascens for distinguishing
morphological features.
Fusarium glycines L. Lombard, Crous & Lampr., sp. nov.
MycoBank MB826840; Fig. 9
Etymology. Name refers to the plant genus Glycine from which this fungus
was isolated.
Typus. south africa, North West Province, from Glycine max, 2010, S.C.
Lamprecht (holotype CBS H-23614 designated here, culture ex-type CBS
144746 = CPC 25808).
Fig. 9 Fusarium glycines (ex-type culture CBS 144746). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light; b.
reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e– f. sporodochia on carnation leaves; g– i. conidiophores and phialides on aerial
mycelium; j –k. sporodochia and sporodochial conidiophores; l. chlamydospore; m. aerial conidia (microconidia); n. sporodochial conidia (macroconidia). — Scale
bars: g– i, l–n = 10 µm; j– k = 50 µm.
26 Persoonia – Volume 43, 2019
Conidiophores carried on the aerial mycelium 5– 45 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
15–25 × 2–4 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: 7–11(–13) × 3– 4 µm (av. 9 × 3
µm); 1-septate conidia: (13 –)14–16(–18) × 3– 4 µm (av. 15 ×
3 µm). Sporodochia bright orange, formed abundantly on car-
nation leaves. Conidiophores in sporodochia verticillately branch-
ed and densely packed, consisting of a short, smooth- and thin-
walled stipe, 4–9 × 2– 4 µm, bearing apical whorls of 2–3
monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 12–14 ×
2–5 µm, smooth- and thin-walled, sometimes showing a
reduced and flared collarette. Sporodochial conidia falcate,
curved dorsiventrally with almost parallel sides tapering slightly
towards both ends, with a blunt to papillate, curved apical cell
and a blunt to foot-like basal cell, (1–)3 5-septate, hyaline,
smooth- and thin-walled; 1-septate conidia: 20–25 × 3– 4 µm
(av. 23 × 3 µm); 3-septate conidia: 37–43(–48) × 45 µm (av.
38 × 4 µm); 4-septate conidia: 44 46(–51) × 4– 5 µm (av. 42 ×
4 µm); 5-septate conidia: 43– 49(–52) × 4–5 µm (av. 46 × 4 µm).
Chlamydospores globose to subglobose, formed terminally,
4–8 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.0–4.4 mm /d at 24 °C. Colony surface
vinaceous, floccose with abundant aerial mycelium; colony
margins irregular, lobate, serrate or filiform. Odour absent.
Reverse vinaceous, lacking diffusible pigment. On SNA, hyphae
hyaline, smooth-walled, with abundant chlamydospores, aerial
mycelium sparse with abundant sporulation on the medium
surface. On CLA, aerial mycelium sparse with abundant bright
orange sporodochia forming on the carnation leaves.
Additional materials examined. arGeNtiNa, substrate unknown, date un-
known, C.J.M. Carrera, CBS 214.49 = NRRL 36356 = LCF F-245. – italy,
from Ocimum basilicum, 1989, G. Tamiette & A. Matta, CBS 200.89. south
africa, North West Province, from Glycine max, 2010, S.C. Lamprecht, CBS
144745 = CPC 25804. – uNkNowN locality, from Linum usitatissium, 1933,
E.C. Stakman, CBS 176.33 = NRRL 36286.
Notes — Fusarium glycines formed a highly-supported sub-
clade in Clade III, closely related to F. carminascens and F. fa-
bacearum. See notes under F. carminascens for distinguishing
morphological features.
Fusarium gossypinum L. Lombard & Crous, sp. nov. — Myco-
Bank MB826841; Fig. 10
Etymology. Name refers to the plant genus Gossypium from which this
fungus was isolated.
Typus. ivory coast, Bouaké, wilted Gossypium hirsutum, Sept. 1995,
K. Abo (holotype CBS H-23615 designated here, culture ex-type CBS 116613).
Conidiophores carried on the aerial mycelium 35–75 µm tall, un-
branched or sparingly branched, bearing terminal or intercalarily
monophialides, often reduced to single phialides; aerial phial-
ides subulate to subcylindrical, smooth- and thin-walled, 3–30
× 2–4 µm, periclinal thickening inconspicuous or absent. Micro-
conidia forming small false heads on the tips of the phia lides,
hyaline, ellipsoidal to falcate, smooth- and thin-walled, 0–1-sep-
tate; 0-septate conidia: (5–)6 –8(–11) × 2–4 µm (av. 7 × 3 µm);
1-septate conidia: (11–)12–14(–15) × 24 µm (av. 15 × 3 µm).
Macroconidia also formed by phialides on aerial mycelium,
falcate, curved dorsiventrally with almost parallel sides taper-
ing slightly towards both ends, with a blunt to papillate, curved
Fig. 10 Fusarium gossypinum (ex-type culture CBS 116613). a –b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e. false head carried on a phialide on aerial mycelium; f– h. conidiophores
and phialides on aerial mycelium; i. aerial conidia (microconidia); j. sporodochial conidia (macroconidia). — Scale bars: e = 20 µm; f– j = 10 µm.
27
L. Lombard et al.: Epitypification of Fusarium oxysporum
apical cell and a blunt to foot-like basal cell, (1–)3-septate, hya-
line, smooth- and thin-walled; 1-septate conidia: 16–18 × 3 µm
(av. 17 × 3 µm); 2-septate conidia: 21–23 × 3 4 µm (av. 22 ×
3 µm); 3-septate conidia: (24–)27–35(– 38) × 3 4 µm (av. 31
× 4 µm). Sporodochia absent. Chlamydospores not observed.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 1.6–2.8 mm /d at 24 °C. Colony surface
white to pale rosy vinaceous, floccose with abundant aerial
mycelium; colony margins irregular, lobate, serrate or filiform.
Odour absent. Reverse pale rosy vinaceous, lacking diffusible
pigment. On SNA, hyphae hyaline, smooth-walled, lacking chla-
mydospores, aerial mycelium sparse with abundant sporulation
on the medium surface. On CLA, aerial mycelium sparse lacking
sporodochia on the carnation leaves.
Additional materials examined. ivory coast, Bouaké, wilted Gossypium
hirsutum, Sept. 1995, K. Abo, CBS 116611 and CBS 116612.
Notes — Fusarium gossypinum formed a unique highly-
supported subclade in Clade III. This species failed to produce
any sporodochia on the carnation leaf pieces, but still produced
abundant 3-septate macroconidia on the aerial mycelium. Other
species included in Clade III, all readily produced sporodochia
on carnation leaves.
Fusarium hoodiae L. Lombard, Crous & Lampr., sp. nov.
MycoBank MB826842; Fig. 11
Etymology. Name refers to the plant genus Hoodia from which this fungus
was isolated.
Typus. south africa, Northern Cape Province, Prieska, root of Hoodia
gordonii, 2002, O.A. Philippou (holotype CBS H-23616 designated here,
culture ex-type CBS 132474).
Conidiophores carried on the aerial mycelium 40 60 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
15–24 × 2–3 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (5–) 6–10(–16) × 2– 4 µm (av.
8 × 3 µm); 1-septate conidia: (11–)12–16(–17) × 3–4 µm (av.
Fig. 11 Fusarium hoodiae (ex-type culture CBS 132474). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e–f. sporodochia on carnation leaves; g– h. conidiophores and phialides on aerial
mycelium; i –k. sporodochia and sporodochial conidiophores; l. chlamydospore; m. aerial conidia (microconidia); n. sporodochial conidia (macroconidia). — Scale
bars: g– h, l–n = 10 µm; i = 50 µm; j– k = 20 µm.
28 Persoonia – Volume 43, 2019
14 × 3 µm). Sporodochia pale vinaceous to light orange, formed
abundantly on carnation leaves. Conidiophores in sporodochia
verticillately branched and densely packed, consisting of a
short, smooth- and thin-walled stipe, 7–11 × 3–5 µm, bearing
apical whorls of 2–3 monophialides or rarely as single lateral
monophialides; sporodochial phialides subulate to subcylindri-
cal, 7–13 × 25 µm, smooth- and thin-walled, sometimes show-
ing a reduced and flared collarette. Sporodochial conidia falcate,
curved dorsiventrally with almost parallel sides tapering slightly
towards both ends, with a blunt to papillate, curved apical cell
and a blunt to foot-like basal cell, (1–)3 (– 4)-septate, hyaline,
smooth- and thin-walled; 1-septate conidia: 20–33 × 3– 5 µm
(av. 25 × 4 µm); 3-septate conidia: (20 –)27–39 (–45) × 3 5 µm
(av. 33 × 4 µm); 4-septate conidia: (35–)36–46(–51) × 4–5
µm (av. 41 × 5 µm). Chlamydospores globose to subglobose,
formed terminally, 4–11 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.5 mm/d at 24 °C. Colony surface
pale vinaceous grey to livid vinaceous, floccose with abundant
aerial mycelium; colony margins irregular, lobate, serrate or
filiform. Odour absent. Reverse livid purple to pale vinaceous
grey, lacking diffusible pigment. On SNA, hyphae hyaline,
smooth-walled, with abundant chlamydospores, aerial myce-
lium sparse with abundant sporulation on the medium surface.
On CLA, aerial mycelium sparse with abundant pale vinaceous
to light orange sporodochia forming on the carnation leaves.
Additional materials examined. south africa, Northern Cape Province,
Prieska, root of Hoodia gordonii, 2002, O.A. Philippou, CBS 132476, CBS
132477.
Notes — Fusarium hoodiae formed a weakly supported
clade constituting Clade IV in this phylogenetic study. All three
isolates studied here, produced pale vinaceous to pale orange
sporodochia on the carnation leaf pieces, unique for all the
isolates studied.
Fusarium languescens L. Lombard & Crous, sp. nov. — Myco-
Bank MB826843; Fig. 12
Etymology. Name refers to the wilting symptoms associated with infec-
tions of this fungus.
Typus. morocco, Solanum lycopersicum, date and collector unknown (holo-
type CBS H-23617 designated here, culture ex-type CBS 645.78 = NRRL
36531).
Conidiophores carried on the aerial mycelium 25 30 µm
tall, unbranched or sparingly branched, bearing terminal or
intercalarily monophialides, often reduced to single phialides;
aerial phialides subulate to subcylindrical, smooth- and thin-
walled, 7–22 × 2–4 µm, periclinal thickening inconspicuous or
absent; aerial conidia forming small false heads on the tips of
the phialides, hyaline, ellipsoidal to falcate, smooth- and thin-
walled, 0–1-septate; 0-septate conidia: (4–) 59(–12) × 2–3
Fig. 12 Fusarium languescens (ex-type culture CBS 645.78). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c. conidiophores on surface of carnation leaf; d. sporodochia on carnation leaves; e– h. conidiophores and phialides on
aerial mycelium; i– k. sporodochia and sporodochial conidiophores; l. chlamydospore; m. aerial conidia (microconidia); n. sporodochial conidia (macroco-
nidia). — Scale bars: e– h, l–n = 10 µm; i– k = 20 µm.
29
L. Lombard et al.: Epitypification of Fusarium oxysporum
µm (av. 7 × 3 µm); 1-septate conidia: (9–)11–15 × 2–4 µm (av.
13 × 3 µm). Sporodochia light orange, formed abundantly on
carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 5–10 × 34 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 10–14 × 2–4
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a
blunt to foot-like basal cell, 1–3(– 5)-septate, hyaline, smooth-
and thin-walled; 1-septate conidia: (15–)18– 23(– 30) × 3 –4 µm
(av. 20 × 3 µm); 2-septate conidia: (14–)16–22(–24) × 4 µm
(av. 19 × 3 µm); 3-septate conidia: (22 –)2638(–47) × 3–5
µm (av. 32 × 4 µm); 5-septate conidia: 32–40 × 4 5 µm (av.
36 × 5 µm). Chlamydospores globose to subglobose, formed
terminally, 6–9 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.5 mm/d at 24 °C. Colony surface
flesh to rosy vinaceous, floccose with abundant aerial myce-
lium; colony margins irregular, lobate, serrate or filiform. Odour
absent. Reverse pale luteous, lacking diffusible pigment. On
SNA, hyphae hyaline, smooth-walled, with abundant chlamydo-
spores, aerial mycelium sparse with abundant sporulation on
the medium surface. On CLA, aerial mycelium sparse with
abundant light orange sporodochia forming on the carnation
leaves.
Additional materials examined. israel, Bet Dagan, Solanum lycopersicum,
1986, R. Cohn, CBS 413.90 = ATCC 66046 = NRRL 36465. – morocco, Sola-
num lycopersicum, date and collector unknown, CBS 646.78 = NRRL 36532.
NetherlaNds, Solanum lycopersicum, 1991, D.H. Elgersma, CBS 300.91 =
NRRL 36416, CBS 302.91 = NRRL 36419. – south africa, Zea mays, date and
collector unknown, CBS 119796 = MRC 8437. – uNkNowN locality, Solanum
lycopersicum, date and collector unknown, CBS 872.95 = NRRL 36570.
Notes — Fusarium languescens forms the highly-supported
Clade VI, which mostly includes strains associated with tomato
wilt. This species displays morphological overlap with several
species treated here. Therefore, phylogenetic inference is
needed to accurately identify this species.
Fusarium libertatis L. Lombard, Crous, sp. nov. — MycoBank
MB826844; Fig. 13
Etymology. Name refers to ‘freedom’. Fusarium libertatis was isolated from
the rock surfaces in the stone quarry on Robben Island where the prisoners
were forced to work. It is named in remembrance of all those who through the
centuries were incarcerated on the Island for their different political beliefs.
Typus. south africa, Western Cape Province, Robben Island, Van Rie-
beeck’s Quarry, from rock surfaces, May 2015, P.W. Crous (holotype CBS
H-23618 designated here, culture ex-type CBS 144749 = CPC 28465).
Conidiophores carried on the aerial mycelium 2– 30 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily phialides, often reduced to single phialides; aerial
phialides mono- and polyphialidic, subulate to subcylindrical,
smooth- and thin-walled, 8–13 × 2–4 µm, sometimes proliferat-
ing percurrently, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (6–)7– 9(–11) × 2 4 µm (av.
8 × 3 µm); 1-septate conidia: (11–)12–14(–15) × 2–4 µm (av.
13 × 3 µm). Sporodochia bright orange, formed abundantly on
carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 4–8 × 3 4 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 6–12 × 2–4
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a
blunt to foot-like basal cell, 1–3-septate, hyaline, smooth- and
thin-walled; 1-septate conidia: (15–)17–21(–23) × 2–4 µm (av.
19 × 3 µm); 2-septate conidia: (18 –)20– 24(–25) × 23(–4) µm
(av. 22 × 4 µm); 3-septate conidia: (24–)30– 38(–40) × (2–)3–5
µm (av. 34 × 4 µm). Chlamydospores globose to subglobose,
formed terminally and intercalarily, carried singly, 59 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 2.3–4.4 mm /d at 24 °C. Colony surface
vinaceous, floccose with abundant aerial mycelium; colony
margins irregular, lobate, serrate or filiform. Odour absent.
Reverse vinaceous, lacking diffusible pigment. On SNA, hyphae
hyaline, smooth-walled, with abundant chlamydospores, aerial
mycelium sparse with abundant sporulation on the medium
surface. On CLA, aerial mycelium sparse with abundant bright
orange sporodochia forming on the carnation leaves.
Additional materials examined. south africa, Western Cape Province,
from Aspalathus sp., 2008, C.M. Bezuidenhout, CBS 144747 = CPC 25788,
CBS 144748 = CPC 25782.
Notes — Fusarium libertatis formed a unique well-supported
clade Clade (II). This species readily produced polyphialidic
conidio genous cells on its aerial mycelium and can be distin-
guished from the other species (F. carminascens, F. curvatum
and F. elaeidis) found to produce polyphialides by only produc-
ing up to 3-septate macroconidia, whereas the other polyphi-
alidic species produce up to 5-septate macroconidia.
Fusarium nirenbergiae L. Lombard & Crous, sp. nov. — Myco-
Bank MB826845; Fig. 14
Etymology. Named in honour of Prof. H.I. Nirenberg for her contribution
to our understanding of Fusarium taxonomy.
Typus.
NetherlaNds, Aalsmeer, from Dianthus caryophyllus, 1988, H. Rat-
tink (holotype CBS H-23619 designated here, culture ex-type CBS 840.88).
Conidiophores carried on the aerial mycelium 18 50 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
8–24 × 2–4 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (5–) 6–10(–11) × 2– 4 µm (av.
8 × 3 µm); 1-septate conidia: (9–)10–14(–15) × 2–4 µm (av.
12 × 3 µm). Sporodochia bright orange, formed abundantly on
carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 6–14 × 35 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 8–18 × 2–4
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a
blunt to foot-like basal cell, 1–5-septate, hyaline, smooth- and
thin-walled; 1-septate conidia: 15–29 (–34) × 3–4 µm (av. 22
× 4 µm); 2-septate conidia: (18 –)19 31(–39) × 2– 4(–5) µm
(av. 25 × 3 µm); 3-septate conidia: (30 –)32– 40(–43) × 3 – 4 µm
(av. 36 × 4 µm); 4-septate conidia: (34 –)36– 44(–48) × 3 – 5 µm
(av. 40 × 4 µm); 5-septate conidia: (36 –)43– 59(–66) × 3 – 5 µm
(av. 51 × 4 µm). Chlamydospores globose to subglobose,
formed terminally, 4–6 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 2.9–4.2 mm /d at 24 °C. Colony surface
30 Persoonia – Volume 43, 2019
pale vinaceous to vinaceous, floccose with abundant aerial
mycelium; colony margins irregular, lobate, serrate or filiform.
Odour absent. Reverse pale vinaceous grey to greyish lilac,
lacking diffusible pigment. On SNA, hyphae hyaline, smooth-
walled, with abundant chlamydospores, aerial mycelium sparse
with moderate sporulation on the medium surface. On CLA,
aerial mycelium sparse with abundant bright orange sporo-
dochia forming on the carnation leaves.
Additional materials examined. BraZil, from Passiflora edulis, 1968,
W. Gerlach, CBS 744.79 = BBA 62355 = NRRL 22549. – italy, Napoli, Castel-
lammare di Stabia, from Bouvardia longiflora, July 1986, B. Aloj, CBS 196.87
= NRRL 26219. – NetherlaNds, Berkel, from Solanum lycopersicum, 16 May
1968, G. Weststeijn, CBS 758.68 = NRRL 36546. – south africa, Western
Cape Province, Riebeeck-Wes, from Agathosma betulina, 2001, K. Lubbe,
CBS 115424 = CPC 5312; Stellenbosch, Elsenberg farm, from Agathosma
betulina, 2001, K. Lubbe, CBS 115416 = CPC 5307, CBS 115417 = CPC
5306, CBS 115419 = CPC 5308. – USA, California, from amputated human
toe, unknown date and collector, CBS 130300 = NRRL 26368; Florida, from
Solanum tuberosum, 1923, H.W. Wollenweber, CBS 181.32 = NRRL 36303;
from Chrysanthemum sp., date unknown, G.M. Armstrong & J.K. Armstrong,
CBS 127.81 = BBA 63924 = NRRL 36229; Florida, from Chrysanthemum sp.,
date unknown, A.W. Engelhard, CBS 129.81 = BBA 63926 = NRRL 22539;
Maryland, Beltsville, from tulip roots, 1991, R.L. Lumsden, CBS 123062 =
GJS 91-17; Florida, Immokalee, from Solanum lycopersicum, date unknown,
J. Swezey, CBS 130303; Texas, San Antonio, from human leg ulcer, date
and collector unknown, CBS 130301 = NRRL 26374. uNkNowN locality,
from Secale cereale, date unknown, H.W. Wollenweber, CBS 129.24; from
Musa sp., date unknown, E.W. Mason, CBS 149.25 = NRRL 36261.
Fig. 13 Fusarium libertatis (ex-type culture CBS 144749). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c –e. conidiophores on surface of carnation leaf; g–k. conidiophores and phialides on aerial mycelium; g –h. monophialides;
i– k. polyphialides; l –n. sporodochia and sporodochial conidiophores; n. phialides of sporodochial conidiophores; o p. chlamydospores; q. aerial conidia
(microconidia); r. sporodochial conidia (macroconidia). — Scale bars: c– r = 10 µm.
31
L. Lombard et al.: Epitypification of Fusarium oxysporum
Notes — Fusarium nirenbergiae formed a well-supported
subclade in Clade VIII, closely related to F. curvatum. See notes
under F. curvatum for distinguishing morphological features.
Fusarium oxysporum Schltdl., Fl. Berol. 2: 139. 1824 — Fig.
15
Synonyms. Fusarium bulbigenum Cooke & Massee, Grevillea 16: 49.
1887.
Fusarium vasinfectum G.F.Atk., Bull. Alabama Agric. Exper. Station 41:
19. 1892.
Fusarium dianthi Prill. & Delacr., Compt. Rend. Acad. Sci. 129: 744. 1899.
Fusarium lini Bolley, Proc. Ann. Meeting Soc. Prom. Agr. Sci. 21: 1– 4.
1902.
Fusarium orthoceras Appel & Wollenw., Arb. Kaiserl. Biol. Anst. Ld.- u.
Forstw. 8: 152. 1910.
Fusarium citrinum Wollenw., Maine Agric. Exp. Sta. Bull. 219: 256. 1913.
Fusarium angustum Sherb., Cornell Univ. Agric. Exp. Sta. Mem. 6: 203.
1915.
Fusarium lutulatum Sherb., Cornell Univ. Agric. Exp. Sta. Mem. 6: 209.
1915.
Fusarium bostrycoides Wollenw. & Reinking, Phytopathology 15: 166.
1925.
Diplosporium vaginae Nann., Atti Reale Accad. Fisiocrit. Siena sér. 4, 17:
491. 1926.
For additional synonyms see Index Fungorum and MycoBank.
Typification. GermaNy, Berlin, from rotten tuber of Solanum tuberosum,
1824, D.L.F. von Schlechtendal, HAL 1612 F, holotype in HAL; (epitype
designated here: GermaNy, Berlin, from rotten tuber of Solanum tuberosum,
17 Oct. 2017, L. Lombard, epitype CBS H-23620, MBT382397, culture ex-
epitype CBS 144134).
Conidiophores carried on the aerial mycelium 15–75 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
11–40 × 2 4 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (4–) 6–10(–11) × 2– 4 µm (av.
8 × 3 µm); 1-septate conidia: 13–15(–16) × 2 4 µm (av. 14
× 3 µm). Sporodochia bright orange, formed abundantly on
carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 4–10 × 45 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 8–13 × 3–5
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a
blunt to foot-like basal cell, (1–)3(–5)-septate, hyaline, smooth-
and thin-walled; 1-septate conidia: (21–)22 –26 × 4–5 µm (av.
24 × 4 µm); 2-septate conidia: 20– 26(–27) × 4– 5 µm (av. 23
× 4 µm); 3-septate conidia: (22–)25 29(– 31) × 4–5 µm (av.
27 × 4 µm); 4-septate conidia: (30–)31– 35 × 4 5 µm (av. 33
× 5 µm); 5-septate conidia: 3538 × 5–6 µm (av. 37 × 5 µm).
Chlamydospores globose to subglobose, formed intercalarily
or terminally, 5–10 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.0–4.0 mm /d at 24 °C. Colony surface
pale vinaceous, floccose with abundant aerial mycelium; colony
margins irregular, lobate, serrate or filiform. Odour absent. Re-
verse vinaceous to rosy vinaceous, lacking diffusible pigment.
On SNA, hyphae hyaline, smooth-walled, producing abundant
chlamydospores, aerial mycelium sparse with abundant sporu-
lation on the medium surface. On CLA, aerial mycelium sparse
with abundant bright orange sporodochia forming on the carna-
tion leaves.
Fig. 14 Fusarium nirenbergiae (ex-type culture CBS 840.88). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c. conidiophores on surface of carnation leaf; d. sporodochia on carnation leaves; e. conidiophores and phialides on aerial
mycelium; f– g. sporodochia and sporodochial conidiophores; h. chlamydospore; i. aerial conidia (microconidia); j. sporodochial conidia (macroconidia). — Scale
bars: e, h– j = 10 µm; f–g = 50 µm.
32 Persoonia – Volume 43, 2019
Additional materials examined. GermaNy, from rotten tuber of Solanum
tuberosum, 17 Oct. 2017, L. Lombard, CBS 144135. – south africa, Western
Cape Province, from Protea sp., date unknown, C.M. Bezuidenhout, CPC
25822. – south east asia, from Camellia sinensis, 1949, F. Bugnicourt, CBS
221.49 = IHEM 4508 = NRRL 22546.
Notes — Fusarium oxysporum formed a well-supported
subclade in Clade V with F. triseptatum as closest relative.
Both species in Clade V displayed some morphological over-
lap. However, the 1-septate ((21–)2226 × 4–5 µm (av. 24 ×
4 µm) and 2-septate (2026(–27) × 4 5 µm (av. 23 × 4 µm)
macroconidia of F. oxysporum are larger than those of F. tri-
septatum ((18–)19– 23(–24) × 3 4 µm (av. 20 × 3 µm) and
17–25(– 26) × 3 µm (av. 21 × 3 µm), respectively), whereas
the 3-septate ((25–)27– 39(–47) × 4–5 µm (av. 33 × 3 µm)),
4-septate ((31–)34 40(– 41) × 4– 5 µm (av. 37 × 4 µm)) and
5-septate ((33–48 (–49) × 4–5 µm (av. 40 × 4 µm)) macroco-
nidia of F. triseptatum are larger than those of F. oxysporum
((22–)25 29(– 31) × 4– 5 µm (av. 27 × 4 µm), (30–)31– 35 ×
4–5 µm (av. 33 × 5 µm) and 35– 38 × 5– 6 µm (av. 37 × 5 µm),
respectively). Additionally, all isolates of F. oxysporum produced
abundant bright orange sporodochia on carnation leaf pieces,
not observed for any of the F. triseptatum isolates studied.
Fusarium pharetrum L. Lombard & Crous, sp. nov. — Myco-
Bank MB826846; Fig. 16
Etymology. Name refers to the practice of the Southern African indige-
nous San people of hollowing out the tubular branches of the host plant,
Aloidendron dichotomum, to form quivers (Latin pharetra) for their arrows.
Fig. 15 Fusarium oxysporum (ex-epitype culture CBS 144134). a–b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e f. sporodochia on carnation leaves; g j. conidiophores and phialides
on aerial mycelium; k– n. sporodochia and sporodochial conidiophores; o– p. chlamydospores; q. aerial conidia (microconidia); r. sporodochial conidia (macro-
conidia). — Scale bars: g– h, m–r = 10 µm; i– l = 50 µm.
33
L. Lombard et al.: Epitypification of Fusarium oxysporum
Typus. south africa, from Aliodendron dichotomum, 2000, F. van der
Walt & G.J. Marais (holotype CBS H-23621 designated here, culture ex-type
CBS 144751 = CPC 30824).
Conidiophores carried on the aerial mycelium 20–75 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
4–28 × 2–5 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: 5– 9(–13) × 2– 3 µm (av. 7 × 3
µm); 1-septate conidia: (10–)12–16 (–18) × 2– 4 µm (av. 14 × 3
µm). Sporodochia rosy vinaceous to orange, formed abundantly
on carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 5–10 × 35 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 7–13 × 3–4
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a
blunt to foot-like basal cell, 3(– 4)-septate, hyaline, smooth- and
thin-walled; 3-septate conidia: (22–)27– 35(–39) × 3–5 µm (av.
31 × 4 µm); 4-septate conidia: (34 –)36–40(–41) × 35 µm (av.
36 × 5 µm). Chlamydospores not observed.
Fig. 16 Fusarium pharetum (ex-type culture CBS 144751). a– b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e f. sporodochia on carnation leaves; g h. false heads carried on a phialide
on aerial mycelium; i– l. conidiophores and phialides on aerial mycelium; m– p. sporodochia and sporodochial conidiophores; q. aerial conidia (microconidia);
r. sporodochial conidia (macroconidia). — Scale bars: g– l, q– r = 10 µm; m–p = 50 µm.
34 Persoonia – Volume 43, 2019
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.5 mm/d at 24 °C. Colony surface
rosy vinaceous, floccose with abundant aerial mycelium; colony
margins irregular, lobate, serrate or filiform. Odour absent.
Reverse rosy vinaceous, lacking diffusible pigment. On SNA,
hyphae hyaline, smooth-walled, lacking chlamydospores, aerial
mycelium sparse with abundant sporulation on the medium
surface. On CLA, aerial mycelium sparse with abundant rosy
vinaceous to orange sporodochia forming on the carnation
leaves.
Additional material examined. south africa, from Aliodendron dichoto-
mum, 2000, F. van der Walt & G.J. Marais, CBS 144750 = CPC 30822.
Notes — Fusarium pharetrum formed a well-supported sub-
clade in Clade VII, closely related to F. contaminatum and F. vete-
rinarium. See notes under F. contaminatum for distinguishing
morphological features.
Fusarium triseptatum L. Lombard & Crous, sp. nov. — Myco-
Bank MB826847; Fig. 17
Etymology. Name refers to the abundant 3-septate macroconidia pro-
duced by this fungus.
Typus. USA, locality unknown, from Ipomoea batatas, 1950, T.T. McClure
(holotype CBS H-23622 designated here, culture ex-type CBS 258.50 =
NRRL 36389).
Conidiophores carried on the aerial mycelium 5– 40 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
6–22 × 2–4 µm, periclinal thickening inconspicuous or absent.
Microconidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (5–) 6–10(–13) × 1– 3 µm (av.
8 × 3 µm); 1-septate conidia: (12–)14–16(–18) × 2 4 µm (av.
15 × 3 µm). Macroconidia also formed by phialides on aerial
mycelium, falcate, curved dorsiventrally with almost paral-
lel sides tapering slightly towards both ends, with a blunt to
papillate, curved apical cell and a blunt to foot-like basal cell,
(1–)3(–5)-septate, hyaline, smooth- and thin-walled; 1-septate
conidia: (18–)19– 23(– 24) × 3 4 µm (av. 20 × 3 µm); 2-septate
conidia: 17–25(– 26) × 3 µm (av. 21 × 3 µm); 3-septate conidia:
(25–)27–39(– 47) × 4–5 µm (av. 33 × 3 µm); 4-septate conidia:
(31–)34 40(– 41) × 4–5 µm (av. 37 × 4 µm); 5-septate conidia:
33–48 (–49) × 4 5 µm (av. 40 × 4 µm). Sporodochia absent.
Chlamydospores globose to subglobose, formed terminally,
5–12 µm diam.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 2.2–3.4 mm /d at 24 °C. Colony surface
pale vinaceous grey to vinaceous grey, floccose with abundant
aerial mycelium; colony margins irregular, lobate, serrate or
filiform. Odour absent. Reverse pale vinaceous grey, lacking
diffusible pigment. On SNA, hyphae hyaline, smooth-walled,
Fig. 17 Fusarium triseptatum (ex-type culture CBS 258.50). a–b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white light;
b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e– f. false heads carried on a phialide on aerial mycelium; g–j. conidiophores
and phialides on aerial mycelium; k– l. chlamydospores; m. microconidia; n. macroconidia. — Scale bars: e, g–n = 10 µm; f = 20 µm.
35
L. Lombard et al.: Epitypification of Fusarium oxysporum
with abundant chlamydospores, aerial mycelium sparse with
abundant sporulation on the medium surface. On CLA, aerial
mycelium sparse lacking sporodochia on the carnation leaves.
Additional materials examined. ivory coast, Béoumi, wilted Gossypium
hirsutum, Oct. 1996, K. Abo, CBS 116619. – PaPua New GuiNea, Suki village,
from sago starch, 2005, A. Greenhill, CBS 119665. – USA, Tennessee, from
human eye, collector and date unknown, CBS 130302 = NRRL 26360 = FRC
755.
Notes — Fusarium triseptatum formed a highly-supported
subclade in Clade V, closely related to F. oxysporum. See notes
under F. oxysporum for distinguishing morphological features.
Fusarium veterinarium L. Lombard & Crous, sp. nov. — Myco-
Bank MB826849; Fig. 18
Etymology. Name refers to the fact that this fungus was isolated mostly
from veterinary samples.
Typus. NetherlaNds, from shark peritoneum, date unknown, C. Hoek (holo-
type CBS H-23623 designated here, culture ex-type CBS 109898 = NRRL
36153).
Conidiophores carried on the aerial mycelium 12 90 µm tall,
unbranched or sparingly branched, bearing terminal or inter-
calarily monophialides, often reduced to single phialides; aerial
phialides subulate to subcylindrical, smooth- and thin-walled,
8–24 × 2–4 µm, periclinal thickening inconspicuous or absent;
aerial conidia forming small false heads on the tips of the phia-
lides, hyaline, ellipsoidal to falcate, smooth- and thin-walled,
0–1-septate; 0-septate conidia: (4–) 68(–11) × 2–4 µm (av.
7 × 3 µm); 1-septate conidia: (9–)10–14(–15) × 2–4 µm (av.
12 × 3 µm). Sporodochia bright orange, formed abundantly on
carnation leaves. Conidiophores in sporodochia verticillately
branched and densely packed, consisting of a short, smooth-
and thin-walled stipe, 8–13 × 34 µm, bearing apical whorls of
2–3 monophialides or rarely as single lateral monophialides;
sporodochial phialides subulate to subcylindrical, 10–15 × 2–4
µm, smooth- and thin-walled, sometimes showing a reduced
and flared collarette. Sporodochial conidia falcate, curved dor-
siventrally with almost parallel sides tapering slightly towards
both ends, with a blunt to papillate, curved apical cell and a blunt
to foot-like basal cell, 1–(2 –) 3-septate, hyaline, smooth- and
Fig. 18 Fusarium veterinarium (ex-type culture CBS 109898). a –b. Colony on PDA; a. surface of colony on PDA after 7 d at 24 °C under continuous white
light; b. reverse of colony on PDA; c– d. conidiophores on surface of carnation leaf; e f. sporodochia on carnation leaves; g i. conidiophores and phialides
on aerial mycelium; j– l. sporodochia and sporodochial conidiophores; m. aerial conidia (microconidia); n. sporodochial conidia (macroconidia). — Scale bars:
g– n = 10 µm.
36 Persoonia – Volume 43, 2019
thin-walled; 1-septate conidia: (12–)15–19(–20) × 3–4 µm (av.
17 × 3 µm); 2-septate conidia: (16–)17–21(– 24) × 3–4 µm (av.
19 × 3 µm); 3-septate conidia: (19 –)20–24(–27) × 34 µm (av.
22 × 3 µm). Chlamydospores not observed.
Culture characteristics — Colonies on PDA with an average
radial growth rate of 3.1–4.5 mm/d at 24 °C. Colony surface
pale vinaceous grey, floccose with moderate aerial mycelium
appearing wet; colony margins irregular, lobate, serrate or
filiform. Odour absent. Reverse pale vinaceous, lacking diffusi-
ble pigment. On SNA, hyphae hyaline, smooth-walled, lacking
chlamydospores, aerial mycelium sparse with abundant sporu-
lation on the medium surface. On CLA, aerial mycelium sparse
with abundant orange sporodochia forming on the carnation
leaves.
Additional materials examined. NetherlaNds, from swab sample near
filling apparatus, Apr. 2005, J. Houbraken, CBS 117787, CBS 117790; from
pasteurized milk-based product, Apr. 2005, J. Houbraken, CBS 117791,
CBS 117792. – USA, California, from endoscope of veterinary clinic, date
and collector unknown, NRRL 62545; from canine stomach, date and col-
lector unknown, NRRL 62547; Massachusetts, from mouse mucosa, date
and collector unknown, NRRL 54984; from little blue penguin foot, date and
collector unknown, NRRL 54996; Texas, from unknown animal faeces, date
and collector unknown, NRRL 62542.
Notes — Fusarium veterinarium formed a highly-supported
subclade in Clade VII, closely related to F. contaminatum and
F. pharetrum. See notes under F. contaminatum for distinguish-
ing morphological features.
DISCUSSION
Fusarium taxonomy and the underlying phylogenetic backbone
on which it is based, is undergoing continuous revision. In
modern day fungal taxonomy, phylogenetic inference plays a
vital role to resolve the identity of cryptic species due to the
paucity of morphological features. However, a key component
of a robust phylogeny is the availability of living ex-type mate-
rial to serve as basic reference point or ‘phylogenetic anchor’
on which comparative taxonomy can be based (Booth 1975).
Epi- and/or neotypification provides a vital means where upon
stability can be enforced into a chaotic classification system as
being applied to F. oxysporum today.
Snyder & Hansen’s (1940) treatment of the section Elegans to
represent only F. oxysporum, has resulted in a much too broad
definition of this species. Based on this, the current morphologi-
cal characters used to define F. oxysporum include aseptate
microconidia forming false heads on short monophialides,
commonly 3-septate macroconidia formed on monophialides
or branched conidiophores in sporodochia, and chlamydo-
spores that are either formed abundantly and quickly or slowly
with some strains not forming them at all (Leslie & Summerell
2006, Fourie et al. 2011). In this study, all isolates were found
to produce not only aseptate microconidia, but abundant 1-sep-
tate microconidia, all of which were carried on false heads.
Several species were also found to form polyphialides (e.g.,
F. carminascens, F. curvatum, F. elaeidis and F. libertatis),
a characteristic not associated with F. oxysporum morpho-
logy (Gerlach & Nirenberg 1982, Nelson et al. 1983, Leslie
& Summerell 2006). Additionally, the majority of the species
introduced here produced 4- to 5-septate macroconidia in the
same abundance as the 3-septate macroconidia. Gerlach &
Nirenberg (1982) also indicated the presence of 7-septate
macroconidia, but these were not observed in this study given
the media and growth conditions we employed. The ex-epitype
strain of F. oxysporum designated here, agrees well with the
morphological characteristics described by Wollenweber &
Reinking (1935), Booth (1971), Gerlach & Nirenberg (1982) and
Nelson et al. (1983). This strain produced abundant aseptate
and 1-septate microconidia on monophialides only, abundant
3-septate macroconidia with much fewer 1-, 2-, 4- and 5-septate
macroconidia on its sporodochia, and smooth-walled globose
chlamydospores carried intercalarily and/ or terminally. Although
this strain was isolated from a potato tuber displaying symptoms
of dry rot, the ability of this strain to induce these symptoms
requires further investigation. Comparisons of the 15 novel
Fusarium taxa introduced here, revealed subtle morphological
distinctions between the species.
Fusarium carminascens, F. curvatum, F. elaeidis and F. libertatis
readily formed polyphialides on the aerial mycelium, a feature
not known for F. oxysporum (Wollenweber & Reinking 1935,
Booth 1971, Gerlach & Nirenberg 1982, Nelson et al. 1983,
Leslie & Summerell 2006). These four species are further
distinguished from each other by the degree of septation and
curvature of their macroconidia. Both F. carminascens and
F. liber tatis readily formed chlamydospores in culture, whereas
no chlamydospores were observed for F. curvatum and F. elaei-
dis. Furthermore, all strains of F. carminascens produced an
almost carmine red exudate on the aerial mycelium on PDA, not
observed for any other strains studied here. The strong curva-
ture of the macroconidia of F. curvatum is also a unique feature.
The remaining 11 novel species introduced here can be dis-
tinguished based on the degree of septation and dimensions
of the macroconidia and the formation of chlamydospores in
culture. Of these, F. contaminatum, F. gossypinum, F. hoodiae,
F. languescens, F. pharetrum, F. triseptatum and F. veterinarium
displayed some morphological overlap with the ex-epitype strain
of F. oxysporum. However, F. contaminatum, F. gossypinum,
F. pharetrum and F. veterinarium did not form chlamydospores
in culture. These four species are easily distinguished based
on macroconidial dimensions with F. contaminatum and F. vete-
rinarium producing the smallest macroconidia. Fusarium hoo-
diae, F. languescens and F. triseptatum readily formed chla-
mydospores in culture and can be distinguished from each
other and F. oxysporum based on their sporodochia. All
strains of F. triseptatum failed to produce any sporodochia on
the carnation leaf pieces, whereas F. hoodiae formed distinct
pale vinaceous to pale orange sporodochia compared to the
only pale orange sporodochia of F. languescens. Fusarium
callistephi, F. fabacearum, F. glycines and F. nirenbergiae are
easily distinguished from each other and F. oxysporum by the
degree of macroconidial septation and dimensions. However,
these subtle morphological differences need to be supported
by phylogenetic inference to accurately discriminate between
these novel species introduced in the FOSC in this study.
Individual analyses of the partial sequences of the four gene re-
gions (cmdA, rpb2, tef1 and tub2) included in this study (results
not shown) revealed that the tef1 gene region provided the best
resolution to discriminate the novel species introduced here.
The rpb2 gene region also provided good resolution, but with
lower statistical support, whereas the cmdA and tub2 provided
little to no support. However, the addition of the latter two gene
regions to either or both the rpb2 and tef1 greatly increased the
statistical support of each Clade (I–VIII) and their underlining
subclades. Genealogical concordance phylogenetic species
recognition analyses also indicated that there was no evidence
of recombination detected between any of the Clades and
subclades resolved in this study. Analysis of the IGS gene re-
gion (results not shown) provided contradictory tree topologies
and support values, with several strains in Clades III, VII and
VIII forming single lineages. Although O’Donnell et al. (2015)
advocates the use of rpb1, rpb2 and tef1 for sequence-based
identification of Fusarium species, attempts to generate rpb1
sequence data in this study failed for the majority of strains
included in this study.
37
L. Lombard et al.: Epitypification of Fusarium oxysporum
Previous studies of FOSC revealed a high phylogenetic di-
versity within this complex, resolving three (O’Donnell et al.
1998, Brankovics et al. 2017), four (O’Donnell et al. 2004) and
five (Laurence et al. 2012) phylogenetic clades, respectively.
Comparisons of all these clades with those resolved in this
study, revealed that Clade I in this study correlates well with
Clade 1 resolved by O’Donnell et al. (1998, 2004), Laurence
et al. (2012) and Brankovics et al. (2017). Similarly, Clade VIII
in this study matched with Clade 3 of each of these studies.
Clade III correlated with Clade 2 resolved by O’Donnell et al.
(2004) and Brankovics et al. (2017), and Clade V correlated
with clades 4 and 5 of Laurence et al. (2012), and Clade 4 of
O’Donnell et al. (2004). Clades II, IV, VI and VII resolved in
this study did not match any of the clades resolved in these
previous studies.
Comparisons of the origin of the strains studied here revealed
some correlation within most of the Clades (and subclades). All
veterinarian strains included in this study clustered together with
some strains originating from equipment used in food process-
ing in a highly-supported subclade representing F. veterinarium.
Similarly, three strains collected from contaminated dairy pro-
ducts and fruit juice clustered together in the highly-supported
(sub)clade representing F. contaminatum. The majority of the
isolates collected from tomato (Solanum lycopersicum) also
cluster together in a clade representing F. languescens, with a
few clustering in the F. nirenbergiae (sub)clade. In contrast to
these few highlighted examples, all medically related strains
clustered in various well- to highly supported clades, represent-
ing F. cugenangense, F. nirenbergiae, F. triseptatum and the
untreated Fusarium clade. The highest host/substrate diversity
was found in the F. nirenbergiae (sub)clade which included
several special forms in addition to the medically related strains.
The application of the special form and pathotype classification
system can only be successfully applied if the species bounda-
ries are well established (Woudenberg et al. 2015), which is
clearly not the case within the FOSC. For the FOSC, special
forms are defined by the accessory chromosome obtained
via horizontal gene transfer, and the pathotype on the type of
virulence genes carried by this chromosome and should not be
confused with the species boundaries within the FOSC. There-
fore, epitypification of F. oxysporum in this study has resulted
in the recognition of 21 phylogenetic species of which 15 are
provided with names here. Although this study includes only
a small subset of strains belonging to the FOSC, the inclusion
of more isolates will provide a much better perspective on the
cryptic diversity within this important species complex, allowing
additional species to be recognised. Furthermore, it is hoped
that with the epitypification of F. oxysporum, the confusing and
sometimes complicated subspecific classification systems that
have been applied to the FOSC in the past will become obsolete
and be replaced by a more stable and convenient species-level
classification system. We believe that such a system will allow
for better communication between Fusarium researchers in the
medical, environmental and phytopathological fields.
Acknowledgements The authors thank the technical staff, A. van Iperen,
D. Vos-Kleyn and Y. Vlug for their valuable assistance with cultures.
REFERENCES
Abd-Elsalam K, Khalil M, Aly AH, et al. 2002. Genetic diversity among
Fusarium oxysporum f. sp. vasinfectum isolates revealed by UP-PCR and
AFLP markers. Phytopathologia Mediterrianae 41: 252– 258.
Abd-Elsalam KA, Asran-Amal A, Schnieder F, et al. 2006. Molecular detec-
tion of Fusarium oxysporum f. sp. vasinfectum in cotton roots by PCR and
real-time PCR assay. Journal of Plant Disease and Protection 113: 14–19.
Abd-Elsalam KA, Omar MR, Migheli Q, et al. 2004. Genetic characterization
of Fusarium oxysporum f. sp. vasinfectum isolates by random amplification
of polymorphic DNA (RAPD) and amplified fragment length polymorphism
(AFLP). Journal of Plant Diseases and Protection 111: 534– 544.
Abo K, Klein KK, Edel-Hermann V, et al. 2005. High genetic diversity among
strains of Fusarium oxysporum f. sp. vasinfectum from cotton in Ivory Coast.
Phytopathology 95: 1391–1396.
Adame-García J, Rodríguez-Guerra R, Iglesias-Andreu LG, et al. 2015. Mo-
lecular identification and pathogenic variation of Fusarium species isolated
from Vanilla planifolia in Papantla Mexico. Botanical Sciences 93: 669– 678.
Aguayo J, Mostert D, Fourrier-Jeandel C, et al. 2017. Development of a hy-
drolysis probe-based real-time assay for the detection of tropical strains of
Fusarium oxysporum f. sp. cubense race 4. PLoS ONE 12 (2): e0171767.
doi: https://doi.org/10.1371/journal.pone.0171767.
Ahn IP, Chung HS, LeeY-H. 1998. Vegetative compatibility groups and
pathogenicity among isolates of Fusarium oxysporum f. sp. cucumerinum.
Plant Disease 82: 244– 246.
Al-Hatmi AMS, Bonifaz A, Ranque S, et al. 2018. Current antifungal treatment
of fusariosis. International Journal of Antimicrobial Agents 51: 326–332.
Al-Husien N, Hamwieh A, Ahmed S, et al. 2017. Genetic diversity of Fusarium
oxysporum f. sp. lentis population affecting lentil in Syria. Journal of Phyto-
pathology 165: 306– 312.
Alexander LJ, Tucker CM. 1945. Physiological specialization in the tomato
wilt fungus Fusarium oxysporum f. sp. lycopersici. Journal of Agricultural
Research 70: 303– 313.
Aloi C, Baayen RP. 1993. Examination of the relationships between vegeta-
tive compatibility groups and races in Fusarium oxysporum f. sp. dianthi.
Plant Pathology 42: 839– 850.
Altinok HH. 2013. Fusarium species isolated from common weeds in egg-
plants fields and symptomless hosts of Fusarium oxysporum f. sp. melon-
genae in Turkey. Journal of Phytopathology 161: 335– 340.
Altinok HH, Can C. 2010. Characterization of Fusarium oxysporum f. sp.
melongenae isolates from eggplant in Turkey by pathogenicity, VCG and
RAPD analysis. Phytoparasitica 38: 149–157.
Altinok HH, Can C, Çolak H. 2013. Vegetative compatibility, pathogenicity and
virulence diversity of Fusarium oxysporum f. sp. melongenae recovered
from eggplant. Journal of Phytopathology 161: 651–660.
Alvarez GLA. 1945. Studies on coffee root disease in Puerto Rico. Journal
of Agriculture of the University of Puerto Rico 29: 1–29.
Alves-Santos FM, Cordeiro-Rodrigues L, Sayagués, et al. 2002a. Patho-
genicity and race characterization of Fusarium oxysporum f. sp. phaseoli
isolates from Spain and Greece. Plant Pathology 51: 605– 611.
Alves-Santos FM, Ramos B, García-Sánchez MA, et al. 2002b. A DNA-based
procedure for in planta detection of Fusarium oxysporum f. sp. phaseoli.
Phytopathology 92: 237–244.
Aoki T, O’Donnell K, Geiser DM. 2014. Systematics of key phytopathogenic
Fusarium species: current status and future challenges. Journal of General
Plant Pathology 80: 189–201.
Arie T, Kaneko I, Yoshida T, et al. 2000. Mating-type genes from asexual
phytopathogenic ascomycetes Fusarium oxysporum and Alternaria alter-
nata. Molecular Plant-Microbe Interactions 13: 1330–1339.
Armstrong GM. 1954. Alfalfa – a common host for the wilt fusaria from alfalfa,
cotton, and cassia. Plant Disease Reporter 38: 221–222.
Armstrong GM, Armstrong JK. 1950. Biological races of the Fusarium cau-
sing wilt of cowpea and soybeans. Phytopathology 40: 181–193.
Armstrong GM, Armstrong JK. 1952. Physiologic races of the fusaria causing
wilts of the Cruciferae. Phytopathology 42: 255– 257.
Armstrong GM, Armstrong JK. 1953. Physiologic races of the wilt fusaria from
cabbage, radish, and stock. Phytopathology 43: 465.
Armstrong GM, Armstrong JK. 1958a. A race of the cotton wilt Fusarium
causing wilt of Yelredo soybean and flue-cured tobacco. Plant Disease
Reporter 42: 147–151.
Armstrong GM, Armstrong JK. 1958b. The Fusarium wilt complex as related
to the sweet potato. Plant Disease Reporter 42: 1319–1329.
Armstrong GM, Armstrong JK. 1960. American, Egyptian, and Indian cotton-
wilt fusaria; their pathogenicity and relationship to other wilt fusaria. U.S.
Department of Agriculture Technical Bulletin 1219: 1–19.
Armstrong GM, Armstrong JK. 1964. Lupinis species – common hosts for
wilt fusaria from alfalfa bean, Cassia, cowpea, lupine, and U.S. cotton.
Phytopathology 54: 1232–1235.
Armstrong GM, Armstrong JK. 1965. A wilt of soybean caused by a new form
of Fusarium oxysporum. Phytopathology 55: 237–239.
Armstrong GM, Armstrong JK. 1966. Races of Fusarium oxysporum f. sp.
conglutinans; race 4, new race; and a new host for race 1, Lychnis chal-
cedonica. Phytopathology 56: 525– 530.
Armstrong GM, Armstrong JK. 1968. Formae speciales and races of Fu-
sarium oxysporum causing a tracheomycosis in the syndrome of disease.
Phytopathology 58: 1242–1246.
38 Persoonia – Volume 43, 2019
Armstrong GM, Armstrong JK. 1971. Races of the aster-wilt Fusarium. Phyto-
pathology 61: 820– 824.
Armstrong GM, Armstrong JK. 1974. Races of Fusarium oxysporum f. sp.
pisi, causal agents of wilt of pea. Phytopathology 64: 849– 857.
Armstrong GM, Armstrong JK. 1976. Common hosts for Fusarium oxyspo-
rum formae speciales spinaciae and betae. Phytopathology 66: 542– 545.
Armstrong GM, Armstrong JK. 1978a. A new race (race 6) of the cotton-wilt
Fusarium from Brazil. Plant Disease Reporter 62: 421–423.
Armstrong GM, Armstrong JK. 1978b. Formae speciales and races of Fusa-
rium oxysporum causing wilts of the Cucurbitaceae. Phytopathology 68:
19– 28.
Armstrong GM, Armstrong JK. 1980. Cowpea wilt Fusarium oxysporum f. sp.
tracheiphilum race 1 from Nigeria. Plant Disease 64: 954– 955.
Armstrong GM, Armstrong JK. 1981. Formae speciales and races in Fu-
sarium oxysporum causing wilting diseases. In: Nelson PE, Toussoun TA,
Cook RJ (eds), Fusarium: diseases, biology, and taxonomy: 391–399. The
Pennsylvania State University Press, Pennsylvania State University, USA.
Armstrong GM, Armstrong JK, Billington RV. 1975. Fusarium oxysporum
forma specialis voandzeiae, a new form species causing wilt of Bambarra
groundnut. Mycologia 67: 709–714.
Armstrong GM, Armstrong JK, Littrell RH. 1970. Wilt of chrysanthemum
caused by Fusarium oxysporum f. sp. chrysanthemi, forma specialis nov.
Phytopathology 60: 496– 498.
Armstrong GM, Armstrong JK, Netzer D. 1978. Pathogenic races of the
cucumber-wilt Fusarium. Plant Disease Reporter 62: 824– 828.
Arya HC, Jain GL. 1962. Fusarium wilt of Eucalyptus. Phytopathology 52:
638– 642.
Assigbetse KB, Fernandez D, Dubois MP, et al. 1994. Differentiation of
Fusarium oxysporum f. sp. vasinfectum races on cotton by random amplified
polymorphic DNA (RAPD) analysis. Phytopathology 84: 622– 626.
Atkinson GF. 1892. Some diseases of cotton. Bulletin of the Alabama Agri-
cultural Experiment Station 41: 1–65.
Baayen RP, Elgersma DM, Demmink JF, et al. 1988. Differences in patho-
genesis observed among susceptible interactions of carnation with four
races of Fusarium oxysporum f. sp. dianthi. Netherlands Journal of Plant
Pathology 94: 81–94.
Baayen RP, Förch MG, Waalwijk C, et al. 1998. Pathogenic, genetic and
molecular characterisation of Fusarium oxysporum f. sp. lilii. European
Journal of Plant Pathology 104: 887–894.
Baayen RP, Kleijn J. 1989. The Elegans fusaria causing wilt disease of
carnation. II. Distinction of vegetative compatibility groups. Netherlands
Journal of Plant Pathology 95: 185–194.
Baayen RP, O’Donnell K, Bonants PJM, et al. 2000. Gene genealogies and
AFLP analyses in the Fusarium oxysporum complex identify monophyletic
and nonmonophyletic formae speciales causing wilt and rot disease. Phyto-
pathology 90: 891–900.
Baayen RP, Van Dreven F, Krijger MC, et al. 1997. Genetic diversity in
Fusarium oxysporum f. sp. dianthi and Fusarium redolens f. sp. dianthi.
European Journal of Plant Pathology 103: 395– 408.
Baker KF. 1948. Fusarium wilt of garden stock (Mattiola incana). Phytopa-
thology 38: 399– 403.
Balmas V, Scherm B, Di Primo P, et al. 2005. Molecular characterisation
of vegetative compatibility groups in Fusarium oxysporum f. sp. radicis-
lycopersici and f. sp. lycopersici by random amplification of polymorphic
DNA and microsatellite-primed PCR. European Journal of Plant Pathology
111: 1–8.
Bao JR, Fravel DR, O’Neill, et al. 2002. Genetic analysis of pathogenic and
nonpathogenic Fusarium oxysporum from tomato plants. Canadian Journal
of Botany 80: 271–279.
Barton E, Borman A, Johnson E, et al. 2016. Pseudo-outbreak of Fusarium
oxysporum associated with bronchoscopy. Journal of Hospital Infection
94: 197–198.
Barve MP, Haware MP, Sainani MN, et al. 2001. Potential of microsatellites
to distinguish four races of Fusarium oxysporum f. sp. ciceri prevalent in
India. Theoretical and Applied Genetics 102: 138–147.
Basirnia T, Banihashemi ZAD. 2005. Vegetative compatibility grouping in
Fusarium oxysporum f. sp. sesami, the causal agent of sesame yellows
and wilt in Fars province. Iranian Journal of Plant Pathology 41: 243– 255.
Bayraktar H, Dolar FS, Maden S. 2008. Use of RAPD and ISSR markers in
detection of genetic variation and population structure among Fusarium
oxysporum f. sp. ciceris isolates on chickpea in Turkey. Journal of Phyto-
pathology 156: 146–154.
Bayraktar H, Türkkan M, Dolar FS. 2010. Characterization of Fusarium
oxysporum f. sp. cepae from onion in Turkey based on vegetative compa-
tibility and rDNA RFLP analysis. Journal of Phytopathology 158: 691–697.
Baysal Ö, Karaaslan Ç, Siragusa M, et al. 2013. Molecular markers reflect
differentiation of Fusarium oxysporum forma speciales on tomato and
forma on eggplant. Biochemical Systematics and Ecology 47: 139–147.
Baysal Ö, Siragusa M, Gümrükcü E, et al. 2010. Molecular characterization
of Fusarium oxysporum f. melongenae by ISSR and RAPD markers on
eggplant. Biochemical Genetics 48: 524– 537.
Beach WS. 1918. The Fusarium wilt of china aster. Annual Report of the
Michigan Academy of Science 29: 281–308.
Belabid L, Baum M, Fortas Z, et al. 2004. Pathogenic and genetic charac-
terization of Algerian isolates of Fusarium oxysporum f. sp. lentis by RAPD
and AFLP analysis. African Journal of Biotechnology 3: 25– 31.
Belabid L, Fortas Z. 2002. Virulence and vegetative compatibility of Algerian
isolates of Fusarium oxysporum f. sp. lentis. Phytopathologia Mediterranea
41: 179–187.
Bell AA, Wheeler MH, Liu J, et al. 2003. United States Department of
Agriculture-Agricultural Research Service studies on polyketide toxins
of Fusarium oxysporum f. sp. vasinfectum: Potential targets for disease
control. Pest Management Science 59: 736–747.
Bennett RS, Hutmacher RB, Davis RM. 2008. Seed transmission of Fusarium
oxysporum f. sp. vasinfectum race 4 in California. Journal of Cotton Science
12: 160–164.
Bennett RS, Scott TZ, Lawrence KS, et al. 2013. Sequence characterization
of race-4-like isolates of Fusarium oxysporum from Alabama and Missis-
sippi. Journal of Cotton Science 17: 125–130.
Bergstrom GC, Kalb DW. 1995. Fusarium oxysporum f. sp. loti: A specific
wilt pathogen of birdsfoot trefoil in New York. Phytopathology 85: 1555.
Bertetti D, Ortu G, Gullino ML, et al. 2014. Contamination of seeds of Iceland
poppy (Papaver nudicaule L.) by Fusarium oxysporum. Phytoparasitica
43: 189–196.
Bertetti D, Ortu G, Gullino ML, et al. 2017. Identification of Fusarium oxyspo-
rum f. sp. opuntiarum on new hosts of the Cactaceae and Euphorbiaceae
families. Journal of Plant Pathology 99: 1–21.
Bertoldo C, Gilardi G, Spadaro D, et al. 2015. Genetic diversity and virulence
of Italian strains of Fusarium oxysporum isolated from Eustoma grandi-
florum. European Journal of Plant Pathology 141: 83– 97.
Bhide VP, Uppal BN. 1948. A new Fusarium disease of lang (Lathyrus sativus).
Phytopathology 38: 560– 567.
Bilaǐ VI. 1955. The Fusaria (biology and systematics). National Academy of
Sciences of Ukraine SSR, Kiev, USSR.
Bilju VC, Fokkens L, Houterman PM, et al. 2017. Multiple evolutionary tra-
jectories have led to the emergence of races in Fusarium oxysporum f. sp.
lycopersici. Applied and Environmental Microbiology 83: e02548-16.
Black LL, Green SK, Hartman GL, et al. 1993. Enfermedades del Chile.
Una Guía de Campo. Ed. Centro Asiatico de Investigación y Desarrollo
Vegetal (AVRDC).
Blok WJ, Bollen GJ. 1997. Host specificity and vegetative compatibility of
Dutch isolates of Fusarium oxysporum f. sp. asparagi. Canadian Journal
of Botany 75: 383– 393.
Boerema GH, Hamers MEC. 1989. Check-list for scientific names of common
parasitic fungi. Series 3b: Fungi on bulbs: Amaryllidaceae and Iridaceae.
Netherlands Journal of Plant Pathology 95 (suppl. 3): 1–32.
Bogale M, Wingfield BD, Wingfield MJ, et al. 2007. Species-specific primers
for Fusarium redolens and a PCR-RFLP technique to distinguish among
three clades of Fusarium oxysporum. FEMS Microbiology Letters 271:
27–32.
Bolley HL. 1901. A preliminary note on the cause of flax-sick soil. Fusarium
lini sp. nov. Proceedings of the Annual Meeting of the Society for the Pro-
motion of Agricultural Science 22: 42– 46.
Bolton AT, Nuttall VW, Lyall LH. 1966. A new race of Fusarium oxysporum f. sp.
pisi. Canadian Journal of Plant Science 46: 343– 347.
Booth C. 1971. The genus Fusarium. Commonwealth Mycological Institute,
Kew, Surrey, United Kingdom.
Booth C. 1975. The present status of Fusarium taxonomy. Annual Review
of Phytopathology 13: 83– 93.
Bosland PW, Williams PH. 1987. An evaluation of Fusarium oxysporum
from crucifers based on pathogenicity, isozyme polymorphism, vegeta-
tive compatibility, and geographic origin. Canadian Journal of Botany 65:
2067–2073.
Brandes EW. 1919. Banana wilt. Phytopathology 9: 339– 390.
Brankovics B, Van Dam P, Rep M, et al. 2017. Mitochondrial genomes reveal
recombination in the presumed asexual Fusarium oxysporum species
complex. BMC Genomics 18: 735.
Bruen TC, Phillippe H, Bryant D. 2006. A simple and robust statistical test
for detecting the presence of recombination. Genetics 172: 2665– 2681.
Bugnicourt F. 1939. Les Fusarium et Cylindrocarpon de l’Indochine. Ency-
clopédie Mycologique 11: 1–206.
Buxton EW. 1955. The taxonomy and variation in culture of Fusarium oxy-
sporum from gladiolus. Transactions of the British Mycological Society
38: 202– 212.
39
L. Lombard et al.: Epitypification of Fusarium oxysporum
Cai G, Gale R, Schneider RW, et al. 2003. Origin of race 3 of Fusarium
oxysporum f. sp. lycopersici at a single site in California. Phytopathology
93: 1014–1022.
Carbone I, Kohn LM. 1999. A method for designing primer sets for speciation
studies in filamentous ascomycetes. Mycologia 91: 553– 556.
Carlesse F, Amaral A-PC, Gonçalves SS, et al. 2017. Outbreak of Fusarium
oxysporum infections in children with cancer: an experience with 7 epi-
sodes of catheter-related fungemia. Antimicrobial Resistance and Infection
Control 6: 93.
Chakrabarti A, Rep M, Wang B, et al. 2011. Variation in potential effector
genes distinguishing Australian and non-Australian isolates of the cotton
wilt pathogen Fusarium oxysporum f. sp. vasinfectum. Plant Pathology
60: 232– 243.
Chang DC, Grant GB, O’Donnell K, et al. 2006. Multistate outbreak of
Fusarium keratitis associated with use of a contact lens solution. JAMA
296: 953– 963.
Chatterjee C, Rai JN. 1974. Fusarium wilt of Eruca vesicaria – observation
on comparative pathogenicity of some strains of Fusarium oxysporum.
Indian Phytopathology 28: 309– 311.
Chen Q, Ji, X, Sun W. 1985. Identification of races of cotton wilt Fusarium in
China. Agricultural Sciences in China 6: 1–6.
Chen WQ, Swart WJ. 2001. Genetic variation among Fusarium oxysporum
isolates associated with root rot of Amaranthus hybridus in South Africa.
Plant Disease 85: 1076–1080.
Chen ZD, Huang RK, Li QQ, et al. 2015. Development of pathogenicity and
AFLP to characterize Fusarium oxysporum f. sp. momordicae isolates from
bitter gourd in China. Journal of Phytopathology 163: 202– 211.
Chiocchetti A, Ghignone S, Minuto A, et al. 1999. Identification of Fusarium
oxysporum f. sp. basilici isolated from soil, basil seed, and plants by RAPD
analysis. Plant Disease 83: 576– 581.
Chiocchetti A, Sciaudone L, Durando F, et al. 2001. PCR detection of Fusa-
rium oxysporum f. sp. basilici on basil. Plant Disease 85: 607–611.
Cianchetta AN, Allen TW, Hutmacher RB, et al. 2015. Survey of Fusarium
oxysporum f. sp. vasinfectum in the United States. Journal of Cotton
Science 19: 328– 336.
Coddington A, Matthews PM, Cullis C, et al. 1987. Restriction digest patterns
of total DNA from different races of Fusarium oxysporum f. sp. pisi – an im-
proved method for race classification. Journal of Phytopathology 118: 9 –20.
Cohen SI. 1946. A wilt and root rot of Asparagus officinalis L. var. altilis L.
Phytopathology 36: 397–397.
Correll JC. 1991. The relationship between formae speciales, races, and
vegetative compatibility groups in Fusarium oxysporum. Phytopathology
81: 1061–1064.
Correll JC, Klittich CJR, Leslie JF. 1987. Nitrate nonutilizing mutants of
Fusarium oxysporum and their use in vegetative compatibility tests. Phyto-
pathology 77: 1640–1646.
Correll JC, Puhalla JE, Schneider RW. 1986. Identification of Fusarium oxy-
sporum f. sp. apii on the basis of colony size, virulence, and vegetative
compatibility. Phytopathology 76: 396–400.
Covey PA, Kuwitzky B, Hanson M, et al. 2014. Multilocus analysis using
putative fungal effectors to describe a population of Fusarium oxysporum
from sugar beet. Phytopathology 104: 886– 896.
Crall JM. 1963 Physiological specialization in Fusarium oxysporum f. sp.
niveum. Phytopathology 53: 873.
Cramer RA, Byrne PF, Brick MA, et al. 2003. Characterization of Fusarium
oxysporum isolates from common bean and sugar beet using pathogeni-
city assays and random-amplified polymorphic DNA markers. Journal of
Phytopathology 151: 352– 360.
Crous PW, Groenewald JZ, Risède JM, et al. 2004. Calonectria species
and their Cylindrocladium anamorphs: species with sphaeropedunculate
vesicles. Studies in Mycology 50: 415– 430.
Crous PW, Verkley GJM, Groenewald JZ, et al. 2009. Fungal Biodiversity.
CBS Laboratory manual Series. Westerdijk Fungal Biodiversity Institute,
Utrecht, The Netherlands.
Crowhurst RN, King FY, Hawthorne BT, et al. 1995. RAPD characterization of
Fusarium oxysporum associated with wilt of angsana (Pterocarpus indicus)
in Singapore. Mycological Research 99: 14–18.
Crutcher FK, Doan HK, Bell AA, et al. 2016. Evaluation of methods to detect
the cotton wilt pathogen Fusarium oxysporum f. sp. vasinfectum race 4.
European Journal of Plant Pathology 144: 225– 230.
Cutuli MT, Gibello A, Rodriquez-Bertos A, et al. 2015. Skin and subcuta-
neous mycoses in tilapia (Oreochromis niloticus) caused by Fusarium
oxysporum in coinfection with Aeromonas hydrophila. Medical Mycology
Case Reports 9: 7–11.
Czislowski E, Fraser-Smith S, Zander M, et al. 2017. Investigation of the
diversity of effector genes in the banana pathogen, Fusarium oxysporum f.
sp. cubense, reveals evidence of horizontal gene transfer. Molecular Plant
Pathology 19: 1155–1171.
Da Silva FP, Vechiato MH, Harakava R. 2014. EF-1α gene and IGS rDNA
sequencing of Fusarium oxysporum f. sp. vasinfectum and F. oxysporum
f. sp. phaseoli reveals polyphyletic origins of strains. Tropical Plant Patho-
logy 39: 64–73.
Datta S, Choudhary RG, Shamin M, et al. 2011. Molecular diversity in Indian
isolates of Fusarium oxysporum f. sp. lentis inciting wilt disease in lentil
(Lens culinaris Medik). African Journal of Biotechnology 10: 7314–7323.
Davis RD, Moore NY, Kochman JK. 1996. Characterisation of a population of
Fusarium oxysporum f. sp. vasinfectum causing wilt of cotton in Australia.
Australian Journal of Agricultural Research 47: 1143–1156.
De Haan LAM, Numansen A, Roebroeck EJA, et al. 2000. PCR detection of
Fusarium oxysporum f. sp. gladioli race 1, causal agent of Gladiolus yellows
disease, from infected corms. Plant Pathology 49: 89–100.
De Sousa MV, Machado J da C, Simmons HE, et al. 2015. Real-time quan-
titative PCR assays for the rapid detection and quantification of Fusarium
oxysporum f. sp. phaseoli in Phaseolus vulgaris (common bean) seeds.
Plant Pathology 64: 478– 488.
De Vega-Bartol JJ, Martín-Dominguez R, Ramos B, et al. 2011. New viru-
lence groups in Fusarium oxysporum f. sp. phaseoli: The expression of
the gene coding for the transcription factor ftf1 correlates with virulence.
Phytopathology 101: 470– 479.
Dean R, Van Kan JAL, Pretorius ZA, et al. 2012. The top 10 fungal pathogens
in molecular plant pathology. Molecular Plant Pathology 13: 414–430.
Deighton FC, Stevenson JA, Cummins GB. 1962. Formae speciales and
the Code. Taxon 11: 70–71.
Demers JE, Garzón CD, Jiménez-Gasco MM. 2014. Striking genetic similarity
between races of Fusarium oxysporum f. sp. ciceris confirms a monophyletic
origin and clonal evolution of the chickpea vascular wilt pathogen. European
Journal of Plant Pathology 139: 309– 324.
Desjardins AE. 2006. Fusarium mycotoxins chemistry, genetics and biology.
APS Press, St. Paul, USA.
Di Pietro A, Madrid MP, Caracuel Z, et al. 2003. Fusarium oxysporum: ex-
ploring the molecular arsenal of a vascular wilt fungus. Molecular Plant
Pathology 4: 315– 325.
Di Primo P, Cappelli C, Katan T. 2002. Vegetative compatibility groupings of
Fusarium oxysporum f. sp. gladioli from saffron. European Journal of Plant
Pathology 108: 869– 875.
Di Primo P, Cartia G, Katan T. 2001. Vegetative compatibility and hetero-
karyon in Fusarium oxysporum f. sp. radicis-lycopersici from Italy. Plant
Pathology 50: 371–382.
Dita MA, Waalwijk C, Buddenhagen IW, et al. 2010. A molecular diagnostic
for tropical race 4 of the banana fusarium wilt pathogen. Plant Pathology
59: 348– 357.
Doan HK, Zhang S, Davis RM. 2014. Development and evaluation of Ampli-
fyRP Acceler8 diagnostic assay for the detection of Fusarium oxysporum
f. sp. vasinfectum race 4 in cotton. Plant Health Progress. doi: https://doi.
org/10.1094 /PHP-RS-13-0115.
Dong Z, Hsiang T, Luo M, et al. 2017. Draft genome sequence of an isolate
of Fusarium oxysporum f. sp. melongenae, the causal agent of Fusarium
wilt of eggplant. Genome Announcements 5: e01597-16.
Dos Santos Silva A, De Oliveira EJ, Haddad F, et al. 2013. Molecular
fingerprinting of Fusarium oxysporum f. sp. passiflorae isolates using AFLP
markers. Scientia Agricola 70: 108–115.
Dubey SC, Priyanka K, Singh V, et al. 2012. Race profiling and molecular
diversity analysis of Fusarium oxysporum f. sp. ciceris causing wilt of
chickpea. Journal of Phytopathology 160: 576– 587.
Dubey SC, Singh SR. 2008. Virulence analysis and oligonucleotide finger-
printing to detect diversity among Indian isolates of Fusarium oxysporum
f. sp. ciceris causing chickpea wilt. Mycopathologia 165: 389– 406.
Dzidzariya OM. 1968. Measures for the control of fusariosis of East Indian
basil [in Russian]. (Translated in Review of Applied Mycology 48: 882
(1969).)
Edel-Hermann V, Sautour M, Gautheron N, et al. 2016. A clonal lineage of
Fusarium oxysporum circulates in the tap water of different French hospitals.
Applied and Environmental Microbiology 82: 6483– 6489.
Egamberdiev SS, Salahutdinov IB, Abdullaev AA, et al. 2014. Detection of
Fusarium oxysporum f. sp. vasinfectum race 3 by single-base extension
method and allele-specific polymerase chain reaction. Canadian Journal
of Plant Pathology 36: 216– 223.
Egamberdiev SS, Ulloa M, Saha S, et al. 2013. Molecular characterization
of Uzbekistan isolates of Fusarium oxysporum f. sp. vasinfectum. Journal
of Plant Science & Molecular Breeding. doi: https:// doi.org/10.7243/ 2050-
2389-2-3.
Elias KS, Schneider RW. 1991. Vegetative compatibility groups in Fusarium
oxysporum f. sp. lycopersici. Phytopathology 81: 159–162.
Elias KS, Schneider RW. 1992. Genetic diversity within and among races
and vegetative compatibility groups of Fusarium oxysporum f. sp. lycoper-
sici as determined by isozyme analysis. Phytopathology 82: 1421–1427.
40 Persoonia – Volume 43, 2019
Elias KS, Zamir D, Lichtman-Pleban T, et al. 1993. Population structure of
Fusarium oxysporum f. sp. lycopersici: restriction fragment length poly-
morphisms provide genetic evidence that vegetative compatibility group
is an indicator of evolutionary origin. Molecular Plant-Microbe Interaction
6: 565– 572.
Elliott ML, Des Jardin EA, Harmon CL, et al. 2017. Confirmation of Fusarium
wilt caused by Fusarium oxysporum f. sp. palmarum on × Butyagrus nabon-
nandii (mule palm) in Florida. Plant Disease 101: 381.
Elliott ML, Des Jardin EA, O’Donnell K, et al. 2010. Fusarium oxysporum f. sp.
palmarum, a novel forma specialis causing a lethal disease of Syagrus
romanzoffiana and Washingtonia robusta in Florida. Plant Disease 94:
31–38.
Elmer WH, Stephens CT. 1989. Classification of Fusarium oxysporum f. sp.
asparagi into vegetatively compatible groups. Phytopathology 79: 88– 93.
Elmer WH, Wick RL, Haviland P. 1994. Vegetative compatibility among
Fusarium oxysporum f. sp. basilicum isolates recovered from basil seed
and infected plants. Plant Disease 78: 789–791.
Elzein A, Brändle F, Cadisch G, et al. 2008. Fusarium oxysporum strains as
potential Striga mycoherbicides: molecular characterization and evidence
for a new forma specialis. The Open Mycology Journal 2: 89– 93.
Elzein A, Kroschel J. 2006. Host range studies of Fusarium oxysporum Foxy
2: an evidence for a new forma specialis and its implications for Striga
control. Journal of Plant Diseases and Protection 20: 875– 887.
Enya J, Togawa M, Takeuchi T, et al. 2008. Biological and phylogenetic
characterization of Fusarium oxysporum complex, which causes yellows on
Brassica spp., and proposal of Fusarium oxysporum f. sp. rapae, a novel
forma specialis pathogenic on Brassica rapa L. in Japan. Phytopathology
98: 475–483.
Epstein L, Kaur S, Chang PL, et al. 2017. Races of the celery pathogen
Fusarium oxysporum f. sp. apii are polyphyletic. Phytopathology 107:
463– 473.
Erwin DC. 1958. Fusarium lateritium f. ciceri incitant of Fusarium wilt of Cicer
arietinum. Phytopathology 48: 498– 501.
Fang X, Jost R, Finnegan PM, et al. 2013. Comparative proteome analysis
of the strawberry-Fusarium oxysporum f. sp. fragariae pathosystem reveals
early activation of defence responses as a crucial determinant of host
resistance. Journal of Proteome 12: 1772–1788.
Feather TV, Ohr HD, Munnecke DE. 1979. Wilt and dieback of Canary Island
palm in California. California Agriculture 33: 19– 20.
Fernandez D, Assigbetse K, Dubois MP, et al. 1994. Molecular characteriza-
tion of races and vegetative compatibility groups in Fusarium oxysporum
f. sp. vasinfectum. Applied and Environmental Microbiology 60: 4039– 4046.
Fernandez D, Ouinten M, Tantaoui A, et al. 1998. Fot 1 insertions in the
Fusarium oxysporum f. sp. albedinis genome provide diagnostic PCR tar-
gets for detection of the date palm pathogen. Applied and Environmental
Microbiology 64: 633– 636.
Fisher NL, Burgess LW, Toussoun TA, et al. 1982. Carnation leaves as a
substrate and for preserving cultures of Fusarium species. Phytopathology
72: 151–153.
Flood J. 2006. A review of Fusarium wilt of oil palm caused by Fusarium
oxysporum f. sp. elaeidis. Phytopathology 96: 660– 662.
Foster V. 1955. Fusarium wilt of cattleyas. Phytopathology 45: 599–602.
Fourie G, Steenkamp ET, Ploetz RC, et al. 2011. Current status of the taxo-
nomic position of Fusarium oxysporum formae specialis cubense within
the Fusarium oxysporum complex. Infection, Genetics and Evolution 11:
533– 542.
Fujinaga M, Ogiso H, Shinohara H, et al. 2005. Phylogenetic relationships
between the lettuce root rot pathogen Fusarium oxysporum f. sp. lactucae
races 1, 2, and 3 based on the sequence of the intergenic spacer region
of its ribosomal DNA. Journal of General Plant Pathology 71: 402– 407.
Fujinaga M, Ogiso H, Tsuchiya N, et al. 2001. Physiological specialization
of Fusarium oxysporum f. sp. lactucae, a causal organism of Fusarium
root rot of crisp head lettuce in Japan. Journal of General Plant Pathology
67: 205– 206.
Fujinaga M, Ogiso H, Tsuchiya N, et al. 2003. Race 3, a new race of Fusa-
rium oxysporum f. sp. lactucae determined by a differential system with
commercial cultivars. Journal of General Plant Pathology 69: 23– 28.
Fujinaga M, Yamagishi N, Ishiyama Y, et al. 2014. PCR-based race differ-
entiation of Fusarium oxysporum f. sp. lactucae. Annual Report of the
Kanto-Tosan Plant Protection Society 61: 47–49.
Gabe HL. 1975. Standardization of nomenclature for pathogenic races of
Fusarium oxysporum f. sp. lycopersici. Transactions of the British Mycologi-
cal Society 64: 156–159.
Galván GA, Koning-Boucoiran CFS, Koopman WJM, et al. 2008. Genetic
variation among Fusarium isolates from onion, and resistance to Fusarium
basal rot in related Allium species. European Journal of Plant Pathology
121: 499– 512.
García-Pedrasjas MD, Bainbridge BW, Heale JB, et al. 1999. A simple PCR-
based method for the detection of the chickpea-wilt pathogen Fusarium
oxysporum f.sp. ciceris in artificial and natural soils. European Journal of
Plant Pathology 105: 251–259.
Gardner DE. 1980. Acacia koa seedling wilt caused by Fusarium oxysporum
f. sp. koae f. sp. nov. Phytopathology 70: 594–597.
Garibaldi A. 1975. Race differentiation in Fusarium oxysporum f. sp. dianthi
(Prill et Del.) Snyd. et Hans. First contribution. Mededelingen Faculteit
Landbouwwetenschappen Rijksuniversiteit Gent 40: 531–537.
Garibaldi A. 1977. Race differentiation in Fusarium oxysporum f. sp. dianthi
and varietal susceptibility. Acta Horticulturae 71: 97–101.
Garibaldi A. 1983. Resistenza di cultivar di garofano nei confronti di otto
patotipi di Fusarium oxysporum f. sp. dianthi (Prill. et Del.) Snyd. et Hans.
Rivista della Ortoflorofrutticoltura Italiana 67: 261–270.
Garibaldi A, Gullino ML. 1985. Wilt of Ranunculus asiaticus caused by
Fusarium oxysporum f. sp. ranunculi, forma specialis nova. Phytopathologia
Mediterranea 24: 213– 214.
Gawehns F, Houterman PM, Ichou FA, et al. 2014. The Fusarium oxysporum
effector Six6 contributes to virulence and suppresses I-2-mediated cell
death. Molecular Plant-Microbe Interactions 27: 336– 348.
Gerdemann JW, Finley AM. 1951. The pathogenicity of races 1 and 2 of
Fusarium oxysporum f. lycopersici. Phytopathology 41: 238– 244.
Geiser DM, Aoki T, Bacon CW, et al. 2013. One fungus, one name: Defining
the genus Fusarium in a scientifically robust way that preserves longstand-
ing use. Phytopathology 103: 400– 408.
Gerlach W. 1954. Untersuchungen über die Welkekrankheit des Alpen-
veilchens. Phytopathologische Zeitschrift 22: 125–176.
Gerlach W, Nirenberg H. 1982. The genus Fusarium – a pictorial atlas. Mit-
teilungen aus der Biologischen Bundesanstalt für Land- und Forstwirtschaft
Berlin-Dahlem 209: 1–406.
Gerlagh M, Blok WJ. 1988. Fusarium oxysporum f. sp. cucurbitacearum n.f.
embracing all formae speciales of F. oxysporum attacking cucurbitaceous
crops. Netherlands Journal of Plant Pathology 94: 17–31.
Gerlagh M, Ester A. 1985. Fusarium oxysporum f. sp. benincasae, a new
adaptation of Fusarium oxysporum to a cucurbitaceous crop. Mededelingen
van de Faculteit Landbouwwetenschappen 37: 1048.
Gherbawy YAMH. 1999. RAPD profile analysis of isolates belonging to dif-
ferent formae speciales of Fusarium oxysporum. Cytologia 64: 269– 276.
Ghosh R, Nagavardhini A, Sengupta A, et al. 2015. Development of loop-
mediated isothermal amplification (LAMP) assay for rapid detection of
Fusarium oxysporum f. sp. ciceris – wilt pathogen of chickpea. BMC
Research Notes 8: 40.
Giesbrecht M, McCarthy M, Elliott ML, et al. 2013. First report of Fusarium
oxysporum f. sp. palmarum in Texas causing Fusarium wilt of Washington
robusta. Plant Disease 97: 1511.
Gilardi G, Franco Ortega S, Van Rijswick PCJ, et al. 2017. A new race
of Fusarium oxysporum f. sp. lactucae of lettuce. Plant Pathology 66:
677–688.
Gordon TR. 2017. Fusarium oxysporum and the Fusarium wilt syndrome.
Annual Review of Phytopathology 55: 23– 39.
Gordon TR, Martyn RD. 1997. The evolutionary biology of Fusarium oxyspo-
rum. Annual Review of Phytopathology 35: 111–128.
Gordon WL. 1965. Pathogenic strains of Fusarium oxysporum. Canadian
Journal of Botany 43: 1309–1318.
Grajal-Martin MJ, Simon CJ, Muehlbauer FJ. 1993. Use of random amplified
polymorphic DNA (RAPD) to characterize race 2 of Fusarium oxysporum
f. sp. pisi. Phytopathology 83: 612– 614.
Groenewald JZ, Nakashima C, Nishikawa J, et al. 2013. Species concepts
in Cercospora: spotting the weeds among the roses. Studies in Mycology
75: 115–170.
Guarro J. 2013. Fusariosis, a complex infection caused by a high diversity
of fungal species refractory to treatment. European Journal of Clinical
Microbiology & Infectious Diseases 32: 1491–1500.
Guarro J, Gene J. 1995. Opportunistic Fusarial infections in humans. Euro-
pean Journal of Clinical Microbiology & Infectious Diseases 14: 741–754.
Gunn LV, Summerell BA. 2002. Differentiation of Fusarium oxysporum iso-
lates from Phoenix canariensis (Canary Island date palm) by vegetative
compatibility grouping and molecular analysis. Australasian Plant Patho-
logy 31: 351–358.
Guo Q, Li S, Lu X, et al. 2015. Identification of a new genotype of Fusarium
oxysporum f. sp. vasinfectum on cotton in China. Plant Disease 99: 1569–
1577.
Gupta D, Chowdhry PN, Padhi B. 1982. A new form of Fusarium oxysporum
on Sansevieria cylindrica. Indian Phytopathology 35: 695– 696.
Gupta M, Jarial K, Vikram A. 2014. Morphological, cultural, pathological and
molecular variability among Fusarium oxysporum f. sp. zingiberti isolates.
International Journal of Bio-resource and Stress Management 5: 375 –380.
41
L. Lombard et al.: Epitypification of Fusarium oxysporum
Gupta PKS. 1974. Plant pathogenic species of the genus Fusarium in India.
Nova Hedwigia 25: 699–717.
Gupta VK. 2012. PCR-RAPD profiling of Fusarium spp. causing guava wilt
disease in India. Journal of Environmental Science and Health 47: 315– 325.
Gurjar G, Barve M, Giri A, et al. 2009. Identification of Indian pathogenic races
of Fusarium oxysporum f. sp. ciceris with gene specific, ITS and random
markers. Mycologia 101: 484– 495.
Hadar E, Katan J, Katan T. 1989. The use of nitrate-nonutilizing mutants and a
selective medium for studies of pathogenic strains of Fusarium oxysporum.
Plant Disease 73: 800– 803.
Haglund WA, Kraft JM. 1979. Fusarium oxysporum f. sp. pisi, race 6: Occur-
rence and distribution. Phytopathology 69: 818– 820.
Hannachi I, Poli A, Rezgui S, et al. 2015. Genetic and phenotypic differences
of Fusarium oxysporum f. sp. citri isolated from sweet orange and tangerine.
European Journal of Plant Pathology 142: 269– 280.
Hansen FT, Gardiner DM, Lysøe E, et al. 2015. An update to polyketide syn-
thase and non-ribosomal synthetase genes and nomenclature in Fusarium.
Fungal Genetics and Biology 75: 20– 29.
Hanzawa J. 1914. Fusarium cepae, ein neuer zweibelpilz Japans, sowie
einige andere pilze an zweibelpflanzen. Mykologie Zentralblatt 5: 4–13.
Hare WW. 1953. A new race causing wilt of cowpea. Phytopathology 43: 291.
Hartig R. 1892. Ein neuer Keimlingspilz. Zeitschrift für Naturwissenschaften
1: 432– 436.
Harveson RM, Rush CM. 1997. Genetic variation among Fusarium oxyspo-
rum isolates from sugar beet as determined by vegetative compatibility.
Plant Disease 81: 85– 88.
Haware MP, Nene YL. 1982. Races of Fusarium oxysporum f. sp. ciceri.
Plant Disease 66: 809– 810.
Hemo I, Pe’Er J, Polacheck I. 1989. Fusarium oxysporum keratitis. Ophthal-
mologica 198: 3–7.
Henrique FH, Carbonell SAM, Ito MF, et al. 2015. Classification of physio-
logical races of Fusarium oxysporum f. sp. phaseoli in common bean.
Bragantia 74: 84– 92.
Henry PM, Kirkpatrick SC, Islas CM, et al. 2017. The population of Fusarium
oxysporum f. sp. fragariae, cause of Fusarium wilt of strawberry, in Cali-
fornia. Plant Disease 101: 550– 556.
Hibar K, Edel-Herman V, Steinberg C, et al. 2007. Genetic diversity of
Fusarium oxysporum populations isolated from tomato plants in Tunisia.
Journal of Phytopathology 155: 136–142.
Hill AL, Reeves PA, Larson RL, et al. 2011. Genetic variability among isolates
of Fusarium oxysporum from sugar beet. Plant Pathology 60: 496– 505.
Hillis DM, Bull JJ. 1993. An empirical test of bootstrapping as a method for
assessing confidence in phylogenetic analysis. Systematic Biology 42:
182–192.
Hirano Y, Arie T. 2006. PCR-based differentiation of Fusarium oxysporum
ff. sp. lycopersici and radicis-lycopersici and races of F. oxysporum f. sp.
lycopersici. Journal of General Plant Pathology 72: 273– 283.
Hirano Y, Arie T. 2009. Variation and phylogeny of Fusarium oxysporum
isolates based on nucleotide sequences of polygalacturonase genes.
Microbes and Environments 24: 113–120.
Hirooka Y, Matsumoto Y, Ebihara Y, et al. 2008. Occurrence of Fusarium wilt
in Japan caused by Fusarium oxysporum f. sp. tanaceti. Japanese Journal
of Phytopathology 74: 7–12.
Holmes EA, Bennett RS, Spurgeon DW, et al. 2009. New genotypes of
Fusarium oxysporum f. sp. vasinfectum from the southeastern United
States. Plant Disease 93: 1298–1304.
Honnareddy N, Dubey SC. 2006. Pathogenic and molecular characterization
of Indian isolates of Fusarium oxysporum f. sp. ciceris causing chickpea
wilt. Current Science 91: 661–666.
Hood JR, Stewart RN. 1957. Factors affecting symptom expression in
Fusarium wilt of Dianthus. Phytopathology 47: 173–178.
Huang CH, Roberts PD, Gale LR, et al. 2013. Population structure of Fusa-
rium oxysporum f. sp. radicis-lycopersici in Florida inferred from vegeta-
tive compatibility groups and microsatellites. European Journal of Plant
Pathology 136: 509– 521.
Huang HC, Phillippe LM, Marshall HH, et al. 1992. Wilt of hardy chrysanthe-
mum caused by a new race of Fusarium oxysporum f. sp. chrysanthemi.
Plant Pathology 1: 57–61.
Huang L-W, Wang C-J, Lin Y-S, et al. 2014. Stem rot of jewel orchids caused
by a new forma specialis, Fusarium oxysporum f. sp. anoectochili in Taiwan.
Plant Pathology 63: 539– 547.
Hubbard JC, Gerik JS. 1993. A new wilt disease of lettuce incited by Fusarium
oxysporum f. sp. lactucum forma specialis nov. Plant Disease 77: 750–754.
Hungerford CW. 1923. A Fusarium wilt of spinach. Phytopathology 13:
205– 209.
Huson DH, Bryant D. 2006. Application of phylogenetic networks in evolution-
ary studies. Molecular Biology and Evolution 23: 254– 267.
Ibrahim FM. 1966. A new race of cotton-wilt Fusarium in the Sudan Gezira.
Empire Cotton Growing Review 43: 296– 299.
Imle EP. 1942. Bulbrot disease of lilies. The Lily Yearbook of the American
Horticultural Society: 30– 41.
Inami K, Yoshioka C, Hirano Y, et al. 2010. Real-time PCR for differential
determination of the tomato wilt fungus, Fusarium oxysporum f. sp. lyco-
persici, and its races. Journal of General Plant Pathology 76: 116–121.
Jacobson DJ, Gordon TR. 1988. Vegetative compatibility and self-incom-
patibility within Fusarium oxysporum f. sp. melonis. Phytopathology 78:
668– 672.
Jacobson DJ, Gordon TR. 1990a. Further investigations of vegetative com-
patibility within Fusarium oxysporum f. sp. melonis. Canadian Journal of
Botany 68: 1245–1248.
Jacobson DJ, Gordon TR. 1990b. Variability of mitochondrial DNA as an in-
dicator of relationships between populations of Fusarium oxysporum f. sp.
melonis. Mycological Research 94: 734–744.
Janardhanan KK, Ganguly D, Husain A. 1964. Fusarium wilt of Rauvolfia
serpentina. Current Science 33: 313.
Janson BF. 1951. A new disease of dill. Phytopathology 41: 19.
Jarvis WR, Shoemaker RA. 1978. Taxonomic status of Fusarium oxysporum
causing foot and root rot of tomato. Phytopathology 68: 1679–1680.
Jelinski NA, Broz K, Jonkers W, et al. 2017. Effector gene suites in some soil
isolates of Fusarium oxysporum are not sufficient predictors of vascular
wilt of tomato. Phytopathology 107: 842– 851.
Jiang Y, Al-Hatmi AMS, Xiang Y, et al. 2016. The concept of ecthyma gan-
grenosum illustrated by a Fusarium oxysporum infection in an immuno-
competent individual. Mycopathologia 181: 759–763.
Jiménez-Gasco MM, Jiménez-Díaz RM. 2003. Development of a specific
polymerase chain reaction-based assay for the identification of Fusarium
oxysporum f. sp. ciceris and its pathogenic races 0, 1A, 5, and 6. Phyto-
pathology 93: 200– 209.
Jiménez-Gasco MM, Milgroom MG, Jiménez-Díaz RM. 2002. Gene gene a-
logies support Fusarium oxysporum f. sp. ciceris as a monophyletic group.
Plant Pathology 51: 72–77.
Jiménez-Gasco MM, Milgroom MG, Jiménez-Díaz RM. 2004a. Stepwise
evolution of races in Fusarium oxysporum f. sp. ciceris inferred from
fingerprinting with repetitive DNA sequences. Phytopathology 94: 228–235.
Jiménez-Gasco MM, Navas-Cortés JA, Jiménez-Díaz RM. 2004b. The
Fusarium oxysporum f. sp. ciceris/ Cicer arietinum pathosystem: a case
study of the evolution of plant-pathogenic fungi into races and pathotypes.
International Microbiology 7: 95–104.
Jiménez-Gasco MM, Pérez-Artés E, Jiménez-Díaz RM. 2001. Identification
of pathogenic races 0, 1B/ C, 5, and 6 of Fusarium oxysporum f. sp. ciceris
with random amplified polymorphic DNA (RAPD). European Journal of
Plant Pathology 107: 237–248.
Johnson J. 1921. Fusarium-wilt of tobacco. Journal of Agriculture Research
20: 515– 535.
Kappelman AJ. 1983. Distribution of races of Fusarium oxysporum f. sp.
vasinfectum within the United States. Plant Disease 67: 1229–1231.
Kashiwa T, Inami K, Teraoka T et al. 2016. Detection of cabbage yellows
fungus Fusarium oxysporum f. sp. conglutinans in soil by PCR and real-time
PCR. Journal of General Plant Pathology 82: 240– 247.
Katan T. 1999. Current status of vegetative compatibility groups in Fusarium
oxysporum. Phytoparasitica 27: 51–64.
Katan T, Di Primo P. 1999. Current status of vegetative compatibility groups
in Fusarium oxysporum: Supplement (1999). Phytoparasitica 27: 273– 277.
Katan T, Katan J. 1988. Vegetative-compatibility grouping of Fusarium oxy-
sporum f. sp. vasinfectum from tissue and the rhizosphere of cotton plants.
Phytopathology 78: 852– 855.
Katan T, Katan J, Gordon TR, et al. 1994. Physiological races and vegeta-
tive compatibility groups of Fusarium oxysporum f. sp. melonis in Israel.
Phytopathology 84: 153–157.
Katan T, Zamir D, Sarfatti M, et al. 1991. Vegetative compatibility groups and
subgroups in Fusarium oxysporum f. sp. radicis-lycopersici. Phytopatho-
logy 81: 255– 262.
Katoh K, Rozewicki J, Yamada KD. 2017. MAFFT online service: multiple
sequence alignment, interactive sequence choice and visualization. Brief-
ings in Bioinformatics: 1–7. doi: https://doi.org/10.1093/bib/ bbx108.
Kawabe M, Katsube K, Yoshida T, et al. 2007. Genetic diversity of Fusarium
oxysporum f. sp. spinaciae in Japan based on phylogenic analyses of
rDNA-IGS and MAT1 sequences. Journal of General Plant Pathology 73:
353– 359.
Kawai I, Suzuki H, Kawai K. 1958. On the pathogenicity of wilt Fusarium of
the cucurbitaceous plants and their forms. Schizuoka Agricultural Experi-
ment Station Bulletin 3: 49– 68.
Kearse M, Moir R, Wilson A, et al. 2012. Geneious Basic: an integrated and
extendable desktop software platform for the organization and analysis of
sequence data. Bioinformatics 28: 1647–1649.
42 Persoonia – Volume 43, 2019
Kelly A, Alcalá-Jiménez AR, Bainbridge BW, et al. 1994. Use of genetic
fingerprinting and random amplified polymorphic DNA to characterize
pathotypes of Fusarium oxysporum f. sp. ciceris infecting chickpea. Phyto-
pathology 84: 1293–1298.
Kelly AG, Bainbridge BW, Heale JB, et al. 1998. In planta-polymerase-chain-
reaction detection of the wilt-inducing pathotype of Fusarium oxysporum
f. sp. ciceris in chickpea (Cicer arietinum L.). Physiological and Molecular
Plant Pathology 52: 397–409.
Kendrick JB, Snyder WC. 1942a. Fusarium wilt of radish. Phytopathology
32: 1031–1033.
Kendrick JB, Snyder WC. 1942b. Fusarium yellows of beans. Phytopatho-
logy 32: 1010–1014.
Khetan S, Khetan P, Katkar V, et al. 2018. Urinary tract infection due to
Fusarium oxysporum in an immunocompetent patient with chronic kidney
disease. Journal of Biomedical Research 32: 157–160.
Killian C, Maire R. 1930. Le Bayoud, maladie du dattier. Bulletin de la Société
d’Histoire Naturelle de l’Afrique du Nord 21: 89–101.
Kim DH, Martyn RD, Magill CW. 1993. Mitochondrial DNA (mtDNA) – Related-
ness among formae speciales of Fusarium oxysporum in the Cucurbitaceae.
Phytopathology 83: 91–97.
Kim H, Hwang SM, Lee JH, et al. 2017. Specific PCR detection of Fusarium
oxysporum f. sp. raphani: a causal agent of Fusarium wilt on radish plants.
Plant Pathology 65: 133–140.
Kim HJ, Choi YK, Min BR. 2001. Variation of the intergenic spacer (IGS)
region of ribosomal DNA among Fusarium oxysporum formae speciales.
Journal of Microbiology 39: 265– 272.
Kim WS, Kim WG, Cho WD, et al. 2002. Wilt of perilla caused by Fusarium
spp. Plant Pathology Journal 18: 293– 299.
Kim Y, Hutmacher RB, Davis RM. 2005. Characterization of California isolates
of Fusarium oxysporum f. sp. vasinfectum. Plant Disease 89: 366– 372.
Kistler HC. 1997. Genetic diversity in the plant-pathogenic fungus Fusarium
oxysporum. Phytopathology 87: 474– 479.
Kistler HC, Alabouvette C, Baayen RP, et al. 1998. Systematic numbering of
vegetative compatibility groups in the plant pathogenic fungus Fusarium
oxysporum. Phytopathology 88: 30– 32.
Kistler HC, Benny U. 1989. The mitochondrial genome of Fusarium oxyspo-
rum. Plasmid 22: 86– 89.
Kistler HC, Bosland PW, Benny U, et al. 1987. Relatedness of strains of
Fusarium oxysporum from crucifers measured by examination of mitochon-
drial and ribosomal DNA. Phytopathology 77: 1289–1293.
Kistler HC, Momol EA, Benny U. 1991. Repetitive genomic sequences for
determining relatedness among strains of Fusarium oxysporum. Phyto-
pathology 81: 331–336.
Kitazawa K, Yanagita K. 1984. Adzuki bean wilt caused by Fusarium oxyspo-
rum Schl.: Re-occurrence and confirmation of the causal organism. Annals
of the Phytopathological Society of Japan 50: 643– 645.
Kitazawa K, Yanagita K. 1989. Fusarium oxysporum Schl. F. sp. adzukicola
n.f. sp., a wilt fungus of Phaseolus angularis. Annals of the Phytopathologi-
cal Society of Japan 55: 76–78.
Klisiewicz JM. 1975. Race 4 of Fusarium oxysporum f. sp. carthami. Plant
Disease Reporter 59: 712–714.
Klisiewicz JM, Houston BR. 1963. A new form of Fusarium oxysporum.
Phytopathology 53: 241.
Klisiewicz JM, Thomas CA. 1970a. Pathogenic races of Fusarium oxysporum
f. sp. carthami. Phytopathology 60: 83– 84.
Klisiewicz JM, Thomas CA. 1970b. Race differentiation in Fusarium oxyspo-
rum f. sp. carthami. Phytopathology 60: 1706.
Kondo T, Chu E, Kageyama K, et al. 2013. Stem canker and wilt of delphinium
caused by Fusarium oxysporum f. sp. delphinii in Japan. Journal of General
Plant Pathology 79: 370– 373.
Koyyappurath S, Atuahiva T, Le Guen R, et al. 2016. Fusarium oxysporum
f. sp. radicis-vanillae is the causal agent of root and stem rot of vanilla.
Plant Pathology 65: 612– 625.
Kraft JM, Haglund WA. 1978. A reappraisal of the race classification of
Fusarium oxysporum f. sp. pisi. Phytopathology 68: 273– 275.
Kulkarni GS. 1934. Studies in the wilt disease of cotton in the Bombay presi-
dency. Indian Journal of Agricultural Sciences 4: 976–1045.
Kumar S, Stecher G, Tamura K. 2016. MEGA7: Molecular Evolutionary
Genetics Analysis version 7.0 for bigger datasets. Molecular Biology and
Evolution 33: 1870–1874.
Larkin RP, Hopkins DL, Martin FN. 1988. Differentiation of strains and patho-
genic races of Fusarium oxysporum f. sp. niveum based on vegetative com-
patibility. Phytopathology 88: 1542.
Larkin RP, Hopkins DL, Martin FN. 1990. Vegetative compatibility within Fusa-
rium oxysporum f. sp. niveum and its relationship to virulence, aggressive-
ness, and race. Canadian Journal of Microbiology 36: 352– 358.
Laskaris T. 1949. Fusarium stem canker and wilt of Delphinium. Phytopa-
thology 39: 913– 919.
Laurence MH, Burgess LW, Summerell BA, et al. 2012. High levels of diver-
sity in Fusarium oxysporum from non-cultivated ecosystems in Australia.
Fungal Biology 116: 289–297.
Laurence MH, Summerell BA, Burgess LW, et al. 2014. Genealogical con-
cordance phylogenetic species recognition in the Fusarium oxysporum
species complex. Fungal Biology 118: 374–384.
Laurence MH, Summerell BA, Liew ECY. 2015. Fusarium oxysporum f. sp.
canariensis: evidence for horizontal gene transfer of putative pathogenicity
genes. Plant Pathology 64: 1068–1075.
Leach JG, Currence TM. 1938. Fusarium wilt of muskmelons in Minnesota.
Minnesota Agricultural Experiment Station Technical Bulletin 129: 1– 32.
Lecomte C, Edel-Hermann V, Cannesan M-A, et al. 2016. Fusarium oxyspo-
rum f. sp. cyclaminis: underestimated genetic diversity. European Journal
of Plant Pathology 145: 421–431.
Leslie JF. 1993. Vegetative compatibility in fungi. Annual Review of Phyto-
pathology 31: 127–151.
Leslie JF, Summerell BA. 2006. The Fusarium laboratory manual. Blackwell
Publishing, Ames.
Li D, Wang L, Zhang Y, et al. 2012. Pathogenic variation and molecular charac-
terization of Fusarium species isolated from wilted sesame in China. African
Journal of Microbiology Research 6: 149–154.
Li E, Ling J, Wang G, et al. 2015. Comparative proteomics analyses of two
races of Fusarium oxysporum f. sp. conglutinans that differ in pathogenicity.
Scientific Reports 5: 13663.
Li E, Wang G, Xiao J, et al. 2016. A SIX1 homolog in Fusarium oxysporum
f. sp. conglutinans is required for full virulence on cabbage. PLoS ONE
11 (3): e0152273.
Li Y, Garibaldi A, Gullino ML. 2010. Molecular detection of Fusarium oxy-
sporum f. sp. chrysanthemi on three host plants: Gerbera jamesonii, Osteo-
spermum sp. and Argyranthemum frutescens. Journal of Plant Pathology
92: 525– 530.
Lievens B, Claes L, Vakalounakis DJ, et al. 2007. A robust identification and
detection assay to discriminate the cucumber pathogens Fusarium oxyspo-
rum f. sp. cucumerinum and f. sp. radicis-cucumerinum. Environmental
Microbiology 9: 2145– 2161.
Lievens B, Houterman PM, Rep M. 2009a. Effector gene screening allows
unambiguous identification of Fusarium oxysporum f. sp. lycopersici races
and discrimination from other formae speciales. FEMS Microbiology Let-
ters 300: 201–215.
Lievens B, Rep M, Thomma BPHJ. 2008. Recent development in the mo-
lecular discrimination of formae speciales of Fusarium oxysporum. Pest
Management Science 64: 781–788.
Lievens B, Van Baarlen P, Verreth C, et al. 2009b. Evolutionary relationships
between Fusarium oxysporum f. sp. lycopersici and Fusarium oxyspo-
rum f. sp. radicis-lypersici isolates inferred from mating type, elongation
factor-1α and exopolygalacturonase sequences. Mycological Research
113: 1181–1191.
Lin B, Shen H. 2017. Fusarium oxysporum f. sp. cubense. In: Wan F, Jiang M,
Zhan A (eds), Biological invasions and its management in China: 225 –236.
Springer, Singapore.
Lin Q, Chen Z. 1994. Identification of the causal agent of root rot of Magnolia
biloba. Acta Phytopathologica Sinica 24: 312.
Lin YH, Chen KS, Chang JY, et al. 2010. Development of the molecular
methods for rapid detection and differentiation of Fusarium oxysporum and
F. oxysporum f. sp. niveum in Taiwan. New Biotechnology 27: 409–418.
Lin YH, Lai PJ, Chang TH, et al. 2014. Genetic diversity and identification of
race 3 of Fusarium oxysporum f. sp. lactucae in Taiwan. European Journal
of Plant Pathology 140: 721–733.
Linfield CA. 1993. A rapid serological test for detecting Fusarium oxysporum
f. sp. narcissi in Narcissus. Annals of Applied Biology 123: 685–693.
Liu YJ, Whelen S, Hall BD. 1999. Phylogenetic relationships among asco-
mycetes: evidence from an RNA polymerse II subunit. Molecular Biology
and Evolution 16: 1799–1808.
Löffler HJM, Rumine P. 1991. Virulence and vegetative compatibility of
Dutch and Italian isolates of Fusarium oxysporum f. sp. lilii. Journal of
Phytopathology 132: 12– 20.
Logrieco A, Moretti A, Castellá G, et al. 1998. Beauvericin production by
Fusarium species. Applied and Environmental Microbiology 64: 3084– 3088.
Lomas-Cano T, Boix-Ruiz A, De Cara-García M, et al. 2016. Etiological and
epidemiological concerns about pepper root and lower stem rot caused
by Fusarium oxysporum f. sp. radicis-capsici f. sp. nova. Phytoparasitica
44: 283– 293.
Lomas-Cano T, Palmero-Liamas D, De Cara M, et al. 2014. First report of
Fusarium oxysporum on sweet pepper seedlings in Almería, Spain. Plant
Disease 98: 1435.
Lombard L, Van der Merwe NA, Groenewald JZ, et al. 2015. Generic concepts
in Nectriaceae. Studies in Mycology 80: 189– 245.
43
L. Lombard et al.: Epitypification of Fusarium oxysporum
López-Berges MS, Hera C, Sulyok M, et al. 2013. The velvet complex governs
mycotoxin production and virulence of Fusarium oxysporum on plant and
mammalian hosts. Molecular Microbiology 87: 49– 65.
López-Díaz C, Rahjoo V, Sulyok M, et al. 2018. Fusaric acid contributes to
virulence of Fusarium oxysporum on plant and mammalian hosts. Molecular
Plant Pathology 19: 440– 453.
Lori GA, Petiet PM, Malbrán I, et al. 2012. Fusarium wilt of cyclamen: Patho-
genicity and vegetative compatibility groups structure of the pathogen in
Argentina. Crop Protection 36: 43– 48.
Louvet J, Toutain G. 1981. Bayoud, Fusarium wilt of date. In: Nelson PE,
Toussoun TA, Cook RJ (eds), Fusarium: Diseases, biology, and taxonomy:
13– 20. The Pennsylvania State University Press, Pennsylvania State
University, USA.
Luongo L, Ferrarini A, Haegi A, et al. 2014. Genetic diversity and pathogenicity
of Fusarium oxysporum f. sp. melonis races from different areas in Italy.
Journal of Phytopathology 163: 73– 83.
Ma LJ, Geiser DM, Proctor RH, et al. 2013. Fusarium pathogenomics. Annual
Review of Microbiology 67: 399– 416.
Ma LJ, Shea T, Young S, et al. 2014. Genome sequence of Fusarium oxy-
sporum f. sp. melonis strain NRRL 26406, a fungus causing wilt disease
on melon. Genome Announcements 2: e00730-14.
Ma LJ, Van der Does HC, Brokovich KA, et al. 2010. Comparative genom-
ics reveals mobile pathogenicity chromosomes in Fusarium. Nature 464:
367–373.
Malençon G. 1934. Nouvelles observations concernant l’étiologie du Bayoud.
Comptes rendus hebdomadaires des séances de l’Acadédes sciences,
Paris 198: 1259–1261.
Manici LM, Caputo F, Saccà ML. 2017. Secondary metabolites released
into the rhizosphere by Fusarium oxysporum and Fusarium spp., as
underestimated component of nonspecific replant disease. Plant and Soil
415: 85– 98.
Manicom BQ, Baayen RP. 1993. Restriction fragment length polymorphism
in Fusarium oxysporum f. sp. dianthi and other fusaria from Dianthus spe-
cies. Plant Pathology 42: 851–857.
Manicom BQ, Bar-Joseph M, Kotze JM, et al. 1990. A restriction fragment
length polymorphism probe relating vegetative compatibility groups and
pathogenicity in Fusarium oxysporum f. sp. dianthi. Phytopathology 80:
336– 339.
Manulis S, Kogan N, Reuven M, et al. 1994. Use of the RAPD technique for
identification of Fusarium oxysporum f. sp. dianthi from carnation. Phyto-
pathology 84: 98–101.
Marasas WFO, Nelson PE, Toussoun TA. 1984. Toxigenic Fusarium species:
Identity and mycotoxicology. The Pennsylvania State University Press,
University Park, Pennsylvania, USA.
Marlatt ML, Correll JC, Kaufmann P, et al. 1996. Two genetically distinct
populations of Fusarium oxysporum f. sp. lycopersici race 3 in the United
States. Plant Disease 80: 1336–1342.
Martyn RD. 1987. Fusarium oxysporum f.sp. niveum race 2: a highly aggres-
sive race new to the United States. Plant Disease 71: 233– 236.
Martyn RD, Bruton BD. 1989. An initial survey of the United States for races
of Fusarium oxysporum f.sp. niveum. HortScience 24: 696– 698.
Maryani N, Lombard L, Poerba YS, et al. 2019. Phylogeny and genetic
diversity of the banana Fusarium wilt pathogen Fusarium oxysporum
f. sp. cubense in the Indonesian centre of origin. Studies in Mycology 92:
155–194.
Marziano F, Aloj B, Noviello C. 1987. Annali della Facolta di Scienze Agrarie
della Universita degli Studi di Napoli Portici 21: 13–19.
Mason-Gamer R, Kellogg E. 1996. Testing for phylogenetic conflict among
molecular datasets in the tribe Triticeae (Graminae). Systematic Biology
45: 524– 545.
Massey LM. 1926. Fusarium rot of gladiolus corms. Phytopathology 16: 509 –
523.
Matuo T, Ishigami K. 1958. On the wilt of Solanum melongena L. and its
causal fungus Fusarium oxysporum f. melongenae n. f. Annals of the
Phytopathological Society of Japan 23: 189–192.
Matuo T, Matsuda A, Ozaki K, et al. 1975. Fusarium oxysporum f. sp. arctii
n. f. causing wilt of Great Burdock. Annuals of the Phytopathological Society
of Japan 41: 77–80.
Matuo T, Miyagawa T, Saito H. 1986. Fusarium oxysporum f. sp. garlic
n. f. sp. causing basal rot of garlic. Annals of the Phytopathological Society
of Japan 52: 860– 864.
Matuo T, Motohashi S. 1967. On Fusarium oxysporum f. sp. lactucae n. f.
causing root rot of lettuce. Transactions of the Mycological Society of
Japan 8: 13–15.
Matuo T, Sato K. 1962. On two new forms of Fusarium lateritium. Transac-
tions of the Mycological Society of Japan 3: 120–126.
Matuo T, Tooyama A, Isaka M. 1979. Fusarium basal rot of Allium bakeri Regel
and its causal fungus, Fusarium oxysporum Schl. f. sp. allii n.f. Annals of
the Phytopathological Society of Japan 45: 305– 312.
Matuo T, Yamamoto I. 1967. On Fusarium oxysporum f. sp. lagenariae n. f.
causing wilt of Lagenaria vulgaris var. hispida. Transactions of the Myco-
logical Society of Japan 8: 61–63.
Mbofung GY, Hong SG, Pryor BM. 2007. Phylogeny of Fusarium oxysporum
f. sp. lactucae inferred from mitochondrial small subunit, elongation factor
1-α, and nuclear ribosomal intergenic spacer sequence data. Phytopathol-
ogy 97: 87–98.
Mbofung GCY, Pryor BM. 2010. A PCR-based assay for detection of Fusarium
oxysporum f. sp. lactucae in lettuce seed. Plant Disease 94: 860– 866.
McFadden HG, Wilson IW, Chapple RM, et al. 2006. Fusarium wilt (Fusarium
oxysporum f. sp. vasinfectum) genes expressed during infection of cotton
(Gossypium hirsutum). Molecular Plant Pathology 7: 87–101.
McNeill J, Turland NJ, Barrie FR, et al. (eds). 2012. International Code of
Nomenclature for algae, fungi, and plants (Melbourne Code). Gantner
Verlag KG [Regnum Vegetabile no. 154].
McRitchie JJ. 1973. Pathogenicity and control of Fusarium oxysporum wilt of
variegated Pyracantha. Plant Disease Reporter 57: 389– 391.
Mercier S, Louvet J. 1973. Recherches sur les fusarioses. X. Une fusariose
vasculaire (Fusarium oxysporum) du palmier des Canaries (Phoenix ca-
nariensis). Annales de Phytopathologie 5: 203– 211.
Mes JJ, Van Doorn J, Roebroeck EJA, et al. 1994. Restriction fragment
length polymorphisms, races and vegetative compatibility groups within a
worldwide collection of Fusarium oxysporum f. sp. gladioli. Plant Pathology
43: 362– 370.
Mes JJ, Weststeijn EA, Herlaar F, et al. 1998. Biological and molecular cha-
racterization of Fusarium oxysporum f. sp. lycopersici divides race 1 isolates
into separate virulence groups. Phytopathology 89: 156–160.
Mirocha CJ, Abbas HK, Kommedahl T, et al. 1989. Mycotoxin production by
Fusarium oxysporum and Fusarium sporotrichioides isolated from Baccha-
ris spp. from Brazil. Applied and Environmental Microbiology 55: 254– 255.
Mirtalebi M, Banihashemi Z. 2014. Genetic relationship among Fusarium
oxysporum f. sp. melonis vegetative compatibility groups and their related-
ness to other F. oxysporum formae speciales. Journal of Agricultural Science
and Technology 16: 931–943.
Mishra RK, Pandey BK, Muthukumar M, et al. 2013a. Detection of Fusarium
wilt pathogens of Psidium guajavae L. in soil using culture independent
PCR (ciPCR). Saudi Journal of Biological Sciences 20: 51–56.
Mishra RK, Pandey BK, Singh V, et al. 2013b. Molecular detection and geno-
typing of Fusarium oxysporum f. sp. psidii isolates from different agro-
ecological regions of India. Journal of Microbiology 51: 405– 412.
Mishra RK, Pandey BK, Singh V, et al. 2013c. Genetic characterization of
Fusarium oxysporum isolated from guava in northern India. African Journal
of Microbiology Research 7: 4228– 4234.
Mishra RK, Verma DK, Pandey BK, et al. 2014. Direct colony nested-PCR
for the detection of Fusarium oxysporum f. sp. psidii causing wilt disease
in Psidium guajavae L. Journal of Horticulture 1: 105. doi: https://doi.
org/10.4172/2376-0354.1000105.
Mohammadi N, Goltapeh EM, Babaie-Ahari A, et al. 2011. Pathogenic and
genetic characterization of Iranian isolates of Fusarium oxysporum f. sp.
lentis by ISSR analysis. International Journal of Agriculture Technology
7: 63–72.
Molnár A, Sulyok L, Hornok L. 1990. Parasexual recombination between
vegetative incompatible strains in Fusarium oxysporum. Mycological Re-
search 94: 393– 398.
Moricca S, Ragazzi A, Kasuga T, et al. 1998. Detection of Fusarium oxyspo-
rum f. sp. vasinfectum in cotton tissue by polymerase chain reaction. Plant
Pathology 47: 486– 494.
Mostert D, Molina AB, Daniells J, et al. 2017. The distribution and host range
of the banana Fusarium wilt fungus, Fusarium oxysporum f. sp. cubense,
in Asia. PLoS ONE 12 (7): e0181630.
Nagarajan G, Kang SW, Nam MH, et al. 2006. Characterization of Fusarium
oxysporum f. sp. fragariae based on vegetative compatibility group, random
amplified polymorphic DNA and pathogenicity. Plant Pathology Journal
22: 222– 229.
Nagarajan G, Nam MH, Song JY, et al. 2004. Genetic variation in Fusarium
oxysporum f. sp. fragariae populations based on RAPD and rDNA RFLP
analyses. Plant Pathology Journal 20: 264– 270.
Namiki F, Matsunaga M, Okuda M, et al. 2001. Mutation of an arginine bio-
synthesis gene causes reduced pathogenicity in Fusarium oxysporum f. sp.
melonis. Molecular Plant-Microbe Interactions 14: 580– 584.
Namiki F, Shiomi T, Kayamura T, et al. 1994. Characterization of the for-
mae speciales of Fusarium oxysporum causing wilts of cucurbits by DNA
fingerprinting with nuclear repetitive DNA sequences. Applied and Envi-
ronmental Microbiology 60: 2684– 2691.
44 Persoonia – Volume 43, 2019
Namiki F, Shiomi T, Nishi K, et al. 1998. Pathogenic and genetic variation in
the Japanese strains of Fusarium oxysporum f. sp. melonis. Phytopatho-
logy 88: 804– 810.
Nash SN, Snyder WC. 1962. Quantitative estimations by plate counts of pro-
pagules of the bean rot Fusarium in field soils. Phytopathology 73: 458 –462.
Nawade B, Talaviya JR, Vyas UM, et al. 2017. Diversity analysis among
Fusarium oxysporum f. sp. cumini isolates using ISSR markers, spore
morphology and pathogenicity. International Journal of Current Microbio-
logy and Applied Sciences 6: 79– 87.
Nelson PE, Toussoun TA, Marasas WFO. 1983. Fusarium species: An illus-
trated manual for identification. The Pennsylvania State University Press,
Pennsylvania, USA.
Netzer D. 1976. Physiological races and soil population level of Fusarium
wilt of watermelon. Phytoparasitica 4: 131–136.
Netzer D, Weintal C. 1987. Fusarium wilt of common heliotrope (Heliotropium
europaeum). Phytoparasitica 15: 139–140.
Ninet BI, Bontems JO, Lechenne O, et al. 2005. Molecular identification of
Fusarium species in onychomycoses. Dermatology 210: 21–25.
Nirenberg HI. 1976. Unterschungen über die morphologische und biolo-
gische Differenzierung in der Fusarium-Sektion Liseola. Mitteilungen der
Biologischen Bundesanstalt für Land- und Forstwirtschaft Berlin-Dahlem
169: 1–117.
Nirenberg HI, Ibrahim G, Michail SH. 1994. Race identity of three isolates
of Fusarium oxysporum Schlecht. f. sp. vasinfectum (Atk.) Snyd. & Hans,
from Egypt and the Sudan. Journal of Plant Diseases and Protection 101:
594– 597.
Nirmaladevi D, Venkataramana M, Srivastava RK, et al. 2016. Molecular
phylogeny, pathogenicity and toxigenicity of Fusarium oxysporum f. sp.
lycopersici. Scientific Reports 6: 21367.
Nishimura N, Kudo K. 1994. Fusarium oxysporum f. sp. colocasiae n. f. sp.
causing dry rot of taro (Colocasia esculenta). Annals of the Phytopathologi-
cal Society of Japan 60: 448– 453.
Nitschke E, Nihlgard M, Varrelmann M. 2009. Differentiation of eleven
Fusarium spp. isolated from sugar beet, using restriction fragment analysis
of a polymerase chain reaction – amplified translation elongation factor 1α
gene fragment. Phytopathology 99: 921–929.
Nourollahi K, Madahjalai M. 2017. Analysis of population genetic structure
of Iranian Fusarium oxysporum f. sp. lentis isolates using microsatellite
markers. Australasian Plant Pathology 46: 35– 42.
Noviello C, Snyder WC. 1962. Fusarium wilt of hemp. Phytopathology 52:
1315–1317.
Nylander JAA. 2004. MrModeltest v. 2. Programme distributed by the author.
Evolutionary Biology Centre, Uppsala University.
O’Donnell K, Cigelnik E. 1997. Two divergent intragenomic rDNA ITS2 types
within a monophyletic lineage of the fungus Fusarium are nonorthologous.
Molecular Phylogenetics and Evolution 7: 103–116.
O’Donnell K, Cigelnik E. 1999. A DNA sequence-based phylogenetic struc-
ture for the Fusarium oxysporum complex. Phytoparasitica 27: 69.
O’Donnel K, Gueidan C, Sink S, et al. 2009. A two-locus DNA sequence
database for typing plant and human pathogens within the Fusarium
oxysporum species complex. Fungal Genetics and Biology 46: 936– 948.
O’Donnell K, Kistler HC, Cigelnik E, et al. 1998. Multiple evolutionary origins
of the fungus causing Panama disease of banana: Concordant evidence
from nuclear and mitochondrial gene genealogies. Proceedings of the Na-
tional Academy of Sciences of the United States of America 95: 2044 –2049.
O’Donnell K, Rooney AP, Proctor RH, et al. 2013. Phylogenetic analyses of
RPB1 and RPB2 support a middle Cretaceous origin for a clade compris-
ing all agriculturally and medically important fusaria. Fungal Genetics and
Biology 52: 20– 31.
O’Donnell K, Sarver BAJ, Brandt M, et al. 2007. Phylogenetic diversity and
microsphere array-based genotyping of human pathogenic fusaria, includ-
ing isolates from the multistate contact lens-associated U.S. keratitis out-
break of 2005 and 2006. Journal of Clinical Microbiology 45: 2235– 2248.
O’Donnell K, Sutton DA, Fothergill A, et al. 2008. Molecular phylogenetic
diversity, multilocus haplotype nomenclature, and in vitro antifungal re-
sistance within the Fusarium solani species complex. Journal of Clinical
Microbiology 46: 2477–2490.
O’Donnell K, Sutton DA, Rinaldi MG, et al. 2004. Genetic diversity of human
pathogenic members of the Fusarium oxysporum complex inferred from
multilocus DNA sequence data and amplified fragment length polymorphism
analyses: Evidence for the recent dispersion of a geographically widespread
clonal lineage and nosocomial origin. Journal of Clinical Microbiology 42:
5109– 5120.
O’Donnell K, Sutton DA, Rinaldi MG, et al. 2010. Internet-accessible DNA
sequence database for identifying fusaria from human and animal infec-
tions. Journal of Clinical Microbiology 48: 3708– 3718.
O’Donnell K, Ward TJ, Robert VARG, et al. 2015. DNA sequence-based
identification of Fusarium: Current status and future directions. Phytopara-
sitica 43: 583– 595.
Ochoa JB, Yangari BF, Ellis MA, et al. 2004. Two new formae specialis of
Fusarium oxysporum, causing vascular wilt on babaco (Carica heilbornii
var. pentagona) and vascular wilt on naranjilla (Solanum quitoense) in
Ecuador. Fitologiya 39: 10–17.
Ogiso H, Fujinaga M, Saito H, et al. 2002. Physiological races and vegetative
compatibility groups of Fusarium oxysporum f. sp. lactucae isolated from
crisp head lettuce in Japan. Journal of General Plant Pathology 68: 292–
299.
Okubara PA, Harrison LA, Gatch EW, et al. 2013. Development and evalu-
ation of a TaqMan real-time PCR assay for Fusarium oxysporum f. sp.
spinaciae. Plant Disease 97: 927–937.
Okuda M, Ikeda K, Namiki F, et al. 1998. Tfo1: an Ac-like transposon from
the plant pathogenic fungus Fusarium oxysporum. Molecular Genetics and
Genomics 258: 599– 607.
Ortiz CS, Bell AA, Magill CW, et al. 2017. Specific PCR detection of Fusarium
oxysporum f. sp. vasinfectum California race 4 based on a unique Tfo1
insertion event in the PHO gene. Plant Disease 101: 34– 44.
Ortu G, Bertetti D, Gullino ML, et al. 2013. A new forma specialis of Fusarium
oxysporum on Crassula ovata. Journal of Plant Pathology 95: 33– 39.
Ortu G, Bertetti D, Gullino ML, et al. 2015a. Fusarium oxysporum f. sp. ech-
everiae, a novel forma specialis causing crown and stem rot of Echeveria
agavoides. Phytopathologia Mediterranea 54: 64–75.
Ortu G, Bertetti D, Martin P, et al. 2015b. Fusarium oxysporum f. sp. pa-
paveris: a new formae specialis isolated from Iceland poppy (Papaver
nudicaule). Phytopathologia Mediterranea 54: 76– 85.
Owen JH. 1956. Cucumber wilt, caused by Fusarium oxysporum f. cucumeri-
num n. f. Phytopathology 46: 153–157.
Padwick GW. 1940. The genus Fusarium. III. A critical study of the fungus
causing wilt of gram (Cicer arietinum L.) and of the related species in the
sub-section Orthocera, with special relation to variability of key characters.
Indian Journal of Agricultural Sciences 10: 241–284.
Palmero D, Rubio-Moraga A, Galvez-Patón L, et al. 2014. Pathogenicity and
genetic diversity of Fusarium oxysporum isolates from corms of Crocus
sativus. Industrial Crops and Products 61: 186–192.
Pande A, Rao VG. 1990. Wilt disease of Tabernaemontana coronaria Willd.
Biovigyanam 16: 58– 61.
Pandotra VR, Gupta JH, Sastry KSM. 1971. Note on wilt disease of Solanum
laciniatum. Indian Journal of Mycology and Plant Pathology 1: 86– 87.
Pappalardo L, Smith MK, Hamill SD, et al. 2009. DNA amplification finger-
printing analysis of genetic variation within Fusarium oxysporum f. sp.
zingiberi. Australasian Plant Pathology 38: 51–54.
Pasquali M, Acquadro A, Balmas V, et al. 2003. RAPD characterization of
Fusarium oxysporum isolates pathogenic on Argyranthemum frutescens
L. Journal of Phytopathology 151: 30– 35.
Pasquali M, Acquadro A, Balmas V, et al. 2004a. Development of PCR
primers for a new Fusarium oxysporum pathogenic on Paris daisy (Argy-
ranthemum frutescens L.). European Journal of Plant Pathology 110: 7–11.
Pasquali M, Dematheis F, Gilardi G, et al. 2005. Vegetative compatibility
groups of Fusarium oxysporum f. sp. lactucae from lettuce. Plant Disease
89: 237–240.
Pasquali M, Dematheis F, Gullino ML, et al. 2007. Identification of race 1 of
Fusarium oxysporum f. sp. lactucae on lettuce by inter-retrotransposon
sequence-characterized amplified region technique. Phytopathology 97:
987–996.
Pasquali M, Marena L, Fiora E, et al. 2004b. Real-time polymerase chain
reaction for identification of highly pathogenic group of Fusarium oxyspo-
rum f. sp. chrysanthemi on Argyranthemum frutescens L. Journal of Plant
Pathology 86: 53– 59.
Pasquali M, Marena L, Gullino ML, et al. 2004c. Vegetative compatibility
grouping of the Fusarium wilt pathogen of Paris daisy (Argyranthemum
frutescens L.). Journal of Phytopathology 152: 257–259.
Pasquali M, Piatti P, Gullino ML, et al. 2006. Development of a real-time poly-
merase chain reaction for the detection of Fusarium oxysporum f. sp. basilici
from basil seeds and roots. Journal of Phytopathology 154: 632– 636.
Pasquali M, Saravanakumar D, Gullino ML, et al. 2008. Sequence-specific
amplified polymorphism (SSAP) technique to analyse Fusarium oxysporum
f. sp. lactucae VCG 0300 isolate from lettuce. Journal of Plant Pathology
90: 527–535.
Pastrana AM, Kirkpatrick SC, Kong M, et al. 2017. Fusarium oxysporum
f. sp. mori, a new forma specialis causing Fusarium wilt of blackberry.
Plant Disease 100: 1018.
Patel PN, Prasad N, Mathur RL, et al. 1957. Fusarium wilt of cumin. Current
Science 26: 181–182.
45
L. Lombard et al.: Epitypification of Fusarium oxysporum
Pinaria AG, Laurence MH, Burgess LW, et al. 2015. Phylogeny and origin
of Fusarium oxysporum f. sp. vanillae in Indonesia. Plant Pathology 64:
1358–1365.
Pintore I, Gilardi G, Gullino ML, et al. 2017. Analysis of vegetative compati-
bility groups of Italian and Dutch isolates of Fusarium oxysporum f. sp.
lactucae. Journal of Plant Pathology 99: 517–521.
Ploetz RC. 2006. Fusarium wilt of banana is caused by several pathogens
re ferred to as Fusarium oxysporum f. sp. cubense. Phytopathology 96:
653– 656.
Ploetz RC. 2015. Fusarium wilt of banana. Phytopathology 105: 1512–1521.
Poli A, Bertetti D, Rapetti S, et al. 2013. Characterization and identification of
Colombian isolates of Fusarium oxysporum f. sp. dianthi. Journal of Plant
Pathology 95: 255– 263.
Poli A, Gilardi G, Spadaro D, et al. 2012. Molecular characterization of Fusa-
rium oxysporum f. sp. cichorii pathogenic on chicory (Cichorium intybus).
Phytoparasitica 40: 383– 391.
Pouralibaba HR, Pérez-de-Luque A, Rubiales D. 2017. Histopathology of the
infection on resistance and susceptible lentil accessions by two contrasting
pathotypes of Fusarium oxysporum f. sp. lentis. European Journal of Plant
Pathology 148: 53– 63.
Pouralibaba HR, Rubiales D, Fondevilla S. 2016. Identification of pathotypes
in Fusarium oxysporum f. sp. lentis. European Journal of Plant Pathology
144: 539– 549.
Prasad MSL, Sujatha M, Raoof MA. 2008. Morphological, pathogenic and
genetic variability in castor wilt isolates. Indian Phytopathology 61: 18– 27.
Prasad N, Mehta PR, Lal SB. 1952. Fusarium wilt of guava (Psidium guajava
L.) in Uttar Pradesh, India. Nature 169: 753–754.
Puhalla JE. 1984a. A visual indicator of heterokaryosis in Fusarium oxyspo-
rum from celery. Canadian Journal of Botany 62: 540–545.
Puhalla JE. 1984b. Races of Fusarium oxysporum f. sp. apii in California and
their genetic interrelationships. Canadian Journal of Botany 62: 546– 550.
Puhalla JE. 1985. Classification of strains of Fusarium oxysporum on the
basis of vegetative compatibility. Canadian Journal of Botany 63: 179–183.
Pyler TR, Simone GW, Fernandez D, et al. 2000. Genetic diversity among
isolates of Fusarium oxysporum f. sp. canariensis. Plant Pathology 49:
155–164.
Quaedvlieg W, Binder M, Groenewald JZ, et al. 2014. Introducing the Con-
solidated Species Concepts to resolve species in the Teratosphaeriaceae.
Persoonia 33: 1–40.
Raabe RD. 1960. Fusarium wilt of Sedum. Phytopathology 50: 651.
Raabe RD. 1985a. Fusarium wilt of Eustoma grandiflora. Phytopathology
75: 1306.
Raabe RD. 1985b. Fusarium wilt of Hebe species. Plant Disease 69: 450
451.
Rafique K, Rauf CA, Naz F, et al. 2015. DNA sequence analysis, morphology
and pathogenicity of Fusarium oxysporum f. sp. lentis isolates inciting lentil
wilt in Pakistan. International Journal of Biosciences 7: 74– 91.
Ramirez-Villupadua J, Endo RM, Bosland P, et al. 1985. A new race of Fusa-
rium oxysporum f. sp. conglutinans that attacks cabbage with type A resis-
tance. Plant Disease 69: 612– 613.
Rataj-Guranowska M, Wiatroszak I, Hornok L. 1984. Serological comparison
of two races of Fusarium oxysporum f. sp. lupini. Journal of Phytopatho-
logy 110: 221–225.
Rayner RW. 1970. A mycological colour chart. CMI and British Mycological
Society, Kew, Surrey, UK.
Reddy JM, Raoof MA, Ulaganathan K. 2012. Development of specific mark-
ers for identification of Indian isolates of Fusarium oxysporum f. sp. ricini.
European Journal of Plant Pathology 134: 713–719.
Reid J. 1958. Studies on the fusaria which cause wilt in melons. 1. The occur-
rence and distribution of races of the muskmelon and watermelon fusaria
and a histological study of the colonization of muskmelon plants susceptible
or resistant to Fusarium wilt. Canadian Journal of Botany 36: 393– 410.
Ren Y, Jiao D, Gong G, et al. 2015. Genetic analysis and chromosome map-
ping of resistance to Fusarium oxysporum f. sp. niveum (FON) race 1 and
race 2 in watermelon (Citrullus lanatus L.). Molecular Breeding 35: 183.
Rheeder JP, Marasas WFO, Vismer HF. 2002. Production of fumonisins
analogs by Fusarium species. Applied and Environmental Microbiology
68: 2101–2105.
Ribeiro RLD. 1977. Race differentiation in Fusarium oxysporum f. sp. pha-
seoli, the causal agent of bean yellows. Proceeding of the American Phyto-
pathology Society 4: 165.
Ribeiro RLD, Hagedorn DJ. 1979. Screening for resistance to and pathogenic
specialization of Fusarium oxysporum f. sp. phaseoli, the causal agent of
bean yellows. Phytopathology 69: 272– 276.
Richter H. 1941. Lupinenfusariosen. Mitteilungen aus der Biologischen
Reichanstalt fur Land-und Forstwirtschaft 64: 50– 61.
Risser G, Banihashemi Z, Davis DW. 1976. A proposed nomenclature of
Fusarium oxysporum f. sp. melonis races and resistance genes in Cucumis
melo. Phytopathology 66: 1105–1106.
Risser G, Mas P. 1965. Mise en évidencede plusieurs races de Fusarium
oxysporum f. melonis. Annales de l’amélioration des plantes 15: 405– 408.
Roebroeck EJA. 2000. Fusarium oxysporum from iridaceous crops: analysis
of genetic diversity and host specialisation. Faculty of Science, University
of Amsterdam, The Netherlands.
Roebroeck EJA, Mes JJ. 1992. Physiological races and vegetative compati-
bility groups within Fusarium oxysporum f. sp. gladioli. Netherlands Journal
of Plant Pathology 98: 57–64.
Romano C, Miracco C, Difonzo EM. 1998. Skin and nail infections due to
Fusarium oxysporum in Tuskany, Italy. Mycoses 41: 433–437.
Ronquist F, Huelsenbeck JP. 2003. MrBayes 3: Bayesian phylogenetic infer-
ence under mixed models. Bioinformatics 19: 1572–1574.
Rosewich UL, Pettway RE, Katan T, et al. 1999. Population genetic analysis
corroborates dispersal of Fusarium oxysporum f. sp. radicis-lycopersici
from Florida to Europe. Phytopathology 89: 623– 630.
Salgado MO, Schwartz HF. 1993. Physiological specialization and effects of
inoculum concentration of Fusarium oxysporum f. sp. phaseoli on common
beans. Plant Disease 77: 492– 496.
Sandoval-Denis M, Guarnaccia V, Polizzi G, et al. 2018. Symptomatic Citrus
trees reveal a new pathogenic lineage in Fusarium and two new Neocos-
mospora species. Persoonia 40: 1–25.
Sands DC, Ford EJ, Miller RV, et al. 1997. Characterization of a vascular
wilt of Erythroxylum coca caused by Fusarium oxysporum f. sp. erythroxyli
forma specialis nova. Plant Disease 81: 501–504.
Sauthoff W, Gerlach W. 1957. Die Fusarium – Welke der Aechmea fasciata
– eine neue pilzkrankheit. Gartenwelt 23: 389– 390.
Sauthoff W, Gerlach W. 1958. Über eine bisher nicht bekannte Fusarium
– welkekrankheit an Aechmea fasciata (Lindl). Bak. Nachrichtenblatt des
Deutschen Pflanzenschutzdienstes (Braunschweig) 10: 1–3.
Scarlett K, Tesoriero L, Daniel R, et al. 2013. Detection and quantification of
Fusarium oxysporum f. sp. cucumerinum in environmental samples using
specific quantitative PCR assay. European Journal of Plant Pathology 137:
315– 324.
Schmidt SM, Lukasiewicz J, Farrer R, et al. 2016. Comparative genomics of
Fusarium oxysporum f. sp. melonis reveals the secreted protein recognized
by the Fom-2 resistance gene in melon. New Phytologist 209: 307–318.
Schneider RW, Norelli JL. 1981. A new race of Fusarium oxysporum f. sp.
apii in California. Phytopathology 71: 108.
Schreuder JC. 1951. Een onderzoek over de Amerikaanse vaatziekte van
de erwten in Nederland. Tijdschrift over Plantenziekten 57 (6): 175– 206.
Sebastiani MS, Bagnaresi P, Sestili S, et al. 2017. Transcriptome analysis
of the melon- Fusarium oxysporum f. sp. melonis race 1.2 pathosystem in
susceptible and resistant plants. Frontiers in Plant Science 8: 362.
Sergent E, Beguet M. 1921. Sur la nature mycosique d’une nouvelle maladies
des dattiers menaçant les oasis marocaines. Comptes rendus hebdoma-
daires des séances de l’Acadédes sciences, Paris 172: 1624–1627.
Sharma KD, Winter P, Kahl G, et al. 2004. Molecular mapping of Fusarium
oxysporum f. sp. ciceris race 3 resistance gene in chickpea. Theoretical
and Applied Genetics 108: 1243–1248.
Sharma M, Nagavardhini A, Thudi M, et al. 2014. Development of DArT
markers and assessment of diversity in Fusarium oxysporum f. sp. ciceris,
wilt pathogen of chickpea (Cicer arietinum L.). BMC Genomics 15: 454.
Sharma M, Sengupta A, Ghosh R, et al. 2016. Genome wide transcriptome
profiling of Fusarium oxysporum f. sp. ciceris conidial germination reveals
new insights into infection-related genes. Science Reports 6: 37353.
Shende SS, Wankhade DJ, Rajurkar AB, et al. 2015. Characterization of
pathogenic species of Fusarium oxysporum isolates of safflower by RAPD
marker. Indian Journal of Plant Protection 43: 251–253.
Shimazu J, Yamauchi N, Hibi T, et al. 2005. Development of sequence tagged
site markers to identify races of Fusarium oxysporum f. sp. lactucae. Journal
of General Plant Pathology 71: 183–189.
Shiraishi A, Leslie JF, Zhong S, et al. 2012. ALFP, pathogenicity, and VCG
analyses of Fusarium oxysporum and Fusarium pseudocircinatum from
Acacia koa. Plant Disease 96: 1111–1117.
Skovgaard K, Nirenberg HI, O’Donnell K, et al. 2001. Evolution of Fusarium
oxysporum f. sp. vasinfectum races inferred from multigene genealogies.
Phytopathology 91: 1231–1237.
Smith EF. 1910. A Cuban banana disease. Science, New Series 31: 754–755.
Smith SN, DeVay JE, Hsieh WH, et al. 2001. Soil-borne populations of Fusa-
rium oxysporum f. sp. vasinfectum, a cotton wilt fungus in California fields.
Mycologia 93: 737–743.
Smith SN, Helms DM, Temple SR, et al. 1999. The distribution of Fusarium
wilt of blackeyed cowpeas within California caused by Fusarium oxysporum
f. sp. tracheiphilum race 4. Plant Disease 83: 694.
46 Persoonia – Volume 43, 2019
Snyder WC, Hansen HN. 1940. The species concept in Fusarium. American
Journal of Botany 27: 64– 67.
Snyder WC, Toole ER, Hepting GH. 1949. Fusaria associated with mimosa
wilt, and pine pitch canker. Journal of Agricultural Research 78: 365–382.
Snyder WC, Walker JC. 1935. Fusarium near wilt of pea. Zentralblatt für
Bakteriologie, Parasitenkunde und Infektionskrankheiten Abteilung II,
Jena 91: 355– 378.
Southwood MJ, Viljoen A, Mostert G, et al. 2012. Molecular identification of
two vegetative compatibility groups of Fusarium oxysporum f. sp. cepae.
Phytopathology 102: 204– 213.
Srinivasan K, Gilardi G, Spadaro D, et al. 2010. Molecular characterization
through IGS sequencing of formae speciales of Fusarium oxysporum
pathogenic on lamb’s lettuce. Phytopathologia Mediterrania 49: 309– 320.
Stamatakis A. 2014. RAxML version 8: a tool for phylogenetic analysis and
post-analysis of large phylogenies. Bioinformatics 30: 1312–1313.
Steinberg C, Laurent J, Edel-Hermann V, et al. 2015. Adaptation of Fusarium
oxysporum and Fusarium dimerum to the specific aquatic environment
provided by the water systems of hospitals. Water Research 76: 53– 65.
Stewart D. 1931. Sugar beet yellows caused by Fusarium conglutinane var.
betae. Phytopathology 19: 59–70.
Stover RH. 1962. Fusarial Wilt (Panama Disease) of bananas and other Musa
species. Commonwealth Mycological Institute, Kew, England.
Suelong H. 1981. A study on the pathogen of blight disease of tungoil trees.
Journal of Nanjing Technological College of Forest Products 9: 53.
Suga H, Hirayama Y, Morishima M, et al. 2013. Development of PCR primers
to identify Fusarium oxysporum f. sp. fragariae. Plant Disease 97: 619 –625.
Summerell BA, Leslie JF, Liew ECY, et al. 2010. Fusarium species associated
with plants in Australia. Fungal Diversity 46: 1–27.
Sun SK, Huang JW. 1983. A new Fusarium wilt of bitter gourd in Taiwan.
Plant Disease 67: 226– 227.
Sung GH, Sung JM, Hywel-Jones NL, et al. 2007. A multi-gene phylogeny of
Clavicipitaceae (Ascomycota, Fungi): Identification of localized incongru-
ence using a combinational bootstrap approach. Molecular phylogenetics
and evolution 44: 1204–1223.
Swanson TA, Van Gundy SD. 1985. Influences of temperature and plant age
on differentiation of races of Fusarium oxysporum f. sp. tracheiphilum on
cowpea. Plant Disease 69: 779–781.
Swett CS, Uchida JY. 2015. Characterization of Fusarium diseases on com-
mercially grown orchids in Hawaii. Plant Pathology 64: 648– 654.
Swift CE, Wickliffe ER, Schwartz HF. 2002. Vegetative compatibility groups
of Fusarium oxysporum f. sp. cepae from onion in Colorado. Plant Disease
86: 606– 610.
Swofford DL. 2003. PAUP*. Phylogenetic analysis using parsimony (*and other
methods), v. 4.0b10. Computer programme. Sinauer Associates, Sunder-
land, Massachusetts, USA.
Taheri N, Rastegar MF, Jafarpour B, et al. 2010. Pathogenic and genetic
characterization of Fusarium oxysporum f. sp. lentis by RAPD and IGS
analysis in Khorasan Province. World Applied Science Journal 9: 239– 244.
Takehara T, Kuniyasu K, Mori M, et al. 2003. Use of a nitrate-nonutilizing
mutant and selective media to examine population dynamics of Fusarium
oxysporum f. sp. spinaciae in soil. Phytopathology 93: 1173–1181.
Takken F, Rep M. 2010. The arms race between tomato and Fusarium oxy-
sporum. Molecular Plant Pathology 11: 309–314.
Talaviya JR, Jadeja KB, Nawade BD. 2014. Variability among different isolates
of Fusarium oxysporum f. sp. cumini. Trends in Biosciences 7: 3611–3616.
Tantaoui A, Boisson C. 1991. Compatibilité végétative d’isolats du Fusarium
oxysporum f. sp. albedinis et des Fusarium oxysporum de la rhizosphère
du palmier dattier et des sols de palmeraies. Phytopathologia Mediterriana
30: 155–163.
Tantaoui A, Fernandez D. 1993. Comparaison entre Fusarium oxysporum
f. sp. albedinis et Fusarium oxysporum des sols palmeraies par l’étude du
polymorphisme de longueur des fragments de restriction (RFLP). Phyto-
pathologia Mediterriana 32: 235– 244.
Tantaoui A, Ouinten M, Geiger J-P, et al. 1996. Characterization of a single
clonal lineage of Fusarium oxysporum f. sp. albedinis causing bayoud
disease of date palm in Morocco. Phytopathology 86: 787–792.
Taylor A, Vágány V, Jackson AC, et al. 2016. Identification of pathogenicity-
related genes in Fusarium oxysporum f. sp. cepae. Molecular Plant Patho-
logy 17: 1032–1047.
Thatcher LF, Gardiner DM, Kazan K, et al. 2012. A highly conserved effec-
tor in Fusarium oxysporum is required for full virulence on Arabidopsis.
Molecular Plant-Microbe Interactions 25: 180–190.
Thatcher LF, Williams AH, Grag G, et al. 2016. Transcriptome analysis of
the fungal pathogen Fusarium oxysporum f. sp. medicaginis during colo-
nisation of resistant and susceptible Medicago truncatula hosts identifies
differential pathogenicity profiles and novel candidate effectors. BMC
Genomics 17: 860.
Thurland NJ, Wiersma JH, Barbie FR, et al. (eds). 2018. International Code
of Nomenclature for algea, fungi, and plants (Shenzhen Code). Koeltz
Botanical Books. [Regnum Vegetabile 159].
Timmer LW. 1982. Host range and host colonization, temperature effects, and
dispersal of Fusarium oxysporum f. sp. citri. Phytopathology 72: 698–702.
Timmer LW, Garnsey SM, Grimm GR, et al. 1979. Wilt and dieback of Mexi-
can lime caused by Fusarium oxysporum. Phytopathology 69: 730–734.
Tok F, Kurt S. 2010. Pathogenicity, vegetative compatibility and amplified
fragment length polymorphism (AFLP) analysis of Fusarium oxysporum f.
radicis-cucumerinum isolates from Turkish greenhouses. Phytoparasitica
38: 253– 260.
Toole ER. 1941. Fusarium wilt of the mimosa tree (Albizzia julibrissin). Phyto-
pathology 31: 599– 616.
Toole ER. 1952. Two races of Fusarium oxysporum f. perniciosum causing
wilt of Albizzia spp. Phytopathology 42: 694.
Toth KF, Lacy ML. 1991. Comparing vegetative compatibility and protein
banding patterns for identification of Fusarium oxysporum f.sp. apii race
2. Canadian Journal Microbiology 37: 669– 674.
Triolo E, Lorenzini G. 1983. Fusarium oxysporum f. sp. fatshederae, a new
forma specialis causing wilt of × Fatshedera lizei. Annals of Applied Bio-
logy 102: 245– 250.
Troisi M, Bertetti D, Gullino ML, et al. 2013. Race differentiation in Fusarium
oxysporum f. sp. chrysanthemi. Journal of Phytopathology 161: 675– 688.
Troisi M, Gullino ML, Garibaldi A. 2010. Gerbera jamesonii, a new host of
Fusarium oxysporum f. sp. tracheiphilum. Journal of Phytopathology 158:
8–14.
Trujillo EE. 1963. Fusarium yellows and rhizome rot of common ginger.
Phytopathology 53: 1370–1371.
Tucker CM. 1927. Vanilla root rot. Journal of Agricultural Research 35:
1121–1136.
Upasani ML, Gurjar GS, Kadoo NY, et al. 2016. Dynamics of colonization and
expression of pathogenicity related genes in Fusarium oxysporum f. sp.
ciceri during chickpea vascular wilt disease progression. PLoS ONE 11:
e0156490. doi: https://doi.org/10.1371/journal.pone.0156490.
Vakalounakis DJ. 1996. Root and stem rot of cucumber caused by Fusarium
oxysporum f. sp. radicis-cucumerinum f. sp. nov. Plant Disease 80: 313
316.
Vakalounakis DJ, Doulis AG, Klironomou E. 2005. Characterization of
Fusarium oxysporum f. sp. radicis-cucumerinum attacking melon under
natural conditions in Greece. Plant Pathology 54: 339– 346.
Vakalounakis DJ, Fragkiadakis GA. 1999. Genetic diversity of Fusarium
oxysporum isolates from cucumber: differentiation by pathogenicity, vegeta-
tive compatibility, and RAPD fingerprinting. Phytopathology 89: 161–168.
Vakalounakis DJ, Wang Z, Fragkiadakis GA, et al. 2004. Characterization
of Fusarium oxysporum isolates obtained from cucumber in China by
pathogenicity, VCG, and RAPD. Plant Disease 88: 645–649.
Van Dam P, Fokkens L, Schmidt SM, et al. 2016. Effector profiles distinguish
formae speciales of Fusarium oxysporum. Environmental Microbiology
18: 4087–4102.
Van Dam P, Rep M. 2017. The distribution of Miniature Impala elements and
SIX genes in the Fusarium genus is suggestive of horizontal gene transfer.
Journal of Molecular Evolution 85: 14– 25.
Van der Does HC, Lievens B, Claes L, et al. 2008. The presence of the viru-
lence locus discriminates Fusarium oxysporum isolates causing tomato wilt
from other isolates. Environmental Microbiology 10: 1475–1485.
Van Diepeningen AD, Feng P, Ahmed S, et al. 2015. Spectrum of Fusarium
infections in tropical dermatology evidenced by multilocus sequencing
typing diagnostics. Mycoses 58: 48– 57.
Van Hall CJJ. 1903. Die Sankt Johaniskrankheit der Erbsen verursacht von
Fusarium vasinfectum Atk. Berichte der Deutschen Botanischen Gesell-
schaft 21: 2– 6.
Vasudeva RS, Srinivasan KV. 1952. Studies on the wilt diseases of lentil.
Indian Phytopathology 5: 23– 32.
Venter SL, Theron DJ, Steyn PJ, et al. 1992. Relationship between vegetative
compatibility and pathogenicity of isolates of Fusarium oxysporum f. sp.
tuberosi from potato. Phytopathology 82: 858– 862.
Von Arx JA. 1952. De voetziekte van Gerbera, veroorzaakt door Fusarium
oxysporum Schlecht. Tijdschrift over Plantziekten 58 (1): 5– 9.
Von Schlechtendahl FK. 1824. Flora berolinensis. Pars secunda. Crypto-
gamia, Berlin.
Wager VA. 1947. Wilt disease of New Zealand flax. Farming in South Africa
22: 871–878.
Waite BH. 1963. Wilt of Heliconia spp. caused by Fusarium oxysporum f.
cubense race 3. Tropical Agriculture 40: 299–305.
Wang B, Brubaker CL, Summerell BA, et al. 2010. Local origin of two vege-
tative compatibility groups of Fusarium oxysporum f. sp. vasinfectum in
Australia. Evolutionary Applications 3 (5– 6): 505– 524. doi: https://doi.
org/10.1111/j.1752-4571.2010.00139.x.
47
L. Lombard et al.: Epitypification of Fusarium oxysporum
Wang B, Brubaker CL, Tate W, et al. 2006. Genetic variation and popula-
tion structure of Fusarium oxysporum f. sp. vasinfectum in Australia. Plant
Pathology 55: 746–755.
Wang PH, Lo HS, Yeh Y. 2001. Identification of F. o. cucumerinum and F. o.
luffae by RAPD-generated DNA probes. Letters in Applied Microbiology
33: 397–401.
Webb KM, Case AJ, Brick MA, et al. 2013. Cross pathogenicity and vegeta-
tive compatibility of Fusarium oxysporum isolated from sugar beet. Plant
Disease 97: 1200–1206.
Weimer JL. 1928. A wilt disease of alfalfa caused by Fusarium oxysporum
var. medicaginis, n. var. Journal of Agricultural Research 37: 419– 433.
Weimer JL. 1944. Some root rots and a foot rot of lupines in the southeastern
part of the United States. Journal of Agricultural Research 68: 441–457.
Wellman FL. 1972. Tropical American plant disease (Neotropical phytopatho-
logy problems). Scarecrow Press, Metuchen, N.J., USA.
Wellman M. 1954. Fusarium oxysporum Shl. f. sp. coffea. Phytopathology
44: 509.
Whitehead DS, Coddington A, Lewis BG. 1992. Classification of races by
DNA polymorphism analysis and vegetative compatibility grouping in
Fusarium oxysporum f. sp. pisi. Physiological and Molecular Plant Patho-
logy 41: 295– 305.
Widodo, Kondo N, Kobayashi K, et al. 2008. Vegetative compatibility groups
within Fusarium oxysporum f. sp. cepae in Hokkaido-Japan. Microbiology
Indonesia 2: 39– 43.
Williams AH, Sharma M, Thatcher LF, et al. 2016. Comparative genomics
and prediction of conditionally dispensable sequences in legume-infecting
Fusarium oxysporum formae speciales facilitates identification of candidate
effectors. BMC Genomics 17: 191.
Winks BL, Williams YN. 1965. A wilt of strawberry caused by a new form
of Fusarium oxysporum. Queensland Journal of Agriculture and Animal
Science 22: 475– 479.
Wollenweber HW. 1913. Studies on the Fusarium problem. Phytopathology
3: 24– 50.
Wollenweber HW. 1914. Identification of species of Fusarium occurring on
the sweet potato, Ipomoea batatas. Journal of Agricultural Research 2:
251–285.
Wollenweber HW. 1931. Fusarium-monographie. Fungi parasitici et sapro-
phytici. Zeitschrift für Parasitenkunde 3: 260– 516.
Wollenweber HW, Reinking OA. 1935. Die Fusarien, ihre Beschreibung,
Schadwirkung und Bekampfung. Paul Parey, Berlin.
Woo SL, Zoina A, Del Sorbo G, et al. 1996. Characterization of Fusarium oxy-
sporum f. sp. phaseoli by pathogenic races, VCGs, RFLPs, and RAPD. Phyto-
pathology 86: 966– 973.
Woudenberg JHC, Seidl MF, Groenewald JZ, et al. 2015. Alternaria section
Alternata: Species, formae speciales or pathotypes? Studies in Mycology
82: 1–21.
Woudt LP, Neuvel A, Sikkema A, et al. 1995. Genetic variation in Fusarium
oxysporum from cyclamen. Phytopathology 85: 1348–1355.
Wunsch MJ, Baker AH, Kalb DW, et al. 2009. Characterization of Fusarium
oxysporum f. sp. loti forma specialis nov., a monophyletic pathogen causing
vascular wilt of birdsfoot trefoil. Plant Disease 93: 58– 66.
Yamauchi N, Horiuchi S, Satou M. 2001. Pathogenicity groups in Fusarium
oxysporum f. sp. lactucae on horticultural types of lettuce cultivars. Journal
of General Plant Pathology 67: 288– 290.
Yamauchi N, Shimazu J, Satou M, et al. 2004. Physiological races and
vege tative compatibility groups of butterhead lettuce isolates of Fusarium
oxysporum f. sp. lactucae in Japan. Journal of General Plant Pathology
70: 308– 313.
Yoo SJ, Watanabe H, Kobayashi K, et al. 1993. Vegetative compatibility
grouping of formae speciales of Fusarium oxysporum pathogenic to the
Liliaceae. Annals of the Phytopathological Society of Japan 59: 3– 9.
Yu TF, Fang CT. 1948. Fusarium diseases of broad beans. III. Root-rot and
wilt of broad beans caused by two new forms of Fusarium. Phytopatho-
logy 38: 587–594.
Yun SH, Arie T, Kaneko I, et al. 2000. Molecular organization of mating type
loci in heterothallic, homothalic, and asexual Gibberella/ Fusarium species.
Fungal Genetics and Biology 31: 7–20.
Zambounis AG, Paplomatas E, Tsaftaris AS. 2007. Intergenic spacer –RFLP
analysis and direct quantification of Australian Fusarium oxysporum f. sp.
vasinfectum isolates from soil and infected cotton tissues. Plant Disease
91: 1564–1573.
Zang W, Zhao W, Huang J, et al. 2014. PEG-mediated genetic transforma-
tion of Fusarium oxysporum f. sp. conglutinans to study pathogenesis in
cabbage. Chiang Mai Journal Science 41: 945– 956.
Zanotti MGS, De Queiroz MV, Dos Santos JK, et al. 2006. Analysis of the
genetic diversity of Fusarium oxysporum f. sp. phaseoli isolates, pathogenic
and non-pathogenic to common bean (Phaseolus vulgaris L.). Journal of
Phytopathology 154: 545– 549.
Zhang Z, Zhang J, Wang Y, et al. 2005. Molecular detection of Fusarium oxy-
sporum f. sp. niveum and Mycosphaerella melonis in infected plant tissues
and soil. FEMS Microbiology Letters 249: 39– 47.
Zhou XG, Everts KL, Bruton BD. 2010. Race 3, a new and highly virulent race
of Fusarium oxysporum f. sp. niveum causing Fusarium wilt in watermelon.
Plant Disease 94: 92– 98.
Zimmermann J, De Klerk M, Musyoki MK, et al. 2015. An explicit AFLP-based
marker for monitoring Fusarium oxysporum f. sp. strigae in tropical soils.
Biological Control 89: 42– 52.
Zimmermann J, Musyoki MK, Cadisch G, et al. 2016. Proliferation of the bio-
control agent Fusarium oxysporum f. sp. strigae and its impact on indige-
nous rhizosphere fungal communities maize under different agro-ecologies.
Rhizosphere 1: 17–25.
... T. cacao causing pod rot in Puerto Rico (Serrato-Diaz et al., 2022 (Leslie & Summerell, 2006;Lombard et al., 2019;Wang et al., 2019). Nevertheless, the morphological approach should not be ruled out as it is the primary method for preliminary species identification. ...
... A polyphasic approach that includes morphological and molecular methods is important for the taxonomic classification of Fusarium (Leslie & Summerell, 2006;Lombard et al., 2019;O'Donnell et al., 2022). In this study, TEF1 and TUB2 were used to support the morphological identification of Fusarium as suggested by many studies (Geiser et al., 2001(Geiser et al., , 2004 (Lateef et al., 2018;Maharachchikumbura et al., 2014). ...
Article
Malaysia stands prominently among Asia’s key cocoa-producing countries. In the cocoa season of 2022/2023, Malaysia demonstrated its contribution to the industry, with an estimated production of around 364,000 tonnes of cacao bean grindings. Nonetheless, fungal diseases pose undeniable challenges to the cocoa sector. Extensive sampling conducted between September 2018 and March 2019 across multiple states in Malaysia revealed concerning symptoms of leaf blight and stem canker affecting Theobroma cacao plants. The aim of this study was to identify and characterize fungal species associated with leaf blight and stem canker of T. cacao in Malaysia through morphological, molecular and pathogenicity analyses. Morphological and molecular phylogenetic analyses using multiple DNA regions (rDNA internal transcribed spacer [ITS], TEF1 and TUB2) were performed and identified 40 fungal isolates found in this study as Diaporthe tulliensis (17 isolates), Fusarium solani (seven isolates), Fusarium proliferatum (six isolates) and Neopestalotiopsis clavispora (10 isolates). Pathogenicity tests with mycelial plugs and wound treatments showed that D. tulliensis and N. clavispora were responsible for causing leaf blight whereas D. tulliensis, F. solani and F. proliferatum caused stem canker of T. cacao. The present study provides insights into disease aetiology and symptomatology that may be useful in planning effective disease management for the host plant.
... However, in the abovementioned studies, F. avenaceum, F. ramigenum, F. culmorum, F. keratoplasticum, F. oxysporum, and F. proliferatum were not found in grapevine nurseries in Türkiye. Interestingly, although F. oxysporum is a large species complex, including 21 species [26], and F. oxysporum has an essential place in this complex, we could not detect F. oxysporum among the Fusarium species we isolated from vines. In contrast to our findings, Zeidan et al. [27] reported that F. oxysporum strains obtained from different grape varieties (cv. ...
Article
Full-text available
Fusarium species are agriculturally important fungi with a broad host range and can be found as endophytic, pathogenic, or opportunistic parasites in many crop plants. This study aimed to identify Fusarium species in bare-rooted, dormant plants in Turkish grapevine nurseries using molecular identification methods and assess their pathogenicity. Asymptomatic dormant plants were sampled from grapevine nurseries (43) in different regions of the country, and fungi were isolated from plant roots and internal basal tissues. The Fusarium strains were identified by performing gene sequencing (TEF1-α, RPB2) and phylogenetic analyses. Pathogenicity tests were carried out by inoculating mycelial agar pieces of strains onto the stem or conidial suspensions into the rhizosphere of vines (1103 Paulsen rootstock). Laboratory tests revealed that Fusarium species were highly prevalent in Turkish grapevine nurseries (41 out of 43)
... F. oxysporum was first introduced by Diederich F.L. von Schlechtendal in 1824. It harbors both pathogenic and non-pathogenic strains, suffers from various classification problems, and was established as an epitype of F. oxysporum by Lombard et al. in 2019 [37]. These are probably also the reasons why there are far-reaching fields of research on this species, but very little on the biodegradation of alkanes. ...
Article
Full-text available
Studying the fates of oil components and their interactions with ecological systems is essential for developing comprehensive management strategies and enhancing restoration following oil spill incidents. The potential expansion of Kazakhstan’s role in the global oil market necessitates the existence of land-specific studies that contribute to the field of bioremediation. In this study, a set of experiments was designed to assess the growth and biodegradation capacities of eight fungal strains sourced from Kazakhstan soil when exposed to the hydrocarbon substrates from which they were initially isolated. The strains were identified as Aspergillus sp. SBUG-M1743, Penicillium javanicum SBUG-M1744, SBUG-M1770, Trichoderma harzianum SBUG-M1750 and Fusarium oxysporum SBUG-1746, SBUG-M1748, SBUG-M1768 and SBUG-M1769 using the internal transcribed spacer (ITS) region. Furthermore, microscopic and macroscopic evaluations agreed with the sequence-based identification. Aspergillus sp. SBUG-M1743 and P. javanicum SBUG-M1744 displayed remarkable biodegradation capabilities in the presence of tetradecane with up to a 9-fold biomass increase in the static cultures. T. harzianum SBUG-M1750 exhibited poor growth, which was a consequence of its low efficiency of tetradecane degradation. Monocarboxylic acids were the main degradation products by SBUG-M1743, SBUG-M1744, SBUG-M1750, and SBUG-M1770 indicating the monoterminal degradation pathway through β-oxidation, while the additional detection of dicarboxylic acid in SBUG-M1768 and SBUG-M1769 cultures was indicative of the fungus’ ability to undertake both monoterminal and diterminal degradation pathways. F. oxysporum SBUG-M1746 and SBUG-M1748 in the presence of cyclohexanone showed a doubling of the biomass with the ability to degrade the substrate almost completely in shake cultures. F. oxysporum SBUG-M1746 was also able to degrade cyclohexane completely and excreted all possible metabolites of the degradation pathway. Understanding the degradation potential of these fungal isolates to different hydrocarbon substrates will help in developing effective bioremediation strategies tailored to local conditions.
... Strain ByF01 was isolated from diseased K. roxburghii roots collected from Xiangyun County, Dali Bai Autonomous prefecture of Yunnan Province, China (25°25′N, 100°40′E), in 2021. This strain was identified as F. oxysporum based on morphological characteristics, PCR amplification [11], a phylogenetic analysis based on the nucleotide sequences of cmdA, rpb2, tef1, and tub2, and a pathogenicity test with 1-year-old healthy seedlings of K. roxburghii, which fulfilled Koch′s postulates [4]. Strain ByF01 was deposited in the Institute of Medicinal Plant Cultivation, Academy of Southern Medicine, Yunnan University of Chinese Medicine, Yunnan, China. ...
Article
Full-text available
Objectives Knoxia roxburghii is a member of the madder (Rubiaceae) family. This plant is cultivated in different areas of China and recognized for its medicinal properties, which leads to its use in traditional Chinese medicine. The incidence of root rot was 10–15%. In June 2023, the causal agent of root rot on K. roxburghii was identified as Fusarium oxysporum. To the best of our knowledge, this is the first report of the complete genome of F. oxysporum strain ByF01 that is the causal agent of root rot of K. roxburghii in China. The results will provide effective resources for pathogenesis on K. roxburghii and the prevention and control of root rot on this host in the future. Data description To understand the molecular mechanisms used by F. oxysporum to cause root rot on K. roxburghii , strain ByF01 was isolated from diseased roots and identified by morphological and molecular methods. The complete genome of strain ByF01 was then sequenced using a combination of the PacBio Sequel IIe and Illumina sequencing platforms. We obtained 54,431,725 bp of nucleotides, 47.46% GC content, and 16,705 coding sequences.
... However, in the abovementioned studies, F. avenaceum, F. ramigenum, F. culmorum, F. keratoplasticum, F. oxysporum, and F. proliferatum were not found in grapevine nurseries in Türkiye. Interestingly, although F. oxysporum is a large species complex, including 21 species [26], and F. oxysporum has an essential place in this complex, we could not detect F. oxysporum among the Fusarium species we isolated from vines. When phylogenetic analyses were performed, it was revealed that many isolates similar to this species were F. curvatum, F. glycines, and F. nirenbergiae. ...
Preprint
Full-text available
Fusarium species are agriculturally important fungi with a broad host range and can be found as endophytic, pathogenic, or opportunistic parasites in many crop plants. This study aimed to identify Fusarium species in bare-rooted, dormant plants in Turkish grapevine nurseries using molecular identification methods and assess their pathogenicity. Asymptomatic dormant plants were sampled from grapevine nurseries (43) in different regions of the country, and fungi were isolated from plant roots and internal basal tissues. The Fusarium isolates were identified by performing gene sequencing (TEF1-α, RPB2) and phylogenetic analyses. Pathogenicity tests were carried out by inoculating mycelial agar pieces of isolates onto the stem or conidial suspensions into the rhizosphere of vines (1103 Paulsen rootstock). Laboratory tests revealed that Fusarium species were highly prevalent in Turkish grapevine nurseries (41 out of 43). Gene sequencing and phylogenetic analyses unraveled that 12 Fusarium species (F. annulatum, F. brachygibbosum, F. clavum, F. curvatum, F. falciforme, F. fredkrugeri, F. glycines, F. nanum, F. nematophilum, F. nirenbergiae, F. solani, and Fusarium spp.) existed in the ready-to-sale plants. Some of these species (F. annulatum, F. curvatum and F. nirenbergiae) consistently caused wood necrosis of seedling stems, rotting of the basal zone and roots, and reduced root biomass. Although the other nine species also caused some root rot and root reduction, their virulence was not as severe as the pathogenic ones, and they were considered opportunistic parasites or endophytic species. This study is the first detailed investigation of Fusarium species in grapevine nurseries in Türkiye and emphasizes the need to consider pathogenic species in producing healthy grapevine seedlings.
... Fusarium sequences (tef1-α) were analyzed using a procedure similar to that described for Trichoderma. In this case, the best model for the ML analysis of the F. oxysporum species complex (FOSC; Lombard et al., 2019) was the Kimura 2-parameter model with a gamma distribution (+G). ...
Article
Full-text available
Banana (Musa acuminata) is the most important crop in the Canary Islands (38.9% of the total cultivated area). The main pathogen affecting this crop is the soil fungal Fusarium oxysporum f. sp. cubense subtropical race 4 (Foc-STR4), for which there is no effective control method under field conditions. Therefore, the use of native biological control agents may be an effective and sustainable alternative. This study aims to: (i) investigate the diversity and distribution of Trichoderma species in the rhizosphere of different banana agroecosystems affected by Foc-STR4 in Tenerife (the island with the greatest bioclimatic diversity and cultivated area), (ii) develop and preserve a culture collection of native Trichoderma species, and (iii) evaluate the influence of soil chemical properties on the Trichoderma community. A total of 131 Trichoderma isolates were obtained from 84 soil samples collected from 14 farms located in different agroecosystems on the northern (cooler and wetter) and southern (warmer and drier) slopes of Tenerife. Ten Trichoderma species, including T. afroharzianum, T. asperellum, T. atrobrunneum, T. gamsii, T. guizhouense, T. hamatum, T. harzianum, T. hirsutum, T. longibrachiatum, and T. virens, and two putative novel species, named T. aff. harzianum and T. aff. hortense, were identified based on the tef1-α sequences. Trichoderma virens (35.89% relative abundance) and T. aff. harzianum (27.48%) were the most abundant and dominant species on both slopes, while other species were observed only on one slope (north or south). Biodiversity indices (Margalef, Shannon, Simpson, and Pielou) showed that species diversity and evenness were highest in the healthy soils of the northern slope. The Spearman analysis showed significant correlations between Trichoderma species and soil chemistry parameters (mainly with phosphorus and soil pH). To the best of our knowledge, six species are reported for the first time in the Canary Islands (T. afroharzianum, T. asperellum, T. atrobrunneum, T. guizhouense, T. hamatum, T. hirsutum) and in the rhizosphere of banana soils (T. afroharzianum, T. atrobrunneum, T. gamsii, T. guizhouense, T. hirsutum, T. virens). This study provides essential information on the diversity/distribution of native Trichoderma species for the benefit of future applications in the control of Foc-STR4.
... Multilocus analysis of the combined tef and rpb2 sequences was used to infer the taxonomic status of these strains. The topologies of the constructed trees for the FSAMSC and FSSC by different methods are concordant with the phylogenetic relationships among Fusarium species reconstructed previously [35][36][37][38][39][40][41][42]. The phylogenetic analysis allowed for accurately identifying the Fusarium species: among the analyzed FSAMSC strains, most of them belong to F. sambucinum and one to F. venenatum, and among the FSSC strains, four species (F. ...
Article
Full-text available
Dry rot of potato tubers is a harmful disease caused by species of the Fusarium genus. Studies on the composition and features of Fusarium spp. that cause the disease in Russia are limited. Thirty-one Fusarium strains belonging to the F. sambucinum species complex (FSAMSC) and F. solani species complex (FSSC) were accurately identified using multilocus phylogenetic analysis of the tef and rpb2 loci, and their physiological characteristics were studied in detail. As a result, 21 strains of F. sambucinum s. str. and 1 strain of F. venenatum within the FSAMSC were identified. Among the analyzed strains within the FSSC, one strain of F. mori, four strains of F. noneumartii, and two strains of both F. stercicola and F. vanettenii species were identified. This is the first record of F. mori on potato as a novel host plant, and the first detection of F. noneumartii and F. stercicola species in Russia. The clear optimal temperature for the growth of the strains belonging to FSAMSC was noted to be 25 °C, with a growth rate of 11.6–15.0 mm/day, whereas, for the strains belonging to FSSC, the optimal temperature range was between 25 and 30 °C, with a growth rate of 5.5–14.1 mm/day. The distinctive ability of F. sambucinum strains to grow at 5 °C has been demonstrated. All analyzed Fusarium strains were pathogenic to potato cv. Gala and caused extensive damage of the tuber tissue at an incubation temperature of 23 °C for one month. Among the fungi belonging to the FSAMSC, the F. sambucinum strains were more aggressive and caused 23.9 ± 2.2 mm of necrosis in the tubers on average compared to the F. venenatum strain—17.7 ± 1.2 mm. Among the fungi belonging to the FSSC, the F. noneumartii strains were the most aggressive and caused 32.2 ± 0.8 mm of necrosis on average. The aggressiveness of the F. mori, F. stercicola, and especially the F. vanettenii strains was significantly lower: the average sizes of damage were 17.5 ± 0.5 mm, 17.2 ± 0.2 mm, and 12.5 ± 1.7 mm, respectively. At an incubation temperature of 5 °C, only the F. sambucinum strains caused tuber necroses in the range of 6.7 ± 0.5–15.9 ± 0.8 mm.
... Ихтисослашган формаларнинг сони жуда ҳам кўп бўлиши мумкин. Масалан, 2018 йил август ойигача F. oxysporum турининг 106 та яхши ўрганилган ва 37 та етарли ўрганилмаган формалари баён қилинган ҳамда ушбу патоген билан зарарланадиган яна 58 та ўсимлик турлари / туркумлари аниқланган, аммо улар учун ҳали форма номлари барпо этилмаган; бир қатор ихтисослашган формалар таркибида физиологик ирқлари дифференциация қилинган (Edel-Hermann, Lecomte, 2019; Lombard et al., 2019). ...
Book
Ҳасанов Б.А., Шеримбетов А.Г., Гулмуродов Р.А., Хайтбаева Н.С., Каримов О.К. Fusarium туркуми ва буғдойнинг фузариозлари. Ўқув қўлланма. Тошкент, 2024, 255 бет, 14 жадвал 35 та расм ва фотосуратлар (ўзбек тилида). Калит сўзлар: Fusarium, туркум, таксономия, номенклатура, морфологик идентификация, молекуляр-генетик идентификация, ген, биохилмахиллик, фитопатоген турлар, инсонлар патогенлари, эндофитлар, буғдой патогенлари, микотоксин. Ушбу ўқув қўлланма 8 бобдан иборат бўлиб, 1-бобда Fusarium туркумининг таксономияси, туркумдаги турлар концепциялари ва турларни идентификация қилишда қўлланиладиган маркер генлар, молекуляр-генетик усуллар ва воситалар баён этилган. 2-бобда Fusarium туркумидаги биохилмахиллик, 3-бобда туркум номенклатураси, 4-бобда турларнинг морфологик белгилари, 5-бобда турларни идентификация қилиш усуллари, 6-бобда Fusarium турларининг яшаш тарзлари, субстратлари, хўжайин ўсимликлари ҳамда ривожланиш цикллари, жумладан туркумнинг ўсимлик, инсон ва ҳайвонларда касаллик қўзғатувчи ҳамда эндофит турлари ҳақида батафсил маълумотлар келтирилган. Кейинги, 7-бобда Fusarium туркумининг дунёда ва Ўзбекистонда буғдой ўсимликларида касаллик қўзғатувчи турлари бўйича аннотацияланган рўйхат ва маълумотларнинг танқидий таҳлили тақдим этилган. Китобнинг охирги, 8-бобида Fusarium туркуми турларининг экинларнинг донлари ва бошқа субстратларда синтез қиладиган микотоксинлари ҳамда маҳсулотларда уларни камайтириш усуллари ҳақида тўла маълумотларнинг таҳлили берилган. Ўқув қўлланмани тайёрлашда 304 та, жумладан 48 та Ўзбекистонда ва бошқа МДҲ мамлакатларида ҳамда 256 та хорижда чоп этилган адабиёт манбаларидаги маълумотлардан фойдаланилган. Китоб охирида глоссарий – биологик, микологик, фитопатологик, генетик, молекуляр-генетик ва филогенетик терминларнинг изоҳли луғати ҳамда микроорганизмлар илмий (лотинча) номларининг индекси келтирилган. Ушбу қўлланма бўйича тақриз, фикр ва мулоҳазаларингизни қуйидаги манзилга жўнатишингизни таклиф қиламиз: khasanov.batyr@gmail.com. Хасанов Б.А., Шеримбетов А.Г., Гулмуродов Р.А., Хайтбаева Н.С., Каримов О.К. Род Fusarium и фузариозы пшеницы. Учебное пособие. Ташкент, 2024, 255 стр., 14 таблиц, 35 рис. и фотографий (на узбекском языке). Ключевые слова: Fusarium, род, таксономия, номенклатура, морфологическая идентификация, молекулярно-генетическая идентификация, ген, биоразнообразие, фитопатогенные виды, патогены людей, эндофиты, патогены пшеницы, микотоксин. Учебное пособие состоит из восьми глав. В 1-главе изложены сведения о таксономии рода Fusarium, концепциях вида, молекулярно-генетических методах, средствах и маркерных генах, используемых при идентификации видов этого рода. Во второй главе приведена подробная информация о биоразнообразии в роде Fusarium, в главе 3 о номенклатуре, главе 4 о морфологических признаках видов, в главе 5 о методах идентификации и в главе 6 – об образе жизни, субстратах, растениях-хозяевах и циклах развития, в том числе о видах рода, вызывающих болезни людей, животных и эндофитных видах. В главе 7 приведены аннотированный список и критический анализ имеющейся информации о видах рода Fusarium, поражающих пшеницу в мире и в Узбекистане. В последней, 8-главе книги изложены полные сведения о микотоксинах, синтезируемых видами рода Fusarium в семенах культурных растений и в других субстратах, а также методах снижения их концентрации в с.-х. продуктах. При подготовке учебного пособия использованы сведения 304-х источников литературы, в том числе 48 работ, опубликованных в Узбекистане и других странах СНГ, и 256 работ зарубежных авторов. В конце книги приведены глоссарий – толковый словарь биологических, микологических, фитопатологических, молекулярно-генетических и филогенетических терминов и индекс (алфавитный указатель) научных (латинских) названий микроорганизмов. Просим отправить ваши критические замечания, отзывы и предложения по адресу: khasanov.batyr@gmail.com. Batyr A. Khasanov, Anvar G. Sherimbetov, Risqiboy A. Gulmurodov, Nodira S. Khaytbaeva, and Otabek K. Karimov. Genus Fusarium and Fusarium diseases of wheat worldwide and in Uzbekistan. Textbook for students. Tashkent, 2024, 255 pp., with 14 Tables, and 35 Figs. (in Uzbek language). Key words: Fusarium, genus, taxonomy, nomenclature, morphologic identification, molecular-genetic identification, gene, biodiversity, plant pathogens, human pathogens, endophytes, wheat pathogens, mycotoxin. The textbook consists of 8 chapters. Chapter 1 contains information on the taxonomy of the genus Fusarium, species concepts, molecular genetic methods, tools and marker genes used in the identification of its species. Chapter 2 provides detailed information on biodiversity in the genus Fusarium, chapter 3 – on nomenclature, chapter 4 – on morphological features of species, chapter 5 – on identification methods, and chapter 6 – on lifestyles, substrates, host plants and development cycles, including species of the genus that cause human and animal diseases, and endophytes. Chapter 7 provides an annotated list and critical analysis of available in the literature information on Fusarium species that infect wheat plants in the world and in Uzbekistan. The last, 8th chapter of the book contains comprehensive information about mycotoxins synthesized by Fusarium species in seeds of cultivated plants and in other substrates, as well as methods for reducing their concentration in agricultural products. For preparing the textbook, information from 304 literature sources was used, including 48 works published in Uzbekistan and other CIS countries, and 256 works of foreign authors. At the end of the book are given a Glossary of biological, mycological, plant pathological, molecular genetic and phylogenetic terms, and an Index of microorganisms’ scientific (Latin) names. Please address your critical comments, feedback and suggestions to: khasanov.batyr@gmail.com.
Preprint
Full-text available
Fusarium spp. are among the most common fungal species associated with diseases both on wild and cultivated plants, including sunflowers. They infect all plant tissues causing damage to roots, bundle vessels, stems, leaves, and seeds, often causing significant yield losses. Because contaminated seeds can spread diseases into new areas and transmit them to growing plants, the quality and sanitary status of the seeds are the key to limit the spread of the disease. This study aimed to identify and determine the prevalence of Fusarium species associated with sunflower seeds and access their transmission to growing plants. A set of 49 Fusarium isolates was collected from seeds of eight sunflower cultivars. They were characterized through morphological, cultural, and genetic features. Genetic diversity was estimated through amplification of the elongation factor gene (EF-1 α), which also served to select representative isolates to perform amplification of the β-tubulin 2 gene (TUB2). There were identified four species of Fusarium (i.e., F. fabacearum , F. proliferatum , F. pseudocircinatum and F. verticillioides ) that caused seed rot, vascular darkening, withering, malformation, and stunting of growing sunflower plants. Among them, F. proliferatum was the most prevalent species. Our results highlight that various species of Fusarium are associated with damage on sunflower seeds and all of them can be transmitted through infected seeds and cause disease in growing plants.
Article
Full-text available
Fusarium, a member of the Ascomycota fungi, encompasses several pathogenic species significant to plants and animals. Some phytopathogenic species have received special attention due to their negative economic impact on the agricultural industry around the world. Traditionally, identification and taxonomic analysis of Fusarium have relied on morphological and phenotypic features, including the fungal host, leading to taxonomic conflicts that have been solved using molecular systematic technologies. In this work, we applied a phylogenomic approach that allowed us to resolve the evolutionary history of the species complexes of the genus and present evidence that supports the F. ventricosum species complex as the most basal lineage of the genus. Additionally, we present evidence that proposes modifications to the previous hypothesis of the evolutionary history of the F. staphyleae, F. newnesense, F. nisikadoi, F. oxysporum, and F. fujikuroi species complexes. Evolutionary analysis showed that the genome GC content tends to be lower in more modern lineages, in both, the whole-genome and core-genome coding DNA sequences. In contrast, genome size gain and losses are present during the evolution of the genus. Interestingly, core genome duplication events positively correlate with genome size. Evolutionary and genome conservation analysis supports the F3 hypothesis of Fusarium as a more compact and conserved group in terms of genome conservation. By contrast, outside of the F3 hypothesis, the most basal clades only share 8.8% of its genomic sequences with the F3 clade. Supplementary Information The online version contains supplementary material available at 10.1186/s12864-024-10200-w.
Article
Full-text available
We present the latest version of the Molecular Evolutionary Genetics Analysis (MEGA) software, which contains many sophisticated methods and tools for phylogenomics and phylomedicine. In this major upgrade, MEGA has been optimized for use on 64-bit computing systems for analyzing bigger datasets. Researchers can now explore and analyze tens of thousands of sequences in MEGA. The new version also provides an advanced wizard for building timetrees and includes a new functionality to automatically predict gene duplication events in gene family trees. The 64-bit MEGA is made available in two interfaces: graphical and command line. The graphical user interface (GUI) is a native Microsoft Windows application that can also be used on Mac OSX. The command line MEGA is available as native applications for Windows, Linux, and Mac OSX. They are intended for use in high-throughput and scripted analysis. Both versions are available from www.megasoftware.net free of charge.
Article
Full-text available
Fusarium oxysporum f. sp. cubense (Foc), the causal agent of Fusarium wilt or Panama disease on banana, is one of the major constraints in banana production worldwide. Indonesia is the centre of origin for wild and cultivated bananas, which likely co-evolved with Foc. This study explored the widest possible genetic diversity of Foc by sampling across Indonesia at 34 geographically and environmentally different locations in 15 provinces at six islands. This resulted in a comprehensive collection of ∼200 isolates from 40 different local banana varieties. Isolates were identified and assessed using sequence analysis of the translation elongation factor-1alpha (tef1), the RNA polymerase II largest subunit (rpb1), and the RNA polymerase II second largest subunit (rpb2). Phylogenetic analyses of these genes allowed the identification of 180 isolates of Fusarium oxysporum f. sp. cubense (Foc), and 20 isolates of the Fusarium fujikuroi species complex (FFSC), the Fusarium incarnatum-equiseti species complex (FIESC), and the Fusarium sambucinum species complex (FSSC). Further analyses, incorporating a worldwide collection of Foc strains, revealed nine independent genetic lineages for Foc, and one novel clade in the Fusarium oxysporum species complex (FOSC). Selected isolates from each lineage were tested on the banana varieties Gros Michel and Cavendish to characterise their pathogenicity profiles. More than 65 % of the isolates were diagnosed as Tropical Race 4 (Foc-TR4) due to their pathogenicity to Cavendish banana, which supports the hypothesis that Foc-TR4 is of Indonesian origin. Nine independent genetic lineages for Foc are formally described in this study. This biodiversity has not been studied since the initial description of Foc in 1919. This study provides a detailed overview of the complexity of Fusarium wilt on banana and its diversity and distribution across Indonesia.
Article
Full-text available
Infections due to Fusarium species are collectively referred to as fusariosis. Fusarium oxysporum has been reported to cause keratitis, onychomycosis, skin infections, catheter associated fungemia and has not been described as a cause of urinary tract infection. Here, we present the first case of fusariosis with urinary tract involvement in a 67 year old male, with chronic kidney disease and type 2 diabetes mellitus. This case illustrates the ever increasing spectrum of rare but offending pathogenic fungi. Early diagnosis of infection with a specific pathogen may lead to changes in antifungal therapy and may be critical for an improved outcome.
Article
Full-text available
This article describes several features in the MAFFT online service for multiple sequence alignment (MSA). As a result of recent advances in sequencing technologies, huge numbers of biological sequences are available and the need for MSAs with large numbers of sequences is increasing. To extract biologically relevant information from such data, sophistication of algorithms is necessary but not sufficient. Intuitive and interactive tools for experimental biologists to semiautomatically handle large data are becoming important. We are working on development of MAFFT toward these two directions. Here, we explain (i) the Web interface for recently developed options for large data and (ii) interactive usage to refine sequence data sets and MSAs.
Article
Full-text available
Background The Fusarium oxysporum species complex (FOSC) contains several phylogenetic lineages. Phylogenetic studies identified two to three major clades within the FOSC. The mitochondrial sequences are highly informative phylogenetic markers, but have been mostly neglected due to technical difficulties. Results A total of 61 complete mitogenomes of FOSC strains were de novo assembled and annotated. Length variations and intron patterns support the separation of three phylogenetic species. The variable region of the mitogenome that is typical for the genus Fusarium shows two new variants in the FOSC. The variant typical for Fusarium is found in members of all three clades, while variant 2 is found in clades 2 and 3 and variant 3 only in clade 2. The extended set of loci analyzed using a new implementation of the genealogical concordance species recognition method support the identification of three phylogenetic species within the FOSC. Comparative analysis of the mitogenomes in the FOSC revealed ongoing mitochondrial recombination within, but not between phylogenetic species. Conclusions The recombination indicates the presence of a parasexual cycle in F. oxysporum. The obstacles hindering the usage of the mitogenomes are resolved by using next generation sequencing and selective genome assemblers, such as GRAbB. Complete mitogenome sequences offer a stable basis and reference point for phylogenetic and population genetic studies.
Article
High populations of Fusarium oxysporum f. sp. vasinfectum, mostly between 100 and 1000 propagules/gram of soil, were found in 10 field soils assayed eight times during 10 yr after susceptible and infected cotton crops had been grown. During some of the years following the cotton crops, cotton and perennial crops such as grapevines and almond trees were grown. Among the isolates of F. oxysporum f. sp. vasinfectum two culture types were found: low sporing and high sporing types. The high sporing type was more frequently isolated from soil. The two types were further characterized by RAPD analysis and pathogenicity tests.
Article
A simple method is described for designing primer sets that can amplify specific protein-encoding sequences in a wide variety of filamentous ascomycetes. Using this technique, we successfully designed primers that amplified the intergenic spacer region of the nuclear ribosomal DNA repeat, portions of the translation elongation factor 1 alpha, calmodulin, and chitin synthase 1 genes, and two other genes encoding actin and ras protein. All amplicons were sequenced and determined to amplify the target gene. Regions were successfully amplified in Sclerotinia sclerotiorum and other sclerotiniaceous species, Neurospora crassa, Trichophyton rubrum, Aspergillus nidulans, Podospora anserina, Fusarium solani, and Ophiostoma novo-ulmi. These regions are a potentially rich source of characters for population and speciation studies in filamentous ascomycetes. Each primer set amplified a DNA product of predicted size from N. crassa.
Article
An isolate of Fusarium oxysporum that caused wilt of the Bambarra groundnut (Voandzeia subterraneo) in Tanzania was used to inoculate 44 different cultivars in numerous genera of plants. The fungus was highly virulent only on groundnut. It caused relatively mild external symptoms of wilt on some cultivars of cowpea, thus indicating a relationship to F. oxysporum f. sp. tracheiphilum. Fifty-six different ff. sp. and races of F. oxysporum that cause wilt were essentially nonpathogenic on groundnut, although twelve of them caused slight vascular discoloration, and f. sp. lupini race 3 caused severe symptoms on two plants. The Fusarium from Voandzeia is a new forma specialis, viz., Fusarium oxysporum f. sp. voandzeiae.
Article
Fusarium oxysporum f. sp. lactucae (FOL), the causal agent of lettuce Fusarium wilt, has spread to several countries where lettuce is grown. To date, four races of FOL have been identified, but only race 1 has been detected in Europe; race 2 and race 3 have been identified in Japan and in Taiwan. A new physiological race has been isolated in the Netherlands. The vegetative compatibility group approach has been used to determine genetic diversity within a group of forty-eight FOL strains of different origin, with special attention to the Italian and Dutch isolates of the pathogen, in order to obtain a better understanding of the recent epidemics that have been observed in Europe. On the basis of the complementation pattern, all the Italian isolates of FOL tested were assigned to VCG 0300. FOL isolates belonging to races 2 and 3 belonged to VCG 0301 and 0302, respectively. The isolates obtained from lettuce in the Netherlands belonged to a new VCG, numbered 0303. The results support the hypothesis that FOL race 4 might have been selected locally in Dutch fields.