You are on page 1of 62

Available online at www.studiesinmycology.org Studies in Mycology 104: 87–148 (2023).

Fusarium diversity associated with diseased cereals in China, with an updated


phylogenomic assessment of the genus

S.L. Han1,2, M.M. Wang1, Z.Y. Ma1,2, M. Raza1, P. Zhao1, J.M. Liang1, M. Gao1,2, Y.J. Li1,2, J.W. Wang3, D.M. Hu4, L. Cai1,2*
1
State Key Laboratory of Mycology, Institute of Microbiology, Chinese Academy of Sciences, Beijing 100101, P. R. China; 2College of Life Science, University
of Chinese Academy of Sciences, Beijing 100049, P. R. China; 3Institute of Biology Co., Ltd., Henan Academy of Science, Zheng Zhou 450008, Henan, P. R.
China; 4College of Bioscience & Engineering, Jiangxi Agricultural University, Nanchang 330045, Jiangxi, P. R. China

*Corresponding author: L. Cai, cail@im.ac.cn

Abstract: Fusarium species are important cereal pathogens that cause severe production losses to major cereal crops such as maize, rice, and wheat.
However, the causal agents of Fusarium diseases on cereals have not been well documented because of the difficulty in species identification and the
debates surrounding generic and species concepts. In this study, we used a citizen science initiative to investigate diseased cereal crops (maize, rice,
wheat) from 250 locations, covering the major cereal-growing regions in China. A total of 2 020 Fusarium strains were isolated from 315 diseased samples.
Employing multi-locus phylogeny and morphological features, the above strains were identified to 43 species, including eight novel species that are described
in this paper. A world checklist of cereal-associated Fusarium species is provided, with 39 and 52 new records updated for the world and China, respectively.
Notably, 56 % of samples collected in this study were observed to have co-infections of more than one Fusarium species, and the detailed associations are
discussed. Following Koch’s postulates, 18 species were first confirmed as pathogens of maize stalk rot in this study. Furthermore, a high-confidence species
tree was constructed in this study based on 1 001 homologous loci of 228 assembled genomes (40 genomes were sequenced and provided in this study),
which supported the “narrow” generic concept of Fusarium (= Gibberella). This study represents one of the most comprehensive surveys of cereal Fusarium
diseases to date. It significantly improves our understanding of the global diversity and distribution of cereal-associated Fusarium species, as well as largely
clarifies the phylogenetic relationships within the genus.

Key words: Cereal pathogens, citizen science, co-infection, new taxa, pathobiome, phylogeny, species complexes, systematics.
Taxonomic novelties: New species: Fusarium erosum S.L. Han, M.M. Wang & L. Cai, Fusarium fecundum S.L. Han, M.M. Wang & L. Cai, Fusarium
jinanense S.L. Han, M.M. Wang & L. Cai, Fusarium mianyangense S.L. Han, M.M. Wang & L. Cai, Fusarium nothincarnatum S.L. Han, M.M. Wang & L. Cai,
Fusarium planum S.L. Han, M.M. Wang & L. Cai, Fusarium sanyaense S.L. Han, M.M. Wang & L. Cai, Fusarium weifangense S.L. Han, M.M. Wang & L. Cai.

Citation: Han SL, Wang MM, Ma ZY, Raza M, Zhao P, Liang JM, Gao M, Li YJ, Wang JW, Hu DM, Cai L (2023). Fusarium diversity associated with diseased
cereals in China, with an updated phylogenomic assessment of the genus. Studies in Mycology 104: 87–148. doi: 10.3114/sim.2022.104.02
Received: 14 September 2022 ; Accepted: 17 January 2023; Effectively published online: 22 February 2023
Corresponding editor: Pedro W. Crous
Studies in Mycology

INTRODUCTION has undergone numerous significant changes based on different


criteria (Crous et al. 2021, Wang et al. 2022a), including eras
Cereal crops, including maize (Zea mays), rice (Oryza sativa) and dominated by morphology (Booth 1971, Nelson et al. 1983),
wheat (Triticum aestivum), are primary staple foods worldwide or phylogenetic inference (Gräfenhan et al. 2011, O’Donnell
(http://www.fao.org/faostat/en). The total grain output of China in et al. 2018, Lombard et al. 2019a). These different taxonomic
2020 was 658 million tons, of which 595 million tons were maize, frameworks are not necessarily consistent in inferring Fusarium
rice and wheat (Ma 2020). The significant factors diminishing species diversity. Historically, species within the F. graminearum
the value and yield of these crops were phytopathogens and clade (Fg clade) have for long been considered as a single species
pests, especially fungi (e.g., Puccinia striiformis and Fusarium using the morphological species concept, until the application of the
graminearum) (Trail 2009, Matny 2015, Savary et al. 2019). genealogical concordance phylogenetic species recognition in the
Fusarium pathogens are notorious, and not only cause yield losses, last two decades (O’Donnell et al. 2000a). Furthermore, for many
but also produce mycotoxins threatening human and animal health decades users had favoured systems with fewer species, such
(Desjardins 2006, Leslie & Summerell 2006, Renev et al. 2021). as the system proposed by Snyder & Hansen (1940, 1941, 1945,
For instance, the annual loss due to wheat scab in China was up to 1954), which reduced the number of Fusarium species to nine with
3.41 million tons from 2000 to 2018 (Su et al. 2021), and from 1993 many formae speciales. In addition, there are several dilemmas
to 2001, Fusarium diseases on crops caused economic losses up in the delimitation of Fusarium species: i) the morphological
to $7.7 billion in the United States (Nganje et al. 2004). Therefore, characteristics of Fusarium species are greatly influenced by
the investigation of species diversity, distribution patterns, host environmental factors, and often overlap with each other; ii) the
range and pathogenicity of Fusarium is of great significance for fungal universal barcode, the internal transcribed spacer region
cereal disease diagnosis and management. cistron (ITS) has a low resolution of Fusarium species; iii) most
Accurate species identification is the first step in Fusarium of the current Fusarium sequences in the public database are still
disease diagnosis and management (O’Donnell et al. 2015, under the species complex name or even under wrong species or
2018, 2022). However, the taxonomic framework of Fusarium genus names (Aoki et al. 2014, Lombard et al. 2019b, Wang et al.

© 2023 Westerdijk Fungal Biodiversity Institute. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/bync-nd/4.0/).

87
Han et al.

2019a, Xia et al. 2019, Crous et al. 2021, Yilmaz et al. 2021). These MATERIALS AND METHODS
dilemmas, together with the changes in the taxonomic system
have resulted in difficulties in the diagnosis and management of Sample collection
Fusarium diseases of cereal crops.
To solve the species delimitation and identification dilemma, Diseased cereals associated with Fusarium spp. were investigated
a polyphasic approach has gradually been applied and several in China from July 2020 to October 2021, including maize ear
online databases (Fusarium-ID, Fusarium MLST and FUSARIOID- rot (Fig. 1), maize stalk rot (Fig. 2), wheat scab (Fig. 3), wheat
ID) have been established (O’Donnell et al. 2012, Crous et al. crown rot (Fig. 4), rice spikelet rot and rice bakanae disease
2021, Torres-Cruz et al. 2022). Evolutionary relationships of (Fig. 5). Geographically, 315 diseased samples (172 diseased
several important Fusarium species complexes, e.g., F. fujikuroi maize samples, Fig. 6A; 105 diseased wheat samples, Fig. 6B;
species complex (FFSC), F. incarnatum-equiseti species complex and 38 diseased rice samples, Fig. 6C) were collected from 250
(FIESC), and F. oxysporum species complex (FOSC) were also sampling sites (several diseased samples were collected from
published (Aoki et al. 2014, Lombard et al. 2019b, Wang et al. various hosts at the same location) which covered 21 provinces,
2019a, Xia et al. 2019, Crous et al. 2021, Yilmaz et al. 2021). four autonomous regions and two municipalities (Fig. 6D). Parts
Despite these significant contributions, debates surrounding the of these samples were collected by local farmers (Supplementary
generic delimitation of Fusarium, especially whether the genus Fig. S1, Supplementary Table S1) who followed a directive protocol
Neocosmospora (also known as F. solani species complex, FSSC) in our Citizen Science Initiative of “Collecting Diseased Cereals”
belongs to Fusarium was disputed (Crous et al. 2021, Geiser et al. posted on social media. Diseased samples in good condition were
2021, Wang et al. 2022a). dried and stored in the Fungarium of the Institute of Microbiology,
According to the USDA fungal database (Farr & Rossman 2022), Chinese Academy of Sciences, Beijing, China (HMAS).
more than 20 % of cereal-associated Fusarium species belong to
the F. sambucinum species complex (FSAMSC). Nevertheless, Strain isolation, preservation and selection
the species relationships within this complex remain to be further
clarified, as several sister species cannot be distinguished based A total of 2 600 fungal strains were isolated from symptomatic tissues
on only RNA polymerase second largest subunit (rpb2) and (i.e., kernels, stems, roots; Figs 1–5) of diseased cereals using single
translation elongation factor 1-alpha (tef1) phylogenetic analyses spore isolation or direct hyphal isolation (Zhang et al. 2013, Raza et
(Laraba et al. 2021). Recently, the phylogenomic approach has al. 2019). Of these, 2 020 strains were assigned to Fusarium based on
been increasingly shown to be able to provide more information colony morphology and tef1 sequences. Strains were further selected
towards understanding species evolution and boundaries following three steps: firstly, for strains with 100 % tef1 similarity from
(Haridas et al. 2020, Liu et al. 2022a). In the case of Fusarium, the same symptomatic tissue of the same host species in each
a phylogenomic approach has also been used to assess generic location, only one strain was retained for subsequent analyses.
boundaries (Geiser et al. 2021), but still suffered from imbalanced Secondly, the genus level phylogenetic analysis was conducted using
generic sampling, and lack of data derived from ex-type cultures. tef1-rpb2 sequences, and the above selected strains were assigned
Previously, Fusarium associated with cereals in China were to six Fusarium species complexes. For each complex, a tef1-rpb2
identified mainly based on morphological features, preliminary phylogeny was re-analysed, and only one isolate in each subclade
nucleotide BLAST, or phylogenetic analyses using single locus (without sequence variation) was selected if multiple isolates from the
datasets (either ITS or tef1) (Huang et al. 2011, Zhang et al. 2012, same symptomatic tissue of each host species in each location were
2015, Xu et al. 2016, Hao et al. 2017). These studies, particularly available. Thirdly, different loci were amplified for the above selected
those applying subspecies, varieties and formae speciales names, strains, and phylogenetic analyses of Fusarium species complexes
proved to be insufficient to identify Fusarium species (Lombard et were conducted using different multi-locus datasets, i.e., calmodulin
al. 2019b, Xia et al. 2019, Yilmaz et al. 2021, Wang et al. 2022a). (CaM), RNA polymerase largest subunit (rpb1), rpb2, tef1 and beta-
For instance, the pathogen of rice bakanae diseases in China has tubulin (tub2) for FFSC; CaM, rpb2 and tef1 for the FIESC and
been identified as three Fusarium species and varieties, i.e., F. FOSC; histone (H3), rpb1, rpb2 and tef1 for the FSAMSC; rpb1, rpb2
moniliforme, F. moniliforme var. subglutinans and F. moniliforme and tef1 for the FNSC and FTSC. Only one isolate of a particular
var. intermedium, based on morphological features (Booth 1971, species from each symptomatic tissue of the same host species in
Ye et al. 1990). However, these epithets have been revealed to each location was selected. After the above selection steps, 608
be synonyms of F. fujikuroi (O’Donnell et al. 1998a, Crous et al. representative strains were included (Supplementary Table S1).
2021). Thus, species diversity and distribution of this group require The type specimens of new species described in this study were
amendment in accordance with the currently used taxonomic deposited in the Fungarium of the Institute of Microbiology, Chinese
system and names. Academy of Sciences, Beijing, China (HMAS), and the living ex-type
The purpose of this study was therefore to examine the Fusarium cultures were deposited in the China General Microbiological Culture
species associated with diseased cereals in China. Specifically to: Collection Centre (CGMCC). All isolates examined in this study were
i) understand the diversity, distribution patterns, host preference deposited in Lei Cai’s personal culture collection (LC), the Institute
and pathogenicity of Fusarium species associated with diseased of Microbiology, Chinese Academy of Sciences, Beijing, China.
cereals; ii) reassess the boundaries of Fusarium s. str. and allied Taxonomic novelties, descriptions and nomenclature were deposited
genera, and assess species boundaries within the FSAMSC; iii) in MycoBank (Crous et al. 2004).
sequence the genomes of new and several known species and
provide an updated phylogenomic overview of Fusarium. DNA extraction and amplification

Total genomic DNAs were extracted from fresh fungal mycelia


grown on potato dextrose agar (PDA; DifcoTM, Becton, Dickinson and

88
Fusarium spp. on cereals in China

Fig. 1. Field symptoms of maize ear rot (MER). A. Kernels covered with salmon-coloured powdery molds and accompanied by insect injury. B. Kernels
covered with white moulds. C. Kernels covered with orchid-coloured mycelia. D, F. Ears covered with whitish or a mixture of pinkish and blackish powdery
moulds coexist with symptoms caused by other fungi. E. Kernels covered with white moulds or violet mycelia. G. Kernels covered with salmon-coloured
powdery moulds. Note: Maize ear rot caused by Fusarium was referred to as Fusarium ear rot of maize (Zhang et al. 2014a), Fusarium maize ear rot (Zhang
et al. 2014b), Gibberella/red ear rot (Lana et al. 2022) and Fusarium/pink ear rot (Zhang et al. 2016) in various previous studies. To avoid confusion, in this
study we refer to this disease as maize ear rot (MER).

www.studiesinmycology.org 89
Han et al.

Fig. 2. Field symptoms of maize stalk rot (MSR). A, B. Maize ear drooping without shedding. C. Stalks turn grey green; internodes turn straw-coloured or
dark brown. D. Stalks covered with whitish mycelia. E, F. Stalks were covered with salmon-coloured powdery moulds and snapped at the nodes. G. Rotted
and brownish stalks. Note: Maize stalk rot caused by Fusarium, known as Fusarium stalk rot (Jiang et al. 2021), Gibberella stalk rot in maize (Ye et al. 2013)
in previous literature, is herein universally referred to as maize stalk rot (MSR) in this paper.

90
Fusarium spp. on cereals in China

Fig. 3. Field symptoms of Fusarium head blight of wheat (FHB). A–E. Head blight of wheat accompanied by pinkish mould and/or whitish mycelia. Note:
The disease name Fusarium head blight (FHB) is adopted in this paper, which was also called wheat head blight and wheat scab in literature (O’Donnell et
al. 2000a).

Company, Sparks, MD, USA) using the cetyltrimethylammonium Phylogenetic analyses


bromide (CTAB) method (Porebski et al. 1997) and stored at -20 °C
until polymerase chain reaction (PCR). PCR amplifications were Consensus sequences were obtained using MEGA v. 7 software
performed in a reaction mixture consisting of 12.5 μL 2 × Taq PCR (Kumar et al. 2016), and sequences for each locus were aligned
Master Mix (Vazyme Biotech Co., Ltd, Nanjing, China), 1 μL each using MAFFT v. 7 (Katoh & Standley 2013). Misalignments were
of 10 μM primers, 2 μL of the undiluted genomic DNA, adjusted to corrected manually where necessary. Phylogenetic analyses were
a final volume of 25 μL with distilled deionised water. Seven loci, performed based on individual and combined datasets, using
including ITS (White et al. 1990), tef1 (O’Donnell et al. 1998b), CaM Maximum-Likelihood (ML) and Bayesian Inference (BI) methods
(O’Donnell et al. 2000b), rpb1 (O’Donnell et al. 2010), rpb2 (Liu et through the CIPRES Science Gateway portal (https://www.phylo.
al. 1999; Reeb et al. 2004), tub2 (O’Donnell & Cigelnik 1997), and org/; Miller et al. 2012).
H3 (Roux et al. 2001) were amplified and sequenced. The PCR The ML analyses were carried out using RAxML-HPC v. 8.2.12
primer pairs and amplification conditions are listed in Table 1. The (Stamatakis 2014), with 1 000 replicates under the GTR+GAMMA
PCR products were visualised using 1 % agarose electrophoresis model. The Bayesian analyses were carried out using MrBayes v.
gels. Sequencing was done by the Tianyi Huiyuan Company 3.2.7a (Huelsenbeck & Ronquist 2001, Ronquist & Huelsenbeck
(Beijing, China) and the SinoGenoMax Company (Beijing, China). 2003), incorporating the best evolutionary models for each marker
as determined by MrModelTest v. 2.4 (Nylander 2004). Bayesian
analyses were computed with four simultaneous Markov Chain

www.studiesinmycology.org 91
Han et al.

Fig. 4. Field symptoms of Fusarium crown rot of wheat (FCR). A, B. “White-head” of wheat tillers. C, D. Reddish-brown discoloration on the lower stems.
E. Dark brown to black discolouration on the crowns and roots. Note: The disease name Fusarium crown rot of wheat (FCR) is adopted in this paper, which
was also referred to as wheat crown rot (Zhang et al. 2015), crown rot of wheat in previous literature (Li et al. 2016).

92
Fusarium spp. on cereals in China

Fig. 5. Field symptoms of Fusarium diseases on rice. A. Rice bakanae disease (RBD), caused by F. fujikuroi, showing elongated barren seedlings. B–E.
Rice spikelet rot (RSR), caused by multiple Fusarium species, showing reddish or brown discoloration on the glumes, sometimes with salmon-coloured and/
or whitish powdery mould. Note: The disease name rice spikelet rot (RSR) is adopted in this paper, which was also known as Fusarium head blight in rice
in literature (Liu et al. 2022d).

www.studiesinmycology.org 93
Han et al.

Fig. 6. Maps showing sampling sites in China, generated by ArcGIS v. 10.5 software (Esri, Redlands, CA, USA). A. Map showing the distribution of 172
diseased maize samples collected in this study. B. Map showing the distribution of 38 diseased rice samples collected in this study. C. Map showing the
distribution of 105 diseased wheat samples collected in this study. D. Map showing the overall distribution of a total of 315 diseased samples collected in
this study.

Monte Carlo chains, 10 M generations, and a sampling frequency of Phylogenetic analyses of different Fusarium species complexes
1 000 generations for the first datasets and 100 generations for the were performed using different multi-locus datasets in accordance
other two datasets, ending the run automatically when the standard with previous studies (Sarver et al. 2011, O’Donnell et al. 2013,
deviation of split frequencies fell below 0.01. The burn-in fraction Gräfenhan et al. 2016, Laurence et al. 2016, Sandoval-Denis et
was set to 0.25, after which the 50 % majority rule consensus trees al. 2018b, Torbati et al. 2018, Lombard et al. 2019c, Maryani et al.
and posterior probability (PP) values were calculated. 2019b, Wang et al. 2019a, Xia et al. 2019, Yilmaz et al. 2021). The
The clade is supported when its RAxML Bootstrap support composition of the multi-locus datasets, outgroup taxa, number of
value is ≥ 70 %, and the Bayesian PP value is ≥ 0.9. The resulting characters and the best models are listed in Supplementary Table
trees were plotted using FigTree v. 1.4.2 (http://tree.bio.ed.ac.uk/ S2. Specifically, the combined CaM, rpb1, rpb2, tef1 and tub2
software/figtree). All sequences and their alignments generated datasets were constructed for phylogenetic analyses for the FFSC,
from 2 020 Fusarium strains in this study were deposited in rooted with F. nirenbergiae CBS 744.97. Phylogenetic analyses of
GenBank (Supplementary Table S1) and TreeBASE (submission the FIESC were performed by using the combined CaM, rpb2 and
ID 30020), respectively. tef1 datasets, rooted with F. concolor NRRL 13459. Phylogenetic

94
Table 1. Primers used in this study, with originating loci, sequences program and references.
Gene/DNA regions Primers

Name Abbreviation Name Direction Sequence (5’→3’) PCR amplification procedures References
Beta tubulin tub T1 Forward AACATGCGTGAGATTGTAAGT 95 °C 3 min; 35 cycles of 94 °C 30 s, 54 °C 45 s, 72 °C 15 s; 72 O’Donnell & Cigelnik (1997)
°C 10 min; 10 °C soak
T2 Reverse TAGTGACCCTTGGCCCAGTTG

Calmodulin CaM CL1 Forward GARTWCAAGGAGGCCTTCTC 95 °C 1 min; 35 cycles of 94 °C 30 s, 55 °C 30 s, 72 °C 15 s; 72 O’Donnell et al. (2000b)
°C 10 min; 10 °C soak
CL2A Reverse TTTTTGCATCATGAGTTGGAC

Histone H3 H3-la Forward ACTAAGCAGACCGCCCGCAGG 96 °C 2 min; 30 cycles of 92 °C 1 min, 60 °C 1 min, 72 °C 10 s; Roux et al. (2001)
72 °C 5 min; 10 °C soak

www.studiesinmycology.org
H3-lb Reverse GCGGGCGAGCTGGATGTCCTT

Internal transcribed spacer region of the ITS ITS1 Forward CCGTAGGTGAACCTGCGG 95 °C 5 min; 35 cycles of 95 °C 30 s, 52 °C 30 s, 72 °C 10 s; 72 White et al. (1990)
rDNA °C 5 min; 10 °C soak
ITS4 Reverse TCCTCCGCTTATTGATATGC

RNA polymerase largest subunit rpb1 F7 Forward CRACACAGAAGAGTTTGAAGG 95 °C 5 min; 5 cycles of 95 °C 2 min, 58 °C 45 s, 72 °C 20 s; 5 O’Donnell et al. (2010)
cycles of 95 °C 2 min, 57 °C 45 s, 72 °C 20 s; 35 cycles of 95
G2R Reverse GTCATYTGDGTDGCDGGYTCDCC °C 2 min, 56 °C 45 s, 72 °C 20 s; 72 °C 10 min; 10 °C soak

RNA polymerase second largest subunit rpb2 5f2 Forward GGGGWGAYCAGAAGAAGGC 94 °C 90 s; 35 cycles of 94 °C 45 s, 57 °C 45 s, 72 °C 20 s; 72 Reeb et al. (2004)
°C 5 min; 10 °C soak
7cf Forward ATGGGYAARCAAGCYATGGG Liu et al. (1999)

7cr Reverse CCCATRGCTTGYTTRCCCAT

11ar Reverse GCRTGGATCTTRTCRTCSACC

translation elongation factor 1-alpha tef1 EF-1 Forward ATGGGTAAGGARGACAAGAC 94 °C 90 s; 35 cycles of 94 °C 45 s, 55 °C 45 s, 72 °C 15 s; 72 O’Donnell et al. (1998b)
°C 10 min; 10 °C soak
EF-2 Reverse GGARGTACCAGTSATCATG

95
Fusarium spp. on cereals in China
Han et al.

Fig. 7. Pathogenicity test of maize stalk rot (MSR) fulfilling Koch’s postulates. A. Diagram of a three-leaf-old seedlings of maize. B. Typical symptoms
of disease severity (0–4). C. Blank control treated with sterile water. D–U. Maize seedlings treated with different Fusarium suspensions. The conidial
concentration (CC) of Fusarium strains and disease severity index (DSI) of infected plants were listed. DSI (%) = [sum (class frequency × score of rating
class)/(total number of plants × maximal disease severity score)] × 100.

96
Fusarium spp. on cereals in China

analyses of the F. nisikadoi species complex (FNSC) were the Sordaromyceta_odb9 gene set. Gene predictions were made
performed by using the combined rpb1, rpb2 and tef1 datasets, using the Funannotate pipeline v. 1.7.0 (Palmer 2016); repetitive
rooted with F. concolor NRRL 13994. Phylogenetic analyses of the elements were initially soft-masked using default parameters;
FOSC were performed by using the combined CaM, rpb2 and tef1 funannotate predict was implemented using Sordariomycetes
datasets, rooted with F. udum CBS 177.31. Phylogenetic analyses BUSCO database and Fusarium graminearum as seed species.
of the FSAMSC were performed by using the combined H3, rpb1, Genome assemblies were deposited in the National Microbiology
rpb2 and tef1 datasets, rooted with F. nelsonii NRRL 13338. Data Centre (NMDC) under BioProject NMDC10018280.
Phylogenetic analyses of the F. tricinctum species complex (FTSC)
were performed by using the combined ITS, rpb1, rpb2 and tef1 Phylogenomic tree construction
datasets, and rooted with F. concolor NRRL 13459.
Predicted proteins were clustered into orthologous groups by
Morphological observations using Orthofinder v. 2.3.3 (Emms & Kelly 2015, 2019). Amino acid
sequences of single-copy orthologs were aligned using MAFFT
Fungal isolates were studied morphologically based on their v. 7 (Katoh & Standley 2013). Conserved sites in the alignment
macroscopic and microscopic features (Crous et al. 2021). Petri were extracted using Gblocks v. 0.91b (Castresana 2000). The
plates were incubated for 7 d at 25 °C. Agar pieces of approximately substitution model and maximum-likelihood tree were inferred
5 × 5 mm were taken from edge of colonies on synthetic nutrient- by IQ-TREE v. 2.2.0.8 (Kalyaanamoorthy et al. 2017, Hoang et
poor agar (SNA; Nirenberg 1976) and transferred onto media for al. 2018, Minh et al. 2020a), and the clade is supportive when its
morphological characterisation. For morphological observations, ultrafast bootstrap (UFBoot) support value is ≥ 95 %. In addition,
PDA and oatmeal agar (OA) media were used. The culture gene concordance factors (gCF) and site concordance factors
characteristics of the colony, including pigmentation and odour, (sCF) were calculated using the “–gcf” and “–scf” options in IQ-
were observed after 7 d of incubation in the dark on PDA and OA TREE v. 2.2.0.8 (Minh et al. 2020b), to quantify genealogical
(Crous et al. 2021). Colours were rated according to the colour concordance. For every branch of a species tree, gCF is defined
charts of Kornerup & Wanscher (1978). as the percentage of “decisive” gene trees containing that branch,
For morphological comparisons, carnation leaf agar (CLA; and sCF is defined as the percentage of decisive alignment sites
Fisher et al. 1982) was used. Micromorphological characteristics, supporting that branch (Minh et al. 2020b).
including sporodochia, conidiophores, phialides, conidia
(sporodochial and aerial conidia) and chlamydospores, were Diversity and distribution analyses
observed after 7–14 d of incubation under a 12/12 h near-ultraviolet
light/dark cycle at 25 °C (Leslie & Summerell 2006, Crous et al. The richness of Fusarium species was defined as recorded species
2021). Morphological characteristics were examined and photo- number associated with different hosts and climate regions (hosts:
documented with water as mounting medium under a Nikon 80i maize, rice, wheat; climate regions: regions affected by plateau
compound microscope with differential interference contrast mountain, subtropical monsoon, temperate continental, temperate
(DIC) optics, and a Nikon SMZ1500 dissecting microscope. For monsoon, and tropical monsoon climate). Co-infection was defined
each species, 30 phialides and chlamydospores, and 50 conidia as a single diseased cereal sample (e.g., a single wheat head,
were randomly measured to calculate the mean value, standard maize ear) from which more than one Fusarium species was
deviation, and minimum–maximum values. isolated. The relative percentage was calculated by counting the
number of each Fusarium species presented in different groups,
Whole-genome sequencing, assembly and gene and was visualised by the ggplot2 package in R v. 4.1.0 (Wickham
annotation 2016). The diversity of Fusarium species and distribution patterns
across different hosts and geographic regions were performed
Whole-genome sequences were generated for eight ex-type using GraphPad Prism v. 8.0.2.
strains of new species (i.e., F. erosum, F. fecundum, F. jinanense,
F. mianyangense, F. nothincarnatum, F. planum, F. sanyaense, and Pathogenicity assays
F. weifangense) and 32 strains of known species (Supplementary
Table S3). All strains were purified using a single-spore isolation Pathogenicity tests were performed to determine whether the
method (Zhang et al. 2013). Hyphae of 7-d-old colonies growing on species isolated from the symptomatic tissues of diseased cereals
PDA were collected and then stored at -80 °C. Genome extraction are pathogens. The Fusarium species isolated from maize stalk rot
and sequencing were done by the Annoroad Gene Technology were used as examples, because maize is the largest grain crop in
Company (Beijing, China). The DNA libraries were sequenced China (http://www.stats.gov.cn/), and maize stalk rot has been one
as 150 bp pair-end reads using Illumina NovaSeq 6000 platform. of the primary diseases that threaten maize production in recent
Reference genomes were retrieved from the public database years (Wang et al. 2010). In brief, we completed pathogenicity
(Supplementary Table S3). tests fulfilling Koch’s postulates of all Fusarium species isolated
The genome assembly and gene annotation were conducted from maize, which have not yet been reported as causal agents of
following the protocol of Ma et al. (2021). Specifically, read quality was maize stalk rot in previous studies.
assessed by using FastQC v. 0.11.8 (Andrews & Babraham 2010). Specifically, representative isolates were inoculated on
Clean reads were assembled with SPAdes v. 3.12.0 (Bankevich et seedlings of the maize cultivar Zhengdan 958, as Ye et al. (2013)
al. 2012), using the “careful” mode and various kmers (21, 33, 55, and Han et al. (2022) described but with slight modifications.
77, 99). Genome assembly quality was assessed using QUAST v. Hyphae of 7-d-old colonies growing on SNA were transferred to 400
5.0.2 (Alexey et al. 2013). Genome completeness was assessed mL of carboxymethylcellulose sodium fluid medium, and cultivated
using genome mode in BUSCO v. 5.3.2 (Seppey et al. 2019), with for 3–7 d at 25 °C at 180 rpm. Then the conidial concentration

www.studiesinmycology.org 97
Han et al.

LC18483
0.009 LC18740
LC18518
LC18623
FFSC LC18416
CaM-rpb1-rpb2-tef1-tub2 LC18470
LC18838
LC18468
CBS 734.97
New Chinese record LC18434
LC18389
New host record (maize) 75/-
LC18645
LC18746
New host record (rice) LC18624
LC18408
LC18490
New host record (wheat) LC18755
LC18430
LC18736
LC18797
LC18491
LC18587
LC18707
LC18836
LC18479
LC18296
LC18830
African clade

LC18754
LC18842
LC18466
LC18495
-/0.98 LC18839
LC18431
LC18599
LC18834
LC18469
Asian clade

LC18519
LC18841
LC18651
LC18843
African clade C
LC18471
LC18388
LC18604
American clade

0.01
LC18760
LC18547
LC18482
LC18500
African clade B LC18526 F. verticillioides
LC18562
LC18418
LC18586
76/0.99
LC18735
LC18494

African clade
LC18433
LC18437
LC18757
LC18465
LC18751
LC18616
LC18467
LC18714
LC18248
LC18459
LC18462
LC18741
LC18615
LC18503
CBS 218.76 ET
LC18729
LC18394
77/-
LC18461
LC18460
LC18399
86/- LC18498
74/-
LC18464
LC18475
LC18391
87/1 LC18730
88/1
-/1 LC18725
LC18607
83/- LC18598
LC18450
LC18525
LC18828
2× LC18456
100/1 LC18727
LC18432
LC18613
76/0.99 LC18458
LC18463
LC18827
LC18489
-/0.97 CBS 119825
100/1 CBS 624.87 T
NRRL 28893 F. musae

Fig. 8. Phylogeny inferred based on the combined CaM-rpb1-rpb2-tef1-tub2 gene regions of the Fusarium fujikuroi species complex (FFSC). Fusarium
nirenbergiae (CBS 744.97) was used as an outgroup. Strains isolated in this study were indicated in red. Pathogenetic strains from previous studies were
indicated in green. The RAxML Bootstrap support values (ML-BS > 70 %) and Bayesian posterior probabilities (BI-PP > 0.9) were displayed at the nodes
(ML-BS / BI-PP). Ex-type, ex-epitype and ex-neotype strains were indicated in bold with T, ET, and NT, respectively.

98
Fusarium spp. on cereals in China

PPRI 20386
0.009 98/1 PPRI 20458 T F. curculicola
92/1 PPRI 20464
FFSC 100/1 98/1 NRRL 25615 F. volatile
CaM-rpb1-rpb2-tef1-tub2 CBS 143874 T
NRRL 66233 T F. coicis
86/1 CMM 3557 T F. gigantea
New Chinese record LC18578
87/1
76/- LC15876 T
LC18411
New host record (maize) R34
86/- F. planum sp. nov.
95/1
LC18487
New host record (rice) LC18445
91/1
LC18541
New host record (wheat) 100/- LC18737
98/1 CBS 119856
CBS 119857 T F. andiyazi
CML 2756
100/1 100/1 CBS 146656
-/0.99 -/0.99 CBS 146669 T F. madaense
100/1 CML 2813
CML 2791
African clade

CML 3858
78/1 CML 3871
CML 3872 F. mirum
CML 3859 T
88/- PPRI 2188.3 T
-/0.97 100/1 PPRI 20462 F. casha
83/1 PPRI 20468
Asian clade

NRRL 66243 T F. tjaetaba


100/1 CBS 135139
99/1
CBS 748.97 T F. napiforme
African clade C 100/1 CBS 526.97 F. ramigenum

African clade
CBS 418.97 T
100/1 CBS 776.96 T F. thapsinum
0.01
American clade

CBS 100312
100/1 CBS 404.97 T F. brevicatenulatum
100/1 CBS 100196
100/1 CBS 414.97 T
African clade B 92/1 CBS 745.97 F. pseudoanthophilum
100/1 CBS 484.94
93/1 CBS 417.97 T F. pseudonygamai
CPC 30825
99/1
CBS 146498 T F. prieskaense
100/1
100/1 CBS 125181 F. ficicrescens
CBS 125178 T
100/1 CBS 675.94 F. sudanense
74/1 CBS 454.97 T
100/1 CBS 119850
92/1 CBS 483.94 T F. terricola
95/-
100/1 CBS 139387 F. nygamai
CBS 749.97 T
100/1 NRRL 66890 F. xyrophilum
NRRL 62721 T
100/1 CBS 449.97 T F. pseudocircinatum
CBS 455.97
100/1 CBS 407.97 T F. denticulatum
86/0.98 CBS 406.97
100/1 CBS 420.97 F. lactis
CBS 411.97 ET
100/1 CML 3163 F. brachiariae
CML 3032 T
100/1 CML 3657 T F. caapi
CML 3658
100/1 RGB5717 T F. mundagurra
99/1 NRRL 25221 T
100/1 NY 001B5 F. chinhoyiense
100/1 CBS 749.79 F. xylarioides
-/0.98 CBS 258.52 T
71/- CBS 216.76 T F. phyllophilum
CBS 246.61
100/1 NRRL 52712 F. longicornicola
100/1 NRRL 52706 T
100/1 LC7502 T F. aquaticum
70/1 LC13615
NRRL 25199 ET F. udum
86/1 CBS 178.32
100/1 NRRL 62594 F. secorum
99/1 NRRL 62593 T
100/1 CBS 402.97 T F. acutatum
CBS 137545

Fig. 8. (Continued).

www.studiesinmycology.org 99
Han et al.

LC18602
0.009 LC18804
LC18512
FFSC LC18622
CaM-rpb1-rpb2-tef1-tub2 LC18802
LC18497
LC18492
New Chinese record LC18807
LC18756
New host record (maize) LC18655
LC18743
New host record (rice) LC18496
LC18818
New host record (wheat) LC18571
LC18677
LC18753
LC18811
LC18528
LC18405
LC18398
LC18786
LC18355
African clade

88/1 LC18390
LC18809
LC18798
LC18806
LC18392
LC18781
ITEM 2341
LC18654
Asian clade

24ALH “F. proliferatum ” in Wulff et al. (2010);


LC18658 causing rice bakanae disease
African clade C
LC18439
LC18829
NRRL 25089
LC18517
American clade

0.01

LC18480
LC18527
LC18706
African clade B CBS 139739
LC18762
LC18734
LC18424 F. annulatum
LC18511
LC18501

Asian clade
F1 “F. proliferatum ” in Okello et al. (2019);
LC18493 causing corn root rot
LC18406
CBS 115.97
LC18417
LC18600
LC18821
LC18627
CBS 217.76 representative isolate of
LC18478 “F. proliferatum ” in previous studies
LC18502
LC18717
LC18521
LC18589
LC18790
-/1 LC18646
LC18653
84/0.99 LC18448
LC18657
-/0.97 LC18472
LC18831
98/1 LC18647
LC18833
97/-
LC18538
CBS 258.54 T
-/0.97
LC18817
LC18803
P2 “F. proliferatum ” in Okello et al. (2019);
NRRL 62905 causing corn root rot
90/1 51ALH “F. proliferatum ” in Wulff et al. (2010);
LC18566 causing rice bakanae disease
LC18591
79/-
LC18713
LC18728
LC18548
CBS 143605
80/1
LC18414
100/1 95/1 LC18726
CBS 792.91
CBS 137537
LC18801
Azerbaijan 71-3 “F. proliferatum ” in Özer et al. (2020);
LC18481 causing wheat crown and root rot
91/-
LC18473
84/1
100/1 LC13644 T F. hechiense
LC13645

Fig. 8. (Continued).

100
Fusarium spp. on cereals in China

0.009 100/1 LC18579


LC15877 T F. erosum sp. nov.
100/1 LC18581
FFSC 88/1 CPC 27189
CBS 142222 T F. siculi
CaM-rpb1-rpb2-tef1-tub2 92/1
CBS 430.97
98/1 CBS 429.97
CBS 431.97 F. globosum
New Chinese record
CBS 428.97 T
LC18791
New host record (maize) LC18819
LC18776
New host record (rice) LC18626
100/1 LC18788
New host record (wheat) LC18789
-/0.96
LC18611
-/0.99 LC18824
LC18446
76/0.98 CBS 221.76 T
CBS 265.54 F. fujikuroi
84/- LC18569
NRRL 13289
LC18814
African clade

LC18499
100/0.99 LC18601
-/0.99 LC18656

Asian clade
96/1 LC18413
LC18367
LC18588
81/1 LC18564
99/1 LC18597
Asian clade

LC18606
-/0.99
LC18590
LC13628
African clade C LC13627 T F. elaeagni
99/1 LC18815
CBS 119853
American clade

0.01
100/1 NRRL 25226
100/1 96/1
F. mangiferae
CBS 120994 T
CBS 480.96 ET
100/1
F026 F. proliferatum
African clade B
LC18540
100/1 LC15882 T F. sanyaense sp. nov.
LC18537
InaCCF872 T
InaCCF993 F. lumajangense
91/1 LC18523
CBS 453.97
-/1 LC18820
100/1 F. concentricum
95/1 CBS 450.97 T
LC18575
90/0.99 LC13656 T F. panlongense
100/1 MUCL 55950
100/1 CBS 148464 T F. chuoi
ZHKUCC_22_0079
100/1 ZHKUCC_22_0078 F. aglaonematis
100/1 ZHKUCC_22_0077 T
NRRL 28852 T F. fractiflexum
LC18574
87/-
LC18536
CBS 139376
99/1 72/1
LC18563 F. sacchari
100/1 LC18533
100/1 CBS 223.76 ET
LC18522
clade C
African

BRIP 71717 T F. dhileepanii


100/1 CBS 146497 T
CBS 146496 F. echinatum

Fig. 8. (Continued).

(CC) of Fusarium suspension were adjusted to 106~107 conidia/mL Disease severity was estimated visually on a 0 to 4 scale as
by haemocytometer, and the suspension were stored at 4 °C until described in previous studies (Jin et al. 1994). Disease severity
maize seedlings became three-leaf-old. The maize kernels were was displayed in Fig. 7B and categorised as follows: 0, no visible
soaked in water for 24 h, moisturised in a wet chamber for 48 h, symptoms; 1, the dry area of the first leaf under 50 %, and the dry
and then incubated in sterilised soil under 16 h of light and 8 h of area of the second leaf under 25 %; 2, the dry area of the first leaf
over 50 % but under 100 %, and the dry area of the second leaf
darkness at 25 °C. After 7 d, the roots (three-leaf-old seedlings, Fig.
under 50 %; 3, the first leaf was completely withered, the dry area
7A) were immersed into the prepared suspension and incubated for of the second leaf over 50 % but under 100 %, and the dry area
6 h at 25 °C, then transferred to soil (25 °C, 16 h light/25 °C, 8 h of the third leaf was under 25 %; 4, the three leaves of inoculation
dark). As a negative control, seedlings were inoculated with sterile plants were completely withered. The disease severity index (DSI)
water. For each tested fungal strain, eight plant replicates were was calculated based on Chiang et al. (2017). Specifically, DSI (%)
used for the pathogenicity test. Pathogenicity was evaluated based = [sum (class frequency × score of rating class)/(total number of
on leaf symptoms of the tested maize plants 3 d post-inoculation. plants × maximal disease severity score)] × 100.
Koch’s postulates were fulfilled by re-isolating the fungus from
symptomatic plants.

www.studiesinmycology.org 101
Han et al.

NRRL 66333
0.009 LC18808
LC18573
73/-
FFSC LC18810
LC18603
CaM-rpb1-rpb2-tef1-tub2 LC18840
LC18532
LC18731 F. subglutinans
New Chinese record 100/1
LC18457
LC18484
New host record (maize) LC18837
74/1 CBS 747.97 NT
New host record (rice) LC18835
84/- LC18249
New host record (wheat) LC18783
LGMF 1930 T F. awaxy
88/0.98 LC18488
LC18785
LC18813
-/1 LC18531
LC18812 F. temperatum
99/1
MUCL 52463 T
African clade

100/1
100/1
CBS 135539
98/-
LC18577
LC18455
98/1 NRRL 20476
CBS 100057 T F. bactridioides
-/1 90/- CBS 222.76 ET
F. anthophilum

American clade
CBS 119858
CBS 187.34
Asian clade

98/1 CBS 219.76 ET


F. succisae

NRRL 25807 T F. babinda
African clade C
CBS 125535 T F. werrikimbe
100/1 CBS 139383
CBS 119849 T F. konzum
CMW 25245 T F. fracticaudum
American clade

0.01

CMW 25267 T F. parvisorum


96/1 NRRL 53984 T F. tupiense
CMW 25243 T F. pininemorale
African clade B CMW 25512 T F. marasasianum
CBS 119864
70/1 100/1 CBS 405.97 T F. circinatum
-/1 CBS 137242 T F. sororula
100/1 NRRL 29123
F. pilosicola
100/1 -/1 NRRL 29124 T
CBS 220.76 T F. bulbicola
100/1 100/1 CBS 118516 T
-/0.96 F. ananatum
100/1 CMW 28598
CBS 409.97 T F. guttiforme
99/1
CBS 452.97 T
CBS 403.97 F. begoniae
100/1
NRRL 25623 T F. sterilihyphosum
2× CBS 118515
F. ophioides
100/1 CBS 118512 T
NRRL 53147 T F. mexicanum
NRRL 53580
100/1 CBS 100193
F. agapanthi
clade B
NRRL 54463 T African
100/1 CBS 119860 T
CBS 481.94 F. dlaminii
CBS 144209 T F. fredkrugeri
CBS 744.97 F. nirenbergiae Outgroup

Fig. 8. (Continued).

RESULTS FNSC (Supplementary Fig. S5), FOSC (Supplementary Fig. S6),


FSAMSC (Supplementary Fig. S7) and FTSC (Supplementary Fig.
Phylogenetic analyses S8). The phylogenetic analyses based on single genes showed that
the rpb2 locus provided a better resolution in species recognition
The preliminary phylogenetic analyses based on combined rpb2 for the six species complexes. Using the three most species-rich
and tef1 loci revealed that the 608 representative isolates were complexes (FFSC, FIESC, FSAMSC) as examples, the rpb2 locus
distributed over six Fusarium species complexes (Supplementary was able to recognise 59, 33, and 35 species among the FFSC
Fig. S2). Phylogenetic analyses were performed respectively for (with 84 species), FIESC (with 44 species), and FSAMSC (with 41
each Fusarium species complex using different datasets. Combined species), respectively.
with morphological characters, these isolates were identified to 12
species in the FFSC (Fig. 8), 17 species in the FIESC (Fig. 9), one
species in the FNSC (Fig. 10), two species in the FOSC (Fig. 11),
Phylogenomic assessment of Fusarium and allied
nine species in the FSAMSC (Fig. 12), and two species in the FTSC genera
(Fig. 13). In summary, the 608 selected strains were identified to 43
species, including 35 known and eight novel species. Employing 228 assembled genomes covering 17 Fusarium
Furthermore, single gene trees were evaluated respectively for species complexes, 11 Fusarium allied genera and one outgroup
FFSC (Supplementary Fig. S3), FIESC (Supplementary Fig. S4), (Stylonectria norvegica IHI 201603), a high-confidence and genome-

102
Fusarium spp. on cereals in China

LC18700
0.03 LC18823
LC18695
FIESC LC18261
CaM-rpb2-tef1 LC18632
LC18444
New Chinese record LC18421
LC18745
New host record (maize) LC18738
LC18799
New host record (rice) LC18698
LC18508
New host record (wheat) LC18419
LC18402
LC18633
LC18425
LC18649
LC18684
LC18689
LC18340
LC18429
Incarnatum clade

LC18642

Incarnatum clade
LC18696
LC18514 F. luffae
LC18693
LC18451
LC18404
LC18794
LC18694
LC18687
LC18635
74/- LC18428
LC18504
Equiseti clade

LC18477
LC18407
10× LC18362
LC18544
DG1301 “F. incarnatum ” in Gai et al. (2016);
LC18644 causing maize stalk rot
LC18673
LC18545
Camptoceras 96/1
LC12167 T
clade NRRL32522
10×
70/- LC18443
0.05 LC18410
83/1 LC18683
LC18686
CQ1038

Fig. 9. Phylogeny inferred based on the combined CaM-rpb2-tef1 gene regions of the Fusarium incarnatum-equiseti species complex (FIESC). Fusarium
concolor (NRRL 13459) was used as an outgroup. Strains isolated in this study were indicated in red. Pathogenetic strains from previous studies were
indicated in green. The RAxML Bootstrap support values (ML-BS > 70 %) and Bayesian posterior probabilities (BI-PP > 0.9) were displayed at the nodes
(ML-BS / BI-PP). Ex-type, ex-epitype and ex-neotype strains were indicated in bold with T, ET, and NT, respectively.

scale phylogenetic tree was generated (Fig. 14). The Gblocks filtered not statistically supported in previous phylogenetic trees (Xia et al.
alignment of 1 001 single-copy orthologs (Supplementary Table S4) 2019, Crous et al. 2021, Wang et al. 2022a), indicating that previous
consisted of 223 451 characters, including alignment gaps. The multi-locus analyses could not provide sufficient phylogenetic
genome sizes of the four cereal-associated species-rich complexes resolution to reveal the evolutionary relationship between the two
(i.e., FFSC, FIESC, FOSC and FSAMSC) differed significantly (P species complexes. However, when the FCAMSC and the FIESC
< 2e-16, Supplementary Fig. S9), with FOSC having the largest were considered together as a whole, this larger group is clearly
genome size (av. ± SD: 52.6 ± 4.2 Mb) and FSAMSC having the a distinct evolutionary lineage that is highly supported in the
smallest genome size (av. ± SD: 37.5 ± 3.7 Mb). In our phylogenomic phylogenomic tree (support value = 100 %; Fig. 14), as well as in
tree, 193 nodes (86 %) received 100 % bootstrap support, including previous multi-locus trees (Crous et al. 2021). Therefore, we herein
the nodes F1, F2, and F3 (Fig. 14). Meanwhile, the gCF and sCF include the FCAMSC in FIESC as the Camptoceras clade (Fig. 14).
values resolved higher in node F3 (gCF 71.1 %, sCF 70.3 %) than Second, Bisifusarium (also known as F. dimerum species complex)
in F2 (gCF 46.5 %, sCF 46.7 %) and F1 (gCF 51.3 %, sCF 47.1 %). and Rectifusarium (also known as F. ventricosum species complex),
There are only two subtle topology differences between our appearing basal in the tree, were paraphyletic in our phylogenomic
phylogenomic tree (Fig. 14) and the multi-locus trees in previous tree (Fig. 14), in agreement with Lombard et al. (2015) and Crous et
studies (Xia et al. 2019, Crous et al. 2021, Wang et al. 2022a). First, al. (2021), but differed from O’Donnell et al. (2013) and Geiser et al.
in our phylogenomic tree, F. camptoceras (previously included in (2021) in which Bisifusarium and Rectifusarium clustered together
F. camptoceras species complex, FCAMSC) clustered within the forming a separate clade. However, in both publications, the clade
Equiseti clade of the FIESC (Fig. 14), while the FCAMSC formed a including Bisifusarium and Rectifusarium was not statistically
separate sister clade to the FIESC in previous phylogenetic studies supported (O’Donnell et al. 2013, Geiser et al. 2021). Overall, our
(Xia et al. 2019, Crous et al. 2021, Wang et al. 2022a). This is not phylogenomic tree is essentially similar with that of Crous et al.
surprising, as when considering the FIESC and the FCAMSC as two (2021), and our data resolved several remaining conflicts among
separate species complexes, the clade representing the FIESC was phylogenies inferred from different studies.

www.studiesinmycology.org 103
Han et al.

LC18595
LC18674
0.03 LC18631
LC18691
FIESC LC18682
CaM-rpb2-tef1 LC18716
LC18584
New Chinese record LC18567
LC18793
New host record (maize) LC18816
LC18513
New host record (rice) LC18648
LC18420
New host record (wheat) LC18629
LC18630
LC18742
LC18454 F. sulawesiense
NRRL43730
LC18510
LC18641
LC18638
LC18267
Incarnatum clade

LC18634
78/- LC18618
LC18685
LC18612
LC18792
-/0.99 LC18426
LC18442
LC18620
LC18614
LC18787
LC18692
-/0.99
LC18699

Incarnatum clade
LC18294
Equiseti clade

84/- LC18608
LC18576
10× LC18688
LC18396
LC18640
LC18539
InaCC F940 T
LC18400
87/1 Indo 167
Camptoceras Fusarium sp. (FIESC32)
10×
clade InaCC F964
LC18625
0.05 LC18295
LC18535
SDBR-CMU425 “F. melonis ” in
SDBR-CMU424 Khuna et al. (2022)
93/1 LC18585
LC18423 F. pernambucanum
96/1
LC18380
NRRL 34070
86/- LC18702
URM 7559 T
NRRL32181
NRRL34001
NRRL 31160
94/0.99
94/1 NRRL32182 F. mianyangense sp. nov.
LC15879 T
99/0.98 LC18549
83/-
100/1 SDBR-CMU423 F. citrullicola
98/1 SDBR-CMU422 T
LC12145
90/1 LC12146
LC7188 T F. irregulare
85/0.98
MRC 2804
98/1 NRRL 54142
NRRL 43639 T F. multiceps
100/1 NRRL 34003
MUM 1859 T F. caatingaense

Fig. 9. (Continued).

Phylogenomic analyses within the FSAMSC better understand the evolutionary relationship of species within
the FSAMSC, especially the Fg clade, a phylogenomic tree of
FSAMSC is one of the most species-rich complexes containing 41 the FSAMSC (Fig. 15) was inferred employing 5 139 single-copy
accepted species, 35 of which have available genome data. Species orthologs (Supplementary Table S5) obtained from 81 assembled
within the FSAMSC particularly the Fg clade, however, exhibit genomes, with Fusarium nelsonii NRRL 13338 as an outgroup (Fig.
almost indistinguishable morphological differences (Sarver et al. 15). Our phylogenomic tree is strongly supported by UFBoot values,
2011) and few nucleotide differences (Laraba et al. 2021), although but weakly supported by gCF and sCF values (Supplementary
they have been recognised as different species in previous studies Fig. S10B), suggesting strongly conflicting signals among genes
(O’Donnell et al. 2000a, 2004, Starkey et al. 2007). Therefore, to (Minh et al. 2020b). Meanwhile, the conflicting topologies were

104
Fusarium spp. on cereals in China

LC18441
75/0.99 LC18452
0.03 -/0.99
LC18675
FIESC LC18636
CaM-rpb2-tef1 LC18795
LC18506
New Chinese record LC18710
73/0.99 LC18534
LC18610
New host record (maize) 96/0.98 InaCC F965 T F. tanahbumbuense
100/1
LC18659
New host record (rice) LC18427
97/1 LC18438
New host record (wheat) Azerbaijan 53-3 “F. cf. incarnatum ” in Özer et al. (2020)
Azerbaijan 43 causing wheat crown and root rot
70/0.97 LC18805
100/1 LC12160 T F. guilinense
85/1
100/1 CBS 161.25 Ex-type of “F. bubalinum ”
LC18436 T in Xia et al. (2019)
NRRL 13379
100/1 LC18382
F. nothincarnatum sp. nov.
Incarnatum clade

97/1 CBS 132907

Incarnatum clade
-/0.95 NRRL 32866
100/1 ITEM 7155
F. incarnatum
CBS 132.73 NT
NRRL 54973 T F. monophialidicum
NRRL 20841 T F. coffeatum
LC18709
99/1 NRRL 22244
73/- LC18596
86/- LC18800 F. nanum
100/1 LC12168 T
89/1
LC18363
98/1 LC18681
Equiseti clade

98/1 ITEM 6748


CBS 143598 F. persicinum
10× 79/0.97
CBS 479.83 T
JS21 “F. incarnatum ” in Wang et al. (2021)
JS9 causing rice spikelet rot
96/1 JS3 F. hainanense
100/1
LC18701
LC11638 T
83/0.99 90/1 NRRL 26417
Camptoceras
clade
CBS 131385 T F. aberrans
10× CBS 119866
0.05 LC18243
LC7922
-/0.98
LC18333 T
CBS 130905 F. weifangense sp. nov.
99/1 LMSF-Id01
CBS 621.87
100/1 CPC 35143
83/-
LC18311
LC18317
98/1 93/1 LC6896 T F. citri
NRRL 52765
94/1 CQ1039 T
100/1 LC18763 F. humuli
LC18553
100/1 CBS 131383
F. fasciculatum
CBS 131382 T

Fig. 9. (Continued).

further confirmed by the comparison of phylogenomic and multi- followed by wheat (23 species) and rice (18 species), and only nine
locus phylogenetic trees (see Supplementary Fig. S11 for more species were shared by all three kinds of cereals (Fig. 16A, C).
details). This discordance is potentially due to incomplete lineage Species with high proportion of isolation in maize samples were
sorting (ILS), horizontal gene transfer (HGT), gene duplication F. verticillioides (24 %) and F. annulatum (17 %), compared to F.
and loss, hybridisation, or recombination (Degnan & Rosenberg asiaticum (30 %) and F. graminearum (22 %) in wheat samples and
2009). In general, both phylogenomic and multi-locus phylogenetic F. sulawesiense (17 %) in rice samples (Fig. 16B). The majority of
trees revealed five major clades (Graminearum, Sporotrichioides, Fusarium species collected in this study (98 % species from 90 %
Sambucinum, Brachygibbosum and Longipes clades), and both samples) were from regions affected by a monsoon climate (Fig.
provide effective resolution for species delimitation within the 16C, Supplementary Table S1).
FSAMSC. Interestingly, a total of 315 diseased samples were used for
fungal isolation. Of these, 176 samples (56 %) were observed
Distribution and pathogenicity of Fusarium spp. on to have a co-infection (Fig. 17), half of which contain previously
diseased cereals reported dominant pathogens, including F. annulatum (previously
usually known as “F. proliferatum”), F. asiaticum, F. graminearum, F.
Based on multi-locus phylogenetic analyses, the 608 representative pseudograminearum, and F. verticillioides (Fig. 17, Supplementary
Fusarium strains were identified as 43 species. Further analyses Table S6). In addition, co-infection by two species can be observed
showed that species richness is higher on maize (34 species), from 59 % of the samples (Supplementary Fig. S12).

www.studiesinmycology.org 105
Han et al.

0.03 LC18422
LC18341
LC18339
FIESC
LC18822
CaM-rpb2-tef1 LC18242
95/1
LC18664
New Chinese record LC18759
LC18619
New host record (maize) LC18639 F. ipomoeae
-/0.95
LC18447
New host record (rice) LC0166
74/1 LC7923
New host record (wheat) mmpp-1 “F. equiseti ” in Li et al. (2014)
100/1
LC18509 causing corn sheath rot
LC18453
98/1
81/1 LC18697
LC18474
80/0.98 98/1 LC18409
LC18378
LC12165 T
LC18561
Incarnatum clade

99/1
NRRL 36448 T
NRRL36401 F. duofalcatisporum
89/0.98 100/1 NRRL 36323 ET
F. compactum
LC18530
LC18379
LC15878 T
89/0.99 -/0.99 LPPC072
99/1

Equiseti clade
LPPC079 F. jinanense sp. nov.
-/0.96 LPPC078
LPPC076 “Fusarium sp.”
98/1 LPPC056 in Lima et al. (2020)
74/0.98
LPPC077
Equiseti clade

98/1 FRCR10113 “FIESC 31a” in Short et al. (2011)


IMI 300797 T
10× MRC 2609
NRRL52753 F. lacertarum
94/1 2Ar097
78/1 NRRL 36123
Azerbaijan 63-6
Azerbaijan 57-5 “F. equiseti ” in Özer et al. (2020)
Azerbaijan 61-5 causing wheat crown and root rot
Camptoceras 80/1 Azerbaijan 75-5
clade
10× Azerbaijan 26-8 F. clavus
Azerbaijan 33-5
0.05 Azerbaijan 16-6
100/1 F16 “FIESC ” in Okello et al. (2019);
CBS 126202 T causing corn root rot
10× LC18293
-/1 LC18605
100/1 LC6026
100/1
LC18440 F. arcuatisporum
71/0.96
LC12147 T
97/1
NRRL 25477 T F. longicaudatum
NRRL 43638 T F. brevicaudatum
-/0.97
CBS 119880 T F. serpentinum
80/1 NRRL 26922 T F. neoscirpi
-/0.95 NRRL 36478 NT F. scirpi
CBS 148244 T F. wereldwijsianum
ATCC 11853 T F. cateniforme
NRRL 36372 F. longifundum
97/1 CBS 307.94 ET
100/1 CPC35134 F. equiseti
NRRL 25796 T
100/1 NRRL43636
F. toxicum
90/1
NRRL 43635 T F. gracilipes
NRRL 36269 T F. flagelliforme
CBS 131777 T
100/1 CBS 131788 F. croceum
CPC 35240
100/1 Indo 175
F. mucidum
CBS 102395 T
Camptoceras

100/1 CBS 190.60


93/1 CBS 189.60 T F. neosemitectum
clade

100/1 IMI 112500 ET F. camptoceras


96/1 InaCC F963 T F. kotabaruense
98/- LC18372
100/1 LC18376
F. fecundum sp. nov.
5× LC15875 T
NRRL 13459 F. concolor Outgroup

Fig. 9. (Continued).

106
Fusarium spp. on cereals in China

LC2823
0.02 100/1
100/1 LC2819 F. paranisikadoi
FNSC
rpb1-rpb2-tef1 88/1 LC2800 T
98/1 CBS 577.97 T F. miscanthi

100/1
NRRL25179
83/1 F. nisikadoi
CBS 456.97 T
CBS 116011 T F. gaditjirri
78/0.99 96/1
CBS 125536 T F. lyarnte
5× LC18609
-/1
LC18652
-/0.97 LC18583
LC18507
F. commune
LC18568
-/0.99
NRRL 28387
CBS 110090 T
LC18486

NRRL 13994 T F. concolor Outgroup

Fig. 10. Phylogeny inferred based on the combined rpb1-rpb2-tef1 gene regions of the Fusarium nisikadoi species complex (FNSC). Fusarium concolor
(NRRL 13994) was used as an outgroup. Strains isolated in this study were indicated in red. The RAxML Bootstrap support values (ML-BS > 70 %) and
Bayesian posterior probabilities (BI-PP > 0.9) were displayed at the nodes (ML-BS / BI-PP). Ex-type strains were indicated in bold with T.

0.004 CBS 130308


FOSC 85/1 LC18297 F. cugenangense
CaM-rpb2-tef1 72/1
CBS 131393
New host
record (maize) 72/1 CBS 130304

99/1
CBS 255.52
CBS 217.49 F. elaeidis
99/1
CBS 218.49
CBS 116612
CBS 116613 T F. gossypinum
97/1
CBS 116611

-/1
-/1 CBS 744.79
LC18485
70/- LC18832 F. nirenbergiae
79/1
CBS 127.81
CBS 840.88 T
95/1
CBS 238.94 T
87/0.9 CBS 247.61 F. curvatum
4× CBS 141.95
CBS 177.31 F. udum Outgroup

Fig. 11. Phylogeny inferred based on the combined CaM-rpb2-tef1 gene regions of the Fusarium oxysporum species complex (FOSC). Fusarium udum (CBS
177.31) was used as an outgroup. Strains isolated in this study were indicated in red. The RAxML Bootstrap support values (ML-BS > 70 %) and Bayesian
posterior probabilities (BI-PP > 0.9) were displayed at the nodes (ML-BS / BI-PP). Ex-type strains were indicated in bold with T.

www.studiesinmycology.org 107
Han et al.

LC18262
LC18291
0.02 LC18572
LC18637
LC18401
FSAMSC LC18305
H3-rpb1-rpb2-tef1 LC18287
LC18251
New Chinese record LC18299
LC18336
LC18826
LC18594
LC18266
LC18325

Graminearum clade
LC18256
LC18250
LC18825
-/0.99 LC18259
LC18300
0.02 LC18308
LC18324
LC18289
LC18265
Sporotrichioides clade

NRRL13818 T
LC18739
2× LC18323
LC18403
LC18690
LC18283
Sambucinum clade
LC18342
4× LC18369
Brachygibbosum clade
LC18329
Longipes clade LC18344
LC18258
LC18338
LC18628 F. asiaticum
LC18650
LC18327
LC18290
LC18288
LC18316

Graminearum clade
LC18678
LC18332
LC18393
LC18271
LC18280
LC18285
LC18302
LC18286
LC18254
LC18397
LC18617
LC18282
LC18253
LC18278
NRRL28720
84/0.93 99/1 LC18274
LC18273
LC18303
LC18582
LC18505
LC18334
LC18257
-/0.98 LC18565
LC18264
LC18343
NRRL26156
NRRL6101
LC18321
LC18345
100/1 LC18660
LC18326
99/1 CBS 123752 T
NRRL45665 F. ussurianum
85/1 -/0.95 NRRL38207
93/0.99 NRRL37605 T
LC18758 F. vorosii
100/1 LC15880
NRRL26755
NRRL26754 T F. acacia-mearnsii
-/0.98 99/1 NRRL54222 T
-/0.99 NRRL54221 F. nepalense
85/1 NRRL46738 T
NRRL46726 F. aethiopicum
NRRL46718
LC18704
NRRL28436
LC18383
LC18542
97/0.97 LC18384
LC18773 F. meridionale
LC18774
NRRL29010
NRRL28723
NRRL28721

Fig. 12. Phylogeny inferred based on the combined H3-rpb1-rpb2-tef1 gene regions of the Fusarium sambucinum species complex (FSAMSC). Fusarium
nelsonii (NRRL 13338) was used as an outgroup. Strains isolated in this study were indicated in red. The RaxML Bootstrap support values (ML-BS > 70 %)
and Bayesian posterior probabilities (BI-PP > 0.9) were displayed at the nodes (ML-BS / BI-PP). Ex-type and ex-epitype strains were indicated in bold with
T and ET, respectively.

108
Fusarium spp. on cereals in China

NRRL28585
NRRL 2903 T F. austroamericanum
0.02 NRRL31238
99/0.99
NRRL31281 T F. brasilicum
FSAMSC NRRL29297 T
NRRL29306 F. cortaderiae
H3-rpb1-rpb2-tef1
NRRL36905 T F. gerlachii
New Chinese record CBS 415.86 T F. mesoamericanum
LC18770
LC18322
LC18364
LC18784
LC18772

Graminearum clade
LC18764
LC18748
LC18284
LC18309
LC18721
0.02
LC18768
LC18346
LC18777
LC18724
LC18476
Sporotrichioides clade

LC18449
LC18263

LC18260
LC18245
LC18679
LC18353
Sambucinum clade LC18314

LC18771
LC18315
Brachygibbosum clade
LC18750
Longipes clade LC18298
90/1 LC18720
LC18708
LC18775
LC18387
NRRL6394
LC18663
NRRL28336

Graminearum clade
NRRL5883
NRRL13383
LC18247
LC18255
LC18246
LC18368
LC18543
LC18335
LC18767
LC18666
LC18766
LC18307
LC18381
LC18765 F. graminearum
LC18747
LC18292
LC18395
NRRL31084
LC18301
LC18705
LC18665
-/1
LC18643
LC18529
LC18672
LC18670
LC18252
LC18435
LC18621
LC18516
LC18668
LC18412
LC18371
LC18359
LC18552
LC18752
LC18373
LC18667
LC18676
LC18385
LC18669
LC18347
78/- LC18310
LC18546
LC18796
LC18769
LC18275
NRRL28063
LC18281
LC18279
88/1 LC18711
LC18761
88/1 LC18244
CBS 136009 T
88/0.97 LC18328
LC18337
LC18330
LC18304
LC18593
LC18272

Fig. 12. (Continued).

www.studiesinmycology.org 109
Han et al.

0.02 100/1 LC18779


FSAMSC LC18778
H3-rpb1-rpb2-tef1 LC18515
LC18749
New Chinese record LC18719
LC18386
LC18703
LC18723
LC18782
LC18715

Graminearum clade
LC18415 F. boothii
100/1 LC18520
NRRL26916 T
88/- LC18733
LC18722
0.02 LC18732
NRRL29020
NRRL29105

Graminearum clade
100/1 NRRL29011
Sporotrichioides clade

NRRL54196 F. louisianense
100/1 CBS 127525 T
NRRL13721
2× 100/0.99 NRRL25491 F. cerealis
NRRL25475 T F. culmorum
NRRL13393 T F. lunulosporum
Sambucinum clade
LC18354
LC18551
4× LC18268
Brachygibbosum clade 100/1 LC18365
Longipes clade 100/0.99 LC18360
LC18318
LC18358
LC18550
LC18276
100/0.96 LC18320
LC18554 F. pseudograminearum
LC18374
LC18375
100/1 LC18671
LC18352
LC18361
100/1 LC18357
NRRL28062 T
100/1 LC18370
LC18377
100/1 LC18366
NRRL29298 T
NRRL29380 F. dactylidis
100/1 NRRL66764 T F. subtropicale
NRRL39664 F. praegraminearum
100/1 NRRL29977
99/1 100/1 100/1 NRRL3299 T F. sporotrichioides
NRRL53431 F. sibiricum
98/1 NRRL53430 T
NRRL53417

Sporotrichioides clade
100/1 CBS113234 T F. langsethiae
100/1 NRRL53409
96/1
NRRL66249
100/1 NRRL66248
F. goolgardi
100/0.98 NRRL66250 T
100/1 CBS 201 63 T
NRRL36351 F. nodosum
100/1 NRRL54056 T
NRRL54058 F. palustre
100/1
100/0.95 NRRL31970
LC18306
100/1 91/1 NRRL26908 T F. armeniacum
99/1 LC18662
100/1 LC15881
LC18661
81/1 NRRL 66748
NRRL 66749 T F. chaquense
LC18351
NRRL 26941 ET
Sambucinum clade

NRRL66297
100/0.99 NRRL13714
LC18712 F. poae
90/0.99
LC18570
-/0.97 75/0.97
LC18718
97/1 100/1 NRRL13392 T F. robustum
100/1 NRRL22187 F. sambucinum
100/1 NRRL22196
100/1 CBS 458 93 T F. venenatum
100/0.96 NRRL25349
100/1 NRRL 3509 T F. kyushuense
LC18277
-/1 NRRL28507 T F. musarum
92/0.99 NRRL13829
100/1 NRRL34033 F. brachygibbosum
98/1 NRRL 20954 T
100/1 MRC 2486 F. subflagellisporum Brachygibbosum
CBS 144211 T clade
100/1 CBS 144217 F. transvaalense
CBS 144212
NRRL 20695 ET F. longipes Longipes clade
NRRL13338 F. nelsonii Outgroup

Fig. 12. (Continued).

110
Fusarium spp. on cereals in China

96/- LC18269
0.03 LC18319
92/1
FTSC LC18556
ITS-rpb1-rpb2-tef1 LC18557
LC18555
89/1
LC18558
LC18270
100/1
LC18313 F. avenaceum
91/1 LC18349
99/1
LC18560
85/1 LC18559
96/0.98
NRRL 25128
92/1 NRRL 54939
LC18592
82/1 CBS 408.86 NT
LC13816
75/-
98/1 LC13817 T F. paeoniae
88/1
LC5166
74/1 LC6037
100/1 LC6043 F. alpinum
LC6045 T
LC4957 T
100/1 LC13813 F. chongqingense
LC13814
LC18348
88/0.96
-/1 LC18331
71/1 LC18312
-/1
LC18680
NRRL 52789
F. acuminatum
LC18744
99/1
75/1 NRRL 36147
LC18780
98/1 LC18350
LC18356
5× -/1
84/1 CBS 253.50
F. tricinctum
NRRL 25481 ET
CBS 122710 T F. sinense
95/1 LC1112
CBS 143608 T F. iranicum

NRRL 13459 F. concolor Outgroup

Fig. 13. Phylogeny inferred based on the combined ITS-rpb1-rpb2-tef1 gene regions of the Fusarium tricinctum species complex (FTSC). Fusarium concolor
(NRRL 13459) was used as an outgroup. Strains isolated in this study were indicated in red. The RAxML Bootstrap support values (ML-BS > 70 %) and
Bayesian posterior probabilities (BI-PP > 0.9) were displayed at the nodes (ML-BS / BI-PP). Ex-type, ex-epitype, and ex-neotype strains were indicated in
bold with T, ET, and NT, respectively.

Furthermore, we summarised the world records of Fusarium species for the world and China were updated, respectively (Table
species associated with cereals (Table 2). The data were retrieved 2). A total of 62 species have been recorded as cereal pathogens
from the USDA fungal database (more than 1 600 records) (Farr (Table 2), of which 18 species (F. acuminatum, F. annulatum, F.
& Rossman 2022), previous studies (168 peer-reviewed papers, arcuatisporum, F. armeniacum, F. awaxy, F. concentricum, F.
Table 2) and this study (2 020 strains, Supplementary Table erosum, F. ipomoeae, F. jinanense, F. luffae, F. mianyangense,
S1). According to the current taxonomy updated by Crous et al. F. nirenbergiae, F. pernambucanum, F. planum, F. sacchari, F.
(2021) and this study, these records correspond to an estimated sanyaense, F. sulawesiense and F. tanahbumbuense) were
97 Fusarium species, belonging to 13 species complexes and proven as pathogens of maize stalk rot for the first time in this
one singleton. Among them, 39 and 52 host records for Fusarium study (Fig. 7). Notably, the DSI varies among seedlings infected

www.studiesinmycology.org 111
Han et al.

100/30.9/90.5 F. verticillioides 7600


0.04 100/35.8/93.1 F. verticillioides LC18525
100/12.3/57.1
Phylogenomic tree 100/9.9/52.6
F. musae NRRL 25059
F. coicis NRRL 66233 T
228 assembled genomes 100/14.3/53.4 100/4.9/39.7 F. planum LC15876 T
100/13.7/72.2 F. ramigenum NRRL 25208 T
1001 single-copy orthologs F. tjaetaba NRRL 66243 T
100/27.6/91.3 100/3.4/45
100/11.9/60.5 F. napiforme NRRL 25196
0.04 FFSC F. brevicatenulatum NRRL 25447
FOSC
100/2.1/40.2 F. pseudoanthophilum NRRL 25211 T
FNSC
100/11.3/44.6
F. pseudonygamai NRRL 13592 T
100/4.4/35.9
FnewSC

100/5.1/49.1
F. thapsinum NRRL 22049
F. xyrophilum NRRL 62721
FRSC

FburSC 100/12.5/59.6 F. pseudocircinatum NRRL 36939


FCOSC

FFBSC
100/5.6/42.6 100/7.8/44.3 F. denticulatum NRRL 25311 T
FSAMSC -/15.3/37.8 F. nygamai MRC8546
FCSC
100/26.4/83.5 F. mundagurra NRRL 66235 T
FASC F. phyllophilum NRRL 13617 T
FIESC 100/4.8/44.9 F. xylarioides NRRL 25486 ET
FTSC 100/13.7/63.5 F. udum NRRL 25194
F. nurragi (singleton)
F. acutatum NRRL 13308
FHSC
F. secorum NRRL 62593 T
F3
FtorSC
-/8/34.7 F. subglutinans NRRL 66333
FLSC 100/23.9/91.6 -/5/48.6 F. subglutinans LC18249
FBSC 100/20.7/93.2 F. awaxy LC18783
Cyanonectria

Neocosmospora
100/13.9/74.8 Fusarium sp. RC 528
100/6.2/49.5 F. temperatum LC18813
Albonectria 100/12.3/69.9
Setofusarium
100/4.5/34.5
F. temperatum CMWF389
F2 Geejayessia
100/5.2/47.4
F. bactridioides NRRL 66639
Nothofusarium
-/1.8/39.3 100/3.7/42.5 F. bulbicola NRRL 25176
Luteonectria F. succisae NRRL 13298
F1 Bisifusarium 100/3.3/51.7 F. circinatum NRRL 25331
Rectifusarium 100/16.9/85.8 F. anthophilum NRRL 25214
Neonectria
98/8.4/36.9 F. pininemorale CMW 25243 T
Corinectria
100/6.1/54.8 100/26.6/79.1 F. sororula FCC 5425
Stylonectria
100/6.2/55.2 F. guttiforme NRRL 22945 FFSC
100/3.8/43.2 F. fracticaudum CBS 137234
F. begoniae NRRL 25300 T
100/28/76.6 F. sterilihyphosum NRRL 25623 T
100/5.1/55.2 F. tupiense NRRL 53984
F. mexicanum NRRL 53147 T
F. agapanthi NRRL 31653
100/8.19/36.4 98/7.6/47.3 F. annulatum NRRL 62905
100/10/65.6 100/7.9/71.5 F. annulatum LC18417
100/14.9/87.8
F. annulatum ITEM 2341

Fusarium
100/5.4/46.6
F. annulatum ET1
100/16.1/76.8 F. annulatum F8_4S_2B
100/5.6/42 F. siculi KOD 1856
100/8.2/76.8 F. erosum LC15877 T
100/5.5/91.2 F. globosum NRRL 26131 T
100/7.6/68.9 100/15/76.9 Fusarium sp. FGSC 8932
Fusarium sp. Augusto2
100/24.5/96.9
100/19.3/53 Fusarium sp. MRC2276
100/20.8/75.8 Fusarium sp. IMI 58289
98/7.6/46.1 F. fujikuroi LC18819
100/24.3/82.8 F. elaeagni LC18815
100/22.4/62.8 F. sanyaense LC15882 T
F. mangiferae NRRL 25226
100/13/60.5 F. concentricum NRRL 25181
100/35.5/99.2 F. concentricum LC18523
100/52.9/97.9 F. sacchari LC18522
100/10.3/94.7 F. sacchari NRRL 66326
-/1.2/32.6 F. dlaminii NRRL 13164 T
-/1.1/41.2 F. cugenangense NRRL26413
100/2.8/70.3 F. cugenangense LC18297
-/6.5/36.5 F. elaeidis RBG6448
100/4.4/84.1
F. duoseptatum Fol-114
-/0.3/50.4 F. glycines Fom009
100/5.7/62.9 -/0.5/30.8 F. gossypinum Foc021
F. fabacearum Fom004
-/0.4/39.4
100/32.6/60.2 F. libertatis RBG5714
F. hoodiae Fo5
100/4.4/61.9 F. trachichlamydosporum 160527
100/17.6/91.3 Fusarium sp. EthFoc-75
F. oxysporum G76
-/1/36.7 F. veterinarium MRL8996 FOSC
-/1.8/30 F. pharetrum Fo1
100/12.4/91.4 100/2.6/52.6 F. languescens KOD888
100/3.6/49.5 F. nirenbergiae LC18485
F. nirenbergiae N139
F. curvatum PHW726
100/8.39/64.3 F. triseptatum 14-004
100/43.2/68.3
100/19.2/86.5 Fusarium sp. Fo24
F. odoratissimum NRRL 54006
-/2.5/39.5 F. foetens NRRL 38302
100/29.1/65.3 100/56.9/94 F. nisikadoi NRRL 25179
100/39.8/81.2 F. miscanthi NRRL 26231
-/4.9/25.9 F. lyarnte NRRL 54252 FNSC
100/47.9/62.2 -/4.9/49.8
F. gaditjirrii NRRL 45417
100/44.2/99.2 F. commune NRRL 28387
F. commune LC18583
100/27.1/61.2 F. newnesense NRRL 66241 FnewSC
100/40.2/82.3 F. redolens NRRL 22901
2× F. spartum NRRL 66896 FRSC
100/40.6/53.6 F. hostae NRRL 29888
100/64.2/88.8 F. algeriense NRRL 66647
100/52.2/65.8 F. burgessii NRRL 66654
F. beomiforme NRRL 25174
FburSC
100/59.4/76.7 F. concolor NRRL 13459
100/36.6/51.2
100/43/55.7 F. austroafricanum NRRL 53441 FCOSC
F. anguioides NRRL 25385
F. falsibabinda NRRL 25539 FFBSC

Fig. 14. Maximum likelihood phylogenomic tree of Fusarium and allied genera. A total of 1 001 single-copy orthologs were employed in the analysis.
Stylonectria norvegica IHI 201603 was used as an outgroup. Strains sequenced in this study were indicated in red. The IQ-TREE ultrafast bootstrap support
values (UFBoot ≥ 95 %), gCF and sCF values were displayed at the nodes (UFBoot / gCF / sCF). Arrows “F1” (= “Terminal Fusarium clade”), “F2” and “F3”
indicate the three alternative Fusarium generic hypotheses sensu Geiser et al. (2013). Ex-type, ex-epitype and ex-neotype strains were indicated in bold with
T, ET, and NT, respectively. Subdivision of the Fusarium clade represents the recognised species complexes, including F. aywerte SC (FASC), F. buharicum
SC (FBSC), F. burgessii SC (FburSC), F. chlamydosporum SC (FCSC), F. concolor SC (FCOSC), F. falsibabinda SC (FFBSC), F. fujikuroi SC (FFSC), F.
heterosporum SC (FHSC), F. incarnatum-equiseti SC (FIESC), F. lateritium SC (FLSC), F. newnesense SC (FnewSC), F. nisikadoi SC (FNSC), F. oxysporum
SC (FOSC), F. redolens SC (FRSC), F. sambucinum SC (FSAMSC), F. torreyae SC (FtorSC) and F. tricinctum SC (FTSC).

112
Fusarium spp. on cereals in China

100/25.8/97.7 F. asiaticum NRRL 28720


0.04
100/2.2/48.6 F. asiaticum LC18397
Phylogenomic tree -/0.6/25.4 F. ussurianum CBS 123752 T
228 assembled genomes -/0.6/39.4 F. nepalense CBS 127943
1001 single-copy orthologs 100/23.1/97.4 F. vorosii LC15880
-/0.3/38
F. vorosii CBS 119177
0.04 FFSC
FOSC
99/1.1/44.3 F. acaciae-mearnsii NRRL 26754 T
FNSC F. aethiopicum CBS 122858 T
100/5.9/47.9 -/0.5/32.6 F. austroamericanum NRRL 2903 T
FnewSC
FRSC
100/3.7/59.3
FburSC F. cortaderiae CBS 119183 T
100/10.4/83.8
F. brasilicum NRRL 31281 T
FCOSC

FFBSC
100/20/98.4
FSAMSC F. meridionale NRRL28721
F. meridionale LC18774
FCSC

FASC
100/19.7/97.7 100/13.9/75.2
FIESC
100/7.9/77.9 F. graminearum LC18796
F. graminearum PH-1
FTSC
F. nurragi (singleton)
100/2.3/40.3
FHSC
F. gerlachii CBS 119176
FtorSC
F3
FLSC 100/22.8/71.6 100/1.1/55.4 F. louisianense CBS 127525 T
FBSC
Cyanonectria 100/27.3/99.3
F. boothii CBS 316.73 T
Neocosmospora F. boothii LC18723
Albonectria
Setofusarium 100/11.2/54.4 100/31.6/68.6 F. mesoamericanum CBS 415.86 T
F2 Geejayessia F. cerealis S18/34
100/52.3/99.2
Nothofusarium

Luteonectria
F. culmorum NRRL 25475 T
F1
100/57.5/80.4 F. pseudograminearum LC18671
Bisifusarium
Rectifusarium

Neonectria
F. pseudograminearum RBG5266
Corinectria F. praegraminearum NRRL 39664
Stylonectria
60/15.1/29.6 100/62.1/94.9 F. subtropicale NRRL 66764
100/22.3/38.8 100/52.1/83.4 F. venenatum NRRL 66329
F. sambucinum NRRL 13708 FSAMSC
100/72/98.2
F. kyushuense LC18277
100/44/68.3 F. kyushuense NRRL 25348
100/31/41 F. poae NRRL 26941 ET
100/71.2/99 F. poae LC18712
100/52.5/69.7 F. sporotrichioides NRRL 3299 T
100/35/86.8 F. sibiricum NRRL 53430 T
100/22.9/70.7 F. langsethiae Fl201059
100/37.9/72.2 100/22.5/67 F. nodosum NRRL 36351
100/31.1/49.4
F. goolgardi NRRL 66250 T
100/40.4/92.3 F. armeniacum NRRL 25141 T
100/51.5/92.8 F. armeniacum LC15881
100/34.4/61.2
F. chaquense NRRL 66749 T

Fusarium
100/20.2/38
F. palustre NRRL 54050
100/24.3/47.7 100/17/35.5
F. transvaalense NRRL 31008
F. brachygibbosum HN-1
F. longipes NRRL 20695 T
100/69.2/89 F. nelsonii NRRL 13338
100/31.5/53.7
F. atrovinosum NRRL 13444 FCSC
F. aywerte NRRL 25410 FASC
100/18/81.7 F. mianyangense LC15879 T
-/6.3/36.9 Fusarium sp. NRRL 31160
100/11.2/70.7 F. sulawesiense LC18688
100/24.5/94.5 F. luffae NRRL 66473
100/62.4/69 100/6.4/53.7 F. luffae LC18508
F. caatingaense NRRL 66470
4× 100/2.5/42.2 F. tanahbumbuense NRRL 66471
100/13.6/67.5
98/2.4/35.5 F. tanahbumbuense LC18534
-/5.4/31.2 F. guilinense LC18805
Incarnatum

100/3.5/50.5
F. incarnatum ITEM 7155
99/2.8/35.6
100/31.9/38.7 F. nothincarnatum LC18436 T
100/30/94.1
100/13.6/59.8 F. nanum LC18709
100/12.2/67.8 F. nanum NRRL 66324
F. hainanense LC18701 FIESC
100/9/43.7
100/19.8/59.5
F. hainanense NRRL 66475
F. coffeatum NRRL 66322
100/45.7/97.6
Fusarium sp. NRRL 66334
100/22.4/65.5
F. weifangense LC18333 T
100/32.7/86.5 F. humuli LC18553
F. clavus NRRL 66337
100/68.4/89
100/13.3/47.7
F. clavus LC18293
Equiseti

100/20.8/36.9 100/25.2/67.1 F. ipomoeae LC18759


85/5.2/35.6 F. jinanense LC15878 T
100/16.1/59.2
100/9.2/49.9 F. flagelliforme NRRL 13405
F. scirpi NRRL 66328
100/9.7/40.2
F. equiseti NRRL 66338
100/33/38.7
F. camptoceras NRRL 13381
Camptoceras
100/29.6/66.2 F. fecundum LC15875 T

Fig. 14. (Continued).

www.studiesinmycology.org 113
Han et al.

0.04 100/54.6/97 F. acuminatum F829


Phylogenomic tree
100/35.5/64.6 F. acuminatum LC18312
97/29.9/41.6
100/54.3/71.6 F. tricinctum NRRL 25481
228 assembled genomes 100/54.7/84.7
F. avenaceum LC18558
100/68.3/70.2
100/63.4/77.8
F. avenaceum Fa05001 FTSC
1001 single-copy orthologs
F. torulosum NRRL 22747
F. nurragi NRRL 36452 Not assigned

Fusarium
0.04 FFSC
100/71.1/70.3
FOSC
F3 F. graminum NRRL 20692
FHSC
FNSC
FnewSC
100/35/40.1
100/83.7/91.3 F. heterosporum NRRL 20693
F. continuum NRRL 66286
FRSC

FburSC
100/62.2/79.5
FCOSC F. zanthoxyli NRRL 66285 FtorSC
FFBSC 100/30.9/42.2 F. torreyae NRRL 54149
100/15.7/31.7
FSAMSC
100/63.8/78.2 F. lateritium NRRL 13622
FLSC
FCSC

FASC F. sarcochroum NRRL 20472


FIESC
F. stilboides NRRL 20429
100/52.5/56.1
FTSC
F. buharicum NRRL 13371
100/72.8/79.1
F. nurragi (singleton)

FHSC F. sublunatum NRRL 13384 FBSC


FtorSC C. cyanostoma NRRL 53998
F3
FLSC
C. buxi NRRL 36148 Cyanonectria
FBSC
Cyanonectria 100/22.5/53 N. floridana NRRL62606
Neocosmospora 100/15.6/52 N. kuroshio UCR3666
Albonectria 100/16.6/48.1 N. ambrosia NRRL 20438 ET
Setofusarium
100/45.4/81.2
F2 Geejayessia 100/10.5/34 N. tuaranensis NRRL 46518
100/41.4/66.4
Nothofusarium
N. obliquiseptata NRRL 62610
F1
Luteonectria

Bisifusarium
100/31.2/58 Neocomospora sp. NRRL 66088
100/24.6/41.1 N. pisi 77-13-4
Neocomospora
Rectifusarium

Neonectria 100/49.6/72.7 N. mori JS-169


100/28.6/47.8
N. rectiphora NRRL 22396
Corinectria

Stylonectria 100/24.2/47.4
-/14.6/31.2 N. metavorans FSSC_6
100/30.7/63.5 N. haematococca LQ1
100/15.9/35.6 100/52/62.7 N. falciformis NRRL 43529 T
100/72/75.6 N. vasinfecta NRRL 22166 ET
100/10.5/41.9 N. protoensiformis NRRL 22178
100/33.9/50 N. illudens NRRL 22090
100/88.3/96.2 A. rigidiuscula NRRL 13412
100/46.5/46.7 A. albosuccinea NRRL 20459 Albonectria
F2
S. setosum NRRL 36526 Setofusarium
Terminal G. atrofusca NRRL 22316
100/80.9/74.1
G. zealandicum NRRL 22465
Geejayessia
Fusarium 100/31.2/44.7
clade
100/26.7/38.5 No. devonianum NRRL 22134 Nothofusarium
100/82.5/84.6 L. nematophila NRRL 54600
L. albida NRRL_22152 Luteonectria
=

100/51.3/47.1 B. penzigii NRRL 20711


F1 100/89.3/87.3
100/86.9/73.4 B. dimerum NRRL 20691 Bisifusarium
2× B. domesticum NRRL 29976
R. robinianum NRRL 25729 Rectifusarium
100/70.7/64.6 Ne. coccinea CBS 119158
100/85.2/75.7 Ne. punicea CBS 119724 Neonectria
100/55.4/45.5 Ne. hederae CBS 714.97
Co. fuckeliana CBS 125109 Corinectria
Stylonectria norvegica IHI 201603 Outgroup
Fig. 14. (Continued).

by different Fusarium species in this study (Fig. 7). For instance, study, 227 strains clustered into three major clades (African clade,
the DSI of seedlings infected by F. annulatum was 68.75 % even American clade, and Asian clade) and 12 subclades, representing
under low conidium concentrations (1 × 106 conidia/mL) (Fig. 7E), nine known and three novel species (Fig. 8).
while those infected by F. awaxy was 6.25 % under low conidium
concentrations (1 × 106 conidia/mL), but appeared 59.38 % under Fusarium annulatum Bugnic., Rev. Gén. Bot. 59: 13. 1952.
higher concentrations (1 × 107 conidia/mL) (Fig. 7H).
Materials examined: (See Supplementary Table S1).
Taxonomy
Notes: Fusarium annulatum was proposed by Bugnicourt (1952),
Fusarium fujikuroi species complex (FFSC) with CBS 258.54 as ex-type strain. Later, a phylogenetic analysis
based on LSU-SSU-tub2 suggested that CBS 258.54 clustered with
As the largest and best-studied species complex in the genus the representative strains (CBS 217.76, NRRL 25089, etc.) of F.
Fusarium, the FFSC comprises 84 described species separated proliferatum (O’Donnell et al. 1998a). Given that F. annulatum was
into five major clades (African clade, African clade B, African clade rarely reported and has only one collection (CBS 258.54), whereas
C, American clade, and the Asian clade) (Yilmaz et al. 2021, Wang F. proliferatum represented by CBS 217.76 has been widely used,
et al. 2022a). This species complex includes many important plant O’Donnell et al. (1998a) recommended treating F. annulatum as
pathogens that cause various diseases and produce numerous synonym of F. proliferatum, but did not formalise the proposal.
mycotoxins (Choi et al. 2018, Duan et al. 2020, Qiu et al. 2020, Recently, a multi-locus phylogenetic analyses including the epitype
Yilmaz et al. 2021). The FFSC can be distinguished from other of F. proliferatum (CBS 480.96) and additional sequences of CBS
complexes by its typical Fusarium macroconidia, abundant 258.54 (CaM, rpb1, rpb2, tef1) demonstrated that CBS 217.76
microconidia and rare chlamydospore formation. In the present clustered together with CBS 258.54, but clustered distant from

114
Fusarium spp. on cereals in China

0.01 100/13.4/91 F. asiaticum CBS 110258


99/6.58/39.4 F. asiaticum NRRL 26156
100/11.3/60.6 F. asiaticum NRRL 6101 F. asiaticum
Phylogenomic tree of FSAMSC 100/40/95.1 F. asiaticum LC18397
81 assembled genomes F. asiaticum NRRL 28720
5139 single copy orthologous genes 100/5.12/44.1
100/13.1/70.4 F. ussurianum CBS 123752 T
100/44.3/97.1 F. ussurianum 29813 F. ussurianum
100/1.54/40.4 F. ussurianum 65202
100/16.6/56.2 F. vorosii CBS 119177
100/38.6/94.9
100/1.87/45.9
F. vorosii CBS 119178 F. vorosii
F. vorosii LC15880
100/49.4/97.3
F. nepalense CBS 127669
F. nepalense
100/2.8/48.1 F. nepalense CBS 127943
100/16.7/44.1 F. acaciae-mearnsii CBS 110253
100/42.1/97.4
F. acaciae-mearnsii CBS 110255 F. acaciae-mearnsii
100/3.41/40.5 F. acaciae-mearnsii NRRL 26754 T
F. aethiopicum CBS 122858 T F. aethiopicum
100/15.9/57.1
100/31.4/91.4
F. graminearum LC18796
F. graminearum PH-1 F. graminearum
100/16.6/80.9
F. graminearum CBS 138562
100/37.1/94.9
F. gerlachii CBS 119175
100/3.79/47 100/31.8/79.4 F. gerlachii
F. gerlachii CBS 119176

Graminearum clade
100/49.5/97.5
F. louisianense CBS 127524
100/9.3/35.5 100/2.41/46.5
F. louisianense CBS 127525 T F. louisianense
100/13.4/55.6
F. boothii CBS 119170
100/48/98.3 95/1.5/35.1 F. boothii CBS 316.73 T F. boothii
100/11.4/62.8 F. boothii CBS 110251
100/1.56/38.6
F. boothii LC18723
F. mesoamericanum CBS 415.86 T F. mesoamericanum
100/28.8/90.5
100/11/33.6 F. meridionale CBS 110249
F. meridionale LC18774
100/21.3/83.4 100/44.9/74.8 F. meridionale NRRL28721 F. meridionale
100/29.2/91.5 F. meridionale 38FSP
100/6.38/56.3 F. brasilicum NRRL 31281 T
100/25.9/93.9 F. brasilicum
F. brasilicum CBS 119179
100/9.92/87.5
100/6.46/48.1
F. cortaderiae CBS 123655
F. cortaderiae CBS 119183 F. cortaderiae
100/24.2/90.7 F. cortaderiae NRRL 29297 T
100/53.4/69.5
F. austroamericanum NRRL 2903 T
100/59.2/98.3 F. austroamericanum CBS 110246 F. austroamericanum
100/14.2/53.3 F. austroamericanum 3FSP
100/23.5/55.8
F. culmorum Class2-1B F. culmorum
F. culmorum NRRL 25475 T
2× F. cerealis S18/34 F. cerealis
100/83/84.5
-/15.2/31.6 F. pseudograminearum CS3096
100/18.2/48 F. pseudograminearum RBG5266
100/71.2/98.1 F. pseudograminearum CS3270 F. pseudograminearum
F. pseudograminearum LC18671
100/83.5/94.6 F. praegraminearum NRRL 39664 F. praegraminearum
100/25/33.6 F. subtropicale NRRL 66764 F. subtropicale

Sporotrichioides clade Sambucinum clade


100/86.1/98.9 F. sambucinum F-4
F. sambucinum
100/72.7/79.5 F. sambucinum NRRL 13708
100/84.5/99.3 F. venenatum A3/5
100/36.1/40.8 F. venenatum
F. venenatum NRRL 66329
100/89.1/98.2 F. kyushuense LC18277
100/68/73.1 F. kyushuense NRRL 25348 F. kyushuense
99/21.8/39.5 F. poae 2516
4× 100/86.6/99 F. poae LC18712
F. poae DAOMC 252244 F. poae
100/80.7/72.3 -/16.5/35.6
F. poae NRRL 26941 ET
100/43.4/76.8 F. sporotrichioides S18/43
100/50.9/86.3 F. sporotrichioides NRRL 3299 T F. sporotrichioides
100/40.5/74.5 F. sibiricum NRRL 53430 T F. sibiricum
100/35.1/64.1 100/74.6/99.5
F. langsethiae Fe2391
F. langsethiae Fl201059 F. langsethiae
100/24.7/42.3
F. goolgardi NRRL 66250 T F. goolgardi
F. nodosum NRRL 36351 F. nodosum
4× 100/63.3/74 -/19.4/37.6 F. chaquense NRRL 66750
100/55.2/95.7 F. chaquense NRRL 66748 F. chaquense
100/74.5/94.5 F. chaquense NRRL 66749 T
F. armeniacum LC15881
100/39.5/49.3 100/55.8/90.2 F. armeniacum NRRL 25141 T F. armeniacum
F. palustre NRRL 54050 F. palustre
100/63.3/87.6 F. longipes NRRL 20695 T Longipes
2× 100/70/82.5 F. longipes NRRL 13317 F. longipes
100/89.6/92.6 Fusarium sp. NRRL 13374 Fusarium sp. clade
Fusarium sp. NRRL 13368 Fusarium sp.
100/41.1/43.8 100/89/98.6 F. brachygibbosum FSAMSC_23
F. brachygibbosum
Brachygibbosum
100/55.5/60.9 F. brachygibbosum HN-1 clade
F. transvaalense NRRL 31008 F. transvaalense
F. nelsonii NRRL 13338 Outgroup

Fig. 15. Maximum likelihood phylogenomic tree of the Fusarium sambucinum species complex (FSAMSC). A total of 5 139 single copy orthologs were
employed in the analysis. F. nelsonii (NRRL 13338) was used as an outgroup. Strains sequenced in this study were indicated in red. The IQ-TREE ultrafast
bootstrap support values (UFBoot ≥ 95 %), gCF and sCF values were displayed at the nodes (UFBoot / gCF / sCF). Ex-type and ex-epitype strains were
indicated in bold with T and ET, respectively.

www.studiesinmycology.org 115
Han et al.

A C
90
F. annulatum
80
70 F. awaxy
Maize Wheat 60 F. concentricum
50 F. elaeagni
6
14 5 40 F. erosum

FFSC
30
F. fujikuroi
20
9 10 F. planum
5 3 0 F. sacchari
F. sanyaense
F. subglutinans
1 F. temperatum
F. verticillioides
F. arcuatisporum
Rice F. clavus
F. compactum
F. fecundum
F. guilinense
B Other species
F. hainanense

FIESC
FIESC F.sulawesiense F. humuli
FFSC F.annulatum FSAMSC F.asiaticum F. ipomoeae
FFSC F.fujikuroi FSAMSC F.boothii F. jinanense
FFSC F.verticillioides FSAMSC F.graminearum
FIESC F.ipomoeae F. luffae
FSAMSC F.pseudograminearum F. mianyangense
FIESC F.luffae
F. nanum
100% F. nothincarnatum
F. pernambucanum
F. sulawesiense
F. tanahbumbuense
F. weifangense
Relative Percentage

SC SC
FO FN
75% 13% F. commune
F. cugenangense
17% F. nirenbergiae
20% F. armeniacum
13%
18% F. asiaticum
F. boothii
50%

FSAMSC
30% 13% F. graminearum
24%
F. kyushuense
14%
F. meridionale
11% F. poae
20% F. pseudograminearum
25% 17% F. vorosii
18%

FTSC
22% 21% F. acuminatum
11% F. avenaceum
15% 20% 16%
M

Te per l m n
Tr per co soo
W

Te rop un
Pl t
R e

Su au

12% 13%
ic
ai

at

op ate nt n
he

11% 11%
m ate on
m ica tai
bt mo
e
z

ic
a

al mo en

0%
m ns t a
on o l
Te ent ate
ze

e
Su nta au
e

at

co em on l

on l
T o ca

so ica

so on
so rat
m p l
ic

m Tr on
m tr n

a
he
ai

in er
i
R

s i

on
on op

on op
ou at

on e

in
M

nt p
m Pl

m
b

Host Climate
Host Climate

Fig. 16. A. The number of unique and shared Fusarium species among different hosts. B. The relative percentage of Fusarium species in different groups
(host: maize, rice, wheat; climate regions: regions affected by plateau mountain, subtropical monsoon, temperate continental, temperate monsoon, and
tropical monsoon climate). The smaller proportion strains of Fusarium were defined as “Other species”, including F. acuminatum, F. arcuatisporum, F.
armeniacum, F. awaxy, F. clavus, F. commune, F. compactum, F. concentricum, F. cugenangense, F. elaeagni, F. erosum, F. fecundum, F. guilinense,
F. hainanense, F. humuli, F. ipomoeae, F. jinanense, F. kyushuense, F. meridionale, F. mianyangense, F. nanum, F. nirenbergiae, F. nothincarnatum, F.
pernambucanum, F. planum, F. poae, F. sacchari, F. sanyaense, F. subglutinans, F. tanahbumbuense, F. temperatum, F. vorosii, and F. weifangense. C.
Detailed information on Fusarium composition among different hosts and climate regions. The number of isolates (0–90) was represented by the shade of
colour (light to dark).

CBS 480.96. Therefore, CBS 217.76 and CBS 480.96 represent Fusarium awaxy Petters-Vandresen et al., Persoonia 43: 363.
two distinct species. In the current study, based on the taxonomic 2019.
treatment of Yilmaz et al. (2021), we further confirmed that many
cereals pathogenic strains (including F1, P2, 24ALH, 51ALH, Material examined: (See Supplementary Table S1).
Azerbaijan 71-3) previously identified as F. proliferatum (Wulff et al.
2010, Okello et al. 2019, Özer et al. 2020), are actually F. annulatum Notes: Fusarium awaxy was originally described from rotten stalks
(Fig. 8). Meanwhile, our results also showed that F. annulatum has of Zea mays in Brazil (Crous et al. 2019), and we herein reported it
a broad host range and geographical distribution (Fig. 16C). In for the first time from China (Fig. 8). This species is closely related
addition, our inoculation test showed its high pathogenicity to maize to the cosmopolitan species F. subglutinans (Figs 8, 14), and both
seedlings (DSI = 68.75 % when CC = 1 × 106 conidia/mL) (Fig. 7E). species have been confirmed for causing maize stalk rot (Leslie &

116
Fusarium spp. on cereals in China

17 Samples
F. g appears in 5 samples
F. p appears in 15 samples

5 Co-infection
samples
F. g appears in 5 samples
F. p appears in 4 samples
F. g & F. p appear in 4 samples
Sa 8
m 8
pl
es
Co 4
- 4
sa infe
m cti
ple on

4 sa ples
18

es
37 s
Co-infection

mpl
am
Samples

les
pea 1 samp s
FCR

23 s les
e
samples

ears in 52 s
mpl

amp
0 sa

in 3
ies
No common pathogens 5

in 3

s
ec

in 3
r in

F. g ppear
were reported from
19 es sp 22
sp

ears
FHB

ars
previous studies
ci ecie

app
RSR spe s

.aa
g ap
e
app
app
spe26 27 ies

F
F. a

& F.
F. g
cie c
s spe

F. a
MSR MER

31 78
Co-infection Co-infection
samples samples

F. an appears in 13 samples F. an appears in 41 samples 126


46 Samples
F. v appears in 14 samples F. v appears in 49 samples
Samples
F. an & F. v appear in 6 samples F. an & F. v appear in 31 samples

F. an appears in 17 samples F. an appears in 45 samples


F. v appears in 18 samples F. v appears in 70 samples

Fig. 17. Summary of co-infections discovered in this study. The white circle represents the number of species isolated from different disease samples (only
one species was isolated from the rice bakanae disease sample, thus information was not shown in this figure). The pentagon represents the number of
co-infection samples, also marked the occurrence of common pathogens, including F. a, F. an, F. g, F. p and F. v. The outside sectors of pentagon represent
the number of whole samples, also marked the occurrence of common pathogens as previously mentioned.
Abbreviations: Fusarium crown rot of wheat (FCR), Fusarium head blight of wheat (FHB), maize ear rot (MER), maize stalk rot (MSR), rice spikelet rot (RSR),
F. asiaticum (F. a), F. annulatum (F. an), F. graminearum (F. g), F. pseudograminearum (F. p), F. verticillioides (F. v).

Summerell 2006: Fig. 7H). Morphologically, they differ from each and Asia (Aoki et al. 2001, Choi et al. 2018, Du et al. 2020). The
other in their sporodochia colour (Crous et al. 2019). In addition, pathogenicity of this species in causing maize ear rot, rice bakanae
F. subglutinans has been reported for causing human keratitis disease and maize stalk rot has been confirmed by Du et al. (2020),
(Al-Hatmi et al. 2014), but F. awaxy does not grow above 32 °C, Choi et al. (2018) and this study (Fig. 7I).
suggesting the inability to cause human infection (Crous et al.
2019). Fusarium elaeagni M.M. Wang & L. Cai, Persoonia 48: 35 2022.

Fusarium concentricum Nirenberg & O’Donnell, Mycologia 90: Material examined: (See Supplementary Table S1).
442. 1998.
Notes: Fusarium elaeagni was originally described from Elaeagnus
Material examined: (See Supplementary Table S1). pungens (Wang et al. 2022a), and subsequently from rice in this
study. This species is closely related to F. fujikuroi (Figs 8, 14)
Notes: Fusarium concentricum was originally described from Musa but could be distinguished by morphological characters, i.e.,
sapientum in Costa Rica, and from Nilaparvata lugens in Korea sporodochial colour, macroconidial septa, microconidial shape, and
(Nirenberg & O’Donnell 1998). It has since been reported from the type of aerial phialides (Wang et al. 2022a).
other hosts, including maize, rice and wheat from central America

www.studiesinmycology.org 117
Table 2. Summary of Fusarium species reported from cereals in literatures and the present study.

118
References of Host records in China2 Host records worldwide (except China)2
Han et al.

species publication
and subsequent
Current name1 Reported name1 Complex1 Pathogens1 adjustment Maize Rice Wheat References3 Maize Rice Wheat References3
F. algeriense F. algeriense FburSC FCR Laraba et al. (2017) n/a n/a n/a n/a n/a n/a √ Özer et al. (2020)
F. beomiforme F. beomiforme FburSC FCR Nelson et al. (1987) n/a n/a n/a n/a √ n/a n/a Laraba et al. (2017)
F. anguioides F. anguioides FCOSC n/a Sherbakoff (1915) n/a n/a n/a n/a √ √ √ Mulenko et al. (2008)
F. concolor F. concolor FCOSC n/a Reinking (1934) n/a n/a n/a n/a n/a n/a √ Ebbels & Allen (1979)
F. atrovinosum F. atrovinosum FCSC n/a Lombard et al. (2019c) n/a n/a n/a n/a n/a n/a √ Lombard et al. (2019c)
F. chlamydosporum F. chlamydosporum FCSC FCR Wollenweber & n/a n/a n/a n/a n/a √ √ Chehri et al. (2010)
Reinking (1925)
F. nelsonii F. nelsonii FCSC MSR Marasas et al. (1998) √ n/a n/a Zhang et al. (2021b) √ n/a √ Marasas et al. (1998),
Chehri et al. (2010)
F. andiyazi F. andiyazi FFSC MER; RBD Marasas et al. (2001) √ √ n/a Zhang et al. (2014a), √ √ n/a Wulff et al. (2010),
Qiu et al. (2020) Leyva-Madrigal et al.
(2015), Venturini et al.
(2017)
F. annulatum F. annulatum FFSC FCR; FHB; Bugnicourt (1952), * * * Huang et al. (2011), * √ * O’Donnell et al. (1998),
MSR; MER; Yilmaz et al. (2021) Qiu et al. (2020), Okello et al. (2019),
RBD present study Özer et al. (2020)
F. anthophilum F. anthophilum FFSC FCR; FHB Wollenweber (1916) n/a n/a n/a n/a n/a n/a √ Chehri et al. (2010)
F. awaxy F. awaxy FFSC MSR Crous et al. (2019) * n/a n/a present study √ n/a n/a Crous et al. (2019)
F. brevicatenulatum F. brevicatenulatum, FFSC MER; RBD Nirenberg et al. (1998) n/a n/a n/a n/a √ √ n/a Amatulli et al. (2010),
F. pseudoanthophilum Tsehaye et al. (2016)

F. concentricum F. concentricum FFSC MER; MSR; Nirenberg & O’Donnell √√ * n/a Du et al. (2020), √ √ √ Aoki et al. (2001), Choi
RBD (1998) present study et al. (2018)
F. dlaminii F. dlaminii FFSC n/a Marasas et al. (1985) n/a n/a n/a n/a √ n/a n/a Scauflaire et al. (2011)
F. elaeagni F. elaeagni FFSC n/a Wang et al. (2022a) n/a * n/a present study n/a n/a n/a n/a
F. erosum F. erosum FFSC MSR present study * n/a n/a present study n/a n/a n/a n/a
F. fujikuroi F. fujikuroi, F. FFSC FCR; FHB; Nirenberg (1976) √√ √√ √√ Duan et al. (2020), √ √ √ McGuire & Crandall
moniliforme, F. sacchari MER; MSR; Qiu et al. (2020), (1967), Byrnes
var. subglutinans, F. RBD; RSR present study & Carroll (1986),
moniliforme var. majus, Adikaram &
Gibberella fujikuroi, Yakandawala (2020)
Lisea fujikuroi
F. globosum F. globosum FFSC n/a Rheeder et al. (1996) n/a n/a n/a n/a √ n/a √ Nirenberg & O’Donnell
(1998); Moses et al.
(2010)
F. madaense F. madaense FFSC MSR Ezekiel et al. (2020) n/a n/a n/a n/a √ √ n/a Costa et al. (2022)
F. napiforme F. napiforme FFSC RBD Marasas et al. (1987) n/a n/a n/a n/a n/a √ n/a Amatulli et al. (2010)
Table 2. (Continued).
References of Host records in China2 Host records worldwide (except China)2
species publication
and subsequent
Current name1 Reported name1 Complex1 Pathogens1 adjustment Maize Rice Wheat References3 Maize Rice Wheat References3
F. nygamai F. nygamai FFSC FHB; MSR Burgess & Trimboli n/a n/a n/a n/a √ √ √ Balmas et al. (2000),
(1986) Leyva-Madrigal et al.
(2015)
F. proliferatum F. proliferatum, FFSC ? Nirenberg (1976) ? ? ? Huang et al. (2010), ? ? ? Wulff et al. (2010),
F. proliferatum var. Qiu et al. (2020) Yasuhara-Bell et al.
minus (2018), Özer et al.
(2020)
F. pseudoanthophilum F. pseudoanthophilum FFSC n/a Nirenberg et al. (1998) n/a n/a n/a n/a √ n/a n/a Tsehaye et al. (2016)

www.studiesinmycology.org
F. pseudonygamai F. pseudonygamai FFSC RBD Nirenberg & O’Donnell n/a n/a n/a n/a n/a √ n/a Bashyal et al. (2016)
(1998)
F. planum F. planum FFSC MSR present study * * * present study n/a n/a n/a n/a
F. sacchari F. sacchari, FFSC FHB; MSR; Gams (1971) √√ * √ Wang et al. (2015), √ √ √ Mulenko et al. (2008),
Cephalosporium MER; RBD Duan et al. (2019), Bashyal et al. (2016)
sacchari present study
F. sanyaense F. sanyaense FFSC MSR present study * n/a n/a present study n/a n/a n/a n/a
F. subglutinans F. subglutinans, FFSC MER; MSR; Nelson et al. (1983) √√ n/a * Gai et al. (2017), √ √ √ Byrnes & Carroll
F. moniliforme RBD present study (1986), Pak et al.
var. subglutinans, (2016)
Gibberella fujikuroi var.
subglutinans
F. temperatum F. temperatum FFSC MER; MSR Scauflaire et al. (2011) √√ n/a n/a Zhang et al. (2014c), √ n/a n/a Varela et al. (2013)
Gai et al. (2017),
present study
F. thapsinum F. thapsinum FFSC MER; MSR; Klittich et al. (1997) √ √ n/a Dong et al. (2020b), √ √ n/a Rahjoo et al. (2008),
RBD Zhang et al. (2021a) Bashyal et al. (2016)

F. verticillioides F. verticillioides, FFSC FCR; FHB; Nirenberg (1976), √√ √√ n/a Gai et al. (2017), √ √ √ Chehri et al. (2010),
Fusarium moniliforme MSR; MER; Yilmaz et al. (2021) Dong et al. (2020b), Wulff et al. (2010),
var. majus, Gibberella RBD Qiu et al. (2020), Gräfenhan et al. (2011)
moniliform, Gibberella present study
moniliformis
F. graminum F. avenaceum var. FHSC n/a Corda (1837) n/a n/a n/a n/a n/a n/a √ Mulenko et al. (2008)
graminum
F. heterosporum F. heterosporum FHSC FCR Nees von Esenbeck n/a n/a n/a n/a √ √ √ Tunali et al. (2008),
(1818) Mulenko et al. (2008)
F. aberrans F. aberrans FIESC n/a Xia et al. (2019) n/a n/a n/a n/a n/a √ n/a Xia et al. (2019)
F. arcuatisporum F. arcuatisporum FIESC MSR Wang et al. (2019a) * n/a n/a present study n/a n/a n/a n/a
F. camptoceras F. camptoceras FIESC n/a Wollenweber & n/a n/a n/a n/a n/a √ √ Chehri et al. (2010)
Reinking (1925)

119
Fusarium spp. on cereals in China
Table 2. (Continued).

120
References of Host records in China2 Host records worldwide (except China)2
Han et al.

species publication
and subsequent
Current name1 Reported name1 Complex1 Pathogens1 adjustment Maize Rice Wheat References3 Maize Rice Wheat References3
F. clavus (as ‘clavum’) F. clavus (as ‘clavum’) FIESC n/a Xia et al. (2019) n/a n/a * present study * n/a * Okello et al. (2019),
Özer et al. (2020)
F. compactum F. compactum FIESC FCR Raillo (1950) * n/a n/a present study n/a n/a √ Tunali et al. (2008)
F. nothincarnatum F. nothincarnatum FIESC n/a present study n/a * * present study n/a * n/a O’Donnell et al. (2009)
F. equiseti F. equiseti, FIESC ? Saccardo (1886) ? n/a ? Li et al. (2014), Xu et ? ? ? Aguín et al. (2014),
Gibberella intricans al. (2016), Gai et al. Barkat et al. (2016)
(2017)
F. fasciculatum F. fasciculatum FIESC n/a Xia et al. (2019) n/a n/a n/a n/a n/a √ n/a Xia et al. (2019)
F. guilinense F. guilinense FIESC n/a Wang et al. (2019a) n/a * n/a present study n/a n/a n/a n/a
F. hainanense F. hainanense FIESC n/a Wang et al. (2019a) * * n/a Wang et al. (2021), n/a √ n/a Xia et al. (2019)
present study
F. humuli F. humuli FIESC n/a Wang et al. (2019a) n/a * * present study n/a n/a n/a n/a
F. incarnatum F. incarnatum, F. FIESC ? Saccardo (1886) ? n/a ? Gai et al. (2016, ? ? ? Tsehaye et al. (2016),
semitectum, F. 2017) Choi et al. (2018),
semitectum var. majus, Özer et al. (2020)
F. diversisporum, F.
pallidoroseum, F.
semitectum
F. ipomoeae F. ipomoeae FIESC MSR Wang et al. (2019a) * √ * Li et al. (2014), n/a n/a n/a n/a
Wang et al. (2019a),
present study
F. jinanense F. jinanense FIESC MSR present study * n/a n/a present study n/a n/a n/a n/a
F. luffae F. luffae FIESC MSR Wang et al. (2019a) * * * Gai et al. (2016), n/a n/a n/a n/a
present study
F. mianyangense F. mianyangense FIESC MSR present study * * n/a present study n/a n/a n/a n/a
F. nanum F. nanum FIESC n/a Wang et al. (2019a) * n/a * present study n/a n/a n/a n/a
F. pernambucanum F. pernambucanum FIESC MSR Santos et al. (2019) * n/a n/a present study n/a n/a n/a n/a
F. sulawesiense (as F. sulawesiense (as FIESC MSR Maryani et al. (2019b) * √√ * Wang et al. (2019a), n/a n/a n/a n/a
‘sulawense’) ‘sulawense’) present study
F. tanahbumbuense F. tanahbumbuense FIESC MSR Maryani et al. (2019b) * * * present study n/a n/a * Özer et al. (2020)
F. weifangense F. weifangense FIESC n/a present study n/a n/a * present study n/a n/a n/a n/a
F. fecundum F. fecundum FIESC n/a present study n/a * * present study n/a n/a n/a n/a
F. lateritium F. lateritium FLSC FCR Nees von Esenbeck n/a n/a n/a n/a √ √ √ Richardson (1979),
(1817) Mulenko et al. (2008),
Tunali et al. (2008)
F. sarcochroum F. sarcochroum FLSC n/a Saccardo (1879) n/a n/a n/a n/a √ n/a n/a Mulenko et al. (2008)
F. miscanthi F. miscanthi FnewSC MER Gams et al. (1999) √ n/a n/a Shang et al. (2020) n/a n/a n/a n/a
F. nisikadoi F. nisikadoi FnewSC n/a Nirenberg (1997) n/a n/a n/a n/a n/a n/a √ Nirenberg (1997)
Table 2. (Continued).
References of Host records in China2 Host records worldwide (except China)2
species publication
and subsequent
Current name1 Reported name1 Complex1 Pathogens1 adjustment Maize Rice Wheat References3 Maize Rice Wheat References3
F. commune F. commune FNSC MSR; RBD Skovgaard et al. √√ * n/a Xi et al. (2019), √ √ n/a Choi et al. (2018),
(2003) present study Husna et al. (2020),
Mezzalama et al.
(2021)
F. carminascens F. carminascens FOSC n/a Lombard et al. n/a n/a n/a n/a √ n/a n/a Lombard et al. (2019b)
(2019b)
F. cugenangense F. cugenangense FOSC n/a Maryani et al. (2019a) * n/a n/a present study n/a n/a n/a n/a
F. fabacearum F. fabacearum FOSC n/a Lombard et al. n/a n/a n/a n/a √ n/a n/a Lombard et al. (2019b)
(2019b)

www.studiesinmycology.org
F. inflexum F. inflexum FOSC FCR Schneider & Dalchow n/a n/a n/a n/a n/a n/a √ Tunali et al. (2008)
(1975)
F. languescens F. languescens FOSC n/a Lombard et al. n/a n/a n/a n/a √ n/a n/a Lombard et al. (2019b)
(2019b)
F. nirenbergiae F. nirenbergiae FOSC MSR Lombard et al. * n/a n/a present study n/a n/a n/a n/a
(2019b)
F. oxysporum F. oxysporum, FOSC FCR; FHB; von Schlechtendal n/a n/a n/a n/a √ √ √ Amatulli et al. (2010),
F. aurantiacum, MER; MSR; (1824) Barkat et al. (2016),
F. oxysporum, RBD Abdul Rahm et al.
F. vasinfectum, (2020)
F. orthoceras
F. hostae F. hostae FRSC FCR Geiser et al. (2001) n/a n/a n/a n/a n/a n/a √ Özer et al. (2019)
F. redolens F. redolens, F. FRSC FCR Wollenweber (1913) n/a √ √ Tai (1979), Wang et √ n/a √ Mulenko et al. (2008),
oxysporum var. al. (2019b) Esmaeili Taheri et al.
redolens (2011)
F. acaciae-mearnsii F. acaciae-mearnsii FSAMSC n/a O’Donnell et al. (2004) n/a n/a √ Liu et al. (2018) n/a n/a √ Beukes et al. (2017)
F. armeniacum F. armeniacum, F. FSAMSC FHB; MSR Burgess et al. (1993) * n/a * present study √ n/a √ Burgess et al. (1993),
acuminatum subsp. Krone et al. (2020)
armeniacum
F. asiaticum F. asiaticum FSAMSC FCR; FHB; O’Donnell et al. (2004) √√ √√ √√ Zhang et al. (2007, √ √ √ Gomes et al. (2015),
MER; MSR; 2015, 2016), Hao Beukes et al. (2017),
RSR et al. (2017), Dong Khaledi et al. (2017)
et al. (2020b, c),
present study

F. boothii F. boothii FSAMSC FHB; MER; O’Donnell et al. (2004) √√ n/a n/a Gai et al. (2017), √ n/a √ Beukes et al. (2017)
MSR present study
F. brachygibbosum F. brachygibbosum FSAMSC FCR; MSR Padwick (1945) √ n/a n/a Shan et al. (2017) n/a n/a √ Özer et al. (2020)
F. cerealis F. cerealis, FSAMSC FHB; MSR Saccardo (1886) √ n/a n/a Shan et al. (2018) √ n/a √ Castañares et al.
F. crookwellense (2019), Xue et al.
(2019)

121
Fusarium spp. on cereals in China
Table 2. (Continued).

122
References of Host records in China2 Host records worldwide (except China)2
Han et al.

species publication
and subsequent
Current name1 Reported name1 Complex1 Pathogens1 adjustment Maize Rice Wheat References3 Maize Rice Wheat References3
F. cortaderiae F. cortaderiae FSAMSC MER; MSR O’Donnell et al. (2004) n/a n/a n/a n/a √ √ √ Gomes et al. (2015),
Kuhnem et al. (2016),
Beukes et al. (2017)
F. culmorum F. culmorum FSAMSC FCR; FHB; Saccardo (1892) √ n/a √ Tai (1979), Zhuang √ n/a √ Aguín et al. (2014),
MSR; MER (2005), Li et al. Touati-Hattab et al.
(2016), Xu et al. (2016)
(2016), Xia et al.
(2021)
F. gerlachii F. gerlachii FSAMSC FHB Starkey et al. (2007) n/a n/a n/a n/a n/a n/a √ Starkey et al. (2007)
F. graminearum F. graminearum, FSAMSC FCR; FHB; Schwabe (1839) √√ √√ √√ Tai (1979), Zhuang √ √ √ Devay et al. (1957),
Gibberella roseum f. MSR; MER; (2005), Zhang et Pennycook (1989),
cerealis, Gibberella RBD; RSR al. (2007, 2015), Cho & Shin (2004)
zeae Dong et al. (2020b),
present study
F. kyushuense F. kyushuense FSAMSC FHB; MER; Aoki & O’Donnell √ √ * Zhao & Lu (2007), n/a n/a √ Sandoval-Denis et al.
MSR (1998) Wang et al. (2014), (2018b)
Cao et al. (2021),
present study
F. langsethiae F. langsethiae FSAMSC n/a Torp & Nirenberg n/a n/a n/a n/a n/a n/a √ Lukanowski et al.
(2004) (2008)
F. louisianense F. louisianense FSAMSC FHB Sarver et al. (2011) n/a n/a n/a n/a n/a n/a √ Yasuhara-Bell et al.
(2018)
F. meridionale F. meridionale FSAMSC FHB; MER; O’Donnell et al. (2004) √√ √ * Zhang et al. (2014b), √ √ √ Gomes et al. (2015),
MSR; RSR Gai et al. (2017), Kuhnem et al. (2016),
Dong et al. (2020a), Khaledi et al. (2017)
present study
F. nepalense F. nepalense FSAMSC RSR Sarver et al. (2011) n/a n/a n/a n/a n/a √ n/a Sarver et al. (2011)
F. nodosum F. nodosum FSAMSC n/a Lombard et al. (2019c) n/a n/a n/a n/a n/a n/a √ Lombard et al. (2019c)
F. poae F. poae FSAMSC FHB; MER Wollenweber (1914) √√ n/a √√ Tai (1979), Xu et √ n/a √ Logrieco et al. (2002a),
al. (2020), present Yli-Mattila et al. (2004)
study
F. pseudograminearum F. pseudograminearum FSAMSC FCR; FHB Aoki & O’Donnell * n/a √√ Li et al. (2012), Xu √ n/a √ Obanor et al. (2010),
(1999) et al. (2015), Zhang Abdallah-Nekache et
et al. (2015), Ji et al. (2019)
al. (2016), present
study
F. sambucinum F. sambucinu, FSAMSC n/a Link (1809), Gams n/a n/a n/a n/a √ √ √ Mulenko et al. (2008),
Gibberella pulicaris, F. (1997) Chehri et al. (2010)
roseum
Table 2. (Continued).
References of Host records in China2 Host records worldwide (except China)2
species publication
and subsequent
Current name1 Reported name1 Complex1 Pathogens1 adjustment Maize Rice Wheat References3 Maize Rice Wheat References3
F. sporotrichioides F. sporotrichioides, FSAMSC FHB; MER; Sherbakoff (1915) √ n/a n/a Wang et al. (2020) √ √ √ Tekauz et al. (2004),
F. sporotrichioides var. RSR Amatulli et al. (2010),
sporotrichioides Wang et al. (2020)
F. venenatum F. venenatum FSAMSC FCR Nirenberg (1995) n/a n/a n/a n/a √ n/a √ Sandoval-Denis et al.
(2018a, b)
F. vorosii F. vorosii FSAMSC FHB; MSR; Starkey et al. (2007) * n/a n/a present study √ √ √ Starkey et al. (2007),
MER Lee et al. (2016)

F. acuminatum F. acuminatum, F. FTSC FCR; FHB; Sherbakoff (1915) * n/a √√ Pan et al. (2015), √ √ √ Desjardins et al.

www.studiesinmycology.org
scirpi, F. scirpi var. MSR Zhang et al. (2015), (2000), Mao et al.
acuminatum present study (1998), Shrestha et al.
(2021)
F. avenaceum F. avenaceum, FTSC FCR; FHB; Saccardo (1886) √ n/a √√ Zhuang (2005), √ √ √ Logrieco et al. (2002b);
Gibberella avenacea MER Zhang et al. (2013, Amatulli et al. (2010)
2015), Ma et al.
(2019), present
study
F. flocciferum F. flocciferum FTSC n/a Sturm (1829) n/a n/a n/a n/a √ n/a √ Gerlach & Ershad
(1970), Sandoval-
Denis et al. (2018a)
F. torulosum F. torulosum FTSC n/a Nirenberg (1995) n/a n/a n/a n/a n/a n/a √ Kristensen et al.
(2005)
F. tricinctum F. tricinctum FTSC n/a Saccardo (1886) n/a √ n/a Li et al. (2019b) √ n/a √ Mulenko et al. (2008),
Castañareset al.
(2011)
F. trichothecioides F. trichothecioides n/a FCR Jamieson & n/a n/a n/a n/a n/a n/a √ Chehri et al. (2010)
Wollenweber (1912)
1
Abbreviations: F. = Fusarium. FburSC = F. burgessii species complex. FCSC = F. chlamydosporum species complex. FCOSC = F. concolor species complex. FFSC = F. fujikuroi species complex. FHSC
= F. heterosporum species complex. FIESC = F. incarnatum-equiseti species complex. FLSC = F. lateritium species complex. FnewSC = F. newnesense species complex. FNSC = F. nisikadoi species
complex. FOSC = F. oxysporum species complex. FRSC = F. redolens species complex. FSAMSC = F. sambucinum species complex. FTSC = F. tricinctum species complex. FCR = Fusarium crown rot of
wheat. FHB = Fusarium head blight of wheat. MSR = maize stalk rot. MER = maize ear rot. RBD = Rice bakanae disease. RSR = Rice spikelet rot. Maize = Zea mays. Rice = Oryza sp. Wheat =Triticum sp.
2
n/a: No record found. √: Species previously reported from cereals in literatures. √√: Species also reported on cereals in the present study. *: New records on cereals in China or updated records on
cereals by re-identification. ?: The records on cereals need to be further confirmed.
3
If there are more than three references for a particular record, only several representative references were listed here.

123
Fusarium spp. on cereals in China
Han et al.

Fusarium erosum S.L. Han, M.M. Wang & L. Cai, sp. nov. Notes: As a pathogen causing rice bakanae disease, F. fujikuroi
MycoBank MB 847021. Fig. 18. has been widely studied, but some are under different synonyms
(e.g., F. moniliforme var. subglutinans, F. moniliforme var.
Etymology: Refers to its colony margin on PDA, erose. intermedium, F. moniliforme and Gibberella fujikuroi) based on
previous morphological identification systems (Booth 1971). With
Sporodochia cream (4A3), formed frequently on carnation leaves, the application of multi-locus analyses, F. fujikuroi was gradually
and often covered by aerial sparse mycelium. Sporodochial revealed to be a pathogen of various crops worldwide (Farr &
conidiophores densely, irregularly and verticillate branched, Rossman 2022), causing maize ear rot, maize stalk rot, wheat
bearing apical whorls of 2–4 phialides. Sporodochial phialides scab, wheat crown rot and rice spikelet rot (Wulff et al. 2010, Duan
subulate to subcylindrical, smooth, thin-walled, 12.4–19.8 × 2.0– et al. 2020, Qiu et al. 2020, this study).
3.8 μm. Sporodochial conidia falcate, slender, straight to slightly
curved dorsiventrally, with a curved to pointed apical cell and a Fusarium planum S.L. Han, M.M. Wang & L. Cai, sp. nov.
well-developed foot-shaped basal cell, hyaline, thin- and smooth- MycoBank MB 847022. Fig. 19.
walled, 3(–5)-septate: 3-septate conidia: 29.3–46.6 × 3.5–4.7
μm (av. ± SD: 39.8 ± 5.8 × 4.0 ± 0.4 μm); 4-septate conidia: Etymology: Referring to its colony elevation in PDA, flat.
42.7–50.6 × 3.5–4.7 μm (av. ± SD: 46.4 ± 2.3 × 4.0 ± 0.3 μm);
5-septate conidia: 39.4–50.5 × 3.5–5.1 μm (av. ± SD: 46.8 ± 2.8 Sporodochia cadmium orange (5A8), formed infrequently inside
× 4.2 ± 0.4 μm). Aerial conidiophores borne on aerial mycelium, agar. Sporodochial conidiophores densely and irregularly
straight or flexuous, erect or prostrate, smooth- and thin-walled, branched, bearing apical whorls of 3–5 phialides. Sporodochial
bearing terminal or lateral phialides. Aerial phialide mono- and phialides subulate to subcylindrical, smooth, thin-walled, 9.0–17.5
polyphialides, subulate to subcylindrical, smooth- and thin-walled, × 2.3–3.9 μm. Sporodochial conidia falcate, slender, straight to
periclinal thickening inconspicuous or absent, 9.7–32.1 × 1.9–3.5 slightly curved dorsiventrally, with a conical to slightly papillate
μm. Aerial conidia single on the tips of phialides, ovoid, hyaline, apical cell and obtuse to barely notched basal cell, hyaline, thin-
smooth- and thin-walled, aseptate, 3.7–11.3 × 1.8–3.5 μm (av. ± and smooth-walled, (1–)3–4(–5)-septate: 1-septate conidia: 14.4–
SD: 7.2 ± 1.7 × 2.7 ± 0.4 μm). Chlamydospores not observed. 22.1 × 3.1–3.9 μm (av. ± SD: 17.6 ± 2.3 × 3.5 ± 0.3 μm); 2-septate
conidia: 21.5–28.5 × 3.2–4.3 μm (av. ± SD: 25 ± 2.4 × 3.7 ± 0.4
Culture characteristics: Colonies on PDA incubated at 25 °C in μm); 3-septate conidia: 25–43 × 3.0–4.5 μm (av. ± SD: 35.8 ± 5.7
the dark reaching 72–82 mm diam in 7 d; convex, with abundant × 3.6 ± 0.4 μm); 4-septate conidia: 41.8–49.4 × 3.3–5.0 μm (av. ±
aerial mycelium, colony margin lightly erose; surface purplish white SD: 44.5 ± 2.8 × 4.2 ± 0.5 μm); 5-septate conidia: 40.5–48.3 × 3.4–
(14A2), reverse pinkish white (10A2) in the centre, white (–A1) at 4.8 μm (av. ± SD: 43.7 ± 4.1 × 4.2 ± 0.5 μm). Aerial conidiophores
the margin; odour absent. On OA in the dark reaching 66–67 mm in borne on aerial mycelium, straight or flexuous, erect or prostrate,
7 d; convex, with abundant aerial mycelium, margin entire; surface smooth- and thin-walled, bearing terminal or lateral phialides. Aerial
pinkish white (10A2), reverse greyish red (10C4); odour absent. phialides monophialide, subulate to subcylindrical, smooth- and
thin-walled, periclinal thickening inconspicuous or absent, 12.6–25
Typus: China, Guangdong Province, Meizhou City (E116.3, N24.52), from × 2–2.8 μm. Aerial conidia often forming false heads, subcylindrical
the symptomatic tissues of maize stalk rot, 7 Jun. 2021, Y.M. Wu (holotype to clavate, hyaline, smooth- and thin-walled, aseptate, 5–10.6 ×
HMAS 351950, ex-type living culture CGMCC3.23518 = LC15877 = 1.7–2.3 μm (av. ± SD: 7.7 ± 1.3 × 2.1 ± 0.2 μm). Chlamydospores
HSL2912). not observed.

Additional materials examined: (See Supplementary Table S1). Culture characteristics: Colonies on PDA incubated at 25 °C in the
dark reaching 68–70 mm diam in 7 d; flat, aerial mycelia scant,
Notes: The three isolates representing F. erosum were resolved colony margin entire; surface greyish orange (6B5) in the centre,
as a strongly supported genealogically exclusive lineage in the pale (2A2) at the margin; reverse peach (7A4) in the centre, pale
phylogeny inferred from combined CaM, rpb1, rpb2, tef1, and (2A2) at the margin; odour absent. On OA in the dark reaching
tub2 loci (Fig. 8). Fusarium erosum is closely related to F. siculi 69–71 mm in 7 d; convex, with abundant aerial mycelium, margin
and F. globosum, but differs by 20 bp and 12 bp from the latter entire; surface white (–A1), reverse pastel pink (11A4) in the centre,
two species respectively in the combined dataset. Morphologically, white (–A1) at the margin; odour absent.
F. erosum differs in the types of aerial conidia production and the
number of septa. For example, F. erosum produces aseptate, Typus: China, Guangdong Province, Qingyuan City (E113.42, N24.19),
ovoid aerial conidia; F. siculi produces 0–1-septate, subcylindrical from the symptomatic tissues of maize ear rot, 21 Nov. 2020, S.Q. Wang
to clavate aerial conidia; while three types of aerial conidia were (holotype HMAS 351949, ex-type living culture CGMCC3.23517 =
found in F. globosum: clavate with a truncate base (0–3-septate), LC15876 = HSL2645).
napiform/pyriform, and globose (0–1-septate) which often have
a distinct papilla (Rheeder et al. 1996, Leslie & Summerell 2006, Additional materials examined: (See Supplementary Table S1).
Sandoval-Denis et al. 2018a). The pathogenicity test fulfilling
Koch’s postulates confirmed its pathogenicity, causing maize stalk Notes: The isolates representing F. planum were resolved as a
rot (Fig. 7J). strongly supported genealogically exclusive lineage in the phylogeny
inferred from combined CaM, rpb1, rpb2, tef1, and tub2 loci (Fig.
Fusarium fujikuroi Nirenberg, Mitt. Biol. Bundesanst. Land- 8). Fusarium planum is closely related to F. andiyazi, F. madaense
Forstw. Berlin-Dahlem 169: 32. 1976. and F. mirum, but differs by 23 bp and 31 bp from F. andiyazi and
Synonyms: see Crous et al. (2021). F. madaense in the 5-locus (CaM-rpb1-rpb2-tef1-tub2) dataset and
11 bp from F. mirum in the 3-locus (rpb2-tef1-tub2) dataset (CaM,
Material examined: (See Supplementary Table S1). rpb1 and the latter half of rpb2 sequence were not available for

124
Fusarium spp. on cereals in China

Fig. 18. Morphology of Fusarium erosum (CGMCC3.23518, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D. Sporodochial
conidiophores and phialides. E. Sporodochial conidia. F. Aerial conidiophores and phialides. G. Phialides and aerial conidia. Scale bars = 10 μm.

F. mirum). Morphologically, this species is distinguished based 0–3-septate in F. madaense) (Marasas et al. 2001, Ezekiel et al.
on the number of septa in sporodochial conidia (1–5-septate in F. 2020, Costa et al. 2022). The pathogenicity test fulfilling the Koch’s
planum vs 3–6-septate in F. andiyazi, 0–6-septate in F. madaense, postulates confirmed its pathogenicity and ability to cause maize
1–6-septate in F. mirum); and the number of septa in aerial conidia stalk rot (Fig. 7Q).
(aseptate in F. planum vs 0–2-septate in F. andiyazi and F. mirum,

www.studiesinmycology.org 125
Han et al.

Fig. 19. Morphology of Fusarium planum (CGMCC3.23517, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D. Sporodochial
conidiophores and phialides. E. Sporodochial conidia. F. Aerial conidiophores and phialides. G. Aerial conidia. Scale bars = 10 μm.

Fusarium sacchari (E.J. Butler) W. Gams, Cephalosporium-artige Material examined: (See Supplementary Table S1).
Schimmelpilze: 218. 1971.
Basionym: Cephalosporium sacchari E.J. Butler, Mem. Dept. Agric. Notes: Fusarium sacchari is a common pathogen of diverse crops
India, Bot. Ser. 6: 185. 1913. worldwide (Farr & Rossman 2022). In China, its correlation with maize
Synonyms: see Crous et al. (2021). ear rot, wheat scab, rice spikelet rot and maize stalk rot was confirmed
by Duan et al. (2019), Wang et al. (2015) and this study (Fig. 7R).

126
Fusarium spp. on cereals in China

Fusarium sanyaense S.L. Han, M.M. Wang & L. Cai, sp. nov. Fusarium subglutinans (Wollenw. & Reinking) P.E. Nelson et
MycoBank MB 847023. Fig. 20. al., Fusarium species. An illustrated manual for identification: 135.
1983.
Etymology: Named after the city, Sanya, where the holotype was Basionym: Fusarium moniliforme var. subglutinans Wollenw. &
collected. Reinking, Phytopathology 15: 163. 1925.
Synonyms: see Crous et al. (2021).
Sporodochia champagne (4B4), formed frequently on carnation
leaves, and infrequently on agar. Sporodochial conidiophores Material examined: (See Supplementary Table S1).
densely and irregularly branched, bearing apical whorls of 2–3
phialides. Sporodochial phialides subulate to subcylindrical, Notes: Fusarium subglutinans is known for causing cereal, animal and
smooth, thin-walled, 9.3–19 × 2.5–4.3 μm. Sporodochial conidia human diseases (e.g., maize ear rot, animal rapid death and human
falcate, slightly slender, straight to slightly curved dorsiventrally, keratitis) and producing various mycotoxins (e.g., moniliformin,
with a conical to slightly papillate apical cell and a well-developed fusaproliferin) (Kriek et al. 1977, Bacon & Hinton 1996, Lew et al.
and foot-shaped basal cell, hyaline, thin- and smooth-walled, 1996, Al-Hatmi et al. 2014, Farr & Rossman 2022). Despite its widely
3–5-septate: 3-septate conidia: 25.3–50.6 × 2.7–4.3 μm (av. ± SD: recognised threats to agricultural and public health, its taxonomy was
38.3 ± 6.2 × 3.6 ± 0.5 μm), 4-septate conidia: 42.4–61.1 × 2.6–5.1 unstable as it lacked a living ex-type culture and holotype specimen,
μm (av. ± SD: 50.6 ± 4.8 × 3.9 ± 0.6 μm), 5-septate conidia: 45.7– which was resolved by Yilmaz et al. (2021), who designated CBS
67.2 × 3.4–4.5 μm (av. ± SD: 55.3 ± 5.2 × 3.9 ± 0.4 μm). Aerial 747.97 as its neotype. Twelve strains obtained from this study
conidiophores borne on aerial mycelium, straight or flexuous, erect clustered together with the neotype, of which one strain was isolated
or prostrate, smooth- and thin-walled, bearing terminal or lateral from wheat, as a new host record in China (Table 2).
phialides. Aerial phialides mono- and polyphialides, subulate
to subcylindrical, smooth- and thin-walled, periclinal thickening Fusarium temperatum Scaufl. & Munaut, Mycologia 103: 593.
inconspicuous or absent, 11–27.6 × 2.2–3.4 μm. Aerial conidia 2011.
single on the tips of phialides, ovoid, hyaline, smooth- and thin-
Material examined: (See Supplementary Table S1).
walled, aseptate: 4.4–11 × 1.9–4.2 μm (av. ± SD: 6.5 ± 1.6 × 2.7 ±
0.5 μm). Chlamydospores not observed.
Notes: Fusarium temperatum was first reported as a pathogen on
maize in Belgium (Scauflaire et al. 2011), and subsequently on
Culture characteristics: Colonies on PDA incubated at 25 °C in humans in Mexico (Al-Hatmi et al. 2014). In China, its ability to
the dark reaching 46–55 mm diam in 7 d; raised, aerial mycelia cause maize ear rot and maize stalk rot have been confirmed by
lightly scant, colony margin undulate; surface flamingo (12A4) in Zhang et al. (2014c) and Liu et al. (2022b). Notably, the existence
the centre, white (–A1) at the margin, reverse greyish ruby (12D5) of sexual reproduction of F. temperatum in the field (Scauflaire et al.
in the centre, white (–A1) at the margin; odour absent. On OA in the 2011), together with the semipermeable species boundary with its
dark reaching 47–58 mm in 7 d; flat, aerial mycelia scant, margin relative species F. subglutinans (Fumero et al. 2021), may facilitate
undulate; surface amaranth (14E7) in the centre, reverse orange the genetic recombination and further contribute to the high genetic
white (5A2); odour absent. diversity of F. temperatum. This species appears to be highly
adaptable and would potentially be widely distributed.
Typus: China, Hainan Province, Sanya City (E109.75, N18.39), from the
symptomatic tissues of maize stalk rot, 6 Feb. 2021, Y.J. Li (holotype Fusarium verticillioides (Sacc.) Nirenberg, Mitt. Biol. Bundesanst.
HMAS 351955, ex-type living culture CGMCC3.23523 = LC15882 = Land-Forstw. 169: 26. 1976.
HSL2737). Basionym: Oospora verticillioides Sacc., Fung. Ital., Fasc. 17–28:
pl. 879. 1881.
Additional materials examined: (See Supplementary Table S1). Synonyms: see Crous et al. (2021).

Notes: The three isolates representing F. sanyaense were resolved Material examined: (See Supplementary Table S1).
as a strongly supported genealogically exclusive lineage in the
phylogeny inferred from combined CaM, rpb1, rpb2, tef1, and Notes: Fusarium verticillioides was first isolated from maize in Italy
tub2 loci (Fig. 8). Fusarium sanyaense is closely related to F. as Oospora verticillioides (Saccardo 1881). The controversy on the
lumajangense, F. mangiferae and F. proliferatum, but differed by taxonomy of this species was recently summarised and resolved
18 bp from F. mangiferae in the 5-locus (CaM-rpb1-rpb2-tef1- (Yilmaz et al. 2021). Fusarium verticillioides is cosmopolitan and
tub2) dataset, 13 bp from F. lumajangense in the 3-locus (rpb2- notorious, being predominately a pathogen of maize, causing a
tef1-tub2) dataset (CaM and rpb1 sequences are not available massive quality and yield reduction (Murillo-Williams & Munkvold
for F. lumajangense) and 10 bp from F. proliferatum in the 4-locus 2008, Blacutt et al. 2018, Schoeman et al. 2018). In our study,
(CaM-rpb2-tef1-tub2) dataset (rpb1 sequence is not available for F. Fusarium verticillioides is also one of the dominant species on
proliferatum). Morphologically, they could be distinguished based maize, with a total of 88 representative strains from diseased maize
on the size of sporodochial conidia (25.3–67.2 × 2.7–4.5 μm in F. and from 88 locations (Supplementary Table S1).
sanyaense vs 30.0–56.0 × 3.0–4.5 μm in F. lumajangense, 43.1–
61.4 × 1.9–3.4 μm in F. mangiferae, and 16.5–60.5 × 1.5–4 μm in Fusarium incarnatum-equiseti species complex (FIESC)
F. proliferatum), and the number of septa in sporodochial conidia
(3–5-septate in F. sanyaense, F. lumajangense and F. madaense Species within the FIESC are cosmopolitan in various ecological
vs 1–4-septate in F. proliferatum) (Britz et al. 2002, Maryani et types. The species relationships within this group have only
al. 2019b, Yilmaz et al. 2021). Pathogenicity tests fulfilling the recently been clarified. Previously, members of this group were
Koch’s postulates confirmed its ability to cause maize stalk rot in most cases identified as F. equiseti or F. incarnatum, based on
(Supplementary Fig. 7S). morphological similarities or ITS/tef1 sequences (Khoa et al. 2004,

www.studiesinmycology.org 127
Han et al.

Fig. 20. Morphology of Fusarium sanyaense (CGMCC3.23523, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D. Sporodochial
conidiophores and phialides. E. Sporodochial conidia. F. Aerial conidiophores and phialides. G. Aerial conidia. Scale bars = 10 μm.

Leslie & Summerell 2006, Marín et al. 2012). With the application of clarified and delimitated, and the majority of the cryptic phylo-species
genealogical concordance phylogenetic species recognition, FIESC have been provided with Latin binomials (Wang et al. 2019a, Xia et
has been revealed to include 32 phylogenetic species separated in al. 2019). Furthermore, our phylogenomic tree (Fig. 14) support the
the Incarnatum and Equiseti clades (O’Donnell et al. 2009, 2012, merger of FCAMSC into the FIESC (as the Camptoceras clade).
Villani et al. 2016). By employing morphological characters and multi- Thus, the FIESC now includes 44 species, which is characterised by a
locus phylogenetic relationships, the species within the FIESC were dorsiventral curvature of its macroconidia, abundant chlamydospores,

128
Fusarium spp. on cereals in China

and mostly lacking microconidia (Wang et al. 2019a). To date, about thin-walled, bearing terminal or lateral phialides. Aerial phialides
half of the species in this complex have been reported from cereals mono- and polyphialides, subulate to subcylindrical, smooth- and
(Table 2). However, the pathogenicity of several species remains thin-walled, periclinal thickening inconspicuous or absent, 8–44
suspicious due to uncertainty of the identities of isolates used, e.g., F. × 2.5–3.7 μm. Aerial conidia falcate, the dorsal side curved than
incarnatum (Fig. 9). In this study, 147 strains clustered into 19 distinct the ventral; with a blunt, conical to pointed apical cell and papillate
clades (13 in Incarnatum clade, five in Equiseti clade, and one in basal cell, hyaline, smooth- and thin-walled, (1–)2–4(–6)-septate:
Camptoceras clade), representing 14 known and five novel species 1-septate conidia: 13–14 × 3.5–4 μm (av. ± SD: 13.6 ± 0.4 × 3.8 ±
(Fig. 9). 0.2 μm); 2-septate conidia: 12.6–26 × 3.6–5.4 μm (av. ± SD: 18 ±
4.6 × 4.3 ± 0.6 μm); 3-septate conidia: 24.1–33.1 × 4.5–6.5 μm (av.
Fusarium arcuatisporum M.M. Wang et al., Persoonia 43: 78. ± SD: 29.6 ± 2.5 × 5.5 ± 0.7 μm); 4-septate conidia: 30.9–38.5 ×
2019. 5.4–5.9 μm (av. ± SD: 33.4 ± 3.1× 5.7 ± 0.2 μm); 5-septate conidia:
30.5–40.1 × 5.4–6.4 μm (av. ± SD: 36.7 ± 5.8 × 6.0 ± 0.4 μm);
Material examined: (See Supplementary Table S1). 6-septate conidia: 33.6–35.8 × 6.1–6.8 μm (av. ± SD: 34.4 ± 1 ×
6.4 ± 0.3 μm). Chlamydospores not observed.
Notes: Fusarium arcuatisporum was first described based on isolation
from Nelumbo nucifera, Brassica campestris, and Oryza sp. from
China (Wang et al. 2019a). Previously, this species has been isolated Culture characteristics: Colonies on PDA incubated at 25 °C in
from a human toenail, as an unnamed phylogenetic species (FIESC 7) the dark reaching 84–90 mm diam in 7 d; raised, felty to velvety,
(Leslie & Summerell 2006). Here we update a new host record (maize), radiate, with abundant aerial mycelia, colony margin entire; surface
and confirm its ability to cause maize stalk rot (Fig. 7F). greyish yellow (2B5) in the centre, white (–A1) at the margin;
reverse milk white (1A2) in the centre, white (–A1) at the margin;
Fusarium clavus J.W. Xia et al., [as ‘clavum’], Persoonia 43: 199. odour absent. On OA in the dark reaching 76–84 mm in 7 d; raised,
2019. felty to dusty, with abundant aerial mycelium, margin entire; surface
white, reverse mimosa (2B8); odour absent.
Material examined: (See Supplementary Table S1).
Typus: China, Shaanxi Province, Hanzhong city (E107.632; N33.211),
Notes: Fusarium clavus was first recognised by O’Donnell et al. obtained from the symptomatic tissues of wheat scab, 13 May 2021, Y.J.
(2009) as an unnamed phylogenetic species (FIESC 5), and later Chen (holotype HMAS 351948, ex-type living culture CGMCC3.23516 =
formally named and described by Xia et al. (2019). To date, F. LC15875 = HSL1587).
clavus has been reported from humans, insects, plants and various
environments (Xia et al. 2019, Matic et al. 2020). Notably, several Additional materials examined: (See Supplementary Table S1).
isolates of this species have been reported from maize (Okello et
al. 2019) and wheat (Özer et al. 2020), but inaccurately identified as Notes: The three isolates representing F. fecundum were resolved
FIESC or F. equiseti, based on tef1 phylogenetic analysis (Okello et as a strongly supported genealogically exclusive lineage in the
al. 2019, Özer et al. 2020, Fig. 9). These inaccurate identification phylogenies inferred from combined CaM, rpb2, and tef1 loci (Fig.
results are common, because species within the FIESC could not 9), and genomic datasets (Fig. 14). Phylogenetically, F. fecundum
be distinguished based on only tef1 sequences, especially when is closely related to the species within the previous FCAMSC (F.
using the previously wrongly named reference sequences from kotabaruense, F. camptoceras and F. neosemitectum), but differs
databases. In this study, we reported it as a new Chinese record. by 83 bp, 84 bp and 84 bp in the combined dataset, respectively.
Morphologically, this species is distinguishable from closely
Fusarium compactum (Wollenw.) Raillo, Fungi of the Genus related species based on its conidial size (13–40.1 × 3.5–6.8 μm
Fusarium: 180. 1950. in F. fecundum vs 21–45 ×5–7.5 μm in F. kotabaruense, 15–51 ×
Basionym: Fusarium scirpi var. compactum Wollenw., Fusaria 4–7 μm in F. camptoceras, 17–41 × 3–6 μm in F. neosemitectum)
Autogr. Delin. 3: no. 924. 1930. and the number of conidial septa (1–6-septate in F. fecundum vs
Synonyms: see Crous et al. (2021). 2–7-septate in F. kotabaruense, 0–7-septate in F. camptoceras,
1–5-septate in F. neosemitectum) (Marasas et al. 1998, Maryani et
Material examined: (See Supplementary Table S1). al. 2019b, Xia et al. 2019).

Notes: The phylogenetic relationship of Fusarium compactum Fusarium guilinense M.M. Wang et al., Persoonia 43: 80. 2019.
was revealed by O’Donnell et al. (2009), which represents a well- New synonym: Fusarium bubalinum J.W. Xia et al., Persoonia 43:
supported phylogenetic species (FIESC 3). Later, Xia et al. (2019) 199. 2019.
designated an epitype for F. compactum to stabilise the use of
this name. To date, this species has been reported from at least Material examined: (See Supplementary Table S1).
18 substrates/hosts, e.g., Austrostipa aristiglumis, Gossypium
barbadense, Triticum aestivum (Bentley et al. 2007, Tunali et al. Notes: Fusarium guilinense was first reported from China from
2008, Schroers et al. 2011), and maize (this study). leaves of Musa nana (Wang et al. 2019a), and in this study from
rice glumes (Supplementary Table S1). In addition, Xia et al. (2019)
Fusarium fecundum S.L. Han, M.M. Wang & L. Cai, sp. nov. in their analyses, did not include the ex-type isolate of F. guilinense
MycoBank MB 847024. Fig. 21. (also F. caatingaense, F. citri, F. hainanense, F. humuli, F. ipomoeae,
F. irregulare, F. luffae, F. nanum and F. pernambucanum), thus could
Etymology: Refers to its high fecundity of aerial conidia in CLA. not reflect on species relationships in this group. Here our multi-
locus phylogenetic analyses clearly showed that “F. bubalinum”
Sporodochia not observed. Aerial conidiophores borne on aerial (Xia et al. 2019) clustered within the F. guilinense clade (Fig. 9).
mycelium, straight or flexuous, erect or prostrate, smooth- and Therefore, we consider F. bubalinum as a synonym of F. guilinense.

www.studiesinmycology.org 129
Han et al.

Fig. 21. Morphology of Fusarium fecundum (CGMCC3.23516, ex-type culture). A. Colony on PDA. B. Colony on OA. C–E. Aerial conidiophores and
phialides. F. Aerial conidia. Scale bars = 10 μm.

Fusarium hainanense M.M. Wang et al., Persoonia 43: 82. 2019. Fusarium humuli M.M. Wang et al., Persoonia 43: 83. 2019.

Material examined: (See Supplementary Table S1). Material examined: (See Table Supplementary S1).

Notes: Fusarium hainanense was first reported from Musa nana Notes: Fusarium humuli was first reported from China on 12
and Oryza sp. (Wang et al. 2019a), and subsequently from Acacia different hosts, including Humulus scandens (Wang et al. 2019a),
sp., Musa acuminata, Oryza australiensis (Xia et al. 2019), and and is herein reported from rice and wheat as new host records.
maize (this study). Three isolates (JS21, JS3, JS9) of rice spikelet Phylogenetically, this species clusters in the Incarnatum clade of
rot clustered with the ex-type strain of F. hainanense in our multi- FIESC, closely related to F. citri, F. fasciculatum and F. weifangense
locus analyses (Fig. 9). (Wang et al. 2019a, Xia et al. 2019, this study). Morphologically,

130
Fusarium spp. on cereals in China

these four species could be distinguished based on the morphology conidia (4–6-septate, dorsiventral curvature conidia in F. jinanense
of their sporodochial conidia (Wang et al. 2019a, Xia et al. 2019, vs 2–4-septate, typical Fusarium conidia in F. lacertarum); the
this study). presence of sporodochia (which are absent in F. lacertarum),
and the absence of aerial conidia (present in F. lacertarum)
Fusarium ipomoeae M.M. Wang et al., Persoonia 43: 83. 2019. (Subrahmanyam 1983, Leslie & Summerell 2006). Pathogenicity
tests confirmed its ability to cause maize stalk rot (Fig. 7L).
Material examined: (See Table Supplementary S1).
Fusarium mianyangense S.L. Han, M.M. Wang & L. Cai, sp. nov.
Notes: Fusarium ipomoeae has been isolated from multiple hosts MycoBank MB 847026. Fig. 23.
(Wang et al. 2019a, Xia et al. 2019, Zhou et al. 2020, Xu et al.
2021b). Notably, one strain of this species has been reported as Etymology: Named after the city, Mianyang, where the holotype
a pathogen of maize but previously misidentified as F. equiseti was collected.
through tef1 BLASTn in GenBank (Li et al. 2014, Fig. 9). Here,
we isolated F. ipomoeae from maize stalks rot (Fig. 2E), maize ear Sporodochia cream (4A3), formed frequently on carnation leaves,
rot (Fig. 1A), wheat scab (Fig. 3D) and rice spikelet rot (Fig. 5E), and infrequently on agar. Sporodochial conidiophores densely
updating two new host (maize and wheat) records (Table 2), and and irregularly branched, bearing apical whorls of 3–5 phialides.
confirmed its pathogenicity to maize (Fig. 7K). Sporodochial phialides subulate to subcylindrical, smooth, thin-
walled, 7.5–13.6 × 1.7–3.8 μm. Sporodochial conidia falcate, the
Fusarium jinanense S.L. Han, M.M. Wang & L. Cai, sp. nov. dorsal side curved than the ventral, with a blunt apical cell and
MycoBank MB 847025. Fig. 22. well-developed foot-shaped basal cell, hyaline, smooth- and thin-
walled, 3(–5)-septate: 3-septate conidia: 24.5–29.9 × 2.8–4.7 μm
Etymology: Named after the city, Jinan, where the holotype was (av. ± SD: 27.3 ± 1.6 × 3.6 ± 0.4 μm); 4-septate conidia: 27.6–34.1
collected. × 2.5–4.5 μm (av. ± SD: 30.9 ± 1.8 × 3.6 ± 1.4 μm); 5-septate
conidia: 29.5–36.6 × 3.0–4.9 μm (av. ± SD: 33.0 ± 2.3 × 3.9 ±
Sporodochia melon (5A6), formed infrequently on carnation leaves 0.5 μm). Aerial conidia not observed. Chlamydospores abundant,
or agar. Sporodochial conidiophores densely and irregularly globose to subglobose, subhyaline, smooth or rough-walled,
branched, bearing apical whorls of 3–6 phialides. Sporodochial intercalary, solitary or forming long chains, 6.0–14.0 μm diam.
phialides subulate to subcylindrical, smooth, thin-walled, 8–14
× 2.5–4 μm. Sporodochial conidia falcate, curved dorsoventrally, Culture characteristics: Colonies on PDA incubated at 25 °C
tapering towards both ends, with an elongated or whip-like in the dark reaching 74–80 mm diam in 7 d; flat, felty to velvety,
curved apical cell and a well-developed to elongate foot-shaped with abundant aerial mycelium, colony margin entire; surface
basal cell, hyaline, thin- and smooth-walled, (4–)5(–6)-septate: white; reverse orange white (5A2) in the centre, white (A1) at the
4-septate conidia: 34.5–43.5 × 3.7–4.7 μm (av. ± SD: 34.6 ± 0.1 margin; odour absent. On OA in the dark reaching 73–80 mm in 7
× 3.9 ± 0.2 μm); 5-septate conidia: 50–58 × 3.9–5.2 μm (av. ± d; convex, with abundant aerial mycelium, margin entire; surface
SD: 54.5 ± 3.5 × 4.6 ± 0.6 μm); 6-septate conidia: 58.7–63.8 × white, reverse apricot (5B6) in the centre, cream (4A3) at the
4.1–4.5 μm (av. ± SD: 61.3 ± 2.8 × 4.2 ± 0.2 μm). Aerial conidia margin; odour absent.
not observed. Chlamydospores abundant, globose, subglobose to
ovoid, subhyaline, smooth or rough-walled, terminal or intercalary, Typus: China, Sichuan Province, Mianyang City (E105.15, N31.78),
solitary, in pairs or forming long chains, 6.2–10.3 μm diam. obtained from the symptomatic tissues of Oryza sativa spikelet rot, 4
Oct. 2020, M.L. Feng (holotype HMAS 351952, ex-type living culture
Culture characteristics: Colonies on PDA incubated at 25 °C in the CGMCC3.23520 = LC15879 = HSL859).
dark reaching 67–73 mm diam in 7 d; flat, felty to velvety, radiate,
with abundant aerial mycelium, colony margin undulate; surface Additional materials examined: (See Supplementary Table S1).
cream (4A3) in the centre, white (–A1) at the margin; reverse
flame yellow (4A8) in the centre, white (–A1) at the margin; odour Notes: Isolates representing F. mianyangense were resolved
absent. On OA in the dark reaching 98–101 mm in 7 d; flat, felty to as a strongly supported genealogically exclusive lineage in the
dusty, with abundant aerial mycelium, margin entire; surface white, phylogenies inferred from combined CaM, rpb2, and tef1 loci (Fig.
reverse curry yellow (4C8) in the centre, cream (4A3) at the margin; 9). Fusarium mianyangense is closely related to F. citrullicola,
odour absent. but differs by 15 bp in the three loci dataset. Morphologically,
they are distinguished from each other in the number of septa
Typus: China, Shandong Province, Jinan City (E117.1; N36.4), obtained in sporodochial conidia (3–5-septate in F. mianyangense vs
from the symptomatic tissues of maize ear rot, 29 Sep. 2020, X.Y. Liu 1–5-septate in F. citrullicola) and the size of sporodochial conidia
(holotype HMAS 351951, ex-type living culture CGMCC3.23519 = (24.5–36.6 × 2.5–4.9 μm in F. mianyangense vs 8–39 × 2–4.9 μm
LC15878 = HSL751). in F. citrullicola) (Khuna et al. 2022). Pathogenicity tests fulfilling
the Koch’s postulates confirmed its ability to cause maize stalk rot
Additional materials examined: (See Supplementary Table S1). (Fig. 7N).

Notes: The isolates representing F. jinanense were resolved Fusarium luffae M.M. Wang et al., Persoonia 43: 85. 2019.
as a strongly supported genealogically exclusive lineage in the
phylogeny inferred from combined CaM, rpb2, and tef1 loci (Fig. 9). Material examined: (See Supplementary Table S1).
Fusarium jinanense is closely related to F. lacertarum, but differs
by 14 bp in the three loci dataset. Morphologically, F. jinanense Notes: Fusarium luffae has been reported as endophyte of Luffa
is distinguished from F. lacertarum based on the morphologies of aegyptiaca, Humulus scandens and Setaria verticillata in previous

www.studiesinmycology.org 131
Han et al.

Fig. 22. Morphology of Fusarium jinanense (CGMCC3.23519, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D, E. Sporodochial
conidiophores and phialides. F. Sporodochial conidia. G. Chlamydospores. Scale bars = 10 μm.

studies (Wang et al. 2019a, Xia et al. 2019). In addition, one tef1) phylogenetic analyses (Fig. 9). We herein confirmed the
pathogenic strain isolated from maize stalk rot was previously pathogenicity of F. luffae (Fig. 7M) and confirmed three new host
misidentified as F. incarnatum through tef1 BLASTn (Gai et al. records (maize, rice and wheat).
2016), but it clustered with F. luffae in our multi-locus (CaM-rpb2-

132
Fusarium spp. on cereals in China

Fig. 23. Morphology of Fusarium mianyangense (CGMCC3.23520, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D, E. Sporodochial
conidiophores and conidiogenous cells. F. Sporodochial conidia. G. Chlamydospores. Scale bars = 10 μm.

Fusarium nanum M.M. Wang et al., Persoonia 43: 85. 2019. includes 19 species, with 16 of them reported from cereals (Wang
et al. 2019a, Xia et al. 2019, this study).
Material examined: (See Supplementary Table S1).
Fusarium nothincarnatum S.L. Han, M.M. Wang & L. Cai, sp.
Notes: Fusarium nanum appears to be a broad host range nov. MycoBank MB 847027. Fig. 24.
species, which has been reported from Musa nana, Oryza sp.,
Solanum lycopersicum, Glycine max, Musa nana, Triticum sp., Etymology: Noth = nothus in Greek, fake, close but different;
maize and wheat (Wang et al. 2019a, Xia et al. 2019, this study). incarnatum = incarnatum-like morphology.
Phylogenetically, it clustered in the Incarnatum clade, which

www.studiesinmycology.org 133
Han et al.

Fig. 24. Morphology of Fusarium nothincarnatum (CGMCC3.24286, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D. Sporodochial
conidiophores and conidiogenous cells. E. Sporodochial conidia. F. Aerial conidiophores and phialides. G. Aerial conidia. H. Chlamydospores. Scale bars
= 10 μm.

Sporodochia cream (4A3), formed abundantly on carnation leaves cell, hyaline, thin- and smooth-walled, (0–)3–5-septate; 0-septate
and the surface of the medium. Sporodochial conidiophores densely conidia: 8.7–23.3 × 1.9–2.9 μm (av. ± SD: 16.5 ± 6.3 × 2.5 ± 0.4 μm),
and irregularly branched, bearing apical whorls of 3–5 phialides. 1-septate conidia: 14.6–19.0 × 2.3–2.8 μm (av. ± SD: 16.9 ± 1.7 × 2.6
Sporodochial phialides subulate to subcylindrical, with a short- ± 0.2 μm), 2-septate conidia: 13.6–26.4 × 2.6–4.0 μm (av. ± SD: 20.1
flared apical collarette, smooth, thin-walled, 6.1–17.8 × 2.3–4.0 μm. ± 3.9 × 3.0 ± 0.4 μm), 3-septate conidia: 25.8–37.7 × 3.3–4.8 μm
Sporodochial conidia falcate, the dorsal side curved than the ventral, (av. ± SD: 32.9 ± 3.3 × 3.9 ± 0.4 μm), 4-septate conidia: 34.7–41.8 ×
with a curved apical cell and poorly-developed foot-shaped basal 3.2–4.6 μm (av. ± SD: 38.2 ± 2.1 × 4.1 ± 0.3 μm), 5-septate conidia:

134
Fusarium spp. on cereals in China

32.5–41.5 × 2.9–4.9 μm (av. ± SD: 37.3 ± 2.2 × 3.7 ± 0.4 μm). 9). Therefore, we consider F. melonis as a later synonym of F.
Aerial conidiophores borne on aerial mycelium, straight or flexuous, pernambucanum.
erect or prostrate, smooth- and thin-walled, unbranched, sympodial
or irregularly branched, bearing terminal or lateral phialides; Aerial Fusarium sulawesiense Maryani et al. (as ‘sulawense’), Persoonia
phialides monophialidic, subulate to subcylindrical, smooth- and thin- 43: 65. 2019.
walled, periclinal thickening inconspicuous or absent, sometimes
proliferating percurrently, 8.2–21.6 × 1.4–3.5 μm. Aerial conidia Material examined: (See Supplementary Table S1).
falcate, curved dorsiventrally, with curved apical cell and blunt
to barely notched basal cell, hyaline, smooth- and thin-walled, Notes: Fusarium sulawesiense was originally reported from Musa
3(–5)-septate: 3-septate conidia: 17.7–34.8 × 3.0–4.5 μm (av. ± SD: acuminata (Maryani et al. 2019b), and subsequently from rice
26.4 ± 4.2 × 3.7 ± 0.4 μm); 4-septate conidia: 30.6–39.9 × 3.2–7.4 μm (Wang et al. 2019a). This study expands its host range (maize and
(av. ± SD: 35.9 ± 2.4 × 4.0 ± 0.7 μm); 5-septate conidia: 32.7–42.4 × wheat) with confirmed pathogenicity (Fig. 7T).
3.2–4.5 μm (av. ± SD: 36.2 ± 3.0 × 3.7 ± 0.4 μm). Chlamydospores
abundant, globose, subglobose to ovoid, subhyaline, smooth-walled, Fusarium tanahbumbuense Maryani et al., Persoonia 43: 63.
2019.
intercalary, solitary, in pairs or forming chains, 7.3–12.4 μm diam.
Material examined: (See Table S1).
Culture characteristics: Colonies on PDA incubated at 25 °C in
the dark reaching 77–79 mm diam in 7 d; flat, felty to velvety, with
Notes: Fusarium tanahbumbuense was first obtained from infected
abundant aerial mycelium, colony margin entire; surface white (A1);
pseudostem of Musa sp. var. Pisang Hawa (ABB), but when
reverse orange white (5A2) in the centre, white (A1) at the margin;
conducting pathogenicity trials, it only caused a slight discoloration
odour absent. On OA in the dark reaching 73–76 mm in 7 d; raised,
in the corm without further disease development (Maryani et al.
with relatively sparse aerial mycelium, margin entire; surface white
2019b). Later, two isolates of this species were recorded as a
(A1), reverse milk white (1A2); odour absent.
pathogen of wheat, but misidentified as F. incarnatum based on
Typus: China, Heilongjiang Province, Daqing City (E124.81, N46.04), tef1 phylogenetic analysis (Özer et al. 2020, Fig. 9). Here we add
obtained from the symptomatic tissues of Oryza sativa spikelet rot, two new host records (maize and rice), and confirm its ability to
7 Oct. 2020, Y.J. Li (holotype HMAS 352343, ex-type living culture cause maize stalk rot (Fig. 7U).
CGMCC3.24286 = LC18436 = HSL221).
Fusarium weifangense S.L. Han, M.M. Wang & L. Cai, sp. nov.
Additional materials examined: (See Supplementary Table S1). MycoBank MB 847028. Fig. 25.

Notes: Isolates representing F. nothincarnatum were resolved Etymology: Named after the city, Weifang, where the holotype was
as a strongly supported genealogically exclusive lineage in the collected.
phylogenies inferred from combined CaM, rpb2, and tef1 (Fig.
9), and genomic datasets (Fig. 14). Fusarium nothincarnatum is Sporodochia milk white (1A2), formed frequently on carnation leaves.
closely related to F. incarnatum, but differs by 23 bp in the three Sporodochial conidiophores densely and irregularly branched,
loci dataset. Morphologically, F. nothincarnatum is distinguished bearing apical whorls of 3–4 phialides. Sporodochial phialides
from F. incarnatum based on the number of septa in sporodochial subulate to subcylindrical, smooth, thin-walled, 8.8–29.9 × 3.1–5.1
conidia (0–5-septate in F. nothincarnatum vs 1–6-septate in F. μm. Sporodochial conidia falcate, the dorsal side curved than the
incarnatum); the morphology of sporodochial conidia (the dorsal ventral, with a blunt apical cell and poorly-developed foot-shaped
side more curved than the ventral, with more curved apical cell in basal cell, hyaline, smooth- and thin-walled, (3–)5(–7)-septate:
F. nothincarnatum); the type of aerial phialides (monophialides in F. 3-septate conidia: 26.5–30.9 × 4.5–5.8 μm (av. ± SD: 28.8 ± 1.7 × 4.9
nothincarnatum vs mono- and polyphialides in F. incarnatum); and ± 0.4 μm); 4-septate conidia: 28.0–35.4 × 4.1–5.5 μm (av. ± SD: 31.0
the morphologies of aerial conidia (F. nothincarnatum produces ± 2.1 × 4.7 ± 0.3 μm); 5-septate conidia: 30.0–49.4 × 4.9–6.7 μm
only falcate aerial conidia with 3–5 septa; while two types of aerial (av. ± SD: 37.2 ± 4.3 × 5.6 ± 0.4 μm); 6-septate conidia: 37.7–49.2 ×
conidia were found in F. incarnatum: ellipsoidal to fusoid-shaped 5.1–7.1 μm (av. ± SD: 44.5 ± 3.7 × 5.9 ± 0.7 μm); 7-septate conidia:
with 0–3 septa, and falcate shaped with 1–7 septa) (Xia et al. 2019). 35.5–36 × 4.8–5.6 μm (av. ± SD: 35.7 ± 0.3 × 5.2 ± 0.6 μm). Aerial
conidia and chlamydospores not observed.
Fusarium pernambucanum A.C.S. Santos et al., Mycologia 111:
253. 2019. Culture characteristics: Colonies on PDA incubated at 25 °C
New synonym: Fusarium melonis S. Khuna et al., J. Fungi 8: 1135. in the dark reaching 69–79 mm diam in 7 d; flat, felty to velvety,
2022, nom. inval., Art. 40.8 (Shenzen). with abundant aerial mycelium, colony margin entire; surface
white (A1); reverse milk white (1A2); odour absent. On OA in the
Material examined: (See Supplementary Table S1).
dark reaching 76–83 mm in 7 d; crateriform, with abundant aerial
mycelium, margin entire; surface white (A1), reverse cream (4A3);
Notes: Fusarium pernambucanum was originally reported from
odour absent.
Aleurocanthus woglumi (Santos et al. 2019), and subsequently from
Cucumis melo (Medeiros Araújo et al. 2020). This study expands Typus: China, Shandong Province, Weifang City (E119.79, N36.23), from
its host range (maize) with confirmed pathogenicity (Fig. 7P). In the symptomatic tissues of wheat scab, 18 May 2021, P. Shen (holotype
addition, the inclusion of a larger sampling of F. pernambucanum HMAS 352342, ex-type living culture CGMCC3.24285 = LC18333 =
isolates for the multi-locus phylogenetic analysis clearly showed HSL1800).
that the ex-type isolate (SDBR-CMU424) of F. melonis (Khuna
et al. 2022) clustered within the F. pernambucanum clade (Fig. Additional materials examined: (See Supplementary Table S1).

www.studiesinmycology.org 135
Han et al.

Fig. 25. Morphology of Fusarium weifangense (CGMCC3.24285, ex-type culture). A. Colony on PDA. B. Colony on OA. C. Sporodochia. D, E. Sporodochial
conidiophores and conidiogenous cells. F. Sporodochial conidia. Scale bars = 10 μm.

Notes: The isolates representing F. weifangense were resolved latter two species in the number of septa in sporodochial conidia
as a strongly supported genealogically exclusive lineage in the (3–7-septate in F. weifangense vs 3–5-septate in F. citri and F.
phylogeny inferred from combined CaM, rpb2, and tef1 loci (Fig. humuli), the size of sporodochial conidia (26.5–49.4 × 4.1–7.1 μm
9). Fusarium weifangense is closely related to F. citri and F. humuli, in F. weifangense vs 25.5–40.5 × 3–5.5 μm in F. citri vs 21–35 ×
but differs by 37 bp and 51 bp in the three loci dataset, respectively. 2–3 μm in F. humuli) (Wang et al. 2019a).
Morphologically, F. weifangense could be distinguished from the

136
Fusarium spp. on cereals in China

Fusarium nisikadoi species complex (FNSC) Fusarium nirenbergiae L. Lombard & Crous, Persoonia 43: 29.
2018 [2019].
FNSC now includes six species (F. commune, F. gaditjirrii, F.
lyarnte, F. miscanthi, F. nisikadoi and F. paranisikadoi), all of which Material examined: (See Supplementary Table S1).
are associated with plants (Nirenberg 1997, Gams et al. 1999,
Skovgaard et al. 2003, Phan et al. 2004, Crous et al. 2021, Wang Notes: Fusarium nirenbergiae has been isolated from at least 20
et al. 2022a). In this study, all six obtained strains belonging to host genera (Lombard et al. 2019b, Wang et al. 2022a), and has
this complex were identified as F. commune based on multi-locus been recorded to contain isolates which were previously classified
analyses (Fig. 10).
in 14 different forma specialis, e.g., F. oxysporum f. sp. bouvardiae,
F. oxysporum f. sp. cubense, F. oxysporum f. sp. lycopersici, etc.
Fusarium commune K. Skovg. et al., Mycologia 95: 632. 2003.
(Lombard et al. 2019b, McTaggart et al. 2021). In this study, we
Material examined: (See Supplementary Table S1). reported a new host record of F. nirenbergiae and confirmed its
pathogenicity (Fig. 7O).
Notes: Fusarium commune has been reported as pathogens of rice
wilt, rice root rot, and maize stalk rot by Husna et al. (2020) and Fusarium sambucinum species complex (FSAMSC)
Xi et al. (2019). This species is pathologically similar to species in
FOSC, causing root rot and wilt diseases on a variety of plants (Ma According to the multi-locus phylogenetic (Fig. 12) and
et al. 2010, Maryani et al. 2019a, McTaggart et al. 2021). phylogenomic support (Fig. 15), the FSAMSC now includes 41
described species (Crous et al. 2021, Laraba et al. 2021, Lombard
Fusarium oxysporum species complex (FOSC) et al. 2022), and almost half of the species in this complex have
been reported from cereals (Table 2). The 205 isolates obtained in
Members of FOSC are mostly soil-borne fungi causing root and this study were identified as nine known species (Fig. 12).
vascular wilt diseases (Ma et al. 2010, Maryani et al. 2019a,
Fusarium armeniacum (G.A. Forbes et al.) L.W. Burgess &
McTaggart et al. 2021). Previously, the FOSC has been classified Summerell, Stud. Mycol. 98: 86. 2021.
as one species which included more than 100 formae speciales, Basionym: Fusarium acuminatum subsp. armeniacum G.A. Forbes
based on subspecific classification systems (Snyder & Hansen et al., Mycologia 85: 120. 1993.
1940, Gordon 1965, Lombard et al. 2019b). The F. oxysporum
strains belonging to the same forma specialis, however, have been Material examined: (See Supplementary Table S1).
shown to represent multiple species based on the multi-locus and
phylogenomic analyses (Zhang & Ma 2017, Lombard et al. 2019b, Notes: Fusarium armeniacum has been recorded as causal agent
McTaggart et al. 2021). For instance, Fusarium oxysporum f. sp. of wheat scab, root rot and seed rot of various plants (Wing et al.
lycopersici, has been shown to be an assemblage of two distinct 1993, Leslie & Summerell 2006, Summerell et al. 2010, Krone et
species, namely F. nirenbergiae and F. languescens (Lombard et al. 2020). Here we update two new host records (maize and wheat)
al. 2019b). With the continuous clarification of the phylogenetic in China, and confirm its ability to cause maize stalk rot (Fig. 7G).
relationships in this species complex, up to 31 cryptic species
have been recognised and formally named (Lombard et al. 2019b, Fusarium asiaticum O’Donnell et al., Fungal Genet. Biol. 41: 619.
Maryani et al. 2019a, Wang et al. 2022a). Recently, population 2004.
genomics analyses based on 410 genomes within the FOSC
suggested that long-term sexual or parasexual reproduction have Material examined: (See Supplementary Table S1).
contributed to the species diversity of this complex, stepping away
from the previous hypothesis suggesting strictly clonal propagation Notes: Fusarium asiaticum is one of the main pathogens of wheat
(McTaggart et al. 2021). scab in the world, especially in Asia (Suga et al. 2008, Zhang et al.
2012, Castañares et al. 2016), and it is also a notorious pathogen
Fusarium cugenangense Maryani et al., Stud. Mycol. 92: 181. of wheat crown rot, maize ear rot, maize stalk rot and rice spikelet
2018 [2019]. rot (Desjardins & Proctor 2011, Gomes et al. 2015, Zhang et al.
2016, Dong et al. 2020b, Xi et al. 2021). In our study, most strains
Material examined: (See Supplementary Table S1). of this species were isolated from blighted wheat spikes (wheat
scab) from southern China, supporting earlier findings that it is a
Notes: The banana Fusarium wilt pathogen, previously called F. dominant agent of wheat scab in China (Zhang et al. 2012, Yang et
oxysporum f. sp. cubense has been confirmed to include nine al. 2018, Xu et al. 2021a).
independent phylogenetic lineages (Maryani et al. 2019a). Among
them, the lineage referred to as Foc Lineage L7 by Fourie et al. Fusarium boothii O’Donnell et al., Fungal Genetics Biol. 41: 618.
(2009), was recently formally described as a new species Fusarium 2004.
cugenangense (Maryani et al. 2019a). According to multi-locus
analyses of available data, F. cugenangense appears to have a very Material examined: (See Supplementary Table S1).
wide host range, including Acer palmatum, Crocus sp., Gossypium
barbadense, Hordeum vulgare, Solanum tuberosum, Smilax sp., Notes: Fusarium boothii has been reported as a pathogen of wheat
Tulipa gesneriana, Musa nana, Musa sp., Vicia faba and Zea mays scab, maize ear rot and stalk rot (Malihipour et al. 2012, Zhang
(Maryani et al. 2019a, Wang et al. 2022a, this study). et al. 2016, Beukes et al. 2017, Gai et al. 2017). The hybrid of F.
graminearum × F. boothii has been reported from maize in France
and South Africa (Boutigny et al. 2011, 2013). Future comparative
genomic analyses of F. graminearum, F. boothii, F. graminearum ×

www.studiesinmycology.org 137
Han et al.

F. boothii would be useful for understanding the speciation of these Synonyms: see Crous et al. (2021).
closely related species.
Material examined: (See Supplementary Table S1).
Fusarium graminearum Schwabe, Fl. Anhalt. 2: 285. 1839.
Synonyms: see Crous et al. (2021). Notes: Fusarium pseudograminearum is a well-known mycotoxin
(3-acetyldeoxynivalenol, zearalenone, nivalenol and culmorin)
Material examined: (See Supplementary Table S1). producing fungus (Farr & Rossman 2022, Laraba et al. 2021). It
is also a major causal agent of wheat crown rot in Northern China
Notes: Fusarium graminearum, with the smallest genome in the (Li et al. 2012, Xu et al. 2015), which was further confirmed in our
whole genus (Supplementary Table S3), has been considered as study (Supplementary Table S1).
one of the top 10 most economically harmful fungal pathogens
(Dean et al. 2012). This species is widely distributed on different Fusarium vorosii B. Tóth et al., Fungal Genet. Biol. 44: 1202.
hosts with different lifestyles (pathogens, endophytes, saprophytes) 2007.
(Cuomo et al. 2007, Starkey et al. 2007, Talas & McDonald
2015, Lofgren et al. 2018, Sarowar et al. 2019). In our study, F. Material examined: (See Supplementary Table S1).
graminearum was also one of the most frequently isolated species
(85 strains) and, in many cases, co-occurred with other pathogens Notes: Fusarium vorosii was first described as a pathogen of wheat
in the same diseased plant, leading to co-infection (Supplementary scab by Starkey et al. (2007). Later, it has been isolated from other
Table S6). cereals, including Hordeum vulgare, Oryza sativa and Zea mays
(Lee et al. 2016). To date, this species has been reported from
Fusarium kyushuense O’Donnell & T. Aoki, Mycoscience 39: 2. five countries (Farr & Rossman 2022), including Hungary, Serbia,
1998. Japan, Korea and China (this study).

Material examined: (See Supplementary Table S1). Fusarium tricinctum species complex (FTSC)

Notes: Fusarium kyushuense clustered within the Sambucinum Species in FTSC are widely distributed in various grains and
clade, closely related to F. venenatum and F. poae (Fig. 12), all of produce a broad range of mycotoxins (Leslie & Summerell 2006).
which have been reported as cereal pathogens (Table 2). In China, Currently, 14 cryptic species have been identified and described in
F. kyushuense has been reported as pathogen causing maize ear this complex (Wang et al. 2022a), among which F. acuminatum and
and stalk rot (Wang et al. 2014, Cao et al. 2021). Here we report F. F. avenaceum have the widest host range (more than 250 different
kyushuense for the first time from diseased wheat (Fig. 3C). hosts). Coincidentally, all 20 strains of this complex isolated in this
study were identified as F. acuminatum or F. avenaceum (Fig. 13).
Fusarium meridionale T. Aoki et al., Fungal Genet. Biol. 41: 618.
2004. Fusarium acuminatum Ellis & Everh., Proc. Acad. Nat. Sci.
Philadelphia 47: 441. 1895.
Material examined: (See Supplementary Table S1). Synonyms: see Crous et al. (2021).

Notes: Fusarium meridionale was originally known from the Material examined: (See Supplementary Table S1).
Southern Hemisphere, but is now known to have a global distribution
(O’Donnell et al. 2004, Farr & Rossman 2022). To date, this species Notes: Fusarium acuminatum was usually reported from temperate
has been reported as pathogens of maize ear rot, maize stalk rot, regions as a soil saprobe or pathogen of root and crown diseases
wheat scab and rice spikelet rot (Zhang et al. 2014b, 2016, Dong of various plants (Leslie & Summerell 2006, Zhang et al. 2015, this
et al. 2020a). Herein we newly record this species from wheat in study: Fig. 7D). In this study, eight strains isolated from diseased
China. cereals clustered with the currently widely used reference strain of
Fusarium acuminatum (NRRL 52789) (O’Donnell et al. 2012, Li et
Fusarium poae (Peck) Wollenw., in Lewis, Bull. Maine. Agric. Exp. al. 2019a, Wang et al. 2022a).
Sta. 219: 254. 1913 [1914].
Basionym: Sporotrichum poae Peck, Bull. New York State Mus. 67: Fusarium avenaceum (Fr.) Sacc., Syll. Fung. 4: 713. 1886.
29. 1904 [1903]. Basionym: Fusisporium avenaceum Fr., Syst. Mycol. 2: 238. 1822,
Synonyms: see Crous et al. (2021). nom. sanct. [Fr., l.c.].
Synonyms: see Crous et al. (2021).
Material examined: (See Supplementary Table S1).
Material examined: (See Supplementary Table S1).
Notes: Fusarium poae produces a series of mycotoxins, such as
diacetoxyscirpenol, beauvericin and nivalenol (Laraba et al. 2021). Notes: Fusarium avenaceum is a cosmopolitan species and has
This species has been widely recorded from cereals, but appears been recorded from more than 250 hosts (Leslie & Summerell
to be pathologically less aggressive compared with F. graminearum 2006, Ma et al. 2019, Crous et al. 2021, Farr & Rossman 2022). On
(Bottalico & Perrone 2002, Xu et al. 2020, Laraba et al. 2021). In cereals it has been associated with maize ear rot, wheat scab and
our investigation, F. poae appeared to be more frequently recorded wheat crown rot (Logrieco et al. 2002b, Vogelgsang et al. 2008a, b,
from alpine and dry areas (Fig. 16C; Supplementary Table S1). Zhang et al. 2015), but due to the lack of sequence data we were
unable to confirm the species identify in previous studies. In this
Fusarium pseudograminearum O’Donnell & T. Aoki, Mycologia investigation, 12 representative strains from wheat scab clustered
91: 604. 1999. together with the ex-neotype of F. avenaceum (Crous et al. 2021).

138
Fusarium spp. on cereals in China

DISCUSSION Co-infections of Fusarium species on cereals

“Broad” and “narrow” generic concept of Fusarium In the present study, there is a notable phenomenon that various
Fusarium species were isolated from the same cereal disease,
Fusarium species have attracted much attention from plant and these co-infection samples account for 56 % of the whole 315
pathologists, food chemists and taxonomists because of their samples (Fig. 17, Supplementary Table S6). Almost all species
widespread distribution, pathogenicity and mycotoxin production isolated from co-infection samples have been previously reported
(Marasas et al. 1984, Aoki et al. 2012, Yang et al. 2018, Torbati as pathogens (Table 2, Supplementary Table S6), and notably, it
et al. 2019, Crous et al. 2021, Tralamazza et al. 2021). However, is impossible to distinguish the co-infection species only through
especially in the past decade, there have been several controversial the disease symptoms on the host. In an extreme example of
proposals on whether Fusarium should remain broadly defined to this study, four different species (i.e., F. luffae, F. sulawesiense,
keep most of the well-known pathogens under the generic name F. tanahbumbuense, and F. verticillioides) were isolated from
Fusarium, or circumscribe Fusarium in a narrower sense, but the same diseased maize ear sample (Fig. 1F) collected from
monophyletic, morphologically meaningful and as a definable Lianyungang city (E119.09, N34.5), Jiangsu Province. This type
unit (Crous et al. 2021, Geiser et al. 2021). The “broad” concept of co-infection (or mixed infection) phenomenon has been noted
emphasised that the “terminal Fusarium clade” (F1) is monophyletic and discussed in previous plant or human diseases (Waner 1994,
(Fig. 14), with the same synapomorphy (fusarium-like macroconidia) Tarashi et al. 2017, Strauss et al. 2021, Wang et al. 2022b, Zhang et
al. 2022). Using wheat scab as an example, a recent study showed
(Geiser et al. 2021). However, fusarium-like asexual macroconidia
that multiple Fusarium species have been isolated from the same
occur in many other genera (e.g., Atractium, Fusicolla, Macroconia)
diseased wheat head samples (Wang et al. 2022b).
outside the “terminal Fusarium clade” (Crous et al. 2021). Under
Previous studies suggested that pathogens occur more
the “narrow” concept, a set of synapomorphies on both sexual and
frequently in an antagonistic or synergistic manner (van Kan et al.
asexual characters could be used to distinguish the genus Fusarium
2014, Bashyal et al. 2016, Gao et al. 2021, Mesny et al. 2021,
from allied genera (Crous et al. 2021). Moreover, recent genomic
de Faria Ferreira 2022). For instance, both F. aglaonematis
analyses have revealed distinct divergence patterns between
and F. elaeidis could cause stem rot in Aglaonema modestum
Fusarium s. str. and other fusarioid taxa in natural selection,
independently, but co-infection enhanced plant disease severity
translational selection and codon usage bias, which further support
(Zhang et al. 2022). In contrast, F. graminearum causes more
the “narrow” Fusarium concept (Hill et al. 2022).
serious wheat scab than co-infection of F. graminearum and F.
In our phylogenomic tree, 193 nodes (86 %) received 100 %
poae (Tan et al. 2020). The concepts of the pathobiome appear to
bootstrap support, including the nodes F1, F2, and F3 (Fig. 14).
provide a more holistic perspective to understand the occurrence
Thus, we calculated concordance factors (gCF and sCF) as
of diseases, rather than simply employing the “one pathogen–
complementary information which may help to improve the accuracy one disease” theory (Vayssier-Taussat et al. 2014, 2015, Sweet
of our interpretations of phylogenetic reconstructions (Fig. 14, & Bulling 2017, Bass et al. 2019). First, the pathobiome reflects
Supplementary Fig. S10A). As we had previously expected, the a more natural state of interaction between the host and its
gCF and sCF values resolved higher in node F3 (gCF 71.1 %, sCF microbes, as the boundaries between pathogens, endophytes
70.3 %) than in F2 (gCF 46.5 %, sCF 46.7 %) and F1 (gCF 51.3 and saprophytes are somewhat ambiguous and relative (Álvarez-
%, sCF 47.1 %). Notably, low concordance values do not mean Loayza et al. 2011, van Kan et al. 2014, Lofgren et al. 2018, Hill et
that the phylogenetic tree is unresolved, but rather gives us further al. 2022). Many so-called “pathogens” only cause plant symptoms
insights on how related or congruent the genes are in resolving the at higher conidial concentrations (Fig. 7), and in many cases they
species phylogeny (Minh et al. 2020b). In our case, among the 1 001 present different lifestyles on different plants (Varga et al. 2015,
single-copy orthologous gene trees, 712 of them support node F3 Lofgren et al. 2018). Secondly, the co-occurrence of different
(Fusarium s.str.), while only 465 and 514 of them support node F2 and species may promote the emergence of new pathogenic varieties
F1, respectively. Our phylogenomic analyses (Fig. 14) thus further through interspecific hybridization (Boutigny et al. 2011, 2013) or
support the rationality of the “narrow” Fusarium concept (assigning gene transfer (Ma 2014, Vlaardingerbroek et al. 2016, Simbaqueba
Fusarium to node F3) as that of Gräfenhan et al. (2011), Lombard et et al. 2018), thus serve as a potential driver of pathogen evolution.
al. (2015) and Crous et al. (2021). Although the selection of a “narrow” For example, F. nirenbergiae (Fo47) assimilated the pathogenicity
concept has inevitably brought in more name changes, in this scheme genes from F. languescens (Fol4287) via horizontal gene transfer
the well-supported genealogically exclusive Fusarium species were (Vlaardingerbroek et al. 2016, McTaggart et al. 2021). Thirdly, the
clearly associated with synapomorphies, including sexual morphs pathobiome vision provides a new perspective to explain why the
(dark purple to black perithecia and subhyaline, smooth, 1–3-septate higher richness of fungal pathogens is usually associated with a
ascospores), asexual morphs (macroconidia with variously shaped more harmful threat to plant health (Liu et al. 2022c), because
apical and basal cells), and trichothecene mycotoxin production co-infection may enhance the reproduction and transmission
(Lombard et al. 2015, Crous et al. 2021). Despite the sampling still capacity of pathogens (Susi et al. 2015). The co-infection widely
being unbalanced (e.g., few from Bisifusarium, and Rectifusarium), detected in this study suggested that Fusarium pathogens may
our 1 001 homologous loci of 228 assembled genomes provided have close interaction with each other in the process of the disease
much more phylogenetic information and a higher resolution than development but the specific mechanisms in different diseases/
previous multi-locus and genomic data (Crous et al. 2021, Geiser hosts remain to be revealed.
et al. 2021). It is hoped that in the future more type material-derived There is no doubt that co-infection will pose diagnostic
genomes will be sequenced and made publicly available (Crous et challenges for plant pathologists and quarantine officials. The
al. 2022), in turn establishing more robust and natural classifications, traditional isolation and culture-based protocols may often
with species and generic names being more meaningful with regards underestimate the diversity of pathogens, ignoring one or more
to their evolutionary relationships and phenotypic traits. other pathogens. Given the vast species number and the difficulty

www.studiesinmycology.org 139
Han et al.

in pathogen recognition, there is an urgent need to develop rapid independent species (O’Donnell et al. 1998b, Fourie et al. 2009,
and efficient diagnosis protocols for the identification of Fusarium Maryani et al. 2019a), and these species are believed to have long
species. The gene chip is likely a useful technical option for the co-evolved with their banana hosts in Asia (Ploetz & Pegg 1997,
recognition of known species (Shi et al. 2012, Yin et al. 2020, Mostert et al. 2017, Maryani et al. 2019a).
Yu et al. 2022), and Ward et al. (2008) has made a successful
attempt in the genus Fusarium. Given the considerable number Checklist of cereal-associated Fusarium species
of Fusarium species causing a same disease as noted in this
study (Supplementary Table S6), more specific primers need to In this study, we updated 39 and 52 host records for Fusarium
be designed in future studies to diagnose these cereal pathogens species for the world and China, respectively, and for the first
more accurately. The high-throughput sequencing is also potentially time (Table 2), confirmed the pathogenicity of 18 species causing
useful but currently, widely used 2nd generation sequencing could maize stalk rot (Fig. 7). According to the updated records presented
only reveal the pathogens at the generic level in most cases. The in this study, about 79 % of the new host records in China were
3rd generation sequencing represented by PacBio technology from new species published since 2019 (Table 2), indicating that
enabling longer reads of sequences would be potentially useful in the description rate of new species in Fusarium has significantly
revealing the diversity of the pathobiome. Meanwhile, considering increased, along with the establishment of the new classification
the complex interactions among pathogenic Fusarium species, it is scheme for the genus. The previous classification was confusing,
necessary to analyse the functional complexities of the pathobiome resulting in many sequences in public databases being assigned
based on genome plasticity, metabolomics and microbial to incorrect names or widely used complex names. For example:
interactions. A better understanding of the Fusarium diversity i) most strains historically identified as F. proliferatum (a well-
and related pathobiome may revolutionize our understanding of known pathogen of maize) in previous studies (Leslie & Summerell
Fusarium pathogenicity and allow us to more effectively prevent 2006, Qiu et al. 2020) have been shown to be F. annulatum based
and control these diseases. on the updated taxonomic system (Yilmaz et al. 2021); ii) many
pathogenic isolates previously recognised as F. incarnatum were
Need for universal barcode re-identified as different species (F. clavus, F. hainanense, F. luffae,
and F. tanahbumbuense) in the current study (Fig. 9). Considering
Based on currently available data (Table 2), the Fusarium diseases the large number of cryptic Fusarium species, there will be more
associated with cereals are mainly caused by members of FFSC, cereal-associated Fusarium species reported in the future. For
FIESC, and FSAMSC. Our results were in agreement with previous instance, there are 74 genealogically exclusive phylogenetically
studies that the combination of CaM, rpb1, rpb2, tef1, and tub2 distinct species in the FSAMSC, whereas only 41 of them were
datasets could recognise all species within the FFSC (Yilmaz et al. formally named to date (Laraba et al. 2021).
2021); and the combination of CaM, rpb2, and tef1 datasets could
recognise all species within the FIESC (Wang et al. 2019a, Xia et
al. 2019). As for the FSAMSC, we recommend to use concatenat-
ed H3-rpb1-rpb2-tef1 loci for species identification. In addition to
ACKNOWLEDGEMENTS
these three species-rich complexes, other cereal-related Fusarium
We are very grateful to all who helped us to contact farmers, collect
species could be discerned by rpb2 and tef1 datasets (Crous et al. diseased samples in their farmland, or provide photographs of diseased
2021, Wang et al. 2022a, this study). As noted above, it is quite cereals (Supplementary Fig. S1). The name of all participants are listed
frustrating that different complexes require different combinations in Supplementary Table S7. Funding for this study has been provided by
of molecular fragments to achieve correct identification, making the National Science and Technology Fundamental Resources Investigation
identification process cumbersome and difficult for trainees and Program of China (2021FY100900) for fungal genome sequencing, NSFC
plant pathologists. Taxonomists should continue to search for more (31725001, 32070023) and Biological Resources Programme, CAS (KFJ-
universal and useful barcodes to facilitate rapid and accurate iden- BRP-009) for fungal diversity survey, the Major Scientific Research Focus
tification of Fusarium species. For the currently widely sequenced Project of Henan Academy of Science (210105002) for academic visiting
of JWW to IMCAS, and the Two Thousand Plan Foundation of Jiangxi
loci, even the most effective species recognition barcode (rpb2)
Province (jxsq2019102026) for LC to visit Jiangxi Agricultural University.
could only recognize 59, 33 and 35 species from the FFSC (with 84
We are very grateful to Prof. P.W. Crous for inspiring our idea of sample
species), FIESC (with 43 species), and FSMSC (with 41 species), collection through a citizen science approach, as well as spending time in
respectively. editing our manuscript.

Host preference of complexes and species DECLARATION ON CONFLICT OF INTEREST


Similar to previous studies, several Fusarium species complexes The authors declare that there is no conflict of interest.
present certain host preferences (Gale et al. 2011, Zhang et al.
2012, 2015, Leyva-Madrigal et al. 2015, Yang et al. 2018, Duan
et al. 2020). For example, in this study, 90 % of diseased wheat
samples were associated with the FSAMSC, while 83 % of diseased
REFERENCES
maize samples were associated with the FFSC (Supplementary
Aamot HU, Ward TJ, Brodal G, et al. (2015). Genetic and phenotypic
Table S1). These host preferences are usually explained as the
diversity within the Fusarium graminearum species complex in
results of the co-evolution of Fusarium species with angiosperms in Norway. European Journal of Plant Pathology 142: 501–519.
warm and humid regions (Leslie & Summerell 2006, Ploetz 2006, Abdallah-Nekache N, Laraba I, Ducos C, et al. (2019). Occurrence of
Przemieniecki et al. 2014, Aamot et al. 2015, Lofgren et al. 2018, Fusarium head blight and Fusarium crown rot in Algerian wheat:
Vorob’eva & Toropova 2020). Taking Fusarium oxysporum f. sp. identification of associated species and assessment of aggressiveness.
cubense as an example, it has been identified and delimited as nine European Journal of Plant Pathology 154: 499–512.

140
Fusarium spp. on cereals in China

Abdul Rahm M, Mohamed Gh K, Essia Reha NA (2020). Prevalence Bottalico A, Perrone G (2002). Toxigenic Fusarium species and mycotoxins
and transmission of seed-borne fungi of maize and their control by associated with head blight in small-grain cereals in Europe. European
phenolic antioxidants. Plant Pathology Journal 19: 176–184. Journal of Plant Pathology 108: 611–624.
Adikaram NKB, Yakandawala DMD (2020). A checklist of plant pathogenic Boutigny AL, Ward TJ, Ballois N, et al. (2013). Diversity of the Fusarium
fungi and Oomycota in Sri Lanka. Ceylon Journal of Science 49: graminearum species complex on French cereals. European Journal
93–123. of Plant Pathology 138: 133–148.
Aguín O, Cao A, Pintos C, et al. (2014). Occurrence of Fusarium species Boutigny AL, Ward TJ, Van Coller GJ, et al. (2011). Analysis of the Fusarium
in maize kernels grown in north western Spain. Plant Pathology 63: graminearum species complex from wheat, barley and maize in South
946–951. Africa provides evidence of species-specific differences in host
Al-Hatmi AM, Bonifaz A, de Hoog GS, et al. (2014). Keratitis by Fusarium preference. Fungal Genetics and Biology 48: 914–920.
temperatum, a novel opportunist. BMC Infectious Diseases 14: 588– Britz H, Steenkamp ET, Coutinho TA, et al. (2002). Two new species of
597. Fusarium section Liseola associated with mango malformation.
Alexey G, Vladislav S, Nikolay V, et al. (2013). QUAST: quality assessment Mycologia 94: 722–730.
tool for genome assemblies. Bioinformatics 29: 1072–1075. Bugnicourt F (1952). Une espèce fusarienne nouvelle, parasite du riz.
Álvarez-Loayza P, White JF Jr, Torres MS, et al. (2011). Light converts Revue Générale de Botanique 59: 13–18.
endosymbiotic fungus to pathogen, influencing seedling survival and Burgess LW, Forbes GA, Windels C, et al. (1993). Characterization and
niche-space filling of a common tropical tree, Iriartea deltoidea. PLoS distribution of Fusarium acuminatum subsp. armeniacum subsp. nov.
ONE 6: e16386. Mycologia 85: 119–124.
Amatulli MT, Spadaro D, Gullino ML, et al. (2010). Molecular identification Burgess LW, Trimboli D (1986). Characterization and distribution of
of Fusarium spp. associated with bakanae disease of rice in Italy and Fusarium nygamai sp. nov. Mycologia 78: 223–229.
assessment of their pathogenicity. Plant Pathology 59: 839–844. Byrnes KJ, Carroll RB (1986). Fungi causing stalk rot of conventional-
Andrews S, Babraham B (2010). FastQC: a quality control tool for high tillage and no-tillage corn in Delaware. Plant Disease 70: 238–239.
throughput sequence data. <http://www.bioinformatics.babraham. Cao Y, Zhang J, Han S, et al. (2021). First report of maize stalk rot caused
ac.uk/projects/fastqc>. by Fusarium kyushuense in China. Plant Disease 105: 3759–3759.
Aoki T, O’Donnell K (1998). Fusarium kyushuense sp. nov. from Japan. Castañares E, Dinolfo MI, Del Ponte EM, et al. (2016). Species composition
Mycoscience 39: 1–6. and genetic structure of Fusarium graminearum species complex
Aoki T, O’Donnell K (1999). Morphological and molecular characterization populations affecting the main barley growing regions of South
of Fusarium pseudograminearum sp. nov., formerly recognized as the America. Plant Pathology 65: 930–939.
Group 1 population of F. graminearum. Mycologia 91: 597–609. Castañares E, Martínez M, Cristos D, et al. (2019). Fusarium species
Aoki T, O’Donnell K, Geiser DM (2014). Systematics of key phytopathogenic and mycotoxin contamination in maize in Buenos Aires Province,
Fusarium species: current status and future challenges. Journal of Argentina. European Journal of Plant Pathology 155: 1265–1275.
General Plant Pathology 80: 189–201. Castañares E, Stenglein SA, Dinolfo MI, et al. (2011). Fusarium tricinctum
Aoki T, O’Donnell K, Ichikawa K (2001). Fusarium fractiflexum sp. nov. associated with head blight on wheat in Argentina. Plant Disease 95:
and two other species within the Gibberella fujikuroi species complex 496–496.
recently discovered in Japan that form aerial conidia in false heads. Castresana J (2000). Selection of conserved blocks from multiple
Mycoscience 42: 461–478. alignments for their use in phylogenetic analysis. Molecular Biology
Aoki T, Ward TJ, Kistler HC, et al. (2012). Systematics, phylogeny and and Evolution 17: 540–552.
trichothecene mycotoxin potential of Fusarium head blight cereal Chehri K, Salleh B, Yli-mattila T, et al. (2010). Occurrence, pathogenicity
pathogens. Mycotoxins 62: 91–102. and distribution of Fusarium spp. in stored wheat seeds Kermanshah
Bacon CW, Hinton DM (1996). Symptomless endophytic colonization Province, Iran. Pakistan Journal of Biological Sciences 13: 1178–1186.
of maize by Fusarium moniliforme. Canadian Journal of Botany 74: Chiang KS, Liu HI, Bock CH (2017). A discussion on disease severity index
1195–1202. values. Part I: warning on inherent errors and suggestions to maximise
Balmas V, Corda P, Marcello A, et al. (2000). Fusarium nygamai associated accuracy. Annals of Applied Biology 171: 139–154.
with Fusarium foot rot of rice in Sardinia. Plant Disease 84: 807–807. Cho WD, Shin HD. (2004). List of plant diseases in Korea. Korean Society
Bankevich A, Nurk S, Antipov D, et al. (2012). SPAdes: a new genome of Plant Pathology, Korea.
assembly algorithm and its applications to single-cell sequencing. Choi HW, Hong SK, Lee YK, et al. (2018). Taxonomy of Fusarium
Journal of Computational Biology 19: 455–477. fujikuroi species complex associated with bakanae on rice in Korea.
Barkat EH, Hardy GESJ, Ren Y, et al. (2016). Fungal contaminants of Australasian Plant Pathology 47: 23–34.
stored wheat vary between Australian states. Australasian Plant Corda ACJ (1837). Icones fungorum hucusque cognitorum. J.G. Calve,
Pathology 45: 621–628. Austria.
Bashyal BM, Aggarwal R, Sharma S, et al. (2016). Single and combined Costa MM, Saleh AA, Melo MP, et al. (2022). Fusarium mirum sp. nov.,
effects of three Fusarium species associated with rice seeds on the intertwining Fusarium madaense and Fusarium andiyazi, pathogens
severity of bakanae disease of rice. Journal of Plant Pathology 98: of tropical grasses. Fungal Biology 126: 250–266.
405–412. Crous PW, Gams W, Stalpers JA, et al. (2004). MycoBank: an online
Bass D, Stentiford GD, Wang HC, et al. (2019). The pathobiome in animal initiative to launch mycology into the 21st century. Studies in Mycology
and plant diseases. Trends in Ecology and Evolution 34: 996–1008. 50: 19–22.
Bentley AR, Petrovic T, Griffiths SP, et al. (2007). Crop pathogens and Crous PW, Lombard L, Sandoval-Denis M, et al. (2021). Fusarium: more
other Fusarium species associated with Austrostipa aristiglumis. than a node or a foot-shaped basal cell. Studies in Mycology 98: 1–184.
Australasian Plant Pathology 36: 434–438. Crous PW, Sandoval-Denis M, Costa MM, et al. (2022). Fusarium and
Beukes I, Rose LJ, van Coller GJ, et al. (2017). Disease development and allied fusarioid taxa (FUSA). 1. Fungal Systematics and Evolution 9:
mycotoxin production by the Fusarium graminearum species complex 161–200.
associated with South African maize and wheat. European Journal of Crous PW, Wingfield MJ, Lombard L, et al. (2019). Fungal planet
Plant Pathology 150: 893–910. description sheets: 951–1041. Persoonia 43: 223–425.
Blacutt AA, Gold SE, Voss KA, et al. (2018). Fusarium verticillioides: Cuomo CA, Gueldener U, Xu JR, et al. (2007). The Fusarium graminearum
advancements in understanding the toxicity, virulence, and genome reveals a link between localized polymorphism and pathogen
niche adaptations of a model mycotoxigenic pathogen of maize. specialization. Science 317: 1400–1402.
Phytopathology 108: 312–326. Dean R, Van Kan JA, Pretorius ZA, et al. (2012). The Top 10 fungal
Booth C (1971). The genus Fusarium. Commonwealth Mycological Institte, pathogens in molecular plant pathology. Molecular Plant Pathology
Kew, Surrey, England. 13: 414–430.

www.studiesinmycology.org 141
Han et al.

Degnan JH, Rosenberg NA (2009). Gene tree discordance, phylogenetic Gams W (1971). Cephalosporium-artige schimmelpilze (Hyphomycetes).
inference and the multispecies coalescent. Trends in Ecology & Gustav Fischer Verlag, Stuttgart, Germany.
Evolution 24: 332–340. Gams W, Klamer M, O’Donnell K (1999). Fusarium miscanthi sp. nov. from
Desjardins AE (2006). Fusarium mycotoxins: chemistry, genetics and Miscanthus litter. Mycologia 91: 263−268.
biology. American Phytopathological Society, USA. Gams W, Nirenberg HI, Seifert KA, et al. (1997). Proposal to conserve the
Desjardins AE, Manandhar HK, Plattner RD, et al. (2000). Fusarium species name Fusarium sambucinum (Hyphomycetes). Taxon 46: 111−113.
from nepalese rice and production of mycotoxins and gibberellic acid Gao M, Xiong C, Gao C, et al. (2021). Disease-induced changes in plant
by selected species. Applied and Environmental Microbiology 66: microbiome assembly and functional adaptation. Microbiome 9: 187.
1020–1025. Geiser DM, Al-Hatmi AMS, Aoki T, et al. (2021). Phylogenomic analysis
Desjardins AE, Proctor RH (2011). Genetic diversity and trichothecene of a 55.1 kb 19-gene dataset resolves a monophyletic Fusarium that
chemotypes of the Fusarium graminearum clade isolated from maize includes the Fusarium solani species complex. Phytopathology 111:
in Nepal and identification of a putative new lineage. Fungal Biology 1064−1079.
115: 38–48. Geiser DM, Juba JH, Wang B, et al. (2001). Fusarium hostae sp. nov.,
Devay JE, Covey RP, Nair PN (1957). Corn diseases and their importance a relative of F. redolens with a Gibberella teleomorph. Mycologia 93:
in Minnesota in 1956. Plant Disease Reporter 41: 505–507. 670−678.
de Faria Ferreira M, Brito-Santos F, Henrique Nascimento Theodoro Gerlach W, Ershad D (1970). Beitrag zur kenntnis der Fusarium-und
P, et al. (2022). Mixed infection by Cryptococcus neoformans and Cylindrocarpon-arten in Iran. Nova Hedwigia 20: 725−784.
Cryptococcus gattii and coinfection with paracoccidioidomycosis in Gomes LB, Ward TJ, Badiale-Furlong E, et al. (2015). Species composition,
PLHIV. Medical Mycology Case Reports 35: 48–50. toxigenic potential and pathogenicity of Fusarium graminearum
Dong F, Xu JH, Shi JR, et al. (2020a). First report of Fusarium head blight species complex isolates from southern Brazilian rice. Plant Pathology
caused by Fusarium meridionale in rice in China. Plant Disease 104: 64: 980−987.
2726–2726. Gordon WL (1965). Pathogenic strains of Fusarium oxysporum. Canadian
Dong F, Zhang X, Xu JH, et al. (2020b). Analysis of Fusarium graminearum Journal of Botany 43: 1309−1318.
species complex from freshly harvested rice in Jiangsu Province Gräfenhan T, Johnston PR, Vaughan MM, et al. (2016). Fusarium
(China). Plant Disease 104: 2138–2143. praegraminearum sp. nov., a novel nivalenol mycotoxin-producing
Dong HY, Qin PW, Gao ZG, et al. (2020c). First report of seedling blight pathogen from New Zealand can induce head blight on wheat.
of maize caused by Fusarium asiaticum in Northeast China. Plant Mycologia 108: 1229−1239.
Disease 105: 1206–1026. Gräfenhan T, Schroers HJ, Nirenberg HI, et al. (2011). An overview of
Du Q, Duan CX, Li SC, et al. (2020). First report of maize ear rot caused the taxonomy, phylogeny, and typification of nectriaceous fungi in
by Fusarium concentricum in China. Plant Disease 104: 1539–1540. Cosmospora, Acremonium, Fusarium, Stilbella, and Volutella. Studies
Duan CX, Du Q, Tang ZL, et al. (2019). First report of maize ear rot caused in Mycology 68: 79−113.
by Fusarium sacchari in China. Plant Disease 103: 2674−2674. Han SB, Cao YY, Zhang J, et al. (2022). First report of Fusarium cf.
Duan CX, Wang BB, Sun FF, et al. (2020). Occurrence of maize ear rot longipes associated with maize stalk rot in China. Plant Disease 106:
caused by Fusarium fujikuroi in China. Plant Disease 104: 587−587. 1064−1064.
Ebbels DL, Allen DJ. (1979). A supplementary and annotated list of plant Hao JJ, Xie SN, Sun J, et al. (2017). Analysis of Fusarium graminearum
diseases, pathogens and associated fungi in Tanzania. Commonwealth species complex from wheat-maize rotation regions in Henan (China).
Mycological Institute, UK. Plant Disease 101: 720−725.
Emms DM, Kelly S (2015). OrthoFinder: solving fundamental biases Haridas S, Albert R, Binder M, et al. (2020). 101 Dothideomycetes
in whole genome comparisons dramatically improves orthogroup genomes: A test case for predicting lifestyles and emergence of
inference accuracy. Genome Biology 16:157−170. pathogens. Studies in Mycology 96: 141−153.
Emms DM, Kelly S (2019). OrthoFinder: phylogenetic orthology inference Hill R, Buggs RJA, Vu DT, et al. (2022). Lifestyle transitions in fusarioid
for comparative genomics. Genome Biology 20: 238−251. fungi are frequent and lack clear genomic signatures. Molecular
Esmaeili Taheri A, Hamel C, Gan Y, et al. (2011). First report of Fusarium Biology and Evolution 39: msac085.
redolens from Saskatchewan and its comparative pathogenicity. Hoang DT, Chernomor O, von Haeseler A, et al. (2018). UFBoot2:
Canadian Journal of Plant Pathology 33: 559−564. improving the ultrafast bootstrap approximation. Molecular Biology
Ezekiel CN, Kraak B, Sandoval-Denis M, et al. (2020). Diversity and toxigenicity and Evolution 35: 518−522.
of fungi and description of Fusarium madaense sp. nov. from cereals, Huang SW, Wang L, Liu LM, et al. (2011). Rice spikelet rot disease in
legumes and soils in north-central Nigeria. MycoKeys 67: 95−124. China–1. Characterization of fungi associated with the disease. Crop
Farr DF, Rossman AY. (2022). Fungal Databases, U.S. National Fungus Protection 30: 1−9.
Collections, ARS, USDA. <https://nt.ars-grin.gov/fungaldatabases/>. Huelsenbeck JP, Ronquist F (2001). MrBayes: Bayesian inference of
Accessed on 13 August 2022. phylogeny. Bioinformatics 17: 754–755.
Fisher NL, Burgess LW, Toussoun TA, et al. (1982). Carnation leaves Husna A, Zakaria L, Mohamed Nor NMI (2020). Fusarium commune
as a substrate and for preserving cultures of Fusarium species. associated with wilt and root rot disease in rice. Plant Pathology 70:
Phytopathology 72: 151−153. 123−132.
Fourie G, Steenkamp ET, Gordon TR, et al. (2009). Evolutionary Jamieson CO, Wollenweber HW (1912). An external dry rot of potato tubers
relationships among the Fusarium oxysporum f. sp. cubense vegetative caused by Fusarium trichothecioides. Journal of the Washington
compatibility groups. Applied and Environmental Microbiology 75: Academy of Sciences 2: 146−152.
4770−4781. Ji LJ, Kong LX, Li QS, et al. (2016). First report of Fusarium
Fumero MV, Yue W, Chiotta ML, et al. (2021). Divergence and gene flow pseudograminearum causing Fusarium head blight of wheat in Hebei
between Fusarium subglutinans and F. temperatum isolated from Province, China. Plant Disease 100: 220−220.
maize in Argentina. Phytopathology 111: 170−183. Jiang W, Han W, Wang R, et al. (2021). Development of an inoculation
Gai XT, Xuan YH, Gao ZG (2017). Diversity and pathogenicity of Fusarium technique for rapidly evaluating maize inbred lines for resistance to
graminearum species complex from maize stalk and ear rot strains in stalk rot caused by Fusarium spp. in the field. Plant Disease 105:
northeast China. Plant Pathology 66: 1267−1275. 2306−2313.
Gai XT, Yang RX, Pan XJ, et al. (2016). First report of Fusarium incarnatum Jin QM, Lu ZZ, Pan SF, et al. (1994). Study on pathogenicities of pathogenic
causing stalk rot on maize in China. Plant Disease 100: 1010–1010. fungi of corn stalk rot in corn seedling stage. Maize Science 1: 73−75.
Gale LR, Harrison SA, Ward TJ, et al. (2011). Nivalenol-type populations Kalyaanamoorthy S, Minh BQ, Wong TKF, et al. (2017). ModelFinder: Fast
of Fusarium graminearum and F. asiaticum are prevalent on wheat in model selection for accurate phylogenetic estimates. Nature Methods
southern Louisiana. Phytopathology 101: 124−134. 14: 587−589.

142
Fusarium spp. on cereals in China

Katoh K, Standley DM (2013). MAFFT multiple sequence alignment postharvest pumpkin caused by Fusarium acuminatum in China. Plant
software version 7: improvements in performance and usability. Disease 103: 1035−1035.
Molecular Biology and Evolution 30: 772−780. Li YG, Zhang X, Zhang R, et al. (2019b). Occurrence of seedling blight
Khaledi N, Taheri P, Falahati Rastegar M (2017). Identification, virulence caused by Fusarium tricinctum on rice in China. Plant Disease 103:
factors characterization, pathogenicity and aggressiveness analysis of 1789-1790.
Fusarium spp., causing wheat head blight in Iran. European Journal of Lima EN, Oster AH, Bordallo PN, et al. (2020). A novel lineage in the
Plant Pathology 147: 897−918. Fusarium incarnatum-equiseti species complex is one of the causal
Khoa LV, Hatai K, Aoki T (2004). Fusarium incarnatum isolated from black agents of Fusarium rot on melon fruits in northeast Brazil. Plant
tiger shrimp, Penaeus monodon Fabricius, with black gill disease Pathology 70: 133−143.
cultured in Vietnam. Journal of Fish Diseases 27: 507−515. Link HF (1809). Observationes in ordines plantarum naturals, Dissetatio I.
Khuna S, Kumla J, Thitla T, et al. (2022). Morphology, molecular Gesellschaft Naturforschender Freunde zu Berlin, Magazin 3: 3−42.
identification, and pathogenicity of two novel Fusarium species Liu F, Ma ZY, Hou LW, et al. (2022a). Updating species diversity of
associated with postharvest fruit rot of cucurbits in northern Thailand. Colletotrichum, with a phylogenomic overview. Studies in Mycology
Journal of Fungi 8: 1135−1152. 101: 1−56.
Klittich CJR, Leslie JF, Nelson PE, et al. (1997). Fusarium thapsinum Liu J, Han Y, Li W, et al. (2022b). Identification of pathogens and evaluation
(Gibberella thapsina): a new species in section Liseola from sorghum. of resistance and genetic diversity of maize inbred lines to stalk rot in
Mycologia 89: 643−652. Heilongjiang Province, China. Plant Disease: https://doi.org/10.1094/
Kornerup A, Wanscher JH. (1978). Methuen handbook of colour. 3rd edn. PDIS-03-22-0525-RE.
Eyre Methuen, London. Liu S, García-Palacios P, Tedersoo L, et al. (2022c). Phylotype diversity
Kriek NPJ, Marasas WFO, Steyn PS, et al. (1977). Toxicity of a moniliformin- within soil fungal functional groups drives ecosystem stability. Nature
producing strain of Fusarium moniliforme var. subglutinans isolated Ecology and Evolution 6: 900−909.
from maize. Food and Cosmetics Toxicology 15: 579−587. Liu X, Fang X, Yu F, et al. (2022d). Improved whole-genome sequence of
Kristensen R, Torp M, Kosiak B, et al. (2005). Phylogeny and toxigenic Fusarium meridionale, the fungal pathogen causing Fusarium head
potential is correlated in Fusarium species as revealed by partial blight in rice. Molecular Plant-Microbe Interactions 35: 85−89.
translation elongation factor 1 alpha gene sequences. Mycological Liu XX, Wang Y, Qiu JF, et al. (2018). Analysis of population, mycotoxin
Research 109: 173−186. chemotypes and mycotoxin pollution of Fusarium from spring wheat in
Krone MJ, Salazar M, Mideros S (2020). First report of Fusarium northern China. Jiangsu Agricultural Sciences 46: 199−202.
armeniacum causing Fusarium head blight on soft red winter wheat in Liu YJ, Whelen S, Hall BD (1999). Phylogenetic relationships among
Illinois. Plant Disease 105: 1199−1199. ascomycetes: evidence from an RNA polymerse II subunit. Molecular
Kuhnem PR, Ward TJ, Silva CN, et al. (2016). Composition and toxigenic Biology and Evolution 16: 1799−1808.
potential of the Fusarium graminearum species complex from maize Lofgren LA, LeBlanc NR, Certano AK, et al. (2018). Fusarium graminearum:
ears, stalks and stubble in Brazil. Plant Pathology 65: 1185−1191. pathogen or endophyte of North American grasses?. New Phytologist
Kumar S, Stecher G, Tamura K (2016). MEGA7: molecular evolutionary 217: 1203−1212.
genetics analysis version 7.0 for bigger datasets. Molecular Biology Logrieco A, Mule G, Moretti A, et al. (2002a). Toxigenic Fusarium species
and Evolution 33: 1870−1874. and mycotoxins associated with maize ear rot in Europe. European
Lana FD, Madden LV, Carvalho CP, et al. (2022). Impact of Gibberella ear Journal of Plant Pathology 108: 597−609.
rot on grain quality and yield components in maize as influenced by Logrieco A, Rizzo A, Ferracane R, et al. (2002b). Occurrence of beauvericin
hybrid reaction. Plant Disease 106: 3061–3075. and enniatins in wheat affected by Fusarium avenaceum head blight.
Laraba I, Keddad A, Boureghda H, et al. (2017). Fusarium algeriense, sp. Applied and Environmental Microbiology 68: 82−85.
nov., a novel toxigenic crown rot pathogen of durum wheat from Algeria Lombard L, Sandoval-Denis M, Cai L, et al. (2019a). Changing the game:
is nested in the Fusarium burgessii species complex. Mycologia 109: resolving systematic issues in key Fusarium species complexes.
935−950. Persoonia 43: i−ii.
Laraba I, McCormick SP, Vaughan MM, et al. (2021). Phylogenetic Lombard L, Sandoval-Denis M, Lamprecht SC, et al. (2019b). Epitypification
diversity, trichothecene potential, and pathogenicity within Fusarium of Fusarium oxysporum − clearing the taxonomic chaos. Persoonia
sambucinum species complex. PLoS ONE 16: e0245037. 43: 1−47.
Laurence MH, Walsh JL, Shuttleworth LA, et al. (2016). Six novel species Lombard L, van der Merwe NA, Groenewald JZ, et al. (2015). Generic
of Fusarium from natural ecosystems in Australia. Fungal Diversity 77: concepts in Nectriaceae. Studies in Mycology 80: 189−245.
349−366. Lombard L, van Doorn R, Crous PW (2019c). Neotypification of Fusarium
Lee T, Paek JS, Lee KA, et al. (2016). Occurrence of toxigenic Fusarium chlamydosporum − a reappraisal of a clinically important species
vorosii among small grain cereals in Korea. The Plant Pathology complex. Fungal Systematics and Evolution 4: 183−200.
Journal 32: 407−413. Lombard L, van Doorn R, Groenewald JZ, et al. (2022). Fusarium diversity
Leslie JF, Summerell BA. (2006). The Fusarium Laboratory Manual. associated with the Sorghum-Striga interaction in Ethiopia. Fungal
Blackwell Publishing Professional, USA. Systematics and Evolution 10: 177–215.
Lew H, Chelkowski J, Pronczuk P, et al. (1996). Occurrence of the Lukanowski A, Lenc L, Sadowski C (2008). First report on the occurrence
mycotoxin moniliformin in maize (Zea mays L.) ears infected by of Fusarium langsethiae isolated from wheat kernels in Poland. Plant
Fusarium subglutinans (Wollenw. & Reinking) Nelson et al. Food Disease 92: 488−488.
Additives and Contaminants 13: 321−324. Ma LJ (2014). Horizontal chromosome transfer and rational strategies to
Leyva-Madrigal KY, Larralde-Corona CP, Apodaca-Sánchez MA, et al. manage Fusarium vascular wilt diseases. Molecular Plant Pathology
(2015). Fusarium species from the Fusarium fujikuroi species complex 15: 763−766.
involved in mixed infections of maize in Northern Sinaloa, Mexico. Ma LJ, van der Does HC, Borkovich KA, et al. (2010). Comparative
Journal of Phytopathology 163: 486−497. genomics reveals mobile pathogenicity chromosomes in Fusarium.
Li HL, He XL, Ding SL, et al. (2016). First report of Fusarium culmorum Nature 464: 367−373.
causing crown rot of wheat in China. Plant Disease 100: 2532−2532. Ma N, Haseeb HA, Xing F, et al. (2019). Fusarium avenaceum: a toxigenic
Li HL, Yuan HX, Fu B, et al. (2012). First report of Fusarium pathogen causing ear rot on maize in Yunnan Province, China. Plant
pseudograminearum causing crown rot of wheat in Henan, China. Disease 103: 1424−1425.
Plant Disease 96: 1065−1065. Ma WF (2020). Analysis of China’s food security situation. Food Processing
Li PP, Cao ZY, Wang K, et al. (2014). First report of Fusarium equiseti 46: 1−4.
causing a sheath rot of corn in China. Plant Disease 98: 998−998. Ma ZY, Wu Q, Zhou X, et al. (2021). Assembly and annotation of fungal
Li YG, Jiang WY, Jiang D, et al. (2019a). First report of fruit rot on genome based on Illumina sequencing. Bio-protocol: e2003704.

www.studiesinmycology.org 143
Han et al.

Malihipour A, Gilbert J, Piercey-Normore M, et al. (2012). Molecular Mulenko W, Majewski T, Ruszkiewicz-Michalska M. (2008). A preliminary
phylogenetic analysis, trichothecene chemotype patterns, and checklist of micromycetes in Poland. W. Szafer Institute of Botany,
variation in aggressiveness of Fusarium isolates causing head blight Polish Academy of Sciences, Poland.
in wheat. Plant Disease 96: 1016−1025. Murillo-Williams A, Munkvold GP (2008). Systemic infection by Fusarium
Mao W, Carroll RB, Whittington DP (1998) Association of Phoma terrestris, verticillioides in maize plants grown under three temperature regimes.
Pythium irregulare, and Fusarium acuminatum in causing red root rot Plant Disease 92: 1695−1700.
of corn. Plant Disease 82: 337−342. Nees von Esenbeck CGD (1817). Das System der Pilze und Schwämme.
Marasas WFO, Nelson PE, Toussoun TA. (1984). Toxigenic Fusarium species: Stahel, Germany.
identity and mycotoxicology. Pennsylvania State University, USA. Nees von Esenbeck CGD, Nees von Esenbeck TFL (1818). De plantis
Marasas WFO, Nelson PE, Toussoun TA (1985). Fusarium dlamini, a new nonnullis e mycetoidearum regno tum nuper detectis, tum minus
species from southern Africa. Mycologia 77: 971−975. cognitis commentatio prior doctoris Nees ab Esenbeck et Friderici
Marasas WFO, Rabie CJ, Lübben A, et al. (1987). Fusarium napiforme, Nees fratrum. Nova Acta Physico-Medica Academiae Caesareae
a new species from millet and sorghum in southern Africa. Mycologia Leopoldino-Carolinae Naturae Curiosorum 9: 226−262.
79: 910−914. Nelson PE, Toussoun TA, Burgess LW (1987). Characterization of
Marasas WFO, Rheeder JP, Lamprecht SC, et al. (2001). Fusarium andiyazi Fusarium beomiforme sp. nov. Mycologia 79: 884−889.
sp. nov., a new species from sorghum. Mycologia 93: 1203−1210. Nelson PE, Toussoun TA, Marasas WFO. (1983). Fusarium species: an
Marasas WFO, Rheeder JP, Logrieco A, et al. (1998). Fusarium nelsonii Illustrated manual for identification. Pennsylvania State University
and F. musarum: two new species in Section Arthrosporiella related to Press, USA.
F. camptoceras. Mycologia 90: 505−513. Nganje WE, Kaitibie S, Wilson WW, et al. (2004). Economic impacts of
Marín P, Moretti A, Ritieni A, et al. (2012). Phylogenetic analyses and toxigenic Fusarium head blight in wheat and barley: 1993-2001. Agribusiness
profiles of Fusarium equiseti and Fusarium acuminatum isolated from & Applied Economics Report 23627, North Dakota State University,
cereals from southern Europe. Food Microbiology 31: 229−237. Department of Agribusiness and Applied Economics.
Maryani N, Lombard L, Poerba YS, et al. (2019a). Phylogeny and genetic Nirenberg H (1976). Untersuchungen über die morphologische und
diversity of the banana Fusarium wilt pathogen Fusarium oxysporum biologische Differenzierung in der Fusarium-Sektion Liseola.
f. sp. cubense in the Indonesian centre of origin. Studies in Mycology Mitteilungen aus der Biologischen Bundesanstalt für Land- und
92: 155−194. Forstwirtschaft Berlin-Dahlem 169: 1−117.
Maryani N, Sandoval-Denis M, Lombard L, et al. (2019b). New endemic Nirenberg HI (1995). Morphological differentiation of Fusarium sambucinum
Fusarium species hitch-hiking with pathogenic Fusarium strains Fuckel sensu stricto, F. torulosum (Berk. & Curt.) Nirenberg comb. nov.
causing Panama disease in small-holder banana plots in Indonesia. and F. venenatum Nirenberg sp. nov. Mycopathologia 129: 131−141.
Persoonia 43: 48−69. Nirenberg HI (1997). Fusarium nisikadoi, a new species from Japan.
Matic S, Tabone G, Gullino ML, et al. (2020). Emerging leafy vegetable Mycoscience 38: 329−333.
crop diseases caused by the Fusarium incarnatum-equiseti species Nirenberg HI, O’Donnell K (1998). New Fusarium species and combinations
complex. Phytopathologia Mediterranea 59: 303−317. within the Gibberella fujikuroi species complex. Mycologia 90: 434−458.
Matny ON (2015). Fusarium head blight and crown rot on wheat & barley: Nirenberg HI, O’Donnell K, Kroschel J, et al. (1998). Two new species of
losses and health risks. Advances in Plants & Agriculture Research Fusarium: Fusarium brevicatenulatum from the noxious weed Striga
2: 38−43. asiatica in Madagascar and Fusarium pseudoanthophilum from Zea
McGuire JU, Crandall BS. (1967). Survey of insect pests and plant diseases mays in Zimbabwe. Mycologia 90: 459−464.
of selected food crops of Mexico, central America and Panama. USDA Nylander JAA (2004). MrModeltest v2. Program distributed by the author.
Int. agric. Development Service, USA. Evolutionary Biology Centre, Uppsala University, Sweden.
McTaggart AR, James TY, Shivas RG, et al. (2021). Population genomics O’Donnell K, Cigelnik E (1997). Two divergent intragenomic rDNA ITS2
reveals historical and ongoing recombination in the Fusarium types within a monophyletic lineage of the fungus Fusarium are
oxysporum species complex. Studies in Mycology 99: 100132. nonorthologous. Molecular Phylogenetics and Evolution 7: 103−116.
Medeiros Araújo MB, Moreira GM, Nascimento LV, et al. (2020). Fusarium O’Donnell K, Cigelnik E, Nirenberg HI (1998a). Molecular systematics
rot of melon is caused by several Fusarium species. Plant Pathology and phylogeography of the Gibberella fujikuroi species complex.
70: 712−721. Mycologia 90: 465−493.
Mesny F, Miyauchi S, Thiergart T, et al. (2021). Genetic determinants O’Donnell K, Humber RA, Geiser DM, et al. (2012). Phylogenetic diversity
of endophytism in the Arabidopsis root mycobiome. Nature of insecticolous fusaria inferred from multilocus DNA sequence data
Communications 12: 7227. and their molecular identification via FUSARIUM-ID and Fusarium
Mezzalama M, Guarnaccia V, Martino I, et al. (2021). First report of MLST. Mycologia 104: 427−445.
Fusarium commune causing root and crown rot on maize in Italy. Plant O’Donnell K, Kistlerr HC, Cigelnik E, et al. (1998b). Multiple evolutionary
Disease 105: 4156. origins of the fungus causing panama disease of banana: Concordant
Miller MA, Pfeiffer W, Schwartz T (2012) The CIPRES science gateway: evidence from nuclear and mitochondrial gene genealogies.
enabling high-impact science for phylogenetics researchers with Proceedings of the National Academy of Sciences of the United States
limited resources. In: Proceedings of the 1st conference of the extreme of America 95: 2044-2049.
science and engineering discovery environment: Bridging from the O’Donnell K, Kistler HC, Tacke BK, et al. (2000a). Gene genealogies
extreme to the campus and beyond. Association for Computing reveal global phylogeographic structure and reproductive isolation
Machinery, USA: 1−8. among lineages of Fusarium graminearum, the fungus causing wheat
Minh BQ, Hahn MW, Lanfear R (2020b). New methods to calculate scab. Proceedings of the National Academy of Sciences of the United
concordance factors for phylogenomic datasets. Molecular Biology States of America 97: 7905−7910.
and Evolution 37: 2727−2733. O’Donnell K, McCormick SP, Busman M, et al. (2018). Marasas et al. 1984
Minh BQ, Schmidt HA, Chernomor O, et al. (2020a). IQ-TREE 2: New “Toxigenic Fusarium Species: Identity and Mycotoxicology” revisited.
models and efficient methods for phylogenetic inference in the Mycologia 110: 1058−1080.
genomic era. Molecular Biology and Evolution 37: 1530−1534. O’Donnell K, Nirenberg H, Aoki T, et al. (2000b). A multigene phylogeny
Moses LM, Marasas WF, Vismer HF, et al. (2010). Molecular characterization of the Gibberella fujikuroi species complex: detection of additional
of Fusarium globosum strains from South African maize and Japanese phylogenetically distinct species. Mycoscience 41: 61−78.
wheat. Mycopathologia 170: 237−249. O’Donnell K, Rooney AP, Proctor RH, et al. (2013). Phylogenetic analyses
Mostert D, Molina AB, Daniells J, et al. (2017). The distribution and host of RPB1 and RPB2 support a middle Cretaceous origin for a clade
range of the banana Fusarium wilt fungus, Fusarium oxysporum f. sp. comprising all agriculturally and medically important fusaria. Fungal
cubense, in Asia. PLoS ONE 12: e0181630. Genetics and Biology 52: 20−31.

144
Fusarium spp. on cereals in China

O’Donnell K, Sutton DA, Rinaldi MG, et al. (2009). Novel multilocus Reeb V, Lutzoni F, Roux C (2004). Contribution of RPB2 to multilocus
sequence typing scheme reveals high genetic diversity of human phylogenetic studies of the Euascomycetes (Pezizomycotina, Fungi)
pathogenic members of the Fusarium incarnatum-F. equiseti and F. with special emphasis on the lichen-forming Acarosporaceae and
chlamydosporum species complexes within the United States. Journal evolution of polyspory. Molecular Phylogenetics and Evolution 32:
of Clinical Microbiology 47: 3851−3861. 1036−1060.
O’Donnell K, Sutton DA, Rinaldi MG, et al. (2010). Internet-accessible DNA Reinking OA (1934). Interesting new Fusaria. Zentralblatt für Bakteriologie
sequence database for identifying fusaria from human and animal und Parasitenkunde, Abteilung 2 89: 509–514.
infections. Journal of Clinical Microbiology 48: 3708−3718. Renev E, Zargaryan NY, Kekalo AY, et al. (2021). Infection of grain crops
O’Donnell K, Ward TJ, Geiser DM, et al. (2004). Genealogical concordance with fungi of the genus Fusarium. BIO Web of Conferences 36:
between the mating type locus and seven other nuclear genes 04008−04015.
supports formal recognition of nine phylogenetically distinct species Rheeder JP, Marasas WFO, Nelson PE (1996). Fusarium globosum, a new
within the Fusarium graminearum clade. Fungal Genetics and Biology species from corn in southern Africa. Mycologia 88: 509−513.
41: 600−623. Richardson MJ (1979). An annotated list of seed-borne diseases. 4th edn.
O’Donnell K, Ward TJ, Robert VARG, et al. (2015). DNA sequence- International Seed Testing Association, Switzerland.
based identification of Fusarium: current status and future directions. Ronquist F, Huelsenbeck JP (2003). MrBayes 3: Bayesian phylogenetic
Phytoparasitica 43: 583−595. inference under mixed models. Bioinformatics 19: 1572−1574.
O’Donnell K, Whitaker B, Laraba I, et al. (2022). DNA sequence-based Roux J, Steenkamp ET, Marasas WFO, et al. (2001). Characterization of
identification of Fusarium: a work in progress. Plant Disease 106: Fusarium graminearum from Acacia and Eucalyptus using β-tubulin
1579−1609 and histone gene sequences. Mycologia 93: 704−711.
Obanor F, Erginbas-Orakci G, Tunali B, et al. (2010). Fusarium culmorum is Saccardo PA (1881). Fungi Italici Autographice Delineati. Patavii, Italy.
a single phylogenetic species based on multilocus sequence analysis. Saccardo PA (1879). Fungi Gallici lecti a cl. viris P. Brunaud, C.C. Gillet et
Fungal Biology 114: 753−765. Abb. Letendre. Michelia 1: 500−538.
Okello PN, Petrović K, Kontz B, et al. (2019). Eight species of Fusarium Saccardo PA (1886). Sylloge Hyphomycetum. Sylloge Fungorum 4: 1−807.
cause root rot of corn (Zea mays) in South Dakota. Plant Health Saccardo PA (1892). Supplementum Universale, Pars II. Discomyceteae-
Progress 20: 38−43. Hyphomyceteae. Sylloge Fungorum 10: 1−964.
Özer G, İmren M, Bayraktar H, et al. (2019). First report of Fusarium Sandoval-Denis M, Guarnaccia V, Polizzi G, et al. (2018a). Symptomatic
hostae causing crown rot on wheat in Azerbaijan. Plant Disease 103: citrus trees reveal a new pathogenic lineage in Fusarium and two new
3278−3278. Neocosmospora species. Persoonia 40: 1−25.
Özer G, Paulitz TC, Imren M, et al. (2020). Identity and pathogenicity of Sandoval-Denis M, Swart WJ, Crous PW (2018b). New Fusarium species
Fungi associated with crown and root rot of dryland winter wheat in from the Kruger National Park, South Africa. MycoKeys 34: 63−92.
Azerbaijan. Plant Disease 104: 2149−2157. Santos A, Trindade JVC, Lima CS, et al. (2019). Morphology, phylogeny, and
Padwick GW (1945). Notes on Indian fungi. III. Mycological Papers 12: sexual stage of Fusarium caatingaense and Fusarium pernambucanum,
1−12. new species of the Fusarium incarnatum-equiseti species complex
Pak D, You MP, Lanoiselet V, et al. (2016). Reservoir of cultivated rice associated with insects in Brazil. Mycologia 111: 244−259.
pathogens in wild rice in Australia. European Journal of Plant Sarowar S, Alam ST, Makandar R, et al. (2019). Targeting the pattern-
Pathology 147: 295−311. triggered immunity pathway to enhance resistance to Fusarium
Palmer JM. (2016). Funannotate: pipeline for genome annotation. <https:// graminearum. Molecular Plant Pathology 20: 626−640.
funannotate.readthedocs.io/en/latest/commands.html>. Sarver BA, Ward TJ, Gale LR, et al. (2011). Novel Fusarium head
Pan XJ, Chen N, Yao Y, et al. (2015). Population and pathogenicity of blight pathogens from Nepal and Louisiana revealed by multilocus
Fusarium spp. causing wheat head blight in Northeast of China. Acta genealogical concordance. Fungal Genetics and Biology 48:
Agriculturae Boreali-Sinica 30: 205−210. 1096−1107.
Pennycook SR. (1989). Plant diseases recorded in New Zealand (Vol. 3). Savary S, Willocquet L, Pethybridge SJ, et al. (2019). The global burden
Plant Diseases Division, DSIR, New Zealand. of pathogens and pests on major food crops. Nature Ecology and
Phan HT, Burgess LW, Summerell BA, et al. (2004). Gibberella gaditjirrii Evolution 3: 430−439.
(Fusarium gaditjirrii) sp. nov., a new species from tropical grasses in Scauflaire J, Gourgue M, Munaut F (2011). Fusarium temperatum sp.
Australia. Studies in Mycology 50: 261−272. nov. from maize, an emergent species closely related to Fusarium
Ploetz RC (2006). Fusarium-induced diseases of tropical, perennial crops. subglutinans. Mycologia 103: 586−597.
Phytopathology 96: 648−652. Schneider R, Dalchow J (1975). Fusarium inflexum spec. nov., als erreger
Ploetz R, Pegg K (1997). Fusarium wilt of banana and Wallace’s line: einer welkekrankheit an Vicia faba L. in Deutschland. Journal of
was the disease originally restricted to his Indo-Malayan region?. Phytopathology 82: 70−82.
Australasian Plant Pathology 26: 239−249. Schoeman A, Flett BC, Janse van Rensburg B, et al. (2018). Pathogenicity
Porebski S, Bailey LG, Baum BR (1997). Modification of a CTAB DNA and toxigenicity of Fusarium verticillioides isolates collected from
extraction protocol for plants containing high polysaccharide and maize roots, stems and ears in South Africa. European Journal of
polyphenol components. Plant Molecular Biology Reporter 15: 8−15. Plant Pathology 152: 677−689.
Przemieniecki SW, Kurowski TP, Korzekwa K (2014). Chemotypes Schroers HJ, Gräfenhan T, Nirenberg HI, et al. (2011). A revision of
and geographic distribution of the Fusarium graminearum species Cyanonectria and Geejayessia gen. nov., and related species with
complex. Environmental Biotechnology 10: 45−54. Fusarium-like anamorphs. Studies in Mycology 68: 115−138.
Qiu JB, Lu YA, He D, et al. (2020). Fusarium fujikuroi species complex Schwabe SH (1839). Flora Anhaltina. Apud Ge. Reimerum, Germany.
associated with rice, maize, and soybean from Jiangsu Province, Seppey M, Manni M, Zdobnov EM (2019). BUSCO: assessing genome
China: phylogenetic, pathogenic, and toxigenic analysis. Plant assembly and annotation completeness. Methods in Molecular
Disease 104: 2193−2201. Biology 1962: 227−245.
Rahjoo V, Zad J, Javan-Nikkhah M, et al. (2008). Morphological and Shan L, Zhang J, Ma N, et al. (2018). First report of maize stalk rot disease
molecular identification of Fusarium isolated from maize ears in Iran. caused by Fusarium cerealis in Yunnan, China. Plant Disease 102:
Journal of Plant Pathology 90: 463−468. 444−444.
Raillo A (1950). Fungi of the genus Fusarium. Publication State Agricultural Shan LY, Cui WY, Zhang DD, et al. (2017). First report of Fusarium
Literature, Russia. brachygibbosum causing maize stalk rot in China. Plant Disease 101:
Raza M, Zhang ZF, Hyde KD, et al. (2019). Culturable plant pathogenic 837−837.
fungi associated with sugarcane in southern China. Fungal Diversity Shang G, Yu H, Yang J, et al. (2020). First report of Fusarium miscanthi
99: 1−104. causing ear rot on maize in China. Plant Disease 105: 1565−1565.

www.studiesinmycology.org 145
Han et al.

Sherbakoff CD (1915). Fusaria of potatoes. New York (Ithaca) Agriculture Torbati M, Arzanlou M, Sandoval-Denis M, et al. (2019). Multigene
6: 87−270. phylogeny reveals new fungicolous species in the Fusarium tricinctum
Shi XC, Liu XQ, Xie XL, et al. (2012). Gene chip array for differentiation species complex and novel hosts in the genus Fusarium from Iran.
of mycobacterial species and detection of drug resistance. Chinese Mycological Progress 18: 119−133.
Medical Journal 125: 3292−3297. Torp M, Nirenberg HI (2004). Fusarium langsethiae sp. nov. on cereals in
Short DP, O’Donnell K, Zhang N, et al. (2011). Widespread occurrence of Europe. International Journal of Food Microbiology 95: 247−256.
diverse human pathogenic types of the fungus Fusarium detected in Torres-Cruz TJ, Whitaker BK, Proctor RH, et al. (2022). FUSARIUM-ID
plumbing drains. Journal of Clinical Microbiology 49: 4264−4272. v.3.0: an updated, downloadable resource for Fusarium species
Shrestha S, Poudel RS, Zhong S (2021). Identification of fungal species identification. Plant Disease 106: 1610−1616.
associated with crown and root rots of wheat and evaluation of plant Touati-Hattab S, Barreau C, Verdal-Bonnin MN, et al. (2016). Pathogenicity
reactions to the pathogens in North Dakota. Plant Disease 105: and trichothecenes production of Fusarium culmorum strains causing
3564−3572. head blight on wheat and evaluation of resistance of the varieties
Simbaqueba J, Catanzariti AM, Gonzalez C, et al. (2018). Evidence for cultivated in Algeria. European Journal of Plant Pathology 145: 797−814.
horizontal gene transfer and separation of effector recognition from Trail F (2009). For blighted waves of grain: Fusarium graminearum in the
effector function revealed by analysis of effector genes shared postgenomics era. Plant Physiology 149: 103−110.
between cape gooseberry- and tomato-infecting formae speciales of Tralamazza SM, Piacentini KC, Savi GD, et al. (2021). Wild rice (O.
Fusarium oxysporum. Molecular Plant Pathology 19: 2302−2318. latifolia) from natural ecosystems in the Pantanal region of Brazil:
Skovgaard K, Rosendahl S, O’Donnell K, et al. (2003). Fusarium Host to Fusarium incarnatum-equiseti species complex and highly
commune is a new species identified by morphological and molecular contaminated by zearalenone. International Journal of Food
phylogenetic data. Mycologia 95: 630−636. Microbiology 345: 109127.
Snyder WC, Hansen HN (1940). The species concept in Fusarium. Tsehaye H, Brurberg MB, Sundheim L, et al. (2016). Natural occurrence
American Journal of Botany 27: 64−67. of Fusarium species and fumonisin on maize grains in Ethiopia.
Snyder WC, Hansen HN (1941). The species concept in Fusarium with European Journal of Plant Pathology 147: 141−155.
reference to section Martiella. American Journal of Botany 28: 738−742. Tunali B, Nicol JM, Hodson D, et al. (2008). Root and crown rot Fungi
Snyder WC, Hansen HN (1945). The species concept in Fusarium with associated with spring, facultative, and winter wheat in Turkey. Plant
reference to Discolour and other sections. American Journal of Botany Disease 92: 1299−1306.
32: 657−666. van Kan JA, Shaw MW, Grant-Downton RT (2014). Botrytis species:
Snyder WC, Hansen HN (1954). Variation and speciation in the genus relentless necrotrophic thugs or endophytes gone rogue?. Molecular
Fusarium. Annals of the New York Academy of Sciences 60: 16−23. Plant Pathology 15: 957−961.
Stamatakis A (2014). RAxML version 8: a tool for phylogenetic analysis Varela CP, Casal OA, Padin MC, et al. (2013). First report of Fusarium
and post-analysis of large phylogenies. Bioinformatics 30: 1312−1313. temperatum causing seedling blight and stalk rot on maize in Spain.
Starkey DE, Ward TJ, Aoki T, et al. (2007). Global molecular surveillance Plant Disease 97: 1252−1253.
reveals novel Fusarium head blight species and trichothecene toxin Varga E, Wiesenberger G, Hametner C, et al. (2015). New tricks of an
diversity. Fungal Genetics and Biology 44: 1191−1204. old enemy: isolates of Fusarium graminearum produce a type A
Strauss AT, Bowerman L, Porath-Krause A, et al. (2021). Mixed infection, trichothecene mycotoxin. Environmental Microbiology 17: 2588−2600.
risk projection, and misdirection: interactions among pathogens alter Vayssier-Taussat M, Albina E, Citti C, et al. (2014). Shifting the paradigm
links between host resources and disease. Ecology and Evolution 11: from pathogens to pathobiome: new concepts in the light of meta-
9599−9609. omics. Frontiers in Cellular and Infection Microbiology 4: 29.
Sturm J (1829). Deutschlands Flora, Abt. III. Die Pilze Deutschlands. Nabu Vayssier-Taussat M, Kazimirova M, Hubalek Z, et al. (2015). Emerging
Press, Germany. horizons for tick-borne pathogens: from the ‘one pathogen-one
Su PS, Ge WY, Wang HW, et al. (2021). Advances in understanding the disease’ vision to the pathobiome paradigm. Future Microbiology 10:
mechanisms of wheat-Fusarium graminearum interactions. Scientia 2033−2043.
Sinca (Vitae) 51: 1493−1507. Venturini G, Toffolatti SL, Quaglino F, et al. (2017). First report of Fusarium
Subrahmanyam A (1983). Fusarium laceratum. Mykosen 26: 478−480. andiyazi causing ear rot on maize in Italy. Plant Disease 101: 839−839.
Suga H, Karugia GW, Ward T, et al. (2008). Molecular characterization of Villani A, Moretti A, De Saeger S, et al. (2016). A polyphasic approach
the Fusarium graminearum species complex in Japan. Phytopathology for characterization of a collection of cereal isolates of the Fusarium
98: 159−166. incarnatum-equiseti species complex. International Journal of Food
Summerell BA, Leslie JF, Liew ECY, et al. (2010). Fusarium species Microbiology 234: 24−35.
associated with plants in Australia. Fungal Diversity 46: 1−27. Vlaardingerbroek I, Beerens B, Rose L, et al. (2016). Exchange of core
Susi H, Barrès B, Vale PF, et al. (2015). Co-infection alters population chromosomes and horizontal transfer of lineage-specific chromosomes
dynamics of infectious disease. Nature Communications 6: 5975. in Fusarium oxysporum. Environmental Microbiology 18: 3702−3713.
Sweet MJ, Bulling MT (2017). On the importance of the microbiome and Vogelgsang S, Sulyok M, Banziger I, et al. (2008a). Effect of fungal strain
pathobiome in coral health and disease. Frontiers in Marine Science and cereal substrate on in vitro mycotoxin production by Fusarium
4: 1−11. poae and Fusarium avenaceum. Food Additives and Contaminants:
Tai FL (1979). Sylloge Fungorum Sinicorum. Science Press, Academia Part A 25: 745−757.
Sinica, China. Vogelgsang S, Sulyok M, Hecker A, et al. (2008b). Toxigenicity and
Talas F, McDonald BA (2015). Genome-wide analysis of Fusarium pathogenicity of Fusarium poae and Fusarium avenaceum on wheat.
graminearum field populations reveals hotspots of recombination. European Journal of Plant Pathology 122: 265−276.
BMC Genomics 16: 996. von Schlechtendal DFL. (1824). Flora Berolinensis, Pars secunda:
Tan J, Ameye M, Landschoot S, et al. (2020). At the scene of the crime: Cryptogamia. Ferdinandi Dümmler, Germany.
New insights into the role of weakly pathogenic members of the Vorob’eva I, Toropova E. (2020). Fungi ecological niches of the genus
Fusarium head blight disease complex. Molecular Plant Pathology 21: Fusarium Link. BIO Web of Conferences 24: 00095.
1559-1572. Waner JL (1994). Mixed viral infections: detection and management.
Tarashi S, Fateh A, Mirsaeidi M, et al. (2017). Mixed infections in Clinical Microbiology Reviews 7: 143−151.
tuberculosis: the missing part in a puzzle. Tuberculosis (Edinb) 107: Wang BB, Guo C, Sun SL, et al. (2020). First report of maize ear rot caused
168−174. by Fusarium sporotrichioides in China. Plant Disease 104: 567−568.
Tekauz A, McCallum B, Ames N, et al. (2004). Fusarium head blight of Wang JH, Li HP, Zhang JB, et al. (2014). First report of Fusarium maize
oat – current status in western Canada. Canadian Journal of Plant ear rot caused by Fusarium kyushuense in China. Plant Disease 98:
Pathology 26: 473−479. 279−280.

146
Fusarium spp. on cereals in China

Wang JH, Peng XD, Lin SH, et al. (2015). First report of Fusarium head Xue AG, Chen Y, Seifert K, et al. (2019). Prevalence of Fusarium species
blight of wheat caused by Fusarium sacchari in China. Plant Disease causing head blight of spring wheat, barley and oat in Ontario during
99: 160−160. 2001–2017. Canadian Journal of Plant Pathology 41: 392−402.
Wang L, Ge SL, Zhao KH, et al. (2021). First report of Fusarium incarnatum Yang MX, Zhang H, Kong XJ, et al. (2018). Host and cropping system shape
causing spikelet rot on rice in China. Plant Disease 105: 3306−3306. the Fusarium population: 3ADON-producers are ubiquitous in wheat
Wang MM, Chen Q, Diao YZ, et al. (2019a). Fusarium incarnatum-equiseti whereas NIV-producers are more prevalent in rice. Toxins 10: 115−127.
complex from China. Persoonia 43: 70−89. Yasuhara-Bell J, Pedley KF, Farman M, et al. (2018). Specific detection
Wang MM, Crous PW, Sandoval-Denis M, et al. (2022a). Fusarium and of the wheat blast pathogen (Magnaporthe oryzae Triticum) by loop-
allied genera from China: species diversity and distribution. Persoonia mediated isothermal amplification. Plant Disease 102: 2550−2559.
48: 1−53. Ye J, Guo Y, Zhang D, et al. (2013). Cytological and molecular
Wang Q, Song R, Fan S, et al. (2022b). Diversity of Fusarium community characterization of quantitative trait locus qRfg1, which confers
assembly shapes mycotoxin accumulation of diseased wheat heads. resistance to Gibberella stalk rot in maize. Molecular Plant-Microbe
Molecular Ecology: https://doi.org/10.1111/mec.16618. Interactions 26: 1417−1428.
Wang S, Sun L, Li WQ, et al. (2019b). First report of seedling blight caused Ye QM, Wang GC, Chen HK (1990). Identification of Fusarium moniliforme
by Fusarium redolens on rice in northeast China. Plant Disease 103: and its similar species. Jiangxi Plant Protection 3: 13−15.
1418−1418. Yilmaz N, Sandoval-Denis M, Lombard L, et al. (2021). Redefining species
Wang XM, Shi J, Jin QM, et al. (2010). Field handbook of corn disease and limits in the Fusarium fujikuroi species complex. Persoonia 46:
insects. Chinese Agricultural Science Technology Press, China. 129−162.
Ward TJ, Clear RM, Rooney AP, et al. (2008). An adaptive evolutionary Yin G, Bie S, Gu H, et al. (2020). Application of gene chip technology in
shift in Fusarium head blight pathogen populations is driving the rapid the diagnostic and drug resistance detection of Helicobacter pylori in
spread of more toxigenic Fusarium graminearum in North America. children. Journal of Gastroenterology and Hepatology 35: 1331−1339.
Fungal Genetics and Biology 45: 473−484. Yli-Mattila T, Paavanen-Huhtala S, Parikka P, et al. (2004). Molecular and
White TJ, Bruns T, Lee S, et al. (1990). Amplification and direct morphological diversity of Fusarium species in Finland and North-
sequencing of fungal ribosomal RNA genes for phylogenetics. In: Western Russia. European Journal of Plant Pathology 110: 573−585.
PCR protocols: a guide to methods and applications (Innis MA, Yu JJ, Zhang X, Tian JH, et al. (2022). Research and development
Gelfand DH, Sninsky JJ, White TJ, eds). Academic Press, San Diego of butterfly microfluidic gene chip for 19 common pathogenic
(California): 315–322. microorganisms of nosocomial infection. Journal of Clinical and
Wickham H (2016). ggplot2: elegant graphics for data analysis. R package Nursing Research 6: 80−85.
version 3.4.0. <https://ggplot2.tidyverse.org/>. Zhang H, Brankovics B, Luo W, et al. (2016). Crops are a main driver for
Wing N, Lauren DR, Bryden WL, et al. (1993). Toxicity and trichothecene species diversity and the toxigenic potential of Fusarium isolates in
production by Fusarium acuminatum subsp. acuminatum and Fusarium maize ears in China. World Mycotoxin Journal 9: 701−715.
acuminatum subsp. armeniacum. Natural Toxins 1: 229−234. Zhang H, Luo W, Pan Y, et al. (2014a). First report of Fusarium ear rot
Wollenweber HW (1913). Studies on the Fusarium problem. Phytopathology of maize caused by Fusarium andiyazi in China. Plant Disease 98:
3: 24–50. 1428−1428.
Wollenweber HW (1914). Identification of species of Fusarium occurring Zhang H, Luo W, Pan Y, et al. (2014b). First report of Fusarium maize
on the sweet potato. Journal of Agricultural Research 2: 251−286. ear rot caused by Fusarium meridionale in China. Plant Disease 98:
Wollenweber HW (1916). Fusaria autographice delineata. Selbstverlag, 1156−1156.
Berlin, Germany. Zhang H, Luo W, Pan Y, et al. (2014c). First report of Fusarium temperatum
Wollenweber HW, Reinking OA (1925). Aliquot Fusaria tropicalia nova vel causing Fusarium ear rot on maize in Northern China. Plant Disease
revisa. Phytopathology 15: 155−169. 98: 1273−1273.
Wulff EG, Sorensen JL, Lubeck M, et al. (2010). Fusarium spp. associated Zhang H, Van der Lee T, Waalwijk C, et al. (2012). Population analysis
with rice bakanae: ecology, genetic diversity, pathogenicity and of the Fusarium graminearum species complex from wheat in China
toxigenicity. Environmental Microbiology 12: 649−657. show a shift to more aggressive isolates. PLoS ONE 7: e31722.
Xi K, Haseeb HA, Shan L, et al. (2019). First report of Fusarium commune Zhang JB, Li HP, Dang FJ, et al. (2007). Determination of the trichothecene
causing stalk rot on maize in Liaoning Province, China. Plant Disease mycotoxin chemotypes and associated geographical distribution and
103: 773−773. phylogenetic species of the Fusarium graminearum clade from China.
Xi K, Shan L, Yang Y, et al. (2021). Species diversity and chemotypes of Mycological Research 111: 967−975.
Fusarium species associated with maize stalk rot in Yunnan Province Zhang J, Cao YY, Han SB, et al. (2021a). First report of Fusarium thapsinum
of Southwest China. Frontiers in Microbiology 12: 652062. causing maize stalk rot in China. Plant Disease 105: 2722−2722.
Xia JW, Sandoval-Denis M, Crous PW, et al. (2019). Numbers to names Zhang K, Su YY, Cai L (2013). An optimized protocol of single spore
− restyling the Fusarium incarnatum-equiseti species complex. isolation for fungi. Cryptogamie, Mycologie 34: 349−356.
Persoonia 43: 186−221. Zhang XJ, Guo C, Wang CM, et al. (2021). First report of maize stalk rot
Xia LK, Cao YY, Wang J, et al. (2021). First report of Fusarium culmorum caused by Fusarium nelsonii in China. Plant Disease 105: 4168−4168.
causing maize stalk rot in China. Plant Disease 106: 1521−1521. Zhang XX, Sun HY, Shen CM, et al. (2015). Survey of Fusarium spp.
Xu F, Liu W, Song Y, et al. (2021a). The distribution of Fusarium causing wheat crown rot in major winter wheat growing regions of
graminearum and Fusarium asiaticum causing Fusarium head blight China. Plant Disease 99: 1610−1615.
of wheat in relation to climate and cropping system. Plant Disease Zhang Y, Chen C, Mai Z, et al. (2022). Co-infection of Fusarium
105: 2830−2835. aglaonematis sp. nov. and Fusarium elaeidis causing stem rot in
Xu F, Song YL, Yang GQ, et al. (2015). First Report of Fusarium Aglaonema modestum in China. Frontiers in Microbiology 13: 1−14.
pseudograminearum from wheat heads with Fusarium head blight in Zhang Y, Ma LJ (2017). Deciphering pathogenicity of Fusarium oxysporum
North China Plain. Plant Disease 99: 156−156. from a phylogenomics perspective. Advances in Genetics 100:
Xu F, Wang JM, Yang GQ, et al. (2020). Occurrence of Fusarium poae from 179−209.
wheat heads with Fusarium head blight in Henan Province of China. Zhao ZH, Lu GZ (2007). Fusarium kyushuense, a newly recorded species
Plant Protection 46: 60−64. from China. Mycotaxon 102: 119−126.
Xu M, Zhang X, Yu J, et al. (2021b). First report of Fusarium ipomoeae Zhou LY, Yang SF, Wang SM, et al. (2020). Identification of Fusarium
causing peanut leaf spot in China. Plant Disease 105: 3754−3754. ipomoeae as the causative agent of leaf spot disease in Bletilla striata
Xu X, Yang GQ, Wang JM, et al. (2016). Composition and variation in in China. Plant Disease 105: 1204−1204.
aggressiveness of Fusarium populations causing wheat head blight in Zhuang WY (2005). Fungi of northwestern China. Mycotaxon Ltd. Ithaca,
Henan Province. Acta Phytopathologica Sinica 46: 294−303. China.

www.studiesinmycology.org 147
Han et al.

Fig. S7. Phylogeny of the Fusarium sambucinum species complex


Supplementary Material: https://studiesinmycology.org/
(FSAMSC) inferred based on the H3 (A), rpb1 (B), rpb2 (C), tef1 (D), and
H3-rpb1-rpb2-tef1 loci (E), respectively. Fusarium nelsonii (NRRL 13338)
Fig. S1. Sample collection by us and other people from various regions was used as an outgroup. GCP clade in Sarver et al. (2011) were indicated
in China. A. S.L. Han collecting rice spikelet rot samples at Jiangsu in green. Strains sequenced in this study were indicated in red. The RAxML
Province, Yangzhou city. B. Y.J. Li collecting diseased maize samples at Bootstrap support values (ML-BS > 70 %) were displayed at the nodes.
Shanxi Province, Yangquan city. C. M.Y. Wang collecting wheat crown Ex-type, ex-epitype and ex-neotype strains were indicated with T, ET, and
rot samples at Shandong Province, Taian city. D. S.Q. Wang collecting NT, respectively.
diseased maize samples at Guangdong Province, Qingyuan city. E. X.Y. Fig. S8. Phylogeny of the Fusarium tricinctum species complex (FTSC)
Liu collecting diseased maize samples at Shandong Province, Jinan city. inferred based on the ITS (A), rpb1 (B), rpb2 (C), and tef1 (D) loci,
F. Y.M. Wu collecting maize ear rot samples at Guangdong Province, respectively. Fusarium udum (CBS 177.31) was used as an outgroup.
Meizhou city. G. M.L. Feng collecting rice spikelet rot samples at Sichuan Strains sequenced in this study were indicated in red. The RAxML
Province, Mianyang city. H. G.J. Han collecting maize ear rot samples at Bootstrap support values (ML-BS > 70 %) were displayed at the nodes.
Shandong Province, Rizhao city. I–N. Local collectors showing photos of Ex-type, ex-epitype, and ex-neotype strains were indicated with T, ET, and
diseased cereals to us to confirm the occurrence of cereal diseases. O–S. NT, respectively.
We received the diseased samples and information from local collectors. Fig. S9. Genome sizes of the four cereal-associated species-rich
Fig. S2. Phylogeny inferred based on the combined rpb2-tef1 gene regions complexes (i.e., FFSC, FIESC, FOSC and FSAMSC) differs significantly
of 608 representative Fusarium strains. Neocosmospora nelsonii (CBS (P < 2e-16), average and standard deviation (av. ± SD) were annotated on
309.75) was used as an outgroup. Ex-type, ex-epitype and ex-neotype the corresponding boxplots.
strains were indicated with T, ET, and NT, respectively. Subdivision of the Fig. S10. The scatter plot of gCF values against sCF values for all branches
Fusarium clade represent the recognised species complexes, including F. of the phylogenetic tree of Fig. 14 (A) and Fig. 15 (B) is shown. The shade
fujikuroi SC (FFSC), F. incarnatum-equiseti SC (FIESC), F. nisikadoi SC of colour (bright to dark) indicated the support values of UFBoot (0–100).
(FNSC), F. oxysporum SC (FOSC), F. sambucinum SC (FSAMSC), and F. Fig. S11. The topology conflicts within the Fg clade, between the
tricinctum SC (FTSC), which were shown in different colours. phylogenomic tree and the muli-locus phylogenetic tree.
Fig. S3. Phylogeny of the Fusarium fujikuroi species complex (FFSC) Fig. S12. The total number of species discovered in samples collected
inferred based on the CaM (A), rpb1 (B), rpb2 (C), tef1 (D), and tub2 from various climate zones. The number on the diagram indicates the
(E) loci, respectively. Fusarium nirenbergiae (CBS 744.97) was used as species number detected from different climate zone, and the percentage
an outgroup. Strains sequenced in this study were indicated in red. The indicates the proportion of samples collected. The number of species co-
RAxML Bootstrap support values (ML-BS > 70 %) were displayed at the occurrence on one sample was represented by different colour.
nodes. Ex-type, ex-epitype and ex-neotype strains were indicated with T, Table S1. Collection details and GenBank accession numbers of 2 020
ET, and NT, respectively. Fusarium strains isolated in this study. The 608 representative Fusarium
Fig. S4. Phylogeny of the Fusarium incarnatum-equiseti species complex
strains are in bold.
(FIESC) inferred based on the CaM (A), rpb2 (B), and tef1 (C) loci,
Table S2. Number of characters/model for BI of each locus in the
respectively. Fusarium concolor (NRRL 13459) was used as an outgroup.
phylogenetic analyses of different Fusarium species complexes.
Strains sequenced in this study were indicated in red. The RAxML
Table S3. Fusarium species for which whole-genome sequences are
Bootstrap support values (ML-BS > 70 %) were displayed at the nodes.
retrieved from public databases, or generated in the current study
Ex-type, ex-epitype and ex-neotype strains were indicated with T, ET, and
(indicated in bold).
NT, respectively.
Table S4. The number of genes in each orthogroup for each species in
Fig. S5. Phylogeny of the Fusarium nisikadoi species complex (FNSC)
Fig. 14.
inferred based on the rpb1 (A), rpb2 (B), and tef1 (C) loci, respectively.
Table S5. The number of genes in each orthogroup for each species in
Fusarium concolor (NRRL 13994) was used as an outgroup. Strains
Fig. 15.
sequenced in this study were indicated in red. The RAxML Bootstrap
Table S6. Detail information of co-infection.
support values (ML-BS > 70 %) were displayed at the nodes. Ex-type
Table S7. Acknowledgement of participants.
strains were indicated with T.
Fig. S6. Phylogeny of the Fusarium oxysporum species complex (FOSC)
inferred based on the CaM (A), rpb2 (B), and tef1 (C) loci, respectively.
Fusarium udum (CBS 177.31) was used as an outgroup. Strains
sequenced in this study were indicated in red. The RAxML Bootstrap
support values (ML-BS > 70 %) were displayed at the nodes. Ex-type
strains were indicated with T.

148

You might also like