Next Article in Journal
Evaluation of Thermography as a Diagnostic Technique in Asymptomatic or Incipient Onychomycosis
Next Article in Special Issue
Genetic Diversity of Lecanosticta acicola in Pinus Ecosystems in Northern Spain
Previous Article in Journal
Insight into the Antifungal Effects of Propolis and Carnosic Acid—Extension to the Pathogenic Yeast Candida glabrata: New Propolis Fractionation and Potential Synergistic Applications
Previous Article in Special Issue
Races CYR34 and Suwon11-1 of Puccinia striiformis f. sp. tritici Played an Important Role in Causing the Stripe Rust Epidemic in Winter Wheat in Yili, Xinjiang, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fusarium mindanaoense sp. nov., a New Fusarium Wilt Pathogen of Cavendish Banana from the Philippines Belonging to the F. fujikuroi Species Complex

1
College of Agriculture, Tamagawa University, 6-1-1 Tamagawa-Gakuen, Machida, Tokyo 194-8610, Japan
2
Cancer Chemotherapy Center, Japanese Foundation for Cancer Research, 3-8-31 Ariake, Koto-ku, Tokyo 135-8550, Japan
3
Unifrutti Tropical Philippines, Inc., Km. 15, Panacan, Davao City 8000, Philippines
4
BaCaDM Project of College of Agriculture, Tamagawa University, 6-1-1 Tamagawa-Gakuen, Machida, Tokyo 194-8610, Japan
5
Unifrutti Japan Corporation, 1-11-1 Marunouchi, Chiyoda-Ku, Tokyo 100-6217, Japan
6
Philippine Center for Postharvest Development and Mechanization, Science City of Muñoz 3120, Philippines
7
Department of Biology, Central Luzon State University, Science City of Muñoz 3120, Philippines
*
Author to whom correspondence should be addressed.
J. Fungi 2023, 9(4), 443; https://doi.org/10.3390/jof9040443
Submission received: 16 February 2023 / Revised: 28 March 2023 / Accepted: 30 March 2023 / Published: 5 April 2023
(This article belongs to the Special Issue Fungal Plant Pathogens)

Abstract

:
The pathogen causing Fusarium wilt in banana is reported to be Fusarium oxysporum f. sp. cubense (FOC). In 2019, wilt symptoms in banana plants (cultivar: Cavendish) in the Philippines were detected, i.e., the yellowing of the leaves and discoloration of the pseudostem and vascular tissue. The fungus isolated from the vascular tissue was found to be pathogenic to Cavendish bananas and was identified as a new species, F. mindanaoense, belonging to the F. fujikuroi species complex (FFSC); species classification was assessed using molecular phylogenetic analyses based on the tef1, tub2, cmdA, rpb1, and rpb2 genes and morphological analyses. A reciprocal blast search using genomic data revealed that this fungus exclusively included the Secreted in Xylem 6 (SIX6) gene among the SIX homologs related to pathogenicity; it exhibited a highly conserved amino acid sequence compared with that of species in the FFSC, but not with that of FOC. This was the first report of Fusarium wilt in Cavendish bananas caused by a species of the genus Fusarium other than those in the F. oxysporum species complex.

1. Introduction

Bananas are one of the most common agricultural exports, though they are also widely consumed in the countries that produce them [1]. Musa sapientum cv. Cavendish (AAA group) is exclusively cultivated in many tropical countries as a commercial crop; 21 million tons of Cavendish bananas were exported in 2019 (FAO 2022). In the 1950s, the planting of Cavendish bananas instead of cv. Gros Michel began to increase worldwide because of an epidemic of Fusarium wilt disease (Panama disease) caused by Fusarium oxysporum f. sp. cubence (FOC) race 1. Thereafter, the causal pathogens of Fusarium wilt disease in bananas were found and characterized as race 1, race 2, race 4, subtropical race 4 (STR4), and tropical race 4 (TR4) based on their pathogenicity. In the 1990s, TR4 was identified as the causal agent of Fusarium wilt disease infecting Cavendish bananas in Taiwan [2]. The disease caused by TR4 has been reported in 23 countries (predominantly in Southeast Asia, South Asia, Africa, and Latin America) [3]. Moreover, TR4 affected the banana industry and reduced banana yield in the Philippines (FAO 2022). Mostert et al. [4] and Solpot et al. [5] investigated the Fusarium wilt pathogen in the Philippines; mostly TR4 (VCG01223/16), and less commonly R4 (VCG0122), was detected in Mindanao. However, information on FOC other than TR4 is scarce and no reports of other Fusarium species are available.
In this study, during a survey of the Fusarium wilt disease in Mindanao conducted in September 2019, a new species was found that belonged to the F. fujikuroi species complex (FFSC); it caused symptoms of leaf yellowing (Figure 1A) to emerge in older leaves and a reddish-brown discoloration of the pseudostem and vascular tissues of bananas (Cavendish) (Figure 1B,C). This pathogen had not been previously reported to cause Fusarium wilt in banana. Therefore, we aimed to identify this causal agent, conducted molecular and morphological analyses, and proposed an isolate as a new pathogenic species of Fusarium wilt.
Furthermore, to provide fundamental information relating to factors of its pathogenicity, we searched the Secreted in Xylem (SIX) genes from the whole-genome data of the new species. In addition, we predicted whether the genome of the new Fusarium species obtained SIX genes via horizontal transfer from FOC or by other means; it is reported that SIX genes are one of the most important factors for infecting banana plants [6,7,8].

2. Materials and Methods

2.1. Sample Collection and Fungal Isolation

Symptomatic banana plants were collected from a farm in Mindanao in 2019. The discolored vascular tissues (Figure 1C) of the pseudostem were cut into pieces of approximately 3 mm2, which were then sterilized with 0.6% (v/v) sodium hypochlorite for 1 min, washed with sterilized water, dried with sterilized paper, and placed on a water agar (WA) plate. The hyphae that emerged on WA were transferred onto a potato dextrose agar (PDA) plate to produce conidia for monoculture. The PD20-05 isolate was maintained on a PDA plate.

2.2. Genomic DNA Extraction

DNA was extracted from the mycelia of each isolate, which were grown for 7–10 days in yeast glucose medium using the modified CTAB method [9]. After treatment with chloroform–isoamyl alcohol (24:1), 2-Mercaptoethanol and 10% CTAB at 0.2% and 2%, respectively, were added to the supernatant and incubated for 40 min at 60 °C. After incubation, an equal volume of chloroform–isoamyl alcohol (24:1) was added, mixed gently for 10 min, and centrifuged for 10 min at 12,000 rpm for purification. The aqueous phase was carried out, and the above-mentioned purification was again conducted. Precipitation was achieved by adding 2.5 and 0.1 times the volume of ethanol and 3 M sodium acetate, respectively, which was then mixed for a short period of time and centrifuged for 10 min at 12,000 rpm. After removing the liquid, a pellet of DNA at the bottom was dried and dissolved with 30 µL of TE buffer.

2.3. Gene Prediction

Genome DNA was sequenced using the Illumina HiSeq genome analyzer platform and DNA libraries and paired-end (PE) genomic libraries were generated. The libraries were sequenced in PE mode with 150 bp reads on the Illumina HiSeq X instrument. Adaptors were eliminated from reads using the Trimmomatic read trimming tool for Illumina NGS data, with a quality cut-off of 30. The raw mate–pair read sequence quality was checked using FastQC vers. 0.11.8 [10] (http://www.bioinformatics.babraham.ac.uk/projects/fastqc accessed on 12 December 2022). Platanus allee vers. 2.0.2 [11] was used to assemble the reads and obtain contig data. The N50 values were calculated to measure the quality of the assemblies. Augustus 3.3.3 [12] was used to perform gene predictions using F. graminearum data as a reference.

2.4. Phylogenetic Analyses

Molecular analyses were conducted to identify the pathogen. To select the DNA sequences, the translation elongation factor 1-alpha (tef1), beta-tubulin (tub2), calmodulin (cmdA), RNA polymerase large subunit (rpb1), and RNA polymerase second-largest subunit (rpb2) genes were amplified according to the method reported by Yilmaz et al. [13]; the genes were then sequenced using the following primer pairs: EF1 and EF2 [14], T1 and T2 [15], CL1 and CL2A [16], Fa [17] and R8 [18], and 5F2 [19] and 7cr [20], respectively. The sequence data were deposited in the DNA Data Bank of Japan (Table 1). One hundred and three sequences (Table 1) of each DNA region were aligned using Clustal W in MEGA 7 [21], concatenated, and subjected to phylogenetic analyses using the maximum-likelihood (ML), maximum-parsimony (MP), and neighbor-joining (NJ) methods. The reliability of the branches on the phylogenetic tree was evaluated using the bootstrap (BS) [22] test with 1000 replicates.

2.5. Morphological Analyses

Mycelial plugs (φ7 mm) of the isolate were placed in the center of the potato dextrose agar (PDA), synthetic nutrient-poor agar (SNA) [23], and oatmeal agar (OA) [24] plates and incubated for 6 days at 25 °C in the dark. The colony character on the surface and reverse sides was observed. The isolate (PD20-05) was cultured on carnation leaf agar (CLA) [25] and SNA, inducing sporodochial conidia and microconidia to observe its asexual morphological characteristics. Thirty conidia and conidiophores were observed under a light microscope (BX51, Olympus, Tokyo, Japan) to record their shape and size. For the mycelial growth test, 6 d cultures of the isolate grown at 25 °C on PDA plates were used. Mycelial plugs (φ7 mm) were then placed on the center of the PDA plates. These plates were incubated at 4 °C, 10 °C, 15 °C, 20 °C, 25 °C, 30 °C, and 35 °C in the dark. After incubation for 6 days, mycelial growth per day was calculated. The average growth rate per day for each temperature was determined from five replicates.

2.6. Pathogenicity Test

A pathogenicity test was conducted using a conidial suspension in sterilized water adjusted to 1 × 107 conidia/mL. Six Cavendish seedlings were used in this experiment, in which the roots of three seedlings were soaked in 500 mL of the conidial suspension for 3 h, before being planted in pots with a 1:1 mixture of red ball earth and humus. A 5 g/L solution of NPK 8-8-8 was added as a chemical fertilizer. The remaining three seedlings were treated with sterilized water as a control. The treated and control plants were inoculated for 34 days at 25 °C with an 8 h light/16 h dark cycle.

2.7. Detection of Secreted in Xylem Genes among the Whole-Genome Data

A tBLASTn analysis was conducted to identify the SIX genes using BLAST 2.11.0+ software [26,27]. In this analysis, previously reported SIX gene protein sequences were obtained from the NCBI protein database and were used as queries against the assembled whole-genome sequence (e-value = 0.001). The identified new SIX gene sequence was reciprocally BLAST searched (BLASTP) against the NCBI-deposited protein sequences of SIX genes to estimate their sequence similarity. The SignalP program (v. 5.0) [28] was used to identify the new SIX protein code signal peptides.

2.8. Identification of the Homologous SIX Genes of PD20-05 in the Foc TR4 Genome

To search for the homologous SIX genes of PD20-05 in Fusarium oxysporum f. sp. cubense (FOC) TR4, we initially constructed the genome sequence of FOC TR4 using the NGS data (SRR10054447) [29] and Platanus-allee. Subsequently, the SIX gene of PD20-05 was used as a query against the assembled FOC TR4 genome sequence (e-value ≤ 0.001).
To clarify whether the SIX6 gene identified in our isolate was from FOC, the putative SIX6 gene protein sequences of our isolate and those of FFSC and FOC were aligned using clustalW in MEGA7 [21], followed by the construction of an NJ tree with the option to completely delete the gap. The reliability of the branches of the phylogenetic tree was evaluated via the BS test [22] with 1000 replicates.

3. Results

3.1. Phylogenetic Analysis

For the phylogenetic analysis using the five loci, the final dataset included 2606 positions (excluding gaps and including sites), comprising 462, 280, 521, 594, and 749 positions from tef1, tub2, cmdA, rpb1, and rpb2 gene sequences, respectively. The PD20-05 isolate was independent of known species and a sister lineage of the F. sacchari clade (BS value = 76; Figure 2).

3.2. Taxonomy

  • Fusarium mindanaoense Nozawa & Watanabe, sp. nov.
  • Mycobank MB 848129; Figure 3.
  • Etymology: the name refers to Mindanao, the region where the ex-type strain was obtained.
  • Holotype: PD20-05S.
  • Ex-holotype: PD20-05.
Figure 3. Colony morphology of F. mindanaoense (PD20-05T; ex-type culture PD20-05) after 6 days growth at 25 °C in the dark: (A). PDA; (B). OA; (C). SNA. Colony surface is shown on left half of each plate and colony undersurface on right half. (D). Conidiophore on carnation leaf; (E). sporodochia on carnation leaf; (F,G). microconidia on a conidiophore on aerial hyphae on SNA; (H,I). microconidia on a conidiophore on hyphae inside SNA; (J). microconidia on a carnation leaf; (K). microconidia on SNA; (L). conidiophores and phialides on sporodochia; (MO). sporodochial conidia (macroconidia): (M). 3-septate conidia; (N). 4-septate conidia; (O). 5-septaate conidia. Scale bars: 20 μm.
Figure 3. Colony morphology of F. mindanaoense (PD20-05T; ex-type culture PD20-05) after 6 days growth at 25 °C in the dark: (A). PDA; (B). OA; (C). SNA. Colony surface is shown on left half of each plate and colony undersurface on right half. (D). Conidiophore on carnation leaf; (E). sporodochia on carnation leaf; (F,G). microconidia on a conidiophore on aerial hyphae on SNA; (H,I). microconidia on a conidiophore on hyphae inside SNA; (J). microconidia on a carnation leaf; (K). microconidia on SNA; (L). conidiophores and phialides on sporodochia; (MO). sporodochial conidia (macroconidia): (M). 3-septate conidia; (N). 4-septate conidia; (O). 5-septaate conidia. Scale bars: 20 μm.
Jof 09 00443 g003
Colonies on PDA reached 50–54 mm diam. after 6 d at 25 °C in the dark; the colonies were raised, aerial mycelia dense, covered the entire margin and surface in the center, and were white at the edge. Colonies were also reverse pink in the center and white at the margin. Hyphae grew at 10–35 °C, with an optimum temperature of 25 °C (avg. ± sd. 4.4 ± 0.16 mm/day; Figure 4). Colonies on OA reached 66–68 mm diam. after 6 d at 25 °C in the dark were raised, aerial mycelia dense, and covered colony margin entire; they were also surface white and reverse white. Colonies on SNA reached 59–61 mm diam. after 6 days at 25 °C in the dark; these colonies were raised, aerial mycelia sparse, covered the entire colony margin entire, and were surface white and reverse white.
Sporodochia milk white formed on carnation leaves deficiently. Conidiophores in sporodochia were verticillately branched; bearing apical pairs were monophialide, while sporodochial phialides subulate to subcylindrical. Sporodochial macroconidia falcate were moderately curved and slender with parallel side tapering slightly toward both ends, as well as being papillate, 3–5-septate, hyaline, thin- and smooth-walled. The 3-septate conidia had dimensions of 37.7–52.6 × 3.5–5.0 (av. ± sd. 45.9 ± 3.8 × 4.2 ± 0.35) µm, while 4-septate conidia had dimensions of 50.4–66.4 × 3.1–4.6 (av. ± sd. 57.6 ± 3.4 × 3.7± 0.39) µm; 5-septate conidia had dimensions of 55.5–67.8 × 3.2–4.1 (av. ± sd. 62.1 ± 4.3 × 4.1 ± 0.7) µm. Conidiophores borne on aerial mycelia on carnation leaf were branched, while those borne on aerial mycelia SNA, bearing chained microconidia, or with values of 15.7–42 (av. ± sd. 24.2 ± 7) μm tall were either unbranched or rarely branched, instead bearing terminal monophialide. Those borne inside SNA were 0–24 (av. ± sd. 8 ± 7.5) µm tall and unbranched; they were had microconidia hyaline, oval, pyriform, smooth- and thin-walled aseptate. The microconidia on carnation leaf was 6.6–13 × 2–3.4 (av. ± sd. 8.6 ± 1.5 × 2.7 ± 0.38) µm, while microconidia on SNA was 7.6–15.8 × 2.2–3.8 (av. ± sd. 10.1 ± 2 × 2.9 ± 0.39) µm. Chlamydospores were not observed.
Note: F. mindanaoense resembled F. concentricum regarding the size of sporodochial conidia (Table 2). However, F. mindanaoense could be distinguished by the characteristics of colonies. F. mindanaoense did not produce concentric aerial hyphae in its mycelium (Figure 3A), while F. concentricum did produce this symptom of fungal infection. A holotype and ex-holotype strain were deposited at Flora and Fauna Analytical and Diagnostic Center at Central Luzon State University.

3.3. Pathogenicity Test

Yellow leaves appeared on inoculated plants after 20–34 days; one dried-up seedling and leaves of other seedlings closed around the main veins (Figure 5A,B). Part of the internal tissues of the corms turned black, while the tissues of pseudostem just above the corn were reddish-brown (Figure 5C,D). Additionally, the roots turned black all around (Figure 5E,F). The inoculated strain was re-isolated from the discolored roots and vascular lesions, whereas the control plants treated with water exhibited no symptoms.

3.4. Detection of Secreted in Xylem Genes in Whole-Genome Data

In this study, the genomic DNA of PD20-05 was sequenced and assembled into 3377 contigs with an N50 of 53.7 kb and a maximum length of 164.3 kb. In the tBLASTn analysis, SIX gene sequences were searched in the PD20-05 genome using the 1186 NCBI-deposited SIX protein sequences. The analysis showed that the SIX6 gene was the only SIX gene found in the PD20-05 genome sequence. Moreover, the SIX6 protein of PD20-05 was predicted to contain signal peptides (Figure S1). Furthermore, a reciprocal BLASTp analysis showed that the SIX6 protein sequence of PD20-05 was identical to that of Fusarium sp. NRRL 25303, F. proliferatum, F. globosum, F. agapanthi, F. denticulatum, F. tjaetaba, F. napiforme, F. pseudocircinatum, F. circinatum, F. phyllophilum, F. mundagurra, and F. pseudoanthophilum (Table 3), which all belong to the FFSC species and made one clade in the phylogenetic tree (Figure 6). These SIX6 gene sequences were greatly different from those of two FOC strains (accession nos. KX435007 and KX435008) [8], as assessed based on the alignment (Figure S1).
Two types of SIX6 genes in the F. hostae (HY9) genome were obtained by conducting a BLASTp using the SIX6 gene sequences of FOC (BRIP628956) and F. mindanaoense (PD20-05) as query sequences with low e-values (3 × 10−82 and 2 × 10−120, respectively). Van Dam and Rep [33] reported that the strain acquired one SIX6 gene via horizontal transfer from the FOC. The two types of SIX6 genes fell into different clades in the phylogenetic tree (Figure 6).

4. Discussion

Fusarium sacchari (leaf blight on AAA genome group and fruit rot on AAA), F. proliferatum (fruit rot on AAB and sheath rot on ABB), F. fujikuroi (fruit rot on AA), F. concentricum (fruit rot on AAA), F. verticillioides (fruit rot on Musa sp.), and F. musae (fruit rot on Musa sp.) belonging to FFSC were reported as banana pathogens [34,35,36,37,38,39,40,41]. These species do not cause Fusarium wilt of bananas. Maryani et al. [39] also isolated F. proliferatum from a symptomatic tissue of Fusarium wilt of banana (AA) in 2019, concluding that the fungus was not a pathogen of Fusarium wilt of banana (Cavendish: AAA); rather, it was an endophyte because it caused only a slight discoloration in the corm without any further disease development. Moreover, in 2022, Thi et al. [42] also isolated FFSC species (F. fujikuroi) from symptomatic tissues of Fusarium wilt of banana (ABB). However, a pathogenicity assay was not carried out. To the best of our knowledge, no FFSC species have been reported as the pathogen underlying Fusarium wilt in banana. In this study, we identified a new causal agent, F. mindanaoense (which belongs to FFSC), of Fusarium wilt in banana in the Philippines. This is the first report of Fusarium wilt in banana caused by a fungus belonging to the FFSC.
As FOC affects Cavendish bananas, research on FOC has focused on managing Fusarium wilt disease. Therefore, rapid detection methods for FOC, such as loop-mediated isothermal amplification and PCR detection, have been developed for diagnosis and occurrence monitoring [43,44,45]. Our study reveals that a pathogen belonging to the FFSC also caused Fusarium wilt in the Cavendish banana. Focusing on FOC and other pathogenetic fungi to acquire basic knowledge that may contribute to controlling Fusarium wilt is necessary.
The SIX genes play a role in the pathogenicity of Fusarium wilt; SIX1, SIX2, SIX6, SIX7, SIX9G1, SIX11, and SIX13 were detected in the FFSC species [34,37]. The present study showed that the F. mindanaoense genome possessed the SIX6 gene exclusively, which matched with those of the FFSC with low e-values (Table 2; 0–3.92 × 10−125). Van Dam and Rep [33] reported that the SIX6 gene from F. hostae (HY9), which belongs to the FFSC species, was horizontally transferred from FOC. We found that F. hostae (HY9) has two types of SIX6 genes: the FOC and FFSC groups (Figure 6). Because the gene sequence of F. mindanaoense that was identified as the SIX6 gene did not belong to a clade of FOC, and one of the SIX6 genes obtained from F. hostae genome belonged to the FFSC in the phylogenetic tree, F. mindanaoense was thought not to have acquired its pathogenicity through horizontal gene transfer from FOC (Figure 6). However, a functional analysis of the SIX6 gene of the FFSC is warranted to clarify whether the SIX6 gene acts as a functional gene in the pathogenicity of Fusarium wilt.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/jof9040443/s1, Figure S1: Multiple sequence alignment of SIX6 genes.

Author Contributions

Conceptualization, S.N. and K.W.; methodology, S.N. and K.W.; validation, S.N., K.W. and Y.S.; formal analysis, S.N. and Y.S.; investigation, S.N., K.W., Y.T., L.A.N., R.R.V., K.O., S.T., R.G.R. and D.G.A.; writing—original draft preparation, S.N.; writing—review and editing, K.W., S.N., R.R.V., R.G.R. and D.G.A.; supervision, K.W.; project administration, K.W.; funding acquisition, K.W. All authors have read and agreed to the published version of the manuscript.

Funding

JST SATREPS Grant Number JPMJSA2007 supported this study.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are available upon reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Nayar, N.M. The bananas: Botany, origin, dispersal. Hortic. Res. 2010, 36, 118–164. [Google Scholar]
  2. Ploetz, R.C. Panama disease: Return of the first banana menace. Int. J. Pest. Manag. 1994, 40, 326–336. [Google Scholar] [CrossRef]
  3. Catambacan, D.G.; Cumagun, C.J. Weed-associated fungal endophytes as biocontrol agents of Fusarium oxysporum f. sp. cubense TR4. J. Fungus 2021, 7, 224. [Google Scholar] [CrossRef]
  4. Mostert, D.; Molina, A.B.; Daniells, J.; Fourie, G.; Hermanto, C.; Chao, C.P.; Fabregar, E.; Sinohin, V.G.; Masdek, N.; Thangavelu, R.; et al. The distribution and host range of the banana Fusarium wilt fungus, Fusarium oxysporum f. sp. cubense, in Asia. PLoS ONE 2017, 12, e0181630. [Google Scholar] [CrossRef] [Green Version]
  5. Solpot, T.C.; Pangga, I.B.; Baconguis, R.D.; Cumagun, C.J. Occurrence of Fusarium oxysporum f. sp. cubense tropical race 4 and other genotypes in banana in south-Central Mindanao, Philippines. Philipp. Agric. Sci. 2016, 99, 370–378. [Google Scholar]
  6. Fraser-Smith, S.; Czislowski, E.; Meldrum, R.A.; Zander, M.; Balali, G.R.; Aitken, E.A.B. Sequence variation in the putative effector gene SIX8 facilitates molecular differentiation of Fusarium oxysporum f. sp. cubense. Plant Pathol. 2014, 63, 1044–1052. [Google Scholar] [CrossRef]
  7. Guo, L.; Han, L.; Yang, L.; Zeng, H.; Fan, D.; Zhu, Y.; Feng, Y.; Wang, G.; Peng, C.; Jiang, X.; et al. Genome and transcriptome analysis of the fungal pathogen Fusarium oxysporum f. sp. cubense causing banana vascular wilt disease. PLoS ONE 2014, 9, e95543. [Google Scholar] [CrossRef]
  8. Farah, N.; Belaffif, M.B.; Aryantha, I.N.P.; Esyanti, R.R. SIX6 shows a high divergence in Fusarium oxysporum f. sp. cubense TR4. Int. J. Agric. Biol. 2021, 25, 1331–1338. [Google Scholar] [CrossRef]
  9. Doyle, J.J.; Doyle, J.L. A rapid DNA isolation procedure for small quantities of fresh leaf tissue. Phytochem. Bull. 1987, 19, 11–15. [Google Scholar]
  10. Brown, J.; Pirrung, M.; McCue, L.A. FQC Dashboard: Integrates FastQC results into a web-based, interactive, and extensible FASTQ quality control tool. Bioinformatics 2017, 33, 3137–3139. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Kajitani, R.; Yoshimura, D.; Okuno, M.; Minakuchi, Y.; Kagoshima, H.; Fujiyama, A.; Kubokawa, K.; Kohara, Y.; Toyoda, A.; Itoh, T. Platanus-allee is a de novo haplotype assembler enabling a comprehensive access to divergent heterozygous regions. Nat. Commun. 2019, 10, 1702. [Google Scholar] [CrossRef] [Green Version]
  12. Stanke, M.; Morgenstern, B. Augustus: A web server for gene prediction in eukaryotes that allows user-defined constraints. Nucleic Acids Res. 2005, 33, W465–W467. [Google Scholar] [CrossRef] [Green Version]
  13. Yilmaz, N.; Sandoval-Denis, M.; Lombard, L.; Visagie, C.M.; Wingfield, B.D.; Crous, P.W. Redefining species limits in the Fusarium fujikuroi species complex. Persoonia 2021, 46, 129–162. [Google Scholar] [CrossRef] [PubMed]
  14. O’Donnell, K.; Cigelnik, E.; Nirenberg, H.I. Molecular systematics and phylogeography of the Gibberella fujikuroi species complex. Mycologia 1998, 90, 465–493. [Google Scholar] [CrossRef]
  15. O’Donnell, K.; Kistler, H.C.; Cigelnik, E.; Ploetz, R.C. Multiple evolutionary origins of the fungus causing Panama disease of banana: Concordant evidence from nuclear and mitochondrial gene genealogies. Proc. Natl. Acad. Sci. USA 1998, 95, 2044–2049. [Google Scholar] [CrossRef] [Green Version]
  16. O’Donnell, K.; Nirenberg, H.I.; Aoki, T.; Cigelnik, E. A multigene phylogeny of the Gibberella fujikuroi species complex: Detection of additional phylogenetically distinct species. Mycoscience 2000, 41, 61–78. [Google Scholar] [CrossRef]
  17. Hofstetter, V.; Miadlikowska, J.; Kauff, F.; Lutzoni, F. Phylogenetic comparison of protein-coding versus ribosomal RNA-coding sequence data: A case study of the Lecanoromycetes (Ascomycota). Mol. Phylogenet. Evol. 2007, 44, 412–426. [Google Scholar] [CrossRef] [PubMed]
  18. O’Donnell, K.; Sutton, D.A.; Rinaldi, M.G.; Sarver, B.A.; Balajee, S.A.; Schroers, H.J.; Summerbell, R.C.; Robert, V.A.; Crous, P.W.; Zhang, N.; et al. Internet-accessible DNA sequence database for identifying fusaria from human and animal infections. J. Clin. Microbiol. 2010, 48, 3708–3718. [Google Scholar] [CrossRef] [Green Version]
  19. Reeb, V.; Lutzoni, F.; Roux, C. Contribution of RPB2 to multilocus phylogenetic studies of the euascomycetes (Pezizomycotina, fungi) with special emphasis on the lichen-forming Acarosporaceae and evolution of polyspory. Mol. Phylogenet. Evol. 2004, 32, 1036–1060. [Google Scholar] [CrossRef]
  20. Liu, Y.J.; Whelen, S.; Hall, B.D. Phylogenetic relationships among ascomycetes: Evidence from an RNA polymerse II subunit. Mol. Biol. Evol. 1999, 16, 1799–1808. [Google Scholar] [CrossRef] [Green Version]
  21. Kumar, S.; Stecher, G.; Tamura, K. MEGA7: Molecular Evolutionary Genetics Analysis Version 7.0 for Bigger Datasets. Mol. Biol. Evol. 2016, 33, 1870–1874. [Google Scholar] [CrossRef] [Green Version]
  22. Felsenstein, J. Confidence limits on phylogenies: An approach using the bootstrap. Evolution 1985, 39, 783–791. [Google Scholar] [CrossRef]
  23. Nirenberg, H.I.; O’Donnell, K. New Fusarium species and combinations within the Gibberella fujikuroi species complex. Mycologia 1998, 90, 434–458. [Google Scholar] [CrossRef]
  24. Booth, C. Fusarium. Laboratory Guide to the Identification of the Major Species; Commonwealth Mycological Institute: Kew, UK, 1977; 58p. [Google Scholar]
  25. Fisher, N.L.; Burgess, W.; Toussoun, T.A.; Nelson, P.E. Carnation leaves as a substrate and for preserving cultures of Fusarium species. Phytopathology 1982, 72, 151–153. [Google Scholar] [CrossRef]
  26. Altschul, S.F.; Gish, W.; Miller, W.; Myers, E.W.; Lipman, D.J. Basic local alignment search tool. J. Mol. Biol. 1990, 215, 403–410. [Google Scholar] [CrossRef]
  27. Camacho, C.; Coulouris, G.; Avagyan, V.; Ma, N.; Papadopoulos, J.; Bealer, K.; Madden, T.L. Blast+: Architecture and applications. BMC Bioinform. 2009, 10, 421. [Google Scholar] [CrossRef] [Green Version]
  28. Almagro Armenteros, J.J.; Tsirigos, K.D.; Sønderby, C.K.; Petersen, T.N.; Winther, O.; Brunak, S.; von Heijne, G.; Nielsen, H. SignalP 5.0 improves signal peptide predictions using deep neural networks. Nat. Biotechnol. 2019, 37, 420–423. [Google Scholar] [CrossRef] [Green Version]
  29. Leiva, A.M.; Rouard, M.; Lopez-Alvarez, D.; Cenci, A.; Breton, C.; Acuña, R.; Rojas, J.C.; Dita, M.; Cuellar, W.J. Draft genome sequence of Fusarium oxysporum f. sp. cubense tropical race 4 from Peru, obtained by nanopore and illumina hybrid assembly. Microbiol. Resoure Announc. 2022, 11, e00347-22. [Google Scholar] [CrossRef]
  30. Bugnicourt, P. Note on the mycoflora of rice seed in the territories of the South Pacific. Rev. Mycol. 1952, 17, 26–29. [Google Scholar]
  31. Britz, H.; Steenkamp, E.T.; Coutinho, T.A.; Wingfield, B.D.; Marasas, W.F.; Wingfield, M.J. Two new species of Fusarium section Liseola associated with mango malformation. Mycologia 2002, 94, 722–730. [Google Scholar] [CrossRef]
  32. Nirenberg, H. Untersuchungen über die morphologische und biologische Differenzierung in der Fusarium-Sektion Liseola. In Biologische Bundesanstalt für Land- und Forstwirtschaft; Mitteilungen aus der Biologischen Bundesanstalt für Land- und Forstwirtschaft, Berlin-Dahlem; Kommissionsverlag Paul Parey: Berlin, Germany, 1976; Volume 169, pp. 1–117. [Google Scholar]
  33. Van Dam, P.; Rep, M. The distribution of miniature impala elements and SIX genes in the Fusarium genus is suggestive of horizontal gene transfer. J. Mol. Evol. 2017, 85, 14–25. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Cui, Y.; Wu, B.; Peng, A.; Song, X.; Chen, X. The genome of banana leaf blight pathogen Fusarium sacchari str. FS66 harbors widespread gene transfer from Fusarium oxysporum. Front. Plant Sci. 2021, 12, 629859. [Google Scholar] [CrossRef]
  35. Hirata, T.; Kimishima, E.; Aoki, T.; Nirenberg, H.I.; O’Donnell, K. Morphological and molecular characterization of Fusarium verticillioides from rotten banana imported into Japan. Mycoscience 2001, 42, 155–166. [Google Scholar] [CrossRef]
  36. Maldonado-Bonilla, L.D.; Calderón-Oropeza, M.A.; Villarruel-Ordaz, J.L.; Sánchez-Espinosa, A.C. Identification of novel potential causal agents of Fusarium wilt of Musa sp. AAB in southern Mexico. J. Plant Pathol. Microbiol. 2019, 10, e479. [Google Scholar] [CrossRef] [Green Version]
  37. Czislowski, E.; Zeil-Rolfe, I.; Aitken, E.A.B. Effector profiles of endophytic Fusarium associated with asymptomatic banana (Musa sp.) Hosts. Int. J. Mol. Sci. 2021, 22, 2508. [Google Scholar] [CrossRef]
  38. Huang, S.P.; Wei, J.G.; Guo, T.X.; Li, Q.L.; Tang, L.H.; Mo, J.Y.; Wei, J.F.; Yang, X.B. First report of sheath rot caused by Fusarium proliferatum on pisang awak banana (Musa ABB) in China. J. Plant Pathol. 2019, 101, 1271–1272. [Google Scholar] [CrossRef] [Green Version]
  39. Maryani, N.; Sandoval-Denis, M.; Lombard, L.; Crous, P.W.; Kema, G.H.J. New endemic Fusarium species hitch-hiking with pathogenic Fusarium strains causing Panama disease in small-holder banana plots in Indonesia. Pers. Mol. Phylogeny Evol. Fungi 2019, 43, 48–69. [Google Scholar] [CrossRef] [Green Version]
  40. Molnár, O.; Bartók, T.; Szécsi, Á. Occurrence of Fusarium verticillioides and Fusarium musae on banana fruits marketed in Hungary. Acta Microbiol. Immunol. Hung. 2015, 62, 109–119. [Google Scholar] [CrossRef] [Green Version]
  41. Van Hove, F.; Waalwijk, C.; Logrieco, A.; Munaut, F.; Moretti, A. Gibberella musae (Fusarium musae) sp. nov., a recently discovered species from banana is sister to F. verticillioides. Mycologia 2011, 103, 570–585. [Google Scholar] [CrossRef] [Green Version]
  42. Thi, L.L.; Mertens, A.; Vu, D.T.; Vu, T.D.; Minh, P.L.A.; Duc, H.N.; De Backer, S.; Swennen, R.; Vandelook, F.; Panis, B.; et al. Diversity of Fusarium associated banana wilt in northern Viet Nam. MycoKeys 2022, 87, 53–76. [Google Scholar] [CrossRef]
  43. Zhang, X.; Zhang, H.; Pu, J.; Qi, Y.; Yu, Q.; Xie, Y.; Peng, J. Development of a real-time fluorescence loop-mediated isothermal amplification assay for rapid and quantitative detection of Fusarium oxysporum f. sp. cubense tropical race 4 in soil. PLoS ONE 2013, 8, e82841. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Ordóñez, N.; Salacinas, M.; Mendes, O.; Seidl, M.F.; Meijer, H.J.G.; Schoen, C.D.; Kema, G.H.J. A loop-mediated isothermal amplification (LAMP) assay based on unique markers derived from genotyping by sequencing data for rapid in planta diagnosis of Panama disease caused by tropical Race 4 in banana. Plant Pathol. 2019, 68, 1682–1693. [Google Scholar] [CrossRef] [Green Version]
  45. Thangavelu, R.; Edwinraj, E.; Gopi, M.; Pushpakanth, P.; Sharmila, K.; Prabaharan, M.; Loganathan, M.; Uma, S. Development of PCR-based race-specific markers for differentiation of Indian Fusarium oxysporum f. sp. cubense, the causal agent of fusarium wilt in banana. J. Fungi 2022, 8, 53. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Infection symptoms of bananas (cv. Cavendish). (A) yellowish leaves; (B) brownish pseudostem; (C) brownish xylem.
Figure 1. Infection symptoms of bananas (cv. Cavendish). (A) yellowish leaves; (B) brownish pseudostem; (C) brownish xylem.
Jof 09 00443 g001
Figure 2. Maximum likelihood (ML) tree based on combined data sets of tef1α, tub2, cmdA, rpb1, and rpb2 sequences. ML, maximum-parsimony (MP), and neighbor-joining (NJ) bootstrap values are indicated at the nodes as ML/MP/NJ. The hyphen (“-”) indicates that a node is not present. “T” indicates the ex-type and ex-epitype strains.
Figure 2. Maximum likelihood (ML) tree based on combined data sets of tef1α, tub2, cmdA, rpb1, and rpb2 sequences. ML, maximum-parsimony (MP), and neighbor-joining (NJ) bootstrap values are indicated at the nodes as ML/MP/NJ. The hyphen (“-”) indicates that a node is not present. “T” indicates the ex-type and ex-epitype strains.
Jof 09 00443 g002aJof 09 00443 g002b
Figure 4. Mycelial growth rate of F. mindanaoense PD20-05 on PDA depending on the temperature in the dark for 6 days.
Figure 4. Mycelial growth rate of F. mindanaoense PD20-05 on PDA depending on the temperature in the dark for 6 days.
Jof 09 00443 g004
Figure 5. Pathogenicity test of F. mindanaoense PD20-05 using bananas (cv. dwarf Cavendish). (A). Wilting symptoms 34 days after inoculation with F. mindanaoense PD20-05. (B). Control plants without inoculation with F. mindanaoense PD20-05. (C). A tuber of the inoculated plant with blackish tissues (red arrows) and discolored tissues (black arrows). (D). A tuber of the control plant. (E). The roots of the inoculated plant. (F). The roots of the control plant.
Figure 5. Pathogenicity test of F. mindanaoense PD20-05 using bananas (cv. dwarf Cavendish). (A). Wilting symptoms 34 days after inoculation with F. mindanaoense PD20-05. (B). Control plants without inoculation with F. mindanaoense PD20-05. (C). A tuber of the inoculated plant with blackish tissues (red arrows) and discolored tissues (black arrows). (D). A tuber of the control plant. (E). The roots of the inoculated plant. (F). The roots of the control plant.
Jof 09 00443 g005
Figure 6. NJ phylogenetic tree based on SIX6 gene sequences. The genes highlighted in red are the SIX6 genes of FFSC, whereas the genes highlighted in green are the SIX6 genes of FOC and F. hostae. F. hostae has two types of SIX6 genes. One belongs to FFSC type (a), and another one belongs to FOC type (b).
Figure 6. NJ phylogenetic tree based on SIX6 gene sequences. The genes highlighted in red are the SIX6 genes of FFSC, whereas the genes highlighted in green are the SIX6 genes of FOC and F. hostae. F. hostae has two types of SIX6 genes. One belongs to FFSC type (a), and another one belongs to FOC type (b).
Jof 09 00443 g006
Table 1. Strains of the Fusarium fujikuroi species complex used in this study with GenBank accession number.
Table 1. Strains of the Fusarium fujikuroi species complex used in this study with GenBank accession number.
SpeciesCulture CollectionsGenBank Accession Number
tef1tub2cmdArpb2rpb1
F. acutatumCBS 402.97 TMW402125MW402323MW402459MW402768MW402653
CBS 739.97AF160276MW402348AF158329MN193883MW402696
CBS 137545MN533987MN534062MN534147MN534228MW402587
F. agapanthiCBS 100193MW401959MW402160MW402363MW402727MW402491
NRRL 54463 TKU900630KU900635KU900611KU900625KU900620
NRRL 54464MN193856KU900637KU900613KU900627MW402718
F. ananatumCBS 118516 TLT996091MN534089MW402376LT996137MW402507
CBS 118517MN533988MN534090MN534157MN534229MW402508
CBS 184.29MW402105MW402303MW402445MW402809MW402629
F. andiyaziCBS 119856MN533989MN534081MN534174MN534286MW402523
CBS 119857 TMN193854LT996113MN534175LT996138MW402524
F. annulatumCBS 258.54 TMT010994MT011041MT010908MT010983MT010944
F. anthophilumCBS 108.92MW401965MW402166MW402368MW402783MW402498
CBS 119858MN533990MN534091MN534158MN534232MW402525
CBS 119859MN533991MN534092MN534164MN534233MW402526
CBS 222.76 ETMW402114MW402312MW402451MW402811MW402641
CBS 737.97MN533992MN534093MN534160MN534234MW402695
F. awaxyCBS 119831MN534056MN534108MN534167MN534237MW402514
CBS 119832MN534057MN534106MN534170MN534240MW402515
CBS 139380MN534058MN534107MN534172MN534238MW402597
F. bactridioidesCBS 100057 TMN533993MN534112MN534173MN534235MW402490
F. begoniaeCBS 403.97MN193858U61543MW402460MN193886MW402654
CBS 452.97 TMN533994MN534101MN534163MN534243MW402675
F. chinhoyienseNY 001B5MN534051MN534083MN534197MN534263MW402725
F. circinatumCBS 405.97 TMN533997MN534097MN534199MN534252MW402656
CBS 119864MW401996MW402196MW402389MW402736MW402528
CBS 141671MW402083MW402282MW402427MW402807MW402610
F. concentricumCBS 450.97 TAF160282MW402334MW402467JF741086MW402674
CBS 453.97MN533998MN534123MN534216MN534264MW402676
CBS 102157MW401963MW402164MW402367MW402728MW402496
F. dlaminiiCBS 175.88MN534002MN534138MN534150MN534256MW402623
CBS 481.94MN534003MN534139MN534151MN534257MW402679
CBS 671.94MN534004MN534136MN534152MN534254MW402690
CBS 672.94MN534005MN534137MN534153MN534255MW402691
CBS 119860 TMW401995MW402195MW402388KU171701KU171681
CBS 119861MN534001MN534135MN534149MN534253MW402527
F. ficicrescensCBS 125177MN534006MN534071MN534176MN534281MW402545
CBS 125181MN534007MN534072MN534177MN534282MW402548
F. fredkrugeriCBS 144209 TLT996097LT996118LT996181LT996147LT996199
CBS 144495LT996096LT996117LT996180LT996146LT996198
F. fujikuroiCBS 186.56MW402108MW402306MW402447MW402765MW402632
CBS 265.54MN534011MN534132MN534222MN534268MW402650
F. globosumCBS 428.97 TKF466417MN534124MN534218KF466406MW402668
CBS 120992MW401998MW402198MW402390MW402788MW402529
F. guttiformeCBS 409.97 TMT010999MT011048MT010901MT010967MT010938
NRRL 22945AF160297U34420AF158350JX171618JX171505
F. konzumCBS 119849 TLT996098MN534095LT996182MW402733MW402519
F. lactisCBS 411.97 ETMN193862MN534077MN534178MN534275MW402659
F. madaenseCBS 146648MW402095MW402294MW402436MW402761MW402616
CBS 146651MW402096MW402295MW402437MW402762MW402617
CBS 146656MW402097MW402296MW402438MW402763MW402618
CBS 146669 TMW402098MW402297MW402439MW402764MW402619
F. mangiferaeCBS 119853MN534016MN534140MN534225MN534270MW402522
CBS 120994 TMN534017MN534128MN534224MN534271MW402530
NRRL 25226AF160281U61561AF158334HM068353MW402712
F. mexicanumNRRL 47473GU737416GU737308GU737389LR792615LR792579
F. mindanaoense
(this study)
PD20-05LC720609LC720611LC720610LC720608LC720612
F. napiformeCBS 748.97 TMN193863MN534085MN534192MN534291MW402701
CBS 135139MN534019MN534084MN534183MN534290MW402572
F. nygamaiCBS 413.97MW402127MW402325MW402462MW402815MW402660
CBS 749.97 TMW402151MW402352MW402479EF470114MW402703
CBS 834.85MW402154MW402355MW402482MW402821MW402707
CBS 119852MW401992MW402192MW402386MW402734MW402521
CBS 139387MW402073MW402272MW402419MW402753MW402601
F. phyllophilumCBS 216.76 TMN193864KF466443KF466333KF466410MW402637
F. pseudonygamaiCBS 416.97MN534030MN534064MN534194MN534283MW402663
CBS 417.97 TAF160263MN534066AF158316MN534285MW402664
CBS 484.94MN534031MN534065MN534195MN534284MW402681
F. ramigenumCBS 418.97 TKF466423MN534145MN534187KF466412MW402665
CBS 526.97MN534032MN534086MN534188MN534292MW402682
F. sacchariCBS 131372MN534033MN534134MN534226MN534293MW402560
NY 001E9MN534034MN534133MN534227MN534294MW402726
F. sterilihyposumNRRL 25623 TMN193869AF160316AF158353MN193897MW402713
F. subglutinansCBS 747.97 NTMW402150MW402351MW402478MW402773MW402700
CBS 136481MW402059MW402258MW402413MW402748MW402585
F. succisaeCBS 219.76 ETAF160291U34419AF158344MW402766MW402639
F. sudanenseCBS 454.97 TMN534037MN534073MN534179MN534278MW402677
CBS 675.94MN534038MN534074MN534182MN534279MW402693
F. temperatumCBS 135538MN534039MN534111MN534168MN534239MW402575
CBS 135539MN534040MN534110MN534169MN534242MW402576
F. thapsinumCBS 539.79MW402140MW402340MW402472MW402818MW402686
CBS 100312MW401961MW402162MW402365MW402780MW402494
F. thapsinumCBS 100313MW401962MW402163MW402366MW402781MW402495
F. tupienseNRRL 53984 TGU737404GU737296GU737377LR792619LR792583
F. udumCBS 747.79MN193872MN534141MN534154MN534258MW402699
F. verticillioidesCBS 125.73MW402012MW402212MW402392MW402791MW402543
CBS 531.95MW402136MW402336MW402468MW402771MW402683
CBS 734.97MW402146MW402346AF158315EF470122MW402694
F. xylarioidesCBS 258.52 TMN193874AY707118MW402455HM068355MW402646
CBS 749.79MN534049MN534143AF158326MN534259MW402702
T Ex-type specimen. ET Ex-epitype specimen. NT Ex-neotype specimen. The sequences deposited to GenBank in this study are shown in bold.
Table 2. Comparison of the size, septation, and shape of sporodochial conidia among related species of FFSC.
Table 2. Comparison of the size, septation, and shape of sporodochial conidia among related species of FFSC.
SpeciesSize (µm)SeptateShapeSubstrate/
Media
References
Fusarium mindanaoense37.7–52.6 × 3.5–5.0
(av. ± sd. 45.9 ± 3.8 × 4.2 ± 0.35; 3-septate)
50.4–66.4 × 3.1–4.6
(av. ± sd. 57.6 ± 3.4 × 3.7 ± 0.39; 4-septate)
55.5–67.8 × 3.2–4.1
(av. ± sd. 62.1 ± 4.3 × 4.1 ± 0.7; 5-septate)
3-5Slightly curvedCLAThis study
F. annulatum13–58 × 1.9–3.33-6Menidiform or annularNot mentionedBugnicourt [30]
F. concentricum53.5–61.4 × 3.7–4 (avg. 57.4 × 3.7)3-5Slightly curvedSNANirenberg and O’Donnell [23]
F. mangiferae43.1–61.4 × 3 1.9–3.4 (avg. 51.8 × 2.3)3-5Slightly curvedCLABritz et al. [31]
F. sacchari35.5–49.5 × 3.3–4.11-5Slightly curvedSNANirenberg [32]
Table 3. Results of the BLASTp analysis using predicted SIX6 of PD20-05.
Table 3. Results of the BLASTp analysis using predicted SIX6 of PD20-05.
Hit_DefintionScoree-ValueQuery_fromQuery_toHit_fromHit_toIdentity
KAF5645217.1 secreted in xylem Fusarium sp. NRRL 25303504.597012461246241
KAG4288609.1 secreted in xylem 6 Fusarium proliferatum503.056012471247240
KAG4277728.1 secreted in xylem 6 Fusarium proliferatum501.13012471247240
KAG4252980.1 secreted in xylem 6 Fusarium proliferatum499.204092471239239
RBA12867.1 secreted in xylem 6 Fusarium proliferatum498.049012471247239
KAF5709672.1 secreted in xylem Fusarium globosum498.049092461238238
KAF4501124.1 secreted in xylem 6 Fusarium agapanthi484.5673.32 × 10−17812461246230
KAF5689079.1 secreted in xylem Fusarium denticulatum476.0936.97 × 10−17512461246224
KAF5626692.1 secreted in xylem 6 Fusarium tjaetaba475.7071.07 × 10−17412461246224
XP_037203386.1 secreted in xylem 6 Fusarium tjaetaba475.7071.07 × 10−17412461246224
KAF5565621.1 secreted in xylem 6 Fusarium napiforme473.7814.55 × 10−17412461246223
KAF5589364.1 secreted in xylem 6 Fusarium pseudocircinatum470.3151.08 × 10−17212461246223
KAF5661873.1 secreted in xylem 6 Fusarium circinatum466.4633.68 × 10−17192461238221
KAF5538596.1 secreted in xylem 6 Fusarium phyllophilum444.1211.90 × 10−16292461238210
KAF5719947.1 secreted in xylem 6 Fusarium mundagurra423.324.06 × 10−15412461242200
KAF5588511.1 secreted in xylem 6 Fusarium pseudoanthophilum403.6752.64 × 10−14612461242203
KAF5973041.1 Secreted in xylem 6 Fusarium coicis348.9773.92 × 10−12512461208180
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Nozawa, S.; Seto, Y.; Takata, Y.; Narreto, L.A.; Valle, R.R.; Okui, K.; Taida, S.; Alvindia, D.G.; Reyes, R.G.; Watanabe, K. Fusarium mindanaoense sp. nov., a New Fusarium Wilt Pathogen of Cavendish Banana from the Philippines Belonging to the F. fujikuroi Species Complex. J. Fungi 2023, 9, 443. https://doi.org/10.3390/jof9040443

AMA Style

Nozawa S, Seto Y, Takata Y, Narreto LA, Valle RR, Okui K, Taida S, Alvindia DG, Reyes RG, Watanabe K. Fusarium mindanaoense sp. nov., a New Fusarium Wilt Pathogen of Cavendish Banana from the Philippines Belonging to the F. fujikuroi Species Complex. Journal of Fungi. 2023; 9(4):443. https://doi.org/10.3390/jof9040443

Chicago/Turabian Style

Nozawa, Shunsuke, Yosuke Seto, Yoshiki Takata, Lalaine Albano Narreto, Reynaldo R. Valle, Keiju Okui, Shigeya Taida, Dionisio G. Alvindia, Renato G. Reyes, and Kyoko Watanabe. 2023. "Fusarium mindanaoense sp. nov., a New Fusarium Wilt Pathogen of Cavendish Banana from the Philippines Belonging to the F. fujikuroi Species Complex" Journal of Fungi 9, no. 4: 443. https://doi.org/10.3390/jof9040443

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop