Open access
Genetics and Molecular Biology
Research Article
12 December 2023

Phylogenetic and functional analyses of N6-methyladenosine RNA methylation factors in the wheat scab fungus Fusarium graminearum

ABSTRACT

In eukaryotes, N6-methyladenosine (m6A) RNA modification plays a crucial role in governing the fate of RNA molecules and has been linked to various developmental processes. However, the phyletic distribution and functions of genetic factors responsible for m6A modification remain largely unexplored in fungi. To get insights into the evolution of m6A machineries, we reconstructed global phylogenies of potential m6A writers, readers, and erasers in fungi. Substantial copy number variations were observed, ranging from up to five m6A writers in early-diverging fungi to a single copy in the subphylum Pezizomycotina, which primarily comprises filamentous fungi. To characterize m6A factors in a phytopathogenic fungus Fusarium graminearum, we generated knockout mutants lacking potential m6A factors including the sole m6A writer MTA1. However, the resulting knockouts did not exhibit any noticeable phenotypic changes during vegetative and sexual growth stages. As obtaining a homozygous knockout lacking MTA1 was likely hindered by its essential role, we generated MTA1-overexpressing strains (MTA1-OE). The MTA1-OE5 strain showed delayed conidial germination and reduced hyphal branching, suggesting its involvement during vegetative growth. Consistent with these findings, the expression levels of MTA1 and a potential m6A reader YTH1 were dramatically induced in germinating conidia, followed by the expression of potential m6A erasers at later vegetative stages. Several genes including transcription factors, transporters, and various enzymes were found to be significantly upregulated and downregulated in the MTA1-OE5 strain. Overall, our study highlights the functional importance of the m6A methylation during conidial germination in F. graminearum and provides a foundation for future investigations into m6A modification sites in filamentous fungi.

IMPORTANCE

N6-methyladenosine (m6A) RNA methylation is a reversible posttranscriptional modification that regulates RNA function and plays a crucial role in diverse developmental processes. This study addresses the knowledge gap regarding phyletic distribution and functions of m6A factors in fungi. The identification of copy number variations among fungal groups enriches our knowledge regarding the evolution of m6A machinery in fungi. Functional characterization of m6A factors in a phytopathogenic filamentous fungus Fusarium graminearum provides insights into the essential role of the m6A writer MTA1 in conidial germination and hyphal branching. The observed effects of overexpressing MTA1 on fungal growth and gene expression patterns of m6A factors throughout the life cycle of F. graminearum further underscore the importance of m6A modification in conidial germination. Overall, this study significantly advances our understanding of m6A modification in fungi, paving the way for future research into its roles in filamentous growth and potential applications in disease control.

INTRODUCTION

Eukaryotic RNA undergoes over 100 chemical modifications that can impact RNA processing and metabolism (13). These modifications include mRNA capping, mRNA polyadenylation, RNA splicing, and RNA methylation (4). Among various RNA methylation, N6-methyladenosine (m6A) is characterized by the methylation of the sixth nitrogen atom on adenosine within RNA. m6A was first described in mammalian cells 50 years ago and is the most prevalent and abundant modification found in eukaryotic mRNA (5, 6). This modification is reversible and regulated by a set of enzymes that function as writers (methyltransferases), erasers (demethylases), and readers (RNA-binding proteins that recognize m6A) (7). Due to its reversible nature, m6A modification serves as a rapid response to environmental stress and regulates various processes including development, immune reactions, and cancer progression in animals, fungi and plants (815).
m6A RNA modification is estimated to occur in approximately 0.1%–0.4% of adenosine nucleotides found in mammalian mRNAs (16, 17). In mammals, the methyltransferase complex responsible for m6A methylation includes proteins, such as methyltransferase-like protein 3 (METTL3), methyltransferase-like protein 14 (METTL 14), and the pre-mRNA-splicing regulator Wilms tumor 1-associated protein (WTAP) (1820). METTL3 serves as the catalytic subunit, forming a heterodimer with its paralogue, METTL14. WTAP is a regulatory subunit whose function is to recruit the m6A methyltransferase complex to the target mRNA in nuclear speckles and is believed to act as a bridge between the METTL3/METTL14 heterodimer and accessory proteins (21). Among the accessory proteins, virilizer-like methyltransferase-associated protein (VIRMA) is implicated in stabilizing the methyltransferase complex and plays a role in the selection of specific sites for the m6A modification (22). YT521-B homology (YTH)-domain proteins are known as m6A readers located in the cytoplasm that influence on translation of methylated mRNAs and their subsequent degradation (23, 24). In humans, the YTH domain family 2 protein (YTHDF2) is involved in controlling mRNA stability by specifically binding to m6A-modified mRNAs and relocating them from ribosomes to processing bodies (25). Demethylation of m6A is catalyzed by 2-oxoglutarate and iron-dependent [2OG-Fe(II)] dioxygenases AlkB-like domain-containing proteins and the fat mass and obesity-associated protein in mammals (26, 27). These diverse m6A factors collectively contribute to m6A RNA modification, playing a significant role in RNA metabolism and gene regulation.
In the budding yeast Saccharomyces cerevisiae, the m6A writer IME4 (Inducer of meiosis 4) is homologous to human METTL3 that catalyzes m6A modification of specific target RNAs involved in the initiation of sexual development and sporulation (28). IME4 is required for proper entry into sexual development and progression through meiotic divisions in diploid cells and is also known to regulate triacylglycerol metabolism, vacuolar morphology, and mitochondrial morphology in haploid cells (29, 30). A yeast two-hybrid screening identified core components of the methyltransferase complex composed of two catalytic factors IME4 and KAR4 (Karyogamy 4, orthologous to human METTL14), as well as two non-catalytic factors MUM2 (Muddled meiosis 2, orthologous to human WTAP), and SLZ1 (Sporulation leucine zipper 1 orthologous to human ZCH13H3) in the budding yeast (31). Recent work revealed that an interacting protein with KAR4, Ygl036wp (thereafter, coined VIR1), in the budding yeast shares a folding pattern similar to VIRMA, despite the lack of discernable protein domains (32, 33). In the budding yeast, SLZ1 enables the methyltransferase complex comprising IME4, KAR4, MUM2 (WTAP), and VIR1 to function in m6A deposition (33).
The genome of the budding yeast encodes a single YTH domain-containing protein, PHO92 (Phosphate metabolism 92), that has been originally described to be involved in phosphate metabolism and response (34, 35). Recently, it became clear that PHO92 recognizes m6A-modified transcripts, facilitates protein synthesis and subsequent decay of m6A-modified transcripts, and promotes meiotic recombination (36, 37). In the fission yeast Schizosaccharomyces pombe, a YTH domain-containing protein, Mmi1 (Meiotic mRNA interception 1), was involved in the selective elimination of meiosis-specific transcripts via posttranscriptional gene silencing (38). However, Mmi1 cannot bind to the m6A consensus motif, suggesting that the function of YTH domain-containing proteins is not limited to m6A recognition and is implicated in diverse cellular functions (39). Although m6A writers and m6A readers have been studied in yeasts, to our best knowledge, m6A erasers await their discovery in fungi.
Fusarium graminearum is a filamentous phytopathogenic fungus and causes devastating diseases in our staple crops, such as wheat, barley, and corn (40). The fungus has served as an excellent model organism for investigating various biological aspects, including host-pathogen interactions, sexual development, mycotoxin production, and RNA editing (4145). Recently, in the rice blast fungus Magnaporthe oryzae, MT-A70 domain protein 1 (MTA1, orthologous to human METTL4) was shown to play an important role in appressorium formation during infection process via regulation of autophagy process (46, 47). However, the roles of the m6A factors are poorly understood in F. graminearum and other filamentous fungi. Thus, the aims of this study were to (i) investigate phyletic distribution and copy number variations of m6A factors in the kingdom Fungi, (ii) functionally characterize putative m6A factors in F. graminearum, and (iii) identify potential targets of the m6A writer MTA1, which is the sole m6A writer found in the Pezizomycotina.

RESULTS

Divergence of m6A writers in the kingdom Fungi

To resolve phylogenetic relationships of genes encoding an MT-A70 protein domain (Pfam domain: PF05063), we collected 1,568 protein sequences from the UniProt database (accessed on 21 May 2023; https://www.uniprot.org/) and curated the list of MT-A70-containing proteins, excluding redundant entries or sequences with an E-value of smaller than 10−5 for PF05063 (see Section Materials and Methods). Following the manual curation, a maximum likelihood tree was reconstructed for 1,116 protein sequences containing an MT-A70 domain in 829 diverse fungal species. This tree revealed the presence of three distinct clades, each including the three previously characterized m6A writers in fungi: IME4, KAR4, and MTA1 (Fig. 1A). Within the IME4 and KAR4 clades, there was a branch exhibiting an early divergence, which exclusively consists of sequences derived from early-diverging fungi, such as species belonging to the phyla Chytridiomycota, Mucoromycota, and Zoopagomycota (see clades labeled with red asterisks in Fig. 1A). Notably, distinct patterns were observed in the phyletic distribution of IME4, KAR4, and MTA1 among different taxonomic groups in fungi (Fig. 1B). Some early-diverging fungi and species within the Pucciniomycotina subphylum were found to possess all the three MT-A70-containing proteins, IME4, KAR4, and MTA1, although many are lacking one of these proteins. The Basidiomycota species (mushroom-forming fungi) and Saccharomycotina species (budding yeasts) typically had IME4 and KAR4, whereas nearly all of the Pezizomycotina species that are comprised of entirely filamentous fungi were found to possess MTA1 as a sole MT-A70-containing protein.
Fig 1
Fig 1 Phylogeny of potential m6A writers in fungi. (A) The maximum-likelihood phylogeny was estimated from 1,116 sequences of MT-A70 domain-containing proteins identified in 829 fungal species. Different phyla were shaded in blue for early-diverging fungi, red for Basidiomycota, and green for Ascomycota. Arrows indicate clades including previously characterized IME4, KAR4, and MTA1 in Ascomycota. Branches marked with red asterisks suggest clades that likely diverged at earlier time points (IME4-like and KAR4-like). The branch lengths in the tree reflect the amount of evolutionary change, with the scale indicating 1.0 amino acid sequence substitution per site. (B) Phyletic distribution of MT-A70 domain-containing proteins in fungi. The bars indicate the standard deviation of gene counts. (C) The phylogeny was estimated from 308 protein sequences that belong to the IME4 subclade and from five METTL3 orthologs found in metazoans and a model plant Arabidopsis thaliana. (D) The phylogeny was estimated from 298 protein sequences that belong to the KAR4 subclade and from five METTL14 orthologs found in metazoans and a model plant Arabidopsis thaliana. (E) The phylogeny was estimated from 510 protein sequences that belong to the MTA1 subclade and from five METTL4 orthologs found in metazoans and a model plant Arabidopsis thaliana. The inner color strip represents the fungal class within the subphylum Pezizomycotina, with each color indicating a distinct taxonomic group as shown in the inset box. (C–E) The color strip outside the tree represents different fungal taxa, with each color indicating a distinct taxonomic group as shown in Fig. 1B. The arrow indicates METTL homologs found in Arabidopsis thaliana, and the scale indicates 1.0 amino acid sequence substitution per site. More detailed trees showing UniProt protein IDs and bootstrap values are shown in Fig. S1.
Independently reconstructed maximum likelihood trees were generated for the IME4, KAR4, and MTA1 clades including their respective homologs of METTL3, METTL14, and METTL4, which were found in humans (Homo sapiens), mouse (Mus musculus), zebrafish (Danio rerio), fruit fly (Drosophila melanogaster), and a model plant (Arabidopsis thaliana) serving as an outgroup (Fig. 1C through E; see Fig. S1 for bootstrap values and protein IDs). The phylogenetic patterns of the IME4 and KAR4 clades exhibited resemblance, encompassing early-diverging fungi and species from Basidiomycota and Saccharomycotina. Among species possessing either IME4 (294 species) or KAR4 (286 species), three-quarters of the species (221 species) had both m6A writers (Table S1). As previously mentioned, some early-diverging fungi exhibited the presence of two copies of IME4 and KAR4. One of these copies appears to have undergone divergence before the emergence of Dikarya (commonly known as “higher fungi”), forming distinct clades here dubbed IME4-like and KAR4-like with robust 100% bootstrap support (Fig. S1). The phylogeny of MTA1 exhibited monophyly and was largely congruent to the phylogeny of Pezizomycotina (Fig. 1E).
Species belonging to Pezizomycotina possessed MTA1 as a sole MT-A70-containing protein. On the other hand, many early-diverging fungi and species from Pucciniomycotina exhibited the presence of three m6A writers, IME4, KAR4, and MTA1 (Fig. 2A). We took a closer look at 26 species that featured the three m6A writers. The increased abundance of MT-A70 domain-containing proteins in early-diverging fungi can be attributable to the existence of IME4-like and KAR4-like genes that diverged from MTA1 at earlier points in time. Although the distribution patterns of MT-A70 domain-containing proteins in early-diverging fungi that belong to the class Chytridiomycetes and Zoopagomycetes are highly variable, they all contained IME4 (or IME4-like), KAR4 (or KAR4-like), and MTA1 (Fig. 2B). Species that belong to the Mucoromycota (except for species from the Endogonales) all lacked IME4 and KAR4, instead having IME4-like and KAR4-like, suggesting that some lineages in the Mucorales have lost IME4 and KAR4 that can be found in species in the Glomeromycetes (the fungal Class sister to the Mucoromycetes). Interestingly, some species in the Mucorales and Mortierellales have two copies of KAR4-like, suggesting gene duplication events. Species that belong to the Pucciniomycotina, the earliest-diverging subphylum within Basidiomycota, have IME4, KAR4, and MTA1 but are devoid of IME4-like and KAR4-like genes (Fig. 2B), suggesting that these early-diverged MT-A70 domain-containing genes were completely lost in Dikarya.
Fig 2
Fig 2 Phyletic distribution of m6A writers in fungi. (A) Copy number variation of potential m6A writers in fungi. The color strip represents different fungal taxa, with each color indicating a distinct taxonomic group as shown in Fig. 1B. (B) Phyletic distribution of IME4, KAR4, and MTA1 in early-diverging fungal phyla. Heatmaps showing the presence (red for one copy, and dark red for two copies) and absence (white) of genes encoding the MT-A70 domain. IME4L: IME4-like gene, and KAR4L: KAR4-like gene.
Recent studies revealed the components of yeast m6A methyltransferase complexes were highly conserved with those found in mammals, insects, and plants, including IME4, KAR4, MUM2 (a homolog of WTAP), and VIR1 (32, 33). Thus, we conducted a search in the UniProt database for relevant Pfam domains, specifically PF17098 for the WTAP/Mum2p family and PF15912 for the N-terminal domain of virilizer (hereafter coined VIRN). We identified 172 and 53 fungal proteins that harbor PF17098 for WTAP and PF15912 for VIRN, respectively, in the current UniProt database (accessed on 4 June 2023; Table S1). After manual curation, we reconstructed phylogenetic trees for WTAP and VIRN, together with homologs found in humans, mice, zebrafish, fruit flies, and Arabidopsis thaliana as an outgroup. WTAP orthologs including Mum2 in yeasts were found only in early-diverging fungi and species from Basidiomycota, and Saccharomycotina, lacking in Pezizomycotina species (Fig. S1). Only 37 fungal species that belong to early-diverging fungi and Basidiomycota were shown to have genes encoding a VIRN domain (Fig. S1). The paucity of homologous sequences in fungi may be attributable to the fact that WTAP and VIRN sequences have diverged significantly from the known sequences, making their identification challenging, especially in Ascomycetous fungi, such as Saccharomycotina and Pezizomycotina. Indeed, a homolog of the virilizer protein, VIR1, was recently found in the budding yeast, which lacks the apparent Pfam domain for VIRN but it exhibits a 3D structure similar to the human homolog VIRMA (32).

Identification of putative m6A reader and eraser in fungi

Among RNA-binding proteins specifically recognizing and binding to m6A, the best studied readers are YTH domain-containing proteins. PHO92 was found as the only protein containing the YTH domain in the budding yeast, while animals can have up to five such proteins and plants can have more than 10. A recent study revealed target transcripts of PHO92 and its important roles during meiosis in budding yeast (37). However, little is known about m6A readers in other fungal taxa. Therefore, we searched for proteins containing a YTH domain (PF04146) in the UniProt database (accessed on 21 May 2023) and found 2,029 fungal sequences. To shorten the list of potential m6A readers, entries with irrelevant protein names (e.g., DNA repair protein RAD51) were excluded, and protein sequences below 300 or above 1,000 amino acids in length were filtered out. After manual curation, we reconstructed a phylogenetic tree for the final list of 868 YTH domain-containing proteins in 695 diverse fungal species (Fig. 3A; see Fig. S1 for bootstrap values and protein IDs). The tree was rooted to YTH domain-containing proteins outside of fungi, including human YTHDF2. The tree was composed of three main clades that were strongly supported by high bootstrap values. One clade contained PHO92 homologs from species belonging to Saccharomycotina and Basidiomycota, as well as YTH domain-containing proteins found in mammals, an insect, and a plant. Another clade exclusively comprised species from Pezizomycotina and was placed sister to the PHO92 clade. The third clade, here dubbed YTH1, consisted of early-diverging fungi and species from Basidiomycota and Ascomycota, displaying a topology that largely matched the species phylogeny. Approximately half of the Agaricomycotina species (67 out of 128) and one-fifth of the Pezizomycotina species (98 out of 498) examined in this study were found to possess both PHO92 and YTH1 while early-diverging fungi and species belonging to Pucciniomycotina, Ustilaginomycotina, and Wallemiomycotina, were found to have only PHO92 (Fig. 3B and C). While the majority of species belonging to Saccharomycotina (30 out of 37) were found to possess PHO92, there were several species including Yarrowia lipolytica (a lipid-producing yeast) that lacked PHO92 and instead exhibited the presence of YTH1 (Fig. 3C; Table S1).
Fig 3
Fig 3 Phylogeny of potential m6A readers in fungi. (A) The maximum-likelihood phylogeny was estimated from 868 sequences of YTH domain-containing proteins identified in 695 fungal species. Red arrows indicate highly supported clades representing PHO92 (Agaricomycotina and Saccharomycotina), PHO92 (Pezizomycotina), and YTH1 families. The branch lengths in the tree reflect the amount of evolutionary change, with the scale indicating 1.0 amino acid sequence substitution per site. The outer color strip represents subphyla, with each color indicating a distinct taxonomic group as shown in (B), and the inner color strip represents fungal class within the subphylum Pezizomycotina, with each color indicating a distinct taxonomic group as shown in the inset box. (B) Phyletic distribution of YTH domain-containing proteins in fungi. (C) Copy number variation of potential m6A readers in fungi. The color strip represents different fungal taxa, with each color indicating a distinct taxonomic group as shown in Fig. 2B.
2OG-Fe[II] dioxygenase AlkB-like domain-containing proteins were known to play roles as m6A erasers by removing the methyl group from m6A in mammals and plants (26, 4850). However, there has been no information on AlkB homologs (ALKBHs) in fungi. Thus, we searched for protein sequences containing a 2OG-Fe(II) dioxygenase superfamily domain in the genome of F. graminearum strain PH-1. We identified five genes harboring a 2OG-Fe(II) dioxygenase domain (PF13532) with an E-value of smaller than 10−10 (gene ID: FGRRES_01255, FGRRES_09872, FGRRES_16456, FGRRES_16652, and FGRRES_20373) (Table 1). 2OG-Fe(II) dioxygenases are involved in diverse cellular processes, such as demethylation of DNA/RNA and modification of histone proteins, as well as secondary metabolite production. To get insights into evolutionary relationships of different types of ALKBH, we reconstructed a phylogenetic tree of ALKBH proteins found in 27 selected fungi, including several functionally characterized ALKBH proteins in humans and Arabidopsis (Fig. 4). Early-diverging fungi, such as Basidiobolus meristosporus and Spizellomyces punctatus, possessed 10 and 9 2OG-Fe(II) dioxygenases, whereas yeasts that belong to Saccharomycotina tend to have only one or two (Table S2). ALKBH2 (FGRRES_09872), ALKBH3 (FGRRES_16456), and ALKBH4 (FGRRES_01255 were closely related to human ALKBH3 and Arabidopsis ALKBH2, which are responsible for repairing DNA damage caused by alkylation (51, 52). ALKBH5 (FGRRES_20373) was placed sister to the clade containing human ALKBH6, which plays a role in DNA repair (53, 54). ALKBH1 (FGRRES_16652) formed a well-supported clade with human and Arabidopsis ALKBH1s, known for their versatile functions in removing methyl groups from DNA/RNA, as well as modifying histone H2A (5557). It was shown that human ALKBH5 and Arabidopsis ALKBH9B and ALKBH10B demethylate m6A (48, 5860). These m6A erasers in mammals and plants formed a clade ancestral to clades including ALKBH1 and ALKBH5 in F. graminearum (Fig. 4).
TABLE 1
TABLE 1 Putative m6A factors found in F. graminearum
NameID (FGRRES)AAPfamE-valuePutative function
MTA106225333PF050631.2e-34writer
WTAPa01626266PF170980.069writer
VIRNa06249143PF159120.073writer
YTH101159613PF041469.9e-66reader
ALKBH116652346PF135323.5e-30eraser
ALKBH209872933PF135321e-29eraser
ALKBH316456443PF135325.6e-35eraser
ALKBH401255326PF135322.8e-28eraser
ALKBH520373228PF135326.2e-15eraser
a
E-value is greater than 10−5.
Fig 4
Fig 4 Phylogeny of potential m6A erasers in fungi. The maximum-likelihood phylogeny was estimated from 125 sequences of 2OG-Fe(II) dioxygenase AlkB-like domain-containing proteins identified in 27 selected fungal species. The color strips on the right side of leaves indicate different fungal taxa: purple—Pezizomycotina, yellow—Saccharomycotina, blown—Basidiomycota, cyan—early-diverging fungi, white—human or Arabidopsis thaliana. Highly supported clades each containing an ALKBH identified in F. graminearum were shaded with different colors. Bootstrap values of greater than 70% were shown. Branch lengths are proportional to the inferred amount of evolutionary change, and the scale represents 1.0 amino acid sequence substitutions per site.

Functional characterization of putative m6A factors in F. graminearum

To examine the functions of m6A factors in F. graminearum during sexual development, we conducted a search for Pfam domains associated with m6A writer, reader, and eraser, and discovered several potential m6A factors (Table 1). Among them, F. graminearum possessed genes homologous to MTA1 and YTH1, along with the previously mentioned five ALKBHs. In addition, we identified two genes that encoded hypothetical proteins harboring PF17098 for WTAP and PF15912 for VIRN, respectively, with E-values of greater than 10−5 (Table 1). Despite the low sequence similarity to WTAP and VIRN found in fungi, we included these potential m6A writers in the knockout study because their expression levels were observed to increase during sexual development (see Fig. 7 in the section “Roles of MTA1 in conidial germination and hyphal growth”). We individually deleted m6A factors in F. graminearum strain PH-1 [wildtype (WT)]. The authenticity of the resulting knockout mutants was verified by diagnostic PCRs (Fig. 5A). The knockouts displayed normal growth and produced perithecia, which are characterized by their dark purple coloration and flask-shaped or spherical structures (Fig. 5B). Morphologies of ascospores (the products of meiosis) enveloped in sac-like structures called asci were also normal in the knockouts, indicating that m6A factors did not affect sexual development in F. graminearum (Fig. S2).
Fig 5
Fig 5 Generation of knockout mutants lacking potential m6A factors in F. graminearum. (A) Diagnostic PCRs confirmed homologous integration of the split marker constructs to the target gene loci. Note the PCR band size between the WT and knockout strains (KO) due to the gene replacement. M—1 kb DNA ladder. (B) Perithecia production of the WT and knockout strains grown on carrot agar media. Photographs were taken with a dissecting microscope 6 days after induction of sexual development (20× magnification). (C) RT-PCR analyses of the WT and Δmta1 strains. Since MTA1 lacks introns, distinguishing RNA expression from possible genomic DNA contamination in RT-PCR analysis was challenging. To verify RNA integrity, a primer set was designed for amplifying EF1α, including two introns. Notably, no band corresponding to genomic DNA amplicon for the EF1α reference gene was observed (658 bp for gDNA, 356 bp for mRNA), indicating the absence of genomic DNA contamination. Expression of MTA1 was confirmed in the WT, whereas no discernible band was observed in the Δmta1 strain. The samples labeled V were collected before sexual induction (i.e., vegetative growth stage), and samples labeled S were collected 6 days after sexual induction (i.e., sexual growth stage). M—100 bp DNA ladder.
Since m6A writer, IME4, in the budding yeast, plays a crucial role in initiating sexual development, we carefully examined the genotype of Δmta1. Although genotyping confirmed the deletion of MTA1 in Δmta1, we were intrigued to find that a PCR band specific to MTA1 could still be amplified from Δmta1. This suggested that Δmta1 is heterokaryotic, indicating that it still contains one or more “wild-type” nuclei. To obtain homokaryons, we performed additional single-spore isolations, using Δmta1, and examined 10 single-spored isolates derived from Δmta1. Diagnostic PCR analysis confirmed the deletion of MTA1 in all 10 isolates; yet, a fragment of MTA1 was still amplified from these isolates (Fig. S3). The expression levels of MTA1 were markedly reduced in Δmta1, suggesting a lower abundance of wild-type nuclei compared to the WT (Fig. 5C). This observation was further supported by RNA-seq analysis. When examining the mapped reads on the MTA1 locus in both WT and Δmta1, we found a decrease in the number of reads mapped to the MTA1 locus in Δmta1 compared to the WT (Fig. S3). The presence of RNA-Seq reads mapping to the MTA1 locus, which should have been absent in homokaryotic Δmta1, served as additional proof for the heterokaryotic nature of Δmta1 and indicated that Δmta1 is, in fact, a knockdown strain. Consequently, we will refer to Δmta1 as the MTA1-KD1 strain hereafter.

MTA1 as an m6A writer in F. graminearum

Extensive efforts were unsuccessful in obtaining knockouts completely lacking MTA1 (Fig. S3), presumably due to its indispensable role in F. graminearum. To get better insight into the roles of m6A RNA methylation in F. graminearum, we generated strains overexpressing MTA1. RT-PCR analysis confirmed that two strains, MTA1-OE4 and MTA1-OE5, exhibited overexpression of MTA1 (Fig. 6A). It was notable that MTA1 was overexpressed to a greater extent in the MTA1-OE5 strain, compared to MTA1-OE4. Both semi-quantitative RT-PCR and RNA-seq analyses indicated that the expression level of the MTA1-OE5 strain was approximately 100 times greater than that of the WT (Fig. S4). The WT and MTA1-OE4 strains produced normal asci containing mature ascospores 6 days after the induction of sexual development, while it was not until 11 days after the induction that the MTA1-OE5 strain produced fully developed asci (Fig. 6B). The difference in the observed phenotypes in the two MTA1-overexpressing strains could likely be attributable to the variation in the degrees of MTA1 expression, and thus the MTA1-OE5 strain which displayed substantial overexpression of MTA1 was used for further analysis and investigation. To determine whether MTA1 is required for m6A RNA methylation in F. graminearum, we compared total m6A RNA methylation levels between the WT and MTA1-OE5 strains. The amount of m6A RNA in WT was 0.075% ± 0.015% of the total RNA, which was approximately half of that in the MTA1-OE5 strain (0.152% ± 0.033% of the total RNA), which indicated that the m6A RNA methylation level was significantly increased in the MTA1-OE5 strain compared to the WT (Fig. 6C).
Fig 6
Fig 6 Overexpression of the m6A writer MTA1. (A) RT-PCR analyses of the WT and MTA1-overexpressing strains (MTA1-OE4 and MTA1-OE5). Since MTA1 lacks introns, distinguishing RNA expression from possible genomic DNA contamination in RT-PCR analysis was challenging. To verify RNA integrity, a primer set was designed for amplifying EF1α, including two introns. Notably, no band corresponding to genomic DNA amplicon for the EF1α reference gene was observed (658 bp for gDNA, 356 bp for mRNA), indicating the absence of genomic DNA contamination. Overexpression of MTA1 was confirmed in the MTA1-OE4 and MTA1-OE5 strains. Note that MTA1 was significantly overexpressed in the MTA1-OE5 strain, compared to the WT, from which no discernible band was observed at the PCR cycle of 30. M—100 bp DNA ladder. (B) m6A methylation level in RNA extracted from the WT and MTA1-OE5 strains. Mann-Whitney U test was performed to compare the means of the ratio for m6A to A between the WT and MTA1-OE5 strains. Box and whisker plots indicate the median, interquartile range between the 25th and 75th percentiles (box), and 1.5 interquartile range (whisker). (C) Perithecia production of the WT, MTA1-OE4, and MTA1-OE5 strains grown on carrot agar media (upper panels, 20× magnification). Photographs were taken with a dissecting microscope at the indicated days after induction of sexual development (DAI). Squash mounts of perithecia were observed with a compound microscope (bar = 50 µm). Mature ascospores were observed in the WT and MTA1-OE4 strains as early as 7 DAI, whereas in the MTA1-OE5 strain, it was not until 11 DAI that ascospore formation became evident.

Roles of MTA1 in conidial germination and hyphal growth

In previous studies, we obtained transcriptome data for conidial germination on the Bird agar medium (61), as well as for perithecial development on the Carrot agar medium (62). By integrating these data sets with the RNA-seq data obtained in the present study (freshly harvested conidia), we conducted a comprehensive analysis of expression level changes in m6A factors throughout the life cycle of F. graminearum, encompassing various stages of vegetative growth and perithecial development (Fig. 7). The m6A writer MTA1 and a potential m6A reader YTH1 demonstrated a synchronized expression pattern, as indicated by a high Pearson correlation coefficient of 0.94. Their expression levels exhibited a significant increase during stage 1 (S1, 15 minutes after incubation of conidia on Bird agar medium), followed by a decline during hyphal growth (S2 and S3), and remained at a basal level during perithecial development (S4–S9). Interestingly, the expression levels of ALKBH1 and ALKBH5, the two most probable m6A erasers in F. graminearum, reached their peak at stage 2 (S2, 3 h after incubation), at which conidia germinated and hyphae started to extend (Fig. S5). During active hyphal growth at stage 3 (S3, 11 h after incubation), the major components of the fungal cytoskeleton, actin (FGRRES_07735) α-tubulin (FGRRES_00639), and β-tubulin (FGRRES_09530) that play a crucial role in polarity establishment, maintenance, and polar growth, displayed the greatest induction during the life cycle of F. graminearum.
Fig 7
Fig 7 Conidial germination affected by the m6A writer MTA1. (A) Gene expression profiles of potential m6A factors (left panel) and housekeeping genes (right panel) in Fusarium graminearum. Average values for reads per kilobase per million mapped reads (RPKM) values for three replicate samples were plotted. Bands surrounding the line plots indicate 95% confidence intervals of the means. The x-axis are different growth stages of F. graminearum (S0–S9). See the right box for the description of vegetative and sexual growth stages. Gene ID for potential m6A factors are ALKBH1 (FGRRES_16652), MTA1 (FGRRES_06225), VIRN (FGRRES_06249), and WTAP (FGRRES_01626). Housekeeping genes examined here are α-tubulin (αTUB, FGRRES_00639), β-tubulin (β TUB, FGRRES_09530), and actin 1 (ACT1, FGRRES_07735). (B) Conidia germination and polar growth of hyphae in quarter-strength potato dextrose broth (q-PDB) medium. Photos were taken 16 h after incubation. Note that shorter hyphae germinated from macroconidia of the MTA1-OE5 strain, compared to the WT strain. (C) The dry weight of mycelia grown in q-PDB medium was measured at 2 days interval. (D) The numbers of differentially expressed genes in fresh macroconidia harvested from carboxymethylcellulose (CMC) medium between the WT and MTA1-OE5 strains, and between the MTA1-KD1 and MTA1-OE5 strains.
Taking into account the expression profile, we proceeded to investigate the phenotypic characteristics of the MTA1-OE5 strain during conidial germination and hyphal growth. In contrast to the WT, the hyphal tips did not exhibit elongation in an overnight culture of the MTA1-OE5 strain cultivated in quarter-strength potato dextrose broth (q-PDB) medium (Fig. 7B). Furthermore, in the MTA1-OE5 strain growing on potato dextrose agar (PDA) medium, the hyphae at the leading edge displayed a tendency to have fewer branches compared to the WT (Fig. S6). Next, we conducted measurements of the total biomass of the WT and MTA1-OE5 strains cultured in q-PDB medium for a duration of 8 days. The MTA1-OE5 strain exhibited a slower growth rate compared to the WT, particularly during the initial 2 days of cultivation (Fig. 7C). To investigate genes responsible for the delayed conidial germination observed in the MTA1-OE5 strain, we performed differential gene expression analyses between the WT, MTA1-KD1, and MTA1-OE5 strains, using RNA-seq data obtained from freshly harvested conidia (stage 0, S0). In this analysis, we identified a total of 20 differentially expressed (DE) genes [fold change > 4 at a false discovery rate (FDR) of 5%], when comparing the WT and MTA1-OE5 strains, as well as 21 DE genes when comparing the MTA1-KD1 and MTA1-OE5 strains (Fig. 7D). Notably, 16 genes were found to be commonly differentially expressed in both comparisons, including MTA1. There were no DE genes observed between the WT and MTA1-KD1 strains (Fig. S7). Among the DE genes, we found that 17 genes were transcriptionally upregulated, and eight genes were downregulated in the MTA1-OE5 strain (Table 2). The DE genes included nine genes encoding hypothetical proteins without any predicted protein domain and 16 functionally annotated genes encoding two transcription factors, two transporters, and diverse enzymes. FGRRES_00725 and FGRRES_05926 encoding GAL4-type transcription factors were significantly downregulated in the MTA1-OE5 strain, while FGRRES_02139 and FGRRES_04188 encoding an ABC multidrug transporter and major facilitator superfamily, respectively, were highly upregulated in the MTA1-OE5 strain (Table 2). In consistent with the expression pattern of MTA1, the expression levels of two transporters were elevated at S1 and reduced thereafter. For two transcription factors, the expression levels of FGRRES_05926 and FGRRES_00725 began to decline at S2 and S1, respectively, and remained basal, suggesting that their roles in early stages of conidial germination and hyphal growth.
TABLE 2
TABLE 2 Differentially expressed genes in the MTA1-overexpressing strain
Gene ID (FGRRES)FCaFDRbPredicted function
11026−5.20.0212Nonribosomal peptide synthetase, malonichrome
16138−6.00.0212Hypothetical protein
11158−8.10.0212Amidohydrolase family
16091c−4.30.0212Hypothetical protein
171365.20.0212Superfamily I DNA and/or RNA helicase
036008.20.0212Hypothetical protein
00251−5.00.0212Probable galactose oxidase precursor
139795.50.0233Acetyltransferase (GNAT) domain
04188c4.00.0233Major facilitator transporter
114125.50.0233Hypothetical protein
128206.90.0245Hypothetical protein
05926−3.80.0245GALl4-type MHR TF (without a zinc cluster domain)
135614.70.0245FAD-dependent oxidoreductase
10598c6.60.0245Hypothetical protein
021395.70.0245ABC transporter
114136.90.0255CoA-transferase family III domain-containing protein
062257.00.0306MTA1, m6A writer
00725c−4.20.0499GAL4-type MHR TF (with a zinc cluster domain)
117364.60.0499Hypothetical protein
058294.90.0499Prolyl oligopeptidase family
03319d4.60.0597AAA family ATPase
11474d5.30.0901Nucleotidyltransferase superfamily, GrpB domain-containing
15648d5.10.0901Hypothetical protein
08659d2.80.1125Hypothetical protein
10446d−4.20.1479Pyruvate decarboxylase
a
log2-transformed fold change in the MTA1-overexpressing strain.
b
False discovery rate.
c
Differentially expressed genes only in comparison between the wild-type and MTA1-overexpressing strain.
d
Differentially expressed genes only in comparison between the ΔMTA1 and MTA1-overexpressing strain.

DISCUSSION

Roles of m6A writer in filamentous fungi

In the phylum Ascomycota, the roles of m6A writers in budding yeasts (Saccharomycotina) and filamentous fungi (Pezizomycotina) may have evolved independently. Species belonging to the Pezizomycotina possess only a single m6A writer MTA1, a homolog of METTL4, while species belonging to the Saccharomycotina have two m6A writers, IME4 and KAR4 that are homologs of METTL3 and METTL14, respectively. In budding yeast, IME4 and KAR4 are crucial for the initiation of meiosis for ascospore formation (29, 63). However, the m6A writer MTA1 is dispensable for perithecial development and ascospore production in filamentous fungi, F. graminearum and M. oryzae (46). The relative abundance of m6A in M. oryzae (0.069% ± 0.003%) is comparable to that in F. graminearum (0.075 ± 0.015%) (47), suggesting that the degree of m6A levels of total mRNA is similar in these two filamentous fungi. In the rice blast fungus M. oryzae, a knockout strain lacking MTA1 showed a defect in appressorium formation during the infection process on the epidermal cells of rice (47). Expression levels of autophagy-related genes in M. oryzae were changed according to the degree of m6A level, which is important for normal appressorium formation. Although the homozygous Δmta1 strain in M. oryzae was viable, our extensive efforts to obtain homozygous Δmta1 were unsuccessful in F. graminearum, suggesting its essential role. Given the significant impact on conidial germination and hyphal growth observed in the MTA1-overexpressing strain in F. graminearum, the role of MTA1 appears to have diverged in the two plant pathogenic fungi.
Most importantly, F. graminearum does not form appressoria and utilizes distinct strategies to penetrate the epidermal cells of the host plants. Both F. graminearum and M. oryzae belong to the Sordariomycetes but fall into different orders, Hypocreales and Magnaporthales, respectively (64). RNA-seq analyses indicated significant divergence in gene expression including autophagy-related genes during conidial germination and infection processes between F. graminearum and M. oryzae (61). The m6A writers IME4 and KAR4 were crucial for the initiation of the sexual cycle and the formation of ascospores in the budding yeast. However, the m6A writer MTA1 does not appear to be responsible for perithecial development and ascospore production in F. graminearum and M. oryzae. Although the MTA1-OE5 strain exhibited delayed perithecial development, it produced normal ascospores. In the Pezizomycotina, A-to-I RNA editing, one of the eukaryotic RNA modifications, is crucial for sexual development and ascospore formation (4345, 6567).

Evolutionary perspectives of m6A factors in fungi

Several ancestral traits observed in early-diverging fungi are shared with metazoans or unicellular opisthokonts, which have been subjected to extensive parallel loss across the Dikarya lineages (68, 69). As with metazoans, many species of early-diverging fungi and species that belong to the Pucciniomycotina (the earliest-diverging subphylum within Basidiomycota) possess all three m6A writers, IME4, MTA1, and KAR4. These suggested that budding yeasts (Saccharomycotina) may have lost the m6A writer MTA1. Conversely, species belonging to Pezizomycotina, which entirely consists of filamentous fungi, may have lost the m6A writers IME4 and KAR4. The phyletic distribution of m6A writers in fungi suggested that Pezizomycotina species including F. graminearum and M. oryzae have likely embraced m6A machineries, such as MTA1, as a means to regulate filamentous growth and facilitate host penetration. This adoption of m6A machinery may have played a crucial role in the development and maintenance of the filamentous morphology that is characteristic of these fungi, allowing them to effectively colonize and interact with their respective hosts.
In contrast to well-characterized yeast m6A methyltransferase complexes including IME4 and KAR4 m6A writers, our understanding of m6A machinery in filamentous fungi is still limited. Despite our functional characterization of potential WTAP and VIR1 homologs in F. graminearum, it is unlikely that these represent genuine homologs. This is because the homologs of WTAP/MUM2 and VIR1 are missing in species within all the Pezizomycotina species. The failure to identify homologs of WTAP and VIR1 in our study is likely due to significant sequence divergence of these m6A factors within the Pezizomycotina. The sequence differences might have hindered their detection using conventional homology search methods, indicating that these m6A factors in Pezizomycotina may have undergone substantial changes in the primary amino acid sequences. Until recently, the existence and function of VIR1 in the budding yeast had remained unknown, but through the application of protein folding prediction tools, its identity was revealed (32). Recent advancements in protein 3D structure prediction tools to uncover hidden homologs (70, 71) hold promise for identifying homologous members of the m6A methyltransferase complexes in Pezizomycotina species. Alternatively, it is possible that components of m6A machinery associated with MTA1 are completely distinct from those found in the yeast m6A methyltransferase complex associated with IME4 and KAR4. In this case, it will be necessary to identify components that interact with MTA1 via co-immunoprecipitation.
In addition, it is worth mentioning, considering the number of putative m6A writers and m6A erasers found in early-diverging fungi and species from the Pucciniomycotina (up to five m6A writers in a species), not many m6A readers containing an YTH domain were identified, suggesting that there may be different types of RNA-binding proteins that recognize m6A modification. It would be interesting to study the possibly divergent roles of m6A factors in these relatively understudied fungal taxa.

Expression levels of m6A factors throughout the life cycle of F. graminearum

The levels of MTA1 and YTH1 expression were significantly elevated at 15 minutes after placing conidia on the Bird medium (S1) and then decreased sharply to a basal level during polar growth and hyphal branching (S2–S3). These findings suggest that the potential m6A writer and m6A reader likely have a role during the initial stage of conidial germination. In line with this hypothesis, we observed slower conidial germination and less hyphal branching in the MTA1-OE5 strain. Although m6A demethylase activities of ALKBH genes have never been confirmed in fungi, the three putative m6A erasers, ALKBH1, ALKBH4, and ALKBH5, were induced during polar growth (S2) or hyphal branching (S3). These possible m6A erasers might be involved in maintaining low m6A levels during active vegetative growth. The significant increase in expression of actin and tubulin genes at S2–S3 reflected extensive hyphal growth and branching. Specifically, the expression level of ALKBH4 reached its peak at S0 (fresh conidia) but experienced a sharp decline at S1 (the initial stage of conidial germination). However, it was gradually induced during polar growth and hyphal branching stages (S2-S3), exhibiting an opposite expression pattern compared to MTA1 and YTH1. Although we did not observe any phenotypic changes in the knockout mutant lacking ALKBH1, it would be possible that other ALKBH genes may be involved in the regulation of conidial germination and hyphal growth by maintaining low m6A level in F. graminearum. Among the five putative m6A erasers, ALKBH2 showed a dramatic increase in expression during the ascospore formation. These expression dynamics of m6A factors highlight the complex regulation and involvement of m6A factors throughout different stages of the F. graminearum life cycle.

Concluding remarks

The largest subphylum Pezizomycotina in fungi underscores their importance and their complex interactions with humans, encompassing a wide range of fungi that have significant influences on humans, both negative and positive. Examples include the opportunistic human pathogen Aspergillus fumigatus and the dermatophyte Trichophyton rubrum, as well as many plant pathogenic fungi causing serious diseases on our staple crops, such as F. graminearum and M. oryzae. The Pezizomycotina also includes several ecologically significant species, encompassing those that play crucial roles in wood and litter decay processes, as well as those that form symbiotic associations with other organisms, including lichens. In filamentous fungi, the m6A writer MTA1 appears to have evolved to have important roles in adapting to diverse ecological niches, particularly in relation to filamentous growth. To identify target genes of MTA1 that caused delayed conidial germination and slower hyphal growth in the MTA1-OE5 strain, we are currently investigating potential m6A sites in transcripts, using an Oxford Nanopore direct RNA sequencing technology.

MATERIALS AND METHODS

Phylogenetic analyses of m6A factors in fungi

We downloaded protein sequences from the UniProt database (https://www.uniprot.org/) that possessed pfam domains associated with m6A factors. These domains include PF05063 (MT-A70), PF04146 (YTH), PF13532 (ALKB), PF17098 (WTAP/MUM2), and PF15912 (Virilizer, N-terminal). However, the initial lists of potential m6A factors contained duplications and likely misannotated sequences. To address this issue, we conducted a manual inspection of the lists, excluding sequences that were either too short or too long, as well as sequences with small E -values, applying specific thresholds (e.g., E -value > 10–5). Whenever a sequence was removed from the lists, we also eliminated all entries from the corresponding species to ensure an accurate estimation of the number of m6A factors per species. The sequences for each pfam domain were aligned using MAFFT (v7.310) with the “auto” setting (72). Poorly aligned regions of the resulting multiple sequence alignment were trimmed, using the Trimal program, with the parameter setting “–gappyout” (73). To determine the best protein substitution model for each pfam domain, we used a perl scrip that can be found on the following website (https://github.com/stamatak/standard-RAxML/blob/master/usefulScripts/ProteinModelSelection.pl). We selected protein substitution models GAMMALG for m6A writers, GAMMAJTTF for m6A readers, GAMMALGF for ALKBHs and VIRN, and GAMMAJTT for WTAP. Maximum likelihood trees were constructed using the RAxML program (v8.2) (74). The outgroup was set to the homologue of m6A factors in Arabidopsis thaliana and nodal supports were evaluated by 1,000 bootstrap replications.

Genetic transformation for gene deletion and overexpression

To generate gene deletion mutants in the F. graminearum PH-1 strain, we employed a split marker strategy (75). This involved amplifying the left and right flanking regions of the target genes and combining them with a minimal gene cassette that carried the hygromycin phosphotransferase (HPH) gene under the control of the trpC promoter from Aspergillus nidulans. To achieve this, we conducted fusion PCR, following the previously described method (76), and the specific primers used for targeted gene deletion can be found in Table S3. In brief, we amplified the left and right flanking regions of the target genes separately, using L5 and L3 primer pairs and R5 and R3 primer pairs, respectively. The L3 and R5 primers contained 27-nucleotide (nt)-long overhang sequences that were complementary to the 5′ and 3′ ends of the minimal HPH cassette (1,376 bp in length). The HPH cassette was obtained from the pCB1004 plasmid (77), and amplified using HYG-F and HYG-R primers. Subsequently, we merged the PCR amplicons through overlap extension, assembling the left flanking region and the HPH cassette, or the right flanking region and the HPH cassette. The split marker constructs were obtained by amplifying the fused amplicons using nested primer pairs (N5 and HY-R primers for the left-half construct and YG-F and N3 primers for the right-half construct). Finally, we introduced the two split marker constructs into protoplasts through polyethylene glycol-mediated transformation (78). Following transformation, transformants resistant to 200 mg/mL of hygromycin were examined for replacement of the target gene with the HPH cassette by diagnostic PCR checks, in which L5 and R3 primers that aneal to flanking sequences of the homologous recombination event were used to confirm homologous integration of the HPH cassette into the target loci.
For the generation of MTA1-overexpressing strains, the coding sequence of MTA1 was cloned to pDS23 plasmid digested with BglII and HindIII restriction enzymes (79), using the In-Fusion HD Cloning kit (Takara Bio, Otsu, Japan). Five micrograms of the plasmid harboring MTA1 was transformed into protoplasts of the WT strain. Following transformation, transformants resistant to 200 mg/mL of nourseothricin (Jena Bioscience, Jena, Germany) were examined for introduction of an additional copy of MTA1 by diagnostic PCR check using primer pair, MTA1_OE_RT_fwd and MTA1_RT_rev. Primers used in this study are listed in Table S3.

Sexual development and hyphal growth measurement

Carrot agar plates (60 mm in diameter) (80) were inoculated by placing an agar block containing hyphae of F. graminearum at the center. The plates were incubated at room temperature under constant light. Six days after incubation, mycelia were removed by gently scraping the surface with a spatula, and then 0.9 mL of 2.5% Tween 60 (Sigma-Aldrich, St. Louis, MO, USA) was applied to the surface to assist the formation of perithecia. Sexual development in knockout mutants was observed by a stereomicroscopy for size and number formed. Squash mounts of young developing perithecia in water were examined using a compound microscope to check the morphology and maturity of ascospores.
To obtain macroconidia of the WT and MTA1-overexpressing strains, small agar blocks containing each strain were inoculated and cultured in carboxymethylcellulose medium for 3 days at 20°C (81). Macroconidia were harvested by filtration through two layers of Miracloth. Concentrations of macroconidia were adjusted to 1 × 108 spores per mL, and 40 µL of spore suspensions was inoculated into 100 mL of q-PDB. Cultures were shaken at 150 rpm at room temperature. Mycelia were collected by filtration through one layer of Miracloth and oven-dried at 55°C for 2 days.

M6A quantification

The m6A RNA methylation level was assessed using the EpiQuikTM m6A RNA methylation quantitative kit (Epigentek, Farmingdale, NY, USA). Briefly, 200 ng of total RNA was added and bound with the antibody in the strip wells, followed by the process of washing, capturing, and detecting antibody. Finally, the signals were detected colorimetrically by reading the absorbance at 450 nm. m6A levels were eventually calculated based on a constructed standard curve.

RT-PCR and quantitative RT-PCR analyses

Total RNA was extracted from fresh conidia, hyphae, and perithecial tissues ground in liquid nitrogen using TRIzol reagent (Thermo Fisher Scientific) according to the manufacturer’s instructions together with the following extraction steps: two phenol (pH 4.6)-chloroform-isoamyl alcohol (25:24:1) extraction steps followed by two chloroform extraction steps after the initial TRIzol-chloroform phase separation. RNA pellets were dissolved in 88 µL of nuclease-free water and subjected to genomic DNA digestion with DNase (Qiagen Inc.). RNA samples were then concentrated using RNA Clean & Concentrator (Zymo Research). For reverse transcription (RT)-PCR analysis of MTA1, 200 ng of RNAs was reverse transcribed and amplified using the OneStep RT-PCR kit (Qiagen Inc.). MTA1 and EF1α (FGRRES_08811) were amplified for 34 cycles in the experiment for Fig. 4C, or 30 cycles in the experiment for Fig. 5A, respectively. The annealing temperature was 62°C. The primer set for amplifying EF1α was designed to amplify flanking exons, including two introns to detect possible genomic DNA contamination. For semi-quantitative RT-PCR assays, first-strand cDNA synthesis was prepared from 200 ng of total RNA using the iScript cDNA Synthesis Kit (Bio-Rad), according to the manufacturer’s instructions. Real-time RT-PCR analyses were performed using the CFX96 Touch Real-Time PCR Detection System (Bio-Rad). RT-PCR mixtures were composed of 1.5 µL of each primer (10 µM), 5 µL of SYBR Green Supermix (Bio-Rad), and 2 µL of cDNA (100 ng/mL). The PCR conditions consisted of an initial denaturing step at 95°C for 3 minutes, a denaturation step at 95°C for 10 s, both annealing and extension steps at 65°C for 30 s for 40 cycles, and 65 to 95°C with a 0.5°C increment and each temperature for 5 s to obtain the melting curve. The quantification of the relative expression levels was performed with the comparative cycle threshold method normalization (82), in that the expression level of MTA1 was normalized against the reference gene EF1α. The averages of the three biological replicates and standard deviations of the relative expression values were presented (Fig. S4). Primers used in the expression analysis are listed in Table S3.

RNA-seq and differential expression analyses

Total RNA samples were extracted from fresh conidia of the WT, MTA1-KD1, and MTA1-OE5 strains that have been harvested from carboxymethylcellulose medium 4 days after incubation at 20°C (81). Three separate experiments were performed to produce conidia, and the conidia samples from each experiment were used as replicates for RNA-Seq analysis. Two micrograms of total RNA was sent to Macrogen Inc. (Seoul, Korea) for cDNA library construction, using the TruSeq Stranded Total RNA Library Prep Gold Kit (Illumina, San Diego, CA, USA), and for sequencing on the HiSeq4000 platform (Illumina). Raw reads (paired-end, 100 bp) were further processed and filtered, using the TrimGalore (v0.6.6) (https://www.bioinformatics.babraham.ac.uk/projects/trim_galore/). Filtered reads were mapped to the genome sequence of F. graminearum (NCBI accession: GCA_000240135.3), using the HISAT2 program (v2.1.0). The gene annotation file used in this study was Ensembl annotation v.32 (83). Mapped reads on genomic features, such as exon and intron, were calculated, using the htseq-count program. Gene expression levels in reads per kilobase per million mapped reads (RPKM) values were computed and normalized by effective library size estimated by trimmed mean of M values, using the edgeR R package (v3.26.8). For differential expression analysis, only genes with CPM values greater than 1 in at least three samples were kept for further analyses (9,792 out of 16,001 gene loci). Then, DE genes showing greater than a fourfold difference at an FDR of 5% were identified between the WT, MTA1-KD1, and MTA1-OE5 strains, using the limma R package (v3.28.21).

DATA accessions

The RNA-seq data generated in the present work have been deposited in NCBI’s Sequence Read Archive and are accessible through SRA accessions from SRX21157089 to SRX21157097, which belong to the BioProject (accession, PRJNA998561). RNA-seq data for investigating the expression level changes in m6A factors throughout the life cycle of F. graminearum can be found in NCBI’s Gene Expression Omnibus GSE109088 for the conidial germination stages (S1–S3) and GSE109094 for the sexual development (S4–S9).

ACKNOWLEDGMENTS

This publication is based upon work supported by the Basic Science Research Program through the National Research Foundation of Korea, funded by the Ministry of Education (2019R1I1A1A01057502 and 2022R1C1C2004118). W.K was partly supported by the BK21 FOUR of the National Research Foundation of Korea (NRF) funded by the Ministry of Education (No. 4299991014422). The funders had no role in study design, data collection and analysis, decision to publish, or preparation of the manuscript.
H.K. performed functional analyses of m6A factors. J.H. and H.K. conducted m6A quantification. W.K. conducted phylogenetic analyses of m6A factors. H.K. and W.K. analyzed and interpreted the data. H.K. and W.K. wrote the manuscript. All authors read and approved the final manuscript.

SUPPLEMENTAL MATERIAL

Supplemental figures. - msphere.00552-23-s0001.pdf
Figures S2 to S7.
Fig. S1 - msphere.00552-23-s0002.pdf
Maximum likelihood trees of potential m6A factors in fungi.
Table S1 - msphere.00552-23-s0003.xlsx
Metadata for UniProt protein sequences related to MTA1, IME4, KAR4, WTAP, VIRN, YTH1, and ALKBHs.
Table S2 - msphere.00552-23-s0004.docx
Numbers of putative m6A erasers found in 27 selected fungi.
Table S3 - msphere.00552-23-s0005.docx
Primers.
ASM does not own the copyrights to Supplemental Material that may be linked to, or accessed through, an article. The authors have granted ASM a non-exclusive, world-wide license to publish the Supplemental Material files. Please contact the corresponding author directly for reuse.

REFERENCES

1.
Hu J, Cai J, Xu T, Kang H. 2022. Epitranscriptomic mRNA modifications governing plant stress responses: underlying mechanism and potential application. Plant Biotechnol J 20:2245–2257.
2.
Manavella PA, Godoy Herz MA, Kornblihtt AR, Sorenson R, Sieburth LE, Nakaminami K, Seki M, Ding Y, Sun Q, Kang H, Ariel FD, Crespi M, Giudicatti AJ, Cai Q, Jin H, Feng X, Qi Y, Pikaard CS. 2023. Beyond transcription: compelling open questions in plant RNA biology. Plant Cell 35:1626–1653.
3.
Motorin Y, Helm M. 2011. RNA nucleotide methylation. WIREs RNA 2:611–631.
4.
Saletore Y, Meyer K, Korlach J, Vilfan ID, Jaffrey S, Mason CE. 2012. The birth of the epitranscriptome: deciphering the function of RNA modifications. Genome Biol 13:175.
5.
Perry RP, Kelley DE. 1974. Existence of methylated messenger RNA in mouse L cells. Cell 1:37–42.
6.
Desrosiers R, Friderici K, Rottman F. 1974. Identification of methylated nucleosides in messenger RNA from novikoff hepatoma cells. Proc Natl Acad Sci U S A 71:3971–3975.
7.
Fu Y, Dominissini D, Rechavi G, He C. 2014. Gene expression regulation mediated through reversible m6A RNA methylation. Nat Rev Genet 15:293–306.
8.
Wang S, Sun C, Li J, Zhang E, Ma Z, Xu W, Li H, Qiu M, Xu Y, Xia W, Xu L, Yin R. 2017. Roles of RNA methylation by means of N6-methyladenosine (m6A) in human cancers. Cancer Letters 408:112–120.
9.
Bodi Z, Button JD, Grierson D, Fray RG. 2010. Yeast targets for mRNA methylation. Nucleic Acids Res 38:5327–5335.
10.
Hsu PJ, Shi H, He C. 2017. Epitranscriptomic influences on development and disease. Genome Biol 18:197.
11.
Cai J, Hu J, Amara U, Park SJ, Li Y, Jeong D, Lee I, Xu T, Kang H. 2023. Arabidopsis N6-methyladenosine methyltransferase FIONA1 regulates floral transition by affecting the splicing of FLC and the stability of floral activators SPL3 and SEP3. J Exp Bot 74:864–877.
12.
Fabian M, Gao M, Zhang X-N, Shi J, Vrydagh L, Kim S-H, Patel P, Hu AR, Lu H. 2023. The flowering time regulator FLK controls pathogen defense in Arabidopsis thaliana. Plant Physiol 191:2461–2474.
13.
Amara U, Hu J, Cai J, Kang H. 2023. FLK is an mRNA m6A reader that regulates floral transition by modulating the stability and splicing of FLC in arabidopsis. Mol Plant 16:919–929.
14.
Hu J, Manduzio S, Kang H. 2019. Epitranscriptomic RNA methylation in plant development and abiotic stress responses. Front Plant Sci 10:500.
15.
Jeon J, Lee SH. 2021. RNA modification and its implication in plant pathogenic fungi. Plant Pathol J 37:505–511.
16.
Yang Y, Hsu PJ, Chen Y-S, Yang Y-G. 2018. Dynamic transcriptomic m6A decoration: writers, erasers, readers and functions in RNA metabolism. Cell Res 28:616–624.
17.
Zaccara S, Ries RJ, Jaffrey SR. 2019. Reading, writing and erasing mRNA methylation. Nat Rev Mol Cell Biol 20:608–624.
18.
Bokar JA, Shambaugh ME, Polayes D, Matera AG, Rottman FM. 1997. Purification and cDNA cloning of the AdoMet-binding subunit of the human mRNA (N6-adenosine)-methyltransferase. RNA 3:1233–1247.
19.
Liu J, Yue Y, Han D, Wang X, Fu Y, Zhang L, Jia G, Yu M, Lu Z, Deng X, Dai Q, Chen W, He C. 2014. A METTL3–METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat Chem Biol 10:93–95.
20.
Ping X-L, Sun B-F, Wang L, Xiao W, Yang X, Wang W-J, Adhikari S, Shi Y, Lv Y, Chen Y-S, Zhao X, Li A, Yang Y, Dahal U, Lou X-M, Liu X, Huang J, Yuan W-P, Zhu X-F, Cheng T, Zhao Y-L, Wang X, Rendtlew Danielsen JM, Liu F, Yang Y-G. 2014. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res 24:177–189.
21.
Horiuchi K, Kawamura T, Iwanari H, Ohashi R, Naito M, Kodama T, Hamakubo T. 2013. Identification of Wilms’ tumor 1-associating protein complex and its role in alternative splicing and the cell cycle. J Biol Chem 288:33292–33302.
22.
Yue Y, Liu J, Cui X, Cao J, Luo G, Zhang Z, Cheng T, Gao M, Shu X, Ma H, Wang F, Wang X, Shen B, Wang Y, Feng X, He C, Liu J. 2018. VIRMA mediates preferential M6A mRNA methylation in 3′UTR and near stop codon and associates with alternative polyadenylation. Cell Discov 4:10.
23.
Zhang Z, Theler D, Kaminska KH, Hiller M, de la Grange P, Pudimat R, Rafalska I, Heinrich B, Bujnicki JM, Allain F-T, Stamm S. 2010. The YTH domain is a novel RNA binding domain. J Biol Chem 285:14701–14710.
24.
Du H, Zhao Y, He J, Zhang Y, Xi H, Liu M, Ma J, Wu L. 2016. YTHDF2 destabilizes m6A-containing RNA through direct recruitment of the CCR4-NOT deadenylase complex. Nat Commun 7:12626.
25.
Wang X, Lu Z, Gomez A, Hon GC, Yue Y, Han D, Fu Y, Parisien M, Dai Q, Jia G, Ren B, Pan T, He C. 2014. N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 505:117–120.
26.
Zheng G, Dahl JA, Niu Y, Fedorcsak P, Huang C-M, Li CJ, Vågbø CB, Shi Y, Wang W-L, Song S-H, et al. 2013. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol Cell 49:18–29.
27.
Jia G, Fu Y, Zhao X, Dai Q, Zheng G, Yang Y, Yi C, Lindahl T, Pan T, Yang Y-G, He C. 2011. N6-Methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat Chem Biol 7:885–887.
28.
Yadav PK, Rajasekharan R. 2017. The m6A methyltransferase Ime4 epitranscriptionally regulates triacylglycerol metabolism and vacuolar morphology in haploid yeast cells. J Biol Chem 292:13727–13744.
29.
Clancy MJ, Shambaugh ME, Timpte CS, Bokar JA. 2002. Induction of sporulation in Saccharomyces cerevisiae leads to the formation of N6-methyladenosine in mRNA: a potential mechanism for the activity of the IME4 gene. Nucleic Acids Res 30:4509–4518.
30.
Yadav PK, Rajasekharan R. 2018. The m6A methyltransferase Ime4 and mitochondrial functions in yeast. Curr Genet 64:353–357.
31.
Agarwala SD, Blitzblau HG, Hochwagen A, Fink GR. 2012. RNA methylation by the MIS complex regulates a cell fate decision in yeast. PLoS Genet 8:e1002732.
32.
Park ZM, Belnap E, Remillard M, Rose MD. 2023. Vir1p, the yeast homolog of virilizer, is required for mRNA m6A methylation and meiosis. Genetics 224:iyad043.
33.
Ensinck I, Maman A, Albihlal WS, Lassandro M, Salzano G, Sideri T, Howell SA, Calvani E, Patel H, Bushkin G, Ralser M, Snijders AP, Skehel M, Casañal A, Schwartz S, van Werven FJ. 2023. The yeast RNA methylation complex consists of conserved yet reconfigured components with m6A-dependent and independent roles. Elife 12:RP87860.
34.
Kang H-J, Jeong S-J, Kim K-N, Baek I-J, Chang M, Kang C-M, Park Y-S, Yun C-W. 2014. A novel protein, Pho92, has a conserved YTH domain and regulates phosphate metabolism by decreasing the mRNA stability of PHO4 in Saccharomyces cerevisiae. Biochem J 457:391–400.
35.
Schwartz S, Agarwala SD, Mumbach MR, Jovanovic M, Mertins P, Shishkin A, Tabach Y, Mikkelsen TS, Satija R, Ruvkun G, Carr SA, Lander ES, Fink GR, Regev A. 2013. High-resolution mapping reveals a conserved, widespread, dynamic mRNA methylation program in yeast meiosis. Cell 155:1409–1421.
36.
Varier RA, Sideri T, Capitanchik C, Manova Z, Calvani E, Rossi A, Edupuganti RR, Ensinck I, Chan VWC, Patel H, Kirkpatrick J, Faull P, Snijders AP, Vermeulen M, Ralser M, Ule J, Luscombe NM, van Werven FJ. 2022. N6-methyladenosine (m6A) reader Pho92 is recruited co-transcriptionally and couples translation to mRNA decay to promote meiotic fitness in yeast. Elife 11:e84034.
37.
Scutenaire J, Plassard D, Matelot M, Villa T, Zumsteg J, Libri D, Séraphin B. 2023. The S. cerevisiae m6A-reader Pho92 promotes timely meiotic recombination by controlling key methylated transcripts. Nucleic Acids Res 51:517–535.
38.
Harigaya Y, Tanaka H, Yamanaka S, Tanaka K, Watanabe Y, Tsutsumi C, Chikashige Y, Hiraoka Y, Yamashita A, Yamamoto M. 2006. Selective elimination of messenger RNA prevents an incidence of untimely meiosis. Nature 442:45–50.
39.
Shichino Y, Otsubo Y, Kimori Y, Yamamoto M, Yamashita A. 2018. YTH-RNA-binding protein prevents deleterious expression of meiotic proteins by tethering their mRNAs to nuclear foci. Elife 7:e32155.
40.
Trail F. 2009. For blighted waves of grain: Fusarium graminearum in the postgenomics era. Plant Physiol 149:103–110.
41.
Wang Z, Kim W, Wang Y-W, Yakubovich E, Dong C, Trail F, Townsend JP, Yarden O. 2023. The sordariomycetes: an expanding resource with big data for mining in evolutionary genomics and transcriptomics. Front Fungal Biol 4:1214537.
42.
Ma L-J, Geiser DM, Proctor RH, Rooney AP, O’Donnell K, Trail F, Gardiner DM, Manners JM, Kazan K. 2013. Fusarium pathogenomics. Annu Rev Microbiol 67:399–416.
43.
Feng C, Cao X, Du Y, Chen Y, Xin K, Zou J, Jin Q, Xu J-R, Liu H, Lin X. 2022. Uncovering cis-regulatory elements important for A-to-I RNA editing in Fusarium graminearum. mBio 13:e0187222.
44.
Xin K, Zhang Y, Fan L, Qi Z, Feng C, Wang Q, Jiang C, Xu J-R, Liu H. 2023. Experimental evidence for the functional importance and adaptive advantage of A-to-I RNA editing in fungi. Proc Natl Acad Sci U S A 120:e2219029120.
45.
Cao S, He Y, Hao C, Xu Y, Zhang H, Wang C, Liu H, Xu J-R. 2017. RNA editing of the AMD1 gene is important for ascus maturation and ascospore discharge in Fusarium graminearum. Sci Rep 7:4617.
46.
Shi Y, Wang H, Wang J, Liu X, Lin F, Lu J. 2019. N6-methyladenosine RNA methylation is involved in virulence of the rice blast fungus Pyricularia oryzae (syn magnaporthe oryzae). FEMS Microbiol Lett 366.
47.
Ren Z, Tang B, Xing J, Liu C, Cai X, Hendy A, Kamran M, Liu H, Zheng L, Huang J, Chen X-L. 2022. MTA1‐mediated RNA m 6 A modification regulates autophagy and is required for infection of the rice blast fungus. New Phytol 235:247–262.
48.
Duan H-C, Wei L-H, Zhang C, Wang Y, Chen L, Lu Z, Chen PR, He C, Jia G. 2017. ALKBH10B is an RNA N 6-methyladenosine demethylase affecting arabidopsis floral transition. Plant Cell 29:2995–3011.
49.
Amara U, Shoaib Y, Kang H. 2022. ALKBH9C, a potential RNA m6A demethylase, regulates the response of arabidopsis to abiotic stresses and abscisic acid. Plant Cell Environ 45:3566–3581.
50.
Shoaib Y, Hu J, Manduzio S, Kang H. 2021. Alpha-ketoglutarate-dependent dioxygenase homolog 10B, an N6-methyladenosine mRNA demethylase, plays a role in salt stress and abscisic acid responses in Arabidopsis thaliana. Physiol Plant 173:1078–1089.
51.
Meza TJ, Moen MN, Vågbø CB, Krokan HE, Klungland A, Grini PE, Falnes PØ. 2012. The DNA dioxygenase ALKBH2 protects Arabidopsis thaliana against methylation damage. Nucleic Acids Res 40:6620–6631.
52.
Liefke R, Windhof-Jaidhauser IM, Gaedcke J, Salinas-Riester G, Wu F, Ghadimi M, Dango S. 2015. The oxidative demethylase ALKBH3 marks hyperactive gene promoters in human cancer cells. Genome Med 7:66.
53.
Zhao S, Devega R, Francois A, Kidane D. 2021. Human ALKBH6 is required for maintenance of genomic stability and promoting cell survival during exposure of alkylating agents in pancreatic cancer. Front Genet 12:635808.
54.
Ma L, Lu H, Tian Z, Yang M, Ma J, Shang G, Liu Y, Xie M, Wang G, Wu W, Zhang Z, Dai S, Chen Z. 2022. Structural insights into the interactions and epigenetic functions of human nucleic acid repair protein ALKBH6. J Biol Chem 298:101671.
55.
Westbye MP, Feyzi E, Aas PA, Vågbø CB, Talstad VA, Kavli B, Hagen L, Sundheim O, Akbari M, Liabakk N-B, Slupphaug G, Otterlei M, Krokan HE. 2008. Human AlkB homolog 1 is a mitochondrial protein that demethylates 3-methylcytosine in DNA and RNA. J Biol Chem 283:25046–25056.
56.
Ougland R, Lando D, Jonson I, Dahl JA, Moen MN, Nordstrand LM, Rognes T, Lee JT, Klungland A, Kouzarides T, Larsen E. 2012. ALKBH1 is a histone H2A dioxygenase involved in neural differentiation. Stem Cells 30:2672–2682.
57.
Liu F, Clark W, Luo G, Wang X, Fu Y, Wei J, Wang X, Hao Z, Dai Q, Zheng G, Ma H, Han D, Evans M, Klungland A, Pan T, He C. 2016. ALKBH1-mediated tRNA demethylation regulates translation. Cell 167:1897.
58.
Zhang C, Samanta D, Lu H, Bullen JW, Zhang H, Chen I, He X, Semenza GL. 2016. Hypoxia induces the breast cancer stem cell phenotype by HIF-dependent and ALKBH5-mediated m6A-demethylation of NANOG mRNA. Proc Natl Acad Sci U S A 113:E2047–E2056.
59.
Zhang S, Zhao BS, Zhou A, Lin K, Zheng S, Lu Z, Chen Y, Sulman EP, Xie K, Bögler O, Majumder S, He C, Huang S. 2017. m6A demethylase ALKBH5 maintains tumorigenicity of glioblastoma stem-like cells by sustaining FOXM1 expression and cell proliferation program. Cancer Cell 31:591–606.
60.
Martínez-Pérez M, Aparicio F, López-Gresa MP, Bellés JM, Sánchez-Navarro JA, Pallás V. 2017. Arabidopsis m6A demethylase activity modulates viral infection of a plant virus and the m6A abundance in its genomic RNAs. Proc Natl Acad Sci U S A 114:10755–10760.
61.
Miguel-Rojas C, Cavinder B, Townsend JP, Trail F. 2023. Comparative transcriptomics of Fusarium graminearum and Magnaporthe oryzae spore germination leading up to infection. mBio 14:e0244222.
62.
Kim W, Miguel-Rojas C, Wang J, Townsend JP, Trail F. 2018. Developmental dynamics of long noncoding RNA expression during sexual fruiting body formation in Fusarium graminearum. mBio 9:e01292-18.
63.
Park ZM, Sporer A, Kraft K, Lum K, Blackman E, Belnap E, Yellman C, Rose MD. 2023. Kar4, the yeast homolog of METTL14, is required for mRNA m6A methylation and meiosis. bioRxiv:2023.01.29.526094.
64.
Kim W, Wang Z, Kim H, Pham K, Tu Y, Townsend JP, Trail F. 2022. Transcriptional divergence underpinning sexual development in the fungal class sordariomycetes. mBio 13:e0110022.
65.
Liu H, Li Y, Chen D, Qi Z, Wang Q, Wang J, Jiang C, Xu J-R. 2017. A-to-I RNA editing is developmentally regulated and generally adaptive for sexual reproduction in Neurospora crassa. Proc Natl Acad Sci U S A 114:E7756–E7765.
66.
Wang C, Xu J-R, Liu H. 2016. A-to-I RNA editing independent of ADARs in filamentous fungi. RNA Biol 13:940–945.
67.
Teichert I, Dahlmann TA, Kück U, Nowrousian M. 2017. RNA editing during sexual development occurs in distantly related filamentous ascomycetes. Genome Biol Evol 9:855–868.
68.
Amses KR, Simmons DR, Longcore JE, Mondo SJ, Seto K, Jerônimo GH, Bonds AE, Quandt CA, Davis WJ, Chang Y, et al. 2022. Diploid-dominant life cycles characterize the early evolution of fungi. Proc Natl Acad Sci U S A 119:e2116841119.
69.
Merényi Z, Krizsán K, Sahu N, Liu X-B, Bálint B, Stajich JE, Spatafora JW, Nagy LG. 2023. Genomes of fungi and relatives reveal delayed loss of ancestral gene families and evolution of key fungal traits. Nat Ecol Evol 7:1221–1231.
70.
van Kempen M, Kim SS, Tumescheit C, Mirdita M, Lee J, Gilchrist CLM, Söding J, Steinegger M. 2023. Fast and accurate protein structure search with foldseek. Nat Biotechnol.
71.
Ruperti F, Papadopoulos N, Musser JM, Mirdita M, Steinegger M, Arendt D. 2023. Cross-phyla protein annotation by structural prediction and alignment. Genome Biol 24:113.
72.
Katoh K, Standley DM. 2013. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol 30:772–780.
73.
Capella-Gutiérrez S, Silla-Martínez JM, Gabaldón T. 2009. trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25:1972–1973.
74.
Stamatakis A. 2014. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30:1312–1313.
75.
Catlett NL, Lee B-N, Yoder OC, Turgeon BG. 2003. Split-marker recombination for efficient targeted deletion of fungal genes. Fungal Genet Newsl 50:9–11.
76.
Yu J-H, Hamari Z, Han K-H, Seo J-A, Reyes-Domínguez Y, Scazzocchio C. 2004. Double-joint PCR: a PCR-based molecular tool for gene manipulations in filamentous fungi. Fungal Genet Biol 41:973–981.
77.
Carroll AM, Sweigard JA, Valent B. 1994. Improved vectors for selecting resistance to hygromycin. Fungal Genet Newsl 41:22.
78.
Hallen-Adams HE, Cavinder BL, Trail F. 2011. Fusarium graminearum from expression analysis to functional assays, p 79–101. In Xu JR, BH Bluhm (ed), Fungal genomics: methods in molecular biology. Humana Press, Totowa, NJ.
79.
Teichert I, Wolff G, Kück U, Nowrousian M. 2012. Combining laser microdissection and RNA-seq to chart the transcriptional landscape of fungal development. BMC Genomics 13:511.
80.
Klittich CJR, Leslie JF. 1988. Nitrate reduction mutants of Fusarium moniliforme (Gibberella fujikuroi). Genetics 118:417–423.
81.
Cappellini RA, Peterson JL. 1965. Macroconidium formation in submerged cultures by a non-sporulating strain of Gibberella zeae. Mycologia 57:962–966.
82.
Schmittgen TD, Livak KJ. 2008. Analyzing real-time PCR data by the comparative CT method. Nat Protoc 3:1101–1108.
83.
King R, Urban M, Hammond-Kosack MCU, Hassani-Pak K, Hammond-Kosack KE. 2015. The completed genome sequence of the pathogenic ascomycete fungus Fusarium graminearum. BMC Genomics 16:544.

Information & Contributors

Information

Published In

cover image mSphere
mSphere
Volume 9Number 130 January 2024
eLocator: e00552-23
Editor: Geraldine Butler, University College Dublin, Dublin, Ireland
PubMed: 38085094

History

Received: 21 September 2023
Accepted: 31 October 2023
Published online: 12 December 2023

Keywords

  1. N6-methyladenosine
  2. m6A RNA methylation
  3. MT-A70
  4. MTA1
  5. Fusarium graminearum

Data Availability

All data generated or analyzed during this study are included in this published article and its supplementary information files.

Contributors

Authors

Hyeonjae Kim
Korean Lichen Research Institute, Sunchon National University, Suncheon, South Korea
Author Contributions: Formal analysis, Investigation, and Writing – original draft.
Jianzhong Hu
Department of Applied Biology, College of Agriculture and Life Sciences, Chonnam National University, Gwangju, South Korea
Author Contributions: Investigation and Methodology.
Hunseung Kang
Department of Applied Biology, College of Agriculture and Life Sciences, Chonnam National University, Gwangju, South Korea
Author Contributions: Methodology and Supervision.
Korean Lichen Research Institute, Sunchon National University, Suncheon, South Korea
Department of Applied Biology, College of Agriculture and Life Sciences, Chonnam National University, Gwangju, South Korea
Author Contributions: Conceptualization, Formal analysis, Funding acquisition, Supervision, Visualization, Writing – original draft, and Writing – review and editing.

Editor

Geraldine Butler
Editor
University College Dublin, Dublin, Ireland

Notes

The authors declare no conflict of interest.

Metrics & Citations

Metrics

Note:

  • For recently published articles, the TOTAL download count will appear as zero until a new month starts.
  • There is a 3- to 4-day delay in article usage, so article usage will not appear immediately after publication.
  • Citation counts come from the Crossref Cited by service.

Citations

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. For an editable text file, please select Medlars format which will download as a .txt file. Simply select your manager software from the list below and click Download.

View Options

Figures and Media

Figures

Media

Tables

Share

Share

Share the article link

Share with email

Email a colleague

Share on social media

American Society for Microbiology ("ASM") is committed to maintaining your confidence and trust with respect to the information we collect from you on websites owned and operated by ASM ("ASM Web Sites") and other sources. This Privacy Policy sets forth the information we collect about you, how we use this information and the choices you have about how we use such information.
FIND OUT MORE about the privacy policy