Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2025 Mar 6.
Published before final editing as: Wiley Interdiscip Rev RNA. 2023 Sep 6:e1810. doi: 10.1002/wrna.1810

Ghost authors revealed: The structure and function of human N6-methyladenosine RNA methyltransferases

Kurtis Breger 1, Charlotte N Kunkler 1, Nathan J O’Leary 1, Jacob P Hulewicz 1, Jessica A Brown 1
PMCID: PMC10915109  NIHMSID: NIHMS1954742  PMID: 37674370

Abstract

Despite the discovery of modified nucleic acids nearly 75 years ago, their biological functions are still being elucidated. N6-methyladenosine (m6A) is the most abundant modification in eukaryotic messenger RNA (mRNA) and has also been detected in non-coding RNAs, including long non-coding RNA, ribosomal RNA, and small nuclear RNA. In general, m6A marks can alter RNA secondary structure and initiate unique RNA–protein interactions that can alter splicing, mRNA turnover, and translation, just to name a few. Although m6A marks in human RNAs have been known to exist since 1974, the structures and functions of methyltransferases responsible for writing m6A marks have been established only recently. Thus far, there are four confirmed human methyltransferases that catalyze the transfer of a methyl group from S-adenosylmethionine (SAM) to the N6 position of adenosine, producing m6A: methyltransferase-like protein (METTL) 3/METTL14 complex, METTL16, METTL5, and zinc-finger CCHC-domain-containing protein 4. Though the methyltransferases have unique RNA targets, all human m6A RNA methyltransferases contain a Rossmann fold with a conserved SAM-binding pocket, suggesting that they utilize a similar catalytic mechanism for methyl transfer. For each of the human m6A RNA methyltransferases, we present the biological functions and links to human disease, RNA targets, catalytic and kinetic mechanisms, and macromolecular structures. We also discuss m6A marks in human viruses and parasites, assigning m6A marks in the transcriptome to specific methyltransferases, small molecules targeting m6A methyltransferases, and the enzymes responsible for hypermodified m6A marks and their biological functions in humans. Understanding m6A methyltransferases is a critical steppingstone toward establishing the m6A epitranscriptome and more broadly the RNome.

This article is categorized under:

RNA Interactions with Proteins and Other Molecules > Protein-RNA Recognition

RNA Interactions with Proteins and Other Molecules > RNA-Protein Complexes

RNA Interactions with Proteins and Other Molecules > Protein-RNA Interactions: Functional Implications

Keywords: N6-methyladenosine, RNA methyltransferase, RNA modification

1 ∣. INTRODUCTION

About 75 years ago, the first nucleic acid modifications were reported when 5-methylcytidine (5mC) was discovered in DNA (Hotchkiss, 1948) followed by pseudouridine in RNA (Cohn & Volkin, 1951). Since the discovery of the first naturally occurring nucleic acid modifications, 58 DNA (Sood et al., 2019) and 143 RNA (Boccaletto et al., 2018; McCown et al., 2020) modifications across all domains of life have been discovered to date. The simplest and most abundant modification in nucleic acids is a methyl group (—CH3), whose transfer from a methyl donor cofactor is catalyzed by a methyltransferase (MTase) or “writer” protein. The most common methyl donor is S-adenosylmethionine (SAM or AdoMet), which is the second most abundant metabolite in human cells after ATP (Cantoni, 1977; J. Gao et al., 2018; German et al., 1983). A single methylation of adenosine at the N6 position produces N6-methyladenosine (6mA) in DNA and (m6A) in RNA. 6mA is the most common DNA modification in prokaryotes, but there is a lack of evidence for 6mA being a bona fide human DNA modification and for the presence of a human 6mA MTase. Though some of the human m6A RNA MTases can methylate DNA in vitro (L. Q. Chen et al., 2022; Qi et al., 2022; Woodcock et al., 2019; Yu, Horton, et al., 2021), 6mA in mammals is currently thought to originate from misincorporation of recycled m6A that was generated by nucleotide-salvage pathway (X. Li, Zhang, et al., 2021; Musheev et al., 2020), a misidentification from a contamination such as bacterial plasmid (Kong et al., 2022), or due to antibody cross reacting with unmethylated adenosine (Douvlataniotis et al., 2020). Therefore, we focus this review on human m6A and the human m6A RNA MTases. We refer you to the following reviews to learn about 6mA in other organisms (O'Brown & Greer, 2016; C. Shen et al., 2022; K. J. Wu, 2020).

The m6A modification was first observed in eukaryotic messenger RNA (mRNA) nearly 50 years ago (Desrosiers et al., 1974) and, a few years later, shown to be present within a specific consensus motif that has since been verified to be DRACH (where D = A/G/U, R = A/G, H = A/C/U, underlined A denotes methylated A here and henceforth; Csepany et al., 1990; Dominissini et al., 2012; Harper et al., 1990; Linder et al., 2015; Meyer et al., 2012; Narayan & Rottman, 1988; Wei & Moss, 1977). For human mRNA, approximately 0.5% of all mRNA adenosines are m6A, which is ~3 m6A marks per transcript, and about half of the m6A marks are detected in multiple cell types (Dominissini et al., 2012, 2016; Körtel et al., 2021; Lavi et al., 1977; Meyer et al., 2012; Wei et al., 1975; Wei & Moss, 1977). Though most m6A marks fall within the DRACH motif, a notable fraction (~1–6%) are found elsewhere (Körtel et al., 2021; C. Liu, Sun, et al., 2022). It was not until the 1990s that the MTase responsible for the mRNA m6A modification was discovered to be MT-A70 (Bokar et al., 1997; Rottman et al., 1994), now known as the methyltransferase-like protein (METTL) 3/METTL14 complex. The expansion of methods for RNA sequencing and for detecting m6A led to the discovery of m6A throughout all domains of life. In addition to mRNA, m6A has been found in most human noncoding RNA (ncRNA) classes, including circular RNA (Y. Yang et al., 2017), chromatin-associated RNAs such as enhancer RNA and promoter upstream transcripts (J. Liu et al., 2020; Louloupi et al., 2018; Xiao et al., 2019; W. Xu et al., 2021; W. Xu, He, et al., 2022), long noncoding RNA (lncRNA; Dominissini et al., 2012; Meyer et al., 2012), micro RNA (miRNA; Alarcón et al., 2015), both the small and large subunits of ribosomal RNA (rRNA) (Taoka et al., 2018), small nuclear RNA (snRNA; Epstein et al., 1980; Harada & Kato, 1980), and small nucleolar RNA (Linder et al., 2015). Only within the past 6 years have more human m6A RNA MTases been confirmed: METTL16, which to date has been confirmed to methylate the methionine adenosyltransferase 2A (MAT2A) mRNA and U6 snRNA (Pendleton et al., 2017); METTL5, which methylates A1832 in 18S rRNA (Ignatova et al., 2020; Rong et al., 2020; Van Tran et al., 2019; Xing et al., 2020), and zinc-finger CCHC-domain-containing protein 4 (ZCCHC4), which methylates A4220 in 28S rRNA (H. Ma et al., 2019; Pinto et al., 2020; Ren et al., 2019; Van Tran et al., 2019; Figure 1). In addition to m6A MTases, there are also m6A demethylases (or “eraser” proteins) and m6A-binding proteins (or “reader” proteins). Human m6A demethylases include the fat mass and obesity-associated protein (FTO; Jia et al., 2011) and α-ketoglutarate-dependent dioxygenase alkB homolog 5, RNA demethylase (ALKBH5) (Zheng et al., 2013). However, it is important to note that the prevalence of m6A demethylation is debatable (Darnell et al., 2018; Mauer & Jaffrey, 2018; Murakami & Jaffrey, 2022; B. S. Zhao et al., 2018). Of the known human m6A-binding proteins, five contain a YT521-B homology (YTH) domain (YTHDC1,2 and YTHDF1,2,3), though the domain is not necessary for binding m6A, as other proteins, including the eukaryotic translation initiation factor 3 subunit A (eIF3A), bind to m6A (Y. Li, Bedi, et al., 2020; Meyer et al., 2015; C. Xu et al., 2014, 2015). Therefore, the m6A modification facilitates dynamic biological outcomes, with MTases writing the mark and binding proteins reading the mark. Though m6A exists in most RNA classes and throughout all domains of life, this review focuses on the structure and function of human m6A RNA MTases (summarized in Table 1). We also highlight m6A MTases in viruses and parasites, methods for detecting m6A in an MTase-specific manner, small molecules that target m6A RNA MTases, hypermodified m6A marks and the enzymes that catalyze them, and future directions of m6A MTase research.

FIGURE 1.

FIGURE 1

Schematics of known RNA substrates and targets of human m6A methyltransferases. RNA probes for METTL3/METTL14 were designed and used by J. Liu et al. (2014). MAT2A mRNA hairpins 1–6 are shown in their predicted “duckbill” structure, along with the structure of hairpin 1 observed when bound to METTL16 (PDB ID: 6DU4). Hairpins are arranged 5′ to 3′, whereby the dashed line represents missing intervening nucleotides between each hairpin. U6 snRNA is shown as the predicted structure along with a proposed METTL16-bound structure that mimics the MAT2A hairpin bound structure at the site of nonamer sequence (Aoyama et al., 2020; Doxtader et al., 2018). The MALAT1 triple helix structure is based on the solved crystal structure (PDB ID: 4PLX). 18S and 28S rRNAs are presented as the structure within the assembled human 80S ribosome (PDB ID: 4UG0). For all RNA secondary structures, a dash (—) represents a canonical Watson–Crick base pair, while a dot (•) represents any non-canonical nucleotide interaction. Red nucleotides show consensus motifs with the modified A in blue. Cm in 18S rRNA and Um in 28S rRNA denote a nucleotide with a 2′-O-methyl modification. MALAT1, metastasis-associated lung adenocarcinoma transcript 1; MAT2A, methionine adenosyl transferase 2A; METTL3, methyltransferase-like protein 3; METTL14, methyltransferase-like protein 14; mRNA, messenger RNA; PDB ID, Protein Data Bank identifier.

TABLE 1.

Summary of features for each human m6A RNA methyltransferase.

METTL3 METTL14 METTL16 METTL5 ZCCHC4
Phylogenetic conservation of MTase domain (C. Liu, Cao, et al., 2022) Eukaryotes Eukaryotes Prokaryotes and eukaryotes Eukaryotes Multicellular eukaryotes
RNA targets for methylation mRNA
lncRNA
miRNA
None, catalytically inactive MAT2A mRNA
U6 snRNA (A43)
18S rRNA (A1832) 28S rRNA (A4220, also known as A4190)
Motif DRACH DRACH ACAGAR in stem loop or bulge UAACA UAACG in loop of stem-loop
Protein–protein interactions Stable complex with HAKAI, KIAA1429/VIRMA, METTL14, RBM15, WTAP, and ZC3H13 Stable complex with HAKAI, KIAA1429/VIRMA, METTL3, RBM15, WTAP, and ZC3H13 Transient binding partners, for example, eIF3A/B, eIF4E2, La, LARP7, MePCE, and those associated with DNA replication and repair, 7SK RNP, spliceosome, rRNA processing and translation Stable complex with TRMT112; transiently binds eIFs and eEFs (Rong et al., 2020) Transient binding partners, for example, nucleolar-localized proteins and those involved in ribosome/RNA metabolism (Pinto et al., 2020)
Subcellular localization Nuclear speckles, nucleus; cytoplasm for some cancers Nuclear speckles, nucleus Nucleolus, nucleus, cytoplasm Nucleolus (enriched), nucleus, cytoplasm Nucleolus (enriched), nucleus, cytoplasm
KO mice Lethal Lethal Lethal Subviable Unknown

Note: Unless indicated in this table, please refer to Section 2 for references.

Abbreviations: eEF, eukaryotic translation elongation factor; eIF3A/B, eukaryotic translation initiation factor 3 subunit A/B; HAKAI, E3 ubiquitin-protein ligase Hakai; LARP7, La ribonucleoprotein 7, transcriptional regulator; lncRNA, long noncoding RNA; m6A, N6-methyladenosine; MePCE, methylphosphate capping enzyme; METTL3, methyltransferase-like protein 3; METTL5, methyltransferase-like protein 5; METTL14, methyltransferase-like protein 14; METTL16, methyltransferase-like protein 16; miRNA, micro RNA; mRNA, messenger RNA; RBM15, RNA-binding motif protein 15; RNP, ribonucleoprotein; VIRMA/KIAA1429, vir-like m6A methyltransferase associated protein; WTAP, Wilms' tumor 1-associating protein; ZCCHC4, zinc-finger CCHC-domain-containing protein 4; ZC3H13, zinc-finger CCCH domain-containing protein 13.

2 ∣. BIOLOGICAL FUNCTIONS OF HUMAN m6A RNA MTases

2.1 ∣. Methyltransferase-like proteins 3/14

METTL3 is a roughly 65-kDa MTase initially isolated in 1994 from HeLa nuclei (Rottman et al., 1994). A comparative sequence analysis suggested that the MTase domain of human METTL3 is conserved within over 40 proteins, including homologs in organisms such as Saccharomyces cerevisiae, Arabidopsis thaliana, and Drosophila melanogaster but not in bacteria (see Table 1; Bujnicki et al., 2002). Human METTL14 was also identified as a possible homolog of METTL3 (Bujnicki et al., 2002). METTL3 and METTL14 were originally thought to be independent, catalytically active m6A MTases that, together, have increased activity (J. Liu et al., 2014). However, it is now understood that METTL3 and METTL14 form a stable heterodimer and only METTL3 is the catalytically active MTase, as METTL14 has a degenerate active site motif of EPPL (see Sections 7.4 and 7.5; Śledź & Jinek, 2016; P. Wang, Doxtader, & Nam, 2016; X. Wang, Feng, et al., 2016; Y. Wang, Li, et al., 2014). Instead of a catalytic function, METTL14 enhances METTL3 methylation capacity by facilitating RNA binding (Śledź & Jinek, 2016; P. Wang, Feng, et al., 2016; X. Wang, Feng, et al., 2016). Though the METTL3/METTL14 heterodimer is catalytically active in a test tube, in vivo it is part of a ~1000-kDa holocomplex that includes other protein-binding partners such as Wilms' tumor 1-associating protein (WTAP), vir-like m6A methyltransferase-associated protein (VIRMA or KIAA1429), RNA-binding motif protein 15 (RBM15), the E3 ubiquitin-protein ligase Cbl proto-oncogene like 1 (CBLL1) more commonly referred to as E3 ubiquitin-protein ligase Hakai (HAKAI, named after “destruction” in Japanese), and zinc-finger CCCH domain-containing protein 13 (ZC3H13; J. Liu et al., 2014, 2018; Patil et al., 2016; Wen et al., 2018). WTAP is a notable adaptor protein because it is required for maintaining m6A abundance (Ping et al., 2014), and it mediates localization of METTL3/METTL14 to nuclear speckles in HeLa cells (Bokar et al., 1997; Ping et al., 2014). However, METTL3 and METTL14 have also been detected in the nucleoplasm and cytoplasm depending on cell type, cellular conditions (stress, cancer, etc.), stoichiometry of components in the holocomplex, and possibly some post-translational modifications (Arcidiacono et al., 2020; S. Lin et al., 2016; Schöller et al., 2018). METTL3 has 10 phosphorylation sites, most of which occur outside of the MTase domain (Matsuoka et al., 2007; Schöller et al., 2018; H. L. Sun et al., 2020). METTL3 phosphorylation is not crucial for MTase activity, interactions with WTAP, or subcellular localization (Schöller et al., 2018), however, phosphorylation of METTL3 by the extracellular signal-regulated kinases pathway allows for easier deubiquitination to stabilize the METTL3/METTL14 complex in mouse embryonic stem cells (mESCs; H. L. Sun et al., 2020). METTL14 appears to have only one phosphorylation site (i.e., S399), whose function is unclear because it is neither important for binding to METTL3 nor MTase activity (Schöller et al., 2018).

With METTL3/METTL14 residing predominantly in the nucleus, this complex methylates nascent pre-mRNA and other RNA polymerase II-transcribed RNAs in a co-transcriptional manner in HeLa cells (Ke et al., 2017). Most m6A marks deposited by METTL3/METTL14 are detected within the DRACH consensus motif in either single- or double-stranded RNA (N. Liu et al., 2013; Meiser et al., 2020; X. Wang, Feng, et al., 2016; Figure 1) and are enriched near stop codons and in 3′ untranslated regions (UTRs; Dominissini et al., 2012; Ke et al., 2017; Meyer et al., 2012) due to the exon junction complex, namely eIF4A3, blocking METTL3 from binding to sites in the mRNA coding sequence (X. Yang et al., 2022). The METTL3/METTL14 complex methylates over 15,000 sites in ~25–60% of mRNAs (human embryonic kidney 293 cells with mutant Simian virus 40 large T antigen [HEK293T]) (Dominissini et al., 2012; He & He, 2021; Koh et al., 2018). Only ç5% of DRACH sites are methylated (Dominissini et al., 2012; He & He, 2021). Although the mechanism of site selection is not well understood, it appears that RBM15 and RBM15B bind to U-rich regions flanking DRACH sites, which subsequently recruits the METTL3/METTL14/WTAP complex (Patil et al., 2016). The plethora of m6A marks raises multiple questions: What is the function of each m6A mark? What percent of each site is methylated? Do all m6A sites on an entire population of a particular transcript need to be there to elicit a biological function? Although context-dependent answers are emerging for select m6A sites, mRNA metabolism may be altered at various steps, such as alternative splicing (Dominissini et al., 2012; Zhou et al., 2019), polyadenylation (J. Liu et al., 2018), nuclear export (Roundtree et al., 2017), decreased mRNA stability/half-life (H. Du, Zhao, et al., 2016; X. Wang, Lu, et al., 2014), and translation (S. Lin et al., 2016). In another study of METTL3 function, alternative splicing events in human liver cancer cell line (HepG2) cells are disrupted when METTL3 is silenced (Dominissini et al., 2012). The METTL3/METTL14 complex can also engage with other cellular machinery, such as the translation initiation factor elF3 to promote eIF4E-independent translation of mRNAs containing m6A in 5′-UTR (Choe et al., 2018; S. Lin et al., 2016; Meyer et al., 2015), the transcription factor CCAAT enhancer-binding protein zeta to aid translation (Barbieri et al., 2017), chromatin to silence retroviral elements in mammalian cells (W. Xu et al., 2021), or for nascent RNA to escape premature termination by the Integrator complex (W. Xu, He, et al., 2022). Further regulation is conferred by “reader” proteins that recognize m6A sites, including those that depend upon a “structural switch” (N. Liu et al., 2015, 2017) or facilitate phase separation (Y. Fu & Zhuang, 2020; Y. Gao et al., 2019; D. Han, Longhini, et al., 2022; Lee et al., 2021; Ries et al., 2019).

In mice, the Mettl3 and Mettl14 genes are both essential, as Mettl3−/− knockout mice and Mettl14−/− knockout mice are embryonic lethal (Geula et al., 2015; Meng et al., 2019; Y. Wang, Li, et al., 2018). Conditional knockout mice and knockout cell lines are viable, although numerous stem cells lacking Mettl3 or Mettl14 exhibit impaired differentiation (Batista et al., 2014; Geula et al., 2015; Meng et al., 2019; Y. Wang, Li, et al., 2018). However, a recent study reports that these so-called “knockouts” may simply be alternatively spliced variants of METTL3 that are still capable of methylation (Poh et al., 2022). Knockout of Mettl3 in mouse cells reduces m6A levels on mRNA by ~70–100% (Geula et al., 2015; Z. Lin et al., 2017; C. X. Wang, Cui, et al., 2018). The loss of Mettl3 and/or Mettl14 impacts a multitude of physiological processes, such as brain development (K. Du et al., 2021; Flamand & Meyer, 2019; Madugalle et al., 2020; C. X. Wang, Cui, et al., 2018; Yoon et al., 2017), the cardiovascular system (Dorn et al., 2019; Yao et al., 2020), immunity (W. Qiu, Zhang, et al., 2021; Tong et al., 2021), memory (Z. Zhang et al., 2018), spermatogenesis (Z. Lin et al., 2017), and ultraviolet-induced DNA damage response (Xiang et al., 2017). Thus, it is not surprising that the METTL3/METTL14 complex and global m6A levels are implicated in human health and disease. Reduced expression of METTL3 and/or METTL14 contribute to Type 2 diabetes (De Jesus et al., 2019), obesity (Y. Wang, Gao, et al., 2020), premature aging (Z. Wu et al., 2020), features of osteoporosis (Y. Wu et al., 2018), and select cancer types such as renal cell carcinoma (X. Li et al., 2017; Y. Zhang, Yao, Qi, et al., 2021). In contrast, the upregulation of METTL3 and/or METTL14 contribute to acute myeloid leukemia (AML; Barbieri et al., 2017; Sorci et al., 2018; Vu et al., 2017; Weng et al., 2018), lung cancer (Choe et al., 2018; S. Lin et al., 2016; Wanna-udom et al., 2020), and pancreatic cancer (M. Wang, Liu, et al., 2020; T. Xia et al., 2019), just to name a few. There is some ambiguity with functions of METTL3/METTL14 in cancers. For example, in human hepatocellular carcinoma, METTL3 upregulation promotes cancer proliferation, tumorigenicity, and metastasis (M. Chen et al., 2018) whereas METTL14 downregulation promotes cancer metastasis and invasion (J. Z. Ma et al., 2017). Likewise, conflicting results regarding whether METTL3 and METTL14 facilitate or suppress tumor growth in glioblastoma stem cells have been reported (Q. Cui et al., 2017; X. Jiang, Liu, et al., 2021; Visvanathan et al., 2018). The implications of METTL3 and METTL14 in various cancers are summarized in recent reviews (L. J. Deng, Deng, et al., 2022; Guan et al., 2022; C. Zeng et al., 2020). Depending upon the disease, both activation and inhibition of the METTL3/METTL14 complex may prove fruitful as possible therapeutics (see Section 6).

2.2 ∣. Methyltransferase-like protein 16

Compared to METTL3/METTL14, much less is known about the biological function of METTL16 since it was 2017 when METTL16 was first shown to have m6A MTase activity (Pendleton et al., 2017). There are various homologs of human METTL16 such as 23S rRNA m6A1618 methyltransferase (RlmF; Sergiev et al., 2008) in Escherichia coli, METT-10 in Caenorhabditis elegans (Dorsett & Schedl, 2009), and FIONA1 in A. thaliana (see Table 1; J. Kim et al., 2008). The cellular localization of endogenous METTL16 is cell-type and cell-cycle dependent for the protein has been detected in the nucleolus, nucleus, and cytoplasm of various human cell lines (Brown et al., 2016; Nance et al., 2020; Sjöstedt et al., 2020; P. Wang, Wang, et al., 2020; Weng et al., 2018). Currently, there are seven RNA substrates of human METTL16 that have been validated in vitro: the U6 snRNA and each of the six hairpins (hp1–6) in the 3′-UTR of MAT2A mRNA (Mendel et al., 2018; Pendleton et al., 2017; Shima et al., 2017; Warda et al., 2017). METTL16 preferentially methylates adenosines near structured RNA regions, such as the nonamer sequence UACAGARAA (R = A or G) present in both U6 snRNA and the MAT2A hp1–6 (Doxtader et al., 2018; Mendel et al., 2018; Pendleton et al., 2017; Figure 1). In mice, two additional mRNA substrates of METTL16 were identified: Fanconi anemia complementation group M and breast cancer 2 protein or BRCA2 (Yoshinaga et al., 2022). For human U6 snRNA, m6A43 is part of the ACAGAGA box that base pairs with the 5′-splice sites of pre-mRNAs during splicing (Wassarman & Steitz, 1992). Indeed, the m6A mark in both Schizosaccharomyces pombe and Arabidopsis thaliana U6 snRNA presumably forms a stable m6A•A+4 base pair with the intron to enable efficient splicing regardless of the 5′ exon sequence (Ishigami et al., 2021; Parker et al., 2022). METTL16 indirectly controls SAM concentrations because METTL16 regulates MAT2A protein levels via methylation of the MAT2A mRNA hp1 (Doxtader et al., 2018; Pendleton et al., 2017; Shima et al., 2017). METTL16 will remain on hp1 if SAM levels are low, allowing for proper splicing of the MAT2A mRNA and subsequently protein synthesis. However, if SAM levels are high, then METTL16 will methylate and dissociate from hp1, leading to intron-detained MAT2A mRNA and less MAT2A protein (Pendleton et al., 2017). This mechanism involves the cleavage factor Im (CFIm) complex; specifically a CFIm25 dimer recruits CFIm59 and CFIm68 to mediate splicing of MAT2A while METTL16 acts as an upstream SAM sensor (Scarborough et al., 2021). However, the interplay between the CFIm complex and METTL16 requires further elucidation. A similar mode of regulation exists in C. elegans, where METT-10, the ortholog of human METTL16, controls the splicing and subsequently abundance of mRNAs for SAM-producing enzymes (Mendel et al., 2021; Watabe et al., 2021). In this case, high levels of SAM lead to methylation of a 3′ splice site (AG) that interferes with splicing factor U2AF35 recognition of the splice site (Mendel et al., 2021; Watabe et al., 2021).

METTL16 has an RNA interactome that is over 4000 mRNAs and also binds to various ncRNA classes (e.g., lncRNA, miRNA, and snRNA) based on a crosslinking and analysis of complementary DNA assay (R. Su, Dong, et al., 2022; Warda et al., 2017). METTL16 binds to the triple helix at the 3′-end of mature metastasis-associated lung adenocarcinoma transcript 1 (MALAT1; Brown et al., 2016; Warda et al., 2017). Although m6A signals have been identified via m6A-seq analysis (methylation [or m6A]-individual-nucleotide resolution crosslinking and immunoprecipitation [miCLIP]; Linder et al., 2015) and nanopore direct RNA sequencing (Krusnauskas et al., 2023), a recent study suggests that METTL16 does not methylate the MALAT1 triple helix (Figure 1) under in vitro conditions (Breger & Brown, 2023). Thus, the methylome of METTL16 is smaller than its interactome. The m6A-crosslinking-exonuclease-sequencing (m6ACE-seq) method (described later in Section 4) identified 136 METTL16-dependent m6A sites (Koh et al., 2018) while methylated RNA immunoprecipitation sequencing (MeRIP-seq) identified 334 mRNAs with METTL16-dependent m6A sites (R. Su, Dong, et al., 2022).

Unlike the METTL3/METTL14 heterodimer, METTL16 engages in dynamic protein–protein interactions (Ignatova et al., 2019). For the methylation of U6 snRNA, METTL16 interacts with the proteins La, La ribonucleoprotein 7, transcriptional regulator (LARP7), and methylphosphate capping enzyme (MePCE) in an RNA-dependent manner (Covelo-Molares et al., 2021; Warda et al., 2017). Other notable proteins that interact with METTL16 include those associated with DNA replication and repair, 7SK ribonucleoprotein (RNP), spliceosome, rRNA processing, and translation (Covelo-Molares et al., 2021; Stixová et al., 2021; R. Su, Dong, et al., 2022; Warda et al., 2017). Interestingly, cytosolic METTL16 enhances translation efficiency in a methylation-independent manner, likely via interactions between the MTase domain of METTL16 and eIF3A/B and/or eIF4E2 proteins that are critical for the formation of 80S ribosomes (R. Su, Dong, et al., 2022; F. Wang et al., 2023). METTL16 was also shown to bind certain RNAs to associate with double-strand break repair protein MRE11 (MRE11) in an RNA-dependent manner, preventing MRE11-dependent DNA-end resection (X. Zeng et al., 2022).

With roles in SAM homeostasis, differentiation, and DNA-damage response, METTL16 is an essential protein in normal human cells (T. Wang et al., 2015) and genetic knockout of Mettl16 in mice results in embryonic lethality (Mendel et al., 2018; Yoshinaga et al., 2022). Recent analyses suggest that METTL16 may play a role in human health and disease because (i) METTL16 is upregulated in multiple cancers including AML (L. Han et al., 2023), breast cancer (F. Ye et al., 2023; B. Zhang et al., 2020), colon cancer (S. Wang, Fan, et al., 2021), esophageal cancer (H. Zhao, Xu, et al., 2021), hepatocellular carcinoma (R. Su, Dong, et al., 2022), osteosarcoma (Cheng et al., 2023), and stomach cancer (X. K. Wang, Zhang, et al., 2021), and (ii) METTL16 is downregulated in liver cancer (W. Wang, Sun, et al., 2020), patients with hepatocellular carcinoma (P. Wang, Wang, et al., 2020), cases of infertility (S. Zhao, Lu, et al., 2021), and instances involving the absence of gut microbiota (Jabs et al., 2020). Interestingly, a recent study found that METTL16 is the most crucial METTL protein for cancer survival out of 25 METTLs and that knockout of METTL16 in hepatocellular carcinoma, where METTL16 is overly abundant, led to the suppression of tumorigenesis (R. Su, Dong, et al., 2022). In some cases, METTL16 appears to mediate cancer independent of its “writer” role because METTL16 acts as a translational regulator via interactions with eIF3A/B, eIF4E2 and 18S rRNA (R. Su, Dong, et al., 2022; F. Wang et al., 2023). Future studies into METTL16 will likely focus on the function of METTL16 as a SAM sensor and its possible interactions with the splicing complex as well as elucidating whether the m6A within U6 snRNA aids base pairing for splicing in humans akin to S. pombe and A. thaliana.

2.3 ∣. A pair of m6A rRNA MTases: METTL5 produces m6A1832 in 18S rRNA and ZCCHC4 produces m6A4220 in 28S rRNA

In human lymphoblast cell line (TK6) cells, 228 RNA modifications have been detected in human rRNA but only two of those modifications are m6A: m6A1832 in 18S rRNA and m6A4220 (also known as m6A4190 using alternative numbering system) in 28S rRNA (Taoka et al., 2018) (Figure 1). Although both of these m6A marks were discovered over 30 years ago (Maden, 1986, 1988), it is only recently that the writers of these marks have been identified: METTL5 methylates A1832 and ZCCHC4 methylates A4220 (Table 1). Both MTases have been rigorously validated using various knockout cell lines that all show a reduction in m6A levels for the appropriate rRNA (Ignatova et al., 2020; H. Ma et al., 2019; Pinto et al., 2020; Rong et al., 2020; Sepich-Poore et al., 2022; Van Tran et al., 2019; Xing et al., 2020). Furthermore, methylated rRNA levels can be restored in the knockout cell lines via expression of catalytically active, but not inactive mutants of, METTL5 (Sepich-Poore et al., 2022) or ZCCHC4 (H. Ma et al., 2019). Knockout cell lines are viable, indicating neither MTase is essential. However, as described below, mechanistic studies have revealed that each MTase contributes to critical roles in translation and human health.

m6A1832 resides in a UAACA motif of helix 44 (Figure 1), which is proximal to the decoding center in the mature human ribosome (Natchiar et al., 2017). However, METTL5, which forms a heterodimer with the protein tRNA methyltransferase activator subunit 112 (TRMT112) (Brūmele et al., 2021; Sepich-Poore et al., 2022; Van Tran et al., 2019), likely adds the m6A mark at a time when A1832 is accessible during the biogenesis of 18S rRNA, which is in accord with endogenous METTL5 being detected predominantly in the nucleoli of HEK293T, HeLa, epidermoid carcinoma (A-431), and human osteosarcoma (U2OS) cell lines (Rong et al., 2020; Sepich-Poore et al., 2022; Sjöstedt et al., 2020). Furthermore, METTL5 appears to methylate only A1832 because anti-METTL5 miCLIP and enhanced crosslinking and immunoprecipitation studies detected METTL5 binding to primarily 18S rRNA and detected no significant changes in m6A levels of mRNA in METTL5-knockout cells (Van Tran et al., 2019). In vitro MTase activity assays showed that METTL5 modifies short, synthetic 18S rRNA mimics and purified 18S rRNA but does not methylate RNAs with AACU, GAACU, GGACU, and UAACG motifs (Rong et al., 2020; Xing et al., 2020).

METTL5 is expressed in all human cell types examined thus far (Sjöstedt et al., 2020) and various Mettl5-knockout cell lines have been generated using human (near-haploid human cell line [HAP1], human colon cancer cell line [HCT116], HEK293T, HeLa, HepG2) and mouse (B16, mESC, mouse embryonic fibroblasts) cells. Although there are some cell-type specific observations, the general trends that have emerged upon loss of METTL5 are as follows: differentiation is impaired (Ignatova et al., 2020; Xing et al., 2020), pluripotency is delayed/lost (Ignatova et al., 2020; Xing et al., 2020), no change in 18S rRNA levels but the number of polysomes is reduced (Ignatova et al., 2020; Rong et al., 2020), and translation is inefficient (Ignatova et al., 2020; Rong et al., 2020). Notably, ribosome profiling or sequencing experiments have revealed that decreased translation occurs on differentiation-promoting genes like F-box/WD repeat-containing protein 7 (Fbxw7; Xing et al., 2020). In addition, translation initiation is attenuated because there are fewer translation initiation/elongation factors associated with the 80S ribosome and phosphorylated ribosomal protein S6 kinase B1, a critical signal for translation initiation, is reduced (Rong et al., 2020). For human cell lines examined to date, it is estimated that ~60–100% of A1832 sites are modified with m6A (Ignatova et al., 2020; N. Liu et al., 2013; Rong et al., 2020; Taoka et al., 2018), a finding that prompts further investigation into the relationship between ribosome heterogeneity and functionality, particularly with respect to the stress response. Ultimately, additional experiments are needed to understand how a single methyl group at the N6 position of A1832 controls both global and gene-specific translational effects.

At the organism level, Mettl5-deficient mice are subviable, displaying several craniofacial abnormalities, cardiac defects, hearing and vision problems, infertility, reduced size, and inactivity (Y. Han, Du, et al., 2022; Ignatova et al., 2020; Sepich-Poore et al., 2022). Such observations are consistent with microcephaly and other developmental disorders observed in human patients having mutations in the Mettl5 gene (Reuter et al., 2017; Riazuddin et al., 2017; Richard et al., 2019; Torun et al., 2022). Phylogenetic analysis shows that METTL5 is conserved in most eukaryotes, except for fungi/yeast (Leismann et al., 2020). Thus, it is not surprising that parallel functional observations have been reported for METTL5 homologs in C. elegans (Liberman et al., 2020) (m6A1717), D. melanogaster (Leismann et al., 2020), and Danio rerio (i.e. zebrafish; Richard et al., 2019). It has also been noted that METTL5 expression decreases in aged human tissues (Rong et al., 2020) and gastric cancer (X. K. Wang, Zhang, et al., 2021) but increases in various tumor types (TCGA Research Network), including breast cancer tissue and cell lines (Rong et al., 2020), liver (Peng et al., 2022; W. Xu, Liu, et al., 2022), lung, and pancreatic cancer (H. Huang et al., 2022; S. Sun et al., 2021; Yan et al., 2021; Z. Zhang, Zhang, Luo, et al., 2021). Because the proliferation of breast and liver cancer cell lines decreased upon short interfering RNA knockdown of METTL5 (Peng et al., 2022; Rong et al., 2020; W. Xu, Liu, et al., 2022), METTL5 is a potential target for anti-cancer therapy.

m6A4220 is located in a UAACG loop motif of helix 81, which is adjacent to the active center of the mature human ribosome (Natchiar et al., 2017; Pinto et al., 2020). A4220 appears to be the major target of ZCCHC4, for both miCLIP and photoactivatable ribonucleoside-enhanced crosslinking and immunoprecipitation against ZCCHC4 did not show enrichment for other RNA targets (H. Ma et al., 2019; Van Tran et al., 2019). ZCCHC4 exhibits the greatest in vitro MTase activity when given a stem-loop RNA substrate containing UAAC in the loop (Pinto et al., 2020; Ren et al., 2019). Depending on cell type, ZCCHC4 has been detected in the nucleolus, nucleus, and cytoplasm (H. Ma et al., 2019; Pinto et al., 2020; Sjöstedt et al., 2020) and the percent of A4220 that is methylated ranges from ~55–100% (N. Liu et al., 2013; H. Ma et al., 2019; Pinto et al., 2020; Taoka et al., 2018; Van Tran et al., 2019). Thus, studies of various ZCCHC4-deficient cell lines (HAP1, HeLa, HepG2, HCT116, mouse embryonic cell line [J1 ES], mouse fibroblast cell line [NIH 3T3]) have some cell-type differences but have converged on the following conclusions: loss of ZCCHC4 does not alter rRNA production (H. Ma et al., 2019; Pinto et al., 2020; Van Tran et al., 2019) but decreases global translation (H. Ma et al., 2019) and ribosome occupancy (Pinto et al., 2020). Although ZCCHC4 orthologs are encoded in a wide range of animals (Pinto et al., 2020), no ZCCHC4-knockout animals have been reported. However, ZCCHC4 may be implicated in cancer and chemotherapy resistance (Z. Zhang, Zhang, Yang, et al., 2021). For example, ZCCHC4 is overexpressed and higher levels of methylated 28S rRNA are detected in hepatocellular carcinoma tumor tissue compared to normal tissue samples (H. Ma et al., 2019). Importantly, xenograft mouse models using Zcchc4-knockout cells showed reduced tumor growth compared to wild-type cells (H. Ma et al., 2019) and similarly, HepG2 cells lacking ZCCHC4 exhibit reduced cell proliferation (H. Ma et al., 2019). We anticipate that future studies will address how the loss of ZCCHC4 affects various model organisms and will establish how the N6-methyl group of A4220 results in specific phenotypes.

3 ∣. ROLES OF m6A MTases IN INFECTIOUS AGENTS: VIRUSES AND PARASITES

N6-methylation is an important modification in the progression of infectious diseases and has been identified in many human pathogens. Although m6A in viral RNA was first identified in the mRNA of influenza virus (Krug et al., 1976) and adenovirus type-5 (Moss & Koczot, 1976; Sommer et al., 1976) in 1976, m6A marks have since been identified in the genomes and/or transcriptomes of various viruses, such as herpes simplex virus 1 (HSV-1; Feng et al., 2022; Moss et al., 1977), Simian virus 40 (SV40; Finkel & Groner, 1983; Tsai et al., 2018), Rous sarcoma virus (RSV; Kane & Beemon, 1985), human immunodeficiency virus 1 (HIV-1; Kennedy et al., 2016; Lichinchi, Gao, et al., 2016), flaviviruses such as Zika virus (ZIKV), West Nile virus, yellow fever, dengue, and others (Gokhale et al., 2016), hepatitis B (Imam et al., 2018) and C (HBV and HCV; Gokhale et al., 2016), Kaposi's sarcoma-associated herpes virus (KSHV; Baquero-Perez et al., 2019; Tan et al., 2017; F. Ye et al., 2017), Epstein–Barr virus (EBV; Lang et al., 2019), enterovirus 71 (EV71; Hao et al., 2019), porcine endemic diarrhea virus (PEDV; J. Chen, Jin, et al., 2020), human metapneumovirus (hMPV; Lu et al., 2020), severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2; Burgess et al., 2021; J. Liu et al., 2021), and human papillomavirus 16 (X. Cui et al., 2022). The percentage of adenosines with m6A marks in viral RNA genomes varies from 0.089% in EV71 to 3% in ZIKV (Lichinchi, Zhao, et al., 2016). Eighteen m6A marks were identified in respiratory syncytial viral mRNAs, the highest for any virus (Xue et al., 2019). During viral infection, m6A levels of host transcripts can increase or decrease depending on the particular m6A mark (Gokhale et al., 2016; Jansens et al., 2022; N. Li, Hui, et al., 2021; Lichinchi, Gao, et al., 2016; Lichinchi, Zhao, et al., 2016; Lichinchi & Rana, 2019; J. Liu et al., 2021; X. Qiu, Hua, et al., 2021; Selberg et al., 2021). Interestingly, cellular m6A methylation is almost completely inhibited in pseudorabies virus-infected cells due to phosphorylation of the METTL3/METTL14/WTAP complex by viral proteins (Jansens et al., 2022).

To date, no instances of virally encoded m6A MTases have been discovered; therefore, m6A modifications are likely mediated via expression and cellular localization changes of host writers and erasers. HIV-1, HCV, KSHV, and HBV all depend on the host m6A writer METTL3, which was confirmed by the knockdown or knockout of METTL3 (Gokhale et al., 2016; Imam et al., 2018; Lichinchi, Gao, et al., 2016; Tan & Gao, 2018; F. Ye et al., 2017). Viruses using host METTL3 are further supported by the retention of the DRACH motif, the consensus sequence of METTL3/METTL14 methylation, or at least RAC in viral genomic RNA and/or viral mRNA isolated from HIV-1, HCV, KSHV, influenza virus, HBV, SV40, respiratory syncytial virus, ZIKV, and hMPV (Bayoumi & Munir, 2021; Courtney et al., 2017; Gokhale et al., 2016; Imam et al., 2018; Lichinchi, Gao, et al., 2016; Lichinchi & Rana, 2019; Tsai et al., 2018; Xue et al., 2019; F. Ye et al., 2017). Conserved methylation sites were also mapped in many Flaviviridae genomes, particularly in the nonstructural protein 3 (NS3) helicase, nonstructural protein 5 (NS5) RNA-dependent RNA-polymerase (RdRp), and envelope genes (Gokhale et al., 2016). m6A marks in viral genomes are also frequently colocalized with hairpins and G-quadruplexes, suggesting that some m6A sites in viruses require RNA structure for recognition by the m6A writer (Fleming et al., 2019). In HBV, a viral protein known as HBV X protein directs METTL3/METTL14 to viral RNA as well as at least one host mRNA, phosphatase and tensin homolog or PTEN (G. W. Kim & Siddiqui, 2021).

m6A marks affect the viral life cycle at various stages. For example, m6A marks dampen viral replication for ZIKV, KSHV, EBV, HBV, HCV, PEDV, and SARS-CoV-2 by as much as fourfold (J. Chen, Jin, et al., 2020; G. W. Kim et al., 2020; Lang et al., 2019; Lichinchi, Zhao, et al., 2016; J. Liu et al., 2021; Tan et al., 2017) but stimulate viral replication for HIV-1, influenza A virus, human cytomegalovirus, SV40, EV71, respiratory syncytial virus, and hMPV by as much as 10-fold (Courtney et al., 2017; Hao et al., 2019; Lichinchi, Gao, et al., 2016; Lu et al., 2020; Rubio et al., 2018; Tsai et al., 2018; Xue et al., 2019). EV71 infection noticeably upregulates expression of METTL3/METTL14 (Hao et al., 2019). The EV71 RdRp 3D interacts with METTL3, which enhances sumoylation and ubiquitination of the RdRp 3D, thereby increasing viral replication (Hao et al., 2019; Y. Liu et al., 2016). For ZIKV, knockdown of METTL3/METTL14 decrease m6A levels in ZIKV RNA and lowered ZIKV RNA release by approximately twofold, while knockdown of FTO and ALKBH5 had the opposite effect (Lichinchi, Zhao, et al., 2016). An increase in m6A marks also increases viral RNA nuclear export (Casu et al., 2013; Lichinchi, Zhao, et al., 2016), enhances splicing (Price et al., 2020; F. Ye, 2017), differentially affects viral RNA stability and protein expression (Imam et al., 2018; Tsai et al., 2018), and allows the viral RNA to avoid detection by the host via pattern recognition receptors such as RIG-I-like receptors or Toll-like receptors (Durbin et al., 2016; Karikó et al., 2005; N. Li, Hui, et al., 2021; Lu et al., 2020; W. Qiu, Zhang, et al., 2021).

Much of the viral-centric research has examined effects upon METTL3 knockout/knockdown, yet only recently have METTL16, METTL5, and ZCCHC4 been validated as m6A MTases. This point presents interesting questions for future research: Are there roles for m6A MTases other than METTL3/METTL14 in methylating viral RNAs? Are other writers, such as METTL16 under SAM-limiting conditions, playing any regulatory roles? The involvement of other m6A writers seems possible given (i) their ability to interact with highly structured RNAs, which are common in viral genomes and transcripts, and (ii) some m6A marks survive disruption of the DRACH motif in SV40 (Tsai et al., 2018). When m6A levels increase or decrease in viral RNAs, what is the regulatory interplay among the m6A machinery of writers, readers, and erasers? Are there instances where methylation at specific sites triggers structural switches (Imam et al., 2018)? Would drugs directly targeting specific m6A MTases be effective antivirals? Although no MTase-specific inhibitors have been used as an antiviral therapeutic, it is worth mentioning that 3-deazaadenosine, a nucleoside analog that inhibits S-adenosyl-L-homocysteine (SAH) hydrolase, has been successful as an antiviral for RSV, HSV-1, reovirus, measles, vesicular stomatitis virus, parainfluenza virus, respiratory syncytial virus, and HIV-1 in various model systems (Bader et al., 1978; De Clercq & Montgomery, 1983; Mayers et al., 1995; Montgomery et al., 1982; Wyde et al., 1990), suggesting that changes to the cellular methylation potential via MTases may be effective.

In addition to viruses, m6A impacts the biology of human parasitic infections in the parasite Plasmodium falciparum, which causes malaria, and the obligate intracellular parasite Toxoplasma gondii, which causes toxoplasmosis. m6A regulates the parasitic life cycle by controlling the intra-erythrocytic development stage in P. falciparum and the growth to the bradyzoite stage in T. gondii (Baumgarten et al., 2019; Holmes et al., 2021). Interestingly, these parasites have their own putative METTL3 ortholog MTases as well as associated METTL14 and WTAP orthologs (Baumgarten et al., 2019; Holmes et al., 2021). Additional binding partners were co-precipitated for both parasites, although, they did not match proteins that are analogous to those in the human METTL3/METTL14 complex. The METTL3 orthologs have their own recognition sequence: GGACA in P. falciparum and YGCAUGCR (Y = pyrimidine and R = purine) in T. gondii (Baumgarten et al., 2019; Holmes et al., 2021). Trypanosoma brucei, which causes trypanosomiasis, does not have a METTL3 ortholog. There is likely another MTase that installs m6A marks, for a CAU motif was identified as a potential methylation sequence in coding sequences, UTRs and intergenic regions (L. Liu, Zeng, et al., 2019; Viegas et al., 2022). T. brucei has a reported threefold increase in m6A sites primarily within its coding sequences between T. brucei life stages: the bloodstream form and procyclic form (L. Liu, Zeng, et al., 2019). T. brucei has also been shown to use m6A marks in poly(A) tails to protect against deadenylation and degradation (Viegas et al., 2022). For parasitic infections involving Cryptosporidium parvum, there is a 30% global increase in the m6A methylome of host cells eight hours postinfection due to down regulation of ALKBH5, FTO, and a nuclear factor kappa-light-chain-enhancer of activated B signaling pathway for innate antimicrobial immune defense (Z. Xia, Xu, et al., 2021). Important future areas of study include how widespread METTL3 orthologs are in human parasites and the prevalence of m6A methylome changes for host and parasitic RNAs. Please refer to a review by Catacalos et al for more details on the biology of m6A in parasites (Catacalos et al., 2022).

4 ∣. WHO MADE THE MARK: TRANSCRIPTOME-WIDE m6A DETECTION METHODS

Due to their improved precision and quantitation, detection methods for m6A sites have played a pivotal role in elucidating the biological functions of m6A MTases (O’Brown & Greer, 2016; Zaccara et al., 2019; Z. Zhang et al., 2019; Zhu et al., 2019). Early methods were limited to single-RNA strategies such as mass spectrometry and 2D thin layer chromatography (Mongan et al., 2019). Current methods can now site specifically detect m6A throughout the entire transcriptome (see the review article by Y. Wang & Jia, 2020, for more information on detection methods; Y. Wang & Jia, 2020). Herein we focus on studies that aim to reveal which m6A RNA MTase, namely METTL3/METTL14 versus METTL16, is responsible for the respective marks.

Antibody based m6A-seq methods are fairly common; however, the anti-m6A antibodies bind to m6A introduced by any MTase. A means to determine which m6A marks are introduced by certain MTases is to knockdown or knockout (if cell viability is not disrupted) the particular MTase of interest. This approach has been coupled with a method called m6ACE-seq, which utilizes 5′-3′ exoribonuclease 1 (XRN1), to achieve single-nucleotide resolution (Koh et al., 2019). The antibody photocrosslinked to an m6A-containing RNA protects the RNA at the m6A site from XRN1 digestion, leaving the reads that start at the m6A site (Koh et al., 2019). Knockout of METTL3 and phosphorylated CTD interacting factor 1 (PCIF1), an enzyme that converts 2’-O-methyladenosine (Am) to N6,2′-O-dimethyladenosine (m6Am; see Section 8 for more details about PCIF1), indicated 14,508 m6A sites and 3798 m6Am sites introduced by the respective MTases in HEK293T cells. In contrast, METTL16 knockdown generated ambiguous results because m6A levels fluctuated for 602 m6A sites, including those assigned to be installed by METTL3 or PCIF1. METTL16 regulates SAM homeostasis; therefore, fluctuations in m6A sites not directly targeted by METTL16 were likely due to the dysregulation of SAM/SAH levels. Only 136 m6A sites were deduced to be solely METTL16 dependent. Another study performed m6A MeRIP-seq on HEK293T-METTL16 knockout cells to further identify m6A marks installed by METTL16 (R. Su, Dong, et al., 2022). This study concluded that METTL16 methylates sites in 334 mRNAs. Similarly, methylation CLICK-degradation sequencing (meCLICK-seq) was performed using short hairpin RNA-mediated knockdown of METTL3 and METTL16 in MOLM-13 cells (Mikutis et al., 2020). For meCLICK-seq, cell cultures are grown in the presence of a methionine analog, propargyl-L-selenohomocysteine, which the cell converts into the SAM surrogate propargylic Se-adenosyl-L-selenomethionine, so that MTases will introduce a propargyl, rather than a methyl, modification onto RNA targets (Mikutis et al., 2020). Then, the Cu(I)-catalyzed azide—alkyne cycloaddition degrades methylated RNA, but not propargyl-modified RNA, to enable measurement of m6A modifications through the depletion of RNA (Mikutis et al., 2020). meCLICK-seq findings were reported with respect to RNAs rather than m6A sites: 6130 RNAs had METTL3-dependent m6A marks, 8545 RNAs had METTL16-dependent m6A marks, and 5721 RNAs had both METTL3- and METTL16-dependent m6A marks.

Some recently developed methods do not use antibodies, thereby mitigating antibody cross reactivity to m6Am (Linder et al., 2015) while enriching for m6A marks catalyzed by certain MTases, usually METTL3. Many of these methods operate under the assumption that a METTL3-dependent m6A mark occurs in the DRACH motif. One such method is the deamination adjacent to RNA modification targets (DART-seq) method (Meyer, 2019). Apolipoprotein B mRNA editing enzyme, catalytic polypeptide 1 (APOBEC1), a cytidine deaminase, is fused to a YTH domain that binds to m6A (Meyer, 2019). Because a majority of m6A marks are upstream of C in the DRACH motif, the YTH domain can bind to m6A and then APOBEC1 will convert the adjacent “C” to “U”, providing single-nucleotide resolution of the DRACH-to-DRAUH motif (Meyer, 2019). DART-seq identified 40,263 DRAUH motifs in 7707 RNAs. It is worth pointing out that METTL16 methylates UACAGARAA motifs; therefore, it would be interesting to know if UAUAGARAA motifs were present in the DART-seq results. Another antibody-free method, MAZTER sequencing, uses the enzyme endoribonuclease toxin MazF, which cleaves unmethylated ACA sites to detect unmethylated and methylated ACA locations at single-nucleotide resolution. However, this method is limited to m6A marks occurring in an ACA context so METTL3 sites would be targeted (Garcia-Campos et al., 2019; Pandey & Pillai, 2019).

Third-generation sequencing technologies, such as single-molecule real-time sequencing and Oxford Nanopore Technologies, are now being adapted for the direct detection of modified nucleotides, such as m6A, at single-nucleotide resolution (Lucas & Novoa, 2023). Traditionally, the DNA or RNA is pulled through a nanopore inside a membrane which causes characteristic disruptions in a measured current. These disruptions are unique to specific nucleotides, including modifications like m6A. Although deconvoluting disruptions of modified nucleotides has been challenging (Garalde et al., 2018; H. Liu, Begik, et al., 2019), recent advances in algorithms have improved the accuracy of m6A detection to roughly 90% for in vitro transcribed RNA and 87% from polyA-selected RNA isolated from S. cerevisiae (H. Liu, Begik, et al., 2019). This methodology could also be applied to cell lines with knockdown/knockouts of specific MTases.

In summary, many detection methods have been developed over the years to identify and to quantitate m6A in RNA (Y. Wang & Jia, 2020), although none of the methods can track m6A marks catalyzed by a specific MTase akin to what is possible for 5mC (Dai et al., 2021). Such technology might provide stronger evidence for whether there are other m6A RNA MTases.

5 ∣. HOW TO WRITE m6A: CATALYTIC AND KINETIC MECHANISMS

A majority of 6mA and m6A MTases possess a similar catalytic mechanism of methyl transfer due to the conserved (D/N/S)PP(F/W/Y) catalytic core (see Section 7.5 for more details; Bheemanaik et al., 2006). As shown in Figure 2a, the side chain of D/N/S within the catalytic core forms a hydrogen bond with the amine of the N6 and the other hydrogen of the amine interacts with the carbonyl of the proline-proline peptide backbone. These interactions enhance the nucleophilic capacity of the nitrogen at N6. During the nucleophilic attack, the amine is deprotonated, leading to a C-N bond with the methyl group of SAM. Once the methyl group is fully transferred to adenosine, the other resulting product is SAH (or sometimes referred to as AdoHcy) with a lone pair of electrons on the sulfur. Please note that the methyl group is directly transferred to N6 of adenosine and does not occur via an intermediate such as an N1-intermediate (Pogolotti et al., 1988). The above-described catalytic mechanism is favored by many, although an alternative mechanism involving amino-imino tautomerization has been proposed (Mashhoon & Reich, 1994).

FIGURE 2.

FIGURE 2

Chemical mechanism of methylation catalyzed by m6A RNA methyltransferases and the possible substrate-binding mechanisms for SAM-dependent methyltransferases. (a) Schematic shows the transfer of the methyl group (blue) from SAM, the methyl donor, to the acceptor adenosine, N6-methyladenosine. SAH is a byproduct. Catalysis depends on a conserved (D/N/S)PP(F/W/Y) motif. Note that the F/W/Y portion orients the acceptor adenosine and is not depicted for clarity. (b) The schemes depict binding order for SAM and RNA for random-order binding mechanism, ordered-sequential binding mechanisms, and a double-displacement reaction or ping-pong binding mechanism. m6A, N6-methyladenosine; SAH, S-adenosylhomocysteine; SAM, S-adenosylmethionine.

Little is confirmed about the transition state of m6A MTases. For MTases in general, the transition state tends to involve the proper arrangement of the methyl acceptor, whether it is an adenosine, cytidine, or other substrate, with the SAM methyl donor in the active site (Lopez et al., 2016). The methyl group of adenosine forms a trigonal bipyramidal arrangement in which the hydrogens of the methyl group are in a planar arrangement and the methyl group is shared between the sulfur of SAM and the recipient nitrogen atom of the methyl acceptor (S. Chen, Kapilashrami, et al., 2020; Q. Du, Wang, & Schramm, 2016; Mahmoodi et al., 2020; Newby et al., 2002). Computational modeling suggests that a hydrogen on the N6 may remain as the nitrogen-methyl group forms a bond; the hydrogen will be subsequently removed by a basic group as the methyl-adenosine intermediate has improved acidity relative to the reactant (Newby et al., 2002). For m6A MTases, structural, biochemical, and computational studies have not yet revealed the actual transition state, distances from the methyl group of SAM to the methyl acceptor, conformational rearrangement of the SAM ligand, and other perturbations to the adenosine and/or RNA in 3D space. Nonetheless, it seems likely that m6A MTases may share transition state features with other MTases.

m6A RNA MTases have two different substrates; therefore, the binding order of SAM and RNA may vary depending upon the MTase and RNA substrate. Most characterized MTases employ either a random sequential mechanism, where SAM and DNA/RNA bind in any order (Figure 2b), or an ordered-sequential mechanism, where SAM binds prior to DNA/RNA or vice versa (Figure 2b; Bheemanaik et al., 2006). A few MTases, such as 23S rRNA m2A2503 methyl-transferase/tRNA m2A37 methyltransferase and 23S rRNA adenosine-C8-methyltransferase Cfr, which methylate positions C2 and C8 of A2503 in 23S rRNA, respectfully, use a ping-pong mechanism (Figure 2b), whereby the MTase will receive the methyl group from SAM, forming a methyl-MTase intermediate, then eject SAH before binding and methylating the methyl acceptor DNA/RNA (Grove et al., 2011). Substrate binding order for METTL16 methylating MAT2A hp1 is reported to be a random-sequential binding mechanism (Figure 2b) based on similar rates obtained from a preincubation assay (Yu, Kaur, et al., 2021); however, a solved structure of METTL16 bound to a minimal MAT2A hp1 RNA suggested that METTL16 requires SAM to bind first because the METTL16•RNA complex showed no possible entry for SAM to enter the active site (Doxtader et al., 2018). For U6 snRNA, METTL16 displays an ordered-sequential binding mechanism with U6 snRNA binding prior to SAM (Figure 2b), suggesting the possibility that binding mechanisms may depend on the substrate (Breger & Brown, 2023). The binding order for the other human m6A RNA MTases remains unknown.

Select kinetic parameters of the known human m6A MTases have been measured and the values revealed a wide range despite the high structural similarity of MTases. The turnover numbers (kcat) for METTL3/METTL14 and METTL5 are rather close at 18 and 17.6 h−1, respectively, while METTL16 has turnover rates of 4.2 h−1 for U6 snRNA (Breger & Brown, 2023) and 30 h−1 for the MAT2A hp1 (Table 2; F. Li et al., 2016; Yu, Kaur, et al., 2021). Substrate binding affinity can be evaluated based on the various equilibrium dissociation constant (KD) or Michaelis–Menten constant (KM) values. METTL3/METTL14, METTL5/TRMT112, and ZCCHC4 all have binding values (either KD or KM) for SAM in the range of 1–7 μM (Table 2) (Ren et al., 2019; X. Wang, Feng, et al., 2016; Yu, Kaur, et al., 2021). In contrast, METTL16 has a KD of 126 μM for the METTL16•U6 snRNA•SAM complex (Breger & Brown, 2023). This unusually weak affinity for SAM may be due to METTL16 regulating intracellular SAM concentrations, which are reported to be approximately 10 μM (Breger & Brown, 2023; Doxtader et al., 2018; Pendleton et al., 2017; Shima et al., 2017; C. Ye & Tu, 2018). For RNA substrate binding affinity, METTL16 displays KD values of 16–18 and 42 nM for U6 snRNA and MAT2A hp1, respectively (Table 2; Aoyama et al., 2020; Breger & Brown, 2023), whereas METTL3/METTL14 has a KD at 23 μM and METTL5/TRMT112 has a KM for its rRNA substrate at 1.1 μM (Table 2; Yu, Kaur, et al., 2021). Interestingly, METTL3/METTL14 exhibits a 75-fold greater catalytic efficiency on damaged dsDNA than ssRNA (0.02 vs. 1.5 μM−1 min−1), raising the intriguing possibility that METTL3/METTL14 functions as a 6mA DNA MTase (L. Q. Chen et al., 2022; Woodcock et al., 2019; Yu, Horton, et al., 2021). In addition, protein-binding partners further impact RNA-binding affinity and MTase activity (S. Su, Li, et al., 2022). Most kinetic parameters are yet to be determined for ZCCHC4.

TABLE 2.

Kinetic parameters of the human m6A RNA methyltransferases.

MTase (RNA substrate) k cat KD (E•SAM) or KM KD (E•RNA) or KM
METTL3/METTL14
(RNA with DRACH motifa)
18 ± 2 h−1
(F. Li et al., 2016)
KD = 1.5 ± 0.2 μM
(X. Wang, Feng, et al., 2016)
KD = 23.0 ± 2.3 μM
(J. Huang et al., 2019)
METTL16
(U6 snRNA)
~3 h−1
(Aoyama et al., 2020)
4.2 ± 1.2 h−1
(Breger & Brown, 2023)
KD = 126 ± 6 μMb
(Breger & Brown, 2023)
KM = 132 ± 8 μM
(Breger & Brown, 2023)
KD = 16 ± 2 nM
(Aoyama et al., 2020)
KD = 18 ± 7 nM
(Breger & Brown, 2023)
KM = 25 ± 10 nM
(Aoyama et al., 2020)
METTL16
(MAT2A hp)
~30 h−1
(Yu, Horton, et al., 2021)
KM >0.4 mM
(Yu, Horton, et al., 2021)
KD = 42 ± 6 nM
(Aoyama et al., 2020)
KM = 29 ± 6 nMc
(Aoyama et al., 2020)
KM = ~10 μMc
(Yu, Horton, et al., 2021)
METTL5/TRMT112
(UCGUAACAAGGUUU)
17.6 ± 0.6 h−1
(vary [SAM])
(Yu, Horton, et al., 2021)
13.1 ± 0.8 h−1
(vary [RNA])
(Yu, Horton, et al., 2021)
KM = 1.0 ± 0.2 μM
(Yu, Horton, et al., 2021)
KM = 1.1 ± 0.2 μM
(Yu, Horton, et al., 2021)
ZCCHC4 Unknown KD = 6.7 ± 0.4 μM
(Ren et al., 2019)
Unknown

Abbreviations: KD, equilibrium dissociation constant; KM, Michaelis–Menten constant; m6A, N6-methyladenosine; METTL3, methyltransferase-like protein 3; METTL5, methyltransferase-like protein 5; METTL14, methyltransferase-like protein 14; METTL16, methyltransferase-like protein 16; MTase, methyltransferase; SAM, S-adenosylmethionine; ZCCHC4, zinc-finger CCHC-domain-containing protein 4.

a

Turnover number, kcat, value measured using UACACUCGAUCUGGACUAAAGCUGCUC; KD value measured using AAAAGGACUAAAA.

b

Value reflects METTL16•U6 snRNA•SAM complex.

c

Reason for KM values differing by ~345-fold is unclear but could be due to shortened hp RNA, SAM, or reaction conditions.

Currently, METTL16 is the only human m6A RNA MTase with information regarding rate-limiting steps. For the methylation of U6 snRNA, the steady-state rate of 0.07 min−1 was eightfold less than the single-turnover rate of 0.56 min−1, indicating that steps following methylation were rate-limiting, such as product release or conformational rearrangements (Breger & Brown, 2023). Although the rate-limiting steps of other human m6A RNA MTases are unknown, other MTases exhibit different rate-limiting steps during methylation. METTL3/METTL14 displays slower methylation activity on higher-affinity substrates, which might suggest that product release is rate limiting (Qi et al., 2022). Metal ions, such as zinc ions, and protein cofactors, such as METTL14 and TRMT112, are another consideration for certain m6A MTases. METTL3 has two CCCH domains (i.e., zinc-finger domain [ZnF] 1 and 2) that bind zinc and are needed for MTase activity in vitro (Huang et al. 2019). Similarly, ZCCHC4 requires zinc due to the GRF, C2H2, and CCHC zinc-finger domains (all named after the conserved residues) (Ren et al., 2019). In contrast, zinc ions inhibit the MTase activity of METTL16 and METTL5 (Breger & Brown, 2023; Yu, Kaur, et al., 2021). METTL16 does not appear to have any zinc fingers (Yu, Kaur, et al., 2021) and neither does METTL5, although its binding partner TRMT112 does have a zinc-binding domain (Van Tran et al., 2019). In theory, methyl transfer could also be limited by conformational changes. An interesting distinction between the 6mA and m6A MTases is base flipping. Because DNA is usually double-stranded, 6mA MTases often need to “flip” the adenosine acceptor into an extra-helical position to receive the methyl group (Bheemanaik et al., 2006). In general, it is thought that m6A RNA MTases do not employ base flipping because the targeted adenosines are not base paired or have the rotational freedom to orient N6 for methylation (J. Liu et al., 2014; Pendleton et al., 2017; Van Tran et al., 2019). However, METTL5 appears to have the A1832 more exposed, which suggests the MTase may be capable of base flipping (Van Tran et al., 2019).

6 ∣. SMALL MOLECULES AND OTHER APPROACHES TARGETING m6A MTases

DNA and RNA MTases have been targeted with an array of small molecules, acting as either inhibitors or activators, for both research and therapeutic purposes. Some of the first inhibitors were products of the methylation reaction: methylated DNA/RNA and SAH (Figure 3). Dead-end analog inhibitors for m6A MTases include a DNA/RNA, whereby the adenosine acceptor is substituted with another nucleobase or the use of SAM analogs like sinefungin (Figure 3). Sinefungin was first isolated from Streptomyces incarnatus and Streptomyces griseolus, where it was used as an antibiotic, but due to its similarity in structure to SAM, it is often used as an MTase competitive inhibitor in kinetic assays (Horiuchi et al., 2013; Nolan, 1987; Selberg et al., 2019). Thus, both product and dead-end analog inhibitors have been useful research tools for kinetic assays. There have also been advancements in bisubstrate analogs, molecules that generally mimic both the SAM and targeted adenosine moieties, which inhibit MTases, such as the bacterial rRNA m6A methyltransferase 23S rRNA m6A2030 methyltransferase (RlmJ, Figure 3; Oerum et al., 2019). The bisubstrate analogs may provide mechanistic insights about RNA recognition and the catalytic mechanism, such as the transition state, utilized by the human m6A MTases.

FIGURE 3.

FIGURE 3

Chemical structures of small molecules that target methyltransferases. Where indicated, IC50 and EC50 values have been determined for the METTL3/METTL14/WTAP complex; however, IC50 values for METTL3 inhibitors STM2457 and UZH1a were determined in the presence of METTL3/METTL14. EC50, half-maximal effective concentration; IC50, half-maximal inhibitory concentration; METTL3, methyltransferase-like protein 3; METTL14, methyltransferase-like protein 14; WTAP, Wilms′ tumor 1-associating protein.

Interest in developing small molecules to target MTases, particularly METTL3 (see reference Fiorentino et al., 2023 for review), for therapeutic purposes is growing. Artificial intelligence and high-throughput screening methods, such as the high-throughput-methyl-reading assay, will enable fast and easy means of testing libraries of compounds for potential activation/inhibition of MTases (Xiao et al., 2022) in addition to virtual high-throughput screening methods that have yielded promising hits for METTL3 (Selberg et al., 2019) and METTL16 (Mitra et al., 2023). Inhibitors of METTL3 would be therapeutically useful because inhibition limits the growth of AML (Barbieri et al., 2017). A few companies, such as STORM therapeutics, have discovered inhibitors for METTL3 and other RNA-modifying enzymes (Cully, 2019; Lee et al., 2022). Inhibitors, such as STM2475 and UZH1a (Figure 3), have been observed to occupy the SAM-binding site of METTL3 based on X-ray crystal structures (Table 3) (Moroz-Omori et al., 2021; Yankova et al., 2021). STM2475 and UZH1a exhibit half-maximal inhibitory concentration (IC50) values of 2.2 and 4.6 μM, respectively, in MOLM-13 cells and prevent proliferation of AML cell lines (Moroz-Omori et al., 2021; Yankova et al., 2021). Other molecules are also currently being investigated for their inhibition of METTL3 in the context of AML (Sabnis, 2021). Currently, the drug STC-15, developed by STORM therapeutics to target METTL3 in a solid tumor, is in Phase 1 clinical trials.

TABLE 3.

All currently available structures of the human m6A methyltransferases with their associated Protein Data Bank identifiers, resolutions, amino acid numbers of recombinant protein, and ligand bound (if any).

Protein name PDB ID Resolution (Å) Amino acids Ligand
METTL3/METTL14 5IL0 1.88 MTase (369–580/109–408) Apo
5IL1 1.71 MTase (369–580/109–408) SAM
5IL2 1.61 MTase (369–580/109–408) SAH
5K7M 1.65 MTase (356–580/108–456) Apo
5K7U 1.70 MTase (356–580/108–456) SAM
5K7W 1.65 MTase (356–580/108–456) SAH
5L6D 1.85 MTase (354–580/107–395) SAH
5L6E 1.90 MTase (354–580/107–395) SAM
5TEY 1.80 MTase (407–568/186–363) SAH
5YZ9 (NMR) METTL3 ZnF1-ZnF2 (259–357) Apo
6TTP 2.0 MTase (354–580/107–395) Adenosinea
6TTT 2.30 MTase (354–580/107–395) ASI_M3M_140a
6TTV 2.14 MTase (354–580/107–395) ASI_M3M_138a
6TTW 2.20 MTase (354–580/107–395) ASI_M3M_047a
6TTX 2.00 MTase (354–580/107–395) ASI_M3M_051a
6TU1 2.31 MTase (354–580/107–395) ASI_M3M_091a
6Y4G 1.90 MTase (354–580/107–395) Sinefungina
7ACD 2.50 MTase (354–580/107–395) T30a
7NHG 2.5 MTase (354–580/107–395) ASI_M3M_041a
7NHH 2.1 MTase (354–580/107–395) UOZ002a
7NHI 1.85 MTase (354–580/107–395) UOZ004a
7NHJ 2.16 MTase (354–580/107–395) DHU_M3M_154a
7NHV 1.91 MTase (354–580/107–395) UOZ016a
7NI7 2.5 MTase (354–580/107–395) UOZ031a
7NI8 2.2 MTase (354–580/107–395) UOZ040aa
7NI9 2.2 MTase (354–580/107–395) UOZ058a
7NIA 2.3 MTase (354–580/107–395) UOZ059aa
7NID 2.3 MTase (354–580/107–395) UOZ078a
7O08 2.0 MTase (354–580/107–395) ADO_AB_075a
7O09 1.8 MTase (354–580/107–395) ADO_AC_074a
7O0L 1.9 MTase (354–580/107–395) ADO_AC_093a
7O0M 2.39 MTase (354–580/107–395) ADO_AD_023a
7O0P 2.7 MTase (354–580/107–395) ADO_AD_022a
7O0Q 2.49 MTase (354–580/107–395) ADO_AD_066a
7O0R 2.3 MTase (354–580/107–395) ADO_AE_026a
7O27 2.4 MTase (354–580/107–395) ADO_AE_005a
7O28 2.47 MTase (354–580/107–395) ADO_AE_009a
7O29 2.75 MTase (354–580/107–395) ADO_AD_044a
7O2E 2.5 MTase (354–580/107–395) ADO_AD_089a
7O2F 2.1 MTase (354–580/107–395) UZH2a
7O2H 2.5 MTase (354–580/107–395) ADO_AD_091a
7O2I 3 MTase (354–580/107–395) STM2457a
7O2X 2.80 MTase (354–580/107–395) T180a
7OED 2.0 MTase (354–580/107–395) UOZ019aa
7OEE 2.7 MTase (354–580/107–395) UOZ019ba
7OEF 2.03 MTase (354–580/107–395) UOZ038a
7OEG 2.79 MTase (354–580/107–395) UOZ040ba
7OEH 2.01 MTase (354–580/107–395) UOZ059ba
7OEI 2.48 MTase (354–580/107–395) UOZ083a
7OEJ 2.3 MTase (354–580/107–395) UOZ090a
7OEK 1.9 MTase (354–580/107–395) UOZ091a
7OEL 1.86 MTase (354–580/107–395) UOZ097a
7OEM 2.2 MTase (354–580/107–395) UOZ120a
7OQL 2.5 MTase (354–580/107–395) UOZ094a
7OQO 3.35 MTase (354–580/107–395) UOZ111a
7OQP 2.0 MTase (354–580/107–395) UOZ113a
34169b 4.4 MTase (1–580/1–456); no residues modeled HWVZb
METTL16 6B91 1.94 MTase (1–291) Apo
6B92 2.10 MTase (1–291) SAH
6DU4 1.70 MTase (2–310) MAT2A HP1
6DU5 3.01 MTase (2–310) MAT2A HP6
6GFK 2.30 MTase (1–251) Apo
6GFN 2.86 MTase (1–291) Apo
6GT5 2.45 MTase (1–291) Apo
6M1U 2.79 VCR1 (310–410); VCR2 (509–562) Apo
METTL5/TRMT112 6H2U 1.60 MTase (1–209/1–118) SAM
6H2V 2.49 MTase (1–209/1–125) SAM
ZCCHC4 6UCA 3.10 GRF-C2H2-MTase-CCHC (23–464) SAH

Abbreviations: HAKAI, E3 ubiquitin-protein ligase Hakai; METTL3, methyltransferase-like protein 3; METTL5, methyltransferase-like protein 5; METTL14, methyltransferase-like protein 14; METTL16, methyltransferase-like protein 16; MTase, methyltransferase; SAH, S-adenosylhomocysteine; SAM, S-adenosylmethionine; WTAP, Wilms' tumor 1-associating protein; VCR1/2, vertebrate conserved region 1/2; VIRMA, vir-like m6A methyltransferase associated protein; ZCCHC4, zinc-finger CCHC-domain-containing protein 4; ZC3H13, zinc-finger CCCH domain-containing protein 13.

a

Ligands designed as METTL3/METTL14 inhibitors; please refer to original published source for details on each ligand.

b

ID for Electron Microscopy Data Bank (EMDB) because there is no model deposited in PDB. HWVZ represents the protein-binding partners HAKAI, WTAP, VIRMA, and ZC3H13, respectively.

In addition to inhibitors, molecules that activate or enhance the activity of MTases represent a potential therapeutic approach; activation of METTL3/METTL14 could increase m6A RNA levels and thereby mitigate symptoms of hypomethylation, such as Type 2 diabetes or tumorigenesis (Q. Cui et al., 2017; De Jesus et al., 2019; J. Li & Gregory, 2021; Z. Zhang, Zhang, Luo, et al., 2021). To identify activators, one study employed a virtual drug screening approach and reported four compounds (methyl piperidine-3-carboxylate hydrochloride, tert-butyl 6-methylpiperidine-3-carboxylate, methyl 6-methylpiperidine-3-carboxylate, and methyl piperazine-2-carboxylate; Figure 3) that putatively target the METTL3/METTL14/WTAP complex (Selberg et al., 2019). Subsequent testing of these compounds, methyl piperidine-3-carboxylate hydrochloride and methyl piperazine-2-carboylate, showed that they enhanced methylation activity by approximately 1.2- and 1.5-fold, respectively (Selberg et al., 2019). The KD for METTL3•SAM is 1.92 μM but improves to 4.7 and 13.7 nM in the presence of 1–100 nM methyl piperidine-3-caroxylate hydrochloride and 25 μM methyl piperazine-2-carboylate, respectively (Selberg et al., 2019). A 16–20% increase in m6A levels was detected after treating HEK293 cells for two hours with three of the four compounds (Selberg et al., 2019). Moreover, these activators demonstrated that enhanced methylation stimulated production of HIV-1 particles (Selberg et al., 2021). Computational molecular docking and other in silico approaches have suggested that the drugs nilotinib and lumacaftor (VX-809) might be potential inhibitors of METTL16; however, no empirical evidence of their inhibition of METTL16 activity has been reported yet (Mitra et al., 2023). Another novel design is the incorporation of a temporal photocage modality via conjugation to the METTL3 activator, methyl piperidine-3-carboxylate (Figure 3; Lan et al., 2021). This photocage is removed upon exposure to light, thereby releasing activator and enhancing the levels of methylation at a specific time (Lan et al., 2021). Activators for other MTases, such as METTL16 decreasing intracellular SAM concentrations, may be possible and their use as possible therapeutics is exciting. Integrating the dynamics of the m6A methylome in the context of specific diseases and drug responses is complex; therefore, resources like m6A-centered regulations of disease development and drug response database (M6AREG), which is a database curating m6A regulators, their RNA targets, and drugs associated with different disease states, may be useful (S. Liu et al., 2023).

Lastly, it is also important to note that macromolecular options for manipulating methylation sites, rather than MTases, are beginning to emerge. One prime example is a Cas13b-METTL14 fusion protein that utilizes a guide RNA (gRNA) to target a particular DRACH motif upon endogenous METTL3 binding to the fusion protein and methylating the target site (M. Gao et al., 2022). Other clustered regularly interspaced short palindromic repeats (CRISPR)-associated protein 9/Cas 13-based systems that enable site-specific methylation exist for demethylation, which similarly uses a fusion protein system except the MTase is replaced with either the ALKBH5 or FTO demethylase (Chang et al., 2022; J. Li, Chen, et al., 2020; Wilson et al., 2020). Clever temporal-controlled strategies have emerged, such as catalytically inactive mutant of Cas13b (dCas13b) in combination with abscisic acid-induced proximity technologies. Here, a gRNA directs dCas13b fused to a pyrabactin-resistance like protein (PYL) to a particular RNA site. Then, the addition of abscisic acid enables PYL to interact with an abscisic acid insensitive domain fused to METTL3, facilitating temporal control of m6A sites through the application of abscisic acid (Shi et al., 2022). We recommend the review paper by X. Sun et al. (2022) to learn more about m6A manipulation via CRISPR-Cas methods. The CRISPR-Cas tools can be delivered via engineered viruses for gene therapy, revealing the potential for these tools to be used for therapeutic purposes in humans (Z. Xia, Tang, et al., 2021).

7 ∣. 3D STRUCTURES OF THE HUMAN m6A MTases AND THEIR SUBSTRATE RECOGNITION

High-resolution 3D structures have provided insights into the functions of MTases. All known human m6A MTases have at least one X-ray crystal structure of the MTase domain in complex with either SAM or SAH (Figure 4 and Table 3). The first of the human m6A MTase structures to be solved was the MTase domains of the METTL3/METTL14 complex, in apo form, bound to SAM, and bound to SAH (Šledź & Jinek, 2016; P. Wang, Feng, et al., 2016; X. Wang, Feng, et al., 2016). Since then, slight variations of the METTL3/METTL14 MTase domains have been solved, as well as multiple structures with SAM-analog inhibitors (see Table 3). The next structure solved was the MTase domain of METTL16 in its apo form, bound to SAH (Ruszkowska et al., 2018), and shortly after, bound to a minimal MAT2A hp substrate (Doxtader et al., 2018). More recently, the crystal structure of the vertebrate conserved regions (VCR) in the C-terminal domain was solved, lending a more complete structural view of METTL16 (Aoyama et al., 2020). Next, the structure of the METTL5/TRMT112 complex bound to SAM was solved (Van Tran et al., 2019). Finally, the MTase domain of ZCCHC4 bound to SAH, along with its three zinc-finger domains, was solved (Ren et al., 2019). All five known human m6A MTases are a part of the Class I SAM-dependent MTases, which contain a Rossmann fold and utilize SAM as the methyl donor. In this section, a comparative structural analysis is presented and the unique structural features for each of the human m6A RNA MTases are highlighted.

FIGURE 4.

FIGURE 4

Domain organization, representative X-ray crystal structure, and Rossmann fold schematic for each human m6A MTase. On the left side of figure, MTase domains (orange/pink) contain the Rossmann fold, NLS indicates nuclear localization signal, RGG indicates arginine-glycine repeats, VCR indicates vertebrate conserved regions, ZnF indicates zinc-finger domains, and GRF, C2H2, and CCHC are specific types of zinc-finger domains. The MTase domain of METTL14 is catalytically inactive (pink), however, it still contains a Rossmann fold. Zinc ions are shown as green spheres. All structures are oriented so that the Rossman fold is displayed to match the same orientation. Displayed on the right side of figure are schematics, which are not drawn to scale, depicting the Class Ia and Class Ig Rossmann folds. Arrows and rectangles represent β-strands and helices, respectively, with colored ones showing the Rossmann fold and gray ones showing other known secondary structures outside of the Rossmann fold. Green circles represent zinc ions bound to the zinc-finger domains. Dashed lines represent unknown structures or disordered regions. The blue “N” and red “C” respectively denote the N- and C-termini while the parenthetical superscript denotes residue number. m6A, N6-methyladenosine; METTL14, methyltransferase-like protein 14; MTase, methyltransferase.

7.1 ∣. Rossmann fold

Rao and Rossmann first observed the Rossmann fold, a well conserved “super-secondary structure” found within nucleoside-binding proteins, 50 years ago (Rao & Rossmann, 1973). Most MTases, including all known human m6A RNA MTases, adopt the Rossmann fold composed of at least six β-strands sandwiched by α-helices (Figure 4). X-ray crystal structures of the human m6A MTases exhibit structural similarities, but also a few unique features in their Rossmann folds and in their SAM/SAH-binding pockets (Table 3 and Figures 4 and 5). Though all known human m6A RNA MTases contain a Rossmann fold, the connectivity of the β-strands and α-helices vary among the five structurally characterized human m6A RNA MTases (Figure 4). METTL16, METTL5, and ZCCHC4 have a Class Ia Rossmann fold, where β-strands interact as 3214576 and all β-strands are unidirectional except β7, while METTL3 and METTL14 contain a Class Ig Rossmann fold, where β-strands interact as 1762354 and all β-strands are unidirectional except β5 (Figure 4; Gana et al., 2013). Throughout the rest of this review, β-strand nomenclature follows the numbering system of the Class Ia Rossmann fold for simplicity (i.e., 3214576). Among all the Rossmann folds of amino-MTases, there are nine conserved sequence motifs (Motifs I–VIII and X; Malone et al., 1995). These nine motifs occur within the human m6A MTases, albeit to varying degrees of conservation (Figure 5a). Each of the nine sequence motifs and their occurrence in human m6A RNA MTases are described below.

FIGURE 5.

FIGURE 5

Sequence alignment of five of the conserved Rossmann fold motifs and visualization of motifs within SAM-binding pocket of the human m6A MTases. (a) Gray-shaded boxes show residues that are structurally conserved between the m6A MTases with consensus sequence above the sequences and the structural elements below. As defined in Figure 4, arrows and rectangles represent β-strands and helices, respectively. Structural elements are labeled based on a Class Ia Rossmann fold. (b) Schematics of SAM-binding pockets showing relative location and residues of Motifs I, II, III, IV, and X. Red-dashed lines denote hydrogen bonds (i.e., side chain and/or backbone interactions) and red arcs denote van der Waals interactions between SAM and specific amino acids. Schematics are based on the following X-ray crystal structures: METTL3•SAM (PDB ID: 5IL1), METTL16•SAH (PDB ID: 6B92), METTL5•SAM (PDB ID: 6H2U), and ZCCHC4•SAH (PDB ID: 6UCA). Water molecules are shown as blue circles. Water molecules were not modeled in X-ray crystal structure of ZCCHC4•SAH. m6A, N6-methyladenosine; MTase, methyltransferase; METTL3, methyltransferase-like protein 3; METTL5, methyltransferase-like protein 5; METTL14, methyltransferase-like protein 14; METTL16, methyltransferase-like protein 16; PDB ID, Protein Data Bank identifier; SAH, S-adenosylhomocysteine; SAM, S-adenosylmethionine; ZCCHC4, zinc-finger CCHC-domain-containing protein 4.

7.2 ∣. Motif I

Motif I is located in the β1-loop-αA region, whereby β1 terminates with a conserved hydrophobic amino acid connected to a GxGxG (or GxG) sequence in the loop (Figures 4 and 5a). The hydrophobic amino acid (leucine in METTL3, METTL14, and METTL5; isoleucine in METTL16; and valine in ZCCHC4) packs against αC and αD within the hydrophobic core of the protein via van der Waals interactions. The GxGxG sequence canonically interacts with the carboxylic acid of the methionine unit of SAM/SAH either directly and/or via water molecule(s) (Figure 5b). Of the five human m6A MTases, only METTL5 (GCGCG) has the full GxGxG sequence while METTL16 (GTGAS) is missing the final glycine. Though METTL3 and ZCCHC4 interact with the carboxylic acid of SAM/SAH with the amino acids following β1, their sequences contain only one glycine (GRPHN and GTPRL, respectively) and the second glycine is replaced with a less conserved proline. METTL14, which has not been shown to bind to SAM/SAH, also contains only one glycine (GRDST). It is unlikely that the missing glycine is responsible for the inability of METTL14 to bind SAM/SAH, as other m6A MTases have been successful catalysts without the entire GxGxG sequence. Beyond interacting with the carboxylic acid of SAM/SAH, the GxGxG sequence sometimes interacts with the amino moiety of the methionine unit of SAM/SAH (METTL3, METTL5, and METTL16) and the 2′- and 3′-hydroxyls of the SAM/SAH ribose (METTL3; Figure 5b).

7.3 ∣. Motif II

Motif II resides in the β2-loop region (Figures 4 and 5a). The loop begins with an acidic amino acid followed by a hydrophobic amino acid (Figure 5a). The negative charge on the acidic amino acid interacts with both the 2′- and 3′-hydroxyls of the SAM/SAH ribose while the hydrophobic amino acid interacts with the SAM/SAH adenine base via van der Waals interactions (Figure 5b). The four catalytically active human m6A MTases exhibit the canonical interactions: METTL5 and ZCCHC4 interact with the ribose hydroxyls using an aspartate (D81 and D225, respectively) and METTL16 uses a glutamate (E133) (Figure 5b). In contrast, METTL3 uses a glutamine (Q550) at this position, and its side chain still seems capable of forming hydrogen bonds with both the 2′- and 3′-hydroxyls (2.2 and 3.2 Å, respectively; Protein Data Bank identifier [PDB ID]: 5IL1) (Figure 5b). METTL14 contains a threonine (T363), which may partially explain why METTL14 does not bind to SAM/SAH (Figure 5a). The hydrophobic amino acid is isoleucine in METTL5 and ZCCHC4 (I82 and I226, respectively), valine in METTL16 (V134), and leucine in METTL3 and METTL14 (L551 and L364, respectively). However, in METTL3, the amino acid that has van der Waals interactions with the SAM/SAH adenosine base is not the leucine (L551) but rather an asparagine (N549) that precedes the glutamine (Q550; Figure 5b).

7.4 ∣. Motif III

Motif III corresponds to β3 and includes two conserved amino acids (Figures 4 and 5a). The first is an amino acid with a side chain oxygen that forms a hydrogen bond with one of the N6 hydrogens of the SAM/SAH adenosine (Figure 5b). METTL3, METTL14, and METTL5 contain an aspartate (D377, D173, and D108, respectively), METTL16 contains a threonine (T164, which interacts via main chain oxygen rather than side chain), and ZCCHC4 contains an asparagine (N243; Figure 5). Immediately after the oxygen-containing amino acid is a hydrophobic amino acid whose side chain interacts with the hydrophobic core of the Rossmann fold while the backbone nitrogen forms a hydrogen bond to SAM/SAH at the N1 position (Figure 5). METTL3 and METTL14 contain an isoleucine (I378 and I174, respectively), METTL16 contains a leucine (L165), METTL5 contains a valine (V109), and ZCCHC4 contains a methionine (M244, Figure 5). Most of the backbone nitrogens directly interact with N1, but the X-ray crystal structure of METTL16 bound to SAH reveals that this interaction is indirect through a water molecule (Figure 5b). Though sequence alignment alone cannot identify Motif III, the amino acid identities and functions are supported by the crystal structures of the individual MTases, except for METTL14, which has not been shown to bind to SAM/SAH (Figure 5).

7.5 ∣. Motif IV

Class I SAM-dependent MTases have a highly conserved catalytic core of (D/N/S)PP(F/W/Y) following β4; this catalytic core constitutes Motif IV (Figures 4 and 5a). It is thought the side-chain oxygen in the D/N/S forms a hydrogen bond with one of the amino hydrogens at the N6 position of the methyl-accepting adenosine while the backbone carbonyl between the two prolines forms a hydrogen bond with the other amino hydrogen (Figures 2a and 5b). These hydrogen bonds are directly contributing to the catalysis of methyl transfer, as one of these interactions results in the deprotonation of the adenosine acceptor at the N6 position, which is necessary for methyl transfer to occur. To date, the amino acid responsible for the deprotonation of the N6-amino is unknown and requires further investigation; however, it is thought to be the D/N/S residue (Bheemanaik et al., 2006). Based on X-ray crystal structures of METTL16 bound to MAT2A hairpin RNAs (PDB ID: 6DU4, 6DU5), the backbone carbonyl between the two prolines is the only electrophile close enough to deprotonate N6, while the asparagine interacts with N6 indirectly through a water molecule. However, the METTL16•MAT2A crystal structures lack SAM/SAH in the active site; therefore, the active site of the ternary complex may still undergo further remodeling that would move the asparagine into position to form a hydrogen bond with the N6 acceptor. The aromatic F/W/Y amino acid interacts with the methyl-accepting adenosine via parallel π–π interactions (i.e., π-stacking) to properly orient the adenosine for methyl transfer. Although all known human m6A MTases contain the central PP, they utilize different amino acids as the N6 hydrogen bond acceptor and aromatic amino acids for stacking (Figure 5a). ZCCHC4 and METTL3 use aspartate (D276 and D395, respectively) while METTL16 and METTL5 use asparagine (N184 and N126, respectively). For the aromatic amino acid, METTL16, METTL5, and ZCCHC4 use a phenylalanine (F187, F129, and F279, respectively) and METTL3 uses a tryptophan (W398). METTL14 has an inactive EPPL catalytic core. Though a glutamate (E192) has hydrogen bond acceptor capabilities, no known catalytically active MTase contains a glutamate at this position in motif IV. Furthermore, leucine (L195) is unable to participate in π–π interactions. Overall, even if METTL14 does bind to SAM, it is unlikely that it has any catalytic activity. In addition to catalysis, the crystal structures of the human m6A MTases in complex with SAM/SAH reveal that the D/N side chain forms a hydrogen bond with the amino moiety of the methionine unit of SAM/SAH in the absence of its RNA substrate (Figure 5b). To date, METTL16 is the only m6A MTase with a solved 3D structure in complex with an RNA substrate and these structures lack SAM/SAH in the binding pocket. The field would benefit from 3D structures of the other m6A MTases in their binary (E•RNA) and ternary (E•RNA•SAM) states.

7.6 ∣. Motifs V–VIII

Although Motifs I–IV are conserved in their active site interactions, Motifs V–VIII are much less conserved among all m6A MTases. This lack of sequence conservation is perhaps expected because these motifs, though structurally similar, have different functions among the MTases. Structurally, Motifs V–VIII include αD, β5, αE, β6, β7, and all loops linking these secondary structures (Figure 4). Characterized functions of these motifs include dimerization interface in METTL3/METTL14 and METTL5/TRMT112, interactions with other intramolecular features of the protein such as the C-terminal zinc-finger domains in ZCCHC4, and RNA recognition by METTL16. These unique structural features are discussed in more detail (see Section 7.8).

7.7 ∣. Motif X

Motif X spans αZ and the loop preceding it (Figure 4). Motif X forms the last side of the SAM/SAH binding pocket by forming a hydrogen bond to the methionine carboxyl- and/or amino-groups in SAM/SAH (Figure 5b). Each m6A MTase has a unique mechanism for interacting with the aforementioned moieties of SAM/SAH, therefore, little sequence conservation among m6A MTases is observed in motif X (Figure 5). To interact with the methionine carboxyl, METTL5 and ZCCHC4 use the backbone amine of T31 and F175, respectively, as a hydrogen bond donor; METTL3 indirectly uses the backbone amine of K510 and the side chain of S508 through a water molecule; and METTL16 indirectly uses the side chain of R82 and the backbone carbonyl of L75 via two water molecules (Figure 5b). Although METTL3 and METTL16 make no interactions with the methionine amino-group within Motif X, METTL5 uses the side chain of T31 as the hydrogen bond acceptor and ZCCHC4 uses the backbone carbonyl of Y173 (Figure 5b). Because of the variations of Motif X within the catalytically active MTases, it is difficult to conclude whether this area of METTL14 is responsible for its inability to bind SAM/SAH (Figure 5a).

7.8 ∣. Structural features unique to each MTase

The MTase domains of the METTL3/METTL14 complex were crystallized and published by multiple labs within 4 months of each other (Table 3; Šledź & Jinek, 2016; P. Wang, Feng, et al., 2016; X. Wang, Feng, et al., 2016). All 3D structures showed a “pseudo-symmetric heterodimer,” where the two Rossmann folds interact through an extensive dimerization interface (over 2000 Å2) from αC to αD (Figure 4). Though it was originally thought that METTL14 was the active MTase in the METTL3/METTL14 complex (J. Liu et al., 2014), structural data and further biochemical analysis established METTL3 as the active MTase (Šledź & Jinek, 2016; P. Wang, Feng, et al., 2016; X. Wang, Feng, et al., 2016). Not only does METTL3 have the conserved catalytic core observed in MTases (i.e., Motif IV) while METTL14 does not, only the METTL3 MTase domain contains electron density for SAM/SAH (Šledź & Jinek, 2016; P. Wang, Feng, et al., 2016; X. Wang, Feng, et al., 2016). Furthermore, modeling a nucleic acid substrate into these catalytic cores reveals that METTL14 lacks the positively charged surface necessary to bind the negatively charged phosphate backbone (Doxtader et al., 2018; Šledź & Jinek, 2016; P. Wang, Feng, et al., 2016). However, biochemical assays show that METTL3 alone is insufficient for methylation, as METTL14 is necessary for METTL3 stability and, therefore, catalysis (Šledź & Jinek, 2016; P. Wang, Feng, et al., 2016). The minimized MTase domains of the METTL3/METTL14 complex used for crystallization of the complex are also inactive, though the addition of the METTL3 zinc-finger domains, which are adjacent to the N-terminus of the MTase domain (Figure 4), restores activity to that of the wildtype METTL3/METTL14 complex (J. Huang et al., 2019; P. Wang, Feng, et al., 2016). Together, the two METTL3 zinc-finger domains, whose structure was solved using NMR (Table 3), are necessary and sufficient to bind GGACU-containing RNA (J. Huang et al., 2019). How the zinc-finger and MTase domains work together to specifically bind and methylate RNA is an exciting structural question that remains unanswered. Recently, a 4.4 Å-resolution structure of the full-length METTL3/METTL14 complex, along with other proteins (i.e., HAKAI, WTAP, VIRMA, and ZC3H13) in the holoenzyme, was solved using cryogenic electron microscopy (cryo-EM; S. Su, Li, et al., 2022). Using this model, protein–protein crosslinking mass spectrometry, and AI-structure characterization, the METTL3/METTL14 complex appears to interact with the rest of the holoenzyme through the N-terminal region of METTL3, including the zinc-finger domains, and the N-terminal region of METTL14 (S. Su, Li, et al., 2022; Yan et al., 2022). These studies reveal the power of integrative structural biology, which is likely to uncover more exciting structure–function relationships for all human m6A RNA MTases.

In METTL16, the N-terminal domain (amino acids 1–288) is conserved from E. coli to humans but the C-terminal domain (amino acids 307–562) is conserved only through vertebrates at the sequence level (Sergiev et al., 2008; Figure 4 and Table 1). The N-terminal domain consists of the highly conserved Rossmann fold, a structured N-terminal region that is needed for RNA binding (amino acids 1–77) (Doxtader et al., 2018; Mendel et al., 2018), a 310 helix preceding β1, and a structured loop following β4 (Figure 4). The C-terminal domain consists of VCR1 and VCR2 separated by a predicted disordered region (Figure 4). The two VCRs fold into a β-sheet with five β-strands (15432), which is structurally and functionally similar to the kinase-associated 1 (KA1) domain of the U6 snRNA-specific terminal uridylyl transferase 1 (Aoyama et al., 2020; Figure 4). Interestingly, AlphaFold indicates that the C-terminal domain of the C. elegans homolog METT-10 folds into a KA1-like domain, suggesting the structure and function of the C-terminal domain of METTL16 is more widely conserved than the sequence (Ju et al., 2023). The VCRs enhance the MTase activity and binding affinity of METTL16 to U6 snRNA by up to ~100-fold compared to the MTase domain alone (Aoyama et al., 2020; Breger & Brown, 2023), while the VCRs are not required for methylation of the MAT2A mRNA hairpins but do improve RNA binding by ~30-fold (Aoyama et al., 2020; Pendleton et al., 2017). Unfortunately, there is no structure of the METTL16 MTase domain and VCRs together. A structure of the full-length METTL16 protein, both apo and RNA bound, will likely lead to a better understanding of how the VCRs differentially contribute to catalysis of different RNA substrates. Unlike the other human m6A MTases, a structure of METTL16 has been solved in complex with an RNA substrate: the minimal MAT2A mRNA hairpin (Doxtader et al., 2018; Table 3). Interestingly, the MTase domain of the METTL16•RNA bound complex is extraordinarily similar to that of the apo and SAH-bound structures (root mean square deviation of α-carbons is ~0.3 Å among the structures). However, the major difference among these structures is that the apo and SAH-bound structures lack electron density for amino acids (188–214) between β4 and αD while the RNA-bound structure shows this region interacting with the RNA, suggesting a conformational change or stabilization (Figure 4). The RNA-bound structure also revealed that K163, a residue located in the K-loop following β3, blocks the SAM binding site in the presence of RNA but is solvent exposed in the absence of RNA (Doxtader et al., 2018; Ruszkowska et al., 2018). The METTL16 K163A mutant increases the binding affinity to SAM and increases METTL16 catalytic activity on MAT2A hp1 by approximately sixfold (Doxtader et al., 2018). METTL16 may have evolved to prefer the weaker binding and methylation activity due to its role in SAM homeostasis. It will be interesting to see how these dynamic regions of METTL16 as well as the C-terminal domain interacts with its other RNA-binding partners.

Composed of only an MTase domain, METTL5 is the simplest validated human m6A MTase (Figure 4). Within its MTase domain, METTL5 has the well-conserved Rossmann fold, a short (25 amino acid) N-terminal structure containing two α-helices, and a structured loop following β4 (Figure 4). However, knowing the sequence and structure of METTL5 alone is not sufficient to understand its activity, as it was shown to be correctly folded and functional only when co-purified with TRMT112, which interacts via a highly hydrophobic surface of METTL5 (1160 Å2) (Van Tran et al., 2019; Figure 4). Though TRMT112 is not known to be necessary for the activity of any other human m6A MTase, TRMT112 is necessary for the functioning of at least seven other human RNA and protein MTases (Brūmele et al., 2021), including metastasis-related methyltransferase 1 (originally WBSCR22 due to it being deleted in Williams-Beuren syndrome), which methylates rRNA G1575 to N7-methylguanosine (Zorbas et al., 2015); AlkB homolog 8, tRNA methyltransferase, which methylates tRNA at the wobble position to 5-methylcarboxymethyluridine (D. Fu et al., 2010); and lysine methyltransferase 9 (originally N6AMT1), which methylates histone H4 lysine 12 (Figaro et al., 2008; W. Li et al., 2019; Metzger et al., 2019).

ZCCHC4 is an evolutionarily conserved MTase within multicellular organisms (H. Ma et al., 2019; Table 1). The Rossmann fold of ZCCHC4 is centralized within three different canonical zinc-finger domains, which collectively bind to six zinc ions (Ren et al., 2019; Figure 4). The two N-terminal zinc-finger domains, GRF- and C2H2-type zinc-fingers (named after the conserved residues), are a part of the putative RNA-binding site (Ren et al., 2019). The C-terminus contains a CCHC-type zinc-finger domain, which is found in proteins that bind to single-stranded nucleic acids. All three zinc-finger domains, in addition to the MTase domain, are necessary for ZCCHC4 to methylate its 28S rRNA substrate. Within the Rossmann fold is a unique auto-inhibitory loop, known as the regulatory loop, that interacts with the cofactor loop so that apo ZCCHC4 adopts a fully closed, inactive conformation (Ren et al., 2019). Although the trigger to activate ZCCHC4 is unknown, Ren et al. propose conformational changes induced by RNA and/or SAM binding. It will be interesting to see how the zinc-finger domains and regulatory loop interact with the RNA to aid in the methylation of the 28S rRNA.

Overall, though some 3D structures are available for each of the human m6A RNA MTases, structural models for all steps in the kinetic pathway (Figure 2b) have not been reported for each of the MTases. Both METTL5 and ZCCHC4 lack structures without SAM/SAH bound in the active site, though we suspect there will be little difference between the apo and SAM/SAH bound structures, as both METTL3 and METTL16 show little differences between apo and SAM/SAH-bound structures. Conversely, the METTL16•MAT2A RNA hp complex does suggest conformational changes between apo and RNA-bound structures. Therefore, the MTases in complex with their RNA substrates is one knowledge gap that needs to be filled. Furthermore, high-resolution structural models of the full-length METTL3/METTL14 complex and of full-length METTL16 are not available. Perhaps the use of cryo-EM, which does not depend on crystallization and is able to provide atomic resolution of large macromolecules, will lead to a better understanding of the functional interplay between the MTase domain, other regions of the MTases, and protein–protein interactions. In addition, AlphaFold2 and the like may provide early insights (Yan et al., 2022). We look forward to learning more about the structures of MTases and the MTase•RNA complexes.

8 ∣. HYPERMODIFIED m6A-CONTAINING RIBONUCLEOTIDE VARIANTS IN HUMANS

Besides m6A, there are other known naturally occurring adenosine modifications that contain an N6-methyl group, including m6Am, N6,N6-dimethyladenosine (m6,6A), N6,N6,2′-O-trimethyladenosine, and 2-methylthio-N6-methyladenosine. To date, only m6Am and m6,6A have been identified in humans. These two hypermodifications will be discussed along with their related MTases, providing a brief comparative analysis with the five m6A MTases.

There are two known human SAM-dependent MTases that catalyze the methyl transfer to the N6 position of 2’-O-methyladenosine (Am) to form m6Am: METTL4 and PCIF1, also known as cap-specific adenosine methyltransferase. METTL4 is conserved from yeast to humans. The C. elegans and mouse homologs of METTL4 methylate DNA to make 6mA marks; however, human METTL4 has shown no evidence for DNA methylation, but instead methylates the N6 position of an internal Am (but not A) in U2 snRNA (H. Chen, Gu, et al., 2020; Goh et al., 2020). Early evidence suggests that the m6Am mark in U2 snRNA plays a role in differential splicing of select mRNAs, though further studies are needed to understand the mechanism (H. Chen, Gu, et al., 2020; Goh et al., 2020). Sequence conservation and AlphaFold predict human METTL4 to adopt a Class Ig Rossmann fold with the conserved DPPW catalytic core, similar to METTL3 (Jumper et al., 2021). Using X-ray crystallography, a 3D structure of Arabidopsis METTL4•Am-containing RNA•SAH complex has been solved, revealing that a Phe residue conserved in METTL4 creates a more hydrophobic binding pocket to favor binding a 2′-O-methyl rather than 2′-OH (Luo et al., 2022). PCIF1 catalyzes the N6-methylation of Am to m6Am at the first position after the 5’-N7-methylguanosine (m7G) cap. The m7G cap alone (also known as cap 0) is necessary for eukaryotic translation, whereas higher eukaryotes also add a N7-methylguanosine followed by a 2′-O-methylnucleotide (Cap 1) to distinguish self from non-self mRNA (Ramanathan et al., 2016). Sometimes, when Cap 1 contains an adenosine (m7G-Am), PCIF1 will methylate Am at the N6-position (m7G-m6Am; Ramanathan et al., 2016). Though the function of the m7G-m6Am cap is not fully understood, it may have an effect on translation, mRNA stability, and stress pathways (Akichika et al., 2019; Boulias et al., 2019; Sendinc et al., 2019; Sikorski et al., 2020). Like most of the human m6A MTases, PCIF1 adopts a Class 1a Rossmann fold and has the conserved NPPF catalytic core (Akichika et al., 2019; PDB ID: 6IRV, 6IRW). The crystal structure also reveals that PCIF1 has a small, positively charged groove that can accommodate the 5′ cap, which contributes to its substrate specificity (Akichika et al., 2019). Readers can learn more about m6Am from this recent review (Cesaro et al., 2023).

There are two known human m6,6A MTases: dimethyladenosine transferase 1 (DIMT1) and the mitochondrial transcription factor B1 (TFB1M, also mtTFB1). DIMT1 is thought to place the m6,6A mark on two neighboring adenosines, A1850 and A1851, in the 18S rRNA, analogous to the yeast homolog Dim1 (Zorbas et al., 2015). DIMT1 has catalytic activity in vitro on a short, synthetic 12-nt hairpin having the same sequence and secondary structure as the predicted m6,6A sites in the 18S rRNA (H. Shen et al., 2020). However, the methylation order of the two adenosines and the binding order of the substrates (RNA and SAM) remains unknown. The MTase activity of DIMT1 is necessary for efficient protein translation (H. Shen et al., 2020). In addition to MTase activity, DIMT1 is a critical scaffold for 40S rRNA assembly, as heterologous expression of a catalytically inactive mutant in a DIMT1+/− heterozygous knockdown cell line (HEK293T) restores the abundance of 40S rRNA to normal levels (H. Shen et al., 2020). Three crystal structures of full-length human DIMT1 (PDB ID: 1ZQ9, 6W6C, 6W6F) have been solved (H. Shen et al., 2020) and show that DIMT1 also contains a Class Ia Rossmann fold but with a noncanonical NLPY catalytic core. TFB1M also catalyzes m6,6A marks in rRNA, specifically on A936 and A937 in helix 45 of mitochondrial 12S rRNA (McCulloch et al., 2002; Metodiev et al., 2009; Seidel-Rogol et al., 2003). Knockout of Tfb1m in mice leads to a destabilization of 12S rRNA and eliminates translation in the mitochondria (Metodiev et al., 2009). Two crystal structures of TFB1M have been solved (PDB ID: 6AAX, 6AJK), showing a Class Ia Rossmann fold with a noncanonical NLPF (141–144) catalytic core in complex with a minimized hairpin substrate (X. Liu, Shen, et al., 2019). More recently, cryo-EM structures showing the formation of the mature mitochondrial small ribosomal subunit shows TFB1M binding to its rRNA substrate (PDB ID: 8CSP, 8CSQ, 8CSR, 8CSU) and the dimethylated product without TFB1M bound (PDB ID: 8CSS, 8CST; Burnside et al., 2023). Both the crystallographic and cryo-EM models show the second adenosine in the MTase active site but not any methylation when bound to TFB1M, so the adenosine methylation order is still not fully understood. It will be interesting to discover how these m6,6A MTases can methylate two adjacent sites with such high specificity.

9 ∣. CONCLUSION

Within the past 7 years, three more proteins (i.e., METTL16, METTL5, and ZCCHC4) have been identified as human m6A MTases, yet many exciting discoveries remain such as a deeper understanding of their biological functions, their roles in rare and neglected diseases, catalytic mechanisms, 3D structures of binary and ternary complexes and perhaps the identification of more m6A RNA MTases. Even E. coli has a large and dynamic methylome with m6A marks installed in mRNA by an unknown m6A MTase (X. Deng et al., 2015). With about 200 SAM-dependent MTases encoded in the human genome, one wonders: are there other MTases that methylate adenosine to produce m6A (Petrossian & Clarke, 2011)? A new twist on RNA methylation is the existence of ribozymes that catalyze methyl transfer reactions, a finding that begs the question: is there a naturally occurring counterpart (Scheitl et al., 2020)? There is a SAM-dependent m7G MTase ribozyme that was created by in vitro selection and similar ribozymes appear to exist in organisms from bacteria to mammals (H. Jiang, Gao, et al., 2021). Recently the first natural RNA MTase was discovered and aptly named methyltransferase ribozyme 1, which utilizes O6-methylguanosine as methyl donor to create N1-methyladenosine as well as a riboswitch that can self-methylate to install 3-methylcytidine using prequeuosine (preQ1) as a methyl donor (J. Deng, Wilson, et al., 2022; Flemmich et al., 2021; Hiller & Strobel, 2022; Scheitl et al., 2022). These recent findings may hint at a number of new RNA ribozymes waiting to be discovered and elucidated.

Among the known human m6A MTases, several key unknowns remain, such as their druggability, the structural basis of their substrate specificity, how m6A RNA MTases impact phase separation (D. Han, Longhini, et al., 2022; Ries et al., 2019), and the physiological relevance of m6A RNA MTases methylating DNA to form 6mA marks. Substrate specificity extends beyond the MTase domain; therefore, more studies will be required to elucidate the unique structural elements that diversify and differentiate these MTases, such as the disordered region in METTL16 or protein-binding partners. Of course, the structure and function of m6A writers must be understood in the context of an m6A methylome that is also regulated by m6A readers and possibly erasers. Therefore, having an improved quantitative view of the m6A methylome is paramount, emphasizing the need for direct, transcriptome-wide readout of m6A marks at single-nucleotide resolution (Alfonzo et al., 2021). By filling-in the aforementioned knowledge gaps, m6A biology and its related MTases will penetrate into new realms of cellular importance and human health.

ACKNOWLEDGMENT

As mentioned in Section 2.3, results were presented from the TCGA Research Network: https://www.cancer.gov/tcga.

FUNDING INFORMATION

This study was supported by startup funds from the University of Notre Dame, the National Institutes of Health grant R35GM133696, and the Clare Boothe Luce Program of the Henry Luce Foundation. Kurtis Breger and Charlotte N. Kunkler are fellows of the Chemistry–Biochemistry–Biology Interface (CBBI) Program at the University of Notre Dame, supported by training grant T32GM075762 from the National Institute of General Medical Sciences. Charlotte N. Kunkler was also supported by a National Institute of General Medical Sciences National Research Service Award Predoctoral Fellowship grant F31GM136163. The content is solely the responsibility of the authors and does not necessarily represent the official views of the National Institute of General Medical Sciences nor the National Institutes of Health.

Funding information

Henry Luce Foundation, Clare Boothe Luce Professorship; National Institute of General Medical Sciences, Grant/Award Numbers: F31GM136163, R35GM133696, T32GM075762; University of Notre Dame, Startup Funds

Abbreviations:

5mC

5-methylcytidine

6mA

N6-methyladenosine (for DNA)

A-431

epidermoid carcinoma cell line

ALKBH5

AlkB homolog 5, RNA demethylase

Am

2′-O-methyladenosine

AML

acute myeloid leukemia

APOBEC1

apolipoprotein B mRNA editing enzyme, catalytic polypeptide 1

B16

mouse melanoma cell line

Cas9

CRISPR-associated protein 9

CFIm

cleavage factor Im

CRISPR

clustered regularly interspaced short palindromic repeats

cryo-EM

cryogenic electron microscopy

DART-seq

deamination adjacent to RNA modification targets

dCas

catalytically inactive mutant of Cas

DIMT1

dimethyladenosine transferase 1

dRNA-seq

direct RNA sequencing

EBV

Epstein–Barr virus

eEF

eukaryotic translation elongation factor

eIF

eukaryotic translation initiation factor

ERK

extracellular signal-regulated kinases

EV71

enterovirus 71

FTO

fat mass and obesity-associated protein

gRNA

guide RNA

HAKAI

E3 ubiquitin-protein ligase Hakai

HAP1

near-haploid human cell line

HBV/HCV

hepatitis B and C viruses

HCT116

human colon cancer cell line

HEK293T

human embryonic kidney 293 cells with mutant SV40 large T antigen

HepG2

human liver cancer cell line

HIV-1

human immunodeficiency virus 1

hMPV

human metapneumovirus

hp

hairpin

HSV-1

herpes simplex virus-1

IC50

half-maximal inhibitory concentration

J1 ES

mouse embryonic cell line

KA1

Kinase-associated 1 domain

k cat

turnover number

K D

equilibrium dissociation constant

K M

Michaelis–Menten constant

KSHV

Kaposi's sarcoma-associated herpes virus

LARP7

La ribonucleoprotein 7, transcriptional regulator

lncRNA

long noncoding RNA

m6,6A

N6,N6-dimethyladenosine

m6A

N6-methyladenosine (for RNA)

m6ACE-seq

m6A-crosslinking-exonuclease-sequencing

m6Am

N6,2′-O-dimethyladenosine

M6AREG

m6A-centered regulations of disease development and drug response database

m7G

N7-methylguanosine

m7G-m6Am

N7-methylguanosine followed by N6,2-O-dimethyladenosine

MALAT1

metastasis-associated lung adenocarcinoma transcript 1

MAT2A

methionine adenosyl transferase 2A

MazF

endoribonuclease toxin MazF

meCLICK-seq

methylation CLICK degradation-sequencing

MEF

mouse embryonic fibroblasts

MePCE

methylphosphate capping enzyme

MeRIP-seq

methylated RNA immunoprecipitation-sequencing

Merm1

metastasis-related methyltransferase 1

mESC

mouse embryonic stem cell

METTL

methyltransferase-like protein

miCLIP

methylation (or m6A)-individual-nucleotide resolution crosslinking and immunoprecipitation

miRNA

micro RNA

MjDim1

Methanocaldococcus jannaschii dimethyltransferase 1

MOLM-13

human acute myeloid leukemia cell line

MRE11

MRE11 homolog, double-strand break repair nuclease

mRNA

messenger RNA

MTase

methyltransferase

ncRNA

noncoding RNA

NIH 3T3

mouse fibroblast cell line

NS

nonstructural protein

nt

nucleotide

PCIF1

phosphorylated CTD interacting factor 1

PDB ID

Protein Data Bank identifier

PEDV

porcine epidemic diarrhea virus

PYL

pyrabactin resistance like

RBM

RNA-binding motif

RdRp

RNA-dependent RNA polymerase

RNP

ribonucleoprotein

rRNA

ribosomal RNA

RSV

Rous sarcoma virus

SAH/AdoHcy

S-adenosyl-L-homocysteine

SAM/AdoMet

S-adenosyl-L-methionine

SARS-CoV-2

severe acute respiratory syndrome coronavirus 2

STM2457

METTL3 inhibitor

SV40

Simian virus 40

TK6

human lymphoblast cell line

TRMT112

tRNA methyltransferase activator subunit 11-2

UTR

untranslated region

UZH1a

METTL3 inhibitor

U2OS

human osteosarcoma cell line

VCR

vertebrate conserved region

VIRMA/KIAA1429

vir-like m6A methyltransferase associated protein

WTAP

Wilms' tumor 1-associating protein

XRN1

5′-3′ exoribonuclease 1

YTH

YT521-B homology domain

YTHDC

YTH domain-containing protein

YTHDF

YTH domain-containing family protein

ZC3H13

zinc-finger CCCH domain-containing protein 13

ZCCHC4

zinc-finger CCHC-domain-containing protein 4

ZIKV

Zika virus

ZnF

zinc-finger domain

Footnotes

CONFLICT OF INTEREST STATEMENT

The authors have declared no conflicts of interest for this article.

RELATED WIREs ARTICLES

RNA methyltransferase METTL16: Targets and function

DATA AVAILABILITY STATEMENT

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

REFERENCES

  1. Akichika S, Hirano S, Shichino Y, Suzuki TT, Nishimasu H, Ishitani R, Sugita A, Hirose Y, Iwasaki S, Nureki O, & Suzuki TT (2019). Cap-specific terminal N6-methylation of RNA by an RNA polymerase II-associated methyltransferase. Science, 363(6423), eaav0080. 10.1126/science.aav0080 [DOI] [PubMed] [Google Scholar]
  2. Alarcón CR, Lee H, Goodarzi H, Halberg N, & Tavazoie SF (2015). N6-methyladenosine marks primary microRNAs for processing. Nature, 519(7544), 482–485. 10.1038/nature14281 [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. Alfonzo JD, Brown JA, Byers PH, Cheung VG, Maraia RJ, & Ross RL (2021). A call for direct sequencing of full-length RNAs to identify all modifications. Nature Genetics, 53(8), 1113–1116. 10.1038/s41588-021-00903-1 [DOI] [PubMed] [Google Scholar]
  4. Aoyama T, Yamashita S, & Tomita K (2020). Mechanistic insights into m6A modification of U6 snRNA by human METTL16. Nucleic Acids Research, 48(9), 5157–5168. 10.1093/nar/gkaa227 [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Arcidiacono OA, Krejčí J, & Bártová E (2020). The distinct function and localization of METTL3/METTL14 and METTL16 enzymes in cardiomyocytes. International Journal of Molecular Sciences, 21(21), 1–19. 10.3390/ijms21218139 [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Bader JP, Brown NR, Chiang PK, & Cantoni GL (1978). 3-Deazaadenosine, an inhibitor of adenosylhomocysteine hydrolase, inhibits reproduction of Rous sarcoma virus and transformation of chick embryo cells. Virology, 89(2), 494–505. 10.1016/0042-6822(78)90191-5 [DOI] [PubMed] [Google Scholar]
  7. Baquero-Perez B, Antanaviciute A, Yonchev ID, Carr IM, Wilson SA, & Whitehouse A (2019). The tudor SND1 protein is an m6A RNA reader essential for replication of Kaposi's sarcoma-associated herpesvirus. eLife, 8, e47261. 10.7554/eLife.47261 [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Barbieri I, Tzelepis K, Pandolfini L, Shi J, Millán-Zambrano G, Robson SC, Aspris D, Migliori V, Bannister AJ, Han N, De Braekeleer E, Ponstingl H, Hendrick A, Vakoc CR, Vassiliou GS, & Kouzarides T (2017). Promoter-bound METTL3 maintains myeloid leukaemia by m6A-dependent translation control. Nature, 552(7683), 126–131. 10.1038/nature24678 [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Batista PJ, Molinie B, Wang J, Qu K, Zhang J, Li L, Bouley DM, Lujan E, Haddad B, Daneshvar K, Carter AC, Flynn RA, Zhou C, Lim KS, Dedon P, Wernig M, Mullen AC, Xing Y, Giallourakis CC, & Chang HY (2014). m6A RNA modification controls cell fate transition in mammalian embryonic stem cells. Cell Stem Cell, 15(6), 707–719. 10.1016/j.stem.2014.09.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Baumgarten S, Bryant JM, Sinha A, Reyser T, Preiser PR, Dedon PC, & Scherf A (2019). Transcriptome-wide dynamics of extensive m6A mRNA methylation during Plasmodium falciparum blood-stage development. Nature Microbiology, 4(12), 2246–2259. 10.1038/s41564-019-0521-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bayoumi M, & Munir M (2021). Evolutionary conservation of the DRACH signatures of potential N6-methyladenosine (m6A) sites among influenza A viruses. Scientific Reports, 11(1), 1–12. 10.1038/s41598-021-84007-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Bheemanaik S, Reddy YVR, & Rao DN (2006). Structure, function and mechanism of exocyclic DNA methyltransferases. Biochemical Journal, 399(2), 177–190. 10.1042/BJ20060854 [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Boccaletto P, MacHnicka MA, Purta E, Pitkowski P, Baginski B, Wirecki TK, De Crécy-Lagard V, Ross R, Limbach PA, Kotter A, Helm M, & Bujnicki JM (2018). MODOMICS: A database of RNA modification pathways. 2017 update. Nucleic Acids Research, 46(D1), D303–D307. 10.1093/nar/gkx1030 [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. 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(11), 1233–1247. [PMC free article] [PubMed] [Google Scholar]
  15. Boulias K, Toczydłowska-Socha D, Hawley BR, Liberman N, Takashima K, Zaccara S, Guez T, Vasseur JJ, Debart F, Aravind L, Jaffrey SR, & Greer EL (2019). Identification of the m6Am methyltransferase PCIF1 reveals the location and functions of m6Am in the transcriptome. Molecular Cell, 75(3), 631–643.e8. 10.1016/j.molcel.2019.06.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Breger K, & Brown JA (2023). Elucidating the kinetic mechanism of human METTL16. Biochemistry, 62(2), 494–506. 10.1021/acs.biochem.2c00601 [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Brown JA, Kinzig CG, Degregorio SJ, & Steitz JA (2016). Methyltransferase-like protein 16 binds the 3′-terminal triple helix of MALAT1 long noncoding RNA. Proceedings of the National Academy of Sciences of the United States of America, 113(49), 14013–14018. 10.1073/pnas.1614759113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Brūmele B, Mutso M, Telanne L, Õunap K, Spunde K, Abroi A, & Kurg R (2021). Human TRMT112-methyltransferase network consists of seven partners interacting with a common co-factor. International Journal of Molecular Sciences, 22(24), 1–17. 10.3390/ijms222413593 [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Bujnicki JM, Feder M, Radlinska M, & Blumenthal RM (2002). Structure prediction and phylogenetic analysis of a functionally diverse family of proteins homologous to the MT-A70 subunit of the human mRNA:m6A methyltransferase. Journal of Molecular Evolution, 55(4), 431–444. 10.1007/S00239-002-2339-8 [DOI] [PubMed] [Google Scholar]
  20. Burgess HM, Depledge DP, Thompson L, Srinivas KP, Grande RC, Vink EI, Abebe JS, Blackaby WP, Hendrick A, Albertella MR, Kouzarides T, Stapleford KA, Wilson AC, & Mohr I (2021). Targeting the m6A RNA modification pathway blocks SARS-CoV-2 and HCoV-OC43 replication. Genes and Development, 35(13), 1005–1019. 10.1101/gad.348320.121 [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Burnside CL, Harper N, & Klinge S (2023). Principles of mitoribosomal small subunit assembly in eukaryotes. Biophysical Journal, 122(3S1), 318a. 10.1016/j.bpj.2022.11.1783 [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Cantoni GL (1977). S-adenosylmethionine: Present status and future perspectives. In The biochemistry of adenosylmethionine (p. 577). Columbia University Press. [Google Scholar]
  23. Casu F, Duggan BM, & Hennig M (2013). The arginine-rich RNA-binding motif of HIV-1 Rev is intrinsically disordered and folds upon RRE binding. Biophysical Journal, 105(4), 1004–1017. 10.1016/J.BPJ.2013.07.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Catacalos C, Krohannon A, Somalraju S, Meyer KD, Janga SC, & Chakrabarti K (2022). Epitranscriptomics in parasitic protists: Role of RNA chemical modifications in posttranscriptional gene regulation. PLoS Pathogens, 18(12), e1010972. 10.1371/JOURNAL.PPAT.1010972 [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Cesaro B, Tarullo M, & Fatica A (2023). Regulation of gene expression by m6Am RNA modification. International Journal of Molecular Sciences, 24(3), 2277. 10.3390/IJMS24032277 [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Chang C, Ma G, Cheung E, & Hutchins AP (2022). A programmable system to methylate and demethylate N6-methyladenosine (m6A) on specific RNA transcripts in mammalian cells. Journal of Biological Chemistry, 298(11), 102525. 10.1016/j.jbc.2022.102525 [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Chen H, Gu L, Orellana EA, Wang Y, Guo J, Liu Q, Wang L, Shen Z, Wu H, Gregory RI, Xing Y, & Shi Y (2020). METTL4 is an snRNA m6Am methyltransferase that regulates RNA splicing. Cell Research, 30(6), 544–547. 10.1038/s41422-019-0270-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Chen J, Jin L, Wang Z, Wang L, Chen Q, Cui Y, & Liu G (2020). N6-methyladenosine regulates PEDV replication and host gene expression. Virology, 548, 59–72. 10.1016/J.VIROL.2020.06.008 [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Chen LQ, Zhang Z, Chen HX, Xi JF, Liu XH, Ma DZ, Zhong YH, Ng WH, Chen T, Mak DW, Chen Q, Chen YQ, & Luo GZ (2022). High-precision mapping reveals rare N6-deoxyadenosine methylation in the mammalian genome. Cell Discovery, 8(1), 1–11. 10.1038/s41421-022-00484-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Chen M, Wei L, Law CT, Tsang FH, Shen J, Cheng CL, Tsang LH, Ho DW, Chiu DK, Lee JM, Wong CC, Ng IO, & Wong CM (2018). RNA N6-methyladenosine methyltransferase-like 3 promotes liver cancer progression through YTHDF2-dependent posttranscriptional silencing of SOCS2. Hepatology, 67(6), 2254–2270. 10.1002/hep.29683 [DOI] [PubMed] [Google Scholar]
  31. Chen S, Kapilashrami K, Senevirathne C, Wang Z, Wang J, Linscott JA, & Luo M (2020). Substrate-differentiated transition states of Set7/9-catalyzed lysine methylation. Journal of the American Chemical Society, 141(20), 8064–8067. 10.1021/jacs.9b02553 [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Cheng J, Xu Z, Tan W, He J, Pan B, Zhang Y, & Deng Y (2023). METTL16 promotes osteosarcoma progression by downregulating VPS33B in an m6A-dependent manner. Journal of Cellular Physiology, 1–13. 10.1002/jcp.31068 [DOI] [PubMed] [Google Scholar]
  33. Choe J, Lin S, Zhang W, Liu Q, Wang L, Ramirez-Moya J, Du P, Kim W, Tang S, Sliz P, Santisteban P, George RE, Richards WG, Wong KK, Locker N, Slack FJ, & Gregory RI (2018). mRNA circularization by METTL3–eIF3h enhances translation and promotes oncogenesis. Nature, 561(7724), 556–560. 10.1038/s41586-018-0538-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Cohn WE, & Volkin E (1951). Nucleoside-5′-phosphates from ribonucleic acid. Nature, 167(4247), 483–484. 10.1038/167483a0 [DOI] [Google Scholar]
  35. Courtney DG, Kennedy EM, Dumm RE, Bogerd HP, Tsai K, Heaton NS, & Cullen BR (2017). Epitranscriptomic enhancement of influenza A virus gene expression and replication. Cell Host and Microbe, 22(3), 377–386.e5. 10.1016/j.chom.2017.08.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Covelo-Molares H, Obrdlik A, Poštulková I, Dohnálková M, Gregorová P, Ganji R, Potěšil D, Gawriyski L, Varjosalo M, & Vaňáčová Š (2021). The comprehensive interactomes of human adenosine RNA methyltransferases and demethylases reveal distinct functional and regulatory features. Nucleic Acids Research, 49(19), 10895–10910. 10.1093/nar/gkab900 [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Csepany T, Lin A, Baldick CJ, & Beemon K (1990). Sequence specificity of mRNA N6-adenosine methyltransferase. Journal of Biological Chemistry, 265(33), 20117–20122. 10.1016/s0021-9258(17)30477-5 [DOI] [PubMed] [Google Scholar]
  38. Cui Q, Shi H, Ye P, Li L, Qu Q, Sun G, Sun G, Lu Z, Huang Y, Yang CG, Riggs AD, He C, & Shi Y (2017). m6A RNA methylation regulates the self-renewal and tumorigenesis of glioblastoma stem cells. Cell Reports, 18(11), 2622–2634. 10.1016/j.celrep.2017.02.059 [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Cui X, Nilsson K, Kajitani N, & Schwartz S (2022). Overexpression of m6A-factors METTL3, ALKBH5, and YTHDC1 alters HPV16 mRNA splicing. Virus Genes, 58(2), 98–112. 10.1007/s11262-022-01889-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Cully M. (2019). Chemical inhibitors make their RNA epigenetic mark. Nature reviews. Drug discovery, 18(12), 892–894. 10.1038/d41573-019-00179-5 [DOI] [PubMed] [Google Scholar]
  41. Dai W, Li A, Yu NJ, Nguyen T, Leach RW, Wühr M, & Kleiner RE (2021). Activity-based RNA-modifying enzyme probing reveals DUS3L-mediated dihydrouridylation. Nature Chemical Biology, 17(11), 1178–1187. 10.1038/s41589-021-00874-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Darnell RR, Shengdong KE, & Darnell JE (2018). Pre-mRNA processing includes N6 methylation of adenosine residues that are retained in mRNA exons and the fallacy of “RNA epigenetics”. RNA, 24(3), 262–267. 10.1261/rna.065219.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. De Clercq E, & Montgomery JA (1983). Broad-spectrum antiviral activity of the carbocyclic analog of 3-deazaadenosine. Antiviral Research, 3(1), 17–24. 10.1016/0166-3542(83)90011-6 [DOI] [PubMed] [Google Scholar]
  44. De Jesus DF, Zhang Z, Kahraman S, Brown NK, Chen M, Hu J, Gupta MK, He C, & Kulkarni RN (2019). m6A mRNA methylation regulates human β-cell biology in physiological states and in type 2 diabetes. Nature Metabolism, 1(8), 765–774. 10.1038/s42255-019-0089-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Deng J, Wilson TJ, Wang J, Peng X, Li M, Lin X, Liao W, Lilley DMJ, & Huang L (2022). Structure and mechanism of a methyltransferase ribozyme. Nature Chemical Biology, 18(5), 556–564. 10.1038/s41589-022-00982-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Deng LJ, Deng WQ, Fan SR, Chen MF, Qi M, Lyu WY, Qi Q, Tiwari AK, Chen JX, Zhang DM, & Chen ZS (2022). m6A modification: Recent advances, anticancer targeted drug discovery and beyond. Molecular Cancer, 21(1), 1–21. 10.1186/s12943-022-01510-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Deng X, Chen K, Luo GZ, Weng X, Ji Q, Zhou T, & He C (2015). Widespread occurrence of N6-methyladenosine in bacterial mRNA. Nucleic Acids Research, 43(13), 6557–6567. 10.1093/nar/gkv596 [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Desrosiers R, Friderici K, & Rottman F (1974). Identification of methylated nucleosides in messenger RNA from Novikoff hepatoma cells. Proceedings of the National Academy of Sciences of the United States of America, 71(10), 3971–3975. 10.1073/pnas.71.10.3971 [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Dominissini D, Moshitch-Moshkovitz S, Schwartz S, Salmon-Divon M, Ungar L, Osenberg S, Cesarkas K, Jacob-Hirsch J, Amariglio N, Kupiec M, Sorek R, & Rechavi G (2012). Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature, 485(7397), 201–206. 10.1038/nature11112 [DOI] [PubMed] [Google Scholar]
  50. Dominissini D, Nachtergaele S, Moshitch-Moshkovitz S, Peer E, Kol N, Ben-Haim MS, Dai Q, Di Segni A, Salmon-Divon M, Clark WC, Zheng G, Pan T, Solomon O, Eyal E, Hershkovitz V, Han D, Doré LC, Amariglio N, Rechavi G, & He C (2016). The dynamic N1-methyladenosine methylome in eukaryotic messenger RNA. Nature, 530(7591), 441–446. 10.1038/nature16998 [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Dorn LE, Lasman L, Chen J, Xu X, Hund TJ, Medvedovic M, Hanna JH, Van Berlo JH, & Accornero F (2019). The N6-methyladenosine mRNA methylase METTL3 controls cardiac homeostasis and hypertrophy. Circulation, 139(4), 533–545. 10.1161/CIRCULATIONAHA.118.036146 [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Dorsett M, & Schedl T (2009). A role for dynein in the inhibition of germ cell proliferative fate. Molecular and Cellular Biology, 29(22), 6128–6139. 10.1128/mcb.00815-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Douvlataniotis K, Bensberg M, Lentini A, Gylemo B, & Nestor CE (2020). No evidence for DNA N6-methyladenine in mammals. Science Advances, 6(12), eaay3335. 10.1126/sciadv.aay3335 [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Doxtader KA, Wang P, Scarborough AM, Seo D, Conrad NK, Nam Y, Brown JA, Bulkley D, Wang J, Valenstein ML, Yario TA, Steitz TA, & Steitz JA (2018). Structural basis for regulation of METTL16, an S-adenosylmethionine homeostasis factor. Molecular Cell, 71(6), 1001–1011.e4. 10.1016/j.molcel.2018.07.025 [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Du H, Zhao Y, He J, Zhang Y, Xi H, Liu M, Ma J, & Wu L (2016). YTHDF2 destabilizes m 6 A-containing RNA through direct recruitment of the CCR4-NOT deadenylase complex. Nature Communications, 7(1), 1–11. 10.1038/ncomms12626 [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Du K, Zhang Z, Zeng Z, Tang J, Lee T, & Sun T (2021). Distinct roles of Fto and Mettl3 in controlling development of the cerebral cortex through transcriptional and translational regulations. Cell Death and Disease, 12(7), 1–12. 10.1038/s41419-021-03992-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Du Q, Wang Z, & Schramm VL (2016). Human DNMT1 transition state structure. Proceedings of the National Academy of Sciences of the United States of America, 113(11), 2916–2921. 10.1073/pnas.1522491113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Durbin AF, Wang C, Marcotrigiano J, & Gehrke L (2016). RNAs containing modified nucleotides fail to trigger RIG-I conformational changes for innate immune signaling. MBio, 7(5), e00833–16. 10.1128/mBio.00833-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Epstein P, Reddy R, Henning D, & Busch H (1980). The nucleotide sequence of nuclear U6 (4.7 S) RNA. Journal of Biological Chemistry, 255(18), 8901–8906. 10.1016/s0021-9258(18)43587-9 [DOI] [PubMed] [Google Scholar]
  60. Feng Z, Zhou F, Tan M, Wang T, Chen Y, Xu W, Li B, Wang X, Deng X, & He ML (2022). Targeting m6A modification inhibits herpes virus 1 infection. Genes and Diseases, 9(4), 1114–1128. 10.1016/j.gendis.2021.02.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Figaro S, Scrima N, Buckingham RH, & Heurgué-Hamard V (2008). HemK2 protein, encoded on human chromosome 21, methylates translation termination factor eRF1. FEBS Letters, 582(16), 2352–2356. 10.1016/j.febslet.2008.05.045 [DOI] [PubMed] [Google Scholar]
  62. Finkel D, & Groner Y (1983). Methylations of adenosine residues (m6A) in pre-mRNA are important for formation of late simian virus 40 mRNAs. Virology, 131(2), 409–425. 10.1016/0042-6822(83)90508-1 [DOI] [PubMed] [Google Scholar]
  63. Fiorentino F, Menna M, Rotili D, Valente S, & Mai A (2023). METTL3 from target validation to the first small-molecule inhibitors: A medicinal chemistry journey. Journal of Medicinal Chemistry, 17, 41–1677. 10.1021/acs.jmedchem.2c01601 [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Flamand MN, & Meyer KD (2019). The epitranscriptome and synaptic plasticity. Current Opinion in Neurobiology, 59, 41–48. 10.1016/j.conb.2019.04.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Fleming AM, Nguyen NLB, & Burrows CJ (2019). Colocalization of m6A and G-Quadruplex-forming sequences in viral RNA (HIV, Zika, Hepatitis B, and SV40) suggests topological control of adenosine N6-methylation. American Chemical Society Central Science, 5(2), 218–228. 10.1021/acscentsci.8b00963 [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Flemmich L, Heel S, Moreno S, Breuker K, & Micura R (2021). A natural riboswitch scaffold with self-methylation activity. Nature Communications, 12(1), 1–10. 10.1038/s41467-021-24193-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Fu D, Brophy JAN, Chan CTY, Atmore KA, Begley U, Paules RS, Dedon PC, Begley TJ, & Samson LD (2010). Human AlkB homolog ABH8 is a tRNA methyltransferase required for wobble uridine modification and DNA damage survival. Molecular and Cellular Biology, 30(10), 2449–2459. 10.1128/mcb.01604-09 [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Fu Y, & Zhuang X (2020). m6A-binding YTHDF proteins promote stress granule formation. Nature Chemical Biology, 16(9), 955–963. 10.1038/s41589-020-0524-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Gana R, Rao S, Huang H, Wu C, & Vasudevan S (2013). Structural and functional studies of S-adenosyl-l-methionine binding proteins: A ligand-centric approach. BMC Structural Biology, 13(6), 1–14. http://www.biomedcentral.com/1472-6807/13/6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Gao J, Cahill CM, Huang X, Roffman JL, Lamon-Fava S, Fava M, Mischoulon D, & Rogers JT (2018). S-adenosyl methionine and transmethylation pathways in neuropsychiatric diseases throughout life. Neurotherapeutics, 15(1), 156–175. 10.1007/s13311-017-0593-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  71. Gao M, Su S, Cao J, Xiang S, Huang Y, Shu X, Ma J, & Liu J (2022). Targeted manipulation of cellular RNA m6A methylation at the single-base level. American Chemical Society Chemical Biology, 17(4), 854–863. 10.1021/acschembio.1c00895 [DOI] [PubMed] [Google Scholar]
  72. Gao Y, Pei G, Li D, Li R, Shao Y, Zhang QC, & Li P (2019). Multivalent m6A motifs promote phase separation of YTHDF proteins. Cell Research, 29(9), 767–769. 10.1038/s41422-019-0210-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  73. Garalde DR, Snell EA, Jachimowicz D, Sipos B, Lloyd JH, Bruce M, Pantic N, Admassu T, James P, Warland A, Jordan M, Ciccone J, Serra S, Keenan J, Martin S, McNeill L, Wallace EJ, Jayasinghe L, Wright C, … Turner DJ (2018). Highly parallel direct RNA sequencing on an array of nanopores. Nature Methods, 15(3), 201–206. 10.1038/nmeth.4577 [DOI] [PubMed] [Google Scholar]
  74. Garcia-Campos MA, Edelheit S, Toth U, Safra M, Shachar R, Viukov S, Winkler R, Nir R, Lasman L, Brandis A, Hanna JH, Rossmanith W, & Schwartz S (2019). Deciphering the “m6A code” via antibody-independent quantitative profiling. Cell, 178(3), 731–747.e16. 10.1016/j.cell.2019.06.013 [DOI] [PubMed] [Google Scholar]
  75. German DC, Bloch CA, & Kredich NM (1983). Measurements of S-adenosylmethionine and L-homocysteine metabolism in cultured human lymphoid cells. Journal of Biological Chemistry, 258(18), 10997–11003. 10.1016/s0021-9258(17)44376-6 [DOI] [PubMed] [Google Scholar]
  76. Geula S, Moshitch-Moshkovitz S, Dominissini D, Mansour AAF, Kol N, Salmon-Divon M, Hershkovitz V, Peer E, Mor N, Manor YS, Ben-Haim MS, Eyal E, Yunger S, Pinto Y, Jaitin DA, Viukov S, Rais Y, Krupalnik V, Chomsky E, … Hanna JH (2015). m6A mRNA methylation facilitates resolution of naïve pluripotency toward differentiation. Science, 347(6225), 1002–1006. 10.1126/science.1261417 [DOI] [PubMed] [Google Scholar]
  77. Goh YT, Koh CWQ, Sim DY, Roca X, & Goh WSS (2020). METTL4 catalyzes m6Am methylation in U2 snRNA to regulate pre-mRNA splicing. Nucleic Acids Research, 48(16), 9250–9261. 10.1093/NAR/GKAA684 [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Gokhale NS, McIntyre ABR, McFadden MJ, Roder AE, Kennedy EM, Gandara JA, Hopcraft SE, Quicke KM, Vazquez C, Willer J, Ilkayeva OR, Law BA, Holley CL, Garcia-Blanco MA, Evans MJ, Suthar MS, Bradrick SS, Mason CE, & Horner SM (2016). N6-methyladenosine in flaviviridae viral RNA genomes regulates infection. Cell Host & Microbe, 20(5), 654–665. 10.1016/j.chom.2016.09.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Grove TL, Benner JS, Radle MI, Ahlum JH, Landgraf BJ, Krebs C, & Booker SJ (2011). A radically different mechanism for S-adenosylmethionine-dependent methyltransferases. Science, 332(6029), 604–607. 10.1126/science.1200877 [DOI] [PubMed] [Google Scholar]
  80. Guan Q, Lin H, Miao L, Guo H, Chen Y, Zhuo Z, & He J (2022). Functions, mechanisms, and therapeutic implications of METTL14 in human cancer. Journal of Hematology and Oncology, 15(1), 1–15. 10.1186/s13045-022-01231-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Han D, Longhini AP, Zhang X, Hoang V, Wilson MZ, & Kosik KS (2022). Dynamic assembly of the mRNA m6A methyltransferase complex is regulated by METTL3 phase separation. PLoS Biology, 20(2), e3001535. 10.1371/journal.pbio.3001535 [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Han L, Dong L, Leung K, Zhao Z, Li Y, Gao L, Chen Z, Xue J, Qing Y, Li W, Pokharel SP, Gao M, Chen M, Shen C, Tan B, Small A, Wang K, Zhang Z, Qin X, … Deng X (2023). METTL16 drives leukemogenesis and leukemia stem cell self-renewal by reprogramming BCAA metabolism. Cell Stem Cell, 30(1), 52–68.e13. 10.1016/j.stem.2022.12.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  83. Han Y, Du T, Guo S, Wang L, Dai G, Long T, Xu T, Zhuang X, Liu C, Li S, Zhang D, Liao X, Dong Y, Lui KO, Tan X, Lin S, Chen Y, & Huang ZP (2022). Loss of m6A methyltransferase METTL5 promotes cardiac hypertrophy through epitranscriptomic control of SUZ12 expression. Frontiers in Cardiovascular Medicine, 9, 350. 10.3389/fcvm.2022.852775 [DOI] [PMC free article] [PubMed] [Google Scholar]
  84. Hao H, Hao S, Chen H, Chen Z, Zhang Y, Wang J, Wang H, Zhang B, Qiu J, Deng F, & Guan W (2019). N6-methyladenosine modification and METTL3 modulate enterovirus 71 replication. Nucleic Acids Research, 47(1), 362–374. 10.1093/nar/gky1007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Harada F, & Kato N (1980). Nucleotide sequences of 4.5S RNAs associated with poly(A)-containing RNAs of mouse and hamster cells. Nucleic Acids Research, 8(6), 1273–1286. 10.1093/nar/8.6.1273 [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Harper JE, Miceli SM, Roberts RJ, & Manley JL (1990). Sequence specificity of the human mRNA N6-adenosine methylase in vitro. Nucleic Acids Research, 18(19), 5735–5741. 10.1093/nar/18.19.5735 [DOI] [PMC free article] [PubMed] [Google Scholar]
  87. He PC, & He C (2021). m6A RNA methylation: From mechanisms to therapeutic potential. The EMBO Journal, 40(3), e105977. 10.15252/embj.2020105977 [DOI] [PMC free article] [PubMed] [Google Scholar]
  88. Hiller DA, & Strobel SA (2022). A new RNA performs old chemistry. Nature Chemical Biology, 18(5), 438–439. 10.1038/S41589-022-00998-5 [DOI] [PubMed] [Google Scholar]
  89. Holmes MJ, Padgett LR, Bastos MS, & Sullivan WJ (2021). m6A RNA methylation facilitates pre-mRNA 3′-end formation and is essential for viability of Toxoplasma gondii. PLoS Pathogens, 17(7), e1009335. 10.1371/journal.ppat.1009335 [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Horiuchi KY, Eason MM, Ferry JJ, Planck JL, Walsh CP, Smith RF, Howitz KT, & Ma H (2013). Assay development for histone methyltransferases. Assay and Drug Development Technologies, 11(4), 227–236. 10.1089/adt.2012.480 [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Hotchkiss RD (1948). The quantitative separation of purines, pyrimidines, and nucleosides by paper chromatography. The Journal of Biological Chemistry, 175(1), 315–332. 10.3891/acta.chem.scand.06-1030 [DOI] [PubMed] [Google Scholar]
  92. Huang H, Li H, Pan R, Wang S, Khan AA, Zhao Y, Zhu H, & Liu X (2022). Ribosome 18S m6A methyltransferase METTL5 promotes pancreatic cancer progression by modulating c-Myc translation. International Journal of Oncology, 60(1), 1–10. 10.3892/ijo.2021.5299 [DOI] [PubMed] [Google Scholar]
  93. Huang J, Dong X, Gong Z, Qin LY, Yang S, Zhu YL, Wang X, Zhang D, Zou T, Yin P, & Tang C (2019). Solution structure of the RNA recognition domain of METTL3-METTL14 N6-methyladenosine methyltransferase. Protein and Cell, 10(4), 272–284. 10.1007/s13238-018-0518-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  94. Ignatova VV, Jansen PWTC, Baltissen MP, Vermeulen M, & Schneider R (2019). The interactome of a family of potential methyltransferases in HeLa cells. Scientific Reports, 9(1), 1–9. 10.1038/s41598-019-43010-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  95. Ignatova VV, Stolz P, Kaiser S, Gustafsson TH, Lastres PR, Sanz-Moreno A, Cho YL, Amarie OV, Aguilar-Pimentel A, Klein-Rodewald T, Calzada-Wack J, Becker L, Marschall S, Kraiger M, Garrett L, Seisenberger C, Hölter SM, Borland K, Van De Logt E, … Schneider R (2020). The rRNA m6A methyltransferase METTL5 is involved in pluripotency and developmental programs. Genes and Development, 34(9–10), 715–729. 10.1101/gad.333369.119 [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Imam H, Khan M, Gokhale NS, McIntyre ABR, Kim GW, Jang JY, Kim SJ, Mason CE, Horner SM, & Siddiqui A (2018). N6-methyladenosine modification of hepatitis B virus RNA differentially regulates the viral life cycle. Proceedings of the National Academy of Sciences of the United States of America, 115(35), 8829–8834. 10.1073/pnas.1808319115 [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Ishigami Y, Ohira T, Isokawa Y, Suzuki Y, & Suzuki T (2021). A single m6A modification in U6 snRNA diversifies exon sequence at the 5′ splice site. Nature Communications, 12(1), 3244. 10.1038/s41467-021-23457-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  98. Jabs S, Biton A, Bécavin C, Nahori MA, Ghozlane A, Pagliuso A, Spanò G, Guérineau V, Touboul D, Giai Gianetto Q, Chaze T, Matondo M, Dillies MA, & Cossart P (2020). Impact of the gut microbiota on the m6A epitranscriptome of mouse cecum and liver. Nature Communications, 11(1), 1–16. 10.1038/s41467-020-15126-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  99. Jansens RJJJ, Verhamme R, Mirza AH, Olarerin-George A, Van Waesberghe C, Jaffrey SR, Favoreel HW, Jansens RJJJ, Verhamme R, Mirza AH, Olarerin-George A, Van Waesberghe C, Van Waesberghe C, Jaffrey SR, & Favoreel HW (2022). Alphaherpesvirus US3 protein-mediated inhibition of the m6A mRNA methyltransferase complex. Cell Reports, 40(3), 111107. 10.1016/j.celrep.2022.111107 [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Jia G, Fu Y, Zhao X, Dai Q, Zheng G, Yang Y, Yi C, Lindahl T, Pan T, Yang YG, & He C (2011). N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nature Chemical Biology, 7(12), 885–887. 10.1038/nchembio.687 [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Jiang H, Gao Y, Zhang L, Chen D, Gan J, & Murchie AIH (2021). The identification and characterization of a selected SAM-dependent methyltransferase ribozyme that is present in natural sequences. Nature Catalysis, 4(10), 872–881. 10.1038/s41929-021-00685-z [DOI] [Google Scholar]
  102. Jiang X, Liu B, Nie Z, Duan L, Xiong Q, Jin Z, Yang C, & Chen Y (2021). The role of m6A modification in the biological functions and diseases. Signal Transduction and Targeted Therapy, 6(1), 1–16. 10.1038/s41392-020-00450-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  103. Ju J, Aoyama T, Yashiro Y, Yamashita S, Kuroyanagi H, & Tomita K (2023). Structure of the Caenorhabditis elegans m6A methyltransferase METT10 that regulates SAM homeostasis. Nucleic Acids Research, 51(5), 2434–2446. 10.1093/nar/gkad081 [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Jumper J, Evans R, Pritzel A, Green T, Figurnov M, Ronneberger O, Tunyasuvunakool K, Bates R, Žídek A, Potapenko A, Bridgland A, Meyer C, Kohl SAA, Ballard AJ, Cowie A, Romera-Paredes B, Nikolov S, Jain R, Adler J, … Hassabis D (2021). Highly accurate protein structure prediction with AlphaFold. Nature, 596(7873), 583–589. 10.1038/s41586-021-03819-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  105. Kane SE, & Beemon K (1985). Precise localization of m6A in Rous sarcoma virus RNA reveals clustering of methylation sites: Implications for RNA processing. Molecular and Cellular Biology, 5(9), 2298–2306. 10.1128/mcb.5.9.2298-2306.1985 [DOI] [PMC free article] [PubMed] [Google Scholar]
  106. Karikó K, Buckstein M, Ni H, & Weissman D (2005). Suppression of RNA recognition by Toll-like receptors: The impact of nucleoside modification and the evolutionary origin of RNA. Immunity, 23(2), 165–175. 10.1016/j.immuni.2005.06.008 [DOI] [PubMed] [Google Scholar]
  107. Ke S, Pandya-Jones A, Saito Y, Fak JJ, Vågbø CB, Geula S, Hanna JH, Black DL, Darnell JE, & Darnell RB (2017). m6A mRNA modifications are deposited in nascent pre-mRNA and are not required for splicing but do specify cytoplasmic turnover. Genes and Development, 31(10), 990–1006. 10.1101/gad.301036.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Kennedy EM, Bogerd HP, Kornepati AVR, Kang D, Ghoshal D, Marshall JB, Poling BC, Tsai K, Gokhale NS, Horner SM, & Cullen BR (2016). Posttranscriptional m6A editing of HIV-1 mRNAs enhances viral gene expression. Cell Host and Microbe, 19(5), 675–685. 10.1016/j.chom.2016.04.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  109. Kim GW, & Siddiqui A (2021). Hepatitis B virus X protein recruits methyltransferases to affect cotranscriptional N6-methyladenosine modification of viral/host RNAs. Proceedings of the National Academy of Sciences of the United States of America, 118(3), e2019455118. 10.1073/pnas.2019455118 [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Kim GW, Imam H, Khan M, & Siddiqui A (2020). N6-ethyladenosine modification of hepatitis B and C viral RNAs attenuates host innate immunity via RIG-I signaling. Journal of Biological Chemistry, 295(37), 13123–13133. 10.1074/jbc.ra120.014260 [DOI] [PMC free article] [PubMed] [Google Scholar]
  111. Kim J, Kim Y, Yeom M, Kim JH, & Nam HG (2008). FIONA1 is essential for regulating period length in the Arabidopsis circadian clock. Plant Cell, 20(2), 307–319. 10.1105/tpc.107.055715 [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Koh CWQ, Goh YT, & Goh WSS (2019). Atlas of quantitative single-base-resolution N6-methyl-adenine methylomes. Nature Communications, 10(1), 5636. 10.1038/s41467-019-13561-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  113. Koh CWQ, Goh YT, Toh JDW, Neo SP, Ng SB, Gunaratne J, Gao YG, Quake SR, Burkholder WF, & Goh WSS (2018). Single-nucleotide-resolution sequencing of human N6-methyldeoxyadenosine reveals strand-asymmetric clusters associated with SSBP1 on the mitochondrial genome. Nucleic Acids Research, 46(22), 11659–11670. 10.1093/nar/gky1104 [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Kong Y, Cao L, Deikus G, Fan Y, Mead EA, Lai W, Zhang Y, Yong R, Sebra R, Wang H, Zhang XS, & Fang G (2022). Critical assessment of DNA adenine methylation in eukaryotes using quantitative deconvolution. Science, 375(6580), 515–522. 10.1126/science.abe7489 [DOI] [PMC free article] [PubMed] [Google Scholar]
  115. Körtel N, Rücklé C, Zhou Y, Busch A, Hoch-Kraft P, Sutandy FXR, Haase J, Pradhan M, Musheev M, Ostareck D, Ostareck-Lederer A, Dieterich C, Hüttelmaier S, Niehrs C, Rausch O, Dominissini D, König J, & Zarnack K (2021). Deep and accurate detection of m6A RNA modifications using miCLIP2 and m6Aboost machine learning. Nucleic Acids Research, 49(16), e92. 10.1093/nar/gkab485 [DOI] [PMC free article] [PubMed] [Google Scholar]
  116. Krug RM, Morgan MA, & Shatkin AJ (1976). Influenza viral mRNA contains internal N6-methyladenosine and 5′-terminal 7-methylguanosine in cap structures. Journal of Virology, 20(1), 45–53. 10.1128/JVI.20.1.45-53.1976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Krusnauskas R, Stakaitis R, Steponaitis G, Almstrup K, & Vaitkiene P (2023). Identification and comparison of m6A modifications in glioblastoma non-coding RNAs with MeRIP-seq and Nanopore dRNA-seq. Epigenetics, 18(1), 2163365. 10.1080/15592294.2022.2163365 [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Lan L, Sun YJ, Jin XY, Xie LJ, Liu L, & Cheng L (2021). A light-controllable chemical modulation of m6A RNA methylation. Angewandte Chemie—International Edition, 60(33), 18116–18121. 10.1002/anie.202103854 [DOI] [PubMed] [Google Scholar]
  119. Lang F, Singh RK, Pei Y, Zhang S, Sun K, & Robertson ES (2019). EBV epitranscriptome reprogramming by METTL14 is critical for viral-associated tumorigenesis. PLoS Pathogens, 15(6), e1007796. 10.1371/journal.ppat.1007796 [DOI] [PMC free article] [PubMed] [Google Scholar]
  120. Lavi U, Fernandez-mufioz R, & Darnell JE (1977). Content of N-6 methyl adenylic acid in heterogeneous nuclear and messenger RNA of HeLa cells. Nucleic Acids Research, 4(1), 63–69. 10.1093/nar/4.1.63 [DOI] [PMC free article] [PubMed] [Google Scholar]
  121. Lee JH, Kim S, Jin MS, & Kim YC (2022). Discovery of substituted indole derivatives as allosteric inhibitors of m6A-RNA methyltransferase, METTL3-14 complex. Drug Development Research, 83(3), 783–799. 10.1002/ddr.21910 [DOI] [PubMed] [Google Scholar]
  122. Lee JH, Wang R, Xiong F, Krakowiak J, Liao Z, Nguyen PT, Moroz-Omori EV, Shao J, Zhu X, Bolt MJ, Wu H, Singh PK, Bi M, Shi CJ, Jamal N, Li G, Mistry R, Jung SY, Tsai KL, … Li W (2021). Enhancer RNA m6A methylation facilitates transcriptional condensate formation and gene activation. Molecular Cell, 81(16), 3368–3385.e9. 10.1016/j.molcel.2021.07.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Leismann J, Spagnuolo M, Pradhan M, Wacheul L, Vu MA, Musheev M, Mier P, Andrade-Navarro MA, Graille M, Niehrs C, Lafontaine DL, & Roignant J (2020). The 18S ribosomal RNA m 6 A methyltransferase Mettl5 is required for normal walking behavior in Drosophila. EMBO Reports, 21(7), e49443. 10.15252/embr.201949443 [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Li F, Kennedy S, Hajian T, Gibson E, Seitova A, Xu C, Arrowsmith CH, & Vedadi M (2016). A radioactivity-based assay for screening human m6A-RNA methyltransferase, METTL3-METTL14 complex, and demethylase ALKBH5. Journal of Biomolecular Screening, 21(3), 290–297. 10.1177/1087057115623264 [DOI] [PubMed] [Google Scholar]
  125. Li J, & Gregory RI (2021). Mining for METTL3 inhibitors to suppress cancer. Nature Structural and Molecular Biology, 28(6), 460–462. 10.1038/s41594-021-00606-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  126. Li J, Chen Z, Chen F, Xie G, Ling Y, Peng Y, Lin Y, Luo N, Chiang CM, & Wang H (2020). Targeted mRNA demethylation using an engineered dCas13b-ALKBH5 fusion protein. Nucleic Acids Research, 48(10), 5684–5694. 10.1093/nar/gkaa269 [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Li N, Hui H, Bray B, Gonzalez GM, Zeller M, Anderson KG, Knight R, Smith D, Wang Y, Carlin AF, & Rana TM (2021). METTL3 regulates viral m6A RNA modification and host cell innate immune responses during SARS-CoV-2 infection. Cell Reports, 35(6), 109091. 10.1016/j.celrep.2021.109091 [DOI] [PMC free article] [PubMed] [Google Scholar]
  128. Li W, Shi Y, Zhang T, Ye J, & Ding J (2019). Structural insight into human N6amt1–Trm112 complex functioning as a protein methyltransferase. Cell Discovery, 5(1), 1–13. 10.1038/s41421-019-0121-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  129. Li X, Tang J, Huang W, Wang F, Li P, Qin C, Qin Z, Zou Q, Wei J, Hua L, Yang H, & Wang Z (2017). The m6A methyltransferase METTL3: Acting as a tumor suppressor in renal cell carcinoma. Oncotarget, 8(56), 96103–96116. 10.18632/oncotarget.21726 [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Li X, Zhang Z, Luo X, Schrier J, Yang AD, & Wu TP (2021). The exploration of N6-deoxyadenosine methylation in mammalian genomes. Protein and Cell, 12(10), 756–768. 10.1007/s13238-021-00866-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Li Y, Bedi RK, Moroz-Omori EV, & Caflisch A (2020). Structural and dynamic insights into redundant function of YTHDF proteins. Journal of Chemical Information and Modeling, 60(12), 5932–5935. 10.1021/acs.jcim.0c01029 [DOI] [PubMed] [Google Scholar]
  132. Liberman N, O’Brown ZK, Earl AS, Boulias K, Gerashchenko MV, Wang SY, Fritsche C, Fady PE, Dong A, Gladyshev VN, & Greer EL (2020). N6-adenosine methylation of ribosomal RNA affects lipid oxidation and stress resistance. Science Advances, 6(17), eaaz4370. 10.1126/sciadv.aaz4370 [DOI] [PMC free article] [PubMed] [Google Scholar]
  133. Lichinchi G, & Rana TM (2019). Profiling of N6-methyladenosine in Zika virus RNA and host cellular mRNA. Methods in Molecular Biology, 1870, 209–218. 10.1007/978-1-4939-8808-2_16 [DOI] [PubMed] [Google Scholar]
  134. Lichinchi G, Gao S, Saletore Y, Gonzalez GM, Bansal V, Wang Y, Mason CE, & Rana TM (2016). Dynamics of the human and viral m6A RNA methylomes during HIV-1 infection of T cells. Nature Microbiology, 1(4), 16011. 10.1038/nmicrobiol.2016.11 [DOI] [PMC free article] [PubMed] [Google Scholar]
  135. Lichinchi G, Zhao BS, Wu Y, Lu Z, Qin Y, He C, & Rana TM (2016). Dynamics of human and viral RNA methylation during Zika virus infection. Cell Host and Microbe, 20(5), 666–673. 10.1016/j.chom.2016.10.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  136. Lin S, Choe J, Du P, Triboulet R, & Gregory RI (2016). The m6A methyltransferase METTL3 promotes translation in human cancer cells. Molecular Cell, 62(3), 335–345. 10.1016/j.molcel.2016.03.021 [DOI] [PMC free article] [PubMed] [Google Scholar]
  137. Lin Z, Hsu PJ, Xing X, Fang J, Lu Z, Zou Q, Zhang KJ, Zhang X, Zhou Y, Zhang T, Zhang Y, Song W, Jia G, Yang X, He C, & Tong MH (2017). Mettl3-/Mettl14-mediated mRNA N6-methyladenosine modulates murine spermatogenesis. Cell Research, 27(10), 1216–1230. 10.1038/cr.2017.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  138. Linder B, Grozhik AV, Olarerin-George AO, Meydan C, Mason CE, & Jaffrey SR (2015). Single-nucleotide-resolution mapping of m6A and m6Am throughout the transcriptome. Nature Methods, 12(8), 767–772. 10.1038/nmeth.3453 [DOI] [PMC free article] [PubMed] [Google Scholar]
  139. Liu C, Cao J, Zhang H, & Yin J (2022). Evolutionary history of RNA modifications at N6-adenosine originating from the R-M system in eukaryotes and prokaryotes. Biology, 11(2), 214. 10.3390/biology11020214 [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Liu C, Sun H, Yi Y, Shen W, Li K, Xiao Y, Li F, Li Y, Hou Y, Lu B, Liu W, Meng H, Peng J, Yi C, & Wang J (2022). Absolute quantification of single-base m6A methylation in the mammalian transcriptome using GLORI. Nature Biotechnology, 41, 355–366. 10.1038/s41587-022-01487-9 [DOI] [PubMed] [Google Scholar]
  141. Liu H, Begik O, Lucas MC, Ramirez JM, Mason CE, Wiener D, Schwartz S, Mattick JS, Smith MA, & Novoa EM (2019). Accurate detection of m6A RNA modifications in native RNA sequences. Nature Communications, 10(1), 1–9. 10.1038/s41467-019-11713-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Liu J, Dou X, Chen C, Chen C, Liu C, Michelle Xu M, Zhao S, Shen B, Gao Y, Han D, & He C (2020). N6-methyladenosine of chromosome-associated regulatory RNA regulates chromatin state and transcription. Science, 367(6477), 580–586. 10.1126/science.aay6018 [DOI] [PMC free article] [PubMed] [Google Scholar]
  143. Liu J, Xu YP, Li K, Ye Q, Zhou HY, Sun H, Li X, Yu L, Deng YQ, Li RT, Cheng ML, He B, Zhou J, Li XF, Wu A, Yi C, & Qin CF (2021). The m6A methylome of SARS-CoV-2 in host cells. Cell Research, 31(4), 404–414. 10.1038/s41422-020-00465-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
  144. 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. Nature Chemical Biology, 10(2), 93–95. 10.1038/nchembio.1432 [DOI] [PMC free article] [PubMed] [Google Scholar]
  145. Liu J, 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 (2018). VIRMA mediates preferential m6A mRNA methylation in 3′UTR and near stop codon and associates with alternative polyadenylation. Cell Discovery, 4(1), 1–17. 10.1038/s41421-018-0019-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  146. Liu L, Zeng S, Jiang H, Zhang Y, Guo X, & Wang Y (2019). Differential m6A methylomes between two major life stages allows potential regulations in Trypanosoma brucei. Biochemical and Biophysical Research Communications, 508(4), 1286–1290. 10.1016/j.bbrc.2018.12.043 [DOI] [PubMed] [Google Scholar]
  147. Liu N, Dai Q, Zheng G, He C, Parisien M, & Pan T (2015). N6-methyladenosine-dependent RNA structural switches regulate RNA-protein interactions. Nature, 518(7540), 560–564. 10.1038/nature14234 [DOI] [PMC free article] [PubMed] [Google Scholar]
  148. Liu N, Parisien M, Dai Q, Zheng G, He C, & Pan T (2013). Probing N6-methyladenosine RNA modification status at single nucleotide resolution in mRNA and long noncoding RNA. RNA, 19(12), 1848–1856. 10.1261/RNA.041178.113 [DOI] [PMC free article] [PubMed] [Google Scholar]
  149. Liu N, Zhou KI, Parisien M, Dai Q, Diatchenko L, & Pan T (2017). N6-methyladenosine alters RNA structure to regulate binding of a low-complexity protein. Nucleic Acids Research, 45(10), 6051–6063. 10.1093/nar/gkx141 [DOI] [PMC free article] [PubMed] [Google Scholar]
  150. Liu S, Chen L, Zhang Y, Zhou Y, He Y, Chen Z, Qi S, Zhu J, Chen X, Zhang H, Luo Y, Qiu Y, Tao L, & Zhu F (2023). m6AREG: m6A-centered regulation of disease development and drug response. Nucleic Acids Research, 51(D1), D1333–D1344. 10.1093/nar/gkac801 [DOI] [PMC free article] [PubMed] [Google Scholar]
  151. Liu X, Shen S, Wu P, Li F, Liu X, Wang C, Gong Q, Wu J, Yao X, Zhang H, & Shi Y (2019). Structural insights into dimethylation of 12S rRNA by TFB1M: Indispensable role in translation of mitochondrial genes and mitochondrial function. Nucleic Acids Research, 47(14), 7648–7665. 10.1093/nar/gkz505 [DOI] [PMC free article] [PubMed] [Google Scholar]
  152. Liu Y, Zheng Z, Shu B, Meng J, Zhang Y, Zheng C, Ke X, Gong P, Hu Q, & Wang H (2016). SUMO modification stabilizes enterovirus 71 polymerase 3D to facilitate viral replication. Journal of Virology, 90(23), 10472–10485. 10.1128/jvi.01756-16 [DOI] [PMC free article] [PubMed] [Google Scholar]
  153. Lopez M, Halby L, & Arimondo PB (2016). DNA methyltransferase inhibitors: Development and applications. Advances in Experimental Medicine and Biology, 945, 431–473. 10.1007/978-3-319-43624-1_16 [DOI] [PubMed] [Google Scholar]
  154. Louloupi A, Ntini E, Conrad T, & Ørom UAV (2018). Transient N6-methyladenosine transcriptome sequencing reveals a regulatory role of m6A in splicing efficiency. Cell Reports, 23(12), 3429–3437. 10.1016/j.celrep.2018.05.077 [DOI] [PubMed] [Google Scholar]
  155. Lu M, Zhang Z, Xue M, Zhao BS, Harder O, Li A, Liang X, Gao TZ, Xu Y, Zhou J, Feng Z, Niewiesk S, Peeples ME, He C, & Li J (2020). N6-methyladenosine modification enables viral RNA to escape recognition by RNA sensor RIG-I. Nature Microbiology, 5(4), 584–598. 10.1038/S41564-019-0653-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Lucas MC, & Novoa EM (2023). Long-read sequencing in the era of epigenomics and epitranscriptomics. Nature Methods, 20(1), 25–29. 10.1038/s41592-022-01724-8 [DOI] [PubMed] [Google Scholar]
  157. Luo Q, Mo J, Chen H, Hu Z, Wang B, Wu J, Liang Z, Xie W, Du K, Peng M, Li Y, Li T, Zhang Y, Shi X, Shen WH, Shi Y, Dong A, Wang H, & Ma J (2022). Structural insights into molecular mechanism for N6-adenosine methylation by MT-A70 family methyltransferase METTL4. Nature Communications, 13(1), 1–11. 10.1038/s41467-022-33277-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  158. Ma H, Wang X, Cai J, Dai Q, Natchiar SK, Lv R, Chen K, Lu Z, Chen H, Shi YGY, Lan F, Fan J, Klaholz BP, Pan T, Shi YGY, & He C (2019). N6-Methyladenosine methyltransferase ZCCHC4 mediates ribosomal RNA methylation. Nature Chemical Biology, 15(1), 88–94. 10.1038/s41589-018-0184-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  159. Ma JZ, Yang F, Zhou CC, Liu F, Yuan JH, Wang F, Wang TT, Xu QG, Zhou WP, & Sun SH (2017). METTL14 suppresses the metastatic potential of hepatocellular carcinoma by modulating N6-methyladenosine-dependent primary microRNA processing. Hepatology, 65(2), 529–543. 10.1002/hep.28885 [DOI] [PubMed] [Google Scholar]
  160. Maden BEH (1986). Identification of the locations of the methyl groups in 18 S ribosomal RNA from Xenopus laevis and man. Journal of Molecular Biology, 189(4), 681–699. 10.1016/0022-2836(86)90498-5 [DOI] [PubMed] [Google Scholar]
  161. Maden BEH (1988). Locations of methyl groups in 28 S rRNA of Xenopus laevis and man. Clustering in the conserved core of molecule. Journal of Molecular Biology, 201(2), 289–314. 10.1016/0022-2836(88)90139-8 [DOI] [PubMed] [Google Scholar]
  162. Madugalle SU, Meyer K, Wang DO, & Bredy TW (2020). RNA N6-methyladenosine and the regulation of RNA localization and function in the brain. Trends in Neurosciences, 43(12), 1011–1023. 10.1016/j.tins.2020.09.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  163. Mahmoodi N, Harijan RK, & Schramm VL (2020). Transition-state analogues of phenylethanolamine N-methyltransferase. Journal of the American Chemical Society, 142(33), 14222–14233. 10.1021/jacs.0c05446 [DOI] [PMC free article] [PubMed] [Google Scholar]
  164. Malone T, Blumenthal RM, & Cheng X (1995). Structure-guided analysis reveals nine sequence motifs conserved among DNA aminomethyltransferases, and suggests a catalytic mechanism for these enzymes. Journal of Molecular Biology, 253(4), 618–632. 10.1006/jmbi.1995.0577 [DOI] [PubMed] [Google Scholar]
  165. Mashhoon N, & Reich NO (1994). Investigation of ionizable residues critical for sequence-specific enzymatic DNA modification: Protein modification and steady-state and pre-steady-state kinetic pH analyses of EcoRI DNA methyltransferase. Biochemistry, 33(23), 7113–7119. 10.1021/bi00189a014 [DOI] [PubMed] [Google Scholar]
  166. Matsuoka S, Ballif BA, Smogorzewska A, McDonald ER, Hurov KE, Luo J, Bakalarski CE, Zhao Z, Solimini N, Lerenthal Y, Shiloh Y, Gygi SP, & Elledge SJ (2007). ATM and ATR substrate analysis reveals extensive protein networks responsive to DNA damage. Science, 316(5828), 1160–1166. 10.1126/science.1140321 [DOI] [PubMed] [Google Scholar]
  167. Mauer J, & Jaffrey SR (2018). FTO, m6Am, and the hypothesis of reversible epitranscriptomic mRNA modifications. Federation of European Biochemical Societies Letters, 592(12), 2012–2022. 10.1002/1873-3468.13092 [DOI] [PubMed] [Google Scholar]
  168. Mayers DL, Mikovits JA, Joshi B, Hewlett IK, Estrada JS, Wolfe AD, Garcia GE, Doctor BP, Burke DS, Gordon RK, Lane JR, & Chiang PK (1995). Anti-human immunodeficiency virus 1 (HIV-1) activities of 3-deazaadenosine analogs: Increased potency against 3′-azido-3′-deoxythymidine-resistant HIV-1 strains. Proceedings of the National Academy of Sciences of the United States of America, 92(1), 215–219. 10.1073/pnas.92.1.215 [DOI] [PMC free article] [PubMed] [Google Scholar]
  169. McCown PJ, Ruszkowska A, Kunkler CN, Breger K, Hulewicz JP, Wang MC, Springer NA, & Brown JA (2020). Naturally occurring modified ribonucleosides. Wiley Interdisciplinary Reviews: RNA, 11(5), e1595. 10.1002/wrna.1595 [DOI] [PMC free article] [PubMed] [Google Scholar]
  170. McCulloch V, Seidel-Rogol BL, & Shadel GS (2002). A human mitochondrial transcription factor is related to RNA adenine methyltransferases and binds S-adenosylmethionine. Molecular and Cellular Biology, 22(4), 1116–1125. 10.1128/mcb.22.4.1116-1125.2002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  171. Meiser N, Mench N, & Hengesbach M (2020). RNA secondary structure dependence in METTL3–METTL14 mRNA methylation is modulated by the N-terminal domain of METTL3. Biological Chemistry, 402(1), 89–98. 10.1515/hsz-2020-0265 [DOI] [PubMed] [Google Scholar]
  172. Mendel M, Chen KM, Homolka D, Gos P, Pandey RR, McCarthy AA, & Pillai RS (2018). Methylation of structured RNA by the m6A writer METTL16 is essential for mouse embryonic development. Molecular Cell, 71(6), 986–1000.e11. 10.1016/j.molcel.2018.08.004 [DOI] [PMC free article] [PubMed] [Google Scholar]
  173. Mendel M, Delaney K, Pandey RR, Chen KM, Wenda JM, Vågbø CB, Steiner FA, Homolka D, & Pillai RS (2021). Splice site m6A methylation prevents binding of U2AF35 to inhibit RNA splicing. Cell, 184(12), 3125–3142.e25. 10.1016/j.cell.2021.03.062 [DOI] [PMC free article] [PubMed] [Google Scholar]
  174. Meng TG, Lu X, Guo L, Hou GM, Ma XS, Li QN, Huang L, Fan LH, Zhao ZH, Ou XH, OuYang YC, Schatten H, Li L, Wang ZB, & Sun QY (2019). Mettl14 is required for mouse postimplantation development by facilitating epiblast maturation. FASEB Journal, 33(1), 1179–1187. 10.1096/fj.201800719R [DOI] [PubMed] [Google Scholar]
  175. Metodiev MD, Lesko N, Park CB, Cámara Y, Shi Y, Wibom R, Hultenby K, Gustafsson CM, & Larsson NG (2009). Methylation of 12S rRNA is necessary for In vivo stability of the small subunit of the mammalian mitochondrial ribosome. Cell Metabolism, 9(4), 386–397. 10.1016/j.cmet.2009.03.001 [DOI] [PubMed] [Google Scholar]
  176. Metzger E, Wang S, Urban S, Willmann D, Schmidt A, Offermann A, Allen A, Sum M, Obier N, Cottard F, Ulferts S, Preca BT, Hermann B, Maurer J, Greschik H, Hornung V, Einsle O, Perner S, Imhof A, … Schüle R (2019). KMT9 monomethylates histone H4 lysine 12 and controls proliferation of prostate cancer cells. Nature Structural and Molecular Biology, 26(5), 361–371. 10.1038/s41594-019-0219-9 [DOI] [PubMed] [Google Scholar]
  177. Meyer KD (2019). DART-seq: An antibody-free method for global m6A detection. Nature Methods, 16(12), 1275–1280. 10.1038/s41592-019-0570-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  178. Meyer KD, Patil DP, Zhou J, Zinoviev A, Skabkin MA, Elemento O, Pestova TV, Qian SB, & Jaffrey SR (2015). 5′ UTR m6A promotes cap-independent translation. Cell, 163(4), 999–1010. 10.1016/j.cell.2015.10.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
  179. Meyer KD, Saletore Y, Zumbo P, Elemento O, Mason CE, & Jaffrey SR (2012). Comprehensive analysis of mRNA methylation reveals enrichment in 3′ UTRs and near stop codons. Cell, 149(7), 1635–1646. 10.1016/j.cell.2012.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  180. Mikutis S, Gu M, Sendinc E, Hazemi ME, Kiely-Collins H, Aspris D, Vassiliou GS, Shi Y, Tzelepis K, & Bernardes GJL (2020). MeCLICK-Seq, a substrate-hijacking and RNA degradation strategy for the study of RNA methylation. American Chemical Society Central Science, 6(12), 2196–2208. 10.1021/acscentsci.0c01094 [DOI] [PMC free article] [PubMed] [Google Scholar]
  181. Mitra A, Manna S, Kundu R, Hazra D, & Roychowdhury A (2023). Brute force virtual drug screening with molecular dynamics simulation and MM/PBSA to find potent inhibitors of METTL16. IEEE/ACM Transactions on Computational Biology and Bioinformatics, 20, 2356–2361. 10.1109/tcbb.2022.3233670 [DOI] [PubMed] [Google Scholar]
  182. Mongan NP, Emes RD, & Archer N (2019). Detection and analysis of RNA methylation. F1000Research, 8, 8. 10.12688/f1000research.17956.1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  183. Montgomery JA, Clayton SJ, Thomas HJ, Shannon WM, Arnett G, Bodner AJ, Kion IK, Cantoni GL, & Chiang PK (1982). Carbocyclic analogue of 3-deazaadenosine: A novel antiviral agent using S-adenosylhomocysteine hydrolase as a pharmacological target. Journal of Medicinal Chemistry, 25(6), 626–629. 10.1021/jm00348a004 [DOI] [PubMed] [Google Scholar]
  184. Moroz-Omori EV, Huang D, Kumar Bedi R, Cheriyamkunnel SJ, Bochenkova E, Dolbois A, Rzeczkowski MD, Li Y, Wiedmer L, & Caflisch A (2021). METTL3 inhibitors for epitranscriptomic modulation of cellular processes. ChemMedChem, 16(19), 3035–3043. 10.1002/cmdc.202100291 [DOI] [PMC free article] [PubMed] [Google Scholar]
  185. Moss B, & Koczot F (1976). Sequence of methylated nucleotides at the 5′-terminus of adenovirus-specific RNA. Journal of Virology, 17(2), 385–392. 10.1128/jvi.17.2.385-392.1976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  186. Moss B, Gershowitz A, Stringer JR, Holland LE, & Wagner EK (1977). 5′-Terminal and internal methylated nucleosides in herpes simplex virus type 1 mRNA. Journal of Virology, 23(2), 234–239. 10.1128/jvi.23.2.234-239.1977 [DOI] [PMC free article] [PubMed] [Google Scholar]
  187. Murakami S, & Jaffrey SR (2022). Hidden codes in mRNA: Control of gene expression by m6A. Molecular Cell, 82(12), 2236–2251. 10.1016/j.molcel.2022.05.029 [DOI] [PMC free article] [PubMed] [Google Scholar]
  188. Musheev MU, Baumgärtner A, Krebs L, & Niehrs C (2020). The origin of genomic N6-methyl-deoxyadenosine in mammalian cells. Nature Chemical Biology, 16(6), 630–634. 10.1038/s41589-020-0504-2 [DOI] [PubMed] [Google Scholar]
  189. Nance DJ, Satterwhite ER, Bhaskar B, Misra S, Carraway KR, & Mansfield KD (2020). Characterization of METTL16 as a cytoplasmic RNA binding protein. PLoS One, 15(1), e0227647. 10.1371/journal.pone.0227647 [DOI] [PMC free article] [PubMed] [Google Scholar]
  190. Narayan P, & Rottman FM (1988). An in vitro system for accurate methylation of internal adenosine residues in messenger RNA. Science, 242(4882), 1159–1162. 10.1126/science.3187541 [DOI] [PubMed] [Google Scholar]
  191. Natchiar SK, Myasnikov AG, Kratzat H, Hazemann I, & Klaholz BP (2017). Visualization of chemical modifications in the human 80S ribosome structure. Nature, 551(7681), 472–477. 10.1038/nature24482 [DOI] [PubMed] [Google Scholar]
  192. Newby ZER, Lau EY, & Bruice TC (2002). A theoretical examination of the factors controlling the catalytic efficiency of the DNA-(adenine-N6)-methyltransferase from Thermus aquaticus. Proceedings of the National Academy of Sciences of the United States of America, 99(12), 7922–7927. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Nolan LL (1987). Molecular target of the antileishmanial action of sinefungin. Antimicrobial Agents and Chemotherapy, 31(10), 1542–1548. 10.1128/AAC.31.10.1542 [DOI] [PMC free article] [PubMed] [Google Scholar]
  194. O’Brown ZK, & Greer EL (2016). N6-methyladenine: A conserved and dynamic DNA mark. Advances in Experimental Medicine and Biology, 945, 213–246. 10.1007/978-3-319-43624-1_10 [DOI] [PMC free article] [PubMed] [Google Scholar]
  195. Oerum S, Catala M, Atdjian C, Brachet F, Ponchon L, Barraud P, Iannazzo L, Droogmans L, Braud E, Ethève-Quelquejeu M, & Tisné C (2019). Bisubstrate analogues as structural tools to investigate m6A methyltransferase active sites. RNA Biology, 16(6), 798–808. 10.1080/15476286.2019.1589360 [DOI] [PMC free article] [PubMed] [Google Scholar]
  196. Pandey RR, & Pillai RS (2019). Counting the cuts: MAZTER-Seq quantifies m6A levels using a methylation-sensitive ribonuclease. Cell, 178(3), 515–517. 10.1016/j.cell.2019.07.006 [DOI] [PubMed] [Google Scholar]
  197. Parker MT, Soanes BK, Kusakina J, Larrieu A, Knop K, Joy N, Breidenbach F, Sherwood AV, Barton GJ, Fica SM, Davies BH, & Simpson GG (2022). m6A modification of U6 snRNA modulates usage of two major classes of pre-mRNA 5′ splice site. eLife, 11, e78808. 10.7554/eLife.78808 [DOI] [PMC free article] [PubMed] [Google Scholar]
  198. Patil DP, Chen CK, Pickering BF, Chow A, Jackson C, Guttman M, & Jaffrey SR (2016). m6A RNA methylation promotes XIST-mediated transcriptional repression. Nature, 537(7620), 369–373. 10.1038/nature19342 [DOI] [PMC free article] [PubMed] [Google Scholar]
  199. Pendleton KE, Chen B, Liu K, Hunter OV, Xie Y, Tu BP, & Conrad NK (2017). The U6 snRNA m6A methyltransferase METTL16 regulates SAM synthetase intron retention. Cell, 169(5), 824–835.e14. 10.1016/j.cell.2017.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  200. Peng H, Chen B, Wei W, Guo S, Han H, Yang C, Ma J, Wang L, Peng S, Kuang M, & Lin S (2022). N6-methyladenosine (m6A) in 18S rRNA promotes fatty acid metabolism and oncogenic transformation. Nature Metabolism, 4(8), 1041–1054. 10.1038/s42255-022-00622-9 [DOI] [PubMed] [Google Scholar]
  201. Petrossian TC, & Clarke SG (2011). Uncovering the human methyltransferasome. Molecular and Cellular Proteomics, 10(1), M110.000976. 10.1074/mcp.M110.000976 [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Ping XL, Sun BF, Wang L, Xiao W, Yang X, Wang WJ, Adhikari S, Shi Y, Lv Y, Chen YS, Zhao X, Li A, Yang Y, Dahal U, Lou XM, Liu X, Huang J, Yuan WP, Zhu XF, … Yang YG (2014). Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Research, 24(2), 177–189. 10.1038/cr.2014.3 [DOI] [PMC free article] [PubMed] [Google Scholar]
  203. Pinto R, Vågbø CB, Jakobsson ME, Kim Y, Baltissen MP, O’Donohue MF, Guzmán UH, Małecki JM, Wu J, Kirpekar F, Olsen JV, Gleizes PE, Vermeulen M, Leidel SA, Slupphaug G, & Falnes PO (2020). The human methyltransferase ZCCHC4 catalyses N6-methyladenosine modification of 28S ribosomal RNA. Nucleic Acids Research, 48(2), 830–846. 10.1093/nar/gkz1147 [DOI] [PMC free article] [PubMed] [Google Scholar]
  204. Pogolotti AL, Ono A, Subramaniam R, & Santi DV (1988). On the mechanism of DNA-adenine methylase. The Journal of Biological Chemistry, 263(16), 7461–7464. 10.1016/s0021-9258(18)68520-5 [DOI] [PubMed] [Google Scholar]
  205. Poh HX, Mirza AH, Pickering BF, & Jaffrey SR (2022). Alternative splicing of METTL3 explains apparently METTL3-independent m6A modifications in mRNA. PLoS Biology, 20(7), e3001683. 10.1371/journal.pbio.3001683 [DOI] [PMC free article] [PubMed] [Google Scholar]
  206. Price AM, Hayer KE, McIntyre ABR, Gokhale NS, Abebe JS, Della Fera AN, Mason CE, Horner SM, Wilson AC, Depledge DP, & Weitzman MD (2020). Direct RNA sequencing reveals m6A modifications on adenovirus RNA are necessary for efficient splicing. Nature Communications, 11(1), 1–17. 10.1038/s41467-020-19787-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  207. Qi S, Mota J, Chan SH, Villarreal J, Dai N, Arya S, Hromas RA, Rao MK, Corrêa IR, & Gupta YK (2022). RNA binding to human METTL3-METTL14 restricts N6-deoxyadenosine methylation of 3 DNA in vitro. eLife, 11, e67150. 10.7554/ELIFE.67150 [DOI] [PMC free article] [PubMed] [Google Scholar]
  208. Qiu W, Zhang Q, Zhang R, Lu Y, Wang X, Tian H, Yang Y, Gu Z, Gao Y, Yang X, Cui G, Sun B, Peng Y, Deng H, Peng H, Yang A, Yang YG, & Yang P (2021). N6-methyladenosine RNA modification suppresses antiviral innate sensing pathways via reshaping double-stranded RNA. Nature Communications, 12(1), 1–16. 10.1038/s41467-021-21904-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  209. Qiu X, Hua X, Li Q, Zhou Q, & Chen J (2021). m6A regulator-mediated methylation modification patterns and characteristics of immunity in blood leukocytes of COVID-19 patients. Frontiers in Immunology, 12, 5025. 10.3389/fimmu.2021.774776 [DOI] [PMC free article] [PubMed] [Google Scholar]
  210. Ramanathan A, Robb GB, & Chan SH (2016). mRNA capping: Biological functions and applications. Nucleic Acids Research, 44(16), 7511–7526. 10.1093/nar/gkw551 [DOI] [PMC free article] [PubMed] [Google Scholar]
  211. Rao ST, & Rossmann MG (1973). Comparison of super-secondary structures in proteins. Journal of Molecular Biology, 76(2), 241–256. 10.1016/0022-2836(73)90388-4 [DOI] [PubMed] [Google Scholar]
  212. Ren W, Lu J, Huang M, Gao L, Li D, Greg Wang G, & Song J (2019). Structure and regulation of ZCCHC4 in m6A-methylation of 28S rRNA. Nature Communications, 10, 5042. 10.1038/s41467-019-12923-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Reuter MS, Tawamie H, Buchert R, Gebril OH, Froukh T, Thiel C, Uebe S, Ekici AB, Krumbiegel M, Zweier C, Hoyer J, Eberlein K, Bauer J, Scheller U, Strom TM, Hoffjan S, Abdelraouf ER, Meguid NA, Abboud A, … Jamra RA (2017). Diagnostic yield and novel candidate genes by exome sequencing in 152 consanguineous families with neurodevelopmental disorders. Journal of the American Medical Association Psychiatry, 74(3), 293–299. 10.1001/jamapsychiatry.2016.3798 [DOI] [PubMed] [Google Scholar]
  214. Riazuddin S, Hussain M, Razzaq A, Iqbal Z, Shahzad M, Polla DL, Song Y, Van Beusekom E, Khan AA, Tomas-Roca L, Rashid M, Zahoor MY, Wissink-Lindhout WM, Basra MAR, Ansar M, Agha Z, Van Heeswijk K, Rasheed F, Van De Vorst M, … Van Bokhoven H (2017). Exome sequencing of Pakistani consanguineous families identifies 30 novel candidate genes for recessive intellectual disability. Molecular Psychiatry, 22(11), 1604–1614. 10.1038/mp.2016.109 [DOI] [PMC free article] [PubMed] [Google Scholar]
  215. Richard EM, Polla DL, Assir MZ, Contreras M, Shahzad M, Khan AA, Razzaq A, Akram J, Tarar MN, Blanpied TA, Ahmed ZM, Abou Jamra R, Wieczorek D, van Bokhoven H, Riazuddin S, & Riazuddin S (2019). Bi-allelic variants in METTL5 cause autosomal-recessive intellectual disability and microcephaly. American Journal of Human Genetics, 105(4), 869–878. 10.1016/j.ajhg.2019.09.007 [DOI] [PMC free article] [PubMed] [Google Scholar]
  216. Ries RJ, Zaccara S, Klein P, Olarerin-George A, Namkoong S, Pickering BF, Patil DP, Kwak H, Lee JH, & Jaffrey SR (2019). m6A enhances the phase separation potential of mRNA. Nature, 571(7765), 424–428. 10.1038/s41586-019-1374-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  217. Rong B, Zhang Q, Wan J, Xing S, Dai R, Li Y, Cai J, Xie J, Song Y, Chen J, Zhang L, Yan G, Zhang W, Gao H, Han JDJ, Qu Q, Ma H, Tian Y, & Lan F (2020). Ribosome 18S m6A methyltransferase METTL5 promotes translation initiation and breast cancer cell growth. Cell Reports, 33(12), 108544. 10.1016/j.celrep.2020.108544 [DOI] [PubMed] [Google Scholar]
  218. Rottman FM, Bokar JA, Narayan P, Shambaugh ME, Ludwiczak R, Rottman FM, Bokar JA, Narayan P, Shambaugh ME, & Ludwiczak R (1994). N6-adenosine methylation in mRNA: Substrate specificity and enzyme complexity. Biochimie, 76(12), 1109–1114. 10.1016/0300-9084(94)90038-8 [DOI] [PubMed] [Google Scholar]
  219. Roundtree IA, Luo GZ, Zhang Z, Wang X, Zhou T, Cui Y, Sha J, Huang X, Guerrero L, Xie P, He E, Shen B, & He C (2017). YTHDC1 mediates nuclear export of N6-methyladenosine methylated mRNAs. eLife, 6, e31311. 10.7554/ELIFE.31311 [DOI] [PMC free article] [PubMed] [Google Scholar]
  220. Rubio RM, Depledge DP, Bianco C, Thompson L, & Mohr I (2018). RNA m6A modification enzymes shape innate responses to DNA by regulating interferon β. Genes and Development, 32(23–24), 1472–1484. 10.1101/gad.319475.118 [DOI] [PMC free article] [PubMed] [Google Scholar]
  221. Ruszkowska A, Ruszkowski M, Dauter Z, & Brown JA (2018). Structural insights into the RNA methyltransferase domain of METTL16. Scientific Reports, 8(1), 5311. 10.1038/s41598-018-23608-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  222. Sabnis RW (2021). Novel METTL3 modulators for treating acute myeloid leukemia (AML). ACS Medicinal Chemistry Letters, 12(7), 1061–1062. 10.1021/acsmedchemlett.1c00280 [DOI] [PMC free article] [PubMed] [Google Scholar]
  223. Scarborough AM, Flaherty JN, Hunter OV, Liu K, Kumar A, Xing C, Tu BP, & Conrad NK (2021). SAM homeostasis is regulated by CFIm- mediated splicing of MAT2A. eLife, 10, e64930. 10.7554/eLife.64930 [DOI] [PMC free article] [PubMed] [Google Scholar]
  224. Scheitl CPM, Ghaem Maghami M, Lenz AK, & Höbartner C (2020). Site-specific RNA methylation by a methyltransferase ribozyme. Nature, 587(7835), 663–667. 10.1038/s41586-020-2854-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Scheitl CPM, Mieczkowski M, Schindelin H, & Höbartner C (2022). Structure and mechanism of the methyltransferase ribozyme MTR1. Nature Chemical Biology, 18(5), 547–555. 10.1038/s41589-022-00976-x [DOI] [PMC free article] [PubMed] [Google Scholar]
  226. Schöller E, Weichmann F, Treiber T, Ringle S, Treiber N, Flatley A, Feederle R, Bruckmann A, & Meister G (2018). Interactions, localization, and phosphorylation of the m6A generating METTL3-METTL14–WTAP complex. RNA, 24(4), 499–512. 10.1261/rna.064063.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  227. Seidel-Rogol BL, McCulloch V, & Shadel GS (2003). Human mitochondrial transcription factor B1 methylates ribosomal RNA at a conserved stem-loop. Nature Genetics, 33(1), 23–24. 10.1038/ng1064 [DOI] [PubMed] [Google Scholar]
  228. Selberg S, Blokhina D, Aatonen M, Koivisto P, Siltanen A, Mervaala E, Kankuri E, & Karelson M (2019). Discovery of small molecules that activate RNA methylation through cooperative binding to the METTL3–14-WTAP complex active site. Cell Reports, 26(13), 3762–3771.e5. 10.1016/j.celrep.2019.02.100 [DOI] [PubMed] [Google Scholar]
  229. Selberg S, Žusinaite E, Herodes K, Seli N, Kankuri E, Merits A, & Karelson M (2021). HIV replication is increased by RNA methylation METTL3/METTL14/WTAP complex activators. ACS Omega, 6(24), 15957–15963. 10.1021/acsomega.1c01626 [DOI] [PMC free article] [PubMed] [Google Scholar]
  230. Sendinc E, Valle-Garcia D, Dhall A, Chen H, Henriques T, Navarrete-Perea J, Sheng W, Gygi SP, Adelman K, & Shi Y (2019). PCIF1 catalyzes m6Am mRNA methylation to regulate gene expression. Molecular Cell, 75(3), 620–630.e9. 10.1016/j.molcel.2019.05.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
  231. Sepich-Poore C, Zheng Z, Schmitt E, Wen K, Zhang ZS, Cui XL, Dai Q, Hu AC, Zhang L, Castillo AS, Tan H, Peng J, Zhuang X, He C, & Nachtergaele S (2022). The METTL5-TRMT112 N6-methyladenosine methyltransferase complex regulates mRNA translation via 18S rRNA methylation. Journal of Biological Chemistry, 298(3), 101590. 10.1016/j.jbc.2022.101590 [DOI] [PMC free article] [PubMed] [Google Scholar]
  232. Sergiev PV, Serebryakova MV, Bogdanov AA, & Dontsova OA (2008). The ybiN gene of Escherichia coli encodes adenine-N6 methyltransferase specific for modification of A1618 of 23 S ribosomal RNA, a methylated residue located close to the ribosomal exit tunnel. Journal of Molecular Biology, 375(1), 291–300. 10.1016/j.jmb.2007.10.051 [DOI] [PubMed] [Google Scholar]
  233. Shen C, Wang K, Deng X, & Chen J (2022). DNA N6-methyldeoxyadenosine in mammals and human disease. Trends in Genetics, 38(5), 454–467. 10.1016/j.tig.2021.12.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  234. Shen H, Stoute J, & Liu KF (2020). Structural and catalytic roles of the human 18S rRNA methyltransferases DIMT1 in ribosome assembly and translation. Journal of Biological Chemistry, 295(34), 12058–12070. 10.1074/jbc.ra120.014236 [DOI] [PMC free article] [PubMed] [Google Scholar]
  235. Shi H, Xu Y, Tian N, Yang M, & Liang FS (2022). Inducible and reversible RNA N6-methyladenosine editing. Nature Communications, 13(1), 1–10. 10.1038/s41467-022-29665-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  236. Shima H, Matsumoto M, Ishigami Y, Ebina M, Muto A, Sato Y, Kumagai S, Ochiai K, Suzuki T, & Igarashi K (2017). S-adenosylmethionine synthesis is regulated by selective N6-adenosine methylation and mRNA degradation involving METTL16 and YTHDC1. Cell Reports, 21(12), 3354–3363. 10.1016/j.celrep.2017.11.092 [DOI] [PubMed] [Google Scholar]
  237. Sikorski PJ, Warminski M, Kubacka D, Ratajczak T, Nowis D, Kowalska J, & Jemielity J (2020). The identity and methylation status of the first transcribed nucleotide in eukaryotic mRNA 5′ cap modulates protein expression in living cells. Nucleic Acids Research, 48(4), 1607–1626. 10.1093/nar/gkaa032 [DOI] [PMC free article] [PubMed] [Google Scholar]
  238. Sjöstedt E, Zhong W, Fagerberg L, Karlsson M, Mitsios N, Adori C, Oksvold P, Edfors F, Limiszewska A, Hikmet F, Huang J, Du Y, Lin L, Dong Z, Yang L, Liu X, Jiang H, Xu X, Wang J, … Mulder, J. (2020). An atlas of the protein-coding genes in the human, pig, and mouse brain. Science, 367(6482), eaay5947. 10.1126/science.aay5947 [DOI] [PubMed] [Google Scholar]
  239. Śledź P, & Jinek M (2016). Structural insights into the molecular mechanism of the m6A writer complex. eLife, 5(September), e18434. 10.7554/eLife.18434 [DOI] [PMC free article] [PubMed] [Google Scholar]
  240. Sommer S, Salditt-Georgieff M, Bachenheimer S, Darnell JE, Furuichi Y, Morgan M, & Shatkin AJ (1976). The methylation of adenovirus-specific nuclear and cytoplasmic RNA. Nucleic Acids Research, 3(3), 749–766. 10.1093/nar/3.3.749 [DOI] [PMC free article] [PubMed] [Google Scholar]
  241. Sood AJ, Viner C, & Hoffman MM (2019). DNamod: The DNA modification database. Journal of Cheminformatics, 11(1), 1–10. 10.1186/s13321-019-0349-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  242. Sorci M, Ianniello Z, Cruciani S, Larivera S, Ginistrelli LC, Capuano E, Marchioni M, Fazi F, & Fatica A (2018). METTL3 regulates WTAP protein homeostasis. Cell Death and Disease, 9(8), 1–12. 10.1038/s41419-018-0843-z [DOI] [PMC free article] [PubMed] [Google Scholar]
  243. Stixová L, Komůrková D, Kovaříková AS, Fagherazzi P, & Bártová E (2021). Localization of METTL16 at the nuclear periphery and the nucleolus is cell cycle-specific and METTL16 interacts with several nucleolar proteins. Life, 11(7), 669. 10.3390/life11070669 [DOI] [PMC free article] [PubMed] [Google Scholar]
  244. Su R, Dong L, Li Y, Gao M, He PC, Liu W, Wei J, Zhao Z, Gao L, Han L, Deng X, Li C, Prince E, Tan B, Qing Y, Qin X, Shen C, Xue M, Zhou K, … Chen J (2022). METTL16 exerts an m6A-independent function to facilitate translation and tumorigenesis. Nature Cell Biology, 24(2), 205–216. 10.1038/s41556-021-00835-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  245. Su S, Li S, Deng T, Gao M, Yin Y, Wu B, Peng C, Liu J, Ma J, & Zhang K (2022). Cryo-EM structures of human m6A writer complexes. Cell Research, 32(11), 982–994. 10.1038/s41422-022-00725-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  246. Sun HL, Zhu AC, Gao Y, Terajima H, Fei Q, Liu S, Zhang L, Zhang Z, Harada BT, He YY, Bissonnette MB, Hung MC, & He C (2020). Stabilization of ERK-phosphorylated METTL3 by USP5 increases m6A methylation. Molecular Cell, 80(4), 633–647.e7. 10.1016/j.molcel.2020.10.026 [DOI] [PMC free article] [PubMed] [Google Scholar]
  247. Sun S, Fei K, Zhang G, Wang J, Yang Y, Guo W, Yang Z, Wang J, Xue Q, Gao Y, & He J (2021). Construction and comprehensive analyses of a METTL5-associated prognostic signature with immune implication in lung adenocarcinomas. Frontiers in Genetics, 11, 617174. 10.3389/fgene.2020.617174 [DOI] [PMC free article] [PubMed] [Google Scholar]
  248. Sun X, Wang DO, & Wang J (2022). Targeted manipulation of m6A RNA modification through CRISPR-Cas-based strategies. Methods, 203, 56–61. 10.1016/j.ymeth.2022.03.006 [DOI] [PubMed] [Google Scholar]
  249. Tan B, & Gao SJ (2018). RNA epitranscriptomics: Regulation of infection of RNA and DNA viruses by N6-methyladenosine (m6A). Reviews in Medical Virology, 28(4), e1983. 10.1002/rmv.1983 [DOI] [PMC free article] [PubMed] [Google Scholar]
  250. Tan B, Liu H, Zhang S, Da Silva SR, Zhang L, Meng J, Cui X, Yuan H, Sorel O, Zhang SW, Huang Y, & Gao SJ (2017). Viral and cellular N6-methyladenosine and N6,2′-O-dimethyladenosine epitranscriptomes in the KSHV life cycle. Nature Microbiology, 3(1), 108–120. 10.1038/s41564-017-0056-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  251. Taoka M, Nobe Y, Yamaki Y, Sato K, Ishikawa H, Izumikawa K, Yamauchi Y, Hirota K, Nakayama H, Takahashi N, & Isobe T (2018). Landscape of the complete RNA chemical modifications in the human 80S ribosome. Nucleic Acids Research, 46(18), 9289–9298. 10.1093/nar/gky811 [DOI] [PMC free article] [PubMed] [Google Scholar]
  252. Tong J, Wang X, Liu Y, Ren X, Wang A, Chen Z, Yao J, Mao K, Liu T, Meng FL, Pan W, Zou Q, Liu J, Zhou Y, Xia Q, Flavell RA, Zhu S, & Li HB (2021). Pooled CRISPR screening identifies m6A as a positive regulator of macrophage activation. Science Advances, 7, eabd4742. 10.1126/sciadv.abd4742 [DOI] [PMC free article] [PubMed] [Google Scholar]
  253. Torun D, Arslan M, Çavdarlı B, Akar H, & Cram DS (2022). Three Afghani siblings with a novel homozygous variant and further delineation of the clinical features of METTL5 related intellectual disability syndrome. Turkish Journal of Pediatrics, 64(5), 956–963. 10.24953/turkjped.2020.3992 [DOI] [PubMed] [Google Scholar]
  254. Tsai K, Courtney DG, & Cullen BR (2018). Addition of m6A to SV40 late mRNAs enhances viral structural gene expression and replication. PLoS Pathogens, 14(2), e1006919. 10.1371/JOURNAL.PPAT.1006919 [DOI] [PMC free article] [PubMed] [Google Scholar]
  255. Van Tran N, Ernst FGM, Hawley BR, Zorbas C, Ulryck N, Hackert P, Bohnsack KE, Bohnsack MT, Jaffrey SR, Graille M, & Lafontaine DLJ (2019). The human 18S rRNA m6A methyltransferase METTL5 is stabilized by TRMT112. Nucleic Acids Research, 47(15), 7719–7733. 10.1093/nar/gkz619 [DOI] [PMC free article] [PubMed] [Google Scholar]
  256. Viegas IJ, de Macedo JP, Serra L, De Niz M, Temporão A, Silva Pereira S, Mirza AH, Bergstrom E, Rodrigues JA, Aresta-Branco F, Jaffrey SR, & Figueiredo LM (2022). N6-methyladenosine in poly(A) tails stabilize VSG transcripts. Nature, 604(7905), 362–370. 10.1038/s41586-022-04544-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
  257. Visvanathan A, Patil V, Arora A, Hegde AS, Arivazhagan A, Santosh V, & Somasundaram K (2018). Essential role of METTL3-mediated m 6 A modification in glioma stem-like cells maintenance and radioresistance. Oncogene, 37(4), 522–533. 10.1038/onc.2017.351 [DOI] [PubMed] [Google Scholar]
  258. Vu LP, Pickering BF, Cheng Y, Zaccara S, Nguyen D, Minuesa G, Chou T, Chow A, Saletore Y, Mackay M, Schulman J, Famulare C, Patel M, Klimek VM, Garrett-Bakelman FE, Melnick A, Carroll M, Mason CE, Jaffrey SR, & Kharas MG (2017). The N6-methyladenosine (m6A)-forming enzyme METTL3 controls myeloid differentiation of normal hematopoietic and leukemia cells. Nature Medicine, 23(11), 1369–1376. 10.1038/nm.4416 [DOI] [PMC free article] [PubMed] [Google Scholar]
  259. Wang CX, Cui GS, Liu X, Xu K, Wang M, Zhang XX, Jiang LY, Li A, Yang Y, Lai WY, Sun BF, Jiang GB, Wang HL, Tong WM, Li W, Wang XJ, Yang YG, & Zhou Q (2018). METTL3-mediated m6A modification is required for cerebellar development. PLoS Biology, 16(6), e2004880. 10.1371/journal.pbio.2004880 [DOI] [PMC free article] [PubMed] [Google Scholar]
  260. Wang F, Zhang J, Lin X, Yang L, Zhou Q, Mi X, Li Q, Wang S, Li D, Liu XM, & Zhou J (2023). METTL16 promotes translation and lung tumorigenesis by sequestering cytoplasmic eIF4E2. Cell Reports, 42(3), 112150. 10.1016/j.celrep.2023.112150 [DOI] [PubMed] [Google Scholar]
  261. Wang M, Liu J, Zhao Y, He R, Xu X, Guo X, Li X, Xu S, Miao J, Guo J, Zhang H, Gong J, Zhu F, Tian R, Shi C, Peng F, Feng Y, Yu S, Xie Y, … Qin R (2020). Upregulation of METTL14 mediates the elevation of PERP mRNA N6 adenosine methylation promoting the growth and metastasis of pancreatic cancer. Molecular Cancer, 19(1), 1–15. 10.1186/s12943-020-01249-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  262. Wang P, Doxtader KA, & Nam Y (2016). Structural basis for cooperative function of Mettl3 and Mettl14 methyltransferases. Molecular Cell, 63(2), 306–317. 10.1016/j.molcel.2016.05.041 [DOI] [PMC free article] [PubMed] [Google Scholar]
  263. Wang P, Wang X, Zheng L, & Zhuang C (2020). Gene signatures and prognostic values of m6A regulators in hepatocellular carcinoma. Frontiers in Genetics, 11, 1211. 10.3389/fgene.2020.540186 [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. Wang S, Fan X, Zhu J, Xu D, Li R, Chen R, Hu J, Shen Y, Hao J, Wang K, Jiang X, Wang Y, Jiang Y, Li J, & Zhang J (2021). The differentiation of colorectal cancer is closely relevant to m6A modification. Biochemical and Biophysical Research Communications, 546, 65–73. 10.1016/j.bbrc.2021.02.001 [DOI] [PubMed] [Google Scholar]
  265. Wang T, Birsoy K, Hughes NW, Krupczak KM, Post Y, Wei JJ, Lander ES, & Sabatini DM (2015). Identification and characterization of essential genes in the human genome. Science, 350(6264), 1096–1101. 10.1126/science.aac7041 [DOI] [PMC free article] [PubMed] [Google Scholar]
  266. Wang X, Feng J, Xue Y, Guan Z, Zhang D, Liu Z, Gong Z, Wang Q, Huang J, Tang C, Zou T, & Yin P (2016). Structural basis of N6-adenosine methylation by the METTL3-METTL14 complex. Nature, 534(7608), 575–578. 10.1038/nature18298 [DOI] [PubMed] [Google Scholar]
  267. Wang Y, & Jia G (2020). Detection methods of epitranscriptomic mark N6-methyladenosine. Essays in Biochemistry, 64(6), 967–979. 10.1042/EBC20200039 [DOI] [PubMed] [Google Scholar]
  268. 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(7481), 117–120. 10.1038/nature12730 [DOI] [PMC free article] [PubMed] [Google Scholar]
  269. Wang Y, Gao M, Zhu F, Li X, Yang Y, Yan Q, Jia L, Xie L, & Chen Z (2020). METTL3 is essential for postnatal development of brown adipose tissue and energy expenditure in mice. Nature Communications, 11(1), 1–14. 10.1038/s41467-020-15488-2 [DOI] [PMC free article] [PubMed] [Google Scholar]
  270. Wang Y, Li Y, Toth JI, Petroski MD, Zhang Z, & Zhao JC (2014). N6-methyladenosine modification destabilizes developmental regulators in embryonic stem cells. Nature Cell Biology, 16(2), 191–198. 10.1038/ncb2902 [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Wang Y, Li Y, Yue M, Wang J, Kumar S, Wechsler-Reya RJ, Zhang Z, Ogawa Y, Kellis M, Duester G, & Zhao JC (2018). N6-methyladenosine RNA modification regulates embryonic neural stem cell self-renewal through histone modifications. Nature Neuroscience, 21(2), 195–206. 10.1038/s41593-017-0057-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  272. Wang W, Sun B, Xia Y, Sun S, & He C (2020). RNA N6-methyladenosine-related gene contribute to clinical prognostic impact on patients with liver cancer. Frontiers in Genetics, 11, 306. 10.3389/fgene.2020.00306 [DOI] [PMC free article] [PubMed] [Google Scholar]
  273. Wang XK, Zhang YW, Wang CM, Li B, Zhang TZ, Zhou WJ, Cheng L-J, Huo MY, Zhang CH, & He YL (2021). METTL16 promotes cell proliferation by up-regulating cyclin D1 expression in gastric cancer. Journal of Cellular and Molecular Medicine, 25(14), 6602–6617. 10.1111/jcmm.16664 [DOI] [PMC free article] [PubMed] [Google Scholar]
  274. Wanna-udom S, Terashima M, Lyu H, Ishimura A, Takino T, Sakari M, Tsukahara T, & Suzuki T (2020). The m6A methyltransferase METTL3 contributes to transforming growth factor-beta-induced epithelial-mesenchymal transition of lung cancer cells through the regulation of JUNB. Biochemical and Biophysical Research Communications, 524(1), 150–155. 10.1016/j.bbrc.2020.01.042 [DOI] [PubMed] [Google Scholar]
  275. Warda AS, Kretschmer J, Hackert P, Lenz C, Urlaub H, Höbartner C, Sloan KE, & Bohnsack MT (2017). Human METTL16 is a N6-methyladenosine (m6A) methyltransferase that targets pre-mRNAs and various non-coding RNAs. EMBO Reports, 18(11), 2004–2014. 10.15252/embr.201744940 [DOI] [PMC free article] [PubMed] [Google Scholar]
  276. Wassarman DA, & Steitz JA (1992). Interactions of small nuclear RNA’s with precursor messenger RNA during in vitro splicing. Science, 257(5078), 1918–1925. 10.1126/science.1411506 [DOI] [PubMed] [Google Scholar]
  277. Watabe E, Togo-Ohno M, Ishigami Y, Wani S, Hirota K, Kimura-Asami M, Hasan S, Takei S, Fukamizu A, Suzuki Y, Suzuki T, & Kuroyanagi H (2021). m6A-mediated alternative splicing coupled with nonsense-mediated mRNA decay regulates SAM synthetase homeostasis. The EMBO Journal, 40(14), e106434. 10.15252/embj.2020106434 [DOI] [PMC free article] [PubMed] [Google Scholar]
  278. Wei CM, Gershowitz A, & Moss B (1975). Methylated nucleotides block 5′ terminus of HeLa cell messenger RNA. Cell, 4(4), 379–386. 10.1016/0092-8674(75)90158-0 [DOI] [PubMed] [Google Scholar]
  279. Wei CM, & Moss B (1977). Nucleotide sequences at the N6-methyladenosine sites of HeLa cell messenger ribonucleic acid. Biochemistry, 16(8), 1672–1676. 10.1021/bi00627a023 [DOI] [PubMed] [Google Scholar]
  280. Wen J, Lv R, Ma H, Shen H, He C, Wang J, Jiao F, Liu H, Yang P, Tan L, Lan F, Shi YG, He C, Shi Y, & Diao J (2018). Zc3h13 regulates nuclear RNA m6A methylation and mouse embryonic stem cell self-renewal. Molecular Cell, 69(6), 1028–1038.e6. 10.1016/j.molcel.2018.02.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  281. Weng H, Huang H, Wu H, Qin X, Zhao BS, Dong L, Shi H, Skibbe J, Shen C, Hu C, Sheng Y, Wang Y, Wunderlich M, Zhang B, Dore LC, Su R, Deng X, Ferchen K, Li C, … Chen J (2018). METTL14 inhibits hematopoietic stem/progenitor differentiation and promotes leukemogenesis via mRNA m6A modification. Cell Stem Cell, 22(2), 191–205.e9. 10.1016/j.stem.2017.11.016 [DOI] [PMC free article] [PubMed] [Google Scholar]
  282. Wilson C, Chen PJ, Miao Z, & Liu DR (2020). Programmable m6A modification of cellular RNAs with a Cas13-directed methyltransferase. Nature Biotechnology, 38(12), 1431–1440. 10.1038/s41587-020-0572-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
  283. Woodcock CB, Yu D, Hajian T, Li J, Huang Y, Dai N, Corrêa IR, Wu T, Vedadi M, Zhang X, & Cheng X (2019). Human MettL3–MettL14 complex is a sequence-specific DNA adenine methyltransferase active on single-strand and unpaired DNA in vitro. Cell Discovery, 5(1), 1–4. 10.1038/s41421-019-0136-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  284. Wu KJ (2020). The epigenetic roles of DNA N6-methyladenine (6mA) modification in eukaryotes. Cancer Letters, 494, 40–46. 10.1016/j.canlet.2020.08.025 [DOI] [PubMed] [Google Scholar]
  285. Wu Y, Xie L, Wang M, Xiong Q, Guo Y, Liang Y, Li J, Sheng R, Deng P, Wang Y, Zheng R, Jiang Y, Ye L, Chen Q, Zhou X, Lin S, & Yuan Q (2018). Mettl3-mediated m6A RNA methylation regulates the fate of bone marrow mesenchymal stem cells and osteoporosis. Nature Communications, 9(1), 1–12. 10.1038/s41467-018-06898-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  286. Wu Z, Shi Y, Lu M, Song M, Yu Z, Wang J, Wang S, Ren J, Yang YG, Liu GH, Zhang W, Ci W, & Qu J (2020). METTL3 counteracts premature aging via m6A-dependent stabilization of MIS12 mRNA. Nucleic Acids Research, 48(19), 11083–11096. 10.1093/nar/gkaa816 [DOI] [PMC free article] [PubMed] [Google Scholar]
  287. Wyde PR, Ambrose MW, Meyer HL, Zolinski CL, & Gilbert BE (1990). Evaluation of the toxicity and antiviral activity of carbocyclic 3-deazaadenosine against respiratory syncytial and parainfluenza type 3 viruses in tissue culture and in cotton rats. Antiviral Research, 14(4–5), 215–225. 10.1016/0166-3542(90)90003-P [DOI] [PubMed] [Google Scholar]
  288. Xia T, Wu X, Cao M, Zhang P, Shi G, Zhang J, Lu Z, Wu P, Cai B, Miao Y, & Jiang K (2019). The RNA m6A methyltransferase METTL3 promotes pancreatic cancer cell proliferation and invasion. Pathology Research and Practice, 215(11), 152666. 10.1016/j.prp.2019.152666 [DOI] [PubMed] [Google Scholar]
  289. Xia Z, Tang M, Ma J, Zhang H, Gimple RC, Prager BC, Tang H, Sun C, Liu F, Lin P, Mei Y, Du R, Rich JN, & Xie Q (2021). Epitranscriptomic editing of the RNA N6-methyladenosine modification by dCasRx conjugated methyltransferase and demethylase. Nucleic Acids Research, 49(13), 7361–7374. 10.1093/nar/gkab517 [DOI] [PMC free article] [PubMed] [Google Scholar]
  290. Xia Z, Xu J, Lu E, He W, Deng S, Gong AY, Strass-Soukup J, Martins GA, Lu G, & Chen XM (2021). m6A mRNA methylation regulates epithelial innate antimicrobial defense against Cryptosporidial infection. Frontiers in Immunology, 12, 2710. 10.3389/fimmu.2021.705232 [DOI] [PMC free article] [PubMed] [Google Scholar]
  291. Xiang Y, Laurent B, Hsu CH, Nachtergaele S, Lu Z, Sheng W, Xu C, Chen H, Ouyang J, Wang S, Ling D, Hsu PH, Zou L, Jambhekar A, He C, & Shi Y (2017). RNA m6A methylation regulates the ultraviolet-induced DNA damage response. Nature, 543(7646), 573–576. 10.1038/nature21671 [DOI] [PMC free article] [PubMed] [Google Scholar]
  292. Xiao S, Cao S, Huang Q, Xia L, Deng M, Yang M, Jia G, Liu X, Shi J, Wang W, Li Y, Liu S, Zhu H, Tan K, Luo Q, Zhong M, He C, & Xia L (2019). The RNA N6-methyladenosine modification landscape of human fetal tissues. Nature Cell Biology, 21(5), 651–661. 10.1038/s41556-019-0315-4 [DOI] [PubMed] [Google Scholar]
  293. Xiao S, Guo S, Han J, Sun Y, Wang M, Chen Y, Fang X, Yang F, Mu Y, Zhang L, Ding Y, Zhang N, Jiang H, Chen K, Zhao K, Luo C, & Chen S (2022). High-throughput-methyl-reading (HTMR) assay: A solution based on nucleotide methyl-binding proteins enables large-scale screening for DNA/RNA methyltransferases and demethylases. Nucleic Acids Research, 50(2), E9. 10.1093/nar/gkab989 [DOI] [PMC free article] [PubMed] [Google Scholar]
  294. Xing M, Liu Q, Mao C, Zeng H, Zhang X, Zhao S, Chen L, Liu M, Shen B, Guo X, Ma H, Chen H, & Zhang J (2020). The 18S rRNA m6A methyltransferase METTL5 promotes mouse embryonic stem cell differentiation. EMBO Reports, 21(10), e49863. 10.15252/embr.201949863 [DOI] [PMC free article] [PubMed] [Google Scholar]
  295. Xu C, Liu K, Ahmed H, Loppnau P, Schapira M, & Min J (2015). Structural basis for the discriminative recognition of N6-methyladenosine RNA by the human YT521-B homology domain family of proteins. Journal of Biological Chemistry, 290(41), 24902–24913. 10.1074/jbc.M115.680389 [DOI] [PMC free article] [PubMed] [Google Scholar]
  296. Xu C, Wang X, Liu K, Roundtree IA, Tempel W, Li Y, Lu Z, He C, & Min J (2014). Structural basis for selective binding of m6A RNA by the YTHDC1 YTH domain. Nature Chemical Biology, 10(11), 927–929. 10.1038/nchembio.1654 [DOI] [PubMed] [Google Scholar]
  297. Xu W, He C, Kaye EG, Li J, Mu M, Nelson GM, Dong L, Wang J, Wu F, Shi YGY, Adelman K, Lan F, Shi YGY, & Shen H (2022). Dynamic control of chromatin-associated m6A methylation regulates nascent RNA synthesis. Molecular Cell, 82(6), 1156–1168.e7. 10.1016/j.molcel.2022.02.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
  298. Xu W, Li J, He C, Wen J, Ma H, Rong B, Diao J, Wang L, Wang J, Wu F, Tan L, Shi YGY, Shi YGY, & Shen H (2021). METTL3 regulates heterochromatin in mouse embryonic stem cells. Nature, 591(7849), 317–321. 10.1038/s41586-021-03210-1 [DOI] [PubMed] [Google Scholar]
  299. Xu W, Liu S, Zhang G, Liu J, & Cao G (2022). Knockdown of METTL5 inhibits the Myc pathway to downregulate PD-L1 expression and inhibits immune escape of hepatocellular carcinoma cells. Journal of Chemotherapy, 1–10, 1–10. 10.1080/1120009X.2022.2143614 [DOI] [PubMed] [Google Scholar]
  300. Xue M, Zhao BS, Zhang Z, Lu M, Harder O, Chen P, Lu Z, Li A, Ma Y, Xu Y, Liang X, Zhou J, Niewiesk S, Peeples ME, He C, & Li J (2019). Viral N6-methyladenosine upregulates replication and pathogenesis of human respiratory syncytial virus. Nature Communications, 10(1), 1–18. 10.1038/s41467-019-12504-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  301. Yan X, Pei K, Guan Z, Liu F, Yan J, Jin X, Wang Q, Hou M, Tang C, & Yin P (2022). AI-empowered integrative structural characterization of m6A methyltransferase complex. Cell Research, 32(12), 1124–1127. 10.1038/s41422-022-00741-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
  302. Yan X, Zhao X, Yan Q, Wang Y, & Zhang C (2021). Analysis of the role of METTL5 as a hub gene in lung adenocarcinoma based on a weighted gene co-expression network. Mathematical Biosciences and Engineering, 18(5), 6608–6619. 10.3934/mbe.2021327 [DOI] [PubMed] [Google Scholar]
  303. Yang X, Triboulet R, Liu Q, Sendinc E, & Gregory RI (2022). Exon junction complex shapes the m6A epitranscriptome. Nature Communications, 13(1), 1–12. 10.1038/s41467-022-35643-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
  304. Yang Y, Fan X, Mao M, Song X, Wu P, Zhang Y, Jin Y, Yang Y, Chen LL, Wang Y, Wong CCL, Xiao X, & Wang Z (2017). Extensive translation of circular RNAs driven by N6-methyladenosine. Cell Research, 27(5), 626–641. 10.1038/cr.2017.31 [DOI] [PMC free article] [PubMed] [Google Scholar]
  305. Yankova E, Blackaby W, Albertella M, Rak J, De Braekeleer E, Tsagkogeorga G, Pilka ES, Aspris D, Leggate D, Hendrick AG, Webster NA, Andrews B, Fosbeary R, Guest P, Irigoyen N, Eleftheriou M, Gozdecka M, Dias JML, Bannister AJ, … Kouzarides T (2021). Small-molecule inhibition of METTL3 as a strategy against myeloid leukaemia. Nature, 593(7860), 597–601. 10.1038/s41586-021-03536-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  306. Yao M-D, Jiang Q, Ma Y, Liu C, Zhu CY, Sun YN, Shan K, Ge HM, Zhang QY, Zhang HY, Yao J, Li XM, & Yan B (2020). Role of METTL3-dependent N6-methyladenosine mRNA modification in the promotion of angiogenesis. Molecular Therapy, 28(10), 2191–2202. 10.1016/j.ymthe.2020.07.022 [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Ye C, & Tu BP (2018). Sink into the epigenome: Histones as repositories that influence cellular metabolism. Trends in Endocrinology and Metabolism, 29(9), 626–637. 10.1016/j.tem.2018.06.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
  308. Ye F. (2017). RNA N6-adenosine methylation (m6A) steers epitranscriptomic control of herpesvirus replication. Inflammation and Cell Signaling, 4(3), e1604. 10.14800/ics.1604 [DOI] [PMC free article] [PubMed] [Google Scholar]
  309. Ye F, Chen ER, & Nilsen TW (2017). Kaposi’s sarcoma-associated herpesvirus utilizes and manipulates RNA N6-adenosine methylation to promote lytic replication. Journal of Virology, 91(16), e00466–17. 10.1128/jvi.00466-17 [DOI] [PMC free article] [PubMed] [Google Scholar]
  310. Ye F, Wu J, & Zhang F (2023). METTL16 epigenetically enhances GPX4 expression via m6A modification to promote breast cancer progression by inhibiting ferroptosis. Biochemical and Biophysical Research Communications, 638, 1–6. 10.1016/j.bbrc.2022.10.065 [DOI] [PubMed] [Google Scholar]
  311. Yoon KJ, Ringeling FR, Vissers C, Jacob F, Pokrass M, Jimenez-Cyrus D, Su Y, Kim NS, Zhu Y, Zheng L, Kim S, Wang X, Doré LC, Jin P, Regot S, Zhuang X, Canzar S, He C, Ming G-L, & Song H (2017). Temporal control of mammalian cortical neurogenesis by m6A methylation. Cell, 171(4), 877–889.e17. 10.1016/j.cell.2017.09.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
  312. Yoshinaga M, Han K, Morgens DW, Horii T, Kobayashi R, Tsuruyama T, Hia F, Yasukura S, Kajiya A, Cai T, Cruz PHC, Vandenbon A, Suzuki Y, Kawahara Y, Hatada I, Bassik MC, & Takeuchi O (2022). The N6-methyladenosine methyltransferase METTL16 enables erythropoiesis through safeguarding genome integrity. Nature Communications, 13(1), 1–17. 10.1038/s41467-022-34078-y [DOI] [PMC free article] [PubMed] [Google Scholar]
  313. Yu D, Horton JR, Yang J, Hajian T, Vedadi M, Sagum CA, Bedford MT, Blumenthal RM, Zhang X, & Cheng X (2021). Human MettL3-MettL14 RNA adenine methyltransferase complex is active on double-stranded DNA containing lesions. Nucleic Acids Research, 49(20), 11629–11642. 10.1093/nar/gkab460 [DOI] [PMC free article] [PubMed] [Google Scholar]
  314. Yu D, Kaur G, Blumenthal RM, Zhang X, & Cheng X (2021). Enzymatic characterization of three human RNA adenosine methyltransferases reveals diverse substrate affinities and reaction optima. Journal of Biological Chemistry, 296, 100270. 10.1016/j.jbc.2021.100270 [DOI] [PMC free article] [PubMed] [Google Scholar]
  315. Zaccara S, Ries RJ, & Jaffrey SR (2019). Reading, writing and erasing mRNA methylation. Nature Reviews Molecular Cell Biology, 20(10), 608–624. 10.1038/s41580-019-0168-5 [DOI] [PubMed] [Google Scholar]
  316. Zeng C, Huang W, Li Y, & Weng H (2020). Roles of METTL3 in cancer: Mechanisms and therapeutic targeting. Journal of Hematology and Oncology, 13(1), 1–15. 10.1186/s13045-020-00951-w [DOI] [PMC free article] [PubMed] [Google Scholar]
  317. Zeng X, Zhao F, Cui G, Zhang Y, Deshpande RA, Chen Y, Deng M, Kloeber JA, Shi Y, Zhou Q, Zhang C, Hou J, Kim W, Tu X, Yan Y, Xu Z, Chen L, Gao H, Guo G, … Lou Z (2022). METTL16 antagonizes MRE11-mediated DNA end resection and confers synthetic lethality to PARP inhibition in pancreatic ductal adenocarcinoma. Nature Cancer, 3(9), 1088–1104. 10.1038/s43018-022-00429-3 [DOI] [PubMed] [Google Scholar]
  318. Zhang B, Gu Y, & Jiang G (2020). Expression and prognostic characteristics of m6A RNA methylation regulators in breast cancer. Frontiers in Genetics, 11, 604597. 10.3389/fgene.2020.604597 [DOI] [PMC free article] [PubMed] [Google Scholar]
  319. Zhang Y, Yao Y, Qi X, Li J, Liu M, Che X, Xu Y, & Wu G (2021). Identification of a new prognostic risk signature of clear cell renal cell carcinoma based on N6-methyladenosine RNA methylation regulators. Journal of Immunology Research, 2021, 6617841. 10.1155/2021/6617841 [DOI] [PMC free article] [PubMed] [Google Scholar]
  320. Zhang Z, Chen LQ, Zhao YL, Yang CG, Roundtree IA, Zhang Z, Ren J, Xie W, He C, & Luo GZ (2019). Single-base mapping of m6A by an antibody-independent method. Science Advances, 5(7), 250–253. 10.1126/sciadv.aax0250 [DOI] [PMC free article] [PubMed] [Google Scholar]
  321. Zhang Z, Wang M, Xie D, Huang Z, Zhang L, Yang Y, Ma D, Li W, Zhou Q, Yang YG, & Wang XJ (2018). METTL3-mediated N6-methyladenosine mRNA modification enhances long-term memory consolidation. Cell Research, 28(11), 1050–1061. 10.1038/s41422-018-0092-9 [DOI] [PMC free article] [PubMed] [Google Scholar]
  322. Zhang Z, Zhang C, Luo Y, Wu P, Zhang G, Zeng Q, Wang L, Yang Z, Xue L, Zheng B, Zeng H, Tan F, Xue Q, Gao S, Sun N, & He J (2021). m6A regulator expression profile predicts the prognosis, benefit of adjuvant chemotherapy, and response to anti-PD-1 immunotherapy in patients with small-cell lung cancer. BioMed Central Medicine, 19(1), 284. 10.1186/s12916-021-02148-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
  323. Zhang Z, Zhang C, Yang Z, Zhang G, Wu P, Luo Y, Zeng Q, Wang L, Xue Q, Zhang Y, Sun N, & He J (2021). m6A regulators as predictive biomarkers for chemotherapy benefit and potential therapeutic targets for overcoming chemotherapy resistance in small-cell lung cancer. Journal of Hematology and Oncology, 14(1), 190. 10.1186/s13045-021-01173-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
  324. Zhao BS, Nachtergaele S, Roundtree IA, & He C (2018). Our views of dynamic N6-methyladenosine RNA methylation. RNA, 24(3), 268–272. 10.1261/rna.064295.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
  325. Zhao H, Xu Y, Xie Y, Zhang L, Gao M, Li S, & Wang F (2021). m6A regulators is differently expressed and correlated with immune response of esophageal cancer. Frontiers in Cell and Developmental Biology, 9, 650023. 10.3389/fcell.2021.650023 [DOI] [PMC free article] [PubMed] [Google Scholar]
  326. Zhao S, Lu J, Chen Y, Wang Z, Cao J, & Dong Y (2021). Exploration of the potential roles of m6A regulators in the uterus in pregnancy and infertility. Journal of Reproductive Immunology, 146, 103341. 10.1016/j.jri.2021.103341 [DOI] [PubMed] [Google Scholar]
  327. Zheng G, Dahl JA, Niu Y, Fedorcsak P, Huang CM, Li CJ, Vågbø CB, Shi Y, Wang WL, Song SH, Lu Z, Bosmans RPG, Dai Q, Hao YJ, Yang X, Zhao WM, Tong WM, Wang XJ, Bogdan F, … He C (2013). ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Molecular Cell, 49(1), 18–29. 10.1016/j.molcel.2012.10.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
  328. Zhou KI, Shi H, Lyu R, Wylder AC, Matuszek Ż, Pan JN, He C, Parisien M, & Pan T (2019). Regulation of Co-transcriptional pre-mRNA splicing by m6A through the low-complexity protein hnRNPG. Molecular Cell, 76(1), 70–81.e9. 10.1016/j.molcel.2019.07.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
  329. Zhu W, Wang JZ, Xu Z, Cao M, Hu Q, Pan C, Guo M, Wei JF, & Yang H (2019). Detection of N6-methyladenosine modification residues (Review). International Journal of Molecular Medicine, 43(6), 2267–2278. 10.3892/ijmm.2019.4169 [DOI] [PMC free article] [PubMed] [Google Scholar]
  330. Zorbas C, Nicolas E, Wacheul L, Huvelle E, Heurgué-Hamard V, & Lafontaine DLJ (2015). The human 18S rRNA base methyltransferases DIMT1L and WBSCR22-TRMT112 but not rRNA modification are required for ribosome biogenesis. Molecular Biology of the Cell, 26(11), 2080–2095. 10.1091/mbc.E15-02-0073 [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

RESOURCES