This review summarizes current knowledge and future directions on the mechanisms that govern the interplay between m6A RNA modification, RNA folding, and protein–RNA interactions, focusing on the mechanisms and available quantitative information.
Abstract
Protein–RNA complexes exist in many forms within the cell, from stable machines such as the ribosome to transient assemblies like the spliceosome. All protein–RNA assemblies rely on spatially and temporally coordinated interactions between specific proteins and RNAs to achieve a functional form. RNA folding and structure are often critical for successful protein binding and protein–RNA complex formation. RNA modifications change the chemical nature of a given RNA and often alter its folding kinetics. Both these alterations can affect how and if proteins or other RNAs can interact with the modified RNA and assemble into complexes. N6-methyladenosine (m6A) is the most common base modification on mRNAs and regulatory noncoding RNAs and has been shown to impact RNA structure and directly modulate protein–RNA interactions. In this review, focusing on the mechanisms and available quantitative information, we discuss first how the METTL3/14 m6A writer complex is specifically targeted to RNA assisted by protein–RNA and other interactions to enable site-specific and co-transcriptional RNA modification and, once introduced, how the m6A modification affects RNA folding and protein–RNA interactions.
Introduction
Protein–RNA complex assemblies are at the core of many essential cellular machineries, such as the spliceosome and the ribosome. In recent years, protein–RNA complexes have gained importance in more cellular processes, and studying them has become increasingly relevant to understanding the regulation of many cellular processes.
Similarly, since the discovery of the first modified nucleotide (pseudouridine) in yeast tRNA (Davis & Allen, 1957), the list of RNA modifications has not only grown in number and diversity but also their impact and role in regulating cellular processes have increasingly been appreciated and have become an integral factor in how RNAs and RNPs function and, for the latter, also how they assemble (Gilbert & Nachtergaele, 2023).
New and advancing technologies enable the discovery and characterization of currently around 170 chemically distinct RNA modifications on highly abundant RNAs such as ribosomal RNA (rRNA) and on rare and short-lived species including mRNAs, a subset of long noncoding RNAs (lncRNAs) and miRNAs (Dunin-Horkawicz et al, 2006; Czerwoniec et al, 2009; Machnicka et al, 2013; Boccaletto et al, 2018, 2022). These modifications and the enzymatic processes involved have context-dependent consequences on the affected RNA species, including effects on structure, stability, and ability to bind protein interaction partners (Watkins & Bohnsack, 2012; Motorin & Helm, 2022).
N6-methyladenosine (m6A) is among the best-studied internal RNA modifications found mainly on RNA polymerase II (Pol II) transcripts, including mRNAs, lncRNA, and primary microRNAs (pri-miRNA) (Alarcón et al, 2015b; Ke et al, 2017; Knuckles et al, 2017; Slobodin et al, 2017; Hong et al, 2022). Within mRNAs, m6A modifications are detected predominately in long introns and exons, terminal exons, and around stop codons (Dominissini et al, 2012; Meyer et al, 2012; Ke et al, 2015). m6A is prominently involved in almost all stages of gene expression, including transcriptional regulation, splicing, mRNA export, mRNA degradation, and translation (Meyer et al, 2015; Xiao et al, 2016; Barbieri et al, 2017; Roundtree et al, 2017; Slobodin et al, 2017; Zaccara et al, 2019).
Installation of the m6A modifications happens mainly co-transcriptionally by functionally distinct methyltransferases, also termed m6A writers, which each display different substrate specificities (Ke et al, 2017; Knuckles et al, 2017; Slobodin et al, 2017; Warda et al, 2017; Louloupi et al, 2018; Akichika et al, 2019; Ma et al, 2019; Sendinc et al, 2019; van Tran et al, 2019; Lee et al, 2021; Xu et al, 2022). The heterodimeric complex formed by methyltransferase-like protein 3 (METTL3) and METTL14 has a relatively broad range of substrates. It is responsible for most of the m6A marks on Pol II transcripts (Balacco & Soller, 2019). Nevertheless, METTL3/METTL14 requires the coordinated assembly of a complex multi-component RNP to ensure recruitment to target sites and correct modification of individual RNAs (Ping et al, 2014; Schöller et al, 2018; Wen et al, 2018; Yue et al, 2018; Balacco & Soller, 2019; Bawankar et al, 2021; Sepich-Poore et al, 2022; Su et al, 2022).
Other m6A-methyltransferases such as METTL16, METTL5/TRMT112, ZCCHC4, and PCIF/CAPAM have narrower substrate requirements and methylate only a very specific set of target RNAs and sequence contexts, but their recruitment to target sites relies, in most cases, on often short-lived but highly coordinated protein–RNA complex assemblies (Table 1) (Warda et al, 2017; Akichika et al, 2019; Ma et al, 2019; Sendinc et al, 2019; van Tran et al, 2019; Sendinc & Shi, 2023). These specialized m6A methyltransferases have been reviewed elsewhere (Huang et al, 2021; Yu et al, 2021b; Satterwhite & Mansfield, 2022) and will not be discussed here.
Table 1.
Substrate specificities and kinetic parameters in m6A RNA modification for writers, readers, and erasers.
| Writers | Substrate | Km (Substrate) | Km (SAM) | KD(1) | KD(2) | kcat | References |
|---|---|---|---|---|---|---|---|
| METTL3/METTL14 | DRACH (D = A,G,U; R = A,G; H = A,C,U) | 22 ± 2 nM | 102 ± 15 nM | n/a | n/a | 18 ± 2 h−1 | Liu et al (2014) and Li et al (2016) |
| WTAP/VIRMA | human ACTB mRNA fragment | n/a | n/a | 826.3 ± 336.0 nM | n/a | n/a | Su et al (2022) |
| WTAP/VIRMA/HAKAI | human ACTB mRNA fragment | n/a | n/a | 562.7 ± 76.6 nM | n/a | n/a | Su et al (2022) |
| WTAP/VIRMA/ZC3H13 | human ACTB mRNA fragment | n/a | n/a | 256.3 ± 27.9 nM | n/a | n/a | Su et al (2022) |
| WTAP/VIRMA/ZC3H13/HAKAI | human ACTB mRNA fragment | n/a | n/a | 214.0 ± 2.6 nM | n/a | n/a | Su et al (2022) |
| METTL16 | UACAGARAA (U6 snRNA, MALAT1) | ˜10 µM | >0.4 mM | 18 ± 7 µM | 126 ± 6 µM | 0.07 ± 0.02 min−1 | Warda et al (2017), Yu et al (2021b), and Breger and Brown (2023) |
| METTL5/TRMT112 | A1832 of 18S rRNA | 1.1 ± 0.2 μM (pH 8.0) | 1.0 ± 0.2 μM (pH 8.0) | n/a | n/a | 13.1 ± 0.8 h−1 (pH 8.0) | van Tran et al (2019) and Yu et al (2021b) |
| ZCCHC4 | A4220 of 28S rRNA | n/a | 6.7 μM | n/a | n/a | n/a | Ren et al (2019) |
| PCIF/CAPAM | m7GpppAm | 3.5 ± 0.71/0.3 ± 0.03 μM (pH 8.0) | 0.65 ± 0.05 μM (pH 8.0) | n/a | n/a | 0.67 ± 0.01 min−1 (pH 8.0) | Akichika et al (2019), Boulias et al (2019), and Yu et al (2021b) |
| Readers | References | ||||||
| YTHDC1 | G G m6A C/G G/A/U | n/a | n/a | 0.39 ± 0.071 µM | n/a | n/a | Arguello et al (2019) |
| YTHDC2 | G G m6A C/U A/G/C/U | n/a | n/a | 321.6 ± 61.9 nM | n/a | n/a | Hsu et al (2017) |
| YTHDF1 | C/G U/G/C m6A G/C/U A/G/C/U | n/a | n/a | 0.51 ± 0.045/0.13 µM | n/a | n/a | Arguello et al (2019) and Zaccara and Jaffrey (2020) |
| YTHDF2 | A/C/G U/G/C m6A G/C/U A/G/C/U | n/a | n/a | 0.79 ± 0.018/0.16 µM | n/a | n/a | Arguello et al (2019) and Zaccara and Jaffrey (2020) |
| YTHDF3 | GGm6ACU | n/a | n/a | 0.14 µM | n/a | n/a | Li et al (2017), Zaccara and Jaffrey (2020), and Zhou et al (2022) |
| IMP1 | UCGGm6ACU | n/a | n/a | 3.7 ± 0.8 nM | n/a | n/a | Nicastro et al (2023) |
| IMP1 | UCGGACU | n/a | n/a | 20.9 ± 4.0 nM | n/a | n/a | Nicastro et al (2023) |
| Erasers | Km (alpha-ketogluarate) | References | |||||
| FTO | m6A | 0.409 ± 0.023 µM | 2.88 µM | n/a | n/a | 0.296 ± 0.004 (min−1) | Jia et al (2008), Jia et al (2011), Ma et al (2012), Mauer et al (2017), Wei et al (2018), and Relier et al (2022) |
| FTO | m6Am | 1.34 µM | 2.88 µM | n/a | n/a | 8.78 (min−1) | Mauer et al (2017) |
| FTO | 3meT | 0.95 ± 0.12 µM | 2.88 µM | n/a | n/a | 0.007 ± 0.0002 (min−1) | Jia et al (2008) and Ma et al (2012) |
| FTO | 3meU | 8.51 ± 3.13 µM | 2.88 µM | n/a | n/a | 0.115 ± 0.022 (min−1) | Jia et al (2008), Jia et al (2011), and Ma et al (2012) |
| ALKBH5 | G(A/G)m6ACU | 1.38 ± 0.27 µM/192 ± 25 nM | 2.5 ± 0.5 µM | n/a | n/a | 0.169 ± 0.0106 (min−1) | Li et al (2016), Kaur et al (2022), and Peng et al (2023) |
The m6A modification has also been shown to either promote or limit interactions with proteins directly or indirectly. A set of specific RNA-binding proteins, termed m6A readers, can bind specifically to m6A-modified RNA and positively or negatively impact the formation of different RNPs at various stages of gene expression regulation (Xu et al, 2014; Hsu et al, 2017; Gilbert & Nachtergaele, 2023; Sikorski et al, 2023).
The m6A modification is one of very few RNA modifications that are reversible (Boo & Kim, 2020). Therefore, many modification-associated effects on RNAs and RNPs can be reversed. Two m6A demethylases mediate the removal of the modification, also termed m6A erasers, fat mass and obesity-associated protein (FTO) or alpha-ketoglutarate–dependent dioxygenase alkB homolog 5 (ALKBH5) (Zhang et al, 2019; Kaur et al, 2022).
The m6A RNA modification requires the formation of protein–RNA complexes for its deposition and removal, and it itself modulates the formation of functionally distinct RNPs on modified RNAs by either enabling or preventing the binding of a defined set of protein interactors or other RNAs.
In this review, we aim to give, whenever possible, quantitative descriptions of mechanisms that govern how RNA modifications, specifically m6A, are modulated through the dynamic assembly of RNP complexes during the “writing” process and how the modification affects RNA structure, RNP assembly, and function (Table 1). We specifically focus on mechanisms at the basis of RNP assembly and RNA folding related to mRNAs and lncRNAs, not reviewing tRNA and rRNA modifications. The biological roles and regulatory pathways which are modulated by the m6A modification and their dysregulation in cancer and other pathologies have been extensively reviewed elsewhere (Barbieri & Kouzarides, 2020; Shi et al, 2020; Huang et al, 2021; Arzumanian et al, 2022; Hong et al, 2022; Boulias & Greer, 2023; Gilbert & Nachtergaele, 2023; Orsolic et al, 2023) and are not part of this review.
Site-specific deposition of m6A modifications on nascent RNA is enabled by protein–RNA and protein–protein interactions
The catalytic core enzyme
The correct and successful installation of each m6A modification on a specific transcript requires the spatially and temporally coordinated assembly of different dynamic RNPs. At the center of these RNP assemblies is the m6A methyltransferase core enzyme METTL3/METTL14, also referred to as m6A–METTL complex (MAC) (Liu et al, 2014; Wang et al, 2016; Śledź & Jinek, 2016). METTL3 is the active methyltransferase harboring a binding pocket for the methyl donor S-Adenosyl methionine (SAM), whereas METTL14 is catalytically inactive but crucial for substrate binding and positioning and is required to yield an active enzyme (Fig 1A) (Wang et al, 2016; Śledź & Jinek, 2016; Schöller et al, 2018). Posttranslational methylation of METTL14 R255 has also been demonstrated to positively impact catalytic activity by enhancing RNA binding (Liu et al, 2021).
Figure 1. Mechanisms of how the m6A core enzyme and its associated regulatory protein domains get recruited to or excluded from different target sites.
(A) Structure of the m6A enzymatic core complex consisting of methyltransferase METTL3 and the critical adapter protein METTL14. Left panel: domain structure of the METTL3 and METTL14 core enzymes. Middle panel: crystal structure of the METTL3/METTL14 complex (MAC) bound to methyl-donor S-Adenosyl methionine (SAM) (gray). The crystal structure contains the METTL3 MTA-70 methyltransferase domain (pink) and the HM and MTA-70 domains of METTL14 (blue); PDB-ID: 5IL1 (Wang et al, 2016). Right panel: AlphaFold 2 model (Mirdita et al, 2022) of the full-length complex between METTL3 and METTL14. Parts of METTL3 and METTL14 that are also present in the X-ray structure are colored in pink and blue, respectively. The CCCH domains of METTL3 are illustrated in green. (B) Regulatory subunit m6A-METTL-associated complex (MACOM). Left panel: domain structures of adapter proteins WTAP, VIRMA, ZC3H13, and HAKAI. Right panel: Cryo-EM structure of the WTAP, VIRMA, and ZC3H13 complex fitted into a low-threshold cryo-EM density map; PDB-ID: 7VF2 (Su et al, 2022). The METTL3/METTL14 X-ray structure (PDB-ID: 5IL1 (Wang et al, 2016)) is fitted into an additional low-threshold density. A potential binding interface for an RNA substrate (black) is illustrated based on protein-RNA crosslinking data from Su et al (2022). The 5′ to 3′ directionality of the RNA is unknown. (C) Model of how the METTL3/METTL14 core enzyme is recruited to mRNA 3′UTRs by VIRMA-meditated interactions with polyadenylation factors CPSF5 and CPSF6; PDB-ID: 5IL1 and 7VF2 (Yue et al, 2018). (D) Model of how METTL3/METTL14 is recruited to DRACH motifs on Xist lncRNA, mediated by RBM15 binding to U-rich stretches; PDB-ID: 5IL1 and 7VF2 (Patil et al, 2016). (E) Model of the exclusion-based mechanism whereby the m6A methylation complex cannot bind if the EJC occludes sites; PDB-ID: 5IL1 and 7VF2 (He et al, 2023; Uzonyi et al, 2023). ChimeraX version 1.6 was used for the visualization of experimental and predicted structures (Goddard et al, 2018; Pettersen et al, 2021). Panel (B) was adapted from Su et al (2022) under the Creative Commons CC By license (license: https://creativecommons.org/licenses/by/4.0/).
METTL3 and METTL14 get assembled into a functional enzyme right after translation and posttranslational processing in the cytosol and are subsequently imported into the nucleus as one functional unit, in which only METTL3 carries a nuclear localization signal (Schöller et al, 2018; Han et al, 2022). To ensure a stoichiometric ratio between METTL3 and METTL14 and to maintain m6A homeostasis, recent findings show that METTL3 competes with E3 ubiquitin ligase STUB1 for binding to METTL14 and enhances METTL14 stability by preventing its degradation (Zeng et al, 2023).
The functional METTL3/METTL14 enzyme localizes mainly to the nucleus and recognizes a distinct five-nucleotide-long consensus motif, the DRACH motif (D = A,G,U; R = A,G; H = A,C,U), within which the central adenosine base gets methylated (Table 1 and Fig 1A) (Csepany et al, 1990; Liu et al, 2014). In vitro, the core enzyme preferentially methylates single-stranded DRACH motifs (Liu et al, 2014; Meiser et al, 2020). Interestingly, the enzyme displays an increased affinity towards double-stranded RNA in vitro and simultaneously a decreased enzymatic turnover (Qi et al, 2022). The preference for single-stranded motifs hints that the METTL3/METTL14 enzyme has no or only a limited ability to remodel structured RNAs.
Besides modifying RNAs, the METTL3/METTL14 enzyme has also been reported to be active on single-stranded DNA in vitro, which has been linked to DNA repair at UV- or X-ray-induced double-stranded DNA lesions in vivo (Yu et al, 2021a; Qi et al, 2022).
Whereas localization of METTL14 seems to be restricted to the nucleus under physiological conditions, METTL3 is also present in the cytoplasm, where it can promote translation independent of its catalytic activity. METTL3 binds to existing m6A marks, directly recruits translation initiation factors CBP80/20 and eIF4E, and facilitates recruitment of eIF3 (Liu et al, 2014; Lin et al, 2016; Schöller et al, 2018).
In summary, the active m6A RNA methyltransferase METTL3 forms a functional and catalytic unit with binding partner METTL14 and methylates preferably the short DRACH consensus motif in single-stranded RNA in the nucleus.
Co-transcriptional m6A deposition
An increasing body of research demonstrates that m6A modification by the METTL3/METTL14 enzyme is installed co-transcriptionally (Ke et al, 2017; Knuckles et al, 2017; Slobodin et al, 2017; Louloupi et al, 2018; Lee et al, 2021; Xu et al, 2022). Efficient co-transcriptional RNA modification requires co-localization and/or recruitment of the m6A core enzyme to actively transcribing genes on chromatin to facilitate its interaction with the nascent RNA emerging from Pol II.
A potential mechanism to promote spatial proximity between the methylation enzyme and nascent RNAs is through direct or indirect interaction of METTL3 or METTL14 with epigenetic marks on histone tails. METTL14, for instance, directly interacts with the histone H3 trimethylation at Lys 36 (H3K36me3), an epigenetic mark for active transcriptional elongation, and with active histone marks H3K27ac and H3K4me3 (Fig 2A) (Barbieri et al, 2017; Huang et al, 2019; Dou et al, 2023). METTL3, on the other hand, gets recruited to repressive histone marks H3K9m3 and H4K20m3. This recruitment depends on the catalytic activity of METTL3 and is maintained by direct interactions between METTL3 and the nuclear m6A reader YTHDC1 (Fig 2C) (Xu et al, 2021).
Figure 2. Mechanisms of MAC/ MACOM recruitment to nascent RNA for co-transcriptional RNA modification.
The methylation complex is recruited to the nascent RNA by the following interactions/mechanisms. (A) Direct interactions between METTL14 and histone marks H3K36me3, H3K27ac, and H3K4me3 (Barbieri et al, 2017; Huang et al, 2019; Dou et al, 2023). (B) Direct interactions between WTAP and the transcription elongation factor Spt6 (Akhtar et al, 2021). (C) Direct interactions between METTL3 and histone marks H3K9me3 and H3K20me3 and maintained by m6A reader YTHDC1 (Xu et al, 2021). (D) Direct interactions between METTL14 and histone mark H3K27me3 leads to the recruitment of histone demethylase KDM6B, which directly interacts with METTL14 (Dou et al, 2023). (E) Direct interaction between the RBM15 SPOC domain with phosphorylation marks on the C-terminus of RNA polymerase II (Appel et al, 2023). (A, C, D) Note that it is not known which parts of the METTL14 (A, D) or METTL3 (C) proteins are interacting with the histone methylation marks. Structures used in this figure are as follows: METTL3/METTL14 (MAC) (PDB-ID: 5IL1), WTAP/ZC3H13/VIRMA (PDB-ID: 7VF2), and human RNA polymerase II (PDB-ID: 3J0K). ChimeraX version 1.6 was used for the visualization of experimental and predicted structures (Goddard et al, 2018; Pettersen et al, 2021).
Furthermore, because of their co-localization to active and repressive histone marks, both METTL3 and METTL14 have been associated with regulating transcriptional activation and repression. METTL14, for example, binds to repressive histone mark H3K27me3 in a DNA- and RNA-independent manner and recruits lysine demethylase 6B (KDM6B) to these sites via a direct interaction. KDM6B can then remove H3K27me3 marks and promote transcriptional activation of the affected genes (Fig 2D) (Dou et al, 2023).
Another study found that co-transcriptional m6A modification of chromatin-associated regulatory RNAs (carRNAs) by METTL3 leads to transcriptional repression of the affected genes. Unmodified carRNAs bind and recruit proteins, such as CBP/EP300 and YY1, to promote open chromatin and activate transcription. Once methylated, the m6A-marked carRNAs are bound by nuclear m6A reader protein YTHDC1 and then targeted for degradation via the nuclear exosome targeting complex, depleting carRNA-associated transcriptional activators from the affected sites (Liu et al, 2020).
A chromatin-independent recruitment of the methyltransferase complex might be facilitated by the C-terminal RGG-rich domain of METTL14, which has the potential to interact with G-quadruplex structures on nascent RNAs. In vitro, enrichment of m6A-modified DRACH motifs close to G-quadruplex structures could be demonstrated, but if this mechanism is relevant in the cellular contexts remains to be understood (Yoshida et al, 2022).
Target selection by specific recruitment of the core methyltransferase complex to RNA through adapter proteins
Strikingly the number of DRACH motifs in the transcriptome far outnumbers the motifs that are experimentally verified to be methylated in cells. On mRNAs, m6A marks are enriched in long exons and introns, close to stop codons, and in the 3′UTRs, whereas DRACH motifs show no specific enrichment (Dominissini et al, 2012, 2013; Meyer et al, 2012; Ke et al, 2017). The mechanism by which the METTL3/METTL14 complex selects or is recruited to these specific sites is still unclear. To date, selectivity in vivo has been proposed to occur via two distinct mechanisms: (1) via specific recruitment of the core methyltransferase complex through adapter proteins or (2) via the exclusion of certain sites through competitive binders.
Within the cell, the METTL3/METTL14 core enzyme interacts with a growing number of protein interactors that recruit the methyltransferase to different cellular locations and prevent or facilitate its interaction with target sites on different RNAs (Bertero et al, 2018; Covelo-Molares et al, 2021).
Precursor-mRNA (pre-mRNA) splicing factor Wilm’s tumor 1-associated protein (WTAP) is frequently found to be associated with the m6A core enzyme (Fig 1B) (Liu et al, 2014). WTAP is an important key factor modulating the enzymatic activity of the METTL3/METTL14 complex in vivo by enhancing its RNA-binding ability, but it is dispensable for efficient enzymatic activity in vitro (Ping et al, 2014; Wang et al, 2016; Śledź & Jinek, 2016; Yan et al, 2023). The interaction between the core enzyme and WTAP is at least in part mediated by direct interactions with the N-terminal domain of METTL3 and is not affected by posttranslational phosphorylation on METTL3 (Ping et al, 2014; Schöller et al, 2018; Su et al, 2022). The association of WTAP with the core enzyme is favored in phase-separated states, such as nuclear speckles. This is in line with the observation that the localization of the core enzyme into nuclear speckles is dependent on its association with WTAP (Schöller et al, 2018; Han et al, 2022).
In Drosophila, the WTAP homolog FI(2)d recruits the METTL3/METTL14 complex to actively transcribing genes via direct interactions with the transcription elongation factor and histone chaperone Spt6 (Akhtar et al, 2021). In this context, the core enzyme recruitment to nascent RNAs is facilitated via the adapter protein WTAP, rather than direct interaction with histone marks (Fig 2B). During hepatitis C infection, WTAP relocalizes to the cytosol and recruits the METTL3/METTL14 complex to hepatitis C virus RNA, which promotes viral RNA methylation. A direct interaction between WTAP and the viral RNA seems required for this recruitment, but the details of this interaction remain to be described (Sacco et al, 2022).
Besides recruiting the core enzyme to nuclear speckles and nascent RNAs, WTAP also acts as the main interaction hub for the METTL3/METTL14 enzyme and mediates many crucial interactions with other adapter proteins. One such WTAP-dependent interactor is virilized homolog (VIRMA) (Fig 1B) (Yue et al, 2018; Su et al, 2022). VIRMA itself directly interacts with the polyadenylation factors cleavage and polyadenylation specificity factor subunit 5 (CPSF5) and CPSF6 in an RNA-dependent manner and recruits the METTL3/METTL14/WTAP complex to the 3′ ends of target mRNAs and thereby regulates the preferential installment of the m6A modification on 3′UTRs and near stop codons (Fig 1C) (Yue et al, 2018). Another recent publication found that WTAP and VIRMA together also counteract the interaction of the METTL3/METTL14 enzyme with double-stranded DNA and potentially also double-stranded RNA and increase methylation efficiency on single-stranded RNA (Yan et al, 2023).
The composite interaction interface created by WTAP and VIRMA recruits Zinc-finger CCCH-type containing 13 (ZC3H13) protein and E3 ubiquitin ligase HAKAI to the core enzyme (Fig 1B) (Su et al, 2022). ZC3H13 itself is an RNA-binding protein that greatly enhances the RNA-binding affinity of the core enzymes towards RNA in vitro and therefore appears to be an integral partner in regulating m6A modification by facilitating the m6A RNP assembly (Table 1) (Wen et al, 2018; Su et al, 2022). Furthermore, ZC3H13 knockdown leads to translocation of most of the WTAP, VIRMA, and HAKAI into the cytoplasm, suggesting that its presence is important for maintaining the nuclear localization of these factors, which are critical to regulating the METTL3/METTL14 core enzyme (Wen et al, 2018).
Not much is known about the function of HAKAI in association with the m6A methylation machinery, but it has been reported that it has a stabilizing effect on the other components of the m6A machinery, including METTL3, METTL14, WTAP, and VIRMA, for which its catalytic activity as E3 ubiquitin ligase is not required (Bawankar et al, 2021).
Because of their frequently observed association and regulatory functions in complex with the METTL3/METTL14 enzyme, the adapter proteins WTAP, VIRMA, HAKAI, and ZC3H13 are referred to as regulatory subunit m6A–METTL-associated complex (MACOM) and are considered integral players in regulating m6A deposition.
In addition to components of the MACOM, proteomics analysis revealed that WTAP also directly interacts with RNA-binding motif protein 15 (RBM15) and its paralog RBM15B and recruits it to the METTL3/METTL14 core enzyme (Horiuchi et al, 2013). RBM15/15B interacts with U-rich sequences in RNAs, and by this mechanism, recruits the m6A core enzyme to specific DRACH motifs on lncRNA Xist (Fig 1D) (Patil et al, 2016). Furthermore, the SPOC domain of RBM15 has been demonstrated to bind phosphoserines and to act as a reader of phosphorylation marks on the C-terminal domain of Pol II, which poses another potential mechanism by which the core enzyme gets recruited to nascent RNAs for co-transcriptional modification (Fig 2E) (Appel et al, 2023).
Furthermore, ERK1-mediated phosphorylation of METTL3 S43 has recently been shown to induce m6A modification of small nuclear RNA 7SK, by promoting METTL3 release from inhibitor HEXIM (Perez-Pepe et al, 2023). A connected study showed that the 7SK snRNA contains eight m6A sites, which are subject to methylation and demethylation by METTL3/METTL14 and ALKBH5, respectively. The presence of methylation at these sites seems to induce a conformational change in 7SK, that favors the release of Pol II transcriptional activator P-TEFb and transcriptional up-regulation (Wang et al, 2023).
Target selection via DRACH-motif exclusion through competitive binders
Though a subset of methylated DRACH sites can so far be explained by site-specific recruitment of the core enzyme by adapter proteins to defined sites, it does not give a sufficient explanation for methylation target selection.
Two recent studies suggest an alternative mechanism by which the distinct distribution of m6A marks on mRNAs can be partially explained by an exclusion-based mechanism through the binding of the exon junction complex (EJC) on the pre-mRNA (Fig 1E). The EJC thereby masks sites of around 100 nucleotides upstream and downstream of the splice junction and makes them physically inaccessible for the methyltransferase complex. This still permits methylation in longer exons, close to stop codons, and at the 3′UTR and noncoding RNAs, that are not spliced (He et al, 2023; Uzonyi et al, 2023). In addition, both studies find that the METTL3/METTL14 complex does not seem to have any selectivity on which sites to methylate besides the previously mentioned DRACH motif, proposing a mechanism by which mRNA methylation might be guided mainly by excluding or suppressing certain DRACH motifs through physical barriers from being methylated. A similar exclusion-based process has been described in heat shock response. During heat shock, the m6A reader protein YTHDF2 localizes to the nucleus and preferentially binds m6A marks installed on the 5′UTR of mRNAs and, thereby, prevents their removal by demethylase FTO by masking the methylation marks (Zhou et al, 2015).
As the assembly of the EJC happens after splicing has been completed, those studies also challenge the notion that m6A deposition happens immediately on nascent transcripts, as suggested by the direct recruitment of the methylation complex to specific chromatin sites and additional research showing that m6A is deposited in nascent RNAs, in introns, and that m6A modifications at splice junctions can lead to a more efficient splicing (Louloupi et al, 2018; Xu et al, 2022).
These seemingly contradictory findings may depict a more complex picture of m6A target site selection and modification timing by the METTL3/METTL14 enzyme. Certain mechanisms might be limited to certain transcripts, for example, noncoding RNAs, or cellular conditions; in other cases, multiple mechanisms might affect the target selection on the same transcript. Future investigations will consolidate current literature and give an updated model.
In brief, the m6A core enzyme METTL3/METTL14 relies on the coordinated association with adapter proteins or the timely assembly of trans-acting machineries for defining site-specific methylation. These interactions are in many cases facilitated by splicing factor WTAP but also occur via direct interactions with other components of the core enzyme.
Effect of m6A modification on RNA structure and protein binding
Through the addition of a methyl group on the base of any given nucleotide, the chemical and physical properties of the affected nucleotides are changed. These changes affect an RNA molecule by either influencing its structure and/or enabling or preventing interactions with cellular interactions partners. Both mechanisms have the potential to facilitate or inhibit the assembly and downstream function of protein–RNA complexes in the cell. In the following sections, we aim to summarize what is known about how altered physiochemical properties of m6A-modified RNAs influence their structure and interactions.
m6A affects RNA structure formation
The m6A modifications introduce an additional methyl group at the nitrogen at position six of the nucleobase. The presence of this additional methyl group at the Watson–Crick interface leads to a decrease of the rate constant for duplex association (kon) by ∼fourfold to ninefold for an RNA–RNA or RNA–DNA duplex containing an m6A-U/T base pair. In contrast, the rate constant for duplex dissociation (koff) changes only by 0.7–1.7fold (Shi et al, 2019). This means that an RNA duplex structure containing an m6A-modified base is less readily and efficiently folded, resulting in a local destabilization of RNA structures. Atomistically, the methyl group on the N6 of the modified adenosine base has to adopt the less energetically favorable anti-conformation to form a Watson–Crick H6--O4 hydrogen bond. This results in the destabilization of a double-stranded GGACU by 1 kcal/mol (Fig 3A) (Roost et al, 2015; Shi et al, 2019). This subtle destabilization is sufficient to loosen up RNA duplexes enough to allow the binding of single-stranded RNA-binding proteins or impair cellular checkpoint processes that rely on base complementarity but will most likely only fine-tune interactions rather than act as yes-or-no switches (see below). As a comparison, the Gibbs standard free energies for G-C and A-U Watson–Crick base-pairs are −5.53 and −4.42 kcal/mol, respectively (Vendeix et al, 2009). Therefore, it is also unlikely that the key effects of the m6A modification are major m6A-induced RNA structural rearrangements.
Figure 3. Effects of m6A modifications on RNA base-pairing and structure.
(A) Left panel: stable Watson–Crick base-pairing between m6A-U is only possible in the energetically less favorable anti-conformation of the m6A base. Right panel: two RNA strands containing an m6A-U instead of an A-U base-pair have a fourfold to ninefold decreased annealing rate constant but a dissociation rate constant which is not significantly changed by the methylation (Shi et al, 2019). (B) Left panel: Hoogsteen–Hoogsteen m6A-A base-pair is stabilized compared with Hoogsteen–Hoogsteen A–A (Roost et al, 2015). Right panel: base-pairing between pre-mRNA 5′splice site (5′SS) and modified (green) and unmodified (red) U6 snRNA (yellow) and U5 snRNA (orange). Successful base-pairing between modified U6 snRNA and 5′SS does not require a conserved AAG motif upstream of the 5′SS, but this conserved motif is required with unmodified U6 snRNA for stable base-pairing (Ishigami et al, 2021; Parker et al, 2022). (C) Left panel: depletion of m6A marks during METTL3 knock-out favors secondary structure formation of otherwise single-stranded endogenous RNAs, which triggers the recognition of these aberrant double-stranded RNAs by RIG-I and MDA5 resulting in an innate immune response (Gao et al, 2020). Right panel: double-stranded virus RNA from vesicular stomatitis virus gets methylated by METTL3, which leads to loss of structure and an impaired RLR recognition and innate immune response (Qiu et al, 2021). Structures used in this figure are as follows: METTL3/METTL14 (PDB-ID: 5IL1). ChimeraX version 1.6 was used for the visualization of experimental and predicted structures (Goddard et al, 2018; Pettersen et al, 2021). Panel (B) was adapted from Ishigami et al (2021) under the Creative Commons CC By license (license: https://creativecommons.org/licenses/by/4.0/).
The destabilizing effect induced by the m6A modifications, furthermore, depends on the structural context within the RNA. One study showed that m6A can stabilize an m6A-U base pair in the presence of Mg2+ when it is directly adjacent to a 5′ bulge (Liu et al, 2018). This finding is supported by the fact that m6A modifications induce a structural transition in the immediate proximity of the modification. Nucleotides directly 5′adjacent to the modified base tend to adopt a single-stranded conformation, whereas nucleotides directly 3′ to m6A are more likely to be base-paired (Roost et al, 2015).
Interestingly, another study showed that the m6A modification also has a destabilizing effect on the m6A – 8-oxo-G base-pairing at the Hoogsteen interface, leading to a drop in a DNA duplex melting temperature of a 13-base pair-long duplex from 47°C to 44°C. However, by which mechanism this influences nucleic acid folding and which are the potential downstream consequences are still unexplored (Wang et al, 2017).
These subtle changes in the base-paring kinetics can affect biological processes in the cell mainly in two ways (1) by altering processes that are directly dependent on base-pairing or (2) by altering the RNA structure to enable or prevent protein association.
Altered base-pairing kinetics directly affect cellular processes that rely on base complementarity
Pre-mRNA splicing in eukaryotes is an essential cellular process that depends, among other things, on base-pairing kinetics. In humans, adenosine 43 (A43) of the U6 snRNA is specifically modified by methyltransferase METTL16 (Pendleton et al, 2017; Warda et al, 2017). A43 is stringently conserved and base-pairs with the 5′-splice site (5′SS) of pre-mRNAs. Interestingly, in higher eukaryotes, which carry the m6A modification in U6 snRNA, the 5′SS is less conserved than in lower eukaryotes, especially at the +4 position. Two studies, in A. thaliana and S. pombe, report that in the presence of the m6A in the U6 snRNA, 5′SSs with an A in position 4 are favored as the modification slightly stabilizes the m6A-A base pair by 0.7 kcal/mol as compared with an A–A base pair. A 5′SS with a U at position 4, on the other hand, gets destabilized by 0.5–1.7 kcal/mol and preferentially base-pairs with the loop 1 region of U5 snRNA instead of the U6 snRNA (Roost et al, 2015; Ishigami et al, 2021; Parker et al, 2022) (Fig 3B). These small thermodynamic changes in base-pairing stability induced by the m6A modification lead to a relaxation in the 5′ exon constraints and allow for increased protein diversity in higher eukaryotes.
Followed by successful splicing, base-pairing kinetics is also crucial for mRNA decoding during translation. The presence of an m6A-modified nucleotide influences efficient base-pairing between the codon on the mRNA and the anticodon on the cognate tRNA leading to only near-cognate codon-anticodon interactions, thereby slowing down proper tRNA accommodation. Under high-accuracy conditions (1.3 mM MgCl2), this results in a reduction of the Michaelis–Menten parameter (kcat/Km)pep for peptide bond formation at m6AAA codons by a factor of 18 as compared with AAA codons (Choi et al, 2016). This effect of the m6A modification during decoding depends on the position and sequence context within the codon. Another study recently confirmed these findings by showing that m6A does not prevent canonical codon–anticodon base-pairing but favors alternative conformations leading to a lower stability and enhancing the tRNA drop-off. This consequently results in a smaller number of ribosomes that complete the decoding process, resulting in a less efficient translation of the affected transcripts (Jain et al, 2023).
A related study connected the reduced translational efficiency caused by the m6A modification to the transcriptional speed of Pol II (Slobodin et al, 2017). The researchers showed that faster transcription rates result in fewer methylation marks on mRNAs, resulting in a more efficient translation of the affected transcripts and vice versa. This mechanism allows an indirect coupling between transcription and translation in an m6A-dependent manner in eukaryotes.
The propensity of m6A-modified nucleotides to alter base-pairing kinetics also reduces the formation of endogenous double-stranded RNAs. Upon the deletion of the m6A methyltransferase METTL3 in murine fetal liver, RNAs, which are normally highly m6A-modified and show a low folding propensity, form double-stranded structures. These aberrantly formed endogenous double-stranded RNAs lead to an anomalous activation of the innate immune response by activating pattern recognition receptors such as RIG-I and MDA5 (Fig 3C) (Gao et al, 2020).
In addition, the RNA duplex destabilizing effect of m6A can also be hijacked by viruses. Infection by double-stranded virus RNA vesicular stomatitis virus leads to translocation of m6A methyltransferase METTL3 to the cytoplasm and an increase in m6A marks on viral RNAs. This reduces the proportion of double-stranded viral RNA and thereby the efficiency by which it is sensed and cleared by retinoic acid-inducible gene I-like receptors (RLRs) (Fig 3C) (Qiu et al, 2021).
Overall, the m6A modification leads to subtle changes in the RNA folding kinetics that result in small changes in the local RNA structure. Those changes carefully fine-tune cellular processes sensitive to base-pairing such as splicing and translation, and can affect the immunogenicity of endogenous or pathogenetic RNAs.
m6A-mediated RNA structural changes create new protein binding sites
The altered base-pairing characteristics induced by the m6A modification are also associated with RNA structural rearrangements that facilitate protein–RNA interactions.
This is exemplified by the interaction of splicing factors heterogeneous nuclear proteins C and G (hnRNP-C and hnRNP-G) with pre-mRNAs and lncRNAs (Liu et al, 2015, 2017).
The protein hnRNP-C binds U-rich single-stranded RNA stretches in pre-mRNA and lncRNAs. These can be masked within double-stranded RNA structures. Upon modification, the A-U Watson-Crick base-pair gets destabilized, thereby releasing U-rich binding sites for hnRNP-C. The ability of hnRNP-C to bind such U-rich stretches in pre-mRNA in an m6A-dependent manner subsequently leads to the generation of hnRNP-C-dependent splicing variations (Fig 4A) (Liu et al, 2015).
Figure 4. Direct and indirect effects of m6A modifications on reader protein binding.
Panels (A, B) show indirect m6A readers, whereas panels (C, D, E) show direct m6A readers or “anti-readers.” (A) “m6A-switch” mechanism 1: methylation of DRACH motifs by the m6A writer complex within RNA hairpins leads to the release of U-rich stretches, which can then be bound by RNA-binding protein hnRNP-C (Liu et al, 2015). (B) “m6A-switch” mechanism 2: Methylation of DRACH motifs by the m6A writer complex within RNA hairpins leads to the release of the DRACH motif from the RNA structure, which allows RNA-binding protein hnRNP-G to bind the single-stranded AGRAC motifs (R = A/G) with its low-complexity domain (Liu et al, 2017). The RRM domain of hnRNP-G can bind CC(A/C)-rich sequences independent of the m6A modification (Heinrich et al, 2009). (C) m6A modifications by the m6A writer complex create a new binding epitope for m6A reader proteins such as YTHDC1-2 and YTHDF1-3. (D) m6A marks within exons in pre-mRNA can be bound by m6A reader YTHDC-1 (orange), which recruits splicing factor SRSF3 (purple) that promotes exon inclusion and prevents binding of splicing factor SRSF10 (brown), which would promote exon skipping in the absence of the m6A mark. (E) m6A modifications by the m6A writer complex can also disrupt protein binding interfaces. The affected class of proteins is termed m6A “anti-readers.”
The protein hnRNP-G is a regulator of alternative splicing and preferentially binds single-stranded AGRAC motifs (R = A/G), via its intrinsically disordered or low-complexity domain. Strikingly, the hnRNP-G binding motif strongly overlaps with the m6A DRACH motif. Methylation of the DRACH motif can release the hnRNP-G binding motifs from double-stranded RNA structures allowing hnRNP-G to bind (Liu et al, 2017). The binding of hnRNP-G to its motif happens independently of the m6A modification in single-stranded RNAs. Similarly to the binding by hnRNP-C, the m6A-dependent accessibility of the consensus motif enables hnRNP-G–specific regulation of alternative splicing (Fig 4B) (Zhou et al, 2019). The RNA-recognition motif (RRM) domain of hnRNP-G, on the other hand, binds single-stranded CC(A/C)-rich motifs m6A-independently (Heinrich et al, 2009), thereby providing additional binding specificity.
Heterogeneous nuclear protein A2/B1 (hnRNPA2B1) binds AGG and UAG motifs via its two RNA-recognition motif domains through a similar mechanism as hnRNP-G. hnRNPA2B1 binds these motifs in a subset of pri-miRNAs and interacts with Dgcr8, a component of the microprocessor complex, thereby promoting miRNA processing (Alarcón et al, 2015a; Wu et al, 2018).
The release of specific single-stranded recognition sequences by m6A-induced base-pairing changes has been termed an “m6A switch.”
An example of how altered base-pairing kinetics can prevent the formation of a protein–RNA complex are box C/D small nucleolar RNAs (snoRNAs). Box C/D snoRNAs are the core component of the Box C/D snoRNP, which installs 2′-O-methylation marks in rRNA during ribosome biogenesis. For correct assembly of this multicomponent RNP, the Box C and D consensus motifs on the snoRNA have to fold into a characteristic kink-turn conformation to allow the binding of the protein core components. The Box D motif contains a GAC motif which is potent for m6A methylation by the METTL3/METTL14 complex. An m6A modification within the Box D motif interferes with formation of a structurally critical trans Hoogsteen-sugar A-G base-pair, disrupting the kink-turn fold. The kink-turn fold is required for binding of the Box C/D-associated proteins. Therefore, the Box C/D snoRNP cannot be assembled in the presence of an m6A modification in the Box D motif (Huang et al, 2017).
In summary, the subtly altered base-pairing kinetics resulting from the presence of an m6A modification also affects the availability of single-stranded RNA-binding interfaces. Therefore, the modification can fine-tune the binding of specific RNA-binding proteins and affect cellular processes such as splicing or miRNA processing.
The m6A modification creates new binding epitopes for proteins
Besides altering base-pairing properties, which directly affect the RNA structure and trans-interactions with other RNAs and proteins, the additional methyl group also introduces a new binding epitope for protein interaction partners. The m6A modifications can influence the protein interactome of a given modified RNA either by (1) serving as an “m6A-switch” as described in the previous section, (2) by providing m6A-specific binding pockets (Dominissini et al, 2012, 2013) or (3) by weakening the interaction with a protein binding partner (Arguello et al, 2017).
YTH domain-containing proteins
YTH domain-containing proteins are the best-known family of m6A-interacting proteins, or m6A readers. This protein family has been associated with almost all cellular processes linked to the m6A modifications to date, including transcriptional regulation, pre-mRNA splicing, RNA nuclear export, translation regulation, and RNA stability and decay (Boulias & Greer, 2023).
Proteins of this family bind the m6A modification specifically through their conserved YTH-domain (Fig 4C). Within the YTH domain, two crucial aromatic residues form the m6A binding pocket. Removal of either residue leads to loss of m6A binding (Theler et al, 2014; Xu et al, 2014; Zhu et al, 2014). The YTH domain is characteristically flanked by two intrinsically disordered regions at the N- and C-terminus (Sikorski et al, 2023).
YTHDC1 is the only exclusively nuclear YTH-domain-containing m6A reader protein. In contrast to the other YTH domain-containing m6A readers, YTHDC1 shows a sequence preference towards the canonical DRACH motif, with a preference for C after the m6A residue and purines at the n-1 and n-2 positions around the methylation site (Table 1). Because of its nuclear localization, YTHDC1 is involved in almost all nuclear processes that involve m6A-modified RNAs, including chromatin remodeling, transcriptional regulation, X-chromosome inactivation, mRNA processing, and nuclear mRNA export (Widagdo et al, 2022).
YTHDC1 can be recruited to co-transcriptionally m6A-modified nascent RNAs and recruits the histone H3K9me3 demethylase KDM3B by an unknown mechanism. This leads to the demethylation of H3K9 and, thereby, the removal of a repressive histone mark (Li et al, 2020).
Furthermore, YTHDC1 has the propensity to initiate phase separation through its arginine-rich C-terminal disordered domain (Cheng et al, 2021; Lee et al, 2021). One study showed that m6A modifications deposited on enhancer RNAs are bound by YTHDC1 and form m6A–enhancer RNA/YTHDC1 condensates. These condensates can undergo mixing with BRD4 coactivator condensates and facilitate the formation of transcriptional activator condensates (Lee et al, 2021). Together with splicing factor hnRNPG, YTHDC1 can also prevent premature transcription termination by binding to co-transcriptionally installed m6A-marks on 5′-ends of RNAs thereby preventing binding of the integrator complex (Xu et al, 2022). During pre-mRNA splicing, YTHDC1 is furthermore involved in alternative splicing regulation. YTHDC1 is recruited to m6A-modified exons and recruits splicing factor SRSF3 via direct interactions between its C-terminus and the C-terminus of SRSF3, resulting in exon inclusion. At the same time, binding of YTHDC1 inhibits the binding of splicing factor SRSF10 and thereby interferes with SRSF10-associated exon skipping. In the absence of an m6A modification within a given exon, SRSF10 can bind and promote exon skipping (Fig 4D) (Xiao et al, 2016). Finally, at the 3′-end of mRNAs, YTHDC1 can interfere with alternative polyadenylation leading to longer 3′UTRs (Chen et al, 2022).
YTHDC1 stays associated with mature m6A-modified mRNA to facilitate their nuclear export through the direct interaction with splicing factor and nuclear export adapter SRSF3. SRSF3, on the other hand, can interact with the nuclear export receptor NXF1, which facilitates the nuclear export of m6A-modified mRNA through protein–protein interactions (Roundtree et al, 2017).
The m6A marks on lncRNA Xist are also recognized and bound by YTHDC1. This interaction with Xist is crucial for X-chromosomal inactivation and gene silencing (Patil et al, 2016). Mechanistically, YTHDC1 binds to m6A-modified highly conserved AUCG tetraloops in the A-repeats, which leads to partial melting of the hairpin and modulation of the RNA structure (Jones et al, 2022).
The YTHDF paralogs YTHDF 1–3 are preferentially localized to the cytosol. Two contradictory models have been proposed regarding the function of these three different proteins. The canonical model assigns each YTHDF paralog-defined functions and targets, whereas a more recent model proposes that all three proteins have redundant functions and targets in mRNA degradation (Zaccara & Jaffrey, 2020).
Within the canonical model, YTHDF1 enhances the translation of m6A-modified transcripts by binding m6A marks at the stop codon and 3′ UTRs of mRNAs. YTHDF1 then directly interacts with and recruits translation initiation factor eIF3 to promote cap-dependent translation (Wang et al, 2015).
YTHDF2, on the other hand, is involved in mediating mRNA decay (Wang et al, 2014). To promote m6A-modified mRNA decay, YTHDF2 binds the m6A marks on mRNAs and either directly recruits the CCR4–NOT deadenylation complex (Du et al, 2016) or interacts with HRSP12, which in turn recruits the RNaseP/MRP endoribonuclease (Park et al, 2019).
In the nucleus, YTHDF2 is also involved in the clearance of R-loop structures. R-loops are three-stranded nucleic acid structures, consisting of a DNA:RNA hybrid and a single-stranded DNA, that are formed at the transcription bubble during transcription. RNA in R-loop structures can be m6A-modified by the METTL3/METTL14 methyltransferase, which leads to the recruitment of reader YTHDF2 to promote mRNA degradation, and R-loop clearance (Abakir et al, 2020; Kang et al, 2021).
Lastly, YTHDF3 is a modulator of YTHDF1 and YTHDF2 functions. YTHDF3 enhances translation through direct interaction with YTHDF1 (Shi et al, 2017). Similarly, YTHDF3 also interacts with YTHDF2 and seems to be involved in mRNA decay through this interaction (Shi et al, 2017). In addition, YTHDF3 binds and recognizes stress-induced newly methylated mRNA and drives their translocation into stress granules during oxidative stress (Anders et al, 2018). It is important to point out that a more recent study claims that m6A marks on mRNAs have only a limited effect on their translocation into stress granules (Khong et al, 2022).
The more recent model proposes that all three paralogs share similar targets and mediate mRNA degradation via association with CNOT, which is a scaffolding subunit of the CCR4vNOT deadenylase complex (Zaccara & Jaffrey, 2020).
The detailed functions and an attempt to consolidate the two existing models have recently been reviewed (Sikorski et al, 2023).
In contrast to the other members of the YTH domain family, YTHDC2 comprises several well-folded domains, including the YTH domain. YTHDC2 can interact with the small ribosomal subunit close to the mRNA entry and exit sites via its YTH and R3H domains. Therefore, YTHDC2 is thought to facilitate efficient translation by bridging m6A-marked mRNAs with the ribosome (Kretschmer et al, 2018). Furthermore, it can recruit 5′-3′ exonuclease XRN1 via direct interaction through its ankyrin domain and potentially promote mRNA decay (Kretschmer et al, 2018).
In summary, YTH domain-containing proteins bind m6A-modified RNA via similar mechanisms but lead to different downstream effects. How substrate specificity is achieved outside of the modification itself still needs to be better understood and will need further investigation to elucidate distinct functions and functional redundancies.
Non-YTH domain m6A readers
Non-YTH domain-containing m6A readers recognize and bind m6A-modified RNAs via alternative modes of recognition and are often dependent on the cellular context.
One such m6A reader is the translation initiation factor eIF3. eIF3 can directly read m6A marks in the 5′ UTRs of mRNA and recruits the 43S pre-initiation complex to the translation start site. Through this mechanism, eIF3 can initiate m6A-dependent/cap-independent translation initiation under stress conditions (Meyer et al, 2015).
Insulin-like growth factor 2 mRNA-binding proteins 1, 2, and 3 (IGF2BP1/2/3) are a family of distinct m6A readers, which specifically bind and recognize m6A modifications within the GG(m6A)CU motif (Huang et al, 2018). IGF2BPs bind the m6A modifications via their K homology (KH) domain, though how exactly the binding specificity is achieved is still unclear. In contrast to YTHDF proteins, IGF2BP proteins promote the stability of their mRNA targets in the cytosol and promote their translation.
A novel pair of m6A reader proteins is FMR1 and its paralogs FXR1 and FXR2. These proteins have been known to bind RNAs with the consensus motifs ACUG/U or U/A/GGGA, which show strong overlap with the METTL3/METTL14 DRACH consensus sequence (Ascano et al, 2012). Indeed, FXR1 was recently identified as a sequence context-specific m6A reader (Edupuganti et al, 2017). Furthermore, FXR1 has been described to bind to m6A marks in nascent RNAs and recruit DNA 5-methylcytosine dioxygenase TET1 to active chromatin loci leading to DNA demethylation and reprogrammed chromatin accessibility (Deng et al, 2022).
IMP1, a recently described m6A reader, interacts with the modification via a dedicated hydrophobic platform, enabling a high-affinity interaction (Table 1). This interaction is sequence-independent but is embedded into the methylation-independent sequence preference of IMP1, GGAC, which has significant similarity with the METTL3/METTL14 DRACH motif (Nicastro et al, 2023). The methyl group within the binding sequence results only in a small change in the on-rate from kon = 1.3 × 105 ± 0.1 × 105 M−1s−1 to kon = 8.7 × 104 ± 0.1 × 104 M−1s−1 but a more significant decrease in the off-rate from koff = 2.7 × 10−3 ± 0.4 × 10−3 s−1 to koff = 3.2 × 10−4 ± 0.7 × 10−4 s−1, highlighting how the modification can specifically increase the lifetime of a protein–RNA complex.
Lastly, proline-rich coiled-coil 2A (PRRC2A) is a recently discovered m6A reader critical for male spermatogenesis. It recognizes m6A marks within DRACH sequences, but little is known yet on the molecular details that govern this interaction (Wu et al, 2019; Tan et al, 2023).
Interactions negatively affected by the m6A modification
In the previous sections, we described examples in which the modification facilitates interactions, but RNP assembly can similarly be negatively affected by the m6A modification. Protein–RNA interactions that are negatively affected by the m6A modification include human Pumilio 2 (hPUM2), stress granule proteins G3BP1, G3BP2, RBM42, USP10, CAPRIN1, and RNA-binding proteins LIN28A, and EWSR1 (Fig 4E). Such proteins are termed m6A “anti-readers.”
One of the best-described m6A “anti-readers” is single-stranded RNA-binding protein hPUM2 (Vaidyanathan et al, 2017). hPUM2 consists of eight PUF repeats, each recognizing and making specific interactions with one base within its consensus motif. The presence of an m6A-modified base within this sequence weakens the binding of hPUM2. This effect is amplified with increasing m6A modifications, showing a quantitative negative effect of hPUM2 binding to its target in an m6A-dependent manner. The presence of m6A-modified bases within the 11-nucleotide-long hPUM2 recognition sequence (CCUGUAUAUAU) results in an increase of the KD from 0.14 to 0.34 nM with one m6A and 5.3 nM with three m6A bases (Vaidyanathan et al, 2017).
A proteome-wide screen has identified stress granule proteins G3BP1, G3BP2, RBM42, USP10, and CAPRIN1 as m6A “anti-readers” (Arguello et al, 2017). The abundance of m6A “anti-readers” in the stress granule proteome suggests that m6A-marked RNAs are not specifically targeted into stress granules or might even be actively excluded (Edupuganti et al, 2017; Khong et al, 2022). The exact mechanism through which these stress granule proteins sense the m6A modification remains to be elucidated.
LIN28A is a potent cellular inhibitor of pri-miRNA processing and antagonizes m6A reader hnRNPA2B1, which is associated with promoting pre-miRNA processing (Viswanathan et al, 2008; Alarcón et al, 2015a). LIN28A has been termed an m6A “anti-reader” as it shows decreased enrichment on m6A-marked RNAs, hinting a reduced affinity to modified RNAs (Sun et al, 2019). In the current model, LIN28A binds unmodified pri-miRNAs and antagonizes the installation of the m6A mark by METTL3/METTL14 to prevent binding of hnRNPA2B1 and the maturation of pri-miRNAs.
In summary, m6A “anti-readers” are a group of proteins whose interactions with their RNA targets are negatively affected by the presence of the m6A modification. Only a few examples of m6A “anti-readers” are known, and only for a small fraction of those the mechanism of modification recognition is described.
Reversing m6A-facilitated RNA interactions by “erasing” the modification
Another feature of m6A RNA modifications is that they can be removed during the life cycle of an RNA. The methyl group removal is achieved through one of two characterized m6A demethylases: FTO or ALKBH5 (Zhang et al, 2019; Kaur et al, 2022).
Both FTO and ALKBH5 are alpha-ketoglutarate-dependent dioxygenases and require oxygen, alpha-ketoglutarate, and Fe(II) to demethylate RNA. FTO localizes mainly in the nucleus and nuclear speckles but is also present in the cytosol (Jia et al, 2011; Wei et al, 2018). ALKBH5, on the other hand, localizes mainly into nuclear speckles (Thalhammer et al, 2011; Zheng et al, 2013). ALKBH5 acts only on m6A-modified bases, whereas FTO also acts on N6,2′-O-dimethyladenosine (m6Am), N1-methyladenosine (m1A), 3-meThymine (3-meT), and 3-meUracil (3meU)-modified nucleotides (Table 1) (Jia et al, 2011; Mauer et al, 2017; Wei et al, 2018; Relier et al, 2022).
The main substrates for FTO are internal m6A and m6Am modifications on mRNA and snRNA and m1A modifications on tRNA (Wei et al, 2018). The influence of specific sequences or RNA structural motifs on demethylation remains largely unclear (Han et al, 2010; Zhang et al, 2019). A large screening for cellular interaction partners showed that FTO interacts with a wide variety of proteins involved in transcription, RNA binding, splicing, DNA repair, and chromatin remodeling, some of which might also be involved in determining substrate specificity (Covelo-Molares et al, 2021). The splicing factor SFPQ, for example, has been identified as a binding partner of FTO and associates with the demethylase via its C-terminal domain. Available data show that SFPQ binds the RNA motif CUGUG and recruits FTO to m6A sites nearby to promote demethylation (Song et al, 2020). SFPQ-directed demethylation accounts for only 20% of FTO-mediated demethylation targets.
ALKBH5 seems to act only on internal m6A-modified nucleotides. Furthermore, it shows a sequence preference toward m6ACU motifs, which notably overlaps with the known METTL3/METTL14 DRACH motif (Kaur et al, 2022). ALKBH5 interacts with factors involved in chromatin remodeling, suggesting that similar to m6A methyltransferase METTL3/METTL14, it also localizes to actively transcribing genes. A recent article showed that ALKBH5 associates with the newly identified m6A reader RBM33, which activates the demethylase by suppressing SUMOylation and regulates the demethylation target selection (Yu et al, 2023). ALKBH5 further associates with components of the TREX complex, the EJC, and components involved in pre-mRNA splicing and miRNA-mediated RNA decay (Covelo-Molares et al, 2021).
So far, only little is known about FTO and ALKBH5 in their function as m6A demethylases. Recent years have unveiled the structural and catalytic basis for the demethylation, and important binding partners have been identified. Nevertheless, target specificity, interaction modes, specific functions, and potential functional redundancies are poorly understood. It is also unclear if specific RNA molecules can get methylated and demethylated multiple times. Future research in these directions will shed more light on the importance of reversing RNA modifications in the coming years.
Conclusions and Outlook
Protein–RNA assemblies play an integral part in many cellular processes, and the coordinated installment of RNA modifications can modulate their assembly. In this review, we aimed to summarize how m6A modifications, from their deposition until their removal, are functionally coupled to RNA folding and protein–RNA complex formation.
The process of the modification itself is a highly coordinated chain of events that relies on protein–RNA and protein–protein complex assembly. It starts with the site-specific and direct recruitment of the m6A core enzyme METTL3/METTL14 to sites of active transcription. This process is mainly mediated by direct protein–protein interactions between the core enzyme and defined histone marks or the C-terminus of Pol-II. The close spatial proximity between the enzymatic machinery and the nascent RNA allows for co-transcriptional installation of the modification (Huang et al, 2019; Wu et al, 2020; Xu et al, 2021). The core enzyme methylates preferentially single-stranded DRACH motifs within the newly synthesized RNA (Wang et al, 2016; Śledź & Jinek, 2016; Qi et al, 2022). Interestingly, only a subset of DRACH motifs gets modified in vivo, with a bias towards long exons and introns, and stop codons and 3′UTRs in mRNA (Dominissini et al, 2012, 2013). Two complementary mechanisms can partially explain this discrepancy: (1) site-specific recruitment of the core enzyme by protein–protein interactions with adapter proteins and (2) masking of DRACH motifs by competitive trans-acting factors (Ping et al, 2014; Yue et al, 2018; Su et al, 2022; He et al, 2023; Uzonyi et al, 2023). For site-specific recruitment, the METTL3/METTL14 core enzyme is targeted to specific DRACH motifs by different adapter proteins, including the MACOM complex components: WTAP, VIRMA, ZC3H13, and HAKAI. Complementary to protein-based recruitment, two very recent studies proposed a mechanism that excludes certain DRACH motifs through competitive binders, such as the exon–junction complex, and also challenge the notion that the modification is mainly installed co-transcriptionally (He et al, 2023; Uzonyi et al, 2023).
After installation, m6A changes the base-pairing kinetics, by weakening Watson–Crick m6A-U base-pairs and stabilizing Hoogsteen–Hoogsteen/Watson–Crick m6A-A base pairs (Roost et al, 2015; Shi et al, 2019). This alters the local RNA-folding kinetics, which modulates the binding of proteins such as hnRNP-G and hnRNP-C to newly available binding sites on the modified RNA (Liu et al, 2015, 2017; Zhou et al, 2019). Furthermore, the additional methyl group also introduces a new or modified binding epitope for RNA-binding proteins. YTH domain-containing proteins are the most common specific m6A readers and bind RNA preferentially in the presence of the modification.
Writers and readers can also cooperate. For example, both the m6A reader YTHDC1 and the METTL3/METTL14 core enzyme localize to chromatin, and can thereby cooperate to maintain a repressive chromatin state at murine retroviral elements (Xu et al, 2021).
Different writers can also compete with each other. For example, during pri-miRNA processing, the m6A modifications allow for regulated competition between inhibitors and facilitators, as is the case for m6A “anti-reader” LIN28A and m6A reader hnRNPA2B1 (Alarcón et al, 2015a; Sun et al, 2019).
Through the examples illustrated in this review, it becomes clear that RNA modifications, here specifically m6A, are regulated by the coordinated assembly of protein–RNA assemblies and play an integral role in modulating the assembly of diverse sets of protein–RNA complexes and thereby affect their downstream functions. Although more and more roles are being proposed for m6A in different pathways and the interactome is constantly expanding, only a few and often very specific cases describe the mechanistic details of m6A-mediated regulation in protein–RNA assemblies.
We still do not have high-resolution structural information on the RNA bound to MAC or MAC/MACOM. Furthermore, questions such as m6A target-site selection, how the modification and the process of modification affect RNA folding, and temporal information on the coordinated action of m6A writers, readers, and erasers remain to be answered. Also, little is still known about the mechanistic details of how m6A modification is functionally coupled to cellular processes such as transcription and the effect of RNA modification on co-transcriptional splicing (Martinez & Gilbert, 2018). Quantitative methods such as single-molecule fluorescence microscopy and other emerging biophysical and structural approaches in vitro and in vivo (Gor & Duss, 2023) will be key to understanding and tracking these multiple-step processes and to uncovering and understanding the crosstalk between RNA modification, RNA folding, and protein–RNA complex formation and function.
Acknowledgements
The authors would like to thank the entire Duss group for helpful discussions. The authors thank Dr. Kyungmin Baeg for carefully reading and providing feedback on this article.
Author Contributions
S Höfler: conceptualization, data curation, investigation, visualization, and writing—original draft, review, and editing.
O Duss: conceptualization, supervision, funding acquisition, visualization, and writing—review and editing.
Conflict of Interest Statement
The authors declare that they have no conflict of interest.
References
- Abakir A, Giles TC, Cristini A, Foster JM, Dai N, Starczak M, Rubio-Roldan A, Li M, Eleftheriou M, Crutchley J, et al. (2020) N6-methyladenosine regulates the stability of RNA:DNA hybrids in human cells. Nat Genet 52: 48–55. 10.1038/s41588-019-0549-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Akhtar J, Renaud Y, Albrecht S, Ghavi-Helm Y, Roignant J-Y, Silies M, Junion G (2021) m6A RNA methylation regulates promoter- proximal pausing of RNA polymerase II. Mol Cell 81: 3356–3367.e6. 10.1016/j.molcel.2021.06.023 [DOI] [PubMed] [Google Scholar]
- Akichika S, Hirano S, Shichino Y, Suzuki T, Nishimasu H, Ishitani R, Sugita A, Hirose Y, Iwasaki S, Nureki O, et al. (2019) Cap-specific terminal N 6-methylation of RNA by an RNA polymerase II-associated methyltransferase. Science 363: eaav0080. 10.1126/science.aav0080 [DOI] [PubMed] [Google Scholar]
- Alarcón CR, Goodarzi H, Lee H, Liu X, Tavazoie S, Tavazoie SF (2015. a) HNRNPA2B1 is a mediator of m(6)A-dependent nuclear RNA processing events. Cell 162: 1299–1308. 10.1016/j.cell.2015.08.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alarcón CR, Lee H, Goodarzi H, Halberg N, Tavazoie SF (2015. b) N6-methyladenosine marks primary microRNAs for processing. Nature 519: 482–485. 10.1038/nature14281 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anders M, Chelysheva I, Goebel I, Trenkner T, Zhou J, Mao Y, Verzini S, Qian S-B, Ignatova Z (2018) Dynamic m6A methylation facilitates mRNA triaging to stress granules. Life Sci Alliance 1: e201800113. 10.26508/lsa.201800113 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Appel L-M, Franke V, Benedum J, Grishkovskaya I, Strobl X, Polyansky A, Ammann G, Platzer S, Neudolt A, Wunder A, et al. (2023) The SPOC domain is a phosphoserine binding module that bridges transcription machinery with co- and post-transcriptional regulators. Nat Commun 14: 166. 10.1038/s41467-023-35853-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arguello AE, DeLiberto AN, Kleiner RE (2017) RNA chemical proteomics reveals the N6-methyladenosine (m6A)-regulated protein-RNA interactome. J Am Chem Soc 139: 17249–17252. 10.1021/jacs.7b09213 [DOI] [PubMed] [Google Scholar]
- Arguello AE, Leach RW, Kleiner RE (2019) In vitro selection with a site-specifically modified RNA library reveals the binding preferences of N6-methyladenosine reader proteins. Biochemistry 58: 3386–3395. 10.1021/acs.biochem.9b00485 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arzumanian VA, Dolgalev GV, Kurbatov IY, Kiseleva OI, Poverennaya EV (2022) Epitranscriptome: Review of top 25 most-studied RNA modifications. Int J Mol Sci 23: 13851. 10.3390/ijms232213851 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ascano M Jr., Mukherjee N, Bandaru P, Miller JB, Nusbaum JD, Corcoran DL, Langlois C, Munschauer M, Dewell S, Hafner M, et al. (2012) FMRP targets distinct mRNA sequence elements to regulate protein expression. Nature 492: 382–386. 10.1038/nature11737 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Balacco DL, Soller M (2019) The m6A writer: Rise of a machine for growing tasks. Biochemistry 58: 363–378. 10.1021/acs.biochem.8b01166 [DOI] [PubMed] [Google Scholar]
- Barbieri I, Kouzarides T (2020) Role of RNA modifications in cancer. Nat Rev Cancer 20: 303–322. 10.1038/s41568-020-0253-2 [DOI] [PubMed] [Google Scholar]
- Barbieri I, Tzelepis K, Pandolfini L, Shi J, Millán-Zambrano G, Robson SC, Aspris D, Migliori V, Bannister AJ, Han N, et al. (2017) Promoter-bound METTL3 maintains myeloid leukaemia by m6A-dependent translation control. Nature 552: 126–131. 10.1038/nature24678 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bawankar P, Lence T, Paolantoni C, Haussmann IU, Kazlauskiene M, Jacob D, Heidelberger JB, Richter FM, Nallasivan MP, Morin V, et al. (2021) Hakai is required for stabilization of core components of the m6A mRNA methylation machinery. Nat Commun 12: 3778. 10.1038/s41467-021-23892-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bertero A, Brown S, Madrigal P, Osnato A, Ortmann D, Yiangou L, Kadiwala J, Hubner NC, de Los Mozos IR, Sadée C, et al. (2018) The SMAD2/3 interactome reveals that TGFβ controls m6A mRNA methylation in pluripotency. Nature 555: 256–259. 10.1038/nature25784 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boccaletto P, Machnicka MA, Purta E, Piatkowski P, Baginski B, Wirecki TK, de Crécy-Lagard V, Ross R, Limbach PA, Kotter A, et al. (2018) MODOMICS: A database of RNA modification pathways. 2017 update. Nucleic Acids Res 46: D303–D307. 10.1093/nar/gkx1030 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boccaletto P, Stefaniak F, Ray A, Cappannini A, Mukherjee S, Purta E, Kurkowska M, Shirvanizadeh N, Destefanis E, Groza P, et al. (2022) MODOMICS: A database of RNA modification pathways. 2021 update. Nucleic Acids Res 50: D231–D235. 10.1093/nar/gkab1083 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boo SH, Kim YK (2020) The emerging role of RNA modifications in the regulation of mRNA stability. Exp Mol Med 52: 400–408. 10.1038/s12276-020-0407-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boulias K, Greer EL (2023) Biological roles of adenine methylation in RNA. Nat Rev Genet 24: 143–160. 10.1038/s41576-022-00534-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boulias K, Toczydłowska-Socha D, Hawley BR, Liberman N, Takashima K, Zaccara S, Guez T, Vasseur JJ, Debart F, Aravind L, et al. (2019) Identification of the m6Am methyltransferase PCIF1 reveals the location and functions of m6Am in the transcriptome. Mol Cell 75: 631–643.e8. 10.1016/j.molcel.2019.06.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Breger K, Brown JA (2023) Elucidating the kinetic mechanism of human METTL16. Biochemistry 62: 494–506. 10.1021/acs.biochem.2c00601 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen L, Fu Y, Hu Z, Deng K, Song Z, Liu S, Li M, Ou X, Wu R, Liu M, et al. (2022) Nuclear m6 A reader YTHDC1 suppresses proximal alternative polyadenylation sites by interfering with the 3′ processing machinery. EMBO Rep 23: e54686. 10.15252/embr.202254686 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cheng Y, Xie W, Pickering BF, Chu KL, Savino AM, Yang X, Luo H, Nguyen DT, Mo S, Barin E, et al. (2021) N6-methyladenosine on mRNA facilitates a phase-separated nuclear body that suppresses myeloid leukemic differentiation. Cancer Cell 39: 958–972.e8. 10.1016/j.ccell.2021.04.017 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi J, Ieong K-W, Demirci H, Chen J, Petrov A, Prabhakar A, O’Leary SE, Dominissini D, Rechavi G, Soltis SM, et al. (2016) N(6)-methyladenosine in mRNA disrupts tRNA selection and translation-elongation dynamics. Nat Struct Mol Biol 23: 110–115. 10.1038/nsmb.3148 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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 Res 49: 10895–10910. 10.1093/nar/gkab900 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Csepany T, Lin A, Baldick CJ Jr., Beemon K (1990) Sequence specificity of mRNA N6-adenosine methyltransferase. J Biol Chem 265: 20117–20122. 10.1016/s0021-9258(17)30477-5 [DOI] [PubMed] [Google Scholar]
- Czerwoniec A, Dunin-Horkawicz S, Purta E, Kaminska KH, Kasprzak JM, Bujnicki JM, Grosjean H, Rother K (2009) MODOMICS: A database of RNA modification pathways. 2008 update. Nucleic Acids Res 37: D118–D121. 10.1093/nar/gkn710 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davis FF, Allen FW (1957) Ribonucleic acids from yeast which contain a fifth nucleotide. J Biol Chem 227: 907–915. 10.1016/S0021-9258(18)70770-9 [DOI] [PubMed] [Google Scholar]
- Deng S, Zhang J, Su J, Zuo Z, Zeng L, Liu K, Zheng Y, Huang X, Bai R, Zhuang L, et al. (2022) RNA m6A regulates transcription via DNA demethylation and chromatin accessibility. Nat Genet 54: 1427–1437. 10.1038/s41588-022-01173-1 [DOI] [PubMed] [Google Scholar]
- Dominissini D, Moshitch-Moshkovitz S, Schwartz S, Salmon-Divon M, Ungar L, Osenberg S, Cesarkas K, Jacob-Hirsch J, Amariglio N, Kupiec M, et al. (2012) Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature 485: 201–206. 10.1038/nature11112 [DOI] [PubMed] [Google Scholar]
- Dominissini D, Moshitch-Moshkovitz S, Salmon-Divon M, Amariglio N, Rechavi G (2013) Transcriptome-wide mapping of N(6)-methyladenosine by m(6)A-seq based on immunocapturing and massively parallel sequencing. Nat Protoc 8: 176–189. 10.1038/nprot.2012.148 [DOI] [PubMed] [Google Scholar]
- Dou X, Huang L, Xiao Y, Liu C, Li Y, Zhang X, Yu L, Zhao R, Yang L, Chen C, et al. (2023) METTL14 is a chromatin regulator independent of its RNA N6-methyladenosine methyltransferase activity. Protein Cell 14: 683–697. 10.1093/procel/pwad009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. Nat Commun 7: 12626. 10.1038/ncomms12626 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dunin-Horkawicz S, Czerwoniec A, Gajda MJ, Feder M, Grosjean H, Bujnicki JM (2006) MODOMICS: A database of RNA modification pathways. Nucleic Acids Res 34: D145–D149. 10.1093/nar/gkj084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Edupuganti RR, Geiger S, Lindeboom RGH, Shi H, Hsu PJ, Lu Z, Wang S-Y, Baltissen MPA, Jansen PWTC, Rossa M, et al. (2017) N6-methyladenosine (m6A) recruits and repels proteins to regulate mRNA homeostasis. Nat Struct Mol Biol 24: 870–878. 10.1038/nsmb.3462 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gao Y, Vasic R, Song Y, Teng R, Liu C, Gbyli R, Biancon G, Nelakanti R, Lobben K, Kudo E, et al. (2020) m6A modification prevents formation of endogenous double-stranded RNAs and deleterious innate immune responses during hematopoietic development. Immunity 52: 1007–1021.e8. 10.1016/j.immuni.2020.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gilbert WV, Nachtergaele S (2023) mRNA regulation by RNA modifications. Annu Rev Biochem 92: 175–198. 10.1146/annurev-biochem-052521-035949 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goddard TD, Huang CC, Meng EC, Pettersen EF, Couch GS, Morris JH, Ferrin TE (2018) UCSF chimeraX: Meeting modern challenges in visualization and analysis. Protein Sci 27: 14–25. 10.1002/pro.3235 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gor K, Duss O (2023) Emerging quantitative biochemical, structural, and biophysical methods for studying ribosome and protein-RNA complex assembly. Biomolecules 13: 866. 10.3390/biom13050866 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han Z, Niu T, Chang J, Lei X, Zhao M, Wang Q, Cheng W, Wang J, Feng Y, Chai J (2010) Crystal structure of the FTO protein reveals basis for its substrate specificity. Nature 464: 1205–1209. 10.1038/nature08921 [DOI] [PubMed] [Google Scholar]
- 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 Biol 20: e3001535. 10.1371/journal.pbio.3001535 [DOI] [PMC free article] [PubMed] [Google Scholar]
- He PC, Wei J, Dou X, Harada BT, Zhang Z, Ge R, Liu C, Zhang L-S, Yu X, Wang S, et al. (2023) Exon architecture controls mRNA m6A suppression and gene expression. Science 379: 677–682. 10.1126/science.abj9090 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heinrich B, Zhang Z, Raitskin O, Hiller M, Benderska N, Hartmann AM, Bracco L, Elliott D, Ben-Ari S, Soreq H, et al. (2009) Heterogeneous nuclear ribonucleoprotein G regulates splice site selection by binding to CC(A/C)-rich regions in pre-mRNA. J Biol Chem 284: 14303–14315. 10.1074/jbc.M901026200 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong J, Xu K, Lee JH (2022) Biological roles of the RNA m6A modification and its implications in cancer. Exp Mol Med 54: 1822–1832. 10.1038/s12276-022-00897-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Horiuchi K, Kawamura T, Iwanari H, Ohashi R, Naito M, Kodama T, Hamakubo T (2013) Identification of Wilms' tumor 1-associating protein complex and its role in alternative splicing and the cell cycle. J Biol Chem 288: 33292–33302. 10.1074/jbc.M113.500397 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hsu PJ, Zhu Y, Ma H, Guo Y, Shi X, Liu Y, Qi M, Lu Z, Shi H, Wang J, et al. (2017) Ythdc2 is an N6-methyladenosine binding protein that regulates mammalian spermatogenesis. Cell Res 27: 1115–1127. 10.1038/cr.2017.99 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang L, Ashraf S, Wang J, Lilley DM (2017) Control of box C/D snoRNP assembly by N6-methylation of adenine. EMBO Rep 18: 1631–1645. 10.15252/embr.201743967 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang H, Weng H, Sun W, Qin X, Shi H, Wu H, Zhao BS, Mesquita A, Liu C, Yuan CL, et al. (2018) Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation. Nat Cell Biol 20: 285–295. 10.1038/s41556-018-0045-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang H, Weng H, Zhou K, Wu T, Zhao BS, Sun M, Chen Z, Deng X, Xiao G, Auer F, et al. (2019) Histone H3 trimethylation at lysine 36 guides m6A RNA modification co-transcriptionally. Nature 567: 414–419. 10.1038/s41586-019-1016-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang W, Chen T-Q, Fang K, Zeng Z-C, Ye H, Chen Y-Q (2021) N6-methyladenosine methyltransferases: Functions, regulation, and clinical potential. J Hematol Oncol 14: 117. 10.1186/s13045-021-01129-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. Nat Commun 12: 3244. 10.1038/s41467-021-23457-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jain S, Koziej L, Poulis P, Kaczmarczyk I, Gaik M, Rawski M, Ranjan N, Glatt S, Rodnina MV (2023) Modulation of translational decoding by m6A modification of mRNA. Nat Commun 14: 4784. 10.1038/s41467-023-40422-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jia G, Fu Y, Zhao X, Dai Q, Zheng G, Yang Y, Yi C, Lindahl T, Pan T, Yang Y-G, et al. (2011) N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO. Nat Chem Biol 7: 885–887. 10.1038/nchembio.687 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jia G, Yang C-G, Yang S, Jian X, Yi C, Zhou Z, He C (2008) Oxidative demethylation of 3-methylthymine and 3-methyluracil in single-stranded DNA and RNA by mouse and human FTO. FEBS Lett 582: 3313–3319. 10.1016/j.febslet.2008.08.019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jones AN, Tikhaia E, Mourão A, Sattler M (2022) Structural effects of m6A modification of the Xist A-repeat AUCG tetraloop and its recognition by YTHDC1. Nucleic Acids Res 50: 2350–2362. 10.1093/nar/gkac080 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kang HJ, Cheon NY, Park H, Jeong GW, Ye BJ, Yoo EJ, Lee JH, Hur J-H, Lee E-A, Kim H, et al. (2021) TonEBP recognizes R-loops and initiates m6A RNA methylation for R-loop resolution. Nucleic Acids Res 49: 269–284. 10.1093/nar/gkaa1162 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaur S, Tam NY, McDonough MA, Schofield CJ, Aik WS (2022) Mechanisms of substrate recognition and N6-methyladenosine demethylation revealed by crystal structures of ALKBH5-RNA complexes. Nucleic Acids Res 50: 4148–4160. 10.1093/nar/gkac195 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ke S, Alemu EA, Mertens C, Gantman EC, Fak JJ, Mele A, Haripal B, Zucker-Scharff I, Moore MJ, Park CY, et al. (2015) A majority of m6A residues are in the last exons, allowing the potential for 3′ UTR regulation. Genes Dev 29: 2037–2053. 10.1101/gad.269415.115 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ke S, Pandya-Jones A, Saito Y, Fak JJ, Vågbø CB, Geula S, Hanna JH, Black DL, Darnell JE Jr., Darnell RB (2017) m6A mRNA modifications are deposited in nascent pre-mRNA and are not required for splicing but do specify cytoplasmic turnover. Genes Dev 31: 990–1006. 10.1101/gad.301036.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khong A, Matheny T, Huynh TN, Babl V, Parker R (2022) Limited effects of m6A modification on mRNA partitioning into stress granules. Nat Commun 13: 3735. 10.1038/s41467-022-31358-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Knuckles P, Carl SH, Musheev M, Niehrs C, Wenger A, Bühler M (2017) RNA fate determination through cotranscriptional adenosine methylation and microprocessor binding. Nat Struct Mol Biol 24: 561–569. 10.1038/nsmb.3419 [DOI] [PubMed] [Google Scholar]
- Kretschmer J, Rao H, Hackert P, Sloan KE, Höbartner C, Bohnsack MT (2018) The m6A reader protein YTHDC2 interacts with the small ribosomal subunit and the 5′-3′ exoribonuclease XRN1. RNA 24: 1339–1350. 10.1261/rna.064238.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee J-H, Wang R, Xiong F, Krakowiak J, Liao Z, Nguyen PT, Moroz-Omori EV, Shao J, Zhu X, Bolt MJ, et al. (2021) Enhancer RNA m6A methylation facilitates transcriptional condensate formation and gene activation. Mol Cell 81: 3368–3385.e9. 10.1016/j.molcel.2021.07.024 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. J Biomol Screen 21: 290–297. 10.1177/1087057115623264 [DOI] [PubMed] [Google Scholar]
- Li A, Chen YS, Ping XL, Yang X, Xiao W, Yang Y, Sun HY, Zhu Q, Baidya P, Wang X, et al. (2017) Cytoplasmic m6A reader YTHDF3 promotes mRNA translation. Cell Res 27: 444–447. 10.1038/cr.2017.10 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li Y, Xia L, Tan K, Ye X, Zuo Z, Li M, Xiao R, Wang Z, Liu X, Deng M, et al. (2020) N6-methyladenosine co-transcriptionally directs the demethylation of histone H3K9me2. Nat Genet 52: 870–877. 10.1038/s41588-020-0677-3 [DOI] [PubMed] [Google Scholar]
- Lin S, Choe J, Du P, Triboulet R, Gregory RI (2016) The m(6)A methyltransferase METTL3 promotes translation in human cancer cells. Mol Cell 62: 335–345. 10.1016/j.molcel.2016.03.021 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J, Yue Y, Han D, Wang X, Fu Y, Zhang L, Jia G, Yu M, Lu Z, Deng X, et al. (2014) A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation. Nat Chem Biol 10: 93–95. 10.1038/nchembio.1432 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu N, Dai Q, Zheng G, He C, Parisien M, Pan T (2015) N(6)-methyladenosine-dependent RNA structural switches regulate RNA-protein interactions. Nature 518: 560–564. 10.1038/nature14234 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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 Res 45: 6051–6063. 10.1093/nar/gkx141 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu B, Merriman DK, Choi SH, Schumacher MA, Plangger R, Kreutz C, Horner SM, Meyer KD, Al-Hashimi HM (2018) A potentially abundant junctional RNA motif stabilized by m6A and Mg2. Nat Commun 9: 2761. 10.1038/s41467-018-05243-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J, Dou X, Chen C, Chen C, Liu C, Xu MM, Zhao S, Shen B, Gao Y, Han D, et al. (2020) N 6-methyladenosine of chromosome-associated regulatory RNA regulates chromatin state and transcription. Science 367: 580–586. 10.1126/science.aay6018 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu X, Wang H, Zhao X, Luo Q, Wang Q, Tan K, Wang Z, Jiang J, Cui J, Du E, et al. (2021) Arginine methylation of METTL14 promotes RNA N6-methyladenosine modification and endoderm differentiation of mouse embryonic stem cells. Nat Commun 12: 3780. 10.1038/s41467-021-24035-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Louloupi A, Ntini E, Conrad T, Ørom UAV (2018) Transient N-6-methyladenosine transcriptome sequencing reveals a regulatory role of m6A in splicing efficiency. Cell Rep 23: 3429–3437. 10.1016/j.celrep.2018.05.077 [DOI] [PubMed] [Google Scholar]
- Ma H, Wang X, Cai J, Dai Q, Natchiar SK, Lv R, Chen K, Lu Z, Chen H, Shi YG, et al. (2019) N6-methyladenosine methyltransferase ZCCHC4 mediates ribosomal RNA methylation. Nat Chem Biol 15: 88–94. 10.1038/s41589-018-0184-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ma M, Harding HP, O’Rahilly S, Ron D, Yeo GSH (2012) Kinetic analysis of FTO (fat mass and obesity-associated) reveals that it is unlikely to function as a sensor for 2-oxoglutarate. Biochem J 444: 183–187. 10.1042/BJ20120065 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Machnicka MA, Milanowska K, Osman Oglou O, Purta E, Kurkowska M, Olchowik A, Januszewski W, Kalinowski S, Dunin-Horkawicz S, Rother KM, et al. (2013) MODOMICS: A database of RNA modification pathways--2013 update. Nucleic Acids Res 41: D262–D267. 10.1093/nar/gks1007 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martinez NM, Gilbert WV (2018) Pre-mRNA modifications and their role in nuclear processing. Quant Biol 6: 210–227. 10.1007/s40484-018-0147-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mauer J, Luo X, Blanjoie A, Jiao X, Grozhik AV, Patil DP, Linder B, Pickering BF, Vasseur J-J, Chen Q, et al. (2017) Reversible methylation of m6Am in the 5′ cap controls mRNA stability. Nature 541: 371–375. 10.1038/nature21022 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. Biol Chem 402: 89–98. 10.1515/hsz-2020-0265 [DOI] [PubMed] [Google Scholar]
- 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: 1635–1646. 10.1016/j.cell.2012.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Meyer KD, Patil DP, Zhou J, Zinoviev A, Skabkin MA, Elemento O, Pestova TV, Qian S-B, Jaffrey SR (2015) 5′ UTR m(6)A promotes cap-independent translation. Cell 163: 999–1010. 10.1016/j.cell.2015.10.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mirdita M, Schütze K, Moriwaki Y, Heo L, Ovchinnikov S, Steinegger M (2022) ColabFold: Making protein folding accessible to all. Nat Methods 19: 679–682. 10.1038/s41592-022-01488-1 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Motorin Y, Helm M (2022) RNA nucleotide methylation: 2021 update. Wiley Interdiscip Rev RNA 13: e1691. 10.1002/wrna.1691 [DOI] [PubMed] [Google Scholar]
- Nicastro G, Abis G, Klein P, Esteban-Serna S, Gallagher C, Chaves-Arquero B, Cai Y, Figueiredo AM, Martin SR, Patani R, et al. (2023) Direct m6A recognition by IMP1 underlays an alternative model of target selection for non-canonical methyl-readers. Nucleic Acids Res 51: 8774–8786. 10.1093/nar/gkad534 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orsolic I, Carrier A, Esteller M (2023) Genetic and epigenetic defects of the RNA modification machinery in cancer. Trends Genet 39: 74–88. 10.1016/j.tig.2022.10.004 [DOI] [PubMed] [Google Scholar]
- Park OH, Ha H, Lee Y, Boo SH, Kwon DH, Song HK, Kim YK (2019) Endoribonucleolytic cleavage of m6A-containing RNAs by RNase P/MRP complex. Mol Cell 74: 494–507.e8. 10.1016/j.molcel.2019.02.034 [DOI] [PubMed] [Google Scholar]
- Parker MT, Soanes BK, Kusakina J, Larrieu A, Knop K, Joy N, Breidenbach F, Sherwood AV, Barton GJ, Fica SM, et al. (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]
- Patil DP, Chen C-K, Pickering BF, Chow A, Jackson C, Guttman M, Jaffrey SR (2016) m(6)A RNA methylation promotes XIST-mediated transcriptional repression. Nature 537: 369–373. 10.1038/nature19342 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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: 824–835.e14. 10.1016/j.cell.2017.05.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Peng Z, Ma J, Christov CZ, Karabencheva-Christove T, Lehnert N, Li D (2023) Kinetic studies on the 2-oxoglutarate/Fe(II)-dependent nucleic acid modifying enzymes from the AlkB and TET families. DNA 3: 65–84. 10.3390/dna3020005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perez-Pepe M, Desotell AW, Li H, Li W, Han B, Lin Q, Klein DE, Liu Y, Goodarzi H, Alarcón CR (2023) 7SK methylation by METTL3 promotes transcriptional activity. Sci Adv 9: eade7500. 10.1126/sciadv.ade7500 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pettersen EF, Goddard TD, Huang CC, Meng EC, Couch GS, Croll TI, Morris JH, Ferrin TE (2021) UCSF ChimeraX: Structure visualization for researchers, educators, and developers. Protein Sci 30: 70–82. 10.1002/pro.3943 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ping X-L, Sun B-F, Wang L, Xiao W, Yang X, Wang W-J, Adhikari S, Shi Y, Lv Y, Chen Y-S, et al. (2014) Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res 24: 177–189. 10.1038/cr.2014.3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qi S, Mota J, Chan S-H, Villarreal J, Dai N, Arya S, Hromas RA, Rao MK, Corrêa IR Jr., Gupta YK (2022) RNA binding to human METTL3-METTL14 restricts N6-deoxyadenosine methylation of DNA in vitro. Elife 11: e67150. 10.7554/eLife.67150 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Qiu W, Zhang Q, Zhang R, Lu Y, Wang X, Tian H, Yang Y, Gu Z, Gao Y, Yang X, et al. (2021) N6-methyladenosine RNA modification suppresses antiviral innate sensing pathways via reshaping double-stranded RNA. Nat Commun 12: 1582. 10.1038/s41467-021-21904-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- Relier S, Rivals E, David A (2022) The multifaceted functions of the fat mass and obesity-associated protein (FTO) in normal and cancer cells. RNA Biol 19: 132–142. 10.1080/15476286.2021.2016203 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ren W, Lu J, Huang M, Gao L, Li D, Wang GG, Song J (2019) Structure and regulation of ZCCHC4 in m6A-methylation of 28S rRNA. Nat Commun 10: 5042. 10.1038/s41467-019-12923-x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roost C, Lynch SR, Batista PJ, Qu K, Chang HY, Kool ET (2015) Structure and thermodynamics of N6-methyladenosine in RNA: A spring-loaded base modification. J Am Chem Soc 137: 2107–2115. 10.1021/ja513080v [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roundtree IA, Luo G-Z, Zhang Z, Wang X, Zhou T, Cui Y, Sha J, Huang X, Guerrero L, Xie P, et al. (2017) YTHDC1 mediates nuclear export of N6-methyladenosine methylated mRNAs. Elife 6: e31311. 10.7554/eLife.31311 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sacco MT, Bland KM, Horner SM (2022) WTAP targets the METTL3 m6A-methyltransferase complex to cytoplasmic hepatitis C virus RNA to regulate infection. J Virol 96: e0099722. 10.1128/jvi.00997-22 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Satterwhite ER, Mansfield KD (2022) RNA methyltransferase METTL16: Targets and function. Wiley Interdiscip Rev RNA 13: e1681. 10.1002/wrna.1681 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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: 499–512. 10.1261/rna.064063.117 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sendinc E, Shi Y (2023) RNA m6A methylation across the transcriptome. Mol Cell 83: 428–441. 10.1016/j.molcel.2023.01.006 [DOI] [PubMed] [Google Scholar]
- 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. Mol Cell 75: 620–630.e9. 10.1016/j.molcel.2019.05.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sepich-Poore C, Zheng Z, Schmitt E, Wen K, Zhang ZS, Cui X-L, Dai Q, Zhu AC, Zhang L, Sanchez Castillo A, et al. (2022) The METTL5-TRMT112 N6-methyladenosine methyltransferase complex regulates mRNA translation via 18S rRNA methylation. J Biol Chem 298: 101590. 10.1016/j.jbc.2022.101590 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi H, Wang X, Lu Z, Zhao BS, Ma H, Hsu PJ, Liu C, He C (2017) YTHDF3 facilitates translation and decay of N6-methyladenosine-modified RNA. Cell Res 27: 315–328. 10.1038/cr.2017.15 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi H, Liu B, Nussbaumer F, Rangadurai A, Kreutz C, Al-Hashimi HM (2019) NMR chemical exchange measurements reveal that N6-methyladenosine slows RNA annealing. J Am Chem Soc 141: 19988–19993. 10.1021/jacs.9b10939 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shi H, Chai P, Jia R, Fan X (2020) Novel insight into the regulatory roles of diverse RNA modifications: Re-defining the bridge between transcription and translation. Mol Cancer 19: 78. 10.1186/s12943-020-01194-6 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sikorski V, Selberg S, Lalowski M, Karelson M, Kankuri E (2023) The structure and function of YTHDF epitranscriptomic m6A readers. Trends Pharmacol Sci 44: 335–353. 10.1016/j.tips.2023.03.004 [DOI] [PubMed] [Google Scholar]
- Śledź P, Jinek M (2016) Structural insights into the molecular mechanism of the m(6)A writer complex. Elife 5: e18434. 10.7554/eLife.18434 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Slobodin B, Han R, Calderone V, Vrielink JAFO, Loayza-Puch F, Elkon R, Agami R (2017) Transcription impacts the efficiency of mRNA translation via co-transcriptional N6-adenosine methylation. Cell 169: 326–337.e12. 10.1016/j.cell.2017.03.031 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Song H, Wang Y, Wang R, Zhang X, Liu Y, Jia G, Chen PR (2020) SFPQ is an FTO-binding protein that facilitates the demethylation substrate preference. Cell Chem Biol 27: 283–291.e6. 10.1016/j.chembiol.2020.01.002 [DOI] [PubMed] [Google Scholar]
- 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 Res 32: 982–994. 10.1038/s41422-022-00725-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sun L, Fazal FM, Li P, Broughton JP, Lee B, Tang L, Huang W, Kool ET, Chang HY, Zhang QC (2019) RNA structure maps across mammalian cellular compartments. Nat Struct Mol Biol 26: 322–330. 10.1038/s41594-019-0200-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tan X, Zheng C, Zhuang Y, Jin P, Wang F (2023) The m6A reader PRRC2A is essential for meiosis I completion during spermatogenesis. Nat Commun 14: 1636. 10.1038/s41467-023-37252-y [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thalhammer A, Bencokova Z, Poole R, Loenarz C, Adam J, O’Flaherty L, Schödel J, Mole D, Giaslakiotis K, Schofield CJ, et al. (2011) Human AlkB homologue 5 is a nuclear 2-oxoglutarate dependent oxygenase and a direct target of hypoxia-inducible factor 1α (HIF-1α). PLoS One 6: e16210. 10.1371/journal.pone.0016210 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Theler D, Dominguez C, Blatter M, Boudet J, Allain FH-T (2014) Solution structure of the YTH domain in complex with N6-methyladenosine RNA: A reader of methylated RNA. Nucleic Acids Res 42: 13911–13919. 10.1093/nar/gku1116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Uzonyi A, Dierks D, Nir R, Kwon OS, Toth U, Barbosa I, Burel C, Brandis A, Rossmanith W, Le Hir H, et al. (2023) Exclusion of m6A from splice-site proximal regions by the exon junction complex dictates m6A topologies and mRNA stability. Mol Cell 83: 237–251.e7. 10.1016/j.molcel.2022.12.026 [DOI] [PubMed] [Google Scholar]
- Vaidyanathan PP, AlSadhan I, Merriman DK, Al-Hashimi HM, Herschlag D (2017) Pseudouridine and N6-methyladenosine modifications weaken PUF protein/RNA interactions. RNA 23: 611–618. 10.1261/rna.060053.116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Tran N, Ernst FGM, Hawley BR, Zorbas C, Ulryck N, Hackert P, Bohnsack KE, Bohnsack MT, Jaffrey SR, Graille M, et al. (2019) The human 18S rRNA m6A methyltransferase METTL5 is stabilized by TRMT112. Nucleic Acids Res 47: 7719–7733. 10.1093/nar/gkz619 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vendeix FAP, Munoz AM, Agris PF (2009) Free energy calculation of modified base-pair formation in explicit solvent: A predictive model. RNA 15: 2278–2287. 10.1261/rna.1734309 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Viswanathan SR, Daley GQ, Gregory RI (2008) Selective blockade of microRNA processing by Lin28. Science 320: 97–100. 10.1126/science.1154040 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Lu Z, Gomez A, Hon GC, Yue Y, Han D, Fu Y, Parisien M, Dai Q, Jia G, et al. (2014) N6-methyladenosine-dependent regulation of messenger RNA stability. Nature 505: 117–120. 10.1038/nature12730 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X, Zhao BS, Roundtree IA, Lu Z, Han D, Ma H, Weng X, Chen K, Shi H, He C (2015) N(6)-methyladenosine modulates messenger RNA translation efficiency. Cell 161: 1388–1399. 10.1016/j.cell.2015.05.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang P, Doxtader KA, Nam Y (2016) Structural basis for cooperative function of Mettl3 and Mettl14 methyltransferases. Mol Cell 63: 306–317. 10.1016/j.molcel.2016.05.041 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang S, Song Y, Wang Y, Li X, Fu B, Liu Y, Wang J, Wei L, Tian T, Zhou X (2017) The m6A methylation perturbs the hoogsteen pairing-guided incorporation of an oxidized nucleotide. Chem Sci 8: 6380–6388. 10.1039/c7sc02340e [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y, Traugot CM, Bubenik JL, Li T, Sheng P, Hiers NM, Fernandez P, Li L, Bian J, Swanson MS, et al. (2023) N6-methyladenosine in 7SK small nuclear RNA underlies RNA polymerase II transcription regulation. Mol Cell 83: 3818–3834.e7. 10.1016/j.molcel.2023.09.020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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 Rep 18: 2004–2014. 10.15252/embr.201744940 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Watkins NJ, Bohnsack MT (2012) The box C/D and H/ACA snoRNPs: key players in the modification, processing and the dynamic folding of ribosomal RNA. Wiley Interdiscip Rev RNA 3: 397–414. 10.1002/wrna.117 [DOI] [PubMed] [Google Scholar]
- Wei J, Liu F, Lu Z, Fei Q, Ai Y, He PC, Shi H, Cui X, Su R, Klungland A, et al. (2018) Differential m6A, m6Am, and m1A demethylation mediated by FTO in the cell nucleus and cytoplasm. Mol Cell 71: 973–985.e5. 10.1016/j.molcel.2018.08.011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wen J, Lv R, Ma H, Shen H, He C, Wang J, Jiao F, Liu H, Yang P, Tan L, et al. (2018) Zc3h13 regulates nuclear RNA m6A methylation and mouse embryonic stem cell self-renewal. Mol Cell 69: 1028–1038.e6. 10.1016/j.molcel.2018.02.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Widagdo J, Anggono V, Wong JJ-L (2022) The multifaceted effects of YTHDC1-mediated nuclear m6A recognition. Trends Genet 38: 325–332. 10.1016/j.tig.2021.11.005 [DOI] [PubMed] [Google Scholar]
- Wu B, Su S, Patil DP, Liu H, Gan J, Jaffrey SR, Ma J (2018) Molecular basis for the specific and multivariant recognitions of RNA substrates by human hnRNP A2/B1. Nat Commun 9: 420. 10.1038/s41467-017-02770-z [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu R, Li A, Sun B, Sun J-G, Zhang J, Zhang T, Chen Y, Xiao Y, Gao Y, Zhang Q, et al. (2019) A novel m6A reader Prrc2a controls oligodendroglial specification and myelination. Cell Res 29: 23–41. 10.1038/s41422-018-0113-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wu C, Chen W, He J, Jin S, Liu Y, Yi Y, Gao Z, Yang J, Yang J, Cui J, et al. (2020) Interplay of m6A and H3K27 trimethylation restrains inflammation during bacterial infection. Sci Adv 6: eaba0647. 10.1126/sciadv.aba0647 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xiao W, Adhikari S, Dahal U, Chen Y-S, Hao Y-J, Sun B-F, Sun H-Y, Li A, Ping X-L, Lai W-Y, et al. (2016) Nuclear m(6)A reader YTHDC1 regulates mRNA splicing. Mol Cell 61: 507–519. 10.1016/j.molcel.2016.01.012 [DOI] [PubMed] [Google Scholar]
- 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. Nat Chem Biol 10: 927–929. 10.1038/nchembio.1654 [DOI] [PubMed] [Google Scholar]
- Xu W, Li J, He C, Wen J, Ma H, Rong B, Diao J, Wang L, Wang J, Wu F, et al. (2021) METTL3 regulates heterochromatin in mouse embryonic stem cells. Nature 591: 317–321. 10.1038/s41586-021-03210-1 [DOI] [PubMed] [Google Scholar]
- Xu W, He C, Kaye EG, Li J, Mu M, Nelson GM, Dong L, Wang J, Wu F, Shi YG, et al. (2022) Dynamic control of chromatin-associated m6A methylation regulates nascent RNA synthesis. Mol Cell 82: 1156–1168.e7. 10.1016/j.molcel.2022.02.006 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yan X, Liu F, Yan J, Hou M, Sun M, Zhang D, Gong Z, Dong X, Tang C, Yin P (2023) WTAP-VIRMA counteracts dsDNA binding of the m6A writer METTL3-METTL14 complex and maintains N6-adenosine methylation activity. Cell Discov 9: 100. 10.1038/s41421-023-00604-5 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yoshida A, Oyoshi T, Suda A, Futaki S, Imanishi M (2022) Recognition of G-quadruplex RNA by a crucial RNA methyltransferase component, METTL14. Nucleic Acids Res 50: 449–457. 10.1093/nar/gkab1211 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu D, Horton JR, Yang J, Hajian T, Vedadi M, Sagum CA, Bedford MT, Blumenthal RM, Zhang X, Cheng X (2021. a) Human MettL3-MettL14 RNA adenine methyltransferase complex is active on double-stranded DNA containing lesions. Nucleic Acids Res 49: 11629–11642. 10.1093/nar/gkab460 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu D, Kaur G, Blumenthal RM, Zhang X, Cheng X (2021. b) Enzymatic characterization of three human RNA adenosine methyltransferases reveals diverse substrate affinities and reaction optima. J Biol Chem 296: 100270. 10.1016/j.jbc.2021.100270 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu F, Zhu AC, Liu S, Gao B, Wang Y, Khudaverdyan N, Yu C, Wu Q, Jiang Y, Song J, et al. (2023) RBM33 is a unique m6A RNA-binding protein that regulates ALKBH5 demethylase activity and substrate selectivity. Mol Cell 83: 2003–2019.e6. 10.1016/j.molcel.2023.05.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yue Y, Liu J, Cui X, Cao J, Luo G, Zhang Z, Cheng T, Gao M, Shu X, Ma H, et al. (2018) VIRMA mediates preferential m6A mRNA methylation in 3′ UTR and near stop codon and associates with alternative polyadenylation. Cell Discov 4: 10. 10.1038/s41421-018-0019-0 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zaccara S, Jaffrey SR (2020) A unified model for the function of YTHDF proteins in regulating m6A-modified mRNA. Cell 181: 1582–1595.e18. 10.1016/j.cell.2020.05.012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zaccara S, Ries RJ, Jaffrey SR (2019) Reading, writing and erasing mRNA methylation. Nat Rev Mol Cell Biol 20: 608–624. 10.1038/s41580-019-0168-5 [DOI] [PubMed] [Google Scholar]
- Zeng Z-C, Pan Q, Sun Y-M, Huang H-J, Chen X-T, Chen T-Q, He B, Ye H, Zhu S-X, Pu K-J, et al. (2023) METTL3 protects METTL14 from STUB1-mediated degradation to maintain m6 A homeostasis. EMBO Rep 24: e55762. 10.15252/embr.202255762 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang X, Wei L-H, Wang Y, Xiao Y, Liu J, Zhang W, Yan N, Amu G, Tang X, Zhang L, et al. (2019) Structural insights into FTO's catalytic mechanism for the demethylation of multiple RNA substrates. Proc Natl Acad Sci USA 116: 2919–2924. 10.1073/pnas.1820574116 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zheng G, Dahl JA, Niu Y, Fedorcsak P, Huang C-M, Li CJ, Vågbø CB, Shi Y, Wang W-L, Song S-H, et al. (2013) ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol Cell 49: 18–29. 10.1016/j.molcel.2012.10.015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou J, Wan J, Gao X, Zhang X, Jaffrey SR, Qian S-B (2015) Dynamic m(6)A mRNA methylation directs translational control of heat shock response. Nature 526: 591–594. 10.1038/nature15377 [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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. Mol Cell 76: 70–81.e9. 10.1016/j.molcel.2019.07.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou W, Han Z, Wu Z, Gong W, Yang S, Chen L, Li C (2022) Specific recognition between YTHDF3 and m6 A-modified RNA: An all-atom molecular dynamics simulation study. Proteins 90: 1965–1972. 10.1002/prot.26389 [DOI] [PubMed] [Google Scholar]
- Zhu T, Roundtree IA, Wang P, Wang X, Wang L, Sun C, Tian Y, Li J, He C, Xu Y (2014) Crystal structure of the YTH domain of YTHDF2 reveals mechanism for recognition of N6-methyladenosine. Cell Res 24: 1493–1496. 10.1038/cr.2014.152 [DOI] [PMC free article] [PubMed] [Google Scholar]




