Next Article in Journal
Comparative Transcriptomic Analysis of Genes in the 20-Hydroxyecdysone Biosynthesis in the Fern Microsorum scolopendria towards Challenges with Foliar Application of Chitosan
Previous Article in Journal
Effect of Epstein–Barr Virus Infection on Selected Immunological Parameters in Children with Type 1 Diabetes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Repertoire of RNA Modifications Orchestrates a Plethora of Cellular Responses

by
Panagiotis G. Adamopoulos
*,
Konstantina Athanasopoulou
,
Glykeria N. Daneva
and
Andreas Scorilas
Department of Biochemistry and Molecular Biology, Faculty of Biology, National and Kapodistrian University of Athens, 15701 Athens, Greece
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(3), 2387; https://doi.org/10.3390/ijms24032387
Submission received: 31 December 2022 / Revised: 21 January 2023 / Accepted: 23 January 2023 / Published: 25 January 2023
(This article belongs to the Special Issue RNA Roles Related to Cellular Growth and Metabolism)

Abstract

:
Although a plethora of DNA modifications have been extensively investigated in the last decade, recent breakthroughs in molecular biology, including high throughput sequencing techniques, have enabled the identification of post-transcriptional marks that decorate RNAs; hence, epitranscriptomics has arisen. This recent scientific field aims to decode the regulatory layer of the transcriptome and set the ground for the detection of modifications in ribose nucleotides. Until now, more than 170 RNA modifications have been reported in diverse types of RNA that contribute to various biological processes, such as RNA biogenesis, stability, and transcriptional and translational accuracy. However, dysfunctions in the RNA-modifying enzymes that regulate their dynamic level can lead to human diseases and cancer. The present review aims to highlight the epitranscriptomic landscape in human RNAs and match the catalytic proteins with the deposition or deletion of a specific mark. In the current review, the most abundant RNA modifications, such as N6-methyladenosine (m6A), N5-methylcytosine (m5C), pseudouridine (Ψ) and inosine (I), are thoroughly described, their functional and regulatory roles are discussed and their contributions to cellular homeostasis are stated. Ultimately, the involvement of the RNA modifications and their writers, erasers, and readers in human diseases and cancer is also discussed.

1. Introduction

Ever since the central dogma of molecular biology was proposed by Francis Crick, our perception of the flow of genetic information has been dramatically broadened. Over the years, the flux of information has revealed numerous processing steps that ensure proper gene expression and has highlighted the high complexity of the human genome, paving the way for deciphering the hidden aspects of life [1]. Moreover, the identification of various transcription factors and the functional clarification of the transcriptional machinery, as well as the rise of genomics, has provided new insights into the genetic programs that govern human development [2]. Although a plethora of chemical DNA modifications had been reported decades before the sequencing of the first nucleic acid, recent breakthroughs in sequencing techniques has enabled the in-depth study of genomes and has led to the introduction of epigenomics [3]. In the same manner, through the rise of epitranscriptomics, RNA sequencing attempts to decode the regulatory layer that rests between genome and proteome, namely transcriptome, and set the ground for the detection of modifications in ribose nucleotides [4,5].
For decades, it was known that RNA molecules were decorated with various chemical modifications, but it was only after the revolution in molecular biology and the advent of RNA sequencing techniques that the emerging field of epitranscriptomics spawned. Until now, more than hundreds of multiple RNA chemical modifications have been reported, whereas only a few of them have been systematically studied [6] (Figure 1). These post-transcriptional changes have been observed in a wide spectrum of newly transcribed RNAs, including transfer RNAs (tRNAs), messenger RNAs (mRNAs), ribosomal RNAs (rRNAs), small non-coding RNAs, circular RNAs (circRNAs) and long non-coding RNAs (lncRNAs) (Supplementary Table S1).
RNA modifications are implicated in the regulation of several facets of RNA processing, thus possessing an indispensable role for the generation of functional RNA molecules [7,8]. These changes can be added or removed by different types of enzymes that catalyze their biosynthetic reactions. More precisely, the deposition of the chemical marks onto the RNAs is mediated by “writers”, a protein family that forms multiprotein complexes. The “writers” constitute a class of enzymes that act on RNAs and covalently introduce methyl and/or acetyl groups into the RNA nucleotides [9,10]. On the contrary, “erasers” represent enzymes that remove these epitranscriptomic marks, whereas “readers” are binding proteins responsible for the selective recognition of RNA modifications [11]. Notably, each modification is deposited, erased, or read by different members of the respective protein group that causes the alteration.
Interestingly, N6-methyladenosine (m6A), N5-methylcytosine (m5C), pseudouridine (Ψ) and inosine (I) are the most abundant RNA modifications in eukaryotic cells and their roles have been extensively investigated. Additionally, 7-methylguanosine (m7G), N1-methyladenosine (m1A), 3-methylcytidine (m3C), queuosine (Q), uridine (U) and ribose methylation (2′-O-Me or Nm) are RNA-related chemical marks that also fall under the scope of modern epitranscriptomics research (Supplementary Table S1) [6]. The present review aims to depict the human epitranscriptomic landscape in various types of RNA molecules and match the different RNA-modifying proteins with a specific mark. An additional goal is to summarize the functional and regulatory roles of each modification, among the different types of RNAs, in order to elucidate their contribution to cellular homeostasis. Ultimately, a thorough investigation of their involvement in human diseases, including cancer, is also discussed.

2. The Landscape of RNA Modifications

The widespread post-transcriptional process of RNA editing affects all nucleotides: A, U, C, and G. Of note, RNA modifications have been reported in all RNA types, including coding and non-coding ones. The abundancy of each modification among the different RNA species varies and depends on the phase of the cell cycle, the cell type, as well as the cellular responses and requirements. Additionally, it should be noted that several RNA modifications (e.g., m1A, I) can affect the Watson–Crick base-pairing, leading to either the misincorporation of nucleotides in the cDNA synthesis procedure or a blockage in the reverse transcriptase (Table 1). Consequently, these chemical marks induce ‘modified base pairs’ that impact on the RNA folding, 3D structures and flexibility of the molecules [12].

2.1. Chemical Modifications of Adenine RNA Base

Adenine, the most heavily modified RNA nucleotide, can be altered in different ways; this includes, for example, methylations on the nitrogen atoms of the adenine base generating m6A and m1A modifications, as well as conversions of adenine to inosine (A-to-I).

2.1.1. N6-Methyladenosine (m6A)

Although it has been over 40 years since m6A was discovered, it is still the center of attention; many studies aim to investigate this type of RNA modification and its functionality, since it constitutes the most prevalent internal RNA change both in yeast and higher eukaryotes [13]. Going back to 1975, different research teams found that approximately 0.3% of total adenine bases are N6-methyl-modified [14,15], indicating its high abundancy in the transcriptome. Recent studies have confirmed that more than 8000 human mRNAs and more than 300 ncRNAs harbor m6A sites [16]. More precisely, Meyer et al. identified >40,000 m6A peaks in human RNAs, while their bioinformatics analysis supported that ~95% of the m6A are located at mRNAs, whereas <2% of the m6A sites are found in ncRNAs [16]. In human cells, m6A is in abundance on mRNAs, highlighting its involvement in multiple biological processes, whereas its presence in rRNAs affects translation. It should be noted that m6A methylation has also been detected in other RNA species, such as microRNAs (miRNAs) and circRNAs, contributing to their biogenesis and maturation; meanwhile, in lncRNAs m6A, bases are responsible for mediating transcriptional repression, thus affecting their functionality (Supplementary Table S1) [17]. On the contrary, although several epitranscriptomic studies focusing on tRNAs support the existence of a plethora of modifications, there is no evidence of m6A in the nucleotide sequence of the human tRNAs.
The deposition of a methyl group onto the sixth nitrogen atom of A is catalyzed by specific enzymes, called m6A-methyltransferases [11]. More precisely, Methyltransferase-like 3 (METTL3) is the first confirmed methyltransferase that can individually incorporate m6A modifications into mRNAs and lncRNAs [18]. In the same manner, the high homolog of METTL3, Methyltransferase-like 14 (METTL14), catalyzes the formation of m6A (Figure 2). Multiple studies have reported that both METTL3 and METTL14 exhibit catalytic functions by themselves, indicating that they constitute the most significant m6A writers [19].
However, these enzymes form a multi-subunit complex that mediates the generation of m6A sites, synergistically targeting specific sequences known as DRACH motifs (D = G/A/U, R = G/A, H = A/U/C). Additional proteins, such as the Wilms tumor 1-associated protein (WTAP), the Vir-like m6A methyltransferase-associated (VIRMA) protein, the RNA binding motif protein 15/15B (RBM15/15B), and the Zinc-finger CCCH-type-containing 13 (ZC3H13) protein, are implicated in the formation and regulation of the m6A methyltransferase complex [20,21]. For instance, WTAP stabilizes the methyltransferase complex, whereas VIRMA guides the complex near the 3′ untranslated regions (UTRs) and the stop codon regions of the mRNAs [22,23,24]. Furthermore, except from these fundamental methyltransferases, it has been recently verified that an additional member of the protein family, Methyltransferase-like 16 (METTL16), can act in an independent manner, catalyzing the m6A formation in both mRNAs and lncRNAs [25,26].
In the case of miRNAs, various scientific reports have supported that pri-miRNAs can be methylated by either METTL3 or METTL14, which facilitate miRNA maturation [27,28]. In the same manner, METTL3 regulates m6A levels in circRNAs [29]. Interestingly, although limited information is known about the enzymes that are responsible for base modifications on human rRNAs, it has been proven that Methyltransferase-like 5 (METTL5) is the m6A “writer” of the 18S rRNA, which is stabilized by TRMT112, whereas ZCCHC4 acts on the 28S rRNA (Table 1) [30,31,32].
On the other hand, only two demethylases have emerged as the universal molecular “erasers” of m6A, the fat mass and obesity-associated protein (FTO), and the Human AlkB homolog H5 (ALKBH5) protein [33,34]; meanwhile, various binding proteins can read m6A alterations, including members of the YT521-B homology domain-containing proteins (YTHDF1-3 and YTHDC1-2), insulin-like growth factor 2 (IGF2), mRNA-binding proteins (IGF2BP1-3), Fragile X messenger ribonucleoprotein 1 (FMR1), Leucine-rich PPR motif-containing protein (LRPPRC), heterogeneous nuclear ribonucleoprotein (HNRNP) protein family members (HNRNPC and HNRNPA2B1) and the eukaryotic initiation factor 3 (eIF3) [10,11,13,35].
Undoubtedly, the existence of m6A in multiple RNA types manifests its involvement in the regulation of almost all phases of RNA metabolism, including RNA structure, localization, stability, and shelf life [21]. In human rRNAs, two m6A sites have been reported: one at position A1832 in 18S rRNA, as well as one at position A4220 in 28S rRNA (Table 2 and Table 3) [32,36,37]. Structural analyses have revealed that these RNA modifications are located at catalytic ribosomal regions and, therefore, may affect the function of the ribosomes being involved in different processes, such as ribosome heterogeneity, and translation rate and efficiency [38].
Recently, multiple studies have focused on the functional role of m6A modification in mRNAs, and its involvement in the mRNA metabolism and fate. m6A sites have been detected in various sites across the mRNAs of eukaryotic cells, such as the 5′ and 3′ UTRs, and the coding sequence (CDS); however, it is particularly enriched in 3′-UTRs around the termination codons. In particular, approximately 50% of the m6A sites are located at the last exon of the transcripts [16]. In addition, m6A modifications have been observed near exonic splice junctions in the CDS region and have been shown to affect the alternative splicing of human precursor mRNAs. More precisely, the mechanism involves the binding of the nuclear reader YTHDC1 to m6A sites, which recruits multiple splicing factors, including the SRSF3 that enhances exon inclusion and regulates the mRNA splicing by directly guiding the spliceosome in the appropriate splice sites [13,39].
The nuclear export of mature transcripts is catalyzed by writers, erasers, and readers of m6A. Specifically, m6A hypermethylated sites facilitate the transportation of the mRNA into the cytosol by their binding to the nuclear protein YTHDC1, which delivers the molecule to nuclear export proteins [40]. Furthermore, into the cytosol, the m6A methylation regulates protein synthesis through a plethora of enzymes that cooperate to increase translation efficiency [41,42]. METTL3 incorporates m6A into the UTRs and binds to eIF3 to form the mRNA loop and promote translation, whereas YTHDF1 and YTHDF3 can also enhance translation through the YTHDF-eIF3 pathway [43,44]. Notably, the recruitment of m6A readers is also necessary for maintaining mRNA stability, while m6A sites regulate secondary structures [45]. Associated studies have supported that, in human cells, the downregulation of the writer protein METTL3 has led to an increase in the half-life of the mRNAs, indicating that m6A levels directly affect the mRNA’s fate [46].

2.1.2. N1-Methyladenosine (m1A)

The m1A constitutes an additional methylation of adenine and the m1A RNA-modifying proteins regulate its dynamic abundancy in both ncRNAs and mRNAs. Most m1A sites have been identified at specific positions in tRNAs, around the GC-rich regions, (Figure 3), whereas m1A is conserved in position 1309 of the human 28S rRNA (Table 3).
M1A has also been found in specific sites in mitochondrial RNA (mt-RNA), as well as in HGGAGRA motifs ((H = A/U/C, R = G/A) of the lncRNAs [47]. In the case of mRNAs, studies have shown that only one m1A modification is present in each transcript within the GUUCNANNC sequence, which is mainly located in the 5′ UTRs and/or in the first splice site [48,49]. In contrast to m6A, the addition of a methyl group into the first atom of the base is catalyzed by TRMT10 and the TRMT6/TRMT61 complex, which are members of the tRNA methyltransferase protein family (TRMT); meanwhile, ALKBH1 and ALKBH3 are the key enzymes that erase this type of modification (Table 1). Notably, Nucleomethylin (NML), an additional methylase, introduces methylations on 28S rRNA [50,51]. As for the proteins that recognize m1A sites, it has been shown that YTHDF2 and YTHDF3 have a low binding affinity for m1A. Although Heat-responsive protein 12 (HRSP12) is characterized as a factor involved in the RNase P/MRP-mediated endoribonucleolytic cleavage of m6A, a recent report supports that HRSP12 functions as a reader protein for m1A [52].
To continue, m1A methylations may play a regulatory or stabilizing role in modified RNAs. As for the tRNA molecules, m1A in position 58 stabilizes the T-loop structure. Especially in tRNAiMet, the m1A58 increases translation by activating polysomes [53]. On the contrary, it is known that the overexpression of m1A erasers decreases the translation levels, while in the case of ALKBH3 overexpression, tRNA fragments (tRFs) can be formed through tRNA cleavage [54]. In rRNAs, m1A participates in the formation of the 60S subunit of the 80S complex [55]. Although m1A alterations in 5′ UTRs of the mRNAs enhance translation efficiency, m1A in the CDS prevents protein synthesis [56]. Finally, the m1A YTHDF2 and YTHDF3 proteins mediate the mRNA destabilization [50,57].

2.1.3. Adenine to Inosine (A-to-I)

In humans, the most common RNA editing procedure that results in the substitution of a specific nucleotide is adenine-to-inosine (A-to-I); in mRNA, this is catalyzed by adenosine deaminase enzymes, namely ADARs, that act on double-stranded RNAs (dsRNAs). The ADAR family includes three proteins, which are encoded by three ADAR genes [58]. The typical protein structure of ADARs includes two distinct domains: the dsRNA binding domain in the N-terminal and the deaminase domain in the C-terminal. More precisely, ADAR1 and ADAR2 are globally expressed, possess a well-characterized adenosine deamination activity and typically act on pre-mRNAs; however, in the case of ADAR3, which is expressed in brain tissues, no deamination activity has been reported and thus its function remains unclear (Table 1) [59]. However, recent studies have connected ADAR3 with RNA editing inhibition, suggesting that ADAR3 acts as a negative regulator of A-to-I editing [60,61]. Especially on tRNAs, the formation of inosine at position 34 is catalyzed by the ADAT2/ADAT3 complex (Figure 3) [62,63].
Recent RNA-seq studies have confirmed a great number of A-to-I RNA editing sites in human mRNAs. Interestingly, the majority of inosine residues are located at UTRs and intronic regions; meanwhile, approximately 1000 editing sites have been found in protein-coding regions, indicating that A-to-I editing has two distinct and critical roles in RNAs [58,64,65]. Firstly, I bases affect the mRNA’s structure and influence the binding affinity of proteins to the mRNA. Additionally, splicing and translation machineries recognize G instead of I, influencing both splicing patterns and translation accuracy [66]. Secondly, I34 of tRNA is located at the wobble position and has been related to the codon recognition mechanism [62,63]. Taken together, A-to-I editing constitutes a significant modification that mediates protein synthesis in different layers, giving birth to various proteins that, therefore, increase the proteome diversity.

2.2. Chemical Modifications of Cytosine RNA Base

Apart from adenine, a great number of post-transcriptional marks have also been found in cytosine, among which m5C, m3C methylations, C-to-U RNA editing and the synthesis of pseudouridine and dihydrouridine participate in various biological processes; thus, these are at the center of epitranscriptome research.

2.2.1. N5-Methylcytosine (m5C)

The most dominant methylation of C in RNAs is the one occurring in the fifth nitrogen atom of the C, namely m5C. This type of modification has been found in diverse types of RNAs. Two distinct methyltransferase groups have been confirmed to incorporate m5C in RNAs: the NOP2/Sun RNA methyltransferase (NSUN) family that includes seven proteins and the DNMT2 [67]. On the contrary, Ten-eleven translocation methylcytosine dioxygenase 1 (TET1) is a recently identified m5C eraser, whereas TET2 and TET3, MBD2/4, ALKBH1 and ALKBH6 are potential m5C demethylases; however, none of them have been confirmed [68,69,70,71,72]. Furthermore, the Aly/REF export factor (ALYREF), Y-box binding protein 1 (Ybx1) and the Radiation sensitive 52 (RAD52) have been identified as the m5C recognition proteins [73,74,75,76,77].
To begin with, m5C is present at multiple positions in human tRNAs, including C38, C48 and C72, which are modified by NSUN2, NSUN6 and DNMT2 (Figure 3) [67]. In particular, the incorporation of the methyl group at C38 in the anticodon loop is catalyzed by DNMT2. Studies have reported that DNMT2 prevents tRNA from degradation and hence m5C38 enhances translation. Methylation in C48 facilitates the binding of m5C38 with nucleoside at position 15, forming the “Levitt pair”. This bond leads to the generation of the characteristic L-structure that stabilizes the molecule. Additionally, C72 in the acceptor stem is critical for tRNA biogenesis and is methylated by NSUN6, whereas the 3′-CCA motif is required [78]. The m5C modifications on tRNAs ensure the increased stability of the molecules and their appropriate folding, and facilitate codon–anticodon interactions [79,80].
In human rRNAs, m5C methylations have only been detected in the 28S rRNA and incorporated by two NSUN proteins, NSUN1 and NSUN5. NSUN1 regulates transcription and targets the C4417 residue, whereas NSUN5 binds to the position 3761 of the 28S rRNA [81,82]. Both modified nucleotides support the translational fidelity and the proper folding of the 28S rRNA. Furthermore, NSUN4 introduces a methyl group into the mt-12S rRNA at position C841 [83]. m5C is also found in lncRNAs, where it facilitates its biogenesis and ensures its stability [84]. Several studies report NSUN7 as the potential writer of m5C in ncRNAs [78,85].
For more than 60 years, it has been known that m5C decorates mRNAs, but the exact locations of this mark were unknown until the advent of liquid chromatography–mass spectrometry and next-generation sequencing, which enabled the mapping of the m5C in single nucleotide resolution [77]. Relevant studies support that UTRs are rich in m5C, while in CDS regions, the m5C sites are depleted [73,86]. Until now, NSUN2 is the only confirmed m5C mRNA writer, whereas ALYREF serves as the reader protein. Remarkably, the m5C status regulates a plethora of cellular responses that affect mRNA fate and its exportation from the nucleus. The wide distribution of m5C within the mRNAs affects the translation efficiency in multiple ways. Firstly, the enriched m5C sites in 5′ UTRs modulate the protein translation, while during cell aging, erasers are activated to demethylate these sites. Additionally, the increased accumulation of m5C modifications in the 3′ UTRs demonstrates an increased translational capacity [78,87]. As for the internal mRNA regions that are subjected to m5C methylations, m5C can reduce translation efficiency by altering the codon–anticodon binding affinity [88]. Overall, m5C in mRNAs are associated with vital biological processes, including nuclear–cytoplasmic shuttling, maternal mRNA stabilization, splicing and the translation rate [89].

2.2.2. N3-Methylcytidine (m3C)

The m3C represents a tRNA modification, which is found in position 32 in different species [90,91]; however, in some human tRNAs, it is also located at the e2 position of the variable loop (Figure 3). Additional reports have mentioned the presence of m3C on mRNAs, but in much lower levels [92]. Although it was known that TRM140 methyltransferase inserts m3C in tRNAs of Saccharomyces cerevisiae, METTL2A, METTL2B, METTL6, and METTL8 have recently been confirmed as human m3C methyltransferases [93,94]. Notably, METTL2A, METTL2B and METTL6 act on tRNAs, whereas METTL8 methylates sites on mRNAs.
On the contrary, the human ALKBH3 erases m3C on tRNAs, but demethylation on mRNAs is achieved by ALKBH1 [95,96]. As the molecular role of m3C32 in tRNAs, it has been shown to interact with the nucleotide at position 38, which leads to the maintenance of the anti-codon loop structure and the increase in the decoding accuracy. Due to its low abundancy, the functionality of m3C in mRNAs remains unclear [97].

2.2.3. Cytidine to Uridine (C-to-U)

Except from A-to-I, another RNA editing mechanism that involves the conversion of bases is the formation of U by C. The Apolipoprotein B mRNA editing enzyme (APOBEC) converts the ribobase C to U in mRNA editing sites [98]. Although C-to-U editing has been suggested to be involved in mRNA stability and translation accuracy, there are no relative studies to prove that statement.

2.2.4. Pseudouridine (Ψ)

The Ψ modification is the well-studied derivative of U and is prevalent in all classes of RNAs. In humans, Ψ is mainly formed by members of the pseudouridine synthase (PUS) family, PUS1-PUS10, while its readers and erasers remain uncertain. Ψ writers can be divided into two separate groups: the guide RNA-dependent synthases that include the small ribonucleoproteins H/ACA sRNPs, and the guide-independent PUS enzymes. Briefly, the guide-dependent process is catalyzed by a two-step reaction that requires a complementary RNA to guide the enzyme in the target region and a protein that forms the modification [99]. On the contrary, the guide-independent pathway utilizes the members of the PUS family, which can directly recognize the target sites [100]. However, the RNA-binding protein PUM2 can recognize the UGUAR motif in human cells and, therefore, is a candidate for being a potential Ψ reader [101,102]. Undoubtedly, the high levels of Ψ in all the types of RNA reflect on its multidimensional implications in the RNAs’ life. Notably, the great thermodynamic stability of the Ψ-modified RNAs is based on the strength of the bond that is created between Ψ and A, whereas Ψ can also stabilize single-stranded RNAs. In tRNAs, Ψ affects a plethora of cellular responses, including tRNA biogenesis, degradation, and the production of tRFs. In particular, Ψ39 in the anticodon arm increases the melting temperature, controlling the tRNA folding process [103].
rRNAs are also subjected to pseudouridylation through Dyskerin Pseudouridine Synthase 1 (DKC1), an alternative enzyme that incorporates Ψ in 28S rRNAs [104]. Notably, although an increased number of Ψ sites had already been detected, seven additional Ψ residues have recently been identified in significant ribosomal regions: the first four at the positions 897, 1045, 1136, 1232 of the 18S rRNA and the rest at positions 1768, 2619 and 4463 of the 28S (Table 2 and Table 3). The 5.8S rRNA is also subjected to Ψ modifications at specific locations (Figure 4). The high density of Ψ in rRNAs underlines its importance in ribosome assembly and translational fidelity [105].
In mRNAs, Ψ is added co-transcriptionally by PUS1, PUS7, and RPUSD4 enzymes, and plays a critical role in the alternative pre-mRNA processing steps that affect gene expression [105,106,107]. Moreover, Ψ sites are distributed throughout mRNA sequences, being present in UTRs and CDS regions, and their regulatory role is to control mRNA metabolism [105]. The mapping of Ψ sites uncovered that human mRNAs are highly modified, ranging from 10–50% [108]. Briefly, Ψ can alter the primary protein sequence through the misincorporation of one or more amino acids, or even promote the termination codon readthrough during protein synthesis [109]. Furthermore, it should be mentioned that snRNAs harbor a plethora of Ψ sites crucial for their interactions with other RNAs and protein molecules. For instance, Ψ6, Ψ7 and Ψ15 in the U2 snRNA sequence are necessary for the assembly of the spliceosome machinery [110,111]. In the same manner, similar modifications in U4, U5 and U6 snRNAs enhance the mechanism of splicing [112].

2.2.5. Dihydrouridine (D)

Besides Ψ, C can be edited into U to synthesize D, a highly conserved alteration found in great abundancy at specific positions in the D-loop of tRNAs [113]. Dihydrouridine synthase (DUS) enzymes are four protein molecules that form D in human tRNAs, with DUS2 being the most dominant one. The incorporation of D into tRNAs has been linked to tRNA folding and the increased flexibility of tRNAs, and it is possible to destabilize their 3D structures [114,115]. Although D was determined to be a tRNA-specific modification, recent sequencing-based studies have supported the existence of D in coding RNAs and lncRNAs; however, its biological role is still ambiguous [116].

2.3. Chemical Modifications of Guanine RNA Base

Even though modern epitranscriptomics is mainly focused on modifications occurring in adenine and cytosine, several guanine-based alterations are vital for the fate of RNAs and are thus worth mentioning.

2.3.1. N7-Methylguanosine (m7G)

Although the m7G mark is widely known due to its presence in tRNAs originating from multiple organisms, it is also associated with the eukaryotic 5′ capping of the mRNA [117,118]. Different methyltransferase complexes act in the RNAs to incorporate this type of modification. Specifically, in humans, the METTL1/WDR4 complex participates in the formation of m7G in tRNAs, whereas WBSCR22 and TRMT112 proteins act synergistically to add m7G on the 18S rRNA [119,120,121]. METTL1 can also methylate miRNAs, such as let-7 miRNA, a critical procedure for their biogenesis [122]. On the other hand, RNA guanine-7 methyltransferase (RNMT) is responsible for the m7G addition to the mRNA cap [47].
As for the physiological role of m7G in tRNAs, the formation of m7G46 promotes the mRNA translation and increases the tRNA stability [117]. It should be noted that, although the catalytic activity of WBSCR22 recruits the m7G1639 in the human 18S rRNA, this modification is not necessary for the biogenesis of the 40S ribosomal subunit [120]. Moreover, the special features of m7G in the 5′ cap of the mRNA enable its involvement in vital biological pathways, including RNA maturation, nuclear export and cap-dependent translation [118].

2.3.2. N1-Methylguanosine (m1G)

In human RNAs, the addition of a methyl group into the first atom of the ribose ring of G is mediated by various writers that act independently. More precisely, tRNA methyltransferase 5 (TRMT5) is the enzyme that catalyzes the incorporation of m1G at tRNA position 37; meanwhile, in case of mitochondrial tRNAs, TRMT5 demonstrates a notably lower activity and is replaced by TRMT10C, an additional tRNA methyltransferase that introduces the m1G9 mt-tRNA modification [123,124]. Furthermore, the m1G9 methyltransferase TRMT10C can form multiprotein complexes with RG9MTD1 and SDR5C1, in order to catalyze the process of G methylation [125,126]. According to several studies, the formation of m1G in tRNAs is vital for proper tRNA folding and its tertiary structure, and can prevent frameshifting during protein production [127]. Moreover, as far as mRNA is concerned, although m1G writers have not been reported, the incorporation of a single m1G mark into the mRNA sequence destabilizes the translation machinery, leading to reduced levels of the generated protein [128].

2.3.3. Queuosine (Q)

Q represents a tRNA-specific hyper-modified guanosine nucleoside derived from G, which forms a plethora of derivatives, including galactosyl-queuosine (GalQ), mannosyl-queuosine (ManQ), and glutamyl-queuosine (GluQ) [129,130]. It is worth mentioning that Q34 on tRNAs protects tRNAs from ribonuclease degradation and affects the translation accuracy [131,132].

2.4. Chemical Modifications of Uracil RNA Base

Although chemical modifications mainly occur in adenine, cytosine and guanine, few studies have reported the existence of post-transcriptional marks in U.

2.4.1. N5-Methyluridine (m5U)

In RNAs, the methylation of the fifth nitrogen atom of the U ring creates the 5-methyluridine (m5U), which constitutes a common modification both in tRNAs and rRNAs. However, high-throughput sequencing studies have also revealed the existence of m5U in mRNAs. Although TRMT2A and TRMT2B are the m5U-catalyzing protein enzymes that add the methyl group into human tRNAs, rRNAs and mRNAs, neither erasers nor readers have been identified yet [133,134]. Of note, m5U regions have been extensively studied in human tRNAs and are found at position m5U54 in the T-loop, but also in the mitochondrial tRNAs [134]. Its function is to maintain and stabilize the tertiary structure of tRNAs; as a result the absence of the m5U54 mark can cause tRNA’s degradation and the generation of tRFs [135].

2.4.2. N3-Methyluridine (m3U)

m3U is a major rRNA modification in multiple species, including humans, that is detected in human 28S rRNA at position 4500 (Table 3). Until now, although recent reports support that Beta-mannosyltransferases, Bmt5 and Bmt6, are responsible for the addition of methyl groups into the rRNA of Saccharomyces cerevisiae, the human m3U writer remains unknown [136]. Hence, its functional role in the rRNAs is unclear.

2.4.3. Uridylation

In higher eukaryotes, the post-transcriptional addition of nucleotides in the 3′ UTRs is a major procedure for the stabilization of the newly synthesized RNA molecules [137]. The 3′ uridylation constitutes a widespread mechanism that is catalyzed by the terminal uridyltransferases TUT4 and TUT7 on different types of RNAs, in order to mark the molecules for degradation [138,139]. Notably, the TUT4/TUT7 complexes target mRNAs and miRNAs, and control both stability and RNA homeostasis by fine-tuning RNA levels during apoptosis [140]. An additional member of the TUT family, TUT1, has also been reported to catalyze uridylation, while DIS3 Like 3’-5’ Exoribonuclease 2 (DIS3L2) recognizes the uridylated sites [141].

2.5. 2′-O-Methylation (Nm) Modification

The Nm modification occurs post-transcriptionally via the incorporation of a methyl substituent into the 2′-hydroxyl of the ribose in any base. Consequently, the Am, Cm, Gm and Um modifications are generated. The Nm is widely distributed in all RNA types; however, tRNAs and rRNAs are particularly enriched in this type of alteration [142,143]. It has been shown that Nm affects the RNA structure by increasing the thermodynamic stability of the molecule in order to protect it from ribonucleases, and to enhance the RNA:RNA base pairing and the formation of RNA duplexes [144,145,146]. On the contrary, the tertiary RNA structures are disrupted, and the RNA–protein interactions are inhibited.
In tRNAs, Nm marks are mainly deposited by FtsJ RNA 2’-O-Methyltransferase 1 (FTSJ1), which recognizes the C32 and N34 regions in the anticodon loop, and the creation of Cm32 and Nm34 influences the translation [147]. Additionally, TRMT44 constitutes a potential Um writer in tRNASer [148]. In rRNAs, in the same manner as Ψ, Nm modifications are formed in pre-rRNAs and their role is to prevent hydrolysis and thus increase the structural rigidity of rRNAs [149]. Nm methylations are induced by snoRNAs that activate and guide Fibrillarin (FBL), the 2′-O-RNA methyltransferase, to the target rRNA, contributing to the fine-tuning of its function [143,150].
In mRNAs, Nm is present in the UTRs and more precisely in the 5′ mRNA termini, whereas in CDS regions, the AGUA motif has been found to harbor this type of modification [151]. The 5′ mRNA termini of eukaryotic organisms are heavily methylated and can form three types of caps: cap-0, cap-1 and cap-2. Nm is involved in the 5′ cap-1 (m7GpppNm) that is produced by CMTR1 and the 5′ cap-2 that includes the highly conserved m7GpppNmNm region, generated by CMTR2 [152]. The Nm modifications are highly involved in transcription processing, the mRNA stability and the protein synthesis efficiency [152].

3. RNA Modifications in Human Disease

Over the years, systematic efforts to catalog the repertoire of modifications that are embroidered on RNA molecules in different tissues and diverse pathophysiological conditions has illuminated the correlation between epitranscriptome deregulation and disease development (Table 4).
More specifically, mutations, as well as irregular expression patterns in numerous RNA modification enzymes, have been linked with defects in the epitranscriptome and, subsequently, with several human diseases, including cancer and neurological and cardiovascular disorders; thus, research has been propelled towards the elucidation of the enigmatic molecular mechanisms driving these pathological conditions [162].
First and foremost, a wide array of RNA modifiers have been associated with guiding normal cells towards the acquisition of traits distinctive of cancer cells, widely known as the “hallmarks of cancer” [171]. For instance, the m6A writer METTL3 has been found to be overexpressed in acute myeloid leukemia (AML) cell lines; it is considered to be accountable for the increased m6A methylation profiles and the translational activation of the MYC proto-oncogene, phosphatase and tensin homolog (PTEN), and the BCL2 apoptosis regulator mRNA transcripts, thus sustaining cell survival and proliferation [172]. On the other hand, in AML, the upregulated m6A eraser FTO contributes to the bypassing of growth suppressors by demethylating; this reduces the stability and negatively regulates the retinoic acid receptor alpha (RARA), the ankyrin repeat and SOCS box-containing 2 (ASB2) mRNAs, hampering cell differentiation and promoting leukemogenesis [173].
Notably, recent studies have also underlined the involvement of the m6A readers YTHDF1 and YTHDF2 in hindering the efficiency of immune responses against tumor antigens, and assisting leukemia stem cells to evade apoptosis, respectively. Specifically, YTHDF1 recognizes the m6A signature and enhances the translation of the lysosomal proteases’ mRNAs found in dendritic cells, which, in turn, degrade the engulfed antigens, inhibiting their cross-presentation and suppressing the induction of CD8+ T cell responses [174]. On the contrary, overexpressed YTHDF2 in AML cells spots m6A-methylated target transcripts, such as the tumor necrosis factor (TNF) receptor superfamily member 1B (TNFRSF1B) mRNAs, and promotes their degradation, affecting the TNF apoptotic signaling pathway [175].
Despite being the most frequently encountered RNA modification, m6A is not the only epitranscriptomic mark that can promote malignancy. In the cytoplasm, Lin-28 homolog A (LIN28A) normally regulates the expression of let-7 miRNAs, the recruitment of writer TUT4. In turn, TUT4 adds uridines at the 3′ end of the precursor miRNAs (pre-let7), thus inhibiting their cleavage by Dicer and promoting their degradation [176]. In breast cancer, however, an overexpression of LIN28A has been detected to lead to the deregulation of the aforementioned pathway and consequently the downregulation of the tumor-suppressing let-7 miRNAs [177]. Moreover, in chronic myeloid leukemia (CML), the activity of A-to-I writer ADAR1 has been proven to hinder let-7 miRNAs’ biogenesis, enhancing the replicative capacity of leukemia stem cells and promoting cancer progression [178]; meanwhile, in colon and lung cancer, the downregulation of the m7G writer METTL1 interferes with the methylation and maturation of the same miRNA family, resulting in the enhanced migration of cancer cells [122].
As far as the Ψ writer DKC1 is concerned, mutations in this gene have been linked with the reduced pseudouridylation of rRNA molecules, and subsequently with the modified translation of several cancer-associated transcripts, such as the vascular endothelial growth factor (VEGF) and the tumor suppressor p53 mRNAs [179,180,181]. On the other hand, the overexpression of NSUN2 in bladder carcinoma, and thus the m5C aberrant methylation of the oncogenic transcripts of the heparin binding growth factor (HDGF) gene, have been shown to augment mRNA stability and correspond with poor cancer prognosis [95]. Finally, the elevated expression of the eraser ALKBH3 and, therefore, the increased m1A demethylation of tRNA molecules, participate in the production of tRNA-derived small RNAs (tDRs) and contribute to an enhanced cancer cell proliferation [95].
Given its critical role in governing brain development and functionality, it is not surprising that perturbations in the m6A signature could also be implicated in numerous neurological diseases [182]. So far, several alterations in the m6A machinery have been detected, underlining the potentially critical role of defective methylation in the establishment and progression of these diseases [183]. Specifically, the FTO gene has already been correlated with neurodevelopmental and neuropsychiatric disorders, such as structural malformations and functional deficiencies of the brain, growth retardation, psychomotor retardation, attention-deficit/hyperactivity disorder (ADHD) and major depressive disorder (MDD) [184,185,186]; meanwhile, an increasing number of studies also highlight the link between this specific eraser and neurodegenerative disorders, such as Alzheimer’s disease (AD) [187,188,189], Parkinson’s disease (PD) [190,191] and amyotrophic lateral sclerosis (ALS) (Table 4) [192]. Furthermore, recent studies have reported the elevated expression levels of METTL3 in mouse models for AD [189], whereas mutations in the m6A eraser ALKBH5 and the m6A reader YTHDC2 have been shown to participate in MDD and autism spectrum disorder, respectively [193,194].
As for the remaining epitranscriptomic marks, alterations in the NSUN2 gene have been identified as a cause of developing autosomal-recessive intellectual disability (ID) [195] and Dubowitz syndrome (DS) [196], while mutations in PUS7 have been proven to result in the reduced pseudouridylation of specific sites in tRNA molecules, leading to microcephaly and ID [197]. Finally, increased Ψ marks seem to be linked with early-stage AD and myotonic dystrophy [198,199], whereas the reduced editing of AMPA and kainate glutamate receptors, due to the downregulated expression of ADAR2, has been correlated with the etiology of mental disorders such as bipolar disorder and schizophrenia [200].
Last but not least, RNA modifications are the key regulators in multiple cardiovascular diseases, including adipogenesis, obesity, type 2 diabetes, different types of atherosclerosis and limb ischemia (Table 4). In particular, the levels of m6A are responsible for the glucose metabolism by regulating the β-cells in the pancreas and urging liver gluconeogenesis, thus affecting the progression of type 2 diabetes. Both METTL3 and METTL14 are downregulated in patients with type 2 diabetes, whereas FTO demonstrates an increased expression pattern, which enhances the expression of forkhead box O1 (FoxO1), glucose-6-phosphatase (G6PC), and diacylglycerol acyltransferase 2 (DGAT2) enzymes; this finally leads to hyperglycemia, due to insulin secretion dysregulation. According to additional studies, an m6A-dependent pathway, in which FTO and METTL3 possess significant roles, controls lipid production and switches on/off both adipogenesis and obesity.
In the case of atherosclerosis, the expression levels of m6A writers, readers and erasers can also promote and/or inhibit chronic inflammation and lipid deposition, which form the atherosclerotic plaques. For instance, the overexpression of METTL14 in endothelial cells is responsible for monocyte aggregation, via affecting either the m6A levels of FoxO1 mRNA or the binding affinity of the FoxO1 protein, resulting in the progression of atherosclerosis. The dysfunction of endothelial cells is also induced by the m5C RNA modification levels. Notably, NSUN2 can promote the translation of ICAM-1, which increases the adhesion of leukocytes to the surface of endothelial cells. On the contrary, limb ischemia is connected to other RNA modifications, such as A-to-I, Ψ and m7G. Studies support that A-to-I editing and Nm can alter the targets of the tumor suppressor microRNA, miR487b, and hence promote angiogenesis; meanwhile, the increased levels of the Ψ writers, RPUSD3 and RPUSD4, induce mitochondrial protein synthesis, leading to the defective assembly of OXPHOS [201]. Recent studies have pointed out that m7G modification in tRNAs is involved in vascular development due to the significant role of METTL1 in neovascularization [202]. Notably, m7G and the downregulation of METTL1 affect the pluripotency of the human-induced pluripotent stem cells, enable their differentiation to endothelial progenitor cells, and thus promote post-ischemic angiogenesis [203].

4. Conclusions

In summary, the rise of epitranscriptomics has led to the identification of a wide spectrum of RNA modifications that exist in many RNA classes. Each modification is located at specific regions in RNAs and possesses specific roles that are often vital to cellular responses. A variety of enzymes are implicated in the mechanisms that incorporate or erase each modification, thus underlining the complexity of the eukaryotic transcriptomes. The dysfunction of the catalytic enzymes that are responsible for the post-transcriptional marks into the RNAs can lead to human diseases, such as cancer. Modern epitranscriptomics aims to decipher the molecular mechanisms that generate the RNA modifications and decode the involvement of RNA marks in cellular homeostasis. Undoubtedly, breakthroughs in molecular biology, including RNA-sequencing techniques, will enhance our efforts to unveil the mysteries of the features of RNA molecules.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24032387/s1.

Author Contributions

P.G.A. and K.A. drafted the text of the article; K.A. designed all figures and tables; P.G.A. and G.N.D. provided additional text and important discussions for the enhancement of the text; P.G.A. and A.S. critically reviewed the final version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data was created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Costa Dos Santos, G., Jr.; Renovato-Martins, M.; de Brito, N.M. The remodel of the “central dogma”: A metabolomics interaction perspective. Metabolomics 2021, 17, 48. [Google Scholar] [CrossRef] [PubMed]
  2. Garcia, H.G.; Berrocal, A.; Kim, Y.J.; Martini, G.; Zhao, J. Lighting up the central dogma for predictive developmental biology. Curr. Top. Dev. Biol. 2020, 137, 1–35. [Google Scholar] [CrossRef] [PubMed]
  3. Zhao, L.Y.; Song, J.; Liu, Y.; Song, C.X.; Yi, C. Mapping the epigenetic modifications of DNA and RNA. Protein Cell 2020, 11, 792–808. [Google Scholar] [CrossRef] [PubMed]
  4. Koks, G.; Pfaff, A.L.; Bubb, V.J.; Quinn, J.P.; Koks, S. At the dawn of the transcriptomic medicine. Exp. Biol. Med. 2021, 246, 286–292. [Google Scholar] [CrossRef] [PubMed]
  5. Lian, H.; Wang, Q.H.; Zhu, C.B.; Ma, J.; Jin, W.L. Deciphering the Epitranscriptome in Cancer. Trends Cancer 2018, 4, 207–221. [Google Scholar] [CrossRef] [PubMed]
  6. Wiener, D.; Schwartz, S. The epitranscriptome beyond m(6)A. Nat. Rev. Genet. 2021, 22, 119–131. [Google Scholar] [CrossRef]
  7. Wilkinson, E.; Cui, Y.H.; He, Y.Y. Roles of RNA Modifications in Diverse Cellular Functions. Front. Cell Dev. Biol. 2022, 10, 828683. [Google Scholar] [CrossRef] [PubMed]
  8. Shi, H.; Chai, P.; Jia, R.; Fan, X. Novel insight into the regulatory roles of diverse RNA modifications: Re-defining the bridge between transcription and translation. Mol. Cancer 2020, 19, 78. [Google Scholar] [CrossRef]
  9. Esteve-Puig, R.; Bueno-Costa, A.; Esteller, M. Writers, readers and erasers of RNA modifications in cancer. Cancer Lett. 2020, 474, 127–137. [Google Scholar] [CrossRef] [PubMed]
  10. Shi, H.; Wei, J.; He, C. Where, When, and How: Context-Dependent Functions of RNA Methylation Writers, Readers, and Erasers. Mol. Cell 2019, 74, 640–650. [Google Scholar] [CrossRef]
  11. Zaccara, S.; Ries, R.J.; Jaffrey, S.R. Reading, writing and erasing mRNA methylation. Nat. Rev. Mol. Cell Biol. 2019, 20, 608–624. [Google Scholar] [CrossRef] [PubMed]
  12. Chawla, M.; Oliva, R.; Bujnicki, J.M.; Cavallo, L. An atlas of RNA base pairs involving modified nucleobases with optimal geometries and accurate energies. Nucleic Acids Res. 2015, 43, 6714–6729. [Google Scholar] [CrossRef] [Green Version]
  13. He, L.; Li, H.; Wu, A.; Peng, Y.; Shu, G.; Yin, G. Functions of N6-methyladenosine and its role in cancer. Mol. Cancer 2019, 18, 176. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Perry, R.P.; Kelley, D.E.; Friderici, K.; Rottman, F. The methylated constituents of L cell messenger RNA: Evidence for an unusual cluster at the 5′ terminus. Cell 1975, 4, 387–394. [Google Scholar] [CrossRef] [PubMed]
  15. Wei, C.M.; Gershowitz, A.; Moss, B. Methylated nucleotides block 5′ terminus of HeLa cell messenger RNA. Cell 1975, 4, 379–386. [Google Scholar] [CrossRef]
  16. Meyer, K.D.; Saletore, Y.; Zumbo, P.; Elemento, O.; Mason, C.E.; Jaffrey, S.R. Comprehensive analysis of mRNA methylation reveals enrichment in 3′ UTRs and near stop codons. Cell 2012, 149, 1635–1646. [Google Scholar] [CrossRef] [Green Version]
  17. Roundtree, I.A.; Evans, M.E.; Pan, T.; He, C. Dynamic RNA Modifications in Gene Expression Regulation. Cell 2017, 169, 1187–1200. [Google Scholar] [CrossRef] [Green Version]
  18. Liu, S.; Zhuo, L.; Wang, J.; Zhang, Q.; Li, Q.; Li, G.; Yan, L.; Jin, T.; Pan, T.; Sui, X.; et al. METTL3 plays multiple functions in biological processes. Am. J. Cancer Res. 2020, 10, 1631–1646. [Google Scholar] [CrossRef]
  19. Zhou, H.; Yin, K.; Zhang, Y.; Tian, J.; Wang, S. The RNA m6A writer METTL14 in cancers: Roles, structures, and applications. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188609. [Google Scholar] [CrossRef]
  20. Wang, T.; Kong, S.; Tao, M.; Ju, S. The potential role of RNA N6-methyladenosine in Cancer progression. Mol. Cancer 2020, 19, 88. [Google Scholar] [CrossRef]
  21. Jiang, X.; Liu, B.; Nie, Z.; Duan, L.; Xiong, Q.; Jin, Z.; Yang, C.; Chen, Y. The role of m6A modification in the biological functions and diseases. Signal Transduct. Target Ther. 2021, 6, 74. [Google Scholar] [CrossRef] [PubMed]
  22. Zhang, N.; Zuo, Y.; Peng, Y.; Zuo, L. Function of N6-Methyladenosine Modification in Tumors. J. Oncol. 2021, 2021, 6461552. [Google Scholar] [CrossRef] [PubMed]
  23. Horiuchi, K.; Kawamura, T.; Iwanari, H.; Ohashi, R.; Naito, M.; Kodama, T.; Hamakubo, T. Identification of Wilms’ tumor 1-associating protein complex and its role in alternative splicing and the cell cycle. J. Biol. Chem. 2013, 288, 33292–33302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Yue, Y.; Liu, J.; Cui, X.; Cao, J.; Luo, G.; Zhang, Z.; Cheng, T.; Gao, M.; Shu, X.; Ma, H.; et al. VIRMA mediates preferential m(6)A mRNA methylation in 3′UTR and near stop codon and associates with alternative polyadenylation. Cell Discov. 2018, 4, 10. [Google Scholar] [CrossRef] [Green Version]
  25. Ruszkowska, A. METTL16, Methyltransferase-Like Protein 16: Current Insights into Structure and Function. Int. J. Mol. Sci. 2021, 22, 2176. [Google Scholar] [CrossRef]
  26. Satterwhite, E.R.; Mansfield, K.D. RNA methyltransferase METTL16: Targets and function. Wiley Interdiscip. Rev. RNA 2022, 13, e1681. [Google Scholar] [CrossRef]
  27. Han, X.; Guo, J.; Fan, Z. Interactions between m6A modification and miRNAs in malignant tumors. Cell Death Dis. 2021, 12, 598. [Google Scholar] [CrossRef]
  28. Han, J.; Wang, J.Z.; Yang, X.; Yu, H.; Zhou, R.; Lu, H.C.; Yuan, W.B.; Lu, J.C.; Zhou, Z.J.; Lu, Q.; et al. METTL3 promote tumor proliferation of bladder cancer by accelerating pri-miR221/222 maturation in m6A-dependent manner. Mol. Cancer 2019, 18, 110. [Google Scholar] [CrossRef] [Green Version]
  29. Yang, Y.; Fan, X.; Mao, M.; Song, X.; Wu, P.; Zhang, Y.; Jin, Y.; Yang, Y.; Chen, L.L.; Wang, Y.; et al. Extensive translation of circular RNAs driven by N(6)-methyladenosine. Cell Res. 2017, 27, 626–641. [Google Scholar] [CrossRef] [Green Version]
  30. van Tran, N.; Ernst, F.G.M.; Hawley, B.R.; Zorbas, C.; Ulryck, N.; Hackert, P.; Bohnsack, K.E.; Bohnsack, M.T.; Jaffrey, S.R.; Graille, M.; et al. The human 18S rRNA m6A methyltransferase METTL5 is stabilized by TRMT112. Nucleic Acids Res. 2019, 47, 7719–7733. [Google Scholar] [CrossRef]
  31. Ren, W.; Lu, J.; Huang, M.; Gao, L.; Li, D.; Wang, G.G.; Song, J. Structure and regulation of ZCCHC4 in m(6)A-methylation of 28S rRNA. Nat. Commun. 2019, 10, 5042. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Pinto, R.; Vagbo, C.B.; Jakobsson, M.E.; Kim, Y.; Baltissen, M.P.; O’Donohue, M.F.; Guzman, U.H.; Malecki, J.M.; Wu, J.; Kirpekar, F.; et al. The human methyltransferase ZCCHC4 catalyses N6-methyladenosine modification of 28S ribosomal RNA. Nucleic Acids Res. 2020, 48, 830–846. [Google Scholar] [CrossRef] [PubMed]
  33. Azzam, S.K.; Alsafar, H.; Sajini, A.A. FTO m6A Demethylase in Obesity and Cancer: Implications and Underlying Molecular Mechanisms. Int. J. Mol. Sci. 2022, 23, 3800. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, J.; Wang, J.; Gu, Q.; Ma, Y.; Yang, Y.; Zhu, J.; Zhang, Q. The biological function of m6A demethylase ALKBH5 and its role in human disease. Cancer Cell Int. 2020, 20, 347. [Google Scholar] [CrossRef]
  35. Meyer, K.D.; Jaffrey, S.R. Rethinking m(6)A Readers, Writers, and Erasers. Annu. Rev. Cell Dev. Biol. 2017, 33, 319–342. [Google Scholar] [CrossRef] [Green Version]
  36. Maden, B.E. Identification of the locations of the methyl groups in 18 S ribosomal RNA from Xenopus laevis and man. J. Mol. Biol. 1986, 189, 681–699. [Google Scholar] [CrossRef]
  37. Maden, B.E. Locations of methyl groups in 28 S rRNA of Xenopus laevis and man. Clustering in the conserved core of molecule. J. Mol. Biol. 1988, 201, 289–314. [Google Scholar] [CrossRef]
  38. Genuth, N.R.; Barna, M. The Discovery of Ribosome Heterogeneity and Its Implications for Gene Regulation and Organismal Life. Mol. Cell 2018, 71, 364–374. [Google Scholar] [CrossRef] [Green Version]
  39. Yang, Y.; Hsu, P.J.; Chen, Y.S.; Yang, Y.G. Dynamic transcriptomic m(6)A decoration: Writers, erasers, readers and functions in RNA metabolism. Cell Res. 2018, 28, 616–624. [Google Scholar] [CrossRef] [Green Version]
  40. Roundtree, I.A.; Luo, G.Z.; Zhang, Z.; Wang, X.; Zhou, T.; Cui, Y.; Sha, J.; Huang, X.; Guerrero, L.; Xie, P.; et al. YTHDC1 mediates nuclear export of N(6)-methyladenosine methylated mRNAs. Elife 2017, 6, e31311. [Google Scholar] [CrossRef]
  41. Lin, X.; Chai, G.; Wu, Y.; Li, J.; Chen, F.; Liu, J.; Luo, G.; Tauler, J.; Du, J.; Lin, S.; et al. RNA m(6)A methylation regulates the epithelial mesenchymal transition of cancer cells and translation of Snail. Nat. Commun. 2019, 10, 2065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Wang, S.; Lv, W.; Li, T.; Zhang, S.; Wang, H.; Li, X.; Wang, L.; Ma, D.; Zang, Y.; Shen, J.; et al. Dynamic regulation and functions of mRNA m6A modification. Cancer Cell Int. 2022, 22, 48. [Google Scholar] [CrossRef]
  43. Li, A.; Chen, Y.S.; Ping, X.L.; Yang, X.; Xiao, W.; Yang, Y.; Sun, H.Y.; Zhu, Q.; Baidya, P.; Wang, X.; et al. Cytoplasmic m(6)A reader YTHDF3 promotes mRNA translation. Cell Res. 2017, 27, 444–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Wang, J.Y.; Lu, A.Q. The biological function of m6A reader YTHDF2 and its role in human disease. Cancer Cell Int. 2021, 21, 109. [Google Scholar] [CrossRef] [PubMed]
  45. Oerum, S.; Meynier, V.; Catala, M.; Tisne, C. A comprehensive review of m6A/m6Am RNA methyltransferase structures. Nucleic Acids Res. 2021, 49, 7239–7255. [Google Scholar] [CrossRef] [PubMed]
  46. Lee, Y.; Choe, J.; Park, O.H.; Kim, Y.K. Molecular Mechanisms Driving mRNA Degradation by m(6)A Modification. Trends Genet. 2020, 36, 177–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Galloway, A.; Kaskar, A.; Ditsova, D.; Atrih, A.; Yoshikawa, H.; Gomez-Moreira, C.; Suska, O.; Warminski, M.; Grzela, R.; Lamond, A.I.; et al. Upregulation of RNA cap methyltransferase RNMT drives ribosome biogenesis during T cell activation. Nucleic Acids Res. 2021, 49, 6722–6738. [Google Scholar] [CrossRef] [PubMed]
  48. Dominissini, D.; Nachtergaele, S.; Moshitch-Moshkovitz, S.; Peer, E.; Kol, N.; Ben-Haim, M.S.; Dai, Q.; Di Segni, A.; Salmon-Divon, M.; Clark, W.C.; et al. The dynamic N(1)-methyladenosine methylome in eukaryotic messenger RNA. Nature 2016, 530, 441–446. [Google Scholar] [CrossRef] [Green Version]
  49. Li, X.; Xiong, X.; Wang, K.; Wang, L.; Shu, X.; Ma, S.; Yi, C. Transcriptome-wide mapping reveals reversible and dynamic N(1)-methyladenosine methylome. Nat. Chem. Biol. 2016, 12, 311–316. [Google Scholar] [CrossRef]
  50. Li, J.; Zhang, H.; Wang, H. N(1)-methyladenosine modification in cancer biology: Current status and future perspectives. Comput. Struct. Biotechnol. J. 2022, 20, 6578–6585. [Google Scholar] [CrossRef]
  51. Waku, T.; Nakajima, Y.; Yokoyama, W.; Nomura, N.; Kako, K.; Kobayashi, A.; Shimizu, T.; Fukamizu, A. NML-mediated rRNA base methylation links ribosomal subunit formation to cell proliferation in a p53-dependent manner. J. Cell Sci. 2016, 129, 2382–2393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Boo, S.H.; Ha, H.; Kim, Y.K. m(1)A and m(6)A modifications function cooperatively to facilitate rapid mRNA degradation. Cell Rep. 2022, 40, 111317. [Google Scholar] [CrossRef] [PubMed]
  53. Oerum, S.; Degut, C.; Barraud, P.; Tisne, C. m1A Post-Transcriptional Modification in tRNAs. Biomolecules 2017, 7, 20. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Liu, B.; Cao, J.; Wang, X.; Guo, C.; Liu, Y.; Wang, T. Deciphering the tRNA-derived small RNAs: Origin, development, and future. Cell Death Dis. 2021, 13, 24. [Google Scholar] [CrossRef] [PubMed]
  55. Sharma, S.; Hartmann, J.D.; Watzinger, P.; Klepper, A.; Peifer, C.; Kotter, P.; Lafontaine, D.L.J.; Entian, K.D. A single N(1)-methyladenosine on the large ribosomal subunit rRNA impacts locally its structure and the translation of key metabolic enzymes. Sci. Rep. 2018, 8, 11904. [Google Scholar] [CrossRef]
  56. Li, X.; Xiong, X.; Zhang, M.; Wang, K.; Chen, Y.; Zhou, J.; Mao, Y.; Lv, J.; Yi, D.; Chen, X.W.; et al. Base-Resolution Mapping Reveals Distinct m(1)A Methylome in Nuclear- and Mitochondrial-Encoded Transcripts. Mol. Cell 2017, 68, 993–1005.e1009. [Google Scholar] [CrossRef] [Green Version]
  57. Seo, K.W.; Kleiner, R.E. YTHDF2 Recognition of N(1)-Methyladenosine (m(1)A)-Modified RNA is Associated with Transcript Destabilization. ACS Chem. Biol. 2020, 15, 132–139. [Google Scholar] [CrossRef] [PubMed]
  58. Licht, K.; Hartl, M.; Amman, F.; Anrather, D.; Janisiw, M.P.; Jantsch, M.F. Inosine induces context-dependent recoding and translational stalling. Nucleic Acids Res. 2019, 47, 3–14. [Google Scholar] [CrossRef] [Green Version]
  59. Tajaddod, M.; Jantsch, M.F.; Licht, K. The dynamic epitranscriptome: A to I editing modulates genetic information. Chromosoma 2016, 125, 51–63. [Google Scholar] [CrossRef] [Green Version]
  60. Raghava Kurup, R.; Oakes, E.K.; Manning, A.C.; Mukherjee, P.; Vadlamani, P.; Hundley, H.A. RNA binding by ADAR3 inhibits adenosine-to-inosine editing and promotes expression of immune response protein MAVS. J. Biol. Chem. 2022, 298, 102267. [Google Scholar] [CrossRef]
  61. Tan, M.H.; Li, Q.; Shanmugam, R.; Piskol, R.; Kohler, J.; Young, A.N.; Liu, K.I.; Zhang, R.; Ramaswami, G.; Ariyoshi, K.; et al. Dynamic landscape and regulation of RNA editing in mammals. Nature 2017, 550, 249–254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Ramos, J.; Han, L.; Li, Y.; Hagelskamp, F.; Kellner, S.M.; Alkuraya, F.S.; Phizicky, E.M.; Fu, D. Formation of tRNA Wobble Inosine in Humans is Disrupted by a Millennia-Old Mutation Causing Intellectual Disability. Mol. Cell Biol. 2019, 39, e00203-19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Torres, A.G.; Rodriguez-Escriba, M.; Marcet-Houben, M.; Santos Vieira, H.G.; Camacho, N.; Catena, H.; Murillo Recio, M.; Rafels-Ybern, A.; Reina, O.; Torres, F.M.; et al. Human tRNAs with inosine 34 are essential to efficiently translate eukarya-specific low-complexity proteins. Nucleic Acids Res. 2021, 49, 7011–7034. [Google Scholar] [CrossRef] [PubMed]
  64. Bazak, L.; Haviv, A.; Barak, M.; Jacob-Hirsch, J.; Deng, P.; Zhang, R.; Isaacs, F.J.; Rechavi, G.; Li, J.B.; Eisenberg, E.; et al. A-to-I RNA editing occurs at over a hundred million genomic sites, located in a majority of human genes. Genome Res. 2014, 24, 365–376. [Google Scholar] [CrossRef] [Green Version]
  65. Picardi, E.; Manzari, C.; Mastropasqua, F.; Aiello, I.; D’Erchia, A.M.; Pesole, G. Profiling RNA editing in human tissues: Towards the inosinome Atlas. Sci. Rep. 2015, 5, 14941. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Song, B.; Shiromoto, Y.; Minakuchi, M.; Nishikura, K. The role of RNA editing enzyme ADAR1 in human disease. Wiley Interdiscip. Rev. RNA 2022, 13, e1665. [Google Scholar] [CrossRef]
  67. Bohnsack, K.E.; Hobartner, C.; Bohnsack, M.T. Eukaryotic 5-methylcytosine (m(5)C) RNA Methyltransferases: Mechanisms, Cellular Functions, and Links to Disease. Genes 2019, 10, 102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Yang, H.; Wang, Y.; Xiang, Y.; Yadav, T.; Ouyang, J.; Phoon, L.; Zhu, X.; Shi, Y.; Zou, L.; Lan, L. FMRP promotes transcription-coupled homologous recombination via facilitating TET1-mediated m5C RNA modification demethylation. Proc. Natl. Acad. Sci. USA 2022, 119, e2116251119. [Google Scholar] [CrossRef] [PubMed]
  69. Shen, H.; Ontiveros, R.J.; Owens, M.C.; Liu, M.Y.; Ghanty, U.; Kohli, R.M.; Liu, K.F. TET-mediated 5-methylcytosine oxidation in tRNA promotes translation. J. Biol. Chem. 2021, 296, 100087. [Google Scholar] [CrossRef]
  70. Yang, J.; Bashkenova, N.; Zang, R.; Huang, X.; Wang, J. The roles of TET family proteins in development and stem cells. Development 2020, 147, dev183129. [Google Scholar] [CrossRef]
  71. Kawarada, L.; Suzuki, T.; Ohira, T.; Hirata, S.; Miyauchi, K.; Suzuki, T. ALKBH1 is an RNA dioxygenase responsible for cytoplasmic and mitochondrial tRNA modifications. Nucleic Acids Res. 2017, 45, 7401–7415. [Google Scholar] [CrossRef] [Green Version]
  72. Ma, L.; Lu, H.; Tian, Z.; Yang, M.; Ma, J.; Shang, G.; Liu, Y.; Xie, M.; Wang, G.; Wu, W.; et al. Structural insights into the interactions and epigenetic functions of human nucleic acid repair protein ALKBH6. J. Biol. Chem. 2022, 298, 101671. [Google Scholar] [CrossRef]
  73. Yang, X.; Yang, Y.; Sun, B.F.; Chen, Y.S.; Xu, J.W.; Lai, W.Y.; Li, A.; Wang, X.; Bhattarai, D.P.; Xiao, W.; et al. 5-methylcytosine promotes mRNA export-NSUN2 as the methyltransferase and ALYREF as an m(5)C reader. Cell Res. 2017, 27, 606–625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Yang, Y.; Wang, L.; Han, X.; Yang, W.L.; Zhang, M.; Ma, H.L.; Sun, B.F.; Li, A.; Xia, J.; Chen, J.; et al. RNA 5-Methylcytosine Facilitates the Maternal-to-Zygotic Transition by Preventing Maternal mRNA Decay. Mol. Cell 2019, 75, 1188–1202.e1111. [Google Scholar] [CrossRef] [PubMed]
  75. Chen, X.; Li, A.; Sun, B.F.; Yang, Y.; Han, Y.N.; Yuan, X.; Chen, R.X.; Wei, W.S.; Liu, Y.; Gao, C.C.; et al. 5-methylcytosine promotes pathogenesis of bladder cancer through stabilizing mRNAs. Nat. Cell Biol. 2019, 21, 978–990. [Google Scholar] [CrossRef] [PubMed]
  76. Chen, H.; Yang, H.; Zhu, X.; Yadav, T.; Ouyang, J.; Truesdell, S.S.; Tan, J.; Wang, Y.; Duan, M.; Wei, L.; et al. m(5)C modification of mRNA serves a DNA damage code to promote homologous recombination. Nat. Commun. 2020, 11, 2834. [Google Scholar] [CrossRef] [PubMed]
  77. Guo, G.; Pan, K.; Fang, S.; Ye, L.; Tong, X.; Wang, Z.; Xue, X.; Zhang, H. Advances in mRNA 5-methylcytosine modifications: Detection, effectors, biological functions, and clinical relevance. Mol. Ther. Nucleic Acids 2021, 26, 575–593. [Google Scholar] [CrossRef]
  78. Trixl, L.; Lusser, A. The dynamic RNA modification 5-methylcytosine and its emerging role as an epitranscriptomic mark. Wiley Interdiscip. Rev. RNA 2019, 10, e1510. [Google Scholar] [CrossRef] [Green Version]
  79. Blaze, J.; Navickas, A.; Phillips, H.L.; Heissel, S.; Plaza-Jennings, A.; Miglani, S.; Asgharian, H.; Foo, M.; Katanski, C.D.; Watkins, C.P.; et al. Neuronal Nsun2 deficiency produces tRNA epitranscriptomic alterations and proteomic shifts impacting synaptic signaling and behavior. Nat. Commun. 2021, 12, 4913. [Google Scholar] [CrossRef]
  80. Hussain, S.; Aleksic, J.; Blanco, S.; Dietmann, S.; Frye, M. Characterizing 5-methylcytosine in the mammalian epitranscriptome. Genome Biol. 2013, 14, 215. [Google Scholar] [CrossRef]
  81. Bourgeois, G.; Ney, M.; Gaspar, I.; Aigueperse, C.; Schaefer, M.; Kellner, S.; Helm, M.; Motorin, Y. Eukaryotic rRNA Modification by Yeast 5-Methylcytosine-Methyltransferases and Human Proliferation-Associated Antigen p120. PLoS ONE 2015, 10, e0133321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Schosserer, M.; Minois, N.; Angerer, T.B.; Amring, M.; Dellago, H.; Harreither, E.; Calle-Perez, A.; Pircher, A.; Gerstl, M.P.; Pfeifenberger, S.; et al. Methylation of ribosomal RNA by NSUN5 is a conserved mechanism modulating organismal lifespan. Nat. Commun. 2015, 6, 6158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Metodiev, M.D.; Spahr, H.; Loguercio Polosa, P.; Meharg, C.; Becker, C.; Altmueller, J.; Habermann, B.; Larsson, N.G.; Ruzzenente, B. NSUN4 is a dual function mitochondrial protein required for both methylation of 12S rRNA and coordination of mitoribosomal assembly. PLoS Genet. 2014, 10, e1004110. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Sun, Z.; Xue, S.; Zhang, M.; Xu, H.; Hu, X.; Chen, S.; Liu, Y.; Guo, M.; Cui, H. Aberrant NSUN2-mediated m(5)C modification of H19 lncRNA is associated with poor differentiation of hepatocellular carcinoma. Oncogene 2020, 39, 6906–6919. [Google Scholar] [CrossRef]
  85. Aguilo, F.; Li, S.; Balasubramaniyan, N.; Sancho, A.; Benko, S.; Zhang, F.; Vashisht, A.; Rengasamy, M.; Andino, B.; Chen, C.H.; et al. Deposition of 5-Methylcytosine on Enhancer RNAs Enables the Coactivator Function of PGC-1alpha. Cell Rep. 2016, 14, 479–492. [Google Scholar] [CrossRef] [Green Version]
  86. Amort, T.; Rieder, D.; Wille, A.; Khokhlova-Cubberley, D.; Riml, C.; Trixl, L.; Jia, X.Y.; Micura, R.; Lusser, A. Distinct 5-methylcytosine profiles in poly(A) RNA from mouse embryonic stem cells and brain. Genome Biol. 2017, 18, 1. [Google Scholar] [CrossRef] [Green Version]
  87. Xing, J.; Yi, J.; Cai, X.; Tang, H.; Liu, Z.; Zhang, X.; Martindale, J.L.; Yang, X.; Jiang, B.; Gorospe, M.; et al. NSun2 Promotes Cell Growth via Elevating Cyclin-Dependent Kinase 1 Translation. Mol. Cell Biol. 2015, 35, 4043–4052. [Google Scholar] [CrossRef] [Green Version]
  88. Hoernes, T.P.; Clementi, N.; Faserl, K.; Glasner, H.; Breuker, K.; Lindner, H.; Huttenhofer, A.; Erlacher, M.D. Nucleotide modifications within bacterial messenger RNAs regulate their translation and are able to rewire the genetic code. Nucleic Acids Res. 2016, 44, 852–862. [Google Scholar] [CrossRef] [Green Version]
  89. Song, H.; Zhang, J.; Liu, B.; Xu, J.; Cai, B.; Yang, H.; Straube, J.; Yu, X.; Ma, T. Biological roles of RNA m(5)C modification and its implications in Cancer immunotherapy. Biomark. Res. 2022, 10, 15. [Google Scholar] [CrossRef]
  90. Han, L.; Marcus, E.; D’Silva, S.; Phizicky, E.M.S. cerevisiae Trm140 has two recognition modes for 3-methylcytidine modification of the anticodon loop of tRNA substrates. RNA 2017, 23, 406–419. [Google Scholar] [CrossRef]
  91. Cui, J.; Liu, Q.; Sendinc, E.; Shi, Y.; Gregory, R.I. Nucleotide resolution profiling of m3C RNA modification by HAC-seq. Nucleic Acids Res. 2021, 49, e27. [Google Scholar] [CrossRef] [PubMed]
  92. Xu, L.; Liu, X.; Sheng, N.; Oo, K.S.; Liang, J.; Chionh, Y.H.; Xu, J.; Ye, F.; Gao, Y.G.; Dedon, P.C.; et al. Three distinct 3-methylcytidine (m(3)C) methyltransferases modify tRNA and mRNA in mice and humans. J. Biol. Chem. 2017, 292, 14695–14703. [Google Scholar] [CrossRef] [Green Version]
  93. Mao, X.L.; Li, Z.H.; Huang, M.H.; Wang, J.T.; Zhou, J.B.; Li, Q.R.; Xu, H.; Wang, X.J.; Zhou, X.L. Mutually exclusive substrate selection strategy by human m3C RNA transferases METTL2A and METTL6. Nucleic Acids Res. 2021, 49, 8309–8323. [Google Scholar] [CrossRef] [PubMed]
  94. Huang, M.H.; Peng, G.X.; Mao, X.L.; Wang, J.T.; Zhou, J.B.; Zhang, J.H.; Chen, M.; Wang, E.D.; Zhou, X.L. Molecular basis for human mitochondrial tRNA m3C modification by alternatively spliced METTL8. Nucleic Acids Res. 2022, 50, 4012–4028. [Google Scholar] [CrossRef]
  95. Chen, Z.; Qi, M.; Shen, B.; Luo, G.; Wu, Y.; Li, J.; Lu, Z.; Zheng, Z.; Dai, Q.; Wang, H. Transfer RNA demethylase ALKBH3 promotes cancer progression via induction of tRNA-derived small RNAs. Nucleic Acids Res. 2019, 47, 2533–2545. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Ma, C.J.; Ding, J.H.; Ye, T.T.; Yuan, B.F.; Feng, Y.Q. AlkB Homologue 1 Demethylates N(3)-Methylcytidine in mRNA of Mammals. ACS Chem. Biol. 2019, 14, 1418–1425. [Google Scholar] [CrossRef]
  97. Bohnsack, K.E.; Kleiber, N.; Lemus-Diaz, N.; Bohnsack, M.T. Roles and dynamics of 3-methylcytidine in cellular RNAs. Trends Biochem. Sci. 2022, 47, 596–608. [Google Scholar] [CrossRef]
  98. Liu, Z.; Zhang, J. Human C-to-U Coding RNA Editing is Largely Nonadaptive. Mol. Biol. Evol. 2018, 35, 963–969. [Google Scholar] [CrossRef] [Green Version]
  99. Yu, Y.T.; Meier, U.T. RNA-guided isomerization of uridine to pseudouridine–pseudouridylation. RNA Biol 2014, 11, 1483–1494. [Google Scholar] [CrossRef] [Green Version]
  100. Rintala-Dempsey, A.C.; Kothe, U. Eukaryotic stand-alone pseudouridine synthases-RNA modifying enzymes and emerging regulators of gene expression? RNA Biol. 2017, 14, 1185–1196. [Google Scholar] [CrossRef]
  101. Schwartz, S.; Bernstein, D.A.; Mumbach, M.R.; Jovanovic, M.; Herbst, R.H.; Leon-Ricardo, B.X.; Engreitz, J.M.; Guttman, M.; Satija, R.; Lander, E.S.; et al. Transcriptome-wide mapping reveals widespread dynamic-regulated pseudouridylation of ncRNA and mRNA. Cell 2014, 159, 148–162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Safra, M.; Nir, R.; Farouq, D.; Vainberg Slutskin, I.; Schwartz, S. TRUB1 is the predominant pseudouridine synthase acting on mammalian mRNA via a predictable and conserved code. Genome Res. 2017, 27, 393–406. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Durant, P.C.; Davis, D.R. Stabilization of the anticodon stem-loop of tRNALys,3 by an A+-C base-pair and by pseudouridine. J Mol. Biol. 1999, 285, 115–131. [Google Scholar] [CrossRef] [PubMed]
  104. Taoka, M.; Nobe, Y.; Yamaki, Y.; Sato, K.; Ishikawa, H.; Izumikawa, K.; Yamauchi, Y.; Hirota, K.; Nakayama, H.; Takahashi, N.; et al. Landscape of the complete RNA chemical modifications in the human 80S ribosome. Nucleic Acids Res. 2018, 46, 9289–9298. [Google Scholar] [CrossRef] [PubMed]
  105. Borchardt, E.K.; Martinez, N.M.; Gilbert, W.V. Regulation and Function of RNA Pseudouridylation in Human Cells. Annu. Rev. Genet. 2020, 54, 309–336. [Google Scholar] [CrossRef]
  106. Martinez, N.M.; Su, A.; Burns, M.C.; Nussbacher, J.K.; Schaening, C.; Sathe, S.; Yeo, G.W.; Gilbert, W.V. Pseudouridine synthases modify human pre-mRNA co-transcriptionally and affect pre-mRNA processing. Mol. Cell 2022, 82, 645–659.e649. [Google Scholar] [CrossRef] [PubMed]
  107. Carlile, T.M.; Martinez, N.M.; Schaening, C.; Su, A.; Bell, T.A.; Zinshteyn, B.; Gilbert, W.V. mRNA structure determines modification by pseudouridine synthase 1. Nat. Chem. Biol. 2019, 15, 966–974. [Google Scholar] [CrossRef]
  108. Dai, Q.; Zhang, L.S.; Sun, H.L.; Pajdzik, K.; Yang, L.; Ye, C.; Ju, C.W.; Liu, S.; Wang, Y.; Zheng, Z.; et al. Quantitative sequencing using BID-seq uncovers abundant pseudouridines in mammalian mRNA at base resolution. Nat. Biotechnol. 2022. [Google Scholar] [CrossRef]
  109. Eyler, D.E.; Franco, M.K.; Batool, Z.; Wu, M.Z.; Dubuke, M.L.; Dobosz-Bartoszek, M.; Jones, J.D.; Polikanov, Y.S.; Roy, B.; Koutmou, K.S. Pseudouridinylation of mRNA coding sequences alters translation. Proc. Natl. Acad. Sci. USA 2019, 116, 23068–23074. [Google Scholar] [CrossRef]
  110. Donmez, G.; Hartmuth, K.; Luhrmann, R. Modified nucleotides at the 5′ end of human U2 snRNA are required for spliceosomal E-complex formation. RNA 2004, 10, 1925–1933. [Google Scholar] [CrossRef]
  111. Kennedy, S.D.; Bauer, W.J.; Wang, W.; Kielkopf, C.L. Dynamic stacking of an expected branch point adenosine in duplexes containing pseudouridine-modified or unmodified U2 snRNA sites. Biochem. Biophys. Res. Commun. 2019, 511, 416–421. [Google Scholar] [CrossRef] [PubMed]
  112. Wan, R.; Yan, C.; Bai, R.; Wang, L.; Huang, M.; Wong, C.C.; Shi, Y. The 3.8 A structure of the U4/U6.U5 tri-snRNP: Insights into spliceosome assembly and catalysis. Science 2016, 351, 466–475. [Google Scholar] [CrossRef] [PubMed]
  113. Whelan, F.; Jenkins, H.T.; Griffiths, S.C.; Byrne, R.T.; Dodson, E.J.; Antson, A.A. From bacterial to human dihydrouridine synthase: Automated structure determination. Acta Crystallogr. D Biol. Crystallogr. 2015, 71, 1564–1571. [Google Scholar] [CrossRef] [PubMed]
  114. Mittelstadt, M.; Frump, A.; Khuu, T.; Fowlkes, V.; Handy, I.; Patel, C.V.; Patel, R.C. Interaction of human tRNA-dihydrouridine synthase-2 with interferon-induced protein kinase PKR. Nucleic Acids Res. 2008, 36, 998–1008. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Finet, O.; Yague-Sanz, C.; Marchand, F.; Hermand, D. The Dihydrouridine landscape from tRNA to mRNA: A perspective on synthesis, structural impact and function. RNA Biol. 2022, 19, 735–750. [Google Scholar] [CrossRef] [PubMed]
  116. Finet, O.; Yague-Sanz, C.; Kruger, L.K.; Tran, P.; Migeot, V.; Louski, M.; Nevers, A.; Rougemaille, M.; Sun, J.; Ernst, F.G.M.; et al. Transcription-wide mapping of dihydrouridine reveals that mRNA dihydrouridylation is required for meiotic chromosome segregation. Mol. Cell 2022, 82, 404–419.e409. [Google Scholar] [CrossRef]
  117. Tomikawa, C. 7-Methylguanosine Modifications in Transfer RNA (tRNA). Int. J. Mol. Sci. 2018, 19, 4080. [Google Scholar] [CrossRef] [Green Version]
  118. Ramanathan, A.; Robb, G.B.; Chan, S.H. mRNA capping: Biological functions and applications. Nucleic Acids Res. 2016, 44, 7511–7526. [Google Scholar] [CrossRef] [Green Version]
  119. Lin, S.; Liu, Q.; Lelyveld, V.S.; Choe, J.; Szostak, J.W.; Gregory, R.I. Mettl1/Wdr4-Mediated m(7)G tRNA Methylome is Required for Normal mRNA Translation and Embryonic Stem Cell Self-Renewal and Differentiation. Mol. Cell 2018, 71, 244–255.e245. [Google Scholar] [CrossRef] [Green Version]
  120. Haag, S.; Kretschmer, J.; Bohnsack, M.T. WBSCR22/Merm1 is required for late nuclear pre-ribosomal RNA processing and mediates N7-methylation of G1639 in human 18S rRNA. RNA 2015, 21, 180–187. [Google Scholar] [CrossRef]
  121. Khan, A.A.; Huang, H.; Zhao, Y.; Li, H.; Pan, R.; Wang, S.; Liu, X. WBSCR22 and TRMT112 synergistically suppress cell proliferation, invasion and tumorigenesis in pancreatic cancer via transcriptional regulation of ISG15. Int. J. Oncol. 2022, 60, 24. [Google Scholar] [CrossRef] [PubMed]
  122. Pandolfini, L.; Barbieri, I.; Bannister, A.J.; Hendrick, A.; Andrews, B.; Webster, N.; Murat, P.; Mach, P.; Brandi, R.; Robson, S.C.; et al. METTL1 Promotes let-7 MicroRNA Processing via m7G Methylation. Mol. Cell 2019, 74, 1278–1290.e1279. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Powell, C.A.; Kopajtich, R.; D’Souza, A.R.; Rorbach, J.; Kremer, L.S.; Husain, R.A.; Dallabona, C.; Donnini, C.; Alston, C.L.; Griffin, H.; et al. TRMT5 Mutations Cause a Defect in Post-transcriptional Modification of Mitochondrial tRNA Associated with Multiple Respiratory-Chain Deficiencies. Am. J. Hum. Genet. 2015, 97, 319–328. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Bhatta, A.; Dienemann, C.; Cramer, P.; Hillen, H.S. Structural basis of RNA processing by human mitochondrial RNase P. Nat. Struct. Mol. Biol. 2021, 28, 713–723. [Google Scholar] [CrossRef]
  125. Vilardo, E.; Rossmanith, W. Molecular insights into HSD10 disease: Impact of SDR5C1 mutations on the human mitochondrial RNase P complex. Nucleic Acids Res. 2015, 43, 5112–5119. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Oerum, S.; Roovers, M.; Rambo, R.P.; Kopec, J.; Bailey, H.J.; Fitzpatrick, F.; Newman, J.A.; Newman, W.G.; Amberger, A.; Zschocke, J.; et al. Structural insight into the human mitochondrial tRNA purine N1-methyltransferase and ribonuclease P complexes. J. Biol. Chem. 2018, 293, 12862–12876. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Hong, S.; Sunita, S.; Maehigashi, T.; Hoffer, E.D.; Dunkle, J.A.; Dunham, C.M. Mechanism of tRNA-mediated +1 ribosomal frameshifting. Proc. Natl. Acad. Sci. USA 2018, 115, 11226–11231. [Google Scholar] [CrossRef] [Green Version]
  128. Zhou, H.; Kimsey, I.J.; Nikolova, E.N.; Sathyamoorthy, B.; Grazioli, G.; McSally, J.; Bai, T.; Wunderlich, C.H.; Kreutz, C.; Andricioaei, I.; et al. m(1)A and m(1)G disrupt A-RNA structure through the intrinsic instability of Hoogsteen base pairs. Nat. Struct. Mol. Biol. 2016, 23, 803–810. [Google Scholar] [CrossRef] [Green Version]
  129. Hillmeier, M.; Wagner, M.; Ensfelder, T.; Korytiakova, E.; Thumbs, P.; Muller, M.; Carell, T. Synthesis and structure elucidation of the human tRNA nucleoside mannosyl-queuosine. Nat. Commun. 2021, 12, 7123. [Google Scholar] [CrossRef]
  130. Tuorto, F.; Legrand, C.; Cirzi, C.; Federico, G.; Liebers, R.; Muller, M.; Ehrenhofer-Murray, A.E.; Dittmar, G.; Grone, H.J.; Lyko, F. Queuosine-modified tRNAs confer nutritional control of protein translation. EMBO J. 2018, 37, e99777. [Google Scholar] [CrossRef]
  131. Nishimura, S. Structure, biosynthesis, and function of queuosine in transfer RNA. Prog. Nucleic Acid Res. Mol. Biol. 1983, 28, 49–73. [Google Scholar] [CrossRef]
  132. Zhang, W.; Pan, T. Quantitative probing of glycosylated queuosine modifications in tRNA. Methods Enzymol. 2021, 658, 73–82. [Google Scholar] [CrossRef]
  133. Li, Z.; Mao, J.; Huang, D.; Song, B.; Meng, J. RNADSN: Transfer-Learning 5-Methyluridine (m(5)U) Modification on mRNAs from Common Features of tRNA. Int. J. Mol. Sci. 2022, 23, 13493. [Google Scholar] [CrossRef] [PubMed]
  134. Powell, C.A.; Minczuk, M. TRMT2B is responsible for both tRNA and rRNA m(5)U-methylation in human mitochondria. RNA Biol. 2020, 17, 451–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Pereira, M.; Ribeiro, D.R.; Pinheiro, M.M.; Ferreira, M.; Kellner, S.; Soares, A.R. m(5)U54 tRNA Hypomodification by Lack of TRMT2A Drives the Generation of tRNA-Derived Small RNAs. Int. J. Mol. Sci. 2021, 22, 2941. [Google Scholar] [CrossRef] [PubMed]
  136. Sharma, S.; Yang, J.; Duttmann, S.; Watzinger, P.; Kotter, P.; Entian, K.D. Identification of novel methyltransferases, Bmt5 and Bmt6, responsible for the m3U methylations of 25S rRNA in Saccharomyces cerevisiae. Nucleic Acids Res. 2014, 42, 3246–3260. [Google Scholar] [CrossRef] [Green Version]
  137. Song, J.; Song, J.; Mo, B.; Chen, X. Uridylation and adenylation of RNAs. Sci. China Life Sci. 2015, 58, 1057–1066. [Google Scholar] [CrossRef] [Green Version]
  138. De Almeida, C.; Scheer, H.; Zuber, H.; Gagliardi, D. RNA uridylation: A key posttranscriptional modification shaping the coding and noncoding transcriptome. Wiley Interdiscip. Rev. RNA 2018, 9, e1440. [Google Scholar] [CrossRef]
  139. Lim, J.; Ha, M.; Chang, H.; Kwon, S.C.; Simanshu, D.K.; Patel, D.J.; Kim, V.N. Uridylation by TUT4 and TUT7 marks mRNA for degradation. Cell 2014, 159, 1365–1376. [Google Scholar] [CrossRef] [Green Version]
  140. Zhang, P.; Frederick, M.I.; Heinemann, I.U. Terminal Uridylyltransferases TUT4/7 Regulate microRNA and mRNA Homeostasis. Cells 2022, 11, 3742. [Google Scholar] [CrossRef]
  141. Ustianenko, D.; Hrossova, D.; Potesil, D.; Chalupnikova, K.; Hrazdilova, K.; Pachernik, J.; Cetkovska, K.; Uldrijan, S.; Zdrahal, Z.; Vanacova, S. Mammalian DIS3L2 exoribonuclease targets the uridylated precursors of let-7 miRNAs. RNA 2013, 19, 1632–1638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Somme, J.; Van Laer, B.; Roovers, M.; Steyaert, J.; Versees, W.; Droogmans, L. Characterization of two homologous 2′-O-methyltransferases showing different specificities for their tRNA substrates. RNA 2014, 20, 1257–1271. [Google Scholar] [CrossRef] [Green Version]
  143. Dimitrova, D.G.; Teysset, L.; Carre, C. RNA 2′-O-Methylation (Nm) Modification in Human Diseases. Genes 2019, 10, 117. [Google Scholar] [CrossRef] [Green Version]
  144. Sproat, B.S.; Lamond, A.I.; Beijer, B.; Neuner, P.; Ryder, U. Highly efficient chemical synthesis of 2′-O-methyloligoribonucleotides and tetrabiotinylated derivatives; novel probes that are resistant to degradation by RNA or DNA specific nucleases. Nucleic Acids Res. 1989, 17, 3373–3386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Majlessi, M.; Nelson, N.C.; Becker, M.M. Advantages of 2′-O-methyl oligoribonucleotide probes for detecting RNA targets. Nucleic Acids Res. 1998, 26, 2224–2229. [Google Scholar] [CrossRef] [PubMed]
  146. Inoue, H.; Hayase, Y.; Imura, A.; Iwai, S.; Miura, K.; Ohtsuka, E. Synthesis and hybridization studies on two complementary nona(2′-O-methyl)ribonucleotides. Nucleic Acids Res. 1987, 15, 6131–6148. [Google Scholar] [CrossRef] [PubMed]
  147. Kawai, G.; Yamamoto, Y.; Kamimura, T.; Masegi, T.; Sekine, M.; Hata, T.; Iimori, T.; Watanabe, T.; Miyazawa, T.; Yokoyama, S. Conformational rigidity of specific pyrimidine residues in tRNA arises from posttranscriptional modifications that enhance steric interaction between the base and the 2′-hydroxyl group. Biochemistry 1992, 31, 1040–1046. [Google Scholar] [CrossRef]
  148. Leschziner, G.D.; Coffey, A.J.; Andrew, T.; Gregorio, S.P.; Dias-Neto, E.; Calafato, M.; Bentley, D.R.; Kinton, L.; Sander, J.W.; Johnson, M.R. Q8IYL2 is a candidate gene for the familial epilepsy syndrome of Partial Epilepsy with Pericentral Spikes (PEPS). Epilepsy Res. 2011, 96, 109–115. [Google Scholar] [CrossRef]
  149. Decatur, W.A.; Fournier, M.J. rRNA modifications and ribosome function. Trends Biochem. Sci. 2002, 27, 344–351. [Google Scholar] [CrossRef]
  150. Elliott, B.A.; Ho, H.T.; Ranganathan, S.V.; Vangaveti, S.; Ilkayeva, O.; Abou Assi, H.; Choi, A.K.; Agris, P.F.; Holley, C.L. Modification of messenger RNA by 2′-O-methylation regulates gene expression in vivo. Nat. Commun. 2019, 10, 3401. [Google Scholar] [CrossRef]
  151. Dai, Q.; Moshitch-Moshkovitz, S.; Han, D.; Kol, N.; Amariglio, N.; Rechavi, G.; Dominissini, D.; He, C. Nm-seq maps 2′-O-methylation sites in human mRNA with base precision. Nat. Methods 2017, 14, 695–698. [Google Scholar] [CrossRef] [PubMed]
  152. Werner, M.; Purta, E.; Kaminska, K.H.; Cymerman, I.A.; Campbell, D.A.; Mittra, B.; Zamudio, J.R.; Sturm, N.R.; Jaworski, J.; Bujnicki, J.M. 2′-O-ribose methylation of cap2 in human: Function and evolution in a horizontally mobile family. Nucleic Acids Res. 2011, 39, 4756–4768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Wu, Y.; Xie, L.; Wang, M.; Xiong, Q.; Guo, Y.; Liang, Y.; Li, J.; Sheng, R.; Deng, P.; Wang, Y.; et al. Mettl3-mediated m(6)A RNA methylation regulates the fate of bone marrow mesenchymal stem cells and osteoporosis. Nat. Commun. 2018, 9, 4772. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Ianniello, Z.; Sorci, M.; Ceci Ginistrelli, L.; Iaiza, A.; Marchioni, M.; Tito, C.; Capuano, E.; Masciarelli, S.; Ottone, T.; Attrotto, C.; et al. New insight into the catalytic -dependent and -independent roles of METTL3 in sustaining aberrant translation in chronic myeloid leukemia. Cell Death Dis. 2021, 12, 870. [Google Scholar] [CrossRef] [PubMed]
  155. Muller, S.; Glass, M.; Singh, A.K.; Haase, J.; Bley, N.; Fuchs, T.; Lederer, M.; Dahl, A.; Huang, H.; Chen, J.; et al. IGF2BP1 promotes SRF-dependent transcription in cancer in a m6A- and miRNA-dependent manner. Nucleic Acids Res. 2019, 47, 375–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Wu, Y.; Chen, Z.; Xie, G.; Zhang, H.; Wang, Z.; Zhou, J.; Chen, F.; Li, J.; Chen, L.; Niu, H.; et al. RNA m(1)A methylation regulates glycolysis of cancer cells through modulating ATP5D. Proc. Natl. Acad. Sci. USA 2022, 119, e2119038119. [Google Scholar] [CrossRef] [PubMed]
  157. Shi, Q.; Xue, C.; Yuan, X.; He, Y.; Yu, Z. Gene signatures and prognostic values of m1A-related regulatory genes in hepatocellular carcinoma. Sci. Rep. 2020, 10, 15083. [Google Scholar] [CrossRef] [PubMed]
  158. Esteve-Puig, R.; Climent, F.; Pineyro, D.; Domingo-Domenech, E.; Davalos, V.; Encuentra, M.; Rea, A.; Espejo-Herrera, N.; Soler, M.; Lopez, M.; et al. Epigenetic loss of m1A RNA demethylase ALKBH3 in Hodgkin lymphoma targets collagen, conferring poor clinical outcome. Blood 2021, 137, 994–999. [Google Scholar] [CrossRef]
  159. Srinivasan, S.; Torres, A.G.; Ribas de Pouplana, L. Inosine in Biology and Disease. Genes 2021, 12, 600. [Google Scholar] [CrossRef]
  160. Xu, L.D.; Ohman, M. ADAR1 Editing and its Role in Cancer. Genes 2018, 10, 12. [Google Scholar] [CrossRef]
  161. Xue, C.; Chu, Q.; Zheng, Q.; Jiang, S.; Bao, Z.; Su, Y.; Lu, J.; Li, L. Role of main RNA modifications in cancer: N(6)-methyladenosine, 5-methylcytosine, and pseudouridine. Signal Transduct. Target Ther. 2022, 7, 142. [Google Scholar] [CrossRef]
  162. Jonkhout, N.; Tran, J.; Smith, M.A.; Schonrock, N.; Mattick, J.S.; Novoa, E.M. The RNA modification landscape in human disease. RNA 2017, 23, 1754–1769. [Google Scholar] [CrossRef] [Green Version]
  163. Wang, S.; Li, H.; Liu, J.; Zhang, Q.; Xu, W.; Xiang, J.; Fang, L.; Xu, P.; Li, Z. Integrative analysis of m3C associated genes reveals METTL2A as a potential oncogene in breast Cancer. J. Transl. Med. 2022, 20, 476. [Google Scholar] [CrossRef] [PubMed]
  164. Ignatova, V.V.; Kaiser, S.; Ho, J.S.Y.; Bing, X.; Stolz, P.; Tan, Y.X.; Lee, C.L.; Gay, F.P.H.; Lastres, P.R.; Gerlini, R.; et al. METTL6 is a tRNA m(3)C methyltransferase that regulates pluripotency and tumor cell growth. Sci. Adv. 2020, 6, eaaz4551. [Google Scholar] [CrossRef]
  165. Feng, B.; Zheng, M.H.; Zheng, Y.F.; Lu, A.G.; Li, J.W.; Wang, M.L.; Ma, J.J.; Xu, G.W.; Liu, B.Y.; Zhu, Z.G. Normal and modified urinary nucleosides represent novel biomarkers for colorectal cancer diagnosis and surgery monitoring. J. Gastroenterol. Hepatol. 2005, 20, 1913–1919. [Google Scholar] [CrossRef]
  166. Teng, P.C.; Liang, Y.; Yarmishyn, A.A.; Hsiao, Y.J.; Lin, T.Y.; Lin, T.W.; Teng, Y.C.; Yang, Y.P.; Wang, M.L.; Chien, C.S.; et al. RNA Modifications and Epigenetics in Modulation of Lung Cancer and Pulmonary Diseases. Int. J. Mol. Sci. 2021, 22, 10592. [Google Scholar] [CrossRef] [PubMed]
  167. Xia, P.; Zhang, H.; Xu, K.; Jiang, X.; Gao, M.; Wang, G.; Liu, Y.; Yao, Y.; Chen, X.; Ma, W.; et al. MYC-targeted WDR4 promotes proliferation, metastasis, and sorafenib resistance by inducing CCNB1 translation in hepatocellular carcinoma. Cell Death Dis. 2021, 12, 691. [Google Scholar] [CrossRef] [PubMed]
  168. Zhang, J.; Lu, R.; Zhang, Y.; Matuszek, Z.; Zhang, W.; Xia, Y.; Pan, T.; Sun, J. tRNA Queuosine Modification Enzyme Modulates the Growth and Microbiome Recruitment to Breast Tumors. Cancers 2020, 12, 628. [Google Scholar] [CrossRef] [Green Version]
  169. Pathak, C.; Jaiswal, Y.K.; Vinayak, M. Hypomodification of transfer RNA in cancer with respect to queuosine. RNA Biol. 2005, 2, 143–148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Kato, T.; Daigo, Y.; Hayama, S.; Ishikawa, N.; Yamabuki, T.; Ito, T.; Miyamoto, M.; Kondo, S.; Nakamura, Y. A novel human tRNA-dihydrouridine synthase involved in pulmonary carcinogenesis. Cancer Res. 2005, 65, 5638–5646. [Google Scholar] [CrossRef]
  171. Barbieri, I.; Kouzarides, T. Role of RNA modifications in cancer. Nat. Rev. Cancer 2020, 20, 303–322. [Google Scholar] [CrossRef]
  172. Vu, L.P.; Pickering, B.F.; Cheng, Y.; Zaccara, S.; Nguyen, D.; Minuesa, G.; Chou, T.; Chow, A.; Saletore, Y.; MacKay, M.; et al. The N(6)-methyladenosine (m(6)A)-forming enzyme METTL3 controls myeloid differentiation of normal hematopoietic and leukemia cells. Nat. Med. 2017, 23, 1369–1376. [Google Scholar] [CrossRef]
  173. Li, Z.; Weng, H.; Su, R.; Weng, X.; Zuo, Z.; Li, C.; Huang, H.; Nachtergaele, S.; Dong, L.; Hu, C.; et al. FTO Plays an Oncogenic Role in Acute Myeloid Leukemia as a N(6)-Methyladenosine RNA Demethylase. Cancer Cell 2017, 31, 127–141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Han, D.; Liu, J.; Chen, C.; Dong, L.; Liu, Y.; Chang, R.; Huang, X.; Liu, Y.; Wang, J.; Dougherty, U.; et al. Anti-tumour immunity controlled through mRNA m(6)A methylation and YTHDF1 in dendritic cells. Nature 2019, 566, 270–274. [Google Scholar] [CrossRef] [PubMed]
  175. Paris, J.; Morgan, M.; Campos, J.; Spencer, G.J.; Shmakova, A.; Ivanova, I.; Mapperley, C.; Lawson, H.; Wotherspoon, D.A.; Sepulveda, C.; et al. Targeting the RNA m(6)A Reader YTHDF2 Selectively Compromises Cancer Stem Cells in Acute Myeloid Leukemia. Cell Stem Cell 2019, 25, 137–148.e136. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Thornton, J.E.; Chang, H.M.; Piskounova, E.; Gregory, R.I. Lin28-mediated control of let-7 microRNA expression by alternative TUTases Zcchc11 (TUT4) and Zcchc6 (TUT7). RNA 2012, 18, 1875–1885. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Piskounova, E.; Polytarchou, C.; Thornton, J.E.; LaPierre, R.J.; Pothoulakis, C.; Hagan, J.P.; Iliopoulos, D.; Gregory, R.I. Lin28A and Lin28B inhibit let-7 microRNA biogenesis by distinct mechanisms. Cell 2011, 147, 1066–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Zipeto, M.A.; Court, A.C.; Sadarangani, A.; Delos Santos, N.P.; Balaian, L.; Chun, H.J.; Pineda, G.; Morris, S.R.; Mason, C.N.; Geron, I.; et al. ADAR1 Activation Drives Leukemia Stem Cell Self-Renewal by Impairing Let-7 Biogenesis. Cell Stem Cell 2016, 19, 177–191. [Google Scholar] [CrossRef] [Green Version]
  179. Ruggero, D.; Grisendi, S.; Piazza, F.; Rego, E.; Mari, F.; Rao, P.H.; Cordon-Cardo, C.; Pandolfi, P.P. Dyskeratosis congenita and cancer in mice deficient in ribosomal RNA modification. Science 2003, 299, 259–262. [Google Scholar] [CrossRef] [PubMed]
  180. Rocchi, L.; Pacilli, A.; Sethi, R.; Penzo, M.; Schneider, R.J.; Trere, D.; Brigotti, M.; Montanaro, L. Dyskerin depletion increases VEGF mRNA internal ribosome entry site-mediated translation. Nucleic Acids Res. 2013, 41, 8308–8318. [Google Scholar] [CrossRef]
  181. Montanaro, L.; Calienni, M.; Bertoni, S.; Rocchi, L.; Sansone, P.; Storci, G.; Santini, D.; Ceccarelli, C.; Taffurelli, M.; Carnicelli, D.; et al. Novel dyskerin-mediated mechanism of p53 inactivation through defective mRNA translation. Cancer Res. 2010, 70, 4767–4777. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  182. Livneh, I.; Moshitch-Moshkovitz, S.; Amariglio, N.; Rechavi, G.; Dominissini, D. The m(6)A epitranscriptome: Transcriptome plasticity in brain development and function. Nat. Rev. Neurosci. 2020, 21, 36–51. [Google Scholar] [CrossRef] [PubMed]
  183. Dermentzaki, G.; Lotti, F. New Insights on the Role of N (6)-Methyladenosine RNA Methylation in the Physiology and Pathology of the Nervous System. Front. Mol. Biosci. 2020, 7, 555372. [Google Scholar] [CrossRef] [PubMed]
  184. Boissel, S.; Reish, O.; Proulx, K.; Kawagoe-Takaki, H.; Sedgwick, B.; Yeo, G.S.; Meyre, D.; Golzio, C.; Molinari, F.; Kadhom, N.; et al. Loss-of-function mutation in the dioxygenase-encoding FTO gene causes severe growth retardation and multiple malformations. Am. J. Hum. Genet. 2009, 85, 106–111. [Google Scholar] [CrossRef] [Green Version]
  185. Choudhry, Z.; Sengupta, S.M.; Grizenko, N.; Thakur, G.A.; Fortier, M.E.; Schmitz, N.; Joober, R. Association between obesity-related gene FTO and ADHD. Obesity 2013, 21, E738–E744. [Google Scholar] [CrossRef]
  186. Samaan, Z.; Anand, S.S.; Zhang, X.; Desai, D.; Rivera, M.; Pare, G.; Thabane, L.; Xie, C.; Gerstein, H.; Engert, J.C.; et al. The protective effect of the obesity-associated rs9939609 A variant in fat mass- and obesity-associated gene on depression. Mol. Psychiatry 2013, 18, 1281–1286. [Google Scholar] [CrossRef]
  187. Keller, L.; Xu, W.; Wang, H.X.; Winblad, B.; Fratiglioni, L.; Graff, C. The obesity related gene, FTO, interacts with APOE, and is associated with Alzheimer’s disease risk: A prospective cohort study. J. Alzheimers Dis. 2011, 23, 461–469. [Google Scholar] [CrossRef]
  188. Li, H.; Ren, Y.; Mao, K.; Hua, F.; Yang, Y.; Wei, N.; Yue, C.; Li, D.; Zhang, H. FTO is involved in Alzheimer’s disease by targeting TSC1-mTOR-Tau signaling. Biochem. Biophys. Res. Commun. 2018, 498, 234–239. [Google Scholar] [CrossRef] [PubMed]
  189. Han, M.; Liu, Z.; Xu, Y.; Liu, X.; Wang, D.; Li, F.; Wang, Y.; Bi, J. Abnormality of m6A mRNA Methylation is Involved in Alzheimer’s Disease. Front. Neurosci. 2020, 14, 98. [Google Scholar] [CrossRef]
  190. Hess, M.E.; Hess, S.; Meyer, K.D.; Verhagen, L.A.; Koch, L.; Bronneke, H.S.; Dietrich, M.O.; Jordan, S.D.; Saletore, Y.; Elemento, O.; et al. The fat mass and obesity associated gene (Fto) regulates activity of the dopaminergic midbrain circuitry. Nat. Neurosci. 2013, 16, 1042–1048. [Google Scholar] [CrossRef]
  191. Chen, X.; Yu, C.; Guo, M.; Zheng, X.; Ali, S.; Huang, H.; Zhang, L.; Wang, S.; Huang, Y.; Qie, S.; et al. Down-Regulation of m6A mRNA Methylation is Involved in Dopaminergic Neuronal Death. ACS Chem. Neurosci. 2019, 10, 2355–2363. [Google Scholar] [CrossRef] [PubMed]
  192. Mitropoulos, K.; Merkouri Papadima, E.; Xiromerisiou, G.; Balasopoulou, A.; Charalampidou, K.; Galani, V.; Zafeiri, K.V.; Dardiotis, E.; Ralli, S.; Deretzi, G.; et al. Genomic variants in the FTO gene are associated with sporadic amyotrophic lateral sclerosis in Greek patients. Hum. Genom. 2017, 11, 30. [Google Scholar] [CrossRef]
  193. Du, T.; Rao, S.; Wu, L.; Ye, N.; Liu, Z.; Hu, H.; Xiu, J.; Shen, Y.; Xu, Q. An association study of the m6A genes with major depressive disorder in Chinese Han population. J. Affect. Disord. 2015, 183, 279–286. [Google Scholar] [CrossRef] [PubMed]
  194. Liu, X.; Shimada, T.; Otowa, T.; Wu, Y.Y.; Kawamura, Y.; Tochigi, M.; Iwata, Y.; Umekage, T.; Toyota, T.; Maekawa, M.; et al. Genome-wide Association Study of Autism Spectrum Disorder in the East Asian Populations. Autism Res. 2016, 9, 340–349. [Google Scholar] [CrossRef] [PubMed]
  195. Khan, M.A.; Rafiq, M.A.; Noor, A.; Hussain, S.; Flores, J.V.; Rupp, V.; Vincent, A.K.; Malli, R.; Ali, G.; Khan, F.S.; et al. Mutation in NSUN2, which encodes an RNA methyltransferase, causes autosomal-recessive intellectual disability. Am. J. Hum. Genet. 2012, 90, 856–863. [Google Scholar] [CrossRef] [Green Version]
  196. Martinez, F.J.; Lee, J.H.; Lee, J.E.; Blanco, S.; Nickerson, E.; Gabriel, S.; Frye, M.; Al-Gazali, L.; Gleeson, J.G. Whole exome sequencing identifies a splicing mutation in NSUN2 as a cause of a Dubowitz-like syndrome. J. Med. Genet. 2012, 49, 380–385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Shaheen, R.; Tasak, M.; Maddirevula, S.; Abdel-Salam, G.M.H.; Sayed, I.S.M.; Alazami, A.M.; Al-Sheddi, T.; Alobeid, E.; Phizicky, E.M.; Alkuraya, F.S. PUS7 mutations impair pseudouridylation in humans and cause intellectual disability and microcephaly. Hum. Genet. 2019, 138, 231–239. [Google Scholar] [CrossRef] [PubMed]
  198. Lee, S.H.; Kim, I.; Chung, B.C. Increased urinary level of oxidized nucleosides in patients with mild-to-moderate Alzheimer’s disease. Clin. Biochem. 2007, 40, 936–938. [Google Scholar] [CrossRef]
  199. deLorimier, E.; Hinman, M.N.; Copperman, J.; Datta, K.; Guenza, M.; Berglund, J.A. Pseudouridine Modification Inhibits Muscleblind-like 1 (MBNL1) Binding to CCUG Repeats and Minimally Structured RNA through Reduced RNA Flexibility. J. Biol. Chem. 2017, 292, 4350–4357. [Google Scholar] [CrossRef] [Green Version]
  200. Kubota-Sakashita, M.; Iwamoto, K.; Bundo, M.; Kato, T. A role of ADAR2 and RNA editing of glutamate receptors in mood disorders and schizophrenia. Mol. Brain 2014, 7, 5. [Google Scholar] [CrossRef]
  201. Zaganelli, S.; Rebelo-Guiomar, P.; Maundrell, K.; Rozanska, A.; Pierredon, S.; Powell, C.A.; Jourdain, A.A.; Hulo, N.; Lightowlers, R.N.; Chrzanowska-Lightowlers, Z.M.; et al. The Pseudouridine Synthase RPUSD4 is an Essential Component of Mitochondrial RNA Granules. J. Biol. Chem. 2017, 292, 4519–4532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  202. Deng, Y.; Zhou, Z.; Ji, W.; Lin, S.; Wang, M. METTL1-mediated m(7)G methylation maintains pluripotency in human stem cells and limits mesoderm differentiation and vascular development. Stem Cell Res. Ther. 2020, 11, 306. [Google Scholar] [CrossRef] [PubMed]
  203. Zhao, Y.; Kong, L.; Pei, Z.; Li, F.; Li, C.; Sun, X.; Shi, B.; Ge, J. m7G Methyltransferase METTL1 Promotes Post-ischemic Angiogenesis via Promoting VEGFA mRNA Translation. Front. Cell Dev. Biol. 2021, 9, 642080. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Venn diagram demonstrating the number of common RNA modifications that exist among many classes of RNA molecules, including tRNAs, mRNAs, and rRNAs, as well as snRNAs.
Figure 1. Venn diagram demonstrating the number of common RNA modifications that exist among many classes of RNA molecules, including tRNAs, mRNAs, and rRNAs, as well as snRNAs.
Ijms 24 02387 g001
Figure 2. Common eukaryotic mRNA modifications that have been identified. The abundancy of each modification is strictly regulated by (a) “writers”, proteins that incorporate the specific modification in specific mRNA sites, (b) “erasers”, which catalyze the removal of a specific modification from the mRNA and (c) “readers”, which interact with specific modifications to regulate a wide spectrum of cellular processes.
Figure 2. Common eukaryotic mRNA modifications that have been identified. The abundancy of each modification is strictly regulated by (a) “writers”, proteins that incorporate the specific modification in specific mRNA sites, (b) “erasers”, which catalyze the removal of a specific modification from the mRNA and (c) “readers”, which interact with specific modifications to regulate a wide spectrum of cellular processes.
Ijms 24 02387 g002
Figure 3. Graphical representation of the identified nucleotide modifications in eukaryotic tRNAs and their corresponding position in the tRNA sequence.
Figure 3. Graphical representation of the identified nucleotide modifications in eukaryotic tRNAs and their corresponding position in the tRNA sequence.
Ijms 24 02387 g003
Figure 4. Schematic demonstration of the detected RNA modifications in human 5.8S RNA, as well as their respective positions in the RNA sequence.
Figure 4. Schematic demonstration of the detected RNA modifications in human 5.8S RNA, as well as their respective positions in the RNA sequence.
Ijms 24 02387 g004
Table 1. Enzymes catalyzing the addition, removal and recognition of common RNA modifications in human cells.
Table 1. Enzymes catalyzing the addition, removal and recognition of common RNA modifications in human cells.
ModificationBase PairingEnzyme TypeEnzyme Name
m6ACanonicalWritersMETTL3, METTL14, METTL16, METTL5, ZCCHC4
ErasersALKBH5, FTO
ReadersYTHD, IGF2, IGF2BP1-3, eIF3, FMR1, LRPPRC, HNRNPC, HNRNPA2B1
m1ANon-canonicalWritersTRM6, TRMT10C, TRM61A, TRM61B, NML
ErasersALKBH1, ALKBH3
ReadersYTHDF2?, YTHDF3
A-to-INon-canonicalWritersADAT2-ADAT3, ADAT1, ADAR2, ADAR1
ErasersDAP3?
Readers?
m5CCanonicalWritersNSUN1 to NSUN7, DNMT2, p120, TRDMT1
ErasersTET1 to TET3?, MBD2/4?, ALKBH1?, ALKBH6?
ReadersALYREF, YBX1, RAD52
m3CNon-canonicalWritersMETTL2A, METTL2B, METTL6, METTL8
ErasersALKBH1, ALKBH3
Readers?
m1GNon-canonicalWritersTRMT5, TRMT10, RG9MTD2, RG9MTD1, RG9MTD3, SDR5C1
Erasers?
Readers?
m7GCanonicalWritersWBSCR22/TRMT112, RNMT, WDR4, METTL1
Erasers?
Readers?
QNon-canonicalWriters?
Erasers?
Readers?
C-to-UNon-canonicalWritersAPOBEC1, APOBEC3A?, APOBEC3G?
Erasers?
Readers?
DNon-canonicalWritersDUS1 to DUS4
Erasers?
Readers?
ΨNon-canonicalWritersPUS1 to PUS10, PUS7L, RPUSD1 to RPUSD4, DKC1
Erasers?
ReadersPUM2?
UridylationNon-canonicalWritersTUT4, TUT7
Erasers?
ReadersDIS3L2
NmCanonicalAm WriterFTSJ1 to FTSJ3, FBL
Cm WriterFTSJ1 to FTSJ3, CCDC76, FBL
Gm WriterTARBP1, FTSJ1, MRM1, RNMTL1, TrmH, FBL
Um WriterFTSJ1 to FTSJ3, FBL, TRMT44
Table 2. Position and type of RNA modifications that have been detected in human 18S rRNA.
Table 2. Position and type of RNA modifications that have been detected in human 18S rRNA.
PositionModificationPositionModificationPositionModificationPositionModification
27Am484Am822Ψ1328Gm
34Ψ509Gm863Ψ1337ac4C
36Ψ512Am866Ψ1347Ψ
93Ψ517Cm867Gm1367Ψ
99Am572Ψ897Ψ1383Am
105Ψ576Am918Ψ1391Cm
109Ψ590Am966Ψ1442Um
116Um601Gm1004Ψ1445Ψ
119Ψ609Ψ1031Am1447Gm
121Um621Cm1045Ψ1490Gm
159Am627Um1046Ψ1625Ψ
166Am644Gm1056Ψ1639m7G
172Um649Ψ1081Ψ1643Ψ
174Cm651Ψ1136Ψ1668Um
210Ψ668Am1174Ψ1678Am
218Ψ681Ψ1177Ψ1692Ψ
296Ψ683Gm1232Ψ1703Cm
354Um686Ψ1238Ψ1804Um
406Ψ797Cm1244Ψ1832m6A
428Um799Um1248m1acp3Ψ1842ac4C
436Gm801Ψ1272Cm1850m62A
462Cm814Ψ1288Um1851m62A
468Am815Ψ1326Um
Table 3. Position and type of RNA modifications that have been detected in human 28S rRNA.
Table 3. Position and type of RNA modifications that have been detected in human 28S rRNA.
PositionModificationPositionModificationPositionModificationPositionModification
389Am2495Ψ3804Am4417m5C
391Am2619Ψ3809Am4426Cm
1303Gm2774Am3820Cm4427Ψ
1309m1A2791Cm3823Ψ4441Ψ
1310Am2802Am3830Ψ4463Ψ
1313Am2811Cm3832Ψ4464Gm
1327Cm2824Um3846Am4468Um
1509Gm2826Ψ3848Cm4469Gm
1511Am2830Ψ3863Ψ4470Ψ
1521Am2848Cm3866Cm4491Ψ
1523Ψ2863Gm3878Gm4493Am
1569Ψ3606Gm3899Ψ4500m3U
1612Gm3616Ψ3904Um4502Ψ
1664Ψ3618Ψ3923Gm4401Ψ
1670Ψ3674Ψ3938Ψ4506Cm
1731Ψ3680Cm4020Gm4522Ψ
1747Gm3694Ψ4032Cm4541Am
1760Um3697Am4166Gm4546Ψ
1766Ψ3703Am4197Um4549Ψ
1768Ψ3709Ψ4198Gm4560Am
1769Ψ3713Ψ4220m6A4588Gm
1779Ψ3723Gm4263Ψ4590Um
1847Ψ3737Ψ4266Ψ4593Gm
1849Ψ3739Am4269Ψ4598Ψ
1858Am3741Ψ4276Um4606Ψ
1868Cm3743Ψ4282Ψ4607Gm
2338Cm3747Ψ4323Ψ4643Ψ
2350Am3749Ψ4331Ψ4659Ψ
2351Gm3761m5C4340Gm4937Ψ
2352Cm3764Am4362Gm4966Ψ
2388Am3771Gm4373Ψ4975Ψ
2402Um3787Cm4390Ψ4506Cm
2409Cm3797Ψm4393Ψ
2411Gm3801Ψ4412Ψ
Table 4. Regulatory roles of RNA modifications in various human diseases.
Table 4. Regulatory roles of RNA modifications in various human diseases.
ModificationPhysiological RoleHuman DiseasesReferences
m6AmRNA splicing, translation efficiency, transcriptional repression by lncRNAsAML, CML, Obesity, Osteoporosis, hepatocellular carcinoma, ADHD, AD, PD[21,153,154,155]
m1ATranslation efficiency, rRNA folding, tRNA folding and stabilityBreast, ovarian, cervical, pancreatic and hepatocellular cancer, leukemia[95,156,157,158]
A-to-IWobble codon recognition in tRNAs, mRNA stability and localizationColorectal, gastric, esophageal and lung cancer, HCC[159,160]
m5CmRNA stability, translation efficiency, rRNA folding, tRNA folding and stability, lncRNA stabilityBreast and bladder cancer, hypotonia, acidosis, ARID, DS[161,162]
m3CtRNA:mRNA stability, translation efficiencyAsthma, neoplastic and developmental pathologies[163,164]
m1GReduces translation efficiencyColorectal cancer[165]
m7GTranslation regulators, tRNA stabilityLung cancer, HCC[166,167]
QProtects tRNA from ribonuclease cleavageColorectal cancer, lymphoma[130,168,169]
DmRNA splicing, translation efficiency, tRNA structureLung cancer[170]
ΨRibosome assembly, translational efficiency, tRNA stabilityBreast, prostate and lung cancer, HCC[161]
NmStability of RNAsAD, asthma[143]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Adamopoulos, P.G.; Athanasopoulou, K.; Daneva, G.N.; Scorilas, A. The Repertoire of RNA Modifications Orchestrates a Plethora of Cellular Responses. Int. J. Mol. Sci. 2023, 24, 2387. https://doi.org/10.3390/ijms24032387

AMA Style

Adamopoulos PG, Athanasopoulou K, Daneva GN, Scorilas A. The Repertoire of RNA Modifications Orchestrates a Plethora of Cellular Responses. International Journal of Molecular Sciences. 2023; 24(3):2387. https://doi.org/10.3390/ijms24032387

Chicago/Turabian Style

Adamopoulos, Panagiotis G., Konstantina Athanasopoulou, Glykeria N. Daneva, and Andreas Scorilas. 2023. "The Repertoire of RNA Modifications Orchestrates a Plethora of Cellular Responses" International Journal of Molecular Sciences 24, no. 3: 2387. https://doi.org/10.3390/ijms24032387

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

Article Metrics

Back to TopTop