N6-Methyladenosine

Last updated
N6-Methyladenosine
N6-Methyladenosine.svg
Names
IUPAC name
N6-Methyladenosine
Systematic IUPAC name
(2R,3S,4R,5R)-2-(Hydroxymethyl)-5-[6-(methylamino)-9H-purin-9-yl]oxolane-2,3-diol
Other names
m6A
Identifiers
3D model (JSmol)
ChEBI
ChemSpider
PubChem CID
UNII
  • InChI=1S/C11H15N5O4/c1-12-9-6-10(14-3-13-9)16(4-15-6)11-8(19)7(18)5(2-17)20-11/h3-5,7-8,11,17-19H,2H2,1H3,(H,12,13,14)/t5-,7-,8-,11-/m1/s1
    Key: VQAYFKKCNSOZKM-IOSLPCCCSA-N
  • n2c1c(ncnc1NC)n(c2)[C@@H]3O[C@@H]([C@@H](O)[C@H]3O)CO
Properties
C11H15N5O4
Molar mass 281.272 g·mol−1
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

N6-Methyladenosine (m6A) was originally identified and partially characterised in the 1970s, [1] [2] [3] [4] and is an abundant modification in mRNA and DNA. [5] It is found within some viruses, [4] [3] [6] [7] and most eukaryotes including mammals, [2] [1] [8] [9] insects, [10] plants [11] [12] [13] and yeast. [14] [15] It is also found in tRNA, rRNA, and small nuclear RNA (snRNA) as well as several long non-coding RNA, such as Xist . [16] [17]

Contents

The methylation of adenosine is directed by a large m6A methyltransferase complex containing METTL3, which is the subunit that binds S-adenosyl-L-methionine (SAM). [18] In vitro, this methyltransferase complex preferentially methylates RNA oligonucleotides containing GGACU [19] and a similar preference was identified in vivo in mapped m6A sites in Rous sarcoma virus genomic RNA [20] and in bovine prolactin mRNA. [21] More recent studies have characterized other key components of the m6A methyltransferase complex in mammals, including METTL14, [22] [23] Wilms tumor 1 associated protein (WTAP), [22] [24] VIRMA [25] and METTL5. [26] Following a 2010 speculation of m6A in mRNA being dynamic and reversible, [27] the discovery of the first m6A demethylase, fat mass and obesity-associated protein (FTO) in 2011 [28] confirmed this hypothesis and revitalized the interests in the study of m6A. A second m6A demethylase alkB homolog 5 (ALKBH5) was later discovered as well. [29]

The biological functions of m6A are mediated through a group of RNA binding proteins that specifically recognize the methylated adenosine on RNA. These binding proteins are named m6A readers. The YT521-B homology (YTH) domain family of proteins (YTHDF1, YTHDF2, YTHDF3 and YTHDC1) have been characterized as direct m6A readers and have a conserved m6A-binding pocket. [17] [30] [31] [32] [33] Insulin-like growth factor-2 mRNA-binding proteins 1, 2, and 3 (IGF2BP1–3) are reported as a novel class of m6A readers. [34] IGF2BPs use K homology (KH) domains to selectively recognize m6A-containing RNAs and promote their translation and stability. [34] These m6A readers, together with m6A methyltransferases (writers) and demethylases (erasers), establish a complex mechanism of m6A regulation in which writers and erasers determine the distributions of m6A on RNA, whereas readers mediate m6A-dependent functions. m6A has also been shown to mediate a structural switch termed m6A switch. [35]

The specificity of m6A installation on mRNA is controlled by exon architecture and exon junction complexes. Exon junction complexes suppress m6A methylation near exon-exon junctions by packaging nearby RNA and protecting it from methylation by the m6A methyltransferase complex. m6A regions in long internal and terminal exons, away from exon-exon junctions and exon junction complexes, escape suppression and can be methylated by the methyltransferase complex. [36]

Species distribution

Yeast

In budding yeast (Saccharomyces cerevisiae), the expression of the homologue of METTL3, IME4, is induced in diploid cells in response to nitrogen and fermentable carbon source starvation and is required for mRNA methylation and the initiation of correct meiosis and sporulation. [14] [15] mRNAs of IME1 and IME2, key early regulators of meiosis, are known to be targets for methylation, as are transcripts of IME4 itself. [15]

Plants

In plants, the majority of the m6A is found within 150 nucleotides before the start of the poly(A) tail. [37]

Mutations of MTA, the Arabidopsis thaliana homologue of METTL3, results in embryo arrest at the globular stage. A >90% reduction of m6A levels in mature plants leads to dramatically altered growth patterns and floral homeotic abnormalities. [37]

Mammals

Mapping of m6A in human and mouse RNA has identified over 18,000 m6A sites in the transcripts of more than 7,000 human genes with a consensus sequence of [G/A/U][G>A]m6AC[U>A/C] [16] [17] [38] consistent with the previously identified motif. The localization of individual m6A sites in many mRNAs is highly similar between human and mouse, [16] [17] and transcriptome-wide analysis reveals that m6A is found in regions of high evolutionary conservation. [16] m6A is found within long internal exons and is preferentially enriched within 3' UTRs and around stop codons. m6A within 3' UTRs is also associated with the presence of microRNA binding sites; roughly 2/3 of the mRNAs which contain an m6A site within their 3' UTR also have at least one microRNA binding site. [16] By integrating all m6A sequencing data, a novel database called RMBase has identified and provided ~200,000 sites in the human and mouse genomes corresponding to N6-Methyladenosines (m6A) in RNA. [38]

Precise m6A mapping by m6A-CLIP/IP [39] (briefly m6A-CLIP) revealed that a majority of m6A locates in the last exon of mRNAs in multiple tissues/cultured cells of mouse and human, [39] and the m6A enrichment around stop codons is a coincidence that many stop codons locate round the start of last exons where m6A is truly enriched. [39] The major presence of m6A in last exon (>=70%) allows the potential for 3'UTR regulation, including alternative polyadenylation. [39] The study combining m6A-CLIP with rigorous cell fractionation biochemistry reveals that m6A mRNA modifications are deposited in nascent pre-mRNA and are not required for splicing but do specify cytoplasmic turnover. [40] [41]

m6A is susceptible to dynamic regulation both throughout development and in response to cellular stimuli. Analysis of m6A in mouse brain RNA reveals that m6A levels are low during embryonic development and increase dramatically by adulthood. [16] In mESCs and during mouse development, FTO has been shown to mediated LINE1 RNA m6A demethylation and consequently affect local chromatin state and nearby gene transcription. [42] Additionally, silencing the m6A methyltransferase significantly affects gene expression and alternative RNA splicing patterns, resulting in modulation of the p53 (also known as TP53) signalling pathway and apoptosis. [17]

m6A is also found on the RNA components of R-loops in human cells, where it is involved in regulation of stability of RNA:DNA hybrids. [43]

The importance of m6A methylation for physiological processes was recently demonstrated. Inhibition of m6A methylation via pharmacological inhibition of cellular methylations or more specifically by siRNA-mediated silencing of the m6A methylase Mettl3 led to the elongation of the circadian period. In contrast, overexpression of Mettl3 led to a shorter period. The mammalian circadian clock, composed of a transcription feedback loop tightly regulated to oscillate with a period of about 24 hours, is therefore extremely sensitive to perturbations in m6A-dependent RNA processing, likely due to the presence of m6A sites within clock gene transcripts. [44] [45] The effects of global methylation inhibition on the circadian period in mouse cells can be prevented by ectopic expression of an enzyme from the bacterial methyl metabolism. Mouse cells expressing this bacterial protein were resistant to pharmacological inhibition of methyl metabolism, showing no decrease in mRNA m6A methylation or protein methylation. [46]

Clinical significance

Considering the versatile functions of m6A in various physiological processes, it is thus not surprising to find links between m6A and numerous human diseases; many originated from mutations or single nucleotide polymorphisms (SNPs) of cognate factors of m6A. The linkages between m6A and numerous cancer types have been indicated in reports that include stomach cancer, prostate cancer, breast cancer, pancreatic cancer, kidney cancer, mesothelioma, sarcoma, and leukaemia. [47] [48] [49] [50] [51] [52] [53] [54] [55] [56] [57] [58] The impacts of m6A on cancer cell proliferation might be much more profound with more data emerging. The depletion of METTL3 is known to cause apoptosis of cancer cells and reduce invasiveness of cancer cells, [59] [60] while the activation of ALKBH5 by hypoxia was shown to cause cancer stem cell enrichment. [61] m6A has also been indicated in the regulation of energy homeostasis and obesity, as FTO is a key regulatory gene for energy metabolism and obesity. SNPs of FTO have been shown to associate with body mass index in human populations and occurrence of obesity and diabetes. [62] [63] [64] [65] [66] The influence of FTO on pre-adipocyte differentiation has been suggested. [67] [68] [69] The connection between m6A and neuronal disorders has also been studied. For instance, neurodegenerative diseases may be affected by m6A as the cognate dopamine signalling was shown to be dependent on FTO and correct m6A methylation on key signalling transcripts. [70] The mutations in HNRNPA2B1, a potential reader of m6A, have been known to cause neurodegeneration. [71] The IGF2BP1–3, a novel class of m6A reader, has oncogenic functions. IGF2BP1–3 knockdown or knockout decreased MYC protein expression, cell proliferation and colony formation in human cancer cell lines. [34] The ZC3H13, a member of the m6A methyltransferase complex, markedly inhibited colorectal cancer cells growth when knocked down. [72]

Additionally, m6A has been reported to impact viral infections. Many RNA viruses including SV40, adenovirus, herpes virus, Rous sarcoma virus, and influenza virus have been known to contain internal m6A methylation on virus genomic RNA. [73] Several more recent studies have revealed that m6A regulators govern the efficiency of infection, replication, translation and transport of RNA viruses such as human immunodeficiency virus (HIV), hepatitis B virus (HBV), hepatitis C virus (HCV), and Zika virus (ZIKV). [74] [75] [76] [77] [78] [79] These results suggest m6A and its cognate factors play crucial roles in regulating virus life cycles and host-viral interactions.

Aside from affecting viruses themselves, m6A modifications can also disrupt the innate immune response. For example, in HBV, m6A modifications were shown to disrupt the recognition of viruses by RIG-1, a pattern recognition receptor in the immune system. Modifications can also disrupt downstream signaling pathways via mechanisms including ubiquitination and changes in the levels of protein expression. [79]

In bacteria

M6A methylation is also widespread in bacteria, influencing functions such as DNA replication, repair, and gene expression, and prokaryotic defense.

In replication, M6A modifications mark DNA regions where the initiation stage takes place as well as regulates precise timing via the Dam methyltransferase in E. coli. [80] [81] Another enzyme, Dam DNA methylase regulates mismatch repair using M6A modifications which influence other repair proteins by recognizing specific mismatches. [82]

In some cases of DNA protection, M6A methylations (along with M4C modifications) play a role in the protection of bacterial DNA by influencing certain endonucleases via the restriction-modification system, decreasing the influence of bacteriophages. One such role is introducing a methyltransferase which recognizes the same target site that restriction enzymes (Type 1 restriction enzymes) attack and modifying it in order to stop such enzymes from attacking bacteria DNA. [83] [84]

In Development

m6A modifications, along with other epigenetic changes, have been shown to play important roles during eukaryotic development. Hematopoietic Stem Cells (HSCs), Neuronal Stem Cells (NSCs) and Primordial Germ Cells (PCGs) have all been shown to undergo m6A modifications during growth and differentiation. Depending on the stage of development, modifications to HSCs can either promote or inhibit stem cell differentiation by affecting the epithelial-to-hemopoietic transition via METTL3 inhibition or depletion. m6A modifications to NSCs can causes changes in brain size, neuron formation, long-term memory, and learning ability. These changes are often caused by inhibition of either METTL or YTHDF readers and writers. In the reproductive system, m6A modifications have been shown to disrupt the maternal-to-zygotic mRNA transition and negatively affect both gamete formation and fertility. Similar to NSCs, inhibition of the METTL and YTHDF families of proteins is often a catalyst for these changes. [85]

Related Research Articles

<span class="mw-page-title-main">Epigenetics</span> Study of DNA modifications that do not change its sequence

In biology, epigenetics is the study of heritable traits, or a stable change of cell function, that happen without changes to the DNA sequence. The Greek prefix epi- in epigenetics implies features that are "on top of" or "in addition to" the traditional genetic mechanism of inheritance. Epigenetics usually involves a change that is not erased by cell division, and affects the regulation of gene expression. Such effects on cellular and physiological phenotypic traits may result from environmental factors, or be part of normal development. They can lead to cancer.

Methylation, in the chemical sciences, is the addition of a methyl group on a substrate, or the substitution of an atom by a methyl group. Methylation is a form of alkylation, with a methyl group replacing a hydrogen atom. These terms are commonly used in chemistry, biochemistry, soil science, and biology.

<span class="mw-page-title-main">DNA methyltransferase</span> Class of enzymes

In biochemistry, the DNA methyltransferase family of enzymes catalyze the transfer of a methyl group to DNA. DNA methylation serves a wide variety of biological functions. All the known DNA methyltransferases use S-adenosyl methionine (SAM) as the methyl donor.

<span class="mw-page-title-main">DNA methylation</span> Biological process

DNA methylation is a biological process by which methyl groups are added to the DNA molecule. Methylation can change the activity of a DNA segment without changing the sequence. When located in a gene promoter, DNA methylation typically acts to repress gene transcription. In mammals, DNA methylation is essential for normal development and is associated with a number of key processes including genomic imprinting, X-chromosome inactivation, repression of transposable elements, aging, and carcinogenesis.

<span class="mw-page-title-main">RNA editing</span> Molecular process

RNA editing is a molecular process through which some cells can make discrete changes to specific nucleotide sequences within an RNA molecule after it has been generated by RNA polymerase. It occurs in all living organisms and is one of the most evolutionarily conserved properties of RNAs. RNA editing may include the insertion, deletion, and base substitution of nucleotides within the RNA molecule. RNA editing is relatively rare, with common forms of RNA processing not usually considered as editing. It can affect the activity, localization as well as stability of RNAs, and has been linked with human diseases.

<span class="mw-page-title-main">Methyltransferase</span> Group of methylating enzymes

Methyltransferases are a large group of enzymes that all methylate their substrates but can be split into several subclasses based on their structural features. The most common class of methyltransferases is class I, all of which contain a Rossmann fold for binding S-Adenosyl methionine (SAM). Class II methyltransferases contain a SET domain, which are exemplified by SET domain histone methyltransferases, and class III methyltransferases, which are membrane associated. Methyltransferases can also be grouped as different types utilizing different substrates in methyl transfer reactions. These types include protein methyltransferases, DNA/RNA methyltransferases, natural product methyltransferases, and non-SAM dependent methyltransferases. SAM is the classical methyl donor for methyltransferases, however, examples of other methyl donors are seen in nature. The general mechanism for methyl transfer is a SN2-like nucleophilic attack where the methionine sulfur serves as the leaving group and the methyl group attached to it acts as the electrophile that transfers the methyl group to the enzyme substrate. SAM is converted to S-Adenosyl homocysteine (SAH) during this process. The breaking of the SAM-methyl bond and the formation of the substrate-methyl bond happen nearly simultaneously. These enzymatic reactions are found in many pathways and are implicated in genetic diseases, cancer, and metabolic diseases. Another type of methyl transfer is the radical S-Adenosyl methionine (SAM) which is the methylation of unactivated carbon atoms in primary metabolites, proteins, lipids, and RNA.

AlkB (Alkylation B) is a protein found in E. coli, induced during an adaptive response and involved in the direct reversal of alkylation damage. AlkB specifically removes alkylation damage to single stranded (SS) DNA caused by SN2 type of chemical agents. It efficiently removes methyl groups from 1-methyl adenines, 3-methyl cytosines in SS DNA. AlkB is an alpha-ketoglutarate-dependent hydroxylase, a superfamily non-haem iron-containing proteins. It oxidatively demethylates the DNA substrate. Demethylation by AlkB is accompanied with release of CO2, succinate, and formaldehyde.

<span class="mw-page-title-main">FTO gene</span> Protein-coding gene in the species Homo sapiens

Fat mass and obesity-associated protein also known as alpha-ketoglutarate-dependent dioxygenase FTO is an enzyme that in humans is encoded by the FTO gene located on chromosome 16. As one homolog in the AlkB family proteins, it is the first mRNA demethylase that has been identified. Certain alleles of the FTO gene appear to be correlated with obesity in humans.

<span class="mw-page-title-main">DNA (cytosine-5)-methyltransferase 3A</span> Protein-coding gene in the species Homo sapiens

DNA (cytosine-5)-methyltransferase 3A (DNMT3A) is an enzyme that catalyzes the transfer of methyl groups to specific CpG structures in DNA, a process called DNA methylation. The enzyme is encoded in humans by the DNMT3A gene.

<span class="mw-page-title-main">RBM15</span> Protein-coding gene in the species Homo sapiens

Putative RNA-binding protein 15 is a protein that in humans is encoded by the RBM15 gene. It is an RNA-binding protein that acts as a key regulator of N6-Methyladenosine (m6A) methylation of RNAs

<span class="mw-page-title-main">METTL3</span> Gene encoding part of N6-adenosine-methyltransferase

N6-adenosine-methyltransferase 70 kDa subunit (METTL3) is an enzyme that in humans is encoded by the METTL3 gene. METTL3 is located on the human chromosome 14q11.2 and out of the METTL protein family, it is the most studied.

<span class="mw-page-title-main">HBx</span>

HBx is a hepatitis B viral protein. It is 154 amino acids long and interferes with transcription, signal transduction, cell cycle progress, protein degradation, apoptosis and chromosomal stability in the host. It forms a heterodimeric complex with its cellular target protein, and this interaction dysregulates centrosome dynamics and mitotic spindle formation. It interacts with DDB1 redirecting the ubiquitin ligase activity of the CUL4-DDB1 E3 complexes, which are intimately involved in the intracellular regulation of DNA replication and repair, transcription and signal transduction.

High-throughput sequencing of RNA isolated by crosslinking immunoprecipitation (HITS-CLIP) is a variant of CLIP for genome-wide mapping protein–RNA binding sites or RNA modification sites in vivo. HITS-CLIP was originally used to generate genome-wide protein-RNA interaction maps for the neuron-specific RNA-binding protein and splicing factor NOVA1 and NOVA2; since then a number of other splicing factor maps have been generated, including those for PTB, RbFox2, SFRS1, hnRNP C, and even N6-Methyladenosine (m6A) mRNA modifications.

<span class="mw-page-title-main">Cancer epigenetics</span> Field of study in cancer research

Cancer epigenetics is the study of epigenetic modifications to the DNA of cancer cells that do not involve a change in the nucleotide sequence, but instead involve a change in the way the genetic code is expressed. Epigenetic mechanisms are necessary to maintain normal sequences of tissue specific gene expression and are crucial for normal development. They may be just as important, if not even more important, than genetic mutations in a cell's transformation to cancer. The disturbance of epigenetic processes in cancers, can lead to a loss of expression of genes that occurs about 10 times more frequently by transcription silencing than by mutations. As Vogelstein et al. points out, in a colorectal cancer there are usually about 3 to 6 driver mutations and 33 to 66 hitchhiker or passenger mutations. However, in colon tumors compared to adjacent normal-appearing colonic mucosa, there are about 600 to 800 heavily methylated CpG islands in the promoters of genes in the tumors while these CpG islands are not methylated in the adjacent mucosa. Manipulation of epigenetic alterations holds great promise for cancer prevention, detection, and therapy. In different types of cancer, a variety of epigenetic mechanisms can be perturbed, such as the silencing of tumor suppressor genes and activation of oncogenes by altered CpG island methylation patterns, histone modifications, and dysregulation of DNA binding proteins. There are several medications which have epigenetic impact, that are now used in a number of these diseases.

In molecular biology, the protein domain YTH refers to a member of the YTH family that has been shown to selectively remove transcripts of meiosis-specific genes expressed in mitotic cells.They also play a role in the epitranscriptome as reader proteins for m6A.

Within the field of molecular biology, the epitranscriptome includes all the biochemical modifications of the RNA within a cell. In analogy to epigenetics that describes "functionally relevant changes to the genome that do not involve a change in the nucleotide sequence", epitranscriptomics involves all functionally relevant changes to the transcriptome that do not involve a change in the ribonucleotide sequence. Thus, the epitranscriptome can be defined as the ensemble of such functionally relevant changes.

<span class="mw-page-title-main">Chuan He</span> Chinese-American chemical biologist

Chuan He is a Chinese-American chemical biologist. He currently serves as the John T. Wilson Distinguished Service Professor at the University of Chicago, and an Investigator of the Howard Hughes Medical Institute. He is best known for his work in discovering and deciphering reversible RNA methylation in post-transcriptional gene expression regulation. He was awarded the 2023 Wolf Prize in Chemistry for his work in discovering and deciphering reversible RNA methylation in post-transcriptional gene expression regulation in addition to his contributions to the invention of TAB-seq, a biochemical method that can map 5-hydroxymethylcytosine (5hmC) at base-resolution genome-wide, as well as hmC-Seal, a method that covalently labels 5hmC for its detection and profiling.

<span class="mw-page-title-main">Epitranscriptomic sequencing</span>

In epitranscriptomic sequencing, most methods focus on either (1) enrichment and purification of the modified RNA molecules before running on the RNA sequencer, or (2) improving or modifying bioinformatics analysis pipelines to call the modification peaks. Most methods have been adapted and optimized for mRNA molecules, except for modified bisulfite sequencing for profiling 5-methylcytidine which was optimized for tRNAs and rRNAs.

HSV epigenetics is the epigenetic modification of herpes simplex virus (HSV) genetic code.

The viral epitranscriptome includes all modifications to viral transcripts, studied by viral epitranscriptomics. Like the more general epitranscriptome, these modifications do not affect the sequence of the transcript, but rather have consequences on subsequent structures and functions.

References

  1. 1 2 Adams JM, Cory S (May 1975). "Modified nucleosides and bizarre 5'-termini in mouse myeloma mRNA". Nature. 255 (5503): 28–33. Bibcode:1975Natur.255...28A. doi: 10.1038/255028a0 . PMID   1128665. S2CID   4199864.
  2. 1 2 Desrosiers R, Friderici K, Rottman F (October 1974). "Identification of methylated nucleosides in messenger RNA from Novikoff hepatoma cells". Proceedings of the National Academy of Sciences of the United States of America. 71 (10): 3971–3975. Bibcode:1974PNAS...71.3971D. doi: 10.1073/pnas.71.10.3971 . PMC   434308 . PMID   4372599.
  3. 1 2 Aloni Y, Dhar R, Khoury G (October 1979). "Methylation of nuclear simian virus 40 RNAs". Journal of Virology. 32 (1): 52–60. doi:10.1128/JVI.32.1.52-60.1979. PMC   353526 . PMID   232187.
  4. 1 2 Beemon K, Keith J (June 1977). "Localization of N6-methyladenosine in the Rous sarcoma virus genome". Journal of Molecular Biology. 113 (1): 165–179. doi:10.1016/0022-2836(77)90047-X. PMID   196091.
  5. Ji P, Wang X, Xie N, Li Y (2018). "N6-Methyladenosine in RNA and DNA: An Epitranscriptomic and Epigenetic Player Implicated in Determination of Stem Cell Fate". Stem Cells International. 2018: 3256524. doi: 10.1155/2018/3256524 . PMC   6199872 . PMID   30405719.
  6. Courtney DG, Kennedy EM, Dumm RE, Bogerd HP, Tsai K, Heaton NS, Cullen BR (September 2017). "Epitranscriptomic Enhancement of Influenza A Virus Gene Expression and Replication". Cell Host & Microbe. 22 (3): 377–386.e5. doi:10.1016/j.chom.2017.08.004. PMC   5615858 . PMID   28910636.
  7. Gokhale NS, McIntyre AB, McFadden MJ, Roder AE, Kennedy EM, Gandara JA, et al. (November 2016). "N6-Methyladenosine in Flaviviridae Viral RNA Genomes Regulates Infection". Cell Host & Microbe. 20 (5): 654–665. doi:10.1016/j.chom.2016.09.015. PMC   5123813 . PMID   27773535.
  8. Wei CM, Gershowitz A, Moss B (January 1976). "5'-Terminal and internal methylated nucleotide sequences in HeLa cell mRNA". Biochemistry. 15 (2): 397–401. doi:10.1021/bi00647a024. PMID   174715.
  9. Perry RP, Kelley DE, Friderici K, Rottman F (April 1975). "The methylated constituents of L cell messenger RNA: evidence for an unusual cluster at the 5' terminus". Cell. 4 (4): 387–394. doi: 10.1016/0092-8674(75)90159-2 . PMID   1168101.
  10. Levis R, Penman S (April 1978). "5'-terminal structures of poly(A)+ cytoplasmic messenger RNA and of poly(A)+ and poly(A)- heterogeneous nuclear RNA of cells of the dipteran Drosophila melanogaster". Journal of Molecular Biology. 120 (4): 487–515. doi:10.1016/0022-2836(78)90350-9. PMID   418182.
  11. Nichols JL (1979). "In maize poly(A)-containing RNA". Plant Science Letters. 15 (4): 357–361. doi:10.1016/0304-4211(79)90141-X.
  12. Kennedy TD, Lane BG (June 1979). "Wheat embryo ribonucleates. XIII. Methyl-substituted nucleoside constituents and 5'-terminal dinucleotide sequences in bulk poly(AR)-rich RNA from imbibing wheat embryos". Canadian Journal of Biochemistry. 57 (6): 927–931. doi:10.1139/o79-112. PMID   476526.
  13. Zhong S, Li H, Bodi Z, Button J, Vespa L, Herzog M, Fray RG (May 2008). "MTA is an Arabidopsis messenger RNA adenosine methylase and interacts with a homolog of a sex-specific splicing factor". The Plant Cell. 20 (5): 1278–1288. doi:10.1105/tpc.108.058883. PMC   2438467 . PMID   18505803.
  14. 1 2 Clancy MJ, Shambaugh ME, Timpte CS, Bokar JA (October 2002). "Induction of sporulation in Saccharomyces cerevisiae leads to the formation of N6-methyladenosine in mRNA: a potential mechanism for the activity of the IME4 gene". Nucleic Acids Research. 30 (20): 4509–4518. doi:10.1093/nar/gkf573. PMC   137137 . PMID   12384598.
  15. 1 2 3 Bodi Z, Button JD, Grierson D, Fray RG (September 2010). "Yeast targets for mRNA methylation". Nucleic Acids Research. 38 (16): 5327–5335. doi:10.1093/nar/gkq266. PMC   2938207 . PMID   20421205.
  16. 1 2 3 4 5 6 Meyer KD, Saletore Y, Zumbo P, Elemento O, Mason CE, Jaffrey SR (June 2012). "Comprehensive analysis of mRNA methylation reveals enrichment in 3' UTRs and near stop codons". Cell. 149 (7): 1635–1646. doi:10.1016/j.cell.2012.05.003. PMC   3383396 . PMID   22608085.
  17. 1 2 3 4 5 Dominissini D, Moshitch-Moshkovitz S, Schwartz S, Salmon-Divon M, Ungar L, Osenberg S, et al. (April 2012). "Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq". Nature. 485 (7397): 201–206. Bibcode:2012Natur.485..201D. doi:10.1038/nature11112. PMID   22575960. S2CID   3517716.
  18. Bokar JA, Shambaugh ME, Polayes D, Matera AG, Rottman FM (November 1997). "Purification and cDNA cloning of the AdoMet-binding subunit of the human mRNA (N6-adenosine)-methyltransferase". RNA. 3 (11): 1233–1247. PMC   1369564 . PMID   9409616.
  19. Harper JE, Miceli SM, Roberts RJ, Manley JL (October 1990). "Sequence specificity of the human mRNA N6-adenosine methylase in vitro". Nucleic Acids Research. 18 (19): 5735–5741. doi:10.1093/nar/18.19.5735. PMC   332308 . PMID   2216767.
  20. Kane SE, Beemon K (September 1985). "Precise localization of m6A in Rous sarcoma virus RNA reveals clustering of methylation sites: implications for RNA processing". Molecular and Cellular Biology. 5 (9): 2298–2306. doi:10.1128/mcb.5.9.2298. PMC   366956 . PMID   3016525.
  21. Horowitz S, Horowitz A, Nilsen TW, Munns TW, Rottman FM (September 1984). "Mapping of N6-methyladenosine residues in bovine prolactin mRNA". Proceedings of the National Academy of Sciences of the United States of America. 81 (18): 5667–5671. Bibcode:1984PNAS...81.5667H. doi: 10.1073/pnas.81.18.5667 . PMC   391771 . PMID   6592581.
  22. 1 2 Liu J, Yue Y, Han D, Wang X, Fu Y, Zhang L, et al. (February 2014). "A METTL3-METTL14 complex mediates mammalian nuclear RNA N6-adenosine methylation". Nature Chemical Biology. 10 (2): 93–95. doi:10.1038/nchembio.1432. PMC   3911877 . PMID   24316715.
  23. Wang Y, Li Y, Toth JI, Petroski MD, Zhang Z, Zhao JC (February 2014). "N6-methyladenosine modification destabilizes developmental regulators in embryonic stem cells". Nature Cell Biology. 16 (2): 191–198. doi:10.1038/ncb2902. PMC   4640932 . PMID   24394384.
  24. Ping XL, Sun BF, Wang L, Xiao W, Yang X, Wang WJ, et al. (February 2014). "Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase". Cell Research. 24 (2): 177–189. doi:10.1038/cr.2014.3. PMC   3915904 . PMID   24407421.
  25. Schwartz S, Mumbach MR, Jovanovic M, Wang T, Maciag K, Bushkin GG, et al. (July 2014). "Perturbation of m6A writers reveals two distinct classes of mRNA methylation at internal and 5' sites". Cell Reports. 8 (1): 284–296. doi:10.1016/j.celrep.2014.05.048. PMC   4142486 . PMID   24981863.
  26. van Tran N, Ernst FG, Hawley BR, Zorbas C, Ulryck N, Hackert P, et al. (September 2019). "The human 18S rRNA m6A methyltransferase METTL5 is stabilized by TRMT112". Nucleic Acids Research. 47 (15): 7719–7733. doi:10.1093/nar/gkz619. PMC   6735865 . PMID   31328227.
  27. He C (December 2010). "Grand challenge commentary: RNA epigenetics?". Nature Chemical Biology. 6 (12): 863–865. doi:10.1038/nchembio.482. PMID   21079590.
  28. Jia G, Fu Y, Zhao X, Dai Q, Zheng G, Yang Y, et al. (October 2011). "N6-methyladenosine in nuclear RNA is a major substrate of the obesity-associated FTO". Nature Chemical Biology. 7 (12): 885–887. doi:10.1038/nchembio.687. PMC   3218240 . PMID   22002720.
  29. Zheng G, Dahl JA, Niu Y, Fedorcsak P, Huang CM, Li CJ, et al. (January 2013). "ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility". Molecular Cell. 49 (1): 18–29. doi:10.1016/j.molcel.2012.10.015. PMC   3646334 . PMID   23177736.
  30. Wang X, Lu Z, Gomez A, Hon GC, Yue Y, Han D, et al. (January 2014). "N6-methyladenosine-dependent regulation of messenger RNA stability". Nature. 505 (7481): 117–120. Bibcode:2014Natur.505..117W. doi:10.1038/nature12730. PMC   3877715 . PMID   24284625.
  31. Wang X, Zhao BS, Roundtree IA, Lu Z, Han D, Ma H, et al. (June 2015). "N(6)-methyladenosine Modulates Messenger RNA Translation Efficiency". Cell. 161 (6): 1388–1399. doi:10.1016/j.cell.2015.05.014. PMC   4825696 . PMID   26046440.
  32. Xu C, Wang X, Liu K, Roundtree IA, Tempel W, Li Y, et al. (November 2014). "Structural basis for selective binding of m6A RNA by the YTHDC1 YTH domain". Nature Chemical Biology. 10 (11): 927–929. doi:10.1038/nchembio.1654. PMID   25242552.
  33. Xiao W, Adhikari S, Dahal U, Chen YS, Hao YJ, Sun BF, et al. (February 2016). "Nuclear m(6)A Reader YTHDC1 Regulates mRNA Splicing". Molecular Cell. 61 (4): 507–519. doi: 10.1016/j.molcel.2016.01.012 . PMID   26876937.
  34. 1 2 3 Huang H, Weng H, Sun W, Qin X, Shi H, Wu H, et al. (March 2018). "Recognition of RNA N6-methyladenosine by IGF2BP proteins enhances mRNA stability and translation". Nature Cell Biology. 20 (3): 285–295. doi:10.1038/s41556-018-0045-z. PMC   5826585 . PMID   29476152.
  35. Liu N, Dai Q, Zheng G, He C, Parisien M, Pan T (February 2015). "N(6)-methyladenosine-dependent RNA structural switches regulate RNA-protein interactions". Nature. 518 (7540): 560–564. Bibcode:2015Natur.518..560L. doi:10.1038/nature14234. PMC   4355918 . PMID   25719671.
  36. He PC, Wei J, Dou X, Harada BT, Zhang Z, Ge R, et al. (February 2023). "Exon architecture controls mRNA m6A suppression and gene expression". Science. 379 (6633): 677–682. Bibcode:2023Sci...379..677H. doi:10.1126/science.abj9090. PMC   9990141 . PMID   36705538.
  37. 1 2 Bodi Z, Zhong S, Mehra S, Song J, Graham N, Li H, et al. (2012). "Adenosine Methylation in Arabidopsis mRNA is Associated with the 3' End and Reduced Levels Cause Developmental Defects". Frontiers in Plant Science. 3: 48. doi: 10.3389/fpls.2012.00048 . PMC   3355605 . PMID   22639649.
  38. 1 2 Sun WJ, Li JH, Liu S, Wu J, Zhou H, Qu LH, Yang JH (January 2016). "RMBase: a resource for decoding the landscape of RNA modifications from high-throughput sequencing data". Nucleic Acids Research. 44 (D1): D259–D265. doi:10.1093/nar/gkv1036. PMC   4702777 . PMID   26464443.
  39. 1 2 3 4 Ke S, Alemu EA, Mertens C, Gantman EC, Fak JJ, Mele A, et al. (October 2015). "A majority of m6A residues are in the last exons, allowing the potential for 3' UTR regulation". Genes & Development. 29 (19): 2037–2053. doi:10.1101/gad.269415.115. PMC   4604345 . PMID   26404942.
  40. Ke S, Pandya-Jones A, Saito Y, Fak JJ, Vågbø CB, Geula S, et al. (May 2017). "m6A mRNA modifications are deposited in nascent pre-mRNA and are not required for splicing but do specify cytoplasmic turnover". Genes & Development. 31 (10): 990–1006. doi:10.1101/gad.301036.117. PMC   5495127 . PMID   28637692.
  41. Rosa-Mercado NA, Withers JB, Steitz JA (May 2017). "Settling the m6A debate: methylation of mature mRNA is not dynamic but accelerates turnover". Genes & Development. 31 (10): 957–958. doi:10.1101/gad.302695.117. PMC   5495124 . PMID   28637691.
  42. Wei J, Yu X, Yang L, Liu X, Gao B, Huang B, et al. (May 2022). "FTO mediates LINE1 m6A demethylation and chromatin regulation in mESCs and mouse development". Science. 376 (6596): 968–973. Bibcode:2022Sci...376..968W. doi:10.1126/science.abe9582. PMC   9746489 . PMID   35511947.
  43. Abakir A, Giles TC, Cristini A, Foster JM, Dai N, Starczak M, et al. (January 2020). "N6-methyladenosine regulates the stability of RNA:DNA hybrids in human cells". Nature Genetics. 52 (1): 48–55. doi:10.1038/s41588-019-0549-x. PMC   6974403 . PMID   31844323.
  44. Fustin JM, Doi M, Yamaguchi Y, Hida H, Nishimura S, Yoshida M, et al. (November 2013). "RNA-methylation-dependent RNA processing controls the speed of the circadian clock". Cell. 155 (4): 793–806. doi: 10.1016/j.cell.2013.10.026 . PMID   24209618.
  45. Hastings MH (November 2013). "m(6)A mRNA methylation: a new circadian pacesetter". Cell. 155 (4): 740–741. doi: 10.1016/j.cell.2013.10.028 . PMID   24209613.
  46. Fustin JM, Ye S, Rakers C, Kaneko K, Fukumoto K, Yamano M, et al. (May 2020). "Methylation deficiency disrupts biological rhythms from bacteria to humans". Communications Biology. 3 (1): 211. doi: 10.1038/s42003-020-0942-0 . PMC   7203018 . PMID   32376902.
  47. Akilzhanova A, Nurkina Z, Momynaliev K, Ramanculov E, Zhumadilov Z, Rakhypbekov T, et al. (September 2013). "Genetic profile and determinants of homocysteine levels in Kazakhstan patients with breast cancer". Anticancer Research. 33 (9): 4049–4059. PMID   24023349.
  48. Reddy SM, Sadim M, Li J, Yi N, Agarwal S, Mantzoros CS, Kaklamani VG (August 2013). "Clinical and genetic predictors of weight gain in patients diagnosed with breast cancer". British Journal of Cancer. 109 (4): 872–881. doi:10.1038/bjc.2013.441. PMC   3749587 . PMID   23922112.
  49. Heiliger KJ, Hess J, Vitagliano D, Salerno P, Braselmann H, Salvatore G, et al. (June 2012). "Novel candidate genes of thyroid tumourigenesis identified in Trk-T1 transgenic mice". Endocrine-Related Cancer. 19 (3): 409–421. doi: 10.1530/ERC-11-0387 . PMID   22454401.
  50. Ortega A, Niksic M, Bachi A, Wilm M, Sánchez L, Hastie N, Valcárcel J (January 2003). "Biochemical function of female-lethal (2)D/Wilms' tumor suppressor-1-associated proteins in alternative pre-mRNA splicing". The Journal of Biological Chemistry. 278 (5): 3040–3047. doi: 10.1074/jbc.M210737200 . hdl: 10261/162976 . PMID   12444081.
  51. Jin DI, Lee SW, Han ME, Kim HJ, Seo SA, Hur GY, et al. (December 2012). "Expression and roles of Wilms' tumor 1-associating protein in glioblastoma". Cancer Science. 103 (12): 2102–2109. doi: 10.1111/cas.12022 . PMC   7659328 . PMID   22957919.
  52. Lin Y, Ueda J, Yagyu K, Ishii H, Ueno M, Egawa N, et al. (July 2013). "Association between variations in the fat mass and obesity-associated gene and pancreatic cancer risk: a case-control study in Japan". BMC Cancer. 13: 337. doi: 10.1186/1471-2407-13-337 . PMC   3716552 . PMID   23835106.
  53. Casalegno-Garduño R, Schmitt A, Wang X, Xu X, Schmitt M (October 2010). "Wilms' tumor 1 as a novel target for immunotherapy of leukemia". Transplantation Proceedings. 42 (8): 3309–3311. doi:10.1016/j.transproceed.2010.07.034. PMID   20970678.
  54. Linnebacher M, Wienck A, Boeck I, Klar E (2010-03-18). "Identification of an MSI-H tumor-specific cytotoxic T cell epitope generated by the (-1) frame of U79260(FTO)". Journal of Biomedicine & Biotechnology. 2010: 841451. doi: 10.1155/2010/841451 . PMC   2842904 . PMID   20339516.
  55. Machiela MJ, Lindström S, Allen NE, Haiman CA, Albanes D, Barricarte A, et al. (December 2012). "Association of type 2 diabetes susceptibility variants with advanced prostate cancer risk in the Breast and Prostate Cancer Cohort Consortium". American Journal of Epidemiology. 176 (12): 1121–1129. doi:10.1093/aje/kws191. PMC   3571230 . PMID   23193118.
  56. Long J, Zhang B, Signorello LB, Cai Q, Deming-Halverson S, Shrubsole MJ, et al. (2013-04-08). "Evaluating genome-wide association study-identified breast cancer risk variants in African-American women". PLOS ONE. 8 (4): e58350. Bibcode:2013PLoSO...858350L. doi: 10.1371/journal.pone.0058350 . PMC   3620157 . PMID   23593120.
  57. Kaklamani V, Yi N, Sadim M, Siziopikou K, Zhang K, Xu Y, et al. (April 2011). "The role of the fat mass and obesity associated gene (FTO) in breast cancer risk". BMC Medical Genetics. 12: 52. doi: 10.1186/1471-2350-12-52 . PMC   3089782 . PMID   21489227.
  58. Pierce BL, Austin MA, Ahsan H (June 2011). "Association study of type 2 diabetes genetic susceptibility variants and risk of pancreatic cancer: an analysis of PanScan-I data". Cancer Causes & Control. 22 (6): 877–883. doi:10.1007/s10552-011-9760-5. PMC   7043136 . PMID   21445555.
  59. Bokar JA (2005-01-01). "The biosynthesis and functional roles of methylated nucleosides in eukaryotic mRNA". In Grosjean H (ed.). Fine-Tuning of RNA Functions by Modification and Editing. Topics in Current Genetics. Vol. 12. Springer Berlin Heidelberg. pp. 141–177. doi:10.1007/b106365. ISBN   9783540244950.
  60. Lin S, Choe J, Du P, Triboulet R, Gregory RI (May 2016). "The m(6)A Methyltransferase METTL3 Promotes Translation in Human Cancer Cells". Molecular Cell. 62 (3): 335–345. doi:10.1016/j.molcel.2016.03.021. PMC   4860043 . PMID   27117702.
  61. Zhang C, Samanta D, Lu H, Bullen JW, Zhang H, Chen I, et al. (April 2016). "Hypoxia induces the breast cancer stem cell phenotype by HIF-dependent and ALKBH5-mediated m⁶A-demethylation of NANOG mRNA". Proceedings of the National Academy of Sciences of the United States of America. 113 (14): E2047–E2056. Bibcode:2016PNAS..113E2047Z. doi: 10.1073/pnas.1602883113 . PMC   4833258 . PMID   27001847.
  62. Loos RJ, Yeo GS (January 2014). "The bigger picture of FTO: the first GWAS-identified obesity gene". Nature Reviews. Endocrinology. 10 (1): 51–61. doi:10.1038/nrendo.2013.227. PMC   4188449 . PMID   24247219.
  63. Frayling TM, Timpson NJ, Weedon MN, Zeggini E, Freathy RM, Lindgren CM, et al. (May 2007). "A common variant in the FTO gene is associated with body mass index and predisposes to childhood and adult obesity". Science. 316 (5826): 889–894. Bibcode:2007Sci...316..889F. doi:10.1126/science.1141634. PMC   2646098 . PMID   17434869.
  64. Wang L, Yu Q, Xiong Y, Liu L, Zhang X, Zhang Z, et al. (2013). "Variant rs1421085 in the FTO gene contribute childhood obesity in Chinese children aged 3-6 years". Obesity Research & Clinical Practice. 7 (1): e14–e22. doi:10.1016/j.orcp.2011.12.007. PMID   24331679.
  65. Kalnina I, Zaharenko L, Vaivade I, Rovite V, Nikitina-Zake L, Peculis R, et al. (September 2013). "Polymorphisms in FTO and near TMEM18 associate with type 2 diabetes and predispose to younger age at diagnosis of diabetes". Gene. 527 (2): 462–468. doi:10.1016/j.gene.2013.06.079. PMID   23860325.
  66. Karra E, O'Daly OG, Choudhury AI, Yousseif A, Millership S, Neary MT, et al. (August 2013). "A link between FTO, ghrelin, and impaired brain food-cue responsivity". The Journal of Clinical Investigation. 123 (8): 3539–3551. doi:10.1172/jci44403. PMC   3726147 . PMID   23867619.
  67. Zhao X, Yang Y, Sun BF, Shi Y, Yang X, Xiao W, et al. (December 2014). "FTO-dependent demethylation of N6-methyladenosine regulates mRNA splicing and is required for adipogenesis". Cell Research. 24 (12): 1403–1419. doi:10.1038/cr.2014.151. PMC   4260349 . PMID   25412662.
  68. Merkestein M, Laber S, McMurray F, Andrew D, Sachse G, Sanderson J, et al. (April 2015). "FTO influences adipogenesis by regulating mitotic clonal expansion". Nature Communications. 6: 6792. Bibcode:2015NatCo...6.6792M. doi:10.1038/ncomms7792. PMC   4410642 . PMID   25881961.
  69. Zhang M, Zhang Y, Ma J, Guo F, Cao Q, Zhang Y, et al. (2015-07-28). "The Demethylase Activity of FTO (Fat Mass and Obesity Associated Protein) Is Required for Preadipocyte Differentiation". PLOS ONE. 10 (7): e0133788. Bibcode:2015PLoSO..1033788Z. doi: 10.1371/journal.pone.0133788 . PMC   4517749 . PMID   26218273.
  70. Hess ME, Hess S, Meyer KD, Verhagen LA, Koch L, Brönneke HS, et al. (August 2013). "The fat mass and obesity associated gene (Fto) regulates activity of the dopaminergic midbrain circuitry". Nature Neuroscience. 16 (8): 1042–1048. doi:10.1038/nn.3449. PMID   23817550. S2CID   11452560.
  71. Kim HJ, Kim NC, Wang YD, Scarborough EA, Moore J, Diaz Z, et al. (March 2013). "Mutations in prion-like domains in hnRNPA2B1 and hnRNPA1 cause multisystem proteinopathy and ALS". Nature. 495 (7442): 467–473. Bibcode:2013Natur.495..467K. doi:10.1038/nature11922. PMC   3756911 . PMID   23455423.
  72. Wang ZL, Li B, Luo YX, Lin Q, Liu SR, Zhang XQ, et al. (January 2018). "Comprehensive Genomic Characterization of RNA-Binding Proteins across Human Cancers". Cell Reports. 22 (1): 286–298. doi: 10.1016/j.celrep.2017.12.035 . PMID   29298429.
  73. Narayan P, Rottman FM (1992). "Methylation of mRNA". In Nord FF (ed.). Advances in Enzymology and Related Areas of Molecular Biology. Advances in Enzymology - and Related Areas of Molecular Biology. Vol. 65. pp. 255–285. doi:10.1002/9780470123119.ch7. ISBN   9780470123119. PMID   1315118.
  74. Kennedy EM, Bogerd HP, Kornepati AV, Kang D, Ghoshal D, Marshall JB, et al. (May 2016). "Posttranscriptional m(6)A Editing of HIV-1 mRNAs Enhances Viral Gene Expression". Cell Host & Microbe. 19 (5): 675–685. doi:10.1016/j.chom.2016.04.002. PMC   4867121 . PMID   27117054.
  75. Tirumuru N, Zhao BS, Lu W, Lu Z, He C, Wu L (July 2016). "N(6)-methyladenosine of HIV-1 RNA regulates viral infection and HIV-1 Gag protein expression". eLife. 5. doi: 10.7554/eLife.15528 . PMC   4961459 . PMID   27371828.
  76. Lichinchi G, Gao S, Saletore Y, Gonzalez GM, Bansal V, Wang Y, et al. (February 2016). "Dynamics of the human and viral m(6)A RNA methylomes during HIV-1 infection of T cells". Nature Microbiology. 1 (4): 16011. doi:10.1038/nmicrobiol.2016.11. PMC   6053355 . PMID   27572442.
  77. Lichinchi G, Zhao BS, Wu Y, Lu Z, Qin Y, He C, Rana TM (November 2016). "Dynamics of Human and Viral RNA Methylation during Zika Virus Infection". Cell Host & Microbe. 20 (5): 666–673. doi:10.1016/j.chom.2016.10.002. PMC   5155635 . PMID   27773536.
  78. Gokhale NS, McIntyre AB, McFadden MJ, Roder AE, Kennedy EM, Gandara JA, et al. (November 2016). "N6-Methyladenosine in Flaviviridae Viral RNA Genomes Regulates Infection". Cell Host & Microbe. 20 (5): 654–665. doi:10.1016/j.chom.2016.09.015. PMC   5123813 . PMID   27773535.
  79. 1 2 Moon, Jae-Su; Lee, Wooseong; Cho, Yong-Hee; Kim, Yonghyo; Kim, Geon-Woo (2024-02-28). "The Significance of N6-Methyladenosine RNA Methylation in Regulating the Hepatitis B Virus Life Cycle". Journal of Microbiology and Biotechnology. 34 (2): 233–239. doi:10.4014/jmb.2309.09013. ISSN   1738-8872. PMC   10940779 . PMID   37942519.
  80. O’Brown, Zach Klapholz; Greer, Eric Lieberman (2016), Jeltsch, Albert; Jurkowska, Renata Z. (eds.), "N6-Methyladenine: A Conserved and Dynamic DNA Mark", DNA Methyltransferases - Role and Function, vol. 945, Cham: Springer International Publishing, pp. 213–246, doi:10.1007/978-3-319-43624-1_10, ISBN   978-3-319-43624-1, PMC   5291743 , PMID   27826841 , retrieved 2024-04-07
  81. Balzarolo, Melania; Engels, Sander; de Jong, Anja J.; Franke, Katka; van den Berg, Timo K.; Gulen, Muhammet F.; Ablasser, Andrea; Janssen, Edith M.; van Steensel, Bas; Wolkers, Monika C. (March 2021). "m6A methylation potentiates cytosolic dsDNA recognition in a sequence-specific manner". Open Biology. 11 (3). doi:10.1098/rsob.210030. ISSN   2046-2441. PMC   8101014 . PMID   33715389.
  82. Raghunathan, Nalini; Goswami, Sayantan; Leela, Jakku K.; Pandiyan, Apuratha; Gowrishankar, Jayaraman (2019). "A new role for Escherichia coli Dam DNA methylase in prevention of aberrant chromosomal replication". Nucleic Acids Research. 47 (11): 5698–5711. doi:10.1093/nar/gkz242. PMC   6582345 . PMID   30957852 . Retrieved 2024-04-07.
  83. Blow, Matthew J.; Clark, Tyson A.; Daum, Chris G.; Deutschbauer, Adam M.; Fomenkov, Alexey; Fries, Roxanne; Froula, Jeff; Kang, Dongwan D.; Malmstrom, Rex R.; Morgan, Richard D.; Posfai, Janos; Singh, Kanwar; Visel, Axel; Wetmore, Kelly; Zhao, Zhiying (2016-02-12). "The Epigenomic Landscape of Prokaryotes". PLOS Genetics. 12 (2): e1005854. doi: 10.1371/journal.pgen.1005854 . ISSN   1553-7404. PMC   4752239 . PMID   26870957.
  84. Loenen, W. A. M.; Dryden, D. T. F.; Raleigh, E. A.; Wilson, G. G. (2014-01-01). "Type I restriction enzymes and their relatives". Nucleic Acids Research. 42 (1): 20–44. doi:10.1093/nar/gkt847. ISSN   0305-1048. PMC   3874165 . PMID   24068554.
  85. Jiang, Xiulin; Liu, Baiyang; Nie, Zhi; Duan, Lincan; Xiong, Qiuxia; Jin, Zhixian; Yang, Cuiping; Chen, Yongbin (2021-02-21). "The role of m6A modification in the biological functions and diseases". Signal Transduction and Targeted Therapy. 6 (1): 74. doi:10.1038/s41392-020-00450-x. ISSN   2059-3635. PMC   7897327 . PMID   33611339.