Abstract
Alternative splicing (AS) serves as an additional regulatory process for gene expression after transcription, and it generates distinct mRNA species, and even noncoding RNAs (ncRNAs), from one primary transcript. Generally, AS can be coupled with transcription and subjected to epigenetic regulation, such as DNA methylation and histone modifications. In addition, ncRNAs, especially long noncoding RNAs (lncRNAs), can be generated from AS and function as splicing factors (“interactors” or “hijackers”) in AS. Recently, RNA modifications, such as the RNA N6-methyladenosine (m6A) modification, have been found to regulate AS. In this review, we summarize recent achievements related to the epigenetic regulation of AS.
Keywords: Alternative splicing, lncRNAs, m6A modification, epigenetic regulation
Introduction
It was said by the director Eisenstein Sergei that the sum of frame A and frame B will produce C, which contains an absolutely new concept in the film field. This concept is “montage”, which is also known as “splicing”. Similarly, in molecular biology, splicing means the editing of precursor messenger RNA (premRNA) transcripts to yield mature messenger RNAs (mRNAs) or noncoding RNAs (ncRNAs). Splicing removes introns and joins exons together through a series of complicated reactions catalyzed by spliceosomes. Generally, RNA splicing includes constitutive splicing (CS) and alternative splicing (AS). AS refers to the process that selectively splices a set of sites in a premRNA to form variable mature mRNAs, thereby producing proteins with different structures and functions [1,2]. This process occurs for ~95% of transcripts and creates diverse proteins by varying exon composition from a single mRNA [2]. AS effectively increases transcriptomic and proteomic diversity and significantly influences many kinds of cellular processes, as well as tissue specificity or development and disease development [3]. In this review, we will summarize recent findings and provide an overview of approximately 40 years of insights related to AS, focusing mainly on the mechanisms of epigenetic regulation of AS.
History of alternative splicing
AS is important for generating proteome complexity from a limited number of genes. The primary transcripts produced from gene transcription undergo AS, with introns being removed and particular exons included or excluded, resulting in the maturation of mRNAs or ncRNAs (Figure 1A). Mainly based on this process, the concept of “one gene-many polypeptides” has been proposed. The generation of ncRNAs from AS was recently reported, further confirming AS as an efficient and economical mode of regulating gene expression.
Figure 1.

Process of alternative splicing and traditional classification of basic types. A. The brief process of AS that results in the production of multiple proteins from a single gene transcript. In detail, after DNA replication begins, particular exons of primary transcribed RNAs may be alternatively included or excluded under some conditions, generating variant isoforms. As a result, the translated proteins contain differences in amino acid sequence or functional structures. B. The five basic patterns of AS, including 1, exon skipping or cassette exon, in which single or multiple exons are skipped; 2, mutually exclusive exons, in which only one of two exons is retained; 3 & 4, alternative donor/acceptor site or 5’/3’ splice junction is used to alter the boundary of exons; and 5, intron retention indicates that a region may be spliced as an intron or simply retained in the mature RNA. Constitutive exons are represented as dark wide blocks, and alternatively spliced exons are represented as light narrow block.
Generally, AS has five discriminative modes that result from a complicated regulation network, including exon skipping (SE, cassette exon), mutually exclusive exons (MXEs), alternative donor sites (A5’SS), alternative acceptor sites (A3’SS), and intron retention (RI) (Figure 1B). SE is the most prevalent mode, with exons included in mRNAs under certain conditions but omitted from mRNAs under different conditions, similar to the differential frame choice between “TRAILER” and “FEATURE” in films. For example, single exon skipping occurs in CAPN3 (E6) [4], Fas (E6) [5] and PXN-AS1 (E4) [6], and multiple exon skipping occurs in exons 8-12 of TGM2 during leukemia tumorigenesis [7]. MXE indicates a process in which two exons are competitively retained in mRNAs, which occurs in some well-known genes, such as PKM1/2 (E9/10) [8], TPM1 (E2a/b) [9] and FGFR2 (EIIIb/c) [10]. A5’SS involves an alternative 5’ splicing junction due to the uncertain splicing of the 3’ boundary of one upstream exon. Both BCL [11,12] and XIST [13], which remarkably regulate cell growth, have two isoforms, including short (BCL-XS and XIST-S) and long (BCL-X and XIST-L) patterns, which are mediated by the A5’SS splicing mode. On the other hand, A3’SS generates an alternative 3’ splice junction by selectively changing the 5’ boundary of one downstream exon; for example, the PNUTS (also known as PPP1R10) preRNA can develop into an mRNA or a lncRNA through A3’SS in Exon 12 [14]. In the uncommon mode of RI, introns may be completely or partially retained in spliced RNAs, such as ARGLU1, which was found to have a retained intron between Exon 2 and Exon 3 of its preRNA [15]. Notably, multiple splicing modes may exist in one preRNA; for example, SE, MXE and A3’/5’SS together contribute to the various isoforms of RUNX1 [16,17].
In the past forty decades, a number of breakthroughs have been made (Figure 2), and these advances will continue to shape the future [3,18-24]. Splicing was primarily discovered in 1977, followed by basic chemistry research to verify the existence of splicing genes [25,26]. Small nuclear ribonucleoproteins (snRNPs) are RNA-protein complexes that combine with premRNA to form spliceosomes, which are key to the AS process. snRNPs were recognized in 1985 and included at least 5 snRNAs and more than 200 proteins. Then, a decade of studies on RNA binding proteins (RBPs), including the SR and hnRNP proteins, arrived. Nuclear speckles and the interchromatin granule distribution pattern of snRNPs were reported in the early 1990s as the foundation for understanding AS and exploring its precise mechanisms in vitro or in vivo. One landmark event was the 1993 Nobel Prize, which was awarded for the discovery of split genes, i.e., the splicing process. Increasingly, due to GFP technology, the dynamic nature of splicing factors and nuclear speckle organization was visualized in living cells, clarifying transcriptional and splicing activity. LncRNAs such as H19, XIST, SRA1, Evf2 and MALAT1 turned out to be closely linked to AS in 2004 [27], with additional research advances emerging. To better understand AS, the systematic structural elucidation of individual spliceosome components and subcomplexes was reported in 2007 [28]. Research on the N6-methyladenosine (m6A) modification emerged in large numbers in the early 2010s and was considered as potential signal for RNA splicing [29], bringing about a new initiative to study AS in more detail. Deciphering the mechanisms of AS is important for understanding the regulatory network in AS and developing applications for therapeutic approaches in the future.
Figure 2.
The history of AS over the past forty decades.
AS coupled to transcription
AS is a dynamic process that occurs simultaneously with and is inseparable from transcription. AS can be affected by the particular sequence or structure of the promoter [30-32]. For example, the C-terminal major domain (CTD) of RNA polymerase II plays a central role in coordinating transcription and AS, and it could assist SR proteins in binding primary transcripts [33,34].
Emerging research has placed AS in a key position in genetic information flow, which indicates that transcription regulators could also subsequently influence AS decisions. For example, TCERG1 is a transcription factor that is associated with the RNA pol II holoenzyme and influences the elongation efficiency of transcribed RNAs, which was found to be a crucial component of spliceosome regulation in AS [35]. Furthermore, Spi-1 tends to interact with spliceosome proteins and modulates AS of RNAs through splice site selection [36]. Moreover, chromatin conformation [32,37,38], the H3K36me3, H3K4me3 and H3K9me3 histone modifications [39,40] and DNA methylation [41-43] go hand in hand with the regulation of Brm [44], MRG15, Pisp1/p52 [45], HP1γ [46] and CTCF/MeCP2 [47], respectively. In detail, the speed bump model shows that nucleosomes can influence transcriptional elongation through hindering RNAPII and promoting exon inclusion, providing evidence that the chromatin structure can reflect underlying recognition of introns and exons. SWI/SNF, a subfamily of chromatin remodelers, not only modulates gene transcription by altering nucleosome organization but also controls AS through interacting with spliceosome components to determine the splicing events of genes such as CD44 [44,48,49].
DNA methylation and histone modifications determine eukaryotic chromatin structures under conditions in which certain histone marks recruit splicing factors through chromatin-binding proteins. For example, H3K36me3 can recruit MRG15 or Psip1 to promote hnRNP or SR splicing factors binding to specific sites of RNAs, respectively, directing the AS process [39,45,50]. DNA methylation frequently occurs in 5’SS and mediates AS through methyl-binding domain proteins (MBDs), such as HP1, which could interact with the histone chaperone complex FACT to bind H3K9me3 containing methylated DNA [51-53].
Regulation of AS by lncRNAs
The spliceosome comprises at least five kinds of snRNAs (U1, U2, U4/U6 and U5) and more than 200 snRNPs as the AS machinery [54]. The functional spliceosome complex is formed step by step and orchestrates with GU and AG dinucleotides. Splicing factors play a significant role in regulating the interaction between AS and exon splicing enhancer (ESE)/silencer (ESS) or intron splicing enhancer (ISE)/silencer (ISS) to promote binding to the SR or hnRNP protein, respectively. Consequently, alternative splicing occurs step by step with the assistance of SR-ESE or inhibition by hnRNP-ESS [1,55] (Figure 3). To obtain a comprehensive understanding of the AS regulation network, we propose some causes to explain various physiological and pathological changes. Elucidating the mechanism of AS is important for understanding gene expression and the pathogenesis of disease.
Figure 3.

A comprehensive overview of the AS machinery and the epigenetic role of lncRNAs. At least five kinds of snRNAs (U1, U2, U4/U6 and U5) and more than 200 snRNPs are included in spliceosomes. The functional spliceosome complex is formed step by step from the A complex to the P complex and recognizes GU and AG dinucleotides. Splicing factors play a significant role in regulating AS, and exon splicing enhancer (ESE)/silencer (ESS) or intron splicing enhancer (ISE)/silencer (ISS) can bind to SR or hnRNP proteins, respectively. LncRNAs can effectively modulate AS in various ways. A & B. Epigenetically, lncRNAs interacted with SR or hnRNP proteins as “interactors” or “hijackers”. C. Many lncRNAs recruited AS-associated RBPs to splicing sites and changed the AS pattern under specific conditions. D. Some lncRNAs, such as Linc-HELLP, could fine-tune AS through direct binding to ribosomal and splicing complexes. E. Partial lncRNAs, especially antisense lncRNAs, tend to form RNA-RNA duplexes with sense transcripts and block spliceosome interactions. F. LncRNAs effectively changed epigenetic modifications or chromatin modeling to influence the overall AS decision.
Vast numbers of transcripts in humans cannot be further translated into proteins. However, ncRNAs seem to affect diverse biological processes by modulating mRNA transcription, splicing, stability, transportation and epigenetic modification [56]. A plethora of long noncoding RNAs (lncRNAs) have been shown to pivotally regulate AS during disease, and the fine-tuning mechanisms extend across all steps of gene expression, including epigenetic sensu lato, co/posttranscriptional control, miRNA maturation and protein stability [27,57,58]. Herein, we summarized the principle that lncRNAs influence AS [5-14,16,59-70] (Figure 3). First, lncRNAs such as Malat1 [71], PVT1 [60], Xist [13], and PANDAR [11] could interact with specific splicing factors (SFs) and function as “interactors” or “hijackers” of splicing regulators, such as SR or hnRNP proteins (Figure 3A, 3B). For example, Malat1 (Neat2) [71,72] is able to interact with SR proteins as an “interactor” and influences the nuclear speckle distributions of SR proteins through modulating their phosphorylation status or reshaping the 3D genome organization to induce AS-related nuclear speckle formation, leading to changes in the whole AS profile. Gomafu exists in the nucleus as a “hijacker” by binding to multiple SFs (SF1, SRSF1, QKI, etc.), and Gomafu is downregulated under stimulation; thus, SFs can freely direct the AS process [66]. In addition, there are two derivative mechanisms: recruitment of RBPs to regulated splicing sites (Figure 3C) and binding to ribosomal or splicing complexes (Figure 3D). Specifically, SPA, a type of lncRNA that is 5’ snoRNA capped and 3’ polyadenylated, could accumulate to alter the AS pattern in PWS disease through binding to RBPs, such as TDP43, RBFOX2 and hnRNP M [73]. INXS could bind the Sam68 splicing modulator and cause a shift in splicing from BCL-XL to BCL-XS downstream, leading to caspase-related apoptosis in cancer cells [12]. LINC-HELLP has been confirmed to bind spliceosome components or ribosomes as a crucial AS regulator [74]. Second, lncRNAs can sometimes form RNA-RNA duplexes with pre-RNA molecules to affect AS events. One example is Fas-antisense (Saf), which can interact with the Fas premRNA and facilitate the exclusion of exon 6 in Fas through combining with SPF45, resulting in the accumulation of FasΔE6, (sFas) which prevents tumor cells from undergoing FasL-induced apoptosis [75]. Zeb2 NAT [76] and lincRNA-uc002yug.2 [17] modulate AS in a similar manner by interacting with the Zeb2 5’UTR to mask the 5’ splice site or covering the SRSF1/MBNL1 recognition site to facilitate AS of the RUNX1 premRNA, respectively. Third, lncRNAs could also recruit chromatin remodelers to the specific locus to regulate the AS process of target genes (Figure 3E). In addition to FGFR2, the FGFR2 locus encodes an evolutionarily conserved nuclear antisense lncRNA (asFGFR2) to promote epithelium-specific alternative splicing of FGFR2. asFGFR2 can recruit Polycomb-group proteins and the histone demethylase KDM2a to create a chromatin environment that is inaccessible for a repressive chromatin-splicing adaptor complex, which is much more important for mesenchymal-specific splicing [10].
To a large extent, various types of lncRNAs participate in modulating AS patterns, including NATs, UCEs and piRNAs. Historically, efforts to generate small molecular drugs targeting protein activation have been unsatisfactory. Given that lncRNAs could be inhibited much more easily through oligonucleotide-based drugs, such drugs seem preferable to undruggable proteins for cancer therapy. We are hoping to develop lncRNAs as pivotal diagnostic or therapeutic targets in the near future [77-80].
Regulation of AS by the m6A modification
N6-methyladenosine (m6A) is the most prevalent and significant modification of eukaryotic transcripts, affecting various aspects of biological processes, including RNA splicing [81]. This process is catalyzed by methyltransferases, such as METTL3, METTL14, WTAP, KIAA1429 or RBM15/RBM15B, which are called “Writers” in mammals [82,83]. RNA m6A methylation occurs mainly at the “RRACH” consensus sequence motif (R, purine; H, nonguanine base). The methylation process is reversible because of the existence of demethyltransferases, such as FTO and ALKBH5, which serve as “Erasers”. Essentially, m6A modification could contribute to mRNA fate determination by recruiting m6A recognition proteins named “Readers” (Figure 4).
Figure 4.

Schematic diagram of m6A modification, which is essential for AS. m6A methylation of RNAs is catalyzed by the multicomponent methyltransferase complex called “Writers”, which includes three main subunits: METTL3, METTL14 and WTAP, and sometimes also contains KIAA1429 or RBM15/15B. FTO and ALKBH5 are m6A demethylases that catalyze m6A removal as “Erasers”. Some m6A-binding proteins named “Readers” contribute to recognizing m6A modifications and modulating various pathways. Thus, hnRNAPA2B1, hnRNPC and YTHDC1 seem to fine-tune RNA processing or splicing, while YTHDF1/3 and eIF3/4F influence mRNA translation, and YTHDF2, YTHDC2 and HuR play pivotal roles in RNA stability or decay. Perturbation of the dynamic m6A status causes alterations in AS. For example, the m6A removal catalyzed by FTO brings about the failure of SRSF2 binding to methylated RNAs, and exons are skipped in unmethylated transcripts; in contrast, YTHDC1 could effectively bind to m6A methylated RNAs and recruit SRSF3 instead of SRSF10 to pre-RNAs, leading to exon inclusion.
Perturbation of the dynamic status of m6A could affect the levels of a large number of RNAs, since the methylation-related molecules were located in nuclear speckles and confirmed to interact with splicing factors that may be involved in modulating AS [84]. Additionally, dysregulation of methyltransferases, demethyltransferases and readers all remarkably influenced AS. METTL3 knockdown facilitates the MyD88 variant through regulating AS [85] and can alter the whole splicing isoform pattern by combining with WTAP, which was initially identified as the WT1 binding splicing factor through targeting the AS site in multi-isoform genes [86]. Emerging evidence has indicated that METTL16, the U6 snRNA m6A methyltransferase, is required for the AS of MAT2A with retained introns and stabilizes MAT2A mRNA by methylation to modulate SAM synthetase. Furthermore, METTL16 could mediate splicing of diverse cellular RNAs, including ncRNAs [87-89].
With respect to the demethyltransferases, re-cent studies indicated that the downregulation of ALKBH5 could affect the distribution of splicing factors, such as SC35, ASF/SF2 and SRPK1 [90]. Additionally, ALKBH5-mediated m6A erasure effectively ensured the appropriate splicing event to produce mRNAs with longer 3’UTR and prevent degradation, mainly due to the predominant colocalization of ALKBH5 and SC35 in the nucleus [91]. Moreover, FTO could mediate AS of RUNX1 through recruiting SRSF2 (recognizes the ESE (Exon Splicing Enhancer) sequence). Specifically, FTO reduces the capacity of SRSF2 to bind to RUNX1, while exon skipping occurs in the unmethylated premRNA to form the short isoform and promote preadipocyte differentiation. In contrast, m6A-modified premRNA (FTO deficiency) could effectively recruit SRSF2 and induce exon inclusion, while METTL3 could antagonize this effect [92] (Figure 4). However, apart from the abovementioned results, AS is tightly coupled to 3’ end processing, and m6A sites are markedly enriched in the 3’UTR [93]. FTO could potentially influence alternative poly(A) site (APA) usage and 3’UTR length [94], possibly through influencing the U2AF 65-CF Im-PAP interaction or altering the miRNA binding pattern [95,96].
Ultimately, the classical m6A readers include YTHDF-3 and YTHDC1-2 of the YTH domain family [97]. The functions of YTH domain family readers are well established, and they effectively accelerate RNA processing, translation, decay or transportation/localization and even affect viral life cycles [98,99]. YTHDC1, which is a nuclear m6A reader, has been verified to modulate splicing through maintaining a dynamic regulation network [29,100]. In detail, YTHDC1 recognizes m6A modification and effectively recruits SRSF3 to pre-RNAs, accompanied by exon inclusion. Conversely, the pre-RNAs are unmethylated, and YTHDC1 fails to recognize the RNA; therefore, SRSF10 binds to and masks the region near the m6A sites, ultimately resulting in exon skipping owing to the failure of m6A modification [98,100]. Above all, m6A modification could mediate fine-tuning by decreasing the cognate combination of SFs and changing the secondary RNA structure to alter the interaction capacity between SFs and m6A-methylated RNAs [101]. The second class of m6A readers includes heterogeneous nuclear ribonucleoproteins (hnRNPs), such as hnRNPC, hnRNPG and hnRNPA2B1 [99,101]. In addition to eliciting AS in a manner similar to that described for METTL3 by directly binding nuclear transcripts, hnRNPA2B1 plays a pivotal role in AS and in premiRNA processing by interacting with the DGCR8-DROSHA complex [102] (Figure 4). Overall, m6A modification is closely linked to the AS process and may develop into overriding signaling for mRNA processing and splicing.
Conclusions and perspectives
By reprogramming genome expression to confer proteomic and transcriptomic diversity, AS is becoming one of the most important dynamic regulatory processes for adaptation to a changing microenvironment. Therefore, it is not surprising to see so many mechanisms of AS regulation by epigenetic networks, such as the modification of DNA, RNA and histone proteins. In this review, we aim to retrospectively evaluate new research on alternative splicing, which will facilitate the process from discovery to future therapeutic use. The main topics for the study of AS are as follows: history of mRNA splicing, components of the spliceosome, introns/exons in alternative splicing, spliceosome complex regulation, diseases related to splicing and potential therapeutic approaches to splicing-induced diseases [103,104].
With regard to therapeutic options for AS-relevant disease, especially for cancer, there are quite a few achievements that could be used as a reference. Many small molecules that can modulate AS have been developed as potential treatment targets for disease through reversing cancer or drug resistance [105-107]. For instance, BaxΔ2, a novel isoform of Bax with exon 2 alternative splicing, could induce the mutation of Bax, and this AS event effectively increased the sensitivity of colon cancer cells to caspase-8-targeted chemotherapeutic drugs [108]. In addition, clinical trials (phase III) of antisense oligo-targeted therapies have been reported [109].
In conclusion, further exploration of AS is needed to elucidate the full landscape of AS, the complete regulatory networks that control AS, such as ncRNAs or m6A modifications, and the function of the vast majority of AS events associated with diseases or disorders.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (81672723, 91740106, 81761138047).
Disclosure of conflict of interest
None.
References
- 1.Tuck AC, Tollervey D. RNA in pieces. Trends Genet. 2011;27:422–432. doi: 10.1016/j.tig.2011.06.001. [DOI] [PubMed] [Google Scholar]
- 2.Elkon R, Ugalde AP, Agami R. Alternative cleavage and polyadenylation: extent, regulation and function. Nat Rev Genet. 2013;14:496–506. doi: 10.1038/nrg3482. [DOI] [PubMed] [Google Scholar]
- 3.Berk AJ. Discovery of RNA splicing and genes in pieces. Proc Natl Acad Sci U S A. 2016;113:801–805. doi: 10.1073/pnas.1525084113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Boutz PL, Chawla G, Stoilov P, Black DL. MicroRNAs regulate the expression of the alternative splicing factor nPTB during muscle development. Genes Dev. 2007;21:71–84. doi: 10.1101/gad.1500707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Sehgal L, Mathur R, Braun FK, Wise JF, Berkova Z, Neelapu S, Kwak LW, Samaniego F. FAS-antisense 1 lncRNA and production of soluble versus membrane Fas in B-cell lymphoma. Leukemia. 2014;28:2376–2387. doi: 10.1038/leu.2014.126. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Yuan JH, Liu XN, Wang TT, Pan W, Tao QF, Zhou WP, Wang F, Sun SH. The MBNL3 splicing factor promotes hepatocellular carcinoma by increasing PXN expression through the alternative splicing of lncRNA-PXN-AS1. Nat Cell Biol. 2017;19:820–832. doi: 10.1038/ncb3538. [DOI] [PubMed] [Google Scholar]
- 7.Minotti L, Baldassari F, Galasso M, Volinia S, Bergamini CM, Bianchi N. A long non-coding RNA inside the type 2 transglutaminase gene tightly correlates with the expression of its transcriptional variants. Amino Acids. 2018;50:421–438. doi: 10.1007/s00726-017-2528-9. [DOI] [PubMed] [Google Scholar]
- 8.Huang JZ, Chen M, Chen , Gao XC, Zhu S, Huang H, Hu M, Zhu H, Yan GR. A peptide encoded by a putative lncRNA HOXB-AS3 suppresses colon cancer growth. Mol Cell. 2017;68:171–184. e6. doi: 10.1016/j.molcel.2017.09.015. [DOI] [PubMed] [Google Scholar]
- 9.Huang GW, Zhang YL, Liao LD, Li EM, Xu LY. Natural antisense transcript TPM1-AS regulates the alternative splicing of tropomyosin I through an interaction with RNA-binding motif protein 4. Int J Biochem Cell Biol. 2017;90:59–67. doi: 10.1016/j.biocel.2017.07.017. [DOI] [PubMed] [Google Scholar]
- 10.Gonzalez I, Munita R, Agirre E, Dittmer TA, Gysling K, Misteli T, Luco RF. A lncRNA regulates alternative splicing via establishment of a splicing-specific chromatin signature. Nat Struct Mol Biol. 2015;22:370–376. doi: 10.1038/nsmb.3005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Pospiech N, Cibis H, Dietrich L, Muller F, Bange T, Hennig S. Identification of novel PANDAR protein interaction partners involved in splicing regulation. Sci Rep. 2018;8:2798. doi: 10.1038/s41598-018-21105-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.DeOcesano-Pereira C, Amaral MS, Parreira KS, Ayupe AC, Jacysyn JF, Amarante-Mendes GP, Reis EM, Verjovski-Almeida S. Long non-coding RNA INXS is a critical mediator of BCL-XS induced apoptosis. Nucleic Acids Res. 2014;42:8343–8355. doi: 10.1093/nar/gku561. [DOI] [PMC free article] [PubMed] [Google Scholar] [Retracted]
- 13.Yue M, Ogawa Y. CRISPR/Cas9-mediated modulation of splicing efficiency reveals short splicing isoform of Xist RNA is sufficient to induce X-chromosome inactivation. Nucleic Acids Res. 2018;46:e26. doi: 10.1093/nar/gkx1227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Grelet S, Link LA, Howley B, Obellianne C, Palanisamy V, Gangaraju VK, Diehl JA, Howe PH. A regulated PNUTS mRNA to lncRNA splice switch mediates EMT and tumour progression. Nat Cell Biol. 2017;19:1105–1115. doi: 10.1038/ncb3595. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Pirnie SP, Osman A, Zhu Y, Carmichael GG. An ultraconserved element (UCE) controls homeostatic splicing of ARGLU1 mRNA. Nucleic Acids Res. 2017;45:3473–3486. doi: 10.1093/nar/gkw1140. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Huan C, Li Z, Ning S, Wang H, Yu XF, Zhang W. Long noncoding RNA uc002yug.2 activates HIV-1 latency through regulation of mrna levels of various RUNX1 isoforms and increased tat expression. J Virol. 2018;92 doi: 10.1128/JVI.01844-17. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Wu H, Zheng J, Deng J, Zhang L, Li N, Li W, Li F, Lu J, Zhou Y. LincRNA-uc002yug.2 involves in alternative splicing of RUNX1 and serves as a predictor for esophageal cancer and prognosis. Oncogene. 2015;34:4723–4734. doi: 10.1038/onc.2014.400. [DOI] [PubMed] [Google Scholar]
- 18.Kopin AS, Wheeler MB, Nishitani J, McBride EW, Chang TM, Chey WY, Leiter AB. The secretin gene: evolutionary history, alternative splicing, and developmental regulation. Proc Nati Acad Sci U S A. 1991;88:5335–5339. doi: 10.1073/pnas.88.12.5335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Meric F, Searfoss AM, Wormington M, Wolffe AP. Masking and unmasking maternal mRNA. The role of polyadenylation, transcription, splicing, and nuclear history. J Biol Chem. 1996;271:30804–30810. doi: 10.1074/jbc.271.48.30804. [DOI] [PubMed] [Google Scholar]
- 20.Lareau LF, Green RE, Bhatnagar RS, Brenner SE. The evolving roles of alternative splicing. Curr Opin Struct Biol. 2004;14:273–282. doi: 10.1016/j.sbi.2004.05.002. [DOI] [PubMed] [Google Scholar]
- 21.Roy SW, Irimia M. Splicing in the eukaryotic ancestor: form, function and dysfunction. Trends Ecol Evol. 2009;24:447–455. doi: 10.1016/j.tree.2009.04.005. [DOI] [PubMed] [Google Scholar]
- 22.Darnell JE Jr. Reflections on the history of pre-mRNA processing and highlights of current knowledge: a unified picture. RNA. 2013;19:443–460. doi: 10.1261/rna.038596.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 23.Sadek J, Read GS. The splicing history of an mRNA affects its level of translation and sensitivity to cleavage by the virion host shutoff endonuclease during herpes simplex virus infections. J Virol. 2016;90:10844–10856. doi: 10.1128/JVI.01302-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Baralle M, Baralle FE. The splicing code. Biosystems. 2018;164:39–48. doi: 10.1016/j.biosystems.2017.11.002. [DOI] [PubMed] [Google Scholar]
- 25.Chow LT, Gelinas RE, Broker TR, Roberts RJ. An amazing sequence arrangement at the 5’ ends of adenovirus 2 messenger RNA. Cell. 1977;12:1–8. doi: 10.1016/0092-8674(77)90180-5. [DOI] [PubMed] [Google Scholar]
- 26.Berget SM, Moore C, Sharp PA. Spliced segments at the 5’ terminus of adenovirus 2 late mRNA. Proc Natl Acad Sci U S A. 1977;74:3171–3175. doi: 10.1073/pnas.74.8.3171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Deveson IW, Hardwick SA, Mercer TR, Mattick JS. The dimensions, dynamics, and relevance of the mammalian noncoding transcriptome. Trends Genet. 2017;33:464–478. doi: 10.1016/j.tig.2017.04.004. [DOI] [PubMed] [Google Scholar]
- 28.Shi Y. Mechanistic insights into precursor messenger RNA splicing by the spliceosome. Nat Rev Mol Cell Biol. 2017;18:655–670. doi: 10.1038/nrm.2017.86. [DOI] [PubMed] [Google Scholar]
- 29.Adhikari S, Xiao W, Zhao YL, Yang YG. m(6)A: signaling for mRNA splicing. RNA Biol. 2016;13:756–759. doi: 10.1080/15476286.2016.1201628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Goldstrohm AC, Greenleaf AL, Garcia-Blanco MA. Co-transcriptional splicing of pre-messenger RNAs: considerations for the mechanism of alternative splicing. Gene. 2001;277:31–47. doi: 10.1016/s0378-1119(01)00695-3. [DOI] [PubMed] [Google Scholar]
- 31.Cramer P, Pesce CG, Baralle FE, Kornblihtt AR. Functional association between promoter structure and transcript alternative splicing. Proc Natl Acad Sci U S A. 1997;94:11456–11460. doi: 10.1073/pnas.94.21.11456. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Naftelberg S, Schor IE, Ast G, Kornblihtt AR. Regulation of alternative splicing through coupling with transcription and chromatin structure. Annu Rev Biochem. 2015;84:165–198. doi: 10.1146/annurev-biochem-060614-034242. [DOI] [PubMed] [Google Scholar]
- 33.Custodio N, Carmo-Fonseca M. Co-transcriptional splicing and the CTD code. Crit Rev Biochem Mol Biol. 2016;51:395–411. doi: 10.1080/10409238.2016.1230086. [DOI] [PubMed] [Google Scholar]
- 34.Fong N, Bird G, Vigneron M, Bentley DL. A 10 residue motif at the C-terminus of the RNA pol II CTD is required for transcription, splicing and 3’ end processing. EMBO J. 2003;22:4274–4282. doi: 10.1093/emboj/cdg396. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Pearson JL, Robinson TJ, Munoz MJ, Kornblihtt AR, Garcia-Blanco MA. Identification of the cellular targets of the transcription factor TCERG1 reveals a prevalent role in mRNA processing. J Biol Chem. 2008;283:7949–7961. doi: 10.1074/jbc.M709402200. [DOI] [PubMed] [Google Scholar]
- 36.Guillouf C, Gallais I, Moreau-Gachelin F. Spi-1/PU. 1 oncoprotein affects splicing decisions in a promoter binding-dependent manner. J Biol Chem. 2006;281:19145–19155. doi: 10.1074/jbc.M512049200. [DOI] [PubMed] [Google Scholar]
- 37.Kornblihtt AR, Schor IE, Allo M, Blencowe BJ. When chromatin meets splicing. Nat Struct Mol Biol. 2009;16:902–903. doi: 10.1038/nsmb0909-902. [DOI] [PubMed] [Google Scholar]
- 38.Allemand E, Myers MP, Garcia-Bernardo J, Harel-Bellan A, Krainer AR, Muchardt C. A broad set of chromatin factors influences splicing. PLoS Genet. 2016;12:e1006318. doi: 10.1371/journal.pgen.1006318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Luco RF, Pan Q, Tominaga K, Blencowe BJ, Pereira-Smith OM, Misteli T. Regulation of alternative splicing by histone modifications. Science. 2010;327:996–1000. doi: 10.1126/science.1184208. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Yun M, Wu J, Workman JL, Li B. Readers of histone modifications. Cell Res. 2011;21:564–578. doi: 10.1038/cr.2011.42. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Anastasiadou C, Malousi A, Maglaveras N, Kouidou S. Human epigenome data reveal increased CpG methylation in alternatively spliced sites and putative exonic splicing enhancers. DNA Cell Biol. 2011;30:267–275. doi: 10.1089/dna.2010.1094. [DOI] [PubMed] [Google Scholar]
- 42.Maunakea AK, Chepelev I, Cui K, Zhao K. Intragenic DNA methylation modulates alternative splicing by recruiting MeCP2 to promote exon recognition. Cell Res. 2013;23:1256–1269. doi: 10.1038/cr.2013.110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Malousi A, Kouidou S. DNA hypermethylation of alternatively spliced and repeat sequences in humans. Mol Genet Genomics. 2012;287:631–642. doi: 10.1007/s00438-012-0703-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Batsche E, Yaniv M, Muchardt C. The human SWI/SNF subunit Brm is a regulator of alternative splicing. Nat Struct Mol Biol. 2006;13:22–29. doi: 10.1038/nsmb1030. [DOI] [PubMed] [Google Scholar]
- 45.Pradeepa MM, Sutherland HG, Ule J, Grimes GR, Bickmore WA. Psip1/Ledgf p52 binds methylated histone H3K36 and splicing factors and contributes to the regulation of alternative splicing. PLoS Genet. 2012;8:e1002717. doi: 10.1371/journal.pgen.1002717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Saint-Andre V, Batsche E, Rachez C, Muchardt C. Histone H3 lysine 9 trimethylation and HP1gamma favor inclusion of alternative exons. Nat Struct Mol Biol. 2011;18:337–344. doi: 10.1038/nsmb.1995. [DOI] [PubMed] [Google Scholar]
- 47.Shukla S, Kavak E, Gregory M, Imashimizu M, Shutinoski B, Kashlev M, Oberdoerffer P, Sandberg R, Oberdoerffer S. CTCF-promoted RNA polymerase II pausing links DNA methylation to splicing. Nature. 2011;479:74–79. doi: 10.1038/nature10442. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Schwartz S, Ast G. Chromatin density and splicing destiny: on the cross-talk between chromatin structure and splicing. EMBO J. 2010;29:1629–1636. doi: 10.1038/emboj.2010.71. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Clapier CR, Cairns BR. The biology of chromatin remodeling complexes. Annu Rev Biochem. 2009;78:273–304. doi: 10.1146/annurev.biochem.77.062706.153223. [DOI] [PubMed] [Google Scholar]
- 50.Llorian M, Schwartz S, Clark TA, Hollander D, Tan LY, Spellman R, Gordon A, Schweitzer AC, de la Grange P, Ast G, Smith CW. Position-dependent alternative splicing activity revealed by global profiling of alternative splicing events regulated by PTB. Nat Struct Mol Biol. 2010;17:1114–1123. doi: 10.1038/nsmb.1881. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Gelfman S, Cohen N, Yearim A, Ast G. DNA-methylation effect on cotranscriptional splicing is dependent on GC architecture of the exon-intron structure. Genome Res. 2013;23:789–799. doi: 10.1101/gr.143503.112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Laurent L, Wong E, Li G, Huynh T, Tsirigos A, Ong CT, Low HM, Kin Sung KW, Rigoutsos I, Loring J, Wei CL. Dynamic changes in the human methylome during differentiation. Genome Res. 2010;20:320–331. doi: 10.1101/gr.101907.109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Grewal SI, Moazed D. Heterochromatin and epigenetic control of gene expression. Science. 2003;301:798–802. doi: 10.1126/science.1086887. [DOI] [PubMed] [Google Scholar]
- 54.Irimia M, Roy SW. Origin of spliceosomal introns and alternative splicing. Cold Spring Harb Perspect Biol. 2014;6 doi: 10.1101/cshperspect.a016071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Matera AG, Wang Z. A day in the life of the spliceosome. Nat Rev Mol Cell Biol. 2014;15:108–121. doi: 10.1038/nrm3742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.An S, Song JJ. The coded functions of noncoding RNAs for gene regulation. Mol Cells. 2011;31:491–496. doi: 10.1007/s10059-011-1004-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Gagnon KT, Corey DR. Argonaute and the nuclear RNAs: new pathways for RNA-mediated control of gene expression. Nucleic Acid Ther. 2012;22:3–16. doi: 10.1089/nat.2011.0330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Engreitz JM, Haines JE, Perez EM, Munson G, Chen J, Kane M, McDonel PE, Guttman M, Lander ES. Local regulation of gene expression by lncRNA promoters, transcription and splicing. Nature. 2016;539:452–455. doi: 10.1038/nature20149. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Redis RS, Vela LE, Lu W, Ferreira de Oliveira J, Ivan C, Rodriguez-Aguayo C, Adamoski D, Pasculli B, Taguchi A, Chen Y, Fernandez AF, Valledor L, Van Roosbroeck K, Chang S, Shah M, Kinnebrew G, Han L, Atlasi Y, Cheung LH, Huang GY, Monroig P, Ramirez MS, Catela Ivkovic T, Van L, Ling H, Gafà R, Kapitanovic S, Lanza G, Bankson JA, Huang P, Lai SY, Bast RC, Rosenblum MG, Radovich M, Ivan M, Bartholomeusz G, Liang H, Fraga MF, Widger WR, Hanash S, Berindan-Neagoe I, Lopez-Berestein G, Ambrosio ALB, Gomes Dias SM, Calin GA. Allele-specific reprogramming of cancer metabolism by the long non-coding RNA CCAT2. Mol Cell. 2016;61:520–534. doi: 10.1016/j.molcel.2016.01.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Yang T, Zhou H, Liu P, Yan L, Yao W, Chen K, Zeng J, Li H, Hu J, Xu H, Ye Z. lncRNA PVT1 and its splicing variant function as competing endogenous RNA to regulate clear cell renal cell carcinoma progression. Oncotarget. 2017;8:85353–85367. doi: 10.18632/oncotarget.19743. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Rodriguez-Mateo C, Torres B, Gutierrez G, Pintor-Toro JA. Downregulation of Lnc-Spry1 mediates TGF-beta-induced epithelial-mesenchymal transition by transcriptional and posttranscriptional regulatory mechanisms. Cell Death Differ. 2017;24:785–797. doi: 10.1038/cdd.2017.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Aoki K, Harashima A, Sano M, Yokoi T, Nakamura S, Kibata M, Hirose T. A thymus-specific noncoding RNA, Thy-ncR1, is a cytoplasmic riboregulator of MFAP4 mRNA in immature T-cell lines. BMC Molecular Biology. 2010;11:99. doi: 10.1186/1471-2199-11-99. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Ciarlo E, Massone S, Penna I, Nizzari M, Gigoni A, Dieci G, Russo C, Florio T, Cancedda R, Pagano A. An intronic ncRNA-dependent regulation of SORL1 expression affecting abeta formation is upregulated in post-mortem Alzheimer’s disease brain samples. Dis Model Mech. 2013;6:424–433. doi: 10.1242/dmm.009761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Wang J, Li X, Wang L, Li J, Zhao Y, Bou G, Li Y, Jiao G, Shen X, Wei R, Liu S, Xie B, Lei L, Li W, Zhou Q, Liu Z. A novel long intergenic noncoding RNA indispensable for the cleavage of mouse two-cell embryos. EMBO Rep. 2016;17:1452–1470. doi: 10.15252/embr.201642051. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Feng J, Bi C, Clark BS, Mady R, Shah P, Kohtz JD. The Evf-2 noncoding RNA is transcribed from the Dlx-5/6 ultraconserved region and functions as a Dlx-2 transcriptional coactivator. Genes Dev. 2006;20:1470–1484. doi: 10.1101/gad.1416106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Barry G, Briggs JA, Vanichkina DP, Poth EM, Beveridge NJ, Ratnu VS, Nayler SP, Nones K, Hu J, Bredy TW, Nakagawa S, Rigo F, Taft RJ, Cairns MJ, Blackshaw S, Wolvetang EJ, Mattick JS. The long non-coding RNA Gomafu is acutely regulated in response to neuronal activation and involved in schizophrenia-associated alternative splicing. Mol Psychiatry. 2014;19:486–494. doi: 10.1038/mp.2013.45. [DOI] [PubMed] [Google Scholar]
- 67.Massone S, Vassallo I, Fiorino G, Castelnuovo M, Barbieri F, Borghi R, Tabaton M, Robello M, Gatta E, Russo C, Florio T, Dieci G, Cancedda R, Pagano A. 17A, a novel non-coding RNA, regulates GABA B alternative splicing and signaling in response to inflammatory stimuli and in Alzheimer disease. Neurobiol Dis. 2011;41:308–317. doi: 10.1016/j.nbd.2010.09.019. [DOI] [PubMed] [Google Scholar]
- 68.Niemczyk M, Ito Y, Huddleston J, Git A, Abu-Amero S, Caldas C, Moore GE, Stojic L, Murrell A. Imprinted chromatin around DIRAS3 regulates alternative splicing of GNG12-AS1, a long noncoding RNA. Am J Hum Genet. 2013;93:224–235. doi: 10.1016/j.ajhg.2013.06.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Teixeira FK, Okuniewska M, Malone CD, Coux RX, Rio DC, Lehmann R. piRNA-mediated regulation of transposon alternative splicing in the soma and germ line. Nature. 2017;552:268–272. doi: 10.1038/nature25018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Ramos AD, Andersen RE, Liu SJ, Nowakowski TJ, Hong SJ, Gertz C, Salinas RD, Zarabi H, Kriegstein AR, Lim DA. The long noncoding RNA Pnky regulates neuronal differentiation of embryonic and postnatal neural stem cells. Cell Stem Cell. 2015;16:439–447. doi: 10.1016/j.stem.2015.02.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Malakar P, Shilo A, Mogilevsky A, Stein I, Pikarsky E, Nevo Y, Benyamini H, Elgavish S, Zong X, Prasanth KV, Karni R. Long noncoding RNA MALAT1 promotes hepatocellular carcinoma development by SRSF1 upregulation and mTOR activation. Cancer Res. 2017;77:1155–1167. doi: 10.1158/0008-5472.CAN-16-1508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Tripathi V, Ellis JD, Shen Z, Song DY, Pan Q, Watt AT, Freier SM, Bennett CF, Sharma A, Bubulya PA, Blencowe BJ, Prasanth SG, Prasanth KV. The nuclear-retained noncoding RNA MALAT1 regulates alternative splicing by modulating SR splicing factor phosphorylation. Mol Cell. 2010;39:925–938. doi: 10.1016/j.molcel.2010.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Wu H, Yin QF, Luo Z, Yao RW, Zheng CC, Zhang J, Xiang JF, Yang L, Chen LL. Unusual processing generates SPA LncRNAs that sequester multiple RNA binding proteins. Mol Cell. 2016;64:534–548. doi: 10.1016/j.molcel.2016.10.007. [DOI] [PubMed] [Google Scholar]
- 74.van Dijk M, Visser A, Buabeng KM, Poutsma A, van der Schors RC, Oudejans CB. Mutations within the LINC-HELLP non-coding RNA differentially bind ribosomal and RNA splicing complexes and negatively affect trophoblast differentiation. Hum Mol Genet. 2015;24:5475–5485. doi: 10.1093/hmg/ddv274. [DOI] [PubMed] [Google Scholar]
- 75.Villamizar O, Chambers CB, Riberdy JM, Persons DA, Wilber A. Long noncoding RNA Saf and splicing factor 45 increase soluble Fas and resistance to apoptosis. Oncotarget. 2016;7:13810–13826. doi: 10.18632/oncotarget.7329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Beltran M, Puig I, Pena C, Garcia JM, Alvarez AB, Pena R, Bonilla F, de Herreros AG. A natural antisense transcript regulates Zeb2/Sip1 gene expression during Snail1-induced epithelial-mesenchymal transition. Genes Dev. 2008;22:756–769. doi: 10.1101/gad.455708. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Evans JR, Feng FY, Chinnaiyan AM. The bright side of dark matter: lncRNAs in cancer. J Clin Invest. 2016;126:2775–2782. doi: 10.1172/JCI84421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Huarte M. The emerging role of lncRNAs in cancer. Nat Med. 2015;21:1253–1261. doi: 10.1038/nm.3981. [DOI] [PubMed] [Google Scholar]
- 79.Khorkova O, Hsiao J, Wahlestedt C. Basic biology and therapeutic implications of lncRNA. Adv Drug Deliv Rev. 2015;87:15–24. doi: 10.1016/j.addr.2015.05.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Quinn JJ, Chang HY. Unique features of long non-coding RNA biogenesis and function. Nat Rev Genet. 2016;17:47–62. doi: 10.1038/nrg.2015.10. [DOI] [PubMed] [Google Scholar]
- 81.Dai D, Wang H, Zhu L, Jin H, Wang X. N6-methyladenosine links RNA metabolism to cancer progression. Cell Death Dis. 2018;9:124. doi: 10.1038/s41419-017-0129-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Edupuganti RR, Geiger S, Lindeboom RGH, Shi H, Hsu PJ, Lu Z, Wang SY, Baltissen MPA, Jansen PWTC, Rossa M, Müller M, Stunnenberg HG, He C, Carell T, Vermeulen M. N(6)-methyladenosine (m(6)A) recruits and repels proteins to regulate mRNA homeostasis. Nat Struct Mol Biol. 2017;24:870–878. doi: 10.1038/nsmb.3462. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Roignant JY, Soller M. m(6)A in mRNA: an ancient mechanism for fine-tuning gene expression. Trends Genet. 2017;33:380–390. doi: 10.1016/j.tig.2017.04.003. [DOI] [PubMed] [Google Scholar]
- 84.Yang Y, Sun BF, Xiao W, Yang X, Sun HY, Zhao YL, Yang YG. Dynamic m 6 A modification and its emerging regulatory role in mRNA splicing. Science Bulletin. 2015;60:21–32. [Google Scholar]
- 85.Feng Z, Li Q, Meng R, Yi B, Xu Q. METTL3 regulates alternative splicing of MyD88 upon the lipopolysaccharide-induced inflammatory response in human dental pulp cells. J Cell Mol Med. 2018;22:2558–2568. doi: 10.1111/jcmm.13491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Ping XL, Sun BF, Wang L, Xiao W, Yang X, Wang WJ, Adhikari S, Shi Y, Lv Y, Chen YS, Zhao X, Li A, Yang Y, Dahal U, Lou XM, Liu X, Huang J, Yuan WP, Zhu XF, Cheng T, Zhao YL, Wang X, Rendtlew Danielsen JM, Liu F, Yang YG. Mammalian WTAP is a regulatory subunit of the RNA N6-methyladenosine methyltransferase. Cell Res. 2014;24:177–189. doi: 10.1038/cr.2014.3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Shima H, Matsumoto M, Ishigami Y, Ebina M, Muto A, Sato Y, Kumagai S, Ochiai K, Suzuki T, Igarashi K. S-Adenosylmethionine synthesis is regulated by selective N(6)-adenosine methylation and mRNA degradation involving METTL16 and YTHDC1. Cell Rep. 2017;21:3354–3363. doi: 10.1016/j.celrep.2017.11.092. [DOI] [PubMed] [Google Scholar]
- 88.Pendleton KE, Chen B, Liu K, Hunter OV, Xie Y, Tu BP, Conrad NK. The U6 snRNA m(6)A methyltransferase METTL16 regulates SAM synthetase intron retention. Cell. 2017;169:824–835. e14. doi: 10.1016/j.cell.2017.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Warda AS, Kretschmer J, Hackert P, Lenz C, Urlaub H, Hobartner C, Sloan KE, Bohnsack MT. Human METTL16 is a N(6)-methyladenosine (m(6)A) methyltransferase that targets pre-mRNAs and various non-coding RNAs. EMBO Rep. 2017;18:2004–2014. doi: 10.15252/embr.201744940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Zheng G, Dahl JA, Niu Y, Fedorcsak P, Huang CM, Li CJ, Vagbo CB, Shi Y, Wang WL, Song SH, Lu Z, Bosmans RP, Dai Q, Hao YJ, Yang X, Zhao WM, Tong WM, Wang XJ, Bogdan F, Furu K, Fu Y, Jia G, Zhao X, Liu J, Krokan HE, Klungland A, Yang YG, He C. ALKBH5 is a mammalian RNA demethylase that impacts RNA metabolism and mouse fertility. Mol Cell. 2013;49:18–29. doi: 10.1016/j.molcel.2012.10.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Tang C, Klukovich R, Peng H, Wang Z, Yu T, Zhang Y, Zheng H, Klungland A, Yan W. ALKBH5-dependent m6A demethylation controls splicing and stability of long 3’-UTR mRNAs in male germ cells. Proc Natl Acad Sci U S A. 2018;115:E325–E333. doi: 10.1073/pnas.1717794115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Zhao X, Yang Y, Sun BF, Shi Y, Yang X, Xiao W, Hao YJ, Ping XL, Chen YS, Wang WJ, Jin KX, Wang X, Huang CM, Fu Y, Ge XM, Song SH, Jeong HS, Yanagisawa H, Niu Y, Jia GF, Wu W, Tong WM, Okamoto A, He C, Rendtlew Danielsen JM, Wang XJ, Yang YG. FTO-dependent demethylation of N6-methyladenosine regulates mRNA splicing and is required for adipogenesis. Cell Res. 2014;24:1403–1419. doi: 10.1038/cr.2014.151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Meyer KD, Saletore Y, Zumbo P, Elemento O, Mason CE, Jaffrey SR. Comprehensive analysis of mRNA methylation reveals enrichment in 3’ UTRs and near stop codons. Cell. 2012;149:1635–1646. doi: 10.1016/j.cell.2012.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Bartosovic M, Molares HC, Gregorova P, Hrossova D, Kudla G, Vanacova S. N6-methyladenosine demethylase FTO targets pre-mRNAs and regulates alternative splicing and 3’-end processing. Nucleic Acids Res. 2017;45:11356–11370. doi: 10.1093/nar/gkx778. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Ke S, Alemu EA, Mertens C, Gantman EC, Fak JJ, Mele A, Haripal B, Zucker-Scharff I, Moore MJ, Park CY, Vågbø CB, Kusśnierczyk A, Klungland A, Darnell JE Jr, Darnell RB. A majority of m6A residues are in the last exons, allowing the potential for 3’ UTR regulation. Genes Dev. 2015;29:2037–2053. doi: 10.1101/gad.269415.115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Martin G, Gruber AR, Keller W, Zavolan M. Genome-wide analysis of pre-mRNA 3’ end processing reveals a decisive role of human cleavage factor I in the regulation of 3’ UTR length. Cell Rep. 2012;1:753–763. doi: 10.1016/j.celrep.2012.05.003. [DOI] [PubMed] [Google Scholar]
- 97.Patil DP, Pickering BF, Jaffrey SR. Reading m(6)A in the transcriptome: m(6)A-binding proteins. Trends Cell Biol. 2018;28:113–127. doi: 10.1016/j.tcb.2017.10.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Kasowitz SD, Ma J, Anderson SJ, Leu NA, Xu Y, Gregory BD, Schultz RM, Wang PJ. Nuclear m6A reader YTHDC1 regulates alternative polyadenylation and splicing during mouse oocyte development. PLoS Genet. 2018;14:e1007412. doi: 10.1371/journal.pgen.1007412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Dominissini D, Moshitch-Moshkovitz S, Schwartz S, Salmon-Divon M, Ungar L, Osenberg S, Cesarkas K, Jacob-Hirsch J, Amariglio N, Kupiec M, Sorek R, Rechavi G. Topology of the human and mouse m6A RNA methylomes revealed by m6A-seq. Nature. 2012;485:201–206. doi: 10.1038/nature11112. [DOI] [PubMed] [Google Scholar]
- 100.Xiao W, Adhikari S, Dahal U, Chen YS, Hao YJ, Sun BF, Sun HY, Li A, Ping XL, Lai WY, Wang X, Ma HL, Huang CM, Yang Y, Huang N, Jiang GB, Wang HL, Zhou Q, Wang XJ, Zhao YL, Yang YG. Nuclear m(6)A Reader YTHDC1 regulates mRNA splicing. Mol Cell. 2016;61:507–519. doi: 10.1016/j.molcel.2016.01.012. [DOI] [PubMed] [Google Scholar]
- 101.Liu N, Zhou KI, Parisien M, Dai Q, Diatchenko L, Pan T. N6-methyladenosine alters RNA structure to regulate binding of a low-complexity protein. Nucleic Acids Res. 2017;45:6051–6063. doi: 10.1093/nar/gkx141. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Alarcon CR, Goodarzi H, Lee H, Liu X, Tavazoie S, Tavazoie SF. HNRNPA2B1 is a mediator of m(6)A-dependent nuclear RNA processing events. Cell. 2015;162:1299–1308. doi: 10.1016/j.cell.2015.08.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Bonomi S, Gallo S, Catillo M, Pignataro D, Biamonti G, Ghigna C. Oncogenic alternative splicing switches: role in cancer progression and prospects for therapy. Int J Cell Biol. 2013;2013:962038. doi: 10.1155/2013/962038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Hallegger M, Llorian M, Smith CW. Alternative splicing: global insights. FEBS J. 2010;277:856–866. doi: 10.1111/j.1742-4658.2009.07521.x. [DOI] [PubMed] [Google Scholar]
- 105.Siegfried Z, Karni R. The role of alternative splicing in cancer drug resistance. Curr Opin Genet Dev. 2018;48:16–21. doi: 10.1016/j.gde.2017.10.001. [DOI] [PubMed] [Google Scholar]
- 106.Lee SC, Abdel-Wahab O. Therapeutic targeting of splicing in cancer. Nat Med. 2016;22:976–986. doi: 10.1038/nm.4165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Wan L, Yu W, Shen E, Sun W, Liu Y, Kong J, Wu Y, Han F, Zhang L, Yu T, Zhou Y, Xie S, Xu E, Zhang H, Lai M. SRSF6-regulated alternative splicing that promotes tumour progression offers a therapy target for colorectal cancer. Gut. 2019;68:118–129. doi: 10.1136/gutjnl-2017-314983. [DOI] [PubMed] [Google Scholar]
- 108.Haferkamp B, Zhang H, Lin Y, Yeap X, Bunce A, Sharpe J, Xiang J. BaxDelta2 is a novel bax isoform unique to microsatellite unstable tumors. J Biol Chem. 2012;287:34722–34729. doi: 10.1074/jbc.M112.374785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Salton M, Misteli T. Small molecule modulators of pre-mRNA splicing in cancer therapy. Trends Mol Med. 2016;22:28–37. doi: 10.1016/j.molmed.2015.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]

