RBM10

Last updated
RBM10
Available structures
PDB Ortholog search: PDBe RCSB
Identifiers
Aliases RBM10 , DXS8237E, GPATC9, GPATCH9, S1-1, TARPS, ZRANB5, RNA binding motif protein 10
External IDs OMIM: 300080 MGI: 2384310 HomoloGene: 31330 GeneCards: RBM10
Orthologs
SpeciesHumanMouse
Entrez
Ensembl
UniProt
RefSeq (mRNA)

NM_001204466
NM_001204467
NM_001204468
NM_005676
NM_152856

Contents

RefSeq (protein)
Location (UCSC) Chr X: 47.15 – 47.19 Mb Chr X: 20.48 – 20.52 Mb
PubMed search [3] [4]
Wikidata
View/Edit Human View/Edit Mouse

RNA-binding motif 10 is a protein that is encoded by the RBM10 gene. [5] [6] [7] [8] This gene maps on the X chromosome at Xp11.23 in humans. RBM10 is a regulator of alternative splicing. [9] [10] [11] Alternative splicing is a process associated with gene expression to produce multiple protein isoforms from a single gene, thereby creating functional diversity and cellular complexity. [12] RBM10 influences the expression of many genes, [9] [10] [13] [14] [15] participating in various cellular processes and pathways such as cell proliferation and apoptosis. [10] [16] Its mutations are associated with various human diseases [17] [18] [19] [20] [21] [22] such as TARP syndrome, [22] [17] an X-linked congenital disorder in males resulting in pre‐ or postnatal lethality, and various cancers in adults. [18] [19]

Gene and protein

The RBM10 gene spans ~41.6 kb and contains 24 exons. This gene is subjected to X-inactivation, [6] [7] in which one of the two RBM10 genes in female cells is transcriptionally silenced by heterochromatin formation.

RBM proteins constitute a large family of RNA-binding proteins (RBPs). There are 52 RBM proteins (HGNC: HUGO Gene Nomenclature Committee), each containing one to several RNA-binding domains called RNA recognition motifs (RRMs). RBM10 contains two RRMs (RRM1 and RRM2) and other domains such as two zinc fingers (ZnFs), an octamer repeat (OCRE), three nuclear localization signals (NLSs), and a glycine-rich domain (G-patch). The amino acid (aa) sequence of RBM10 is conserved among mammals. Human RBM10 isoform 1 shares 96% and 97% sequence homology with those of mice and rats, respectively, indicating that the molecular functions of RBM10 are essentially the same in humans and rodents.

RBM10 has multiple isoforms, generated via alternative splicing events of the RBM10 primary transcript. The main isoforms, 1–4, may contain an exon 4 sequence (77 residues) and/or a Val residue corresponding to the last codon of exon 10. Isoform 1 (930 residues) contains both the exon 4 sequence and V354, whereas isoform 4 (929 residues) does not contain this valine residue. Similarly, the exon 4–minus isoform 3 (853 residues) contains V277, whereas isoform 2 (852 residues) does not. Isoform 5 (995 residues) has a longer 65-aa N-terminus, compared with that of isoform 1. In addition, automated computational analysis using the Gnomon gene prediction tool (NCBI gene) has shown that there may be more than 10 different RBM isoforms.

Function

RBM10 is ubiquitously expressed in almost every type of cell, both growing as well as quiescent (UniProtKB-P98175 [human] and Q99KG3 [mouse]; The Human Protein Atlas). In general, it is more strongly expressed in actively transcribing cells. [23]

In the alternative splicing regulation, RBM10 promotes the exclusion of an exon, called a cassette or alternative exon, from target pre-mRNAs, and less frequently, other alternative splicing events such as alternative 5ʹ-splice site selection. [9] [10] [11] [24] In the exon skipping process, RBM10 binds close to the 3ʹ- and 5ʹ-splice sites of cassette exons and interferes with the recognition and/or the pairing of the splice sites, thereby enhancing the pairing of the splice sites distal to the cassette exons, which ultimately leads to the exclusion of the exons together with the flanking upstream and downstream introns. [9] [10] [24]

The diversity of target RNAs bound by RBM10 in cells suggests that it is involved in various metabolic processes such as oxidative phosphorylation; pathways linked to cell proliferation, apoptosis, cell adhesion, and actin/cytoskeleton reorganization; and various diseases such as cancers and neurodegenerative diseases. [10] [16] [25] These data, together with the ubiquitous expression of RBM10, indicate that it is a fundamental cellular component participating in various cellular processes. In addition to alternative splicing regulation, RBM10 participates in other reactions. Some examples are polyadenylation of cardiac pre-mRNAs of anti-hypertrophy regulators, wherein it acts as a co-regulator of STAR-poly(A) polymerase, [26] stabilization of angiotensin II receptor mRNA by binding to its 3ʹ-UTR, [27] let-7g miRNA biogenesis through interaction with its precursor, [28] p53 stabilization by binding to its negative regulator, MDM2, [29] cell cycle arrest, [30] [31] and anti-viral reactions. [32]

RBM10 localizes to the nucleoplasm, where transcription and splicing occur, as well as in membrane-less nuclear compartments called S1-1 nuclear bodies (S1-1 NBs). [23] The numbers (ca. 10–40 per nucleus) and sizes (ca. 0.5 µm) of S1-1 NBs vary with the cell type and cellular conditions. When RNA polymerase II transcription decreases, RBM10 in the nucleoplasm is sequestered in S1-1 NBs, which become larger and spherical; when transcription is restored, RBM10 and the S1-1 NBs return to their initial states. [23] S1-1 NBs often overlap with nuclear speckles (also known as splicing speckles or interchromatin granule clusters), [23] [33] seemingly indicating a close functional relationship between these nuclear domains, i.e., alternative splicing regulation and splicing reaction.

Regulation

In females, most genes on one of the two X chromosomes are transcriptionally silenced by heterochromatin formation, and RBM10 is subjected to this X-inactivation. [6] [7] [34] In addition, there are mechanisms to control elevated cellular levels of RBM10. RBM10 auto-regulates its overexpressed pre-mRNA by alternative splicing to exclude exon 6 or 12, which generates a premature stop codon in the transcripts, leading to their degradation through nonsense-mediated mRNA decay (NMD). [14] When RNA polymerase II transcription decreases, RBM10 is sequestered in S1-1 NBs until transcription is restored. [23] In addition, RBM10 undergoes post-translational modifications: phosphorylation at many sites in response to various stimuli and changes in cellular conditions (UniProtKB-P98175; PhosphoSitePlus RBM10), as well as ubiquitylation, [35] [36] acetylation, [37] and methylation. [38] However, the molecular and biological significance of these various post-translational modifications of RBM10 is not well understood.

Clinical significance

Mutations in RBM10 are associated with various human diseases. The phenotypes caused by RBM10 mutations differ by the stages of development and affected tissues. Typical examples are TARP syndrome, an X-linked pleiotropic developmental malformation in neonates, [17] [22] and various cancers such as lung adenocarcinoma (LUAD) [18] and bladder carcinoma (BLCA) in adults. [19] These diseases are more common in males than in females. [39] [40] [41] One reason for this is the difference in the copy number of the RBM10 gene in a cell (one in male cells and two in female cells). Mutations in RBM10 occur throughout the molecule, and many of them are null mutations. TARP syndrome is generally pre- or postnatally lethal. [17] [42] [43] However, patients aged 11, 14, and 28 years have been reported to escape these null mutations. [44] [9] [45] RBM10 mutations have also been identified in other cancers [46] such as renal carcinomas, [47] [48] [49] pancreatic cancers, [50] [51] colorectal cancers, [52] [53] thyroid cancers, [54] [55] [56] breast cancers, [57] bile duct cancers, [58] [59] prostate cancer, [57] and brain tumor meningiomas and astroblastomas. [60] [61]

NUMB is the most studied downstream effector of RBM10. RBM10 promotes the skipping of exon 9 of the NUMB transcript, producing a NUMB isoform that causes ubiquitination followed by proteasomal degradation of the Notch receptor, and thereby inhibits the Notch signaling cell-proliferation pathway. [10] [62] [20] In various cancers, RBM10 mutations that inactivate or reduce its alternative splicing regulatory activity enhance the production of the exon 9–including NUMB isoform, which promotes cancer cell proliferation through the Notch pathway. [10] [63] [64]

RBM10 suppresses cell proliferation [10] [27] [63] [64] [65] [66] [29] and promotes apoptosis. [27] [64] [65] [29] [67] [68] Hence, it is generally regarded as a tumor suppressor. However, in certain cases, it may exert an opposite oncogenic function by acting as a tumor promoter or growth enhancer, [16] [69] [70] presumably due to the cellular contexts composed of different constituents and active pathways. A typical example of this is patients with pancreatic ductal adenocarcinoma (PDAC) having RBM10 mutations, who exhibit a survival rate remarkably higher than the general 5-year PDAC survival rate of less than 7–8%. [50] [71] [72]

Paralogs and splicing network

RBM5 and RBM6 are paralogs of RBM10. They were generated by gene duplications during genome evolution. They generally function as tumor suppressors, [10] [73] [74] [75] [76] [77] [78] [79] and their mutations are often identified in lung cancers. [21] RBM5, RBM6, and RBM10 regulate alternative splicing [10] [80] [81] and generally act on different RNAs; however, in certain cases, they act on the same subset of RNAs, likely producing synergistic or antagonistic effects. [10] There is a cross-regulation between RBM5 and RBM10; RBM10 lowers RBM5 transcript levels by alternative splicing–coupled NMD. [14] Furthermore, RBM10 perturbation (knockdown or overexpression) brings about splicing alterations in multiple splicing regulators, including RBM5, and also significantly influences the expression of other splicing regulators, including RBM10 itself. [9] [14] In addition, RBM10 primary transcripts are subjected to alternative splicing at several exons by unidentified splicing regulators, leading to the generation of multiple RBM10 isoforms. These data suggest the existence of an alternative splicing network formed by RBM5, RBM6, and RBM10, as well as other splicing regulators. [82] Studies on such networks are expected to promote our understanding of transcriptomic homeostasis regulated by splicing and the molecular and biological significance of RBM10 in cells.

RBM10 regulates hundreds of genes. [9] [10] [13] [14] [15] Further studies on the various RBM10-mediated processes and pathways may help elucidate the pathogenesis and progression of diseases caused by RBM10 mutations and the mechanisms of the antithetical actions of RBM10 as a tumor suppressor, and in certain cases, a tumor promoter, and provide clues for better treatment of the diseases.

Notes

Related Research Articles

<span class="mw-page-title-main">RNA splicing</span> Process in molecular biology

RNA splicing is a process in molecular biology where a newly-made precursor messenger RNA (pre-mRNA) transcript is transformed into a mature messenger RNA (mRNA). It works by removing all the introns and splicing back together exons. For nuclear-encoded genes, splicing occurs in the nucleus either during or immediately after transcription. For those eukaryotic genes that contain introns, splicing is usually needed to create an mRNA molecule that can be translated into protein. For many eukaryotic introns, splicing occurs in a series of reactions which are catalyzed by the spliceosome, a complex of small nuclear ribonucleoproteins (snRNPs). There exist self-splicing introns, that is, ribozymes that can catalyze their own excision from their parent RNA molecule. The process of transcription, splicing and translation is called gene expression, the central dogma of molecular biology.

<span class="mw-page-title-main">Alternative splicing</span> Process by which a gene can code for multiple proteins

Alternative splicing, or alternative RNA splicing, or differential splicing, is an alternative splicing process during gene expression that allows a single gene to code for multiple proteins. In this process, particular exons of a gene may be included within or excluded from the final, processed messenger RNA (mRNA) produced from that gene. This means the exons are joined in different combinations, leading to different (alternative) mRNA strands. Consequently, the proteins translated from alternatively spliced mRNAs usually contain differences in their amino acid sequence and, often, in their biological functions.

<span class="mw-page-title-main">Protein isoform</span> Forms of a protein produced from different genes

A protein isoform, or "protein variant", is a member of a set of highly similar proteins that originate from a single gene or gene family and are the result of genetic differences. While many perform the same or similar biological roles, some isoforms have unique functions. A set of protein isoforms may be formed from alternative splicings, variable promoter usage, or other post-transcriptional modifications of a single gene; post-translational modifications are generally not considered. Through RNA splicing mechanisms, mRNA has the ability to select different protein-coding segments (exons) of a gene, or even different parts of exons from RNA to form different mRNA sequences. Each unique sequence produces a specific form of a protein.

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

Clusterin is a 75-80 kDa disulfide-linked heterodimeric protein associated with the clearance of cellular debris and apoptosis. In humans, clusterin is encoded by the CLU gene on chromosome 8. CLU is a molecular chaperone responsible for aiding protein folding of secreted proteins, and its three isoforms have been differentially implicated in pro- or antiapoptotic processes. Through this function, CLU is involved in many diseases related to oxidative stress, including neurodegenerative diseases, cancers, inflammatory diseases, and aging.

<span class="mw-page-title-main">L1 (protein)</span> Mammalian protein found in Homo sapiens

L1, also known as L1CAM, is a transmembrane protein member of the L1 protein family, encoded by the L1CAM gene. This protein, of 200-220 kDa, is a neuronal cell adhesion molecule with a strong implication in cell migration, adhesion, neurite outgrowth, myelination and neuronal differentiation. It also plays a key role in treatment-resistant cancers due to its function. It was first identified in 1984 by M. Schachner who found the protein in post-mitotic mice neurons.

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

Tropomyosin is a two-stranded alpha-helical, coiled coil protein found in many animal and fungal cells. In animals, it is an important component of the muscular system which works in conjunction with troponin to regulate muscle contraction. It is present in smooth and striated muscle tissues, which can be found in various organs and body systems, including the heart, blood vessels, respiratory system, and digestive system. In fungi, tropomyosin is found in cell walls and helps maintain the structural integrity of cells.

RNA-binding proteins are proteins that bind to the double or single stranded RNA in cells and participate in forming ribonucleoprotein complexes. RBPs contain various structural motifs, such as RNA recognition motif (RRM), dsRNA binding domain, zinc finger and others. They are cytoplasmic and nuclear proteins. However, since most mature RNA is exported from the nucleus relatively quickly, most RBPs in the nucleus exist as complexes of protein and pre-mRNA called heterogeneous ribonucleoprotein particles (hnRNPs). RBPs have crucial roles in various cellular processes such as: cellular function, transport and localization. They especially play a major role in post-transcriptional control of RNAs, such as: splicing, polyadenylation, mRNA stabilization, mRNA localization and translation. Eukaryotic cells express diverse RBPs with unique RNA-binding activity and protein–protein interaction. According to the Eukaryotic RBP Database (EuRBPDB), there are 2961 genes encoding RBPs in humans. During evolution, the diversity of RBPs greatly increased with the increase in the number of introns. Diversity enabled eukaryotic cells to utilize RNA exons in various arrangements, giving rise to a unique RNP (ribonucleoprotein) for each RNA. Although RBPs have a crucial role in post-transcriptional regulation in gene expression, relatively few RBPs have been studied systematically.It has now become clear that RNA–RBP interactions play important roles in many biological processes among organisms.

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

T-box transcription factor T, also known as Brachyury protein, is encoded for in humans by the TBXT gene. Brachyury functions as a transcription factor within the T-box family of genes. Brachyury homologs have been found in all bilaterian animals that have been screened, as well as the freshwater cnidarian Hydra.

<span class="mw-page-title-main">PAX3</span> Paired box gene 3

The PAX3 gene encodes a member of the paired box or PAX family of transcription factors. The PAX family consists of nine human (PAX1-PAX9) and nine mouse (Pax1-Pax9) members arranged into four subfamilies. Human PAX3 and mouse Pax3 are present in a subfamily along with the highly homologous human PAX7 and mouse Pax7 genes. The human PAX3 gene is located in the 2q36.1 chromosomal region, and contains 10 exons within a 100 kb region.

<span class="mw-page-title-main">Neurofibromin 1</span> Mammalian protein found in Homo sapiens

neurofibromatosis 1 (NF1) is a gene in humans that is located on chromosome 17. NF1 codes for neurofibromin, a GTPase-activating protein that negatively regulates RAS/MAPK pathway activity by accelerating the hydrolysis of Ras-bound GTP. NF1 has a high mutation rate and mutations in NF1 can alter cellular growth control, and neural development, resulting in neurofibromatosis type 1. Symptoms of NF1 include disfiguring cutaneous neurofibromas (CNF), café au lait pigment spots, plexiform neurofibromas (PN), skeletal defects, optic nerve gliomas, life-threatening malignant peripheral nerve sheath tumors (MPNST), pheochromocytoma, attention deficits, learning deficits and other cognitive disabilities.

p16 Mammalian protein found in Homo sapiens

p16, is a protein that slows cell division by slowing the progression of the cell cycle from the G1 phase to the S phase, thereby acting as a tumor suppressor. It is encoded by the CDKN2A gene. A deletion in this gene can result in insufficient or non-functional p16, accelerating the cell cycle and resulting in many types of cancer.

<span class="mw-page-title-main">TP63</span> Protein-coding gene in the species Homo sapiens

Tumor protein p63, typically referred to as p63, also known as transformation-related protein 63 is a protein that in humans is encoded by the TP63 gene.

<span class="mw-page-title-main">TIA1</span> Mammalian protein found in Homo sapiens

TIA1 or Tia1 cytotoxic granule-associated rna binding protein is a 3'UTR mRNA binding protein that can bind the 5'TOP sequence of 5'TOP mRNAs. It is associated with programmed cell death (apoptosis) and regulates alternative splicing of the gene encoding the Fas receptor, an apoptosis-promoting protein. Under stress conditions, TIA1 localizes to cellular RNA-protein conglomerations called stress granules. It is encoded by the TIA1 gene.

<span class="mw-page-title-main">Wilms tumor protein</span> Transcription factor gene involved in the urogenital system

Wilms tumor protein (WT33) is a protein that in humans is encoded by the WT1 gene on chromosome 11p.

<span class="mw-page-title-main">HNRNPA2B1</span> Protein-coding gene in the species Homo sapiens

Heterogeneous nuclear ribonucleoproteins A2/B1 is a protein that in humans is encoded by the HNRNPA2B1 gene.

<span class="mw-page-title-main">STK11</span> Protein-coding gene in the species Homo sapiens

Serine/threonine kinase 11 (STK11) also known as liver kinase B1 (LKB1) or renal carcinoma antigen NY-REN-19 is a protein kinase that in humans is encoded by the STK11 gene.

<span class="mw-page-title-main">RBM4</span> Protein-coding gene in the species Homo sapiens

RNA-binding protein 4 is a protein that in humans is encoded by the RBM4 gene.

<span class="mw-page-title-main">EIF4A3</span> Protein-coding gene in the species Homo sapiens

Eukaryotic initiation factor 4A-III is a protein that in humans is encoded by the EIF4A3 gene.

<span class="mw-page-title-main">MALAT1</span>

MALAT 1 also known as NEAT2 is a large, infrequently spliced non-coding RNA, which is highly conserved amongst mammals and highly expressed in the nucleus. MALAT1 was identified in multiple types of physiological processes, such as alternative splicing, nuclear organization, epigenetic modulating of gene expression, and a number of evidences indicate that MALAT1 also closely relate to various pathological processes, ranging from diabetes complications to cancers. It regulates the expression of metastasis-associated genes. It also positively regulates cell motility via the transcriptional and/or post-transcriptional regulation of motility-related genes. MALAT1 may play a role in temperature-dependent sex determination in the Red-eared slider turtle.

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

A minigene is a minimal gene fragment that includes an exon and the control regions necessary for the gene to express itself in the same way as a wild type gene fragment. This is a minigene in its most basic sense. More complex minigenes can be constructed containing multiple exons and intron(s). Minigenes provide a valuable tool for researchers evaluating splicing patterns both in vivo and in vitro biochemically assessed experiments. Specifically, minigenes are used as splice reporter vectors and act as a probe to determine which factors are important in splicing outcomes. They can be constructed to test the way both cis-regulatory elements and trans-regulatory elements affect gene expression.

References

  1. 1 2 3 GRCh38: Ensembl release 89: ENSG00000182872 - Ensembl, May 2017
  2. 1 2 3 GRCm38: Ensembl release 89: ENSMUSG00000031060 - Ensembl, May 2017
  3. "Human PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  4. "Mouse PubMed Reference:". National Center for Biotechnology Information, U.S. National Library of Medicine.
  5. Nagase T, Seki N, Tanaka A, Ishikawa K, Nomura N (August 1995). "Prediction of the coding sequences of unidentified human genes. IV. The coding sequences of 40 new genes (KIAA0121-KIAA0160) deduced by analysis of cDNA clones from human cell line KG-1". DNA Research. 2 (4): 167–74, 199–210. doi: 10.1093/dnares/2.4.167 . PMID   8590280.
  6. 1 2 3 Coleman MP, Ambrose HJ, Carrel L, Németh AH, Willard HF, Davies KE (January 1996). "A novel gene, DXS8237E, lies within 20 kb upstream of UBE1 in Xp11.23 and has a different X inactivation status". Genomics. 31 (1): 135–8. doi:10.1006/geno.1996.0022. PMID   8808293.
  7. 1 2 3 Thiselton DL, McDowall J, Brandau O, Ramser J, d'Esposito F, Bhattacharya SS, et al. (April 2002). "An integrated, functionally annotated gene map of the DXS8026-ELK1 interval on human Xp11.3-Xp11.23: potential hotspot for neurogenetic disorders". Genomics. 79 (4): 560–72. doi:10.1006/geno.2002.6733. PMID   11944989.
  8. Inoue A, Takahashi KP, Kimura M, Watanabe T, Morisawa S (August 1996). "Molecular cloning of a RNA binding protein, S1-1". Nucleic Acids Research. 24 (15): 2990–7. doi:10.1093/nar/24.15.2990. PMC   146028 . PMID   8760884.
  9. 1 2 3 4 5 6 7 Wang Y, Gogol-Döring A, Hu H, Fröhler S, Ma Y, Jens M, et al. (September 2013). "Integrative analysis revealed the molecular mechanism underlying RBM10-mediated splicing regulation". EMBO Molecular Medicine. 5 (9): 1431–42. doi:10.1002/emmm.201302663. PMC   3799496 . PMID   24000153.
  10. 1 2 3 4 5 6 7 8 9 10 11 12 13 Bechara EG, Sebestyén E, Bernardis I, Eyras E, Valcárcel J (December 2013). "RBM5, 6, and 10 differentially regulate NUMB alternative splicing to control cancer cell proliferation". Molecular Cell. 52 (5): 720–33. doi: 10.1016/j.molcel.2013.11.010 . PMID   24332178.
  11. 1 2 Inoue A, Yamamoto N, Kimura M, Nishio K, Yamane H, Nakajima K (March 2014). "RBM10 regulates alternative splicing". FEBS Letters. 588 (6): 942–7. doi: 10.1016/j.febslet.2014.01.052 . PMID   24530524. S2CID   10303057.
  12. Yang X, Coulombe-Huntington J, Kang S, Sheynkman GM, Hao T, Richardson A, et al. (February 2016). "Widespread Expansion of Protein Interaction Capabilities by Alternative Splicing". Cell. 164 (4): 805–17. doi:10.1016/j.cell.2016.01.029. PMC   4882190 . PMID   26871637.
  13. 1 2 Sutherland LC, Thibault P, Durand M, Lapointe E, Knee JM, Beauvais A, et al. (July 2017). "Splicing arrays reveal novel RBM10 targets, including SMN2 pre-mRNA". BMC Molecular Biology. 18 (1): 19. doi: 10.1186/s12867-017-0096-x . PMC   5520337 . PMID   28728573.
  14. 1 2 3 4 5 Sun Y, Bao Y, Han W, Song F, Shen X, Zhao J, et al. (August 2017). "Autoregulation of RBM10 and cross-regulation of RBM10/RBM5 via alternative splicing-coupled nonsense-mediated decay". Nucleic Acids Research. 45 (14): 8524–8540. doi:10.1093/nar/gkx508. PMC   5737846 . PMID   28586478.
  15. 1 2 Collins KM, Kainov YA, Christodolou E, Ray D, Morris Q, Hughes T, et al. (June 2017). "An RRM-ZnF RNA recognition module targets RBM10 to exonic sequences to promote exon exclusion". Nucleic Acids Research. 45 (11): 6761–6774. doi:10.1093/nar/gkx225. PMC   5499739 . PMID   28379442.
  16. 1 2 3 Loiselle JJ, Roy JG, Sutherland LC (2017). "RBM10 promotes transformation-associated processes in small cell lung cancer and is directly regulated by RBM5". PLOS ONE. 12 (6): e0180258. Bibcode:2017PLoSO..1280258L. doi: 10.1371/journal.pone.0180258 . PMC   5491171 . PMID   28662214.
  17. 1 2 3 4 Johnston JJ, Teer JK, Cherukuri PF, Hansen NF, Loftus SK, Chong K, et al. (May 2010). "Massively parallel sequencing of exons on the X chromosome identifies RBM10 as the gene that causes a syndromic form of cleft palate". American Journal of Human Genetics. 86 (5): 743–8. doi:10.1016/j.ajhg.2010.04.007. PMC   2868995 . PMID   20451169.
  18. 1 2 3 Imielinski M, Berger AH, Hammerman PS, Hernandez B, Pugh TJ, Hodis E, et al. (September 2012). "Mapping the hallmarks of lung adenocarcinoma with massively parallel sequencing". Cell. 150 (6): 1107–20. doi:10.1016/j.cell.2012.08.029. PMC   3557932 . PMID   22980975.
  19. 1 2 3 Seiler M, Peng S, Agrawal AA, Palacino J, Teng T, Zhu P, et al. (April 2018). "Somatic Mutational Landscape of Splicing Factor Genes and Their Functional Consequences across 33 Cancer Types". Cell Reports. 23 (1): 282–296.e4. doi:10.1016/j.celrep.2018.01.088. PMC   5933844 . PMID   29617667.
  20. 1 2 Cieply B, Carstens RP (2015). "Functional roles of alternative splicing factors in human disease". Wiley Interdisciplinary Reviews. RNA. 6 (3): 311–26. doi:10.1002/wrna.1276. PMC   4671264 . PMID   25630614.
  21. 1 2 Coomer AO, Black F, Greystoke A, Munkley J, Elliott DJ (2019). "Alternative splicing in lung cancer". Biochimica et Biophysica Acta (BBA) - Gene Regulatory Mechanisms. 1862 (11–12): 194388. doi:10.1016/j.bbagrm.2019.05.006. PMID   31152916. S2CID   173188639.
  22. 1 2 3 Gorlin RJ, Cervenka J, Anderson RC, Sauk JJ, Bevis WD (February 1970). "Robin's syndrome. A probably X-linked recessive subvariety exhibiting persistence of left superior vena cava and atrial septal defect". American Journal of Diseases of Children. 119 (2): 176–8. doi:10.1001/archpedi.1970.02100050178020. PMID   5410571.
  23. 1 2 3 4 5 Inoue A, Tsugawa K, Tokunaga K, Takahashi KP, Uni S, Kimura M, et al. (September 2008). "S1-1 nuclear domains: characterization and dynamics as a function of transcriptional activity". Biology of the Cell. 100 (9): 523–35. doi:10.1042/BC20070142. PMID   18315527. S2CID   9893063.
  24. 1 2 Zheng S, Damoiseaux R, Chen L, Black DL (June 2013). "A broadly applicable high-throughput screening strategy identifies new regulators of Dlg4 (Psd-95) alternative splicing". Genome Research. 23 (6): 998–1007. doi:10.1101/gr.147546.112. PMC   3668367 . PMID   23636947.
  25. Lim J, Hao T, Shaw C, Patel AJ, Szabó G, Rual JF, et al. (May 2006). "A protein-protein interaction network for human inherited ataxias and disorders of Purkinje cell degeneration". Cell. 125 (4): 801–14. doi: 10.1016/j.cell.2006.03.032 . PMID   16713569. S2CID   13709685.
  26. Mohan N, Kumar V, Kandala DT, Kartha CC, Laishram RS (September 2018). "A Splicing-Independent Function of RBM10 Controls Specific 3' UTR Processing to Regulate Cardiac Hypertrophy". Cell Reports. 24 (13): 3539–3553. doi: 10.1016/j.celrep.2018.08.077 . PMID   30257214.
  27. 1 2 3 Mueller CF, Berger A, Zimmer S, Tiyerili V, Nickenig G (August 2009). "The heterogenous nuclear riboprotein S1-1 regulates AT1 receptor gene expression via transcriptional and posttranscriptional mechanisms". Archives of Biochemistry and Biophysics. 488 (1): 76–82. doi:10.1016/j.abb.2009.06.002. PMID   19508861.
  28. Treiber T, Treiber N, Plessmann U, Harlander S, Daiß JL, Eichner N, et al. (April 2017). "A Compendium of RNA-Binding Proteins that Regulate MicroRNA Biogenesis". Molecular Cell. 66 (2): 270–284.e13. doi: 10.1016/j.molcel.2017.03.014 . PMID   28431233.
  29. 1 2 3 Jung JH, Lee H, Cao B, Liao P, Zeng SX, Lu H (January 2020). "RNA-binding motif protein 10 induces apoptosis and suppresses proliferation by activating p53". Oncogene. 39 (5): 1031–1040. doi:10.1038/s41388-019-1034-9. PMC   6994357 . PMID   31591476.
  30. Guan G, Li R, Tang W, Liu T, Su Z, Wang Y, et al. (March 2017). "Expression of RNA-binding motif 10 is associated with advanced tumor stage and malignant behaviors of lung adenocarcinoma cancer cells". Tumour Biology. 39 (3): 1010428317691740. doi: 10.1177/1010428317691740 . PMID   28347232. S2CID   206612545.
  31. Kunimoto H, Inoue A, Kojima H, Yang J, Zhao H, Tsuruta D, Nakajima K (February 2020). "RBM10 regulates centriole duplication in HepG2 cells by ectopically assembling PLK4-STIL complexes in the nucleus". Genes to Cells. 25 (2): 100–110. doi: 10.1111/gtc.12741 . PMID   31820547. S2CID   209165475.
  32. Pozzi B, Bragado L, Mammi P, Torti MF, Gaioli N, Gebhard LG, et al. (July 2020). "Dengue virus targets RBM10 deregulating host cell splicing and innate immune response". Nucleic Acids Research. 48 (12): 6824–6838. doi:10.1093/nar/gkaa340. PMC   7337517 . PMID   32432721.
  33. Salichs E, Ledda A, Mularoni L, Albà MM, de la Luna S (March 2009). "Genome-wide analysis of histidine repeats reveals their role in the localization of human proteins to the nuclear speckles compartment". PLOS Genetics. 5 (3): e1000397. doi: 10.1371/journal.pgen.1000397 . PMC   2644819 . PMID   19266028.
  34. Goto Y, Kimura H (December 2009). "Inactive X chromosome-specific histone H3 modifications and CpG hypomethylation flank a chromatin boundary between an X-inactivated and an escape gene". Nucleic Acids Research. 37 (22): 7416–28. doi:10.1093/nar/gkp860. PMC   2794193 . PMID   19843608.
  35. Stes E, Laga M, Walton A, Samyn N, Timmerman E, De Smet I, et al. (June 2014). "A COFRADIC protocol to study protein ubiquitination". Journal of Proteome Research. 13 (6): 3107–13. doi:10.1021/pr4012443. PMID   24816145.
  36. Akimov V, Barrio-Hernandez I, Hansen SV, Hallenborg P, Pedersen AK, Bekker-Jensen DB, et al. (July 2018). "UbiSite approach for comprehensive mapping of lysine and N-terminal ubiquitination sites". Nature Structural & Molecular Biology. 25 (7): 631–640. doi:10.1038/s41594-018-0084-y. PMID   29967540. S2CID   49559977.
  37. Choudhary C, Kumar C, Gnad F, Nielsen ML, Rehman M, Walther TC, et al. (August 2009). "Lysine acetylation targets protein complexes and co-regulates major cellular functions". Science. 325 (5942): 834–40. Bibcode:2009Sci...325..834C. doi: 10.1126/science.1175371 . PMID   19608861. S2CID   206520776.
  38. Guo A, Gu H, Zhou J, Mulhern D, Wang Y, Lee KA, et al. (January 2014). "Immunoaffinity enrichment and mass spectrometry analysis of protein methylation". Molecular & Cellular Proteomics. 13 (1): 372–87. doi: 10.1074/mcp.O113.027870 . PMC   3879628 . PMID   24129315.
  39. Cancer Genome Atlas Research Network (July 2014). "Comprehensive molecular profiling of lung adenocarcinoma". Nature. 511 (7511): 543–50. Bibcode:2014Natur.511..543T. doi:10.1038/nature13385. PMC   4231481 . PMID   25079552.
  40. Yuan Y, Liu L, Chen H, Wang Y, Xu Y, Mao H, et al. (May 2016). "Comprehensive Characterization of Molecular Differences in Cancer between Male and Female Patients". Cancer Cell. 29 (5): 711–722. doi:10.1016/j.ccell.2016.04.001. PMC   4864951 . PMID   27165743.
  41. Yin LL, Wen XM, Li M, Xu YM, Zhao XF, Li J, Wang XW (November 2018). "A gene mutation in RNA-binding protein 10 is associated with lung adenocarcinoma progression and poor prognosis". Oncology Letters. 16 (5): 6283–6292. doi:10.3892/ol.2018.9496. PMC   6202477 . PMID   30405763.
  42. Powis Z, Hart A, Cherny S, Petrik I, Palmaer E, Tang S, Jones C (2 June 2017). "Clinical diagnostic exome evaluation for an infant with a lethal disorder: genetic diagnosis of TARP syndrome and expansion of the phenotype in a patient with a newly reported RBM10 alteration". BMC Medical Genetics. 18 (1): 60. doi: 10.1186/s12881-017-0426-3 . PMC   5455125 . PMID   28577551.
  43. Gripp KW, Hopkins E, Johnston JJ, Krause C, Dobyns WB, Biesecker LG (October 2011). "Long-term survival in TARP syndrome and confirmation of RBM10 as the disease-causing gene". American Journal of Medical Genetics. Part A. 155A (10): 2516–20. doi:10.1002/ajmg.a.34190. PMC   3183328 . PMID   21910224.
  44. Niceta M, Barresi S, Pantaleoni F, Capolino R, Dentici ML, Ciolfi A, et al. (June 2019). "TARP syndrome: Long-term survival, anatomic patterns of congenital heart defects, differential diagnosis and pathogenetic considerations". European Journal of Medical Genetics. 62 (6): 103534. doi: 10.1016/j.ejmg.2018.09.001 . PMID   30189253.
  45. Højland AT, Lolas I, Okkels H, Lautrup CK, Diness BR, Petersen MB, Nielsen IK (December 2018). "First reported adult patient with TARP syndrome: A case report". American Journal of Medical Genetics. Part A. 176 (12): 2915–2918. doi:10.1002/ajmg.a.40638. PMC   6587983 . PMID   30462380.
  46. Loiselle JJ, Sutherland LC (May 2018). "RBM10: Harmful or helpful-many factors to consider". Journal of Cellular Biochemistry. 119 (5): 3809–3818. doi:10.1002/jcb.26644. PMC   5901003 . PMID   29274279.
  47. Xia QY, Wang XT, Zhan XM, Tan X, Chen H, Liu Y, et al. (May 2017). "Xp11 Translocation Renal Cell Carcinomas (RCCs) With RBM10-TFE3 Gene Fusion Demonstrating Melanotic Features and Overlapping Morphology With t(6;11) RCC: Interest and Diagnostic Pitfall in Detecting a Paracentric Inversion of TFE3". The American Journal of Surgical Pathology. 41 (5): 663–676. doi:10.1097/PAS.0000000000000837. PMID   28288037. S2CID   205918230.
  48. Argani P, Zhang L, Reuter VE, Tickoo SK, Antonescu CR (May 2017). "RBM10-TFE3 Renal Cell Carcinoma: A Potential Diagnostic Pitfall Due to Cryptic Intrachromosomal Xp11.2 Inversion Resulting in False-negative TFE3 FISH". The American Journal of Surgical Pathology. 41 (5): 655–662. doi:10.1097/PAS.0000000000000835. PMC   5391276 . PMID   28296677.
  49. Kato I, Furuya M, Baba M, Kameda Y, Yasuda M, Nishimoto K, et al. (August 2019). "RBM10-TFE3 renal cell carcinoma characterised by paracentric inversion with consistent closely split signals in break-apart fluorescence in-situ hybridisation: study of 10 cases and a literature review". Histopathology. 75 (2): 254–265. doi:10.1111/his.13866. PMID   30908700. S2CID   85516169.
  50. 1 2 Witkiewicz AK, McMillan EA, Balaji U, Baek G, Lin WC, Mansour J, et al. (April 2015). "Whole-exome sequencing of pancreatic cancer defines genetic diversity and therapeutic targets". Nature Communications. 6: 6744. Bibcode:2015NatCo...6.6744W. doi:10.1038/ncomms7744. PMC   4403382 . PMID   25855536.
  51. Furukawa T, Kuboki Y, Tanji E, Yoshida S, Hatori T, Yamamoto M, et al. (2011). "Whole-exome sequencing uncovers frequent GNAS mutations in intraductal papillary mucinous neoplasms of the pancreas". Scientific Reports. 1: 161. Bibcode:2011NatSR...1E.161F. doi:10.1038/srep00161. PMC   3240977 . PMID   22355676.
  52. Giannakis M, Mu XJ, Shukla SA, Qian ZR, Cohen O, Nishihara R, et al. (April 2016). "Genomic Correlates of Immune-Cell Infiltrates in Colorectal Carcinoma". Cell Reports. 15 (4): 857–865. doi:10.1016/j.celrep.2016.03.075. PMC   4850357 . PMID   27149842.
  53. Lawrence MS, Stojanov P, Mermel CH, Robinson JT, Garraway LA, Golub TR, et al. (January 2014). "Discovery and saturation analysis of cancer genes across 21 tumour types". Nature. 505 (7484): 495–501. Bibcode:2014Natur.505..495L. doi:10.1038/nature12912. PMC   4048962 . PMID   24390350.
  54. Ibrahimpasic T, Xu B, Landa I, Dogan S, Middha S, Seshan V, et al. (October 2017). "RBM10 as Novel Thyroid Cancer Genes Associated with Tumor Virulence". Clinical Cancer Research. 23 (19): 5970–5980. doi:10.1158/1078-0432.CCR-17-1183. PMC   5626586 . PMID   28634282.
  55. Antonello ZA, Hsu N, Bhasin M, Roti G, Joshi M, Van Hummelen P, et al. (October 2017). "V600E". Oncotarget. 8 (49): 84743–84760. doi:10.18632/oncotarget.21262. PMC   5689570 . PMID   29156680.
  56. Ibrahimpasic T, Ghossein R, Shah JP, Ganly I (March 2019). "Poorly Differentiated Carcinoma of the Thyroid Gland: Current Status and Future Prospects". Thyroid. 29 (3): 311–321. doi:10.1089/thy.2018.0509. PMC   6437626 . PMID   30747050.
  57. 1 2 Kan Z, Jaiswal BS, Stinson J, Janakiraman V, Bhatt D, Stern HM, et al. (August 2010). "Diverse somatic mutation patterns and pathway alterations in human cancers". Nature. 466 (7308): 869–73. Bibcode:2010Natur.466..869K. doi:10.1038/nature09208. PMC   3026267 . PMID   20668451.
  58. Tian W, Hu W, Shi X, Liu P, Ma X, Zhao W, et al. (April 2020). "Comprehensive genomic profile of cholangiocarcinomas in China". Oncology Letters. 19 (4): 3101–3110. doi:10.3892/ol.2020.11429. PMC   7074170 . PMID   32256810.
  59. Schwab ME, Song H, Mattis A, Phelps A, Vu LT, Huang FW, Nijagal A (March 2020). "De novo somatic mutations and KRAS amplification are associated with cholangiocarcinoma in a patient with a history of choledochal cyst". Journal of Pediatric Surgery. 55 (12): 2657–2661. doi: 10.1016/j.jpedsurg.2020.03.008 . PMC   7942710 . PMID   32295706.
  60. Juratli TA, McCabe D, Nayyar N, Williams EA, Silverman IM, Tummala SS, et al. (November 2018). "DMD genomic deletions characterize a subset of progressive/higher-grade meningiomas with poor outcome". Acta Neuropathologica. 136 (5): 779–792. doi:10.1007/s00401-018-1899-7. PMID   30123936. S2CID   52039057.
  61. Majd NK, Metrus NR, Santos-Pinheiro F, Trevino CR, Fuller GN, Huse JT, et al. (February 2019). "RBM10 truncation in astroblastoma in a patient with history of mandibular ameloblastoma: A case report". Cancer Genetics. 231–232: 41–45. doi:10.1016/j.cancergen.2019.01.001. PMID   30803556. S2CID   73477150.
  62. Misquitta-Ali CM, Cheng E, O'Hanlon D, Liu N, McGlade CJ, Tsao MS, Blencowe BJ (January 2011). "Global profiling and molecular characterization of alternative splicing events misregulated in lung cancer". Molecular and Cellular Biology. 31 (1): 138–50. doi:10.1128/MCB.00709-10. PMC   3019846 . PMID   21041478.
  63. 1 2 Hernández J, Bechara E, Schlesinger D, Delgado J, Serrano L, Valcárcel J (2016). "Tumor suppressor properties of the splicing regulatory factor RBM10". RNA Biology. 13 (4): 466–72. doi:10.1080/15476286.2016.1144004. PMC   4841610 . PMID   26853560.
  64. 1 2 3 Zhao J, Sun Y, Huang Y, Song F, Huang Z, Bao Y, et al. (January 2017). "Functional analysis reveals that RBM10 mutations contribute to lung adenocarcinoma pathogenesis by deregulating splicing". Scientific Reports. 7: 40488. Bibcode:2017NatSR...740488Z. doi:10.1038/srep40488. PMC   5238425 . PMID   28091594.
  65. 1 2 Han LP, Wang CP, Han SL (October 2018). "Overexpression of RBM10 induces osteosarcoma cell apoptosis and inhibits cell proliferation and migration". Médecine/Sciences. 34 Focus issue F1: 81–86. doi: 10.1051/medsci/201834f114 . PMID   30403180.
  66. Jin X, Di X, Wang R, Ma H, Tian C, Zhao M, et al. (June 2019). "RBM10 inhibits cell proliferation of lung adenocarcinoma via RAP1/AKT/CREB signalling pathway". Journal of Cellular and Molecular Medicine. 23 (6): 3897–3904. doi:10.1111/jcmm.14263. PMC   6533519 . PMID   30955253.
  67. Sutherland LC, Rintala-Maki ND, White RD, Morin CD (January 2005). "RNA binding motif (RBM) proteins: a novel family of apoptosis modulators?". Journal of Cellular Biochemistry. 94 (1): 5–24. doi:10.1002/jcb.20204. PMID   15514923. S2CID   26344717.
  68. Wang K, Bacon ML, Tessier JJ, Rintala-Maki ND, Tang V, Sutherland LC (2012). "RBM10 Modulates Apoptosis and Influences TNF-α Gene Expression". Journal of Cell Death. 5: 1–19. doi:10.4137/JCD.S9073. PMC   4583097 . PMID   26446321.
  69. Rodor J, FitzPatrick DR, Eyras E, Cáceres JF (January 2017). "The RNA-binding landscape of RBM10 and its role in alternative splicing regulation in models of mouse early development". RNA Biology. 14 (1): 45–57. doi:10.1080/15476286.2016.1247148. PMC   5270529 . PMID   27763814.
  70. Sun X, Jia M, Sun W, Feng L, Gu C, Wu T (February 2019). "Functional role of RBM10 in lung adenocarcinoma proliferation". International Journal of Oncology. 54 (2): 467–478. doi:10.3892/ijo.2018.4643. PMC   6317669 . PMID   30483773.
  71. Balachandran VP, Łuksza M, Zhao JN, Makarov V, Moral JA, Remark R, et al. (November 2017). "Identification of unique neoantigen qualities in long-term survivors of pancreatic cancer". Nature. 551 (7681): 512–516. Bibcode:2017Natur.551..512B. doi:10.1038/nature24462. PMC   6145146 . PMID   29132146.
  72. Siegel RL, Miller KD, Jemal A (January 2018). "Cancer statistics, 2018". CA: A Cancer Journal for Clinicians. 68 (1): 7–30. doi: 10.3322/caac.21442 . PMID   29313949.
  73. Mourtada-Maarabouni M, Williams GT (July 2002). "RBM5/LUCA-15--tumour suppression by control of apoptosis and the cell cycle?". TheScientificWorldJournal. 2: 1885–90. doi: 10.1100/tsw.2002.859 . PMC   6009235 . PMID   12920317.
  74. Oh JJ, Razfar A, Delgado I, Reed RA, Malkina A, Boctor B, Slamon DJ (April 2006). "3p21.3 tumor suppressor gene H37/Luca15/RBM5 inhibits growth of human lung cancer cells through cell cycle arrest and apoptosis". Cancer Research. 66 (7): 3419–27. doi: 10.1158/0008-5472.CAN-05-1667 . PMID   16585163.
  75. Fushimi K, Ray P, Kar A, Wang L, Sutherland LC, Wu JY (October 2008). "Up-regulation of the proapoptotic caspase 2 splicing isoform by a candidate tumor suppressor, RBM5". Proceedings of the National Academy of Sciences of the United States of America. 105 (41): 15708–13. Bibcode:2008PNAS..10515708F. doi: 10.1073/pnas.0805569105 . PMC   2572934 . PMID   18840686.
  76. Bonnal S, Martínez C, Förch P, Bachi A, Wilm M, Valcárcel J (October 2008). "RBM5/Luca-15/H37 regulates Fas alternative splice site pairing after exon definition". Molecular Cell. 32 (1): 81–95. doi: 10.1016/j.molcel.2008.08.008 . PMID   18851835.
  77. Sutherland LC, Wang K, Robinson AG (March 2010). "RBM5 as a putative tumor suppressor gene for lung cancer". Journal of Thoracic Oncology. 5 (3): 294–8. doi: 10.1097/JTO.0b013e3181c6e330 . PMID   20186023.
  78. Jamsai D, Watkins DN, O'Connor AE, Merriner DJ, Gursoy S, Bird AD, et al. (November 2017). "In vivo evidence that RBM5 is a tumour suppressor in the lung". Scientific Reports. 7 (1): 16323. Bibcode:2017NatSR...716323J. doi:10.1038/s41598-017-15874-9. PMC   5701194 . PMID   29176597.
  79. Wang Q, Wang F, Zhong W, Ling H, Wang J, Cui J, et al. (May 2019). "RNA-binding protein RBM6 as a tumor suppressor gene represses the growth and progression in laryngocarcinoma". Gene. 697: 26–34. doi:10.1016/j.gene.2019.02.025. PMID   30772516. S2CID   73456531.
  80. Deckert J, Hartmuth K, Boehringer D, Behzadnia N, Will CL, Kastner B, et al. (July 2006). "Protein composition and electron microscopy structure of affinity-purified human spliceosomal B complexes isolated under physiological conditions". Molecular and Cellular Biology. 26 (14): 5528–43. doi:10.1128/MCB.00582-06. PMC   1592722 . PMID   16809785.
  81. Papasaikas P, Tejedor JR, Vigevani L, Valcárcel J (January 2015). "Functional splicing network reveals extensive regulatory potential of the core spliceosomal machinery". Molecular Cell. 57 (1): 7–22. doi: 10.1016/j.molcel.2014.10.030 . PMID   25482510. S2CID   11133534.
  82. Ule J, Blencowe BJ (October 2019). "Alternative Splicing Regulatory Networks: Functions, Mechanisms, and Evolution". Molecular Cell. 76 (2): 329–345. doi: 10.1016/j.molcel.2019.09.017 . PMID   31626751.

Further reading