Open Access

Aberrant mRNA splicing generates oncogenic RNA isoforms and contributes to the development and progression of cholangiocarcinoma (Review)

  • Authors:
    • Juthamas Yosudjai
    • Sopit Wongkham
    • Siwanon Jirawatnotai
    • Worasak Kaewkong
  • View Affiliations

  • Published online on: January 25, 2019     https://doi.org/10.3892/br.2019.1188
  • Pages: 147-155
  • Copyright: © Yosudjai et al. This is an open access article distributed under the terms of Creative Commons Attribution License.

Metrics: Total Views: 0 (Spandidos Publications: | PMC Statistics: )
Total PDF Downloads: 0 (Spandidos Publications: | PMC Statistics: )


Abstract

Cholangiocarcinoma is a lethal biliary cancer, with an unclear molecular pathogenesis. Alternative splicing is a post‑transcriptional modification that generates mature mRNAs, which are subsequently translated into proteins. Aberrant alternative splicing has been reported to serve a role in tumor initiation, maintenance and metastasis in several types of human cancer, including cholangiocarcinoma. In this review, the aberrant splicing of genes and the functional contributions of the spliced genes, in the carcinogenesis, progression and aggressiveness of cholangiocarcinoma are summarized. In addition, factors that influence this aberrant splicing that may be relevant as therapeutic targets or prognosis markers for cholangiocarcinoma are discussed.

1. Introduction

Cholangiocarcinoma (CCA), is a malignant tumor that arises from the biliary epithelial tissue and is highly aggressive, with no effective pharmacological treatment available. This cancer has a poor prognosis and a high mortality rate (1). The highest worldwide incidence of CCA is found in the North and Northeast of Thailand, at ~85 cases per 100,000 individuals per year (2). The major predisposing factors for CCA in Asia are infection by the liver fluke, Opisthorchis viverrini (3,4) and exposure to groups of food-borne carcinogens, especially N-nitrosodimethylamine compounds identified in grilled or fermented foods (5). The only effective treatment for the disease is surgery. For patients who are not eligible for surgical therapy, gemcitabine- or 5-fluoro-uracil (FU)-based treatments are given. These are largely ineffective, since the response rate is only 20-30%.

The molecular pathology of bile duct cancer has been a topic of intense study. The molecular pathogenesis of CCA usually involves abnormal signal transduction and pro-inflammatory secretion, facilitated by gene mutations and epigenetic dysregulations (on a set of oncogenes and tumor suppressor genes) (6). Several lines of evidence also indicate that the abnormal expression of growth factors and receptors, the RAS/RAF/ dual specificity mitogen-activated protein kinase kinase 1 pathway, and the phosphatidylinositol 3 kinase (PI3K)/protein kinase B (AKT)/mammalian target of rapamycin pathway may be involved with CCA initiation, maintenance, and metastasis (7). Several studies reported that specific-target drugs or inhibitors, including epithelial growth factor receptor (EGFR; Lapatinib or Erlotinib), fibroblast (F)GFR and PI3K inhibitor, (8) may be applicable to CCA. A number of novel therapeutics are under evaluation in a phase 2 study (9).

Alternative splicing (AS) is a post-transcription modulation process that can generate a variety of gene isoforms. Spliced mRNA is able to be translated to differential amino acids with various biological functions (10). Pre-mRNA is spliced through the spliceosome; a large macromolecule comprising 5 small nuclear ribonucleoproteins (snRNPs U1, U2, U4/U6 and U5). The AS generates 5 common splicing patterns, including alternative 5' splice site, alternative 3' splice site, exon skipping, intron retention and mutually exclusive exons. Previous data demonstrates that aberrant alternative splicing also includes exonic regulatory element mutation, splice site mutation and altered splice isoform ratios. The differential expression of splicing factors is implicated in various diseases and linked to hallmarks of cancer (11-15). A number of reports demonstrated a correlation between aberrant AS and tumor initiation/progression (16-20). The truncated oncogenic forms of the proteins, resulted from aberrant AS involved in cancer cell growth, apoptosis, drug resistance and angiogenesis.

Aberrant splicing of macrophage-stimulating protein receptor (RON) (21) and Racl (22) promoted angiogenesis and epithelial mesenchymal transition (EMT) phenotypes. In addition, a BRAF (V600E) spliced isoform, lacking exon 4-8 induced vemurafenib drug resistance in melanoma (23). In the present review, evidence is presented that supports important roles for aberrant splicing and the spliced isoforms of the genes, in CCA carcinogenesis and cancer aggressiveness.

2. Relevance of aberrant AS in cholangiocarcinoma development, progression and aggressiveness of phenotypes

A number of articles have summarized the interconnection between AS and cancer progression, including 17 genes in lung cancer (16), 2 reports in breast cancer in which 7 genes (17) and 9 genes (18), respectively were demonstrated, and 9 genes in hepatocellular carcinoma (19,20). The global cancer-specific transcript variants of five cancers demonstrated protein metabolism and modification are the most prevalent functional processes in cancer (24). As mentioned previously, aberrant AS has been discovered and proven to have functional involvement in the initiation and progression of cancer. In CCA, 623 genes presented with alternative splicing in CCA samples when compared with healthy bile duct tissue samples (25). In this review, atypical splicing of nine genes, which have been investigated at the in vitro, in vivo and clinical levels, and their relevance to CCA pathogenicity are summarized. The structure of nine pre-mRNAs that undergo alternative mRNA splicing to generate wild-type mRNA or variant transcripts are presented in Fig. 1. The derived-spliced transcripts or protein isoforms are summarized by how they can facilitate various characteristics of a cancer cell, as presented in Fig. 2 and Table I.

Table I.

Spliced mRNA transcripts and their functions in cholangiocarcinoma.

Table I.

Spliced mRNA transcripts and their functions in cholangiocarcinoma.

Author, yearGeneSpliced transcript/isoformSplicing variantsFunction(Refs.)
Yun et al, 2002CD44CD44v6Retained exon v6Proliferation(32)
Thanee et al, 2016CD44v8-10Retained exon v8-10Anti-apoptosis(33)
Tanaka et al, 2003Wnt-inducible secreted ProteinWISP1vSkipping exon 3Neural and lymphatic invasion(37)
Kokuryo et al, 2007 Serine/threonine-protein kinase Nek2Nek2BSkipping exon 8Function unknown(44)
Kamlua et al, 2012Trefoil factor 2ΔEX2TFF2Skipping exon 2Independent prognostic marker(48)
Harada et al, 2012Forkhead box protein 3Foxp3Δ3Skipping exon 3Function unknown(50)
Nutthasirikul et al, 2013Tumor protein 53Δ133p53Exon 1-4 skippingIndependent prognostic marker(60)
Nutthasirikul et al, 2015   5-Fluorouracil resistance(61)
Yu et al, 2015Pyruvate kinasePKM2Mutually exclusive exons; exon 9 skipping and exon 10 retentionNeural invasion(67)
Du et al, 2015E prostanoid receptor 3EP3-4Exon 2b, 3, 4, 6 and 8 skippingProliferation migration and invasion(71)
Yosudjai et al, 2018Anterior Gradient-2AGR2vHAlternative 3' and 5' splice site and exon 4-7 skippingMigration, invasion and adhesion(74)
Cluster of differentiation (CD)44v6 and CD44v8-10.

CD44 is a transmembrane glycoprotein receptor that specifically binds to extracellular hyaluronan and other extracellular matrix (ECM) proteins to activate signal transduction, and serves important roles in tumor proliferation, migration, and invasion (26,27). CD44 pre-mRNA encodes transmembrane and cytoplasmic-tail regions. The AS of CD44 can generate up to 12 isoforms of proteins with different biological functions. The CD44v isoforms participate in cancer progression: CD44v6 promotes EMT and activates the transforming growth factor-β pathway (28,29), and CD44v8-10 is involved in poor cancer prognosis (30,31). Expression of CD44v6 can be linked to CCA proliferation. CD44v6 is a CCA-specific isoform that has never been detected in normal bile ducts (32). Furthermore, the CD44v8-10 transcript was overexpressed in CCA and was demonstrated to stabilize the xCT system, a cysteine/glutamate transporter, to elevate glutathione synthesis and inhibit reactive oxygen species (ROS) accumulation in CCA cells. This function of CD44v8-10 was demonstrated to facilitate cancer cell survival in cases caused by liver fluke-induced inflammation. In addition, upregulation of CD44v8-10 suppressed p38 mitogen-activated protein kinase 1 (MAPK), which is a signaling protein involved in ROS suppression. Although the mechanism by which the CD44 spliced isoform may suppress p38 is still unclear, this observation appeared to be clinically relevant, since patients with CD44v overexpression and negative-phosphorylated (phospho)-p38MAPK have significantly shorter survival times compared with low CD44v expression and positive-phospho-p38(MAPK) (33).

WISP1v

Wnt-inducible secreted protein 1 [(WISP1) also known as CNN4] is a member of the cysteine-rich CCN family of proteins, which are highly expressed in skeletal tissues and has a role in bone formation and maintenance. Functions of this protein involve cell proliferation, osteoblastic differentiation and migration (34,35). WISP1 comprises 4 domains, including insulin-like growth factor-binding protein (IGFBP), VWC, thrombospondin-1 (TSP-1) and CT domains and is known to have variants with biological functions. A WISP1 variant lacking exon 3 (WISP1v) loses its VWC domain, which is thought to participate in protein complex formation. Ectopic expression demonstrated that the WISP1v is a secreted oncoprotein, which drives cellular transformation and rapid cumulative growth. WISP1v overexpression enhanced the invasive phenotype in gastric carcinoma cells, while wild-type WISP1 exhibited no such potential. These findings suggested that the CCN protein WISP1v was involved in the aggressive progression of scirrhous gastric carcinoma (36). In CCA, the aberrant isoform WISP1v was demonstrated to be overexpressed in patient CCA tissues (37). Furthermore, upregulation of WISP1v was associated with shorter overall survival time among patients following surgical treatment (38). In addition, WISP1v was demonstrated to promote cell invasion in vitro and this process was demonstrated to be mediated by p38 MAPK (37).

Nek2A and Nek2B

Nek2, or NIMA-related kinase 2, is a serine/threonine kinase that regulates cell division through centrosome separation (39). The spliced isoform of Nek consists of three forms, Nek2A, Nek2B and Nek2A-T (40). Isoforms of NEK are demonstrated to be functionally involved with cancer formation. In patients, overexpression of Nek2a was associated with Ki-67 expression, a cell proliferation marker (41). In addition, NEK2A cytoplasmic expression was positively associated with cancer grade and tumor size in breast invasive ductal carcinoma and metastatic potential (42). Cancer cells overexpressing the NEK2A isoform demonstrated a significant increase in colony formation compared with control cells and small interfering (si)RNA-based depletion of NEK2a resulted in the halting of cancer cell proliferation (43). Nek2A/Nek2A-T were demonstrated to be highly upregulated in CCA cell lines, with the predominant isoform being Nek2A/Nek2A-T and Nek2B being the lesser expressed isoform (44). Furthermore, the expression of Nek2B was demonstrated to positively correlate with proliferation potential in breast cancer cells (45).

ΔEX2TFF2

Trefoil factor 2 (TFF2) is a secreted protein that serves important roles in gastrointestinal restitution (46), chronic kidney disease and pulmonary inflammation, through the induction of cell migration and proliferation. Overexpression of TFF2 is commonly identified in several types of cancer, implicating it in carcinogenesis. TFF2 was reported to exert its pro-proliferative activity through the EGFR-MAPK pathway in CCA (47). Previously, ΔEX2TFF2, an exon 2- skipping isoform of TFF2 with a stop codon (TAG) at exon 1, was uncovered as a spliced isoform of TFF2(48). Although, the roles of this transcript have not been clarified, the present study demonstrated that a high expression ratio of ΔEX2TFF2/wtTFF2 in patients was significantly associated with a longer survival time (48). Therefore, the spliced isoform may act as a dominant-negative form of TFF2 that counteracts the cancer promoting wtTFF2 activity in CCA.

Forkhead box protein 3 (FOXP3Δ3)

FOXP3 is a transcription factor in the forkhead protein family that is involved in CD25+ regulatory T cell (Treg) development. Not only does FOXP3 control Treg development, it is also expressed in colorectal cancer cells, which is associated with poor prognosis in patients (49). Exon 3 skipping of FOXP3, resulting in an amino acid frameshift, has been reported in CCA (50). In addition, a FOXP3 splice isoform was also observed in melanoma cells, suggesting it has a role in suppressing immune activity (51).

∆133p53

Tumor protein 53 (TP53 or p53) is one of the most important tumor suppressors, indicated by its high mutation rate across all types of cancer. p53 responds to various stress signals and orchestrates processes including cell cycle arrest, DNA repair, cellular senescence and apoptosis in response to specific stress signals (52). AS generates 12 p53 isoforms, including Tap53, Δ40p53, Δ133p53 and Δ160p53 among others (53,54). The differential regulation of p53 isoforms promotes the aggressiveness of several types of cancer. A study demonstrated that Δ133p53b enhanced breast cancer stemness (55) and protected colorectal cells from camptothecin-induced apoptosis (56).

p53 has been identified as a gene that frequently mutates in a large number of CCA cases (57-59), suggesting that a perturbed p53 pathway facilitates CCA carcinogenesis. A study demonstrated that a high Δ133p53/p53 mRNA expression ratio correlates with a poor overall survival (60). Notably, Δ133p53 is also associated with resistance to certain cancer drugs; an association between Δ133p53 and 5-FU-resistance in CCA cells was demonstrated, and Δ133p53 was upregulated in 5-FU-resistant tumor tissues and CCA cell lines in a dose-dependent manner (61). Given that 5-FU is a cytotoxic drug that interferes with DNA synthesis, the Δ133p53 isoform may act as a dominant-negative p53 that interferes with the activity of wtp53 in the ternary complex (62). Accordingly, suppression of Δ133p53 promoted apoptosis, which correlated with an upregulation of pro-apoptotic Bax and a downregulation of anti-apoptotic Bcl-2(61).

Pyruvate kinase (PKM2)

PKM is a rate-limiting enzyme that catalyzes the conversion of phosphoenolpyruvate to pyruvate during glycolysis. PKM can be generated in 4 isoforms, which are expressed differently in various tissues. One of the isoforms is PKM2, which lacks exon 9 and is a major isoform highly expressed in a number of types of cancer (63). Previously data demonstrate that overexpression of PKM2 is linked to tumor growth, metastasis capability and a poor prognosis in hepatocellular carcinoma, pancreatic ductal adenocarcinoma and gallbladder cancer (64-66). In hilar cholangiocarcinoma, immunohistochemical staining specific to the PKM2 isoform demonstrated a great number of positive-staining cells in the tumor tissue. Patients with high-PKM2-expressing tumors exhibited a higher rate of tumor recurrence and a shorter overall survival time, when compared with patients with low PKM2 expression. However, there is still no conclusive evidence that indicates PKM2 is a cancer driver for CCA. In addition, PKM2 elevation was associated with CCA development and neural invasion (67).

EP3-4

E prostanoid receptor 3 (EP3), or prostaglandin E2 receptor 3 (PTGER3), is a member of a G protein-coupled receptor family, that specifically binds to prostaglandin E2 (PGE2) to activate various responses. EP3 receptor can generate up to 11 spliced isoforms. Previous data demonstrate that EP3-5 and EP3-6 isoforms were associated with cell proliferation in the myometrium in humans (68). Furthermore, overexpression of the EP3-4 receptor promoted cell growth through upregulating FUSE-binding protein 1 in liver cancer (69). In CCA, the truncated EP3-4 isoform, which includes exon 1, 2a, 5 and 10, was detected (70). This EP3-4 isoform is activated through the Src/EGFR/PI3K/AKT/glycogen synthase kinase-3β pathway and promotes cell proliferation, migration, and invasion. This results in enhanced expression of the downstream proteins c-Myc and snail. Therefore, it is believed to serve a regulatory role in CCA cell growth and metastasis (71).

Anterior Gradient-2 (AGR2)vH

The expression profiling of metastasis-associated genes in CCA demonstrated that AGR2 is one of the most-upregulated genes, specific to the metastatic CCA cell line, when compared with the parental cell line (72). The AGR2 gene encodes for a disulfide isomerase enzyme, which is commonly expressed in mucus-secreting tissues. The mRNA splicing of AGR2 was first characterized in prostate cancer (PCa). Spliced isoforms include AGR2vC, AGR2vE, AGR2vF, AGR2vG and AGR2vH. Among the 5 spliced isoforms and the wild-type, AGR2vG and AGR2vH were demonstrated to be significantly upregulated in the exosome from patient's urine sample analysis. These two exhibited high diagnostic value, with higher sensitivity and specificity when compared with the prostate-specific antigen used as a standard clinical biomarker for PCa diagnosis (73). In CCA cell lines, AGR2 RNA isoforms, namely AGR2vE, AGR2vF and AGR2vH, were recently reported that are specific to metastatic CCA cells (74). It was demonstrated that the AGR2vH isoform enables various metastatic-associated phenotypes in CCA cells. Suppression of AGR2vH by the AGR2vH-specific siRNA significantly reduced CCA cell migration and invasion. Concordantly, AGR2vH overexpression promoted cell proliferation, migration, invasion and adhesion potential. In addition, it was demonstrated that the expression of AGR2vH influences metastasis-associated phenotypes through the upregulation of vimentin. Therefore, the results indicated that the metastasis-specific isoform AGR2vH serves an important role in cancer severity (74).

3. Targeting aberrant splicing as a novel concept for cancer treatment

The prominent role of the aberrant AS in carcinogenesis has been demonstrated, indicating that AS may be a good target for cancer therapy. Aberrant AS can be manipulated in several steps: For example, Pre-Trans-Splicing Molecule (PTM) is the artificial sequence that can reprogram mRNA through replacement of the 3'exon, 5'exon and internal exon (75,76). The results demonstrated that the trans-splicing molecule reduced the number of mutant p53 transcripts in the transfected cells, which resulted in cell cycle arrest, apoptosis and tumor xenograft suppression with colorectal cancer and hepatocellular carcinoma cells (77,78). However, the use of PTM for targeting oncogenic AS events is not yet well studied and the PTM modification has limitations for cancer treatment. Therefore, this review discussed the methodologies that may apply to cancer therapy, including small molecule splicing modulators and SSOs, each of which are currently under study in clinical trials.

Small molecules splicing modulators

Splicing factors are key molecules that influence AS regulation and are associated with cancer aggressiveness and pathological phenotypes (79). A previous report demonstrated that an overexpression of serine/arginine-rich splicing factor 1 (SRSF1) can facilitate abnormal splicing of tumor suppressors and proto-oncogenes (80). The results demonstrated that SRSF1 promotes 12A inclusion of an isoform of BIN1, which interferes with the tumor-suppressing activity of this protein. In the same study, the researchers demonstrated an increase in S6K1 isoform 2 expression resulting from SRSF1 overexpression that was associated with colony formation activity (80). An Ov-infected hamster model was used to identify the differentially expressed genes to study the molecular mechanism of CCA carcinogenesis. The results demonstrated that SRSF9 is one of the genes that are overexpressed in Ov-infected hamsters and may be associated with CCA initiation (81).

Aberrant spliceosomal proteins are important factors associated with carcinogenesis

The data revealed that mutations in splicing factor 3B subunit 1A (SF3B1), which encodes the core component of U2 snRNP, is linked to erroneous 3' splice site selection (82-84). The results demonstrated that the SF3B1 K700E mutation led to differential splicing in uveal melanoma and breast cancer (85,86). In addition, luminal B and progesterone receptor-negative breast cancer patients with additional SF3B1 mutations have significantly shorter survival times (87).

It is possible to modulate aberrant AS based on small molecule inhibitors of splicing factors or mutated spliceosomal proteins: For example, it has been demonstrated that a natural product ‘Borrelidin’ can bind to a splicing protein, FBP21, leading to a decrease of the vascular endothelial growth factor (VEGF) pro-angiogenic isoform and an increase of the VEGF anti-angiogenic isoform, in RPE cells (88). Previous studies demonstrated that a natural product, FR901464 and its methylated derivative, spliceostatin A, as well as E7107, specifically inhibit spliceosome assembly through SF3B1 and lead to halted splicing reactions (89-91). The results demonstrated that treatment of these small molecules inhibits cell cycle progression and inhibits the tumor angiogenesis through decreasing the levels of VEGF transcripts (92,93).

Not only does the altered expression of splicing regulators affect AS, but the alteration of the phosphorylation status of the splicing factor/modulator was also implicated in cancer progression. In head and neck squamous cell carcinoma, hyperphosphorylation of SRPK2, a serine/arginine-rich protein-specific kinase that phosphorylates SRSF1/2, was detected in cancer cells; the phosphorylation promotes cell proliferation, migration and invasion (94). Alteration to the kinase alters the AS pattern. A previous study demonstrated that CLKs and SRSF protein kinases (SRPKs) are targets for kinase inhibitors to modulate AS; treatment with Cpd-1, Cpd-2, and Cpd-3 significantly reduced the levels of phosphorylated SR proteins, therefore affecting the splicing pattern of multiple genes and inducing cell apoptosis (95). Furthermore, the other kinase inhibitors, including ceramide, affect splice site selection of Bcl-x and increases pro-apoptotic isoforms through the dephosphorylation of the SR protein (96).

SSOs technology

SSOs are single-stranded nucleic acids, usually 15-25 bases, that are complementary to the mRNA target transcripts or the recognition sequence of the splice sites, that leads to modulated splicing. A number of studies demonstrated that SSO can inhibit aberrant RNA translation: I.e., MDM4 is the protein that contributes to embryonic development and is undetectable in adult tissues. An MDM4 isoform with exon 6 is frequently upregulated in cancer cells, impairing p53 tumor-suppressor function. The SSO-mediated skipping of exon 6 results in decreased MDM4 levels and reduced melanoma growth (97). Similarly, SSO targeting exon 26 of HER4 mRNA, named as SSOe26, demonstrated its capacity on HER4 isoform switching from CYT1 to CYP2. This treatment resulted in the depletion of the proliferation of breast cancer cells and tumor growth in mice xenografts (98). Furthermore, SSO targeted B-cell lymphoma (Bcl)-x pre-mRNA, which increased the Bcl-xS isoform, gaining pro-apoptotic activity, which was verified in the models of murine melanoma and in human glioma cell lines (99,100).

Drug development based on targeting aberrant AS, namely small molecule splicing modulators, is an interesting approach for cancer treatment. Splicing regulators are upstream molecules that control the splicing events of multiple genes. Insight into novel target genes of the splicing regulators, can be used to manipulate the effective inhibitor(s) of these upstream molecules to suppress various downstream oncogenic spliced isoforms. However, the off-target effect, toxicity (101,102) and the effects of small splicing factors interfering with the normal splicing patterns of global genes, should be considered. On the other hand, the specificity of SSO technology overcomes more than small splicing modulators by modulating AS through inhibiting only its oncogenic target which leads to effective treatment. The major problems of oligonucleotides include toxicity, instability against nucleases and delivery limitations.

4. Conclusion

The present review summarized the experimental evidence for and clinical relevance of the verification of significant effects of aberrant mRNA splicing of well-characterized genes with respect to CCA initiation and aggressiveness. The nine genes discussed underwent AS and revealed an intercorrelation with cholangiocarcinogenesis and progression. This information will serve as an opportunity to develop novel strategies for CCA detection and intervention. Interestingly, certain of the cancer-specific variants may serve as potential targets for CCA prognosis including ∆2TFF2 and ∆133p53, which demonstrate their clinical impact on patient survival. These oncogenic isoforms may be used as targets for cancer treatment, using specific antibodies, or the construction of SSOs which can modulate aberrant splicing. The regulatory machinery, including splicing factors and regulators, represents alternative targets of precision strategies, regarding the depletion of oncogenic isoforms. Finally, this summarization provides new ideas for the improvement of CCA diagnosis and treatment. Further studies should aim to investigate the unclear linkages between AS and CCA to unlock the molecular mechanisms governing AS regulation in CCA development and progression.

Acknowledgements

Not applicable.

Funding

JY was supported by The Nuresuan University research scholarship for graduate students. WK is supported by the grant from the Thailand Research Fund and Office of the Higher Education Commission providing to (grant no. TRF-MRG6080014). SJ is supported by The Foundation for Cancer Care, Siriraj Hospital, the Advanced Research on Pharmacology Fund; Siriraj Foundation (grant no. D003421) and the Chalermphrakiat Grant, Faculty of Medicine Siriraj Hospital, Mahidol University.

Availability of data and materials

Not applicable.

Authors' contributions

JY and WK designed, performed and wrote the literature review. SJ and SW revised the manuscript for intellectual content.

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

References

1 

Macias RI: Cholangiocarcinoma: And pharmacological Biology, Clinical management perspectives. ISRN Hepatol. 2014.https://doi.org/10.1155/2014/828074. PubMed/NCBI View Article : Google Scholar

2 

Banales JM, Cardinale V, Carpino G, Marzioni M, Andersen JB, Invernizzi P, Lind GE, Folseraas T, Forbes SJ and Fouassier L: et al Expert consensus document: Cholangiocarcinoma: current knowledge and future perspectives consensus statement from the European Network for the Study of Cholangiocarcinoma (ENS-CCA). Nat Rev Gastroenterol Hepatol 13. 261–280. 2016.PubMed/NCBI View Article : Google Scholar

3 

Sripa B and Pairojkul C: Cholangiocarcinoma: Lessons from Thailand. Curr Opin Gastroenterol 24. 349–356. 2008.PubMed/NCBI View Article : Google Scholar

4 

Thuwajit C, Thuwajit P, Kaewkes S, Sripa B, Uchida K, Miwa M and Wongkham S: Increased cell proliferation of mouse fibroblast NIH-3T3 in vitro induced by excretory/secretory product(s) from Opisthorchis viverrini. Parasitology 129. 455–464. 2004.PubMed/NCBI

5 

Srivatanakul P, Ohshima H, Khlat M, Parkin M, Sukaryodhin S, Brouet I and Bartsch H: Opisthorchis viverrini infestation and endogenous nitrosamines as risk factors for cholangiocarcinoma in Thailand. Int J Cancer 48. 821–825. 1991.PubMed/NCBI View Article : Google Scholar

6 

Patel T: New insights into the molecular pathogenesis of intrahepatic cholangiocarcinoma. J Gastroenterol 49. 165–172. 2014.PubMed/NCBI View Article : Google Scholar

7 

Marks EI and Yee NS: Molecular genetics and targeted therapeutics in biliary tract carcinoma. World J Gastroenterol 22. 1335–1347. 2016.PubMed/NCBI View Article : Google Scholar

8 

Rizvi S, Borad MJ, Patel T and Gores GJ: Cholangiocarcinoma: Molecular pathways and therapeutic opportunities. Semin Liver Dis 34. 456–464. 2014.PubMed/NCBI View Article : Google Scholar

9 

Goldstein D, Lemech C and Valle J: New molecular and immunotherapeutic approaches in biliary cancer. ESMO Open 2 (Suppl 1). (e000152)2017.PubMed/NCBI View Article : Google Scholar

10 

Roy B, Haupt LM and Griffiths LR: Review: Alternative splicing (AS) of genes as an approach for generating protein complexity. Curr Genomics 14. 182–194. 2013.PubMed/NCBI View Article : Google Scholar

11 

Douglas AG and Wood MJ: RNA splicing: Disease and therapy. Brief Funct Genomics 10. 151–164. 2011.PubMed/NCBI View Article : Google Scholar

12 

Ghigna C, Valacca C and Biamonti G: Alternative splicing and tumor progression. Curr Genomics 9. 556–570. 2008.PubMed/NCBI View Article : Google Scholar

13 

Tazi J, Bakkour N and Stamm S: Alternative splicing and disease. Biochim Biophys Acta 1792. 14–26. 2009.PubMed/NCBI View Article : Google Scholar

14 

Venables JP: Aberrant alternative splicing in cancer. Cancer Res 64. 7647–7654. 2004.PubMed/NCBI View Article : Google Scholar

15 

Ladomery M: Aberrant alternative splicing is another hallmark of cancer. Int J Cell Biol 2013. 463786:2013.PubMed/NCBI View Article : Google Scholar

16 

Pio R and Montuenga LM: Alternative splicing in lung cancer. J Thorac Oncol 4. 674–678. 2009.PubMed/NCBI View Article : Google Scholar

17 

Martínez-Montiel N, Anaya-Ruiz M, Pérez-Santos M and Martínez-Contreras RD: Alternative splicing in breast cancer and the potential development of therapeutic tools. Genes (Basel) 8. pii(E217)2017.PubMed/NCBI View Article : Google Scholar

18 

Xiping Z, Qingshan W, Shuai Z, Hongjian Y and Xiaowen D: A summary of relationships between alternative splicing and breast cancer. Oncotarget 8. 51986–51993. 2017.PubMed/NCBI View Article : Google Scholar

19 

Liu L, Xie S, Zhang C and Zhu F: Aberrant regulation of alternative pre-mRNA splicing in hepatocellular carcinoma. Crit Rev Eukaryot Gene Expr 24. 133–149. 2014.PubMed/NCBI View Article : Google Scholar

20 

Zhang L, Liu X, Zhang X and Chen R: Identification of important long non-coding RNAs and highly recurrent aberrant alternative splicing events in hepatocellular carcinoma through integrative analysis of multiple RNA-Seq datasets. Mol Genet Genomics 291. 1035–1051. 2016.PubMed/NCBI View Article : Google Scholar

21 

Ghigna C, Giordano S, Shen H, Benvenuto F, Castiglioni F, Comoglio PM, Green MR, Riva S and Biamonti G: Cell motility is controlled by SF2/ASF through alternative splicing of the Ron protooncogene. Mol Cell 20. 881–890. 2005.PubMed/NCBI View Article : Google Scholar

22 

Stallings-Mann ML, Waldmann J, Zhang Y, Miller E, Gauthier ML, Visscher DW, Downey GP, Radisky ES, Fields AP and Radisky DC: Matrix metalloproteinase induction of Rac1b, a key effector of lung cancer progression. Sci Transl Med 4. 142ra95:2012.PubMed/NCBI View Article : Google Scholar

23 

Poulikakos PI, Persaud Y, Janakiraman M, Kong X, Ng C, Moriceau G, Shi H, Atefi M, Titz B and Gabay MT: et al RAF inhibitor resistance is mediated by dimerization of aberrantly spliced BRAF(V600E). Nature 480. 387–390. 2011.PubMed/NCBI View Article : Google Scholar

24 

He C, Zhou F, Zuo Z, Cheng H and Zhou R: A global view of cancer-specific transcript variants by subtractive transcriptome-wide analysis. PLoS One 4. e4732:2009.PubMed/NCBI View Article : Google Scholar

25 

Chen Y, Liu D, Liu P, Chen Y, Yu H and Zhang Q: Identification of biomarkers of intrahepatic cholangiocarcinoma via integrated analysis of mRNA and miRNA microarray data. Mol Med Rep 15. 1051–1056. 2017.PubMed/NCBI View Article : Google Scholar

26 

Yu Q and Stamenkovic I: Cell surface-localized matrix metalloproteinase-9 proteolytically activates TGF-beta and promotes tumor invasion and angiogenesis. Genes Dev 14. 163–176. 2000.PubMed/NCBI View Article : Google Scholar

27 

Nam K, Oh S, Lee KM, Yoo SA and Shin I: CD44 regulates cell proliferation, migration, and invasion via modulation of c-Src transcription in human breast cancer cells. Cell Signal 27. 1882–1894. 2015.PubMed/NCBI View Article : Google Scholar

28 

Saito S, Okabe H, Watanabe M, Ishimoto T, Iwatsuki M, Baba Y, Tanaka Y, Kurashige J, Miyamoto Y and Baba H: CD44v6 expression is related to mesenchymal phenotype and poor prognosis in patients with colorectal cancer. Oncol Rep 29. 1570–1578. 2013.PubMed/NCBI View Article : Google Scholar

29 

Wang J, Xiao L, Luo CH, Zhou H, Zeng L, Zhong J, Tang Y, Zhao XH, Zhao M and Zhang Y: CD44v6 promotes β-catenin and TGF-β expression, inducing aggression in ovarian cancer cells. Mol Med Rep 11. 3505–3510. 2015.PubMed/NCBI View Article : Google Scholar

30 

Yamaguchi A, Zhang M, Goi T, Fujita T, Niimoto S, Katayama K and Hirose K: Expression of variant CD44 containing variant exon v8-10 in gallbladder cancer. Oncol Rep 7. 541–544. 2000.PubMed/NCBI View Article : Google Scholar

31 

Sosulski A, Horn H, Zhang L, Coletti C, Vathipadiekal V, Castro CM, Birrer MJ, Nagano O, Saya H and Lage K: et al CD44 splice variant v8-10 as a marker of serous ovarian cancer prognosis. PLoS One 11. e0156595:2016.PubMed/NCBI View Article : Google Scholar

32 

Yun KJ, Yoon KH and Han WC: Immunohistochemical study for CD44v6 in hepatocellular carcinoma and cholangiocarcinoma. Cancer Res Treat 34. 170–174. 2002.PubMed/NCBI View Article : Google Scholar

33 

Thanee M, Loilome W, Techasen A, Sugihara E, Okazaki S, Abe S, Ueda S, Masuko T, Namwat N and Khuntikeo N: et al CD44 variant-dependent redox status regulation in liver fluke-associated cholangiocarcinoma: A target for cholangiocarcinoma treatment. Cancer Sci 107. 991–1000. 2016.PubMed/NCBI View Article : Google Scholar

34 

Liu H, Dong W, Lin Z, Lu J, Wan H, Zhou Z and Liu Z: CCN4 regulates vascular smooth muscle cell migration and proliferation. Mol Cells 36. 112–118. 2013.PubMed/NCBI View Article : Google Scholar

35 

Ono M, Inkson CA, Kilts TM and Young MF: WISP-1/CCN4 regulates osteogenesis by enhancing BMP-2 activity. J Bone Miner Res 26. 193–208. 2011.PubMed/NCBI View Article : Google Scholar

36 

Tanaka S and Sugimachi K, Saeki H, Kinoshita J, Ohga T, Shimada M, Maehara Y and Sugimachi K: A novel variant of WISP1 lacking a Von Willebrand type C module overexpressed in scirrhous gastric carcinoma. Oncogene 20. 5525–5532. 2001.PubMed/NCBI View Article : Google Scholar

37 

Tanaka S, Sugimachi K, Kameyama T, Maehara S, Shirabe K, Shimada M, Wands JR and Maehara Y: Human WISP1v, a member of the CCN family, is associated with invasive cholangiocarcinoma. Hepatology 37. 1122–1129. 2003.PubMed/NCBI View Article : Google Scholar

38 

Wu Q, Jorgensen M, Song J, Zhou J, Liu C and Pi L: Members of the Cyr61/CTGF/NOV protein family: Emerging players in hepatic progenitor cell activation and intrahepatic cholangiocarcinoma. Gastroenterol Res Pract 2016. 2313850:2016.PubMed/NCBI View Article : Google Scholar

39 

Helps NR, Luo X, Barker HM and Cohen PT: NIMA-related kinase 2 (Nek2), a cell-cycle-regulated protein kinase localized to centrosomes, is complexed to protein phosphatase 1. Biochem J 349. 509–518. 2000.PubMed/NCBI View Article : Google Scholar

40 

Fardilha M, Wu W, Sá R, Fidalgo S, Sousa C, Mota C, da Cruz e Silva OA and da Cruz e Silva EF: Alternatively spliced protein variants as potential therapeutic targets for male infertility and contraception. Ann N Y Acad Sci 1030. 468–478. 2004.PubMed/NCBI View Article : Google Scholar

41 

Zhong X, Guan X, Dong Q, Yang S, Liu W and Zhang L: Examining Nek2 as a better proliferation marker in non-small cell lung cancer prognosis. Tumour Biol 35. 7155–7162. 2014.PubMed/NCBI View Article : Google Scholar

42 

Wang S, Li W, Liu N, Zhang F, Liu H, Liu F, Liu J, Zhang T and Niu Y: Nek2A contributes to tumorigenic growth and possibly functions as potential therapeutic target for human breast cancer. J Cell Biochem 113. 1904–1914. 2012.PubMed/NCBI View Article : Google Scholar

43 

Lai XB, Nie YQ, Huang HL, Li YF, Cao CY, Yang H, Shen B and Feng ZQ: NIMA-related kinase 2 regulates hepatocellular carcinoma cell growth and proliferation. Oncol Lett 13. 1587–1594. 2017.PubMed/NCBI View Article : Google Scholar

44 

Kokuryo T, Senga T, Yokoyama Y, Nagino M, Nimura Y and Hamaguchi M: Nek2 as an effective target for inhibition of tumorigenic growth and peritoneal dissemination of cholangiocarcinoma. Cancer Res 67. 9637–9642. 2007.PubMed/NCBI View Article : Google Scholar

45 

Wang Y, Shen H, Yin Q, Zhang T, Liu Z, Zhang W and Niu Y: Effect of NIMA-related kinase 2B on the sensitivity of breast cancer to paclitaxel in vitro and vivo. Tumour Biol 39. 1010428317699754:2017.PubMed/NCBI View Article : Google Scholar

46 

Xue L, Aihara E, Podolsky DK, Wang TC and Montrose MH: In vivo action of trefoil factor 2 (TFF2) to speed gastric repair is independent of cyclooxygenase. Gut 59. 1184–1191. 2010.PubMed/NCBI View Article : Google Scholar

47 

Kosriwong K, Menheniott TR, Giraud AS, Jearanaikoon P, Sripa B and Limpaiboon T: Trefoil factors: Tumor progression markers and mitogens via EGFR/MAPK activation in cholangiocarcinoma. World J Gastroenterol 17. 1631–1641. 2011.PubMed/NCBI View Article : Google Scholar

48 

Kamlua S, Patrakitkomjorn S, Jearanaikoon P, Menheniott TR, Giraud AS and Limpaiboon T: A novel TFF2 splice variant (∆EX2TFF2) correlates with longer overall survival time in cholangiocarcinoma. Oncol Rep 27. 1207–1212. 2012.PubMed/NCBI View Article : Google Scholar

49 

Kim M, Grimmig T, Grimm M, Lazariotou M, Meier E, Rosenwald A, Tsaur I, Blaheta R, Heemann U and Germer CT: et al Expression of Foxp3 in colorectal cancer but not in Treg cells correlates with disease progression in patients with colorectal cancer. PLoS One 8. e53630:2013.PubMed/NCBI View Article : Google Scholar

50 

Harada K, Shimoda S, Kimura Y, Sato Y, Ikeda H, Igarashi S, Ren XS, Sato H and Nakanuma Y: Significance of immunoglobulin G4 (IgG4)-positive cells in extrahepatic cholangiocarcinoma: Molecular mechanism of IgG4 reaction in cancer tissue. Hepatology 56. 157–164. 2012.PubMed/NCBI View Article : Google Scholar

51 

Ebert LM, Tan BS Browning J, Svobodova S, Russell SE, Kirkpatrick N, Gedye C, Moss D, Ng SP and MacGregor D: et al The regulatory T cell-associated transcription factor FoxP3 is expressed by tumor cells. Cancer Res 68. 3001–3009. 2008.PubMed/NCBI View Article : Google Scholar

52 

Hu W, Feng Z and Levine AJ: The regulation of multiple p53 stress responses is mediated through MDM2. Genes Cancer 3. 199–208. 2012.PubMed/NCBI View Article : Google Scholar

53 

Khoury MP and Bourdon JC: The isoforms of the p53 protein. Cold Spring Harb Perspect Biol 2. a000927:2010.PubMed/NCBI View Article : Google Scholar

54 

Surget S, Khoury MP and Bourdon JC: Uncovering the role of p53 splice variants in human malignancy: A clinical perspective. Onco Targets Ther 7. 57–68. 2013.PubMed/NCBI View Article : Google Scholar

55 

Arsic N, Gadea G, Lagerqvist EL, Busson M, Cahuzac N, Brock C, Hollande F, Gire V, Pannequin J and Roux P: The p53 isoform Δ133p53β promotes cancer stem cell potential. Stem Cell Reports 4. 531–540. 2015.PubMed/NCBI View Article : Google Scholar

56 

Arsic N, Ho-Pun-Cheung A, Evelyne C, Assenat E, Jarlier M, Anguille C, Colard M, Pezet M, Roux P and Gadea G: The p53 isoform delta133p53ß regulates cancer cell apoptosis in a RhoB-dependent manner. PLoS One 12. e0172125:2017.PubMed/NCBI View Article : Google Scholar

57 

Della Torre G, Pasquini G, Pilotti S, Alasio L, Civelli E, Cozzi G, Milella M, Salvetti M, Pierotti MA and Severini A: TP53 mutations and mdm2 protein overexpression in cholangiocarcinomas. Diagn Mol Pathol 9. 41–46. 2000.PubMed/NCBI

58 

Tullo A, D'Erchia AM, Honda K, Kelly MD, Habib NA, Saccone C and Sbisà E: New p53 mutations in hilar cholangiocarcinoma. Eur J Clin Invest 30. 798–803. 2000.PubMed/NCBI View Article : Google Scholar

59 

Liu XF, Zhang H, Zhu SG, Zhou XT, Su HL, Xu Z and Li SJ: Correlation of p53 gene mutation and expression of P53 protein in cholangiocarcinoma. World J Gastroenterol 12. 4706–4709. 2006.PubMed/NCBI View Article : Google Scholar

60 

Nutthasirikul N, Limpaiboon T, Leelayuwat C, Patrakitkomjorn S and Jearanaikoon P: Ratio disruption of the ∆133p53 and TAp53 isoform equilibrium correlates with poor clinical outcome in intrahepatic cholangiocarcinoma. Int J Oncol 42. 1181–1188. 2013.PubMed/NCBI View Article : Google Scholar

61 

Nutthasirikul N, Hahnvajanawong C, Techasen A, Limpaiboon T, Leelayuwat C, Chau-In S and Jearanaikoon P: Targeting the ∆133p53 isoform can restore chemosensitivity in 5-fluorouracil-resistant cholangiocarcinoma cells. Int J Oncol 47. 2153–2164. 2015.PubMed/NCBI View Article : Google Scholar

62 

Liu K, Zang Y, Guo X, Wei F, Yin J, Pang L and Chen D: The Δ133p53 isoform reduces wtp53-induced stimulation of DNA Pol γ activity in the presence and absence of D4T. Aging Dis 8. 228–239. 2017.PubMed/NCBI View Article : Google Scholar

63 

David CJ, Chen M, Assanah M, Canoll P and Manley JL: HnRNP proteins controlled by c-Myc deregulate pyruvate kinase mRNA splicing in cancer. Nature 463. 364–368. 2010.PubMed/NCBI View Article : Google Scholar

64 

Liu WR, Tian MX, Yang LX, Lin YL, Jin L, Ding ZB, Shen YH, Peng YF, Gao DM and Zhou J: et al PKM2 promotes metastasis by recruiting myeloid-derived suppressor cells and indicates poor prognosis for hepatocellular carcinoma. Oncotarget 6. 846–861. 2015.PubMed/NCBI View Article : Google Scholar

65 

Li C, Zhao Z, Zhou Z and Liu R: PKM2 promotes cell survival and invasion under metabolic stress by enhancing Warburg effect in pancreatic ductal adenocarcinoma. Dig Dis Sci 61. 767–773. 2016.PubMed/NCBI View Article : Google Scholar

66 

Lu W, Cao Y, Zhang Y, Li S, Gao J, Wang XA, Mu J, Hu YP, Jiang L and Dong P: et al Up-regulation of PKM2 promote malignancy and related to adverse prognostic risk factor in human gallbladder cancer. Sci Rep 6. 26351:2016.PubMed/NCBI View Article : Google Scholar

67 

Yu G, Yu W, Jin G, Xu D, Chen Y, Xia T, Yu A, Fang W, Zhang X and Li Z: et al PKM2 regulates neural invasion of and predicts poor prognosis for human hilar cholangiocarcinoma. Mol Cancer 14. 193:2015.PubMed/NCBI View Article : Google Scholar

68 

Kotani M, Tanaka I, Ogawa Y, Suganami T, Matsumoto T, Muro S, Yamamoto Y, Sugawara A, Yoshimasa Y and Sagawa N: et al Multiple signal transduction pathways through two prostaglandin E receptor EP3 subtype isoforms expressed in human uterus. J Clin Endocrinol Metab 85. 4315–4322. 2000.PubMed/NCBI View Article : Google Scholar

69 

Ma J, Chen M, Xia SK, Shu W, Guo Y, Wang YH, Xu Y, Bai XM, Zhang L and Zhang H: et al Prostaglandin E2 promotes liver cancer cell growth by the upregulation of FUSE-binding protein 1 expression. Int J Oncol 42. 1093–1104. 2013.PubMed/NCBI View Article : Google Scholar

70 

Kotelevets L, Foudi N, Louedec L, Couvelard A, Chastre E and Norel X: A new mRNA splice variant coding for the human EP3-I receptor isoform. Prostaglandins Leukot Essent Fatty Acids 77. 195–201. 2007.PubMed/NCBI View Article : Google Scholar

71 

Du M, Shi F, Zhang H, Xia S, Zhang M, Ma J, Bai X, Zhang L, Wang Y and Cheng S: et al Prostaglandin E2 promotes human cholangiocarcinoma cell proliferation, migration and invasion through the upregulation of β-catenin expression via EP3-4 receptor. Oncol Rep 34. 715–726. 2015.PubMed/NCBI View Article : Google Scholar

72 

Uthaisar K, Vaeteewoottacharn K, Seubwai W, Talabnin C, Sawanyawisuth K, Obchoei S, Kraiklang R, Okada S and Wongkham S: Establishment and characterization of a novel human cholangiocarcinoma cell line with high metastatic activity. Oncol Rep 36. 1435–1446. 2016.PubMed/NCBI View Article : Google Scholar

73 

Neeb A, Hefele S, Bormann S, Parson W, Adams F, Wolf P, Miernik A, Schoenthaler M, Kroenig M and Wilhelm K: et al Splice variant transcripts of the anterior gradient 2 gene as a marker of prostate cancer. Oncotarget 5. 8681–8689. 2014.PubMed/NCBI View Article : Google Scholar

74 

Yosudjai J, Inpad C, Chomwong S, Dana P, Sawanyawisuth K, Phimsen S, Wongkham S, Jirawatnotai S and Kaewkong W: An aberrantly spliced isoform of anterior gradient-2, AGR2vH promotes migration and invasion of cholangiocarcinoma cell. Biomed Pharmacother 107. 109–116. 2018.PubMed/NCBI View Article : Google Scholar

75 

Yang Y and Walsh CE: Spliceosome-mediated RNA trans-splicing. Mol Ther 12. 1006–1012. 2005.PubMed/NCBI

76 

Mansfield SG, Chao H and Walsh CE: RNA repair using spliceosome-mediated RNA trans-splicing. Trends Mol Med 10. 263–268. 2004.PubMed/NCBI View Article : Google Scholar

77 

He X, Liao J, Liu F, Yan J, Yan J, Shang H, Dou Q, Chang Y, Lin J and Song Y: Functional repair of p53 mutation in colorectal cancer cells using trans-splicing. Oncotarget 6. 2034–2045. 2015.PubMed/NCBI View Article : Google Scholar

78 

He X, Liu F, Yan J, Zhang Y, Yan J, Shang H, Dou Q, Zhao Q and Song Y: Trans-splicing repair of mutant p53 suppresses the growth of hepatocellular carcinoma cells in vitro and in vivo. Sci Rep 5. 8705:2015.PubMed/NCBI View Article : Google Scholar

79 

Gout S, Brambilla E, Boudria A, Drissi R, Lantuejoul S, Gazzeri S and Eymin B: Abnormal expression of the pre-mRNA splicing regulators SRSF1, SRSF2, SRPK1 and SRPK2 in non small cell lung carcinoma. PLoS One 7. e46539:2012.PubMed/NCBI View Article : Google Scholar

80 

Karni R, de Stanchina E, Lowe SW, Sinha R, Mu D and Krainer AR: The gene encoding the splicing factor SF2/ASF is a proto-oncogene. Nat Struct Mol Biol 14. 185–193. 2007.PubMed/NCBI View Article : Google Scholar

81 

Loilome W, Yongvanit P, Wongkham C, Tepsiri N, Sripa B, Sithithaworn P, Hanai S and Miwa M: Altered gene expression in Opisthorchis viverrini-associated cholangiocarcinoma in hamster model. Mol Carcinog 45. 279–287. 2006.PubMed/NCBI View Article : Google Scholar

82 

Cretu C, Schmitzová J, Ponce-Salvatierra A, Dybkov O, De Laurentiis EI, Sharma K, Will CL, Urlaub H, Lührmann R and Pena V: Molecular Architecture of SF3b and Structural Consequences of Its Cancer-Related Mutations. Mol Cell 64. 307–319. 2016.PubMed/NCBI View Article : Google Scholar

83 

Darman RB, Seiler M, Agrawal AA, Lim KH, Peng S, Aird D, Bailey SL, Bhavsar EB, Chan B and Colla S: et al Cancer-Associated SF3B1 Hotspot Mutations Induce Cryptic 3' Splice Site Selection through Use of a Different Branch Point. Cell Reports 13. 1033–1045. 2015.PubMed/NCBI View Article : Google Scholar

84 

Alsafadi S, Houy A, Battistella A, Popova T, Wassef M, Henry E, Tirode F, Constantinou A, Piperno-Neumann S and Roman-Roman S: et al Cancer-associated SF3B1 mutations affect alternative splicing by promoting alternative branchpoint usage. Nat Commun 7. 10615:2016.PubMed/NCBI View Article : Google Scholar

85 

Furney SJ, Pedersen M, Gentien D, Dumont AG, Rapinat A, Desjardins L, Turajlic S, Piperno-Neumann S, de la Grange P and Roman-Roman S: et al SF3B1 mutations are associated with alternative splicing in uveal melanoma. Cancer Discov 3. 1122–1129. 2013.PubMed/NCBI View Article : Google Scholar

86 

Maguire SL, Leonidou A, Wai P, Marchiò C, Ng CK, Sapino A, Salomon AV, Reis-Filho JS, Weigelt B and Natrajan RC: SF3B1 mutations constitute a novel therapeutic target in breast cancer. J Pathol 235. 571–580. 2015.PubMed/NCBI View Article : Google Scholar

87 

Fu X, Tian M, Gu J, Cheng T, Ma D, Feng L and Xin X: SF3B1 mutation is a poor prognostic indicator in luminal B and progesterone receptor-negative breast cancer patients. Oncotarget 8. 115018–115027. 2017.PubMed/NCBI View Article : Google Scholar

88 

Woolard J, Vousden W, Moss SJ, Krishnakumar A, Gammons MV, Nowak DG, Dixon N, Micklefield J, Spannhoff A and Bedford MT: et al Borrelidin modulates the alternative splicing of VEGF in favour of anti-angiogenic isoforms. Chem Sci (Camb) 2011. 273–278. 2011.PubMed/NCBI View Article : Google Scholar

89 

Kaida D, Motoyoshi H, Tashiro E, Nojima T, Hagiwara M, Ishigami K, Watanabe H, Kitahara T, Yoshida T and Nakajima H: et al Spliceostatin A targets SF3b and inhibits both splicing and nuclear retention of pre-mRNA. Nat Chem Biol 3. 576–583. 2007.PubMed/NCBI View Article : Google Scholar

90 

Folco EG, Coil KE and Reed R: The anti-tumor drug E7107 reveals an essential role for SF3b in remodeling U2 snRNP to expose the branch point-binding region. Genes Dev 25. 440–444. 2011.PubMed/NCBI View Article : Google Scholar

91 

Roybal GA and Jurica MS: Spliceostatin A inhibits spliceosome assembly subsequent to prespliceosome formation. Nucleic Acids Res 38. 6664–6672. 2010.PubMed/NCBI View Article : Google Scholar

92 

Satoh T and Kaida D: Upregulation of p27 cyclin-dependent kinase inhibitor and a C-terminus truncated form of p27 contributes to G1 phase arrest. Sci Rep 6. 27829:2016.PubMed/NCBI View Article : Google Scholar

93 

Furumai R, Uchida K, Komi Y, Yoneyama M, Ishigami K, Watanabe H, Kojima S and Yoshida M: Spliceostatin A blocks angiogenesis by inhibiting global gene expression including VEGF. Cancer Sci 101. 2483–2489. 2010.PubMed/NCBI View Article : Google Scholar

94 

Radhakrishnan A, Nanjappa V, Raja R, Sathe G, Chavan S, Nirujogi RS, Patil AH, Solanki H, Renuse S and Sahasrabuddhe NA: et al Dysregulation of splicing proteins in head and neck squamous cell carcinoma. Cancer Biol Ther 17. 219–229. 2016.PubMed/NCBI View Article : Google Scholar

95 

Araki S, Dairiki R, Nakayama Y, Murai A, Miyashita R, Iwatani M, Nomura T and Nakanishi O: Inhibitors of CLK protein kinases suppress cell growth and induce apoptosis by modulating pre-mRNA splicing. PLoS One 10. e0116929:2015.PubMed/NCBI View Article : Google Scholar

96 

Massiello A, Salas A, Pinkerman RL, Roddy P, Roesser JR and Chalfant CE: Identification of two RNA cis-elements that function to regulate the 5' splice site selection of Bcl-x pre-mRNA in response to ceramide. J Biol Chem 279. 15799–15804. 2004.PubMed/NCBI View Article : Google Scholar

97 

Dewaele M, Tabaglio T, Willekens K, Bezzi M, Teo SX, Low DH, Koh CM, Rambow F, Fiers M and Rogiers A: et al Antisense oligonucleotide-mediated MDM4 exon 6 skipping impairs tumor growth. J Clin Invest 126. 68–84. 2016.PubMed/NCBI View Article : Google Scholar

98 

Nielsen TO, Sorensen S, Dagnæs-Hansen F, Kjems J and Sorensen BS: Directing HER4 mRNA expression towards the CYT2 isoform by antisense oligonucleotide decreases growth of breast cancer cells in vitro and in vivo. Br J Cancer 108. 2291–2298. 2013.PubMed/NCBI View Article : Google Scholar

99 

Bauman JA, Li SD, Yang A, Huang L and Kole R: Anti-tumor activity of splice-switching oligonucleotides. Nucleic Acids Res 38. 8348–8356. 2010.PubMed/NCBI View Article : Google Scholar

100 

Li Z, Li Q, Han L, Tian N, Liang Q, Li Y, Zhao X, Du C and Tian Y: Pro-apoptotic effects of splice-switching oligonucleotides targeting Bcl-x pre-mRNA in human glioma cell lines. Oncol Rep 35. 1013–1019. 2016.PubMed/NCBI View Article : Google Scholar

101 

Eskens FA, Ramos FJ, Burger H, O'Brien JP, Piera A, de Jonge MJ, Mizui Y, Wiemer EA, Carreras MJ and Baselga J: et al Phase I pharmacokinetic and pharmacodynamic study of the first-in-class spliceosome inhibitor E7107 in patients with advanced solid tumors. Clin Cancer Res 19. 6296–6304. 2013.PubMed/NCBI View Article : Google Scholar

102 

Hong DS, Kurzrock R, Naing A, Wheler JJ, Falchook GS, Schiffman JS, Faulkner N, Pilat MJ, O'Brien J and LoRusso P: A phase I, open-label, single-arm, dose-escalation study of E7107, a precursor messenger ribonucleic acid (pre-mRNA) splicesome inhibitor administered intravenously on days 1 and 8 every 21 days to patients with solid tumors. Invest New Drugs 32. 436–444. 2014.PubMed/NCBI View Article : Google Scholar

Related Articles

Journal Cover

March-2019
Volume 10 Issue 3

Print ISSN: 2049-9434
Online ISSN:2049-9442

Sign up for eToc alerts

Recommend to Library

Copy and paste a formatted citation
x
Spandidos Publications style
Yosudjai J, Wongkham S, Jirawatnotai S and Kaewkong W: Aberrant mRNA splicing generates oncogenic RNA isoforms and contributes to the development and progression of cholangiocarcinoma (Review). Biomed Rep 10: 147-155, 2019
APA
Yosudjai, J., Wongkham, S., Jirawatnotai, S., & Kaewkong, W. (2019). Aberrant mRNA splicing generates oncogenic RNA isoforms and contributes to the development and progression of cholangiocarcinoma (Review). Biomedical Reports, 10, 147-155. https://doi.org/10.3892/br.2019.1188
MLA
Yosudjai, J., Wongkham, S., Jirawatnotai, S., Kaewkong, W."Aberrant mRNA splicing generates oncogenic RNA isoforms and contributes to the development and progression of cholangiocarcinoma (Review)". Biomedical Reports 10.3 (2019): 147-155.
Chicago
Yosudjai, J., Wongkham, S., Jirawatnotai, S., Kaewkong, W."Aberrant mRNA splicing generates oncogenic RNA isoforms and contributes to the development and progression of cholangiocarcinoma (Review)". Biomedical Reports 10, no. 3 (2019): 147-155. https://doi.org/10.3892/br.2019.1188