Next Article in Journal
Tipping Growth Inhibition into Apoptosis by Combining Treatment with MDM2 and WIP1 Inhibitors in p53WT Uterine Leiomyosarcoma
Next Article in Special Issue
DNA Methylation Regulates a Set of Long Non-Coding RNAs Compromising Hepatic Identity during Hepatocarcinogenesis
Previous Article in Journal
Radical or Not-So-Radical Prostatectomy: Do Surgical Margins Matter?
Previous Article in Special Issue
The Role of Protein SUMOylation in Human Hepatocellular Carcinoma: A Potential Target of New Drug Discovery and Development
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Impact of Alternative Splicing Variants on Liver Cancer Biology

1
Experimental Hepatology and Drug Targeting (HEVEPHARM), Institute for Biomedical Research of Salamanca (IBSAL), University of Salamanca, 37007 Salamanca, Spain
2
Center for the Study of Liver and Gastrointestinal Diseases (CIBERehd), Carlos III National Institute of Health, 28029 Madrid, Spain
3
Instituto de Investigaciones Sanitarias de Navarra IdiSNA, 31008 Pamplona, Spain
4
Hepatology Program, CIMA, University of Navarra, 31008 Pamplona, Spain
*
Author to whom correspondence should be addressed.
Cancers 2022, 14(1), 18; https://doi.org/10.3390/cancers14010018
Submission received: 19 November 2021 / Revised: 15 December 2021 / Accepted: 16 December 2021 / Published: 21 December 2021

Abstract

:

Simple Summary

Among the top ten deadly solid tumors are the two most frequent liver cancers, hepatocellular carcinoma, and intrahepatic cholangiocarcinoma, whose development and malignancy are favored by multifactorial conditions, which include aberrant maturation of pre-mRNA due to abnormalities in either the machinery involved in the splicing, i.e., the spliceosome and associated factors, or the nucleotide sequences of essential sites for the exon recognition process. As a consequence of cancer-associated aberrant splicing in hepatocytes- and cholangiocytes-derived cancer cells, abnormal proteins are synthesized. They contribute to the dysregulated proliferation and eventually transformation of these cells to phenotypes with enhanced invasiveness, migration, and multidrug resistance, which contributes to the poor prognosis that characterizes these liver cancers.

Abstract

The two most frequent primary cancers affecting the liver, whose incidence is growing worldwide, are hepatocellular carcinoma (HCC) and intrahepatic cholangiocarcinoma (iCCA), which are among the five most lethal solid tumors with meager 5-year survival rates. The common difficulty in most cases to reach an early diagnosis, the aggressive invasiveness of both tumors, and the lack of favorable response to pharmacotherapy, either classical chemotherapy or modern targeted therapy, account for the poor outcome of these patients. Alternative splicing (AS) during pre-mRNA maturation results in changes that might affect proteins involved in different aspects of cancer biology, such as cell cycle dysregulation, cytoskeleton disorganization, migration, and adhesion, which favors carcinogenesis, tumor promotion, and progression, allowing cancer cells to escape from pharmacological treatments. Reasons accounting for cancer-associated aberrant splicing include mutations that create or disrupt splicing sites or splicing enhancers or silencers, abnormal expression of splicing factors, and impaired signaling pathways affecting the activity of the splicing machinery. Here we have reviewed the available information regarding the impact of AS on liver carcinogenesis and the development of malignant characteristics of HCC and iCCA, whose understanding is required to develop novel therapeutical approaches aimed at manipulating the phenotype of cancer cells.

1. Introduction

According to cancer incidence and mortality statistics worldwide (both genders, all ages) registered by Globocan (https://gco.iarc.fr/ accessed on 18 September 2021), in 2020, liver cancer was ranked sixth in incidence with 905,677 new cases (4.7% of new cases) but third regarding mortality with 830,180 deaths (8.3% of all cancer-related deaths). Liver cancer includes several forms of cancers, but in most adult patients, these tumors are either hepatocellular carcinoma (HCC) (85–90%), derived from hepatocytes, or intrahepatic cholangiocarcinoma (iCCA) (10–15%), derived from cholangiocytes.
Because aberrant splicing generates mRNA isoforms involved in liver cancer development and malignancy [1], several studies have attempted to establish a relationship between alternative splicing (AS) signatures and the prognosis of HCC [2,3,4,5] and iCCA [6,7]. Editing immature mRNA (pre-mRNA) by splicing is carried out by the spliceosome, a multi-megadalton ribonucleoprotein complex in charge of introns removal and exons ligation as part of an integrated mechanism involved in gene regulatory network [8]. In “constitutive splicing”, this occurs at canonical sites of pre-mRNA, whereas in AS the cut-and-paste process involves different splicing sites, which result in diverse forms of partial or complete removal or retention of exons and introns [9]. The relevance of AS is easily understood if one considers that a single pre-mRNA can produce more than one splicing variant (SV) of mature mRNA, resulting in structurally and functionally dissimilar proteins. Since approximately 97% of human genes can undergo AS [9,10] under normal physiological conditions, this diversity substantially increases the coding capacity of our genome [11]. Nevertheless, AS is also a source of noxious proteins, which have been associated with the pathogenesis of several diseases, including cancer. Moreover, re-splicing of mature mRNA in cancer cells promotes the activation of distant weak alternative splice sites, potentially generating mutation-independent diversity in cancer transcriptomes. This process might result from the impairment during carcinogenesis of a not-yet well-understood mechanism existing in normal cells to prevent potentially harmful mRNA re-splicing events [12]. Cancer-specific re-splicing, such as that of TSG101 mRNA in nasopharyngeal carcinoma, has been proposed to be associated with enhanced invasiveness and metastasis [13].
The present review deals with the current understanding of the impact of AS on liver carcinogenesis and the development of malignant traits frequently found in HCC and iCCA.

2. The Intron Removal Process

The vast majority of pre-mRNA in humans are processed by the so-called “major spliceosome”, which binds to specific sequences located in the introns and plays an essential role in the exon-recognition process [14]. These are the 5′ donor site (5′SS, GU) and the 3′ acceptor site (3′SS, AG) located at the intron-exon junctions [15]. In the case of a few genes (1%) the “minor spliceosome” binds to equivalent sequences for 5′SS and 3′SS, which in this case are AU and AC, respectively [16,17,18]. Two additional signatures of the intron are the branch point site (BPS), located 18–40 nt from the end of the intron, and the polypyrimidine tract (PPT), usually 15–20 nt long, located about 5–40 nt upstream of the 3′SS [15]. The sequential interaction of spliceosome elements, such as small nuclear ribonucleoproteins (snRNP) consisting of small nuclear RNAs (U1, U2, U4, U5, and U6 snRNA) and Sm proteins, is affected by the balance between SR protein and heterogeneous nuclear ribonucleoproteins (hnRNPs), which determines the outcome of mRNA maturation. During exon-definition, the first step is the formation of the E complex, which starts with the interaction of snRNP with the 5′SS in the pre-mRNA. Then, the splicing factor 1 (Sf1) binds to the BPS, which is followed by the recruitment of the U2 heterodimer (U2AF), whose 35 kDa (U2af1) and 65 kDa (U2af2) subunits bind to the 3’SS and PPT, respectively [19]. The next step is the formation of the A complex, in which U2 snRNP displaces Sf1 in an ATP-dependent manner. The RNA helicases PRP5/DDX46 and DDX39B/UAP56 promote the interaction of U1 snRNP with U2 snRNP [20] and stabilize the interaction between U2 snRNP and pre-mRNA BPS [21], respectively. The recruitment of U4/U6-U5 tri-snRNP generates the pre-catalytic B complex (B1), from which U1 and U4 snRNPs are released, resulting in the activated B complex (B2) [22]. The subsequent step leads to the first reaction of splicing, producing the C complex together with an intermediate pre-mRNA with lariat structure [23]. Another trans-esterification reaction catalyzed by the C complex releases the intron lariat structure, which is rapidly degraded by the cell [24,25,26]. The pre-mRNA processing factors 3, 4, 6, 8, and 31 (PRPF3, PRPF4, PRPF6, PRPF8, PRPF31) participate in the splicing of the pre-mRNA as components of the U4/U6-U5 tri-snRNP complex and the spliceosome B complexes [27].
Besides the difference in guanine and cytosine content between exons and introns together with the existence of tag sequences (5′SS, 3′SS, BPS, and PPT), there are cis-regulatory sequences, which depending on their position and function, are named exonic splicing enhancers (ESE), exonic splicing silencers (ESS), intronic splicing enhancers (ISE), and intronic splicing silencers (ISS). Due to their interaction with regulatory proteins, these sites play a crucial role in defining pre-mRNA fragments, which are then removed as introns from the final mature mRNA [15,18,28,29]. Among the splicing factors involved in this process are SR proteins, characterized by a domain rich in arginine/serine dipeptides, which bind to ESE and ISE, favoring short alternatives of splicing [30]. In contrast, hnRNPs bind to ESS and ISS, inhibiting short options of splicing and hence favoring the formation of long SVs [31].
Besides, several kinases and phosphatases, by modulating the phosphorylation and dephosphorylation of SR proteins and other spliceosome components, determine their activity and cellular localization and hence participate indirectly in regulating the spliceosome [32].

3. Liver Cancer-Associated Alterations in the Spliceosome Machinery

During carcinogenesis, a common feature is the need for a high rate of pre-mRNA processing in combination with the frequent appearance of alterations in spliceosome elements, which altogether account for the generation of a large variety of SVs (more than 15,000) associated with different cancer types [33,34]. Although some of these SVs are noxious for the cells harboring them, in most cases, these variants have a positive impact on tumor biology, enhancing their malignant characteristics [35,36,37]. In the case of liver cancer, the global profile of alternative RNA splicing events indicates the existence of differential AS in HCC and iCCA compared to adjacent non-tumor liver tissue, affecting several hundred genes. This information has been proposed as a valuable tool to distinguish different types of liver cancer [2,38].
Cancer-associated changes in the spliceosome can be due to: (i) altered expression of splicing factors [39]; (ii) mutations affecting genes coding spliceosome components and associated regulatory proteins; (iii) disruption of either SS or splicing regulatory sites (enhancers or silencers) [40,41]; and (iv) impairment of signaling pathways involved in the splicing machinery regulation [40,42]. In this regard, marked alterations in the expression pattern of spliceosome factors, in particular those proteins accounting for exon recognition, have been described both in HCC and iCCA, where they may affect patient outcomes (Table 1) [43,44]. Moreover, transcriptome-wide analysis has revealed the presence in HCC of a high number of aberrant AS events, which has been associated with impaired expression of splicing factors [3].
Besides alterations in their expression levels, mutations in RNA-binding proteins (RBPs) are commonly found in cancer, which contributes to tumor initiation and progression [45]. More precisely, 286 RBP-encoding genes have been found differentially expressed (231 up-regulated and 55 down-regulated) in HCC samples versus non-tumor tissue [3]. These alterations in RBP abundance presumably lead to AS dysregulation.
A meta-analysis has shown that many genes of the spliceosome machinery, such as SF3B proteins, which are essential components of the U2 complex, are up-regulated in HCC [46]. One of these proteins (SF3B1) is particularly relevant in the context of this review because it has been identified as a driver in many cancers [47]. SF3B1 plays a critical role in exon recognition by favoring U2 snRNA binding to the BPS [48] and by allowing the identification of short exons flanked by long introns with differential GC content through interaction with nucleosomes [49]. SF3B1 up-regulation in HCC has been associated with higher tumor aggressiveness and shorter overall survival (OS). This may be due in part to the favored generation of some oncogenic SVs of specific proteins, such as KLF6-SV1 [47]. Recent results obtained in human gastric cancer cells (AGS and MKN28) have revealed that inhibition of SF3B1 reduced their proliferation rate by inducing apoptosis and G2/M phase arrest through altering homeobox A10 (HOXA10) mRNA splicing. Thus, SF3B1 knockdown down-regulated HOXA10 mRNA expression whereas enhanced the expression of long noncoding RNA (lncRNA) isoforms of HOXA10 (exons 1 and 3) and HOXA10 (exons 2 and 3). Moreover, SF3B1 inhibition increased PTEN levels and reduced AKT protein phosphorylation [50].
The pharmacological inhibition of SF3B1, using pladienolide-B, results in decreased cell proliferation, migration, and colony-formation ability in HCC cells in vitro and impaired tumor growth in vivo [47]. These findings support the role of SF3B1 in liver carcinogenesis. They suggest that SF3B1 is a promising novel druggable target as well as a prognostic biomarker in HCC [51], as has been reported in other malignancies, such as pancreas [52], breast [53], and prostate [54] cancers.
Interestingly, not only changes in SF3B1 expression but also somatic mutations affecting this component of the splicing machinery play an essential role in liver cancer progression. Indeed, SF3B1 is the most frequently mutated splicing factor in many cancers [52,53,55]. These mutations have been associated with increased DNA damage and altered response to DNA damage [56]. SF3B1 mutations have also been found in HCC [57], where they are especially frequent at the late stage of carcinogenesis [58].
In addition to SF3B1, other SF3B proteins have been found associated with HCC. Thus, SF3B2 and SF3B4 are overexpressed in HCC and, similarly, in precancerous lesions in early-stage HCC [46]. Aberrant overexpression of SF3B4 inactivates Krüppel-like factor 4 (KLF4), a tumor suppressor. By inducing KLF4 aberrant splicing [59], SF3B4 may promote liver tumorigenesis, which has led to suggest its usefulness as an early-stage marker of HCC. Additionally, reduced expression of miRNA-133b triggers enhanced expression of SF3B4 in HCC, favoring metastasis, poor prognosis, and reduced OS [60,61].
Regarding core proteins involved in E complex formation, no relationship between clinical parameters of HCC and changes in SF1 and U2AF1 expression has been found [43]. In contrast, SNRPB and SNRPD1 expression is higher in HCC than in the adjacent liver tissue and correlates with shorter OS [43]. Furthermore, SNRPB up-regulation has been recently described to be mediated by c-Myc in HCC [62]. In that study, ectopic expression of SNRPB induced HCC cell proliferation and migration. In contrast, gene down-regulation promoted cell apoptosis, suggesting that SNRPB could act as an indirect oncogene in HCC [62]. It is noteworthy that SNRPB expression was found significantly increased in metastatic versus non-metastatic liver tumors, which suggests a potential role of SNRPB in HCC metastasis [62]. This is not a unique HCC characteristic because the expression of SNRPB is also increased in gliomas [63,64] and in non-small cell lung cancer (NSCLC) [65], where the tumorigenic capacity of SNRPB seems to be mediated, in part, by AS promotion leading to the formation of a RAB26 SV that retains intron 7 [65].
In both human and animal models, the step II splicing regulator SLU7 has been found down-regulated in HCC, as well as in preneoplastic conditions, such as liver cirrhosis [66,67]. The reduced SLU7 expression has been associated with the activation of the epidermal growth factor receptor (EGFR) by its ligand amphiregulin (AREG) during inflammation [66], which has also been linked to the down-regulation of other splicing regulators, such as the SR protein SRSF3 [68,69].
Regarding members of the hnRNP family, altered expression of several hnRNPs has been found to favor carcinogenesis in several organs, including the liver. Thus, tumor suppressor hnRNPE1 (PCBP1) is down-regulated in several cancers. In contrast, oncogenes hnRNPE2 (PCBP2) and hnRNPK are up-regulated and promote tumorigenesis [70,71,72,73]. Nevertheless, in HCC, controversial results have been reported in this respect. Thus, both up- and down-regulation of hnRNPE2 in HCC have been reported by different studies [43,74]. Low expression of hnRNPK, hnRNPL, and hnRNPE2 has been associated with shorter OS of HCC patients [43]. The study of the role of hnRNPK in HCC development revealed that, after SUMOylation, hnRNPK could bind to p53 in HCC cells. This interaction ultimately leads to the accumulation and transactivation of p53 and hence inhibition of HCC cell proliferation [75]. Besides, miR-1249-3p could induce HCC cell proliferation and invasion through hnRNPK down-regulation [76]. These data are consistent with the positive impact of enhanced hnRNPK expression on HCC patient survival [43]. The role of hnRNPK in iCCA has also been described. O-glycosyl-N-acetylation of that protein has been associated with shorter survival of iCCA patients and increased iCCA progression and metastasis. The reason is that this post-translational modification promotes hnRNPK nuclear translocation, which acting as a transcription factor, induces the expression of several proteins involved in cell proliferation, migration, and apoptosis inhibition. Based on these findings, hnRNPK O-GlcNAcylation has been suggested as a druggable target to repress iCCA progression [44].
Enhanced expression of hnRNPAB in HCC has been associated with phenotypic epithelial-mesenchymal transition (EMT) in a SNAIL-dependent manner, leading to higher migration capacity and worse outcomes [77]. Furthermore, other members of this family of splicing factors, the polypyrimidine tract binding proteins (PTBP) are involved in HCC progression and metastasis. Thus, PTBP3 has increased expression in HCC, which is correlated with tumor size, metastasis, and lower survival [78]. Moreover, elevated levels of MTR4 promote HCC tumorigenesis through recruitment of PTPBP1 to its target pre-mRNA, which results in a metabolic switch in cancer precursor cells [79].
Among SR proteins, SRSF1, SRSF3, SRSF5, and SRSF9 are considered indirect oncogenes because they promote carcinogenesis through the enhanced generation of more oncogenic SVs of direct oncogenes, and their expression is elevated in several cancers [80,81,82,83,84]. Nevertheless, this does not seem to be the case in HCC, where a general down-regulation of these SR proteins has been found [69]. This is crucial because SRSF3 has been identified as an essential player for hepatocyte differentiation and metabolic function [85]. Thus, the aberrant splicing isoform of SRSF3, including exon 4 SRSF3-Iso2, which participates in the induction of genome instability, is up-regulated not only in HCC but already in the cirrhotic liver [68]. In mice, the loss of Srsf3 induces AS of genes related to EMT and drives spontaneous HCC with aging [86]. A recent study has found a decreased SRSF3 activity in HCC cells due to an overexpression of PPM1G, a serine/threonine phosphatase involved in the dephosphorylation of splicing factors. As a result, cell proliferation, invasion, and metastasis of HCC cells were promoted through changes in the AS pattern of genes related to cell cycle control and transcriptional regulation [87]. Furthermore, SRSF2 expression levels have been shown to have prognostic value because SRSF2 up-regulation was inversely associated with patient survival [37,43]. Consistently, SRSF2 silencing using shRNAs decreased the tumorigenic potential of HCC cell lines in vivo by controlling the AS of target genes in a specific manner [37]. Altogether, these data indicate that SRSF2 protein could act as a cancer driver in human HCC. This seems to be a species-specific characteristic, as the loss of Srsf2 has the opposite effect in mice, triggering hepatic progenitor cell activation and HCC onset [88].
Table 1. Impact of altered expression of spliceosome components and associated regulating proteins on the outcome of patients with liver cancer.
Table 1. Impact of altered expression of spliceosome components and associated regulating proteins on the outcome of patients with liver cancer.
GeneChangeConsequencesClinical ImpactCancerRefs.
ESRP1UpImpaired exon inclusionEnhanced migration; Distant metastasisHCC
iCCA
[89,90]
hnRNPABUpEnhanced migration
Enhanced EMT
Promote tumor recurrenceHCC[77]
hnRNPE2ControversialControversialChange in OSHCC[43,74]
hnRNPKDownTransactivation of p53
Decreased proliferation
Shorter OSHCC[43]
hnRNPKO-glycosyl-N-acetylationProgression and MetastasisShorter OSiCCA[44]
hnRNPLDownUnknownShorter OSHCC[43]
hnRNPMUpEnhanced exon skippingEnhanced migration; Distant metastasisHCCi
CCA
[91,92]
PTBP3UpCorrelated with tumor size and metastasisShorter OSHCC[78]
SF3B1UpEnhanced proliferation, migration, and tumor growthShorter OS; Higher aggressivenessHCC[47]
SF3B4UpPromote tumorigenesis and metastasisReduced OSHCC[59,60]
SLU7DownEnhanced genome instability and de-differentiationMay favor hepatocarcinogenesis from cirrhotic liverHCC[66,67,68,69]
SNRPBUpEnhanced proliferation and migrationShorter OS; MetastasisHCC[43]
SNRPD1UpUnknownShorter OSHCC[43]
SRSF2DownEnhanced proliferation
Liver tumorigenesis in vivo
Longer OSHCC[43,88]
SRSF3DownEnhanced EMT and cell proliferationMetastasisHCC[87]
Changes in gene expression were based on mRNA abundance. Down, down-regulation; EMT, Epithelial-mesenchymal transition; HCC, hepatocellular carcinoma; iCCA, intrahepatic cholangiocarcinoma; OS, overall survival; Up, up-regulation.

4. Altered Splicing of Genes Involved in Liver Carcinogenesis and Metastasis

Besides alterations in the splicing machinery, mutations in SS and cis-regulatory elements of target genes can also trigger cancer onset [93]. Thus, mutations (frequently at 3′SS) in several tumor suppressor genes, such as TP53, ARID1A, PTEN, BRCA1, BRCA2, CDH1, MLH1, and FLCN, are of particular relevance in carcinogenesis [94,95,96,97]. These mutations can affect the integrity of the SS, resulting in aberrant mRNA maturation [94,95,96,97] and subsequent impairment of their tumor suppression function, hence favoring cancer development [98]. Moreover, the cell cycle is highly dependent on pre-mRNA splicing and accurate AS. Several key cell cycle factors and their known SVs have been related to uncontrolled cell proliferation. Besides changes in their expression levels and the presence of somatic mutations, aberrant post-translational modifications, for instance, affecting chromatin dynamics, can also alter proper cell cycle control during carcinogenesis. Two primary connections link pre-mRNA splicing, cell cycle progression, and carcinogenesis. One is derived from the need for particular pre-mRNA splicing to allow the production of essential factors of the cell cycle that harbor introns. The second is AS of those critical factors required to produce the specific isoform needed for cell cycle progression (for a complete review, see [99]). Moreover, the proper function of the spliceosome component SF3B1, whose impact on carcinogenesis has been mentioned above, relies on its phosphorylated state at several sites, which varies during the cell cycle [100]. SF3B1 phosphorylation peaks at the G2/M phase in a CDK1-dependent manner and then decreases owing to PP2A and PP1 phosphatase activities as the cell cycle progresses through the G1/S phase [99].
Metastasis of HCC and iCCA cells has been related to AS events, mainly due to the appearance of SVs that promote EMT (Table 2 and Table 3). Alterations in the expression of different splicing factors, such as ESRP1 and hnRNPM, are critical in EMT activation. Mesenchymal phenotype is favored by hnRNPM due to its effect on enhancing exon skipping and TGFβ signaling, whereas ESRP1 favors epithelial phenotype by enhancing exon inclusion [90] (Table 1). Furthermore, ESRP1 can also act as an oncogene when it is overexpressed [89] (Table 1).

4.1. Role of SVs in the Onset, Progression, and Metastasis of HCC

4.1.1. TP53

HCC development has been associated with aberrant splicing events affecting crucial genes (Table 2). For instance, the essential tumor suppressor gene TP53 is frequently mutated in many types of cancers [101], including HCC [102]; several of its mutations affect SS located between introns 3 and 9, causing skipping of adjacent exons [101,103,104]. On the other hand, the TP53 gene potentially encodes at least 12 p53 naturally occurring isoforms, resulting from the combination of four different N-terminal p53 forms (full-length, Δ40, Δ133, and Δ160) and three different C-terminal domains (α, β, and γ) [105]. The SV of TP53 Δ40p53α, which lacks the 39 N-terminal amino acids corresponding to the first transactivation domain (TAD-I) of FL-p53, can suppress the proliferation of HCC cells [106].

4.1.2. TP73

Two SVs of p73, ΔEx2p73, and ΔEx2/3p73, are highly expressed in HCC but are not found in healthy hepatocytes. Both isoforms are truncated proteins lacking the transactivation domain [66,107,108]. Their up-regulation in HCC accounts for the inhibition of p73 tumor-suppressive activity. Overexpression of human ΔEx2/3p73 in hepatocytes of transgenic mice results in the spontaneous development of HCC, supporting the oncogenic impact of these SVs [108]. Previous studies have unraveled the mechanism implicated in the up-regulation of these oncogenic isoforms and demonstrated that AREG induction in response to stress or inflammatory signals during liver damage is responsible for the down-regulation of SLU7, which in turn results in aberrant exon 2 skipping and hence the expression of the oncogenic isoform ΔEx2p73 [66].

4.1.3. CDH17

Somatic mutations in critical sequences of oncogenes leading to their aberrant splicing also drive HCC development. Thus, the intestinal cell-cell adhesion protein, cadherin 17 (CDH17), is expressed in HCC as an SV lacking exon 7 (ΔEX7CDH17), which correlates with decreased OS and higher tumor recurrence [109]. The appearance of ΔEX7CDH17 has been associated with the presence in CDH17 pre-mRNA of two single nucleotide polymorphisms (SNPs) located in exon 6 (rs 3214050) and the BPS of intron 6 (rs 2514813) [109]. The generation of this SV is considered among the genetic events contributing to the high prevalence of HCC in Asia [110].

4.1.4. KLF6

As mentioned above, aberrant AS of genes regulating cell cycle arrest can also promote HCC development. Thus, the Krüppel-like factor 6 (KLF6) is a zinc-finger transcription factor that inhibits cell proliferation in part by transcriptional activation of p21, a cell cycle checkpoint protein. KLF6 function is abrogated in human cancers owing to increased AS that yields a dominant-negative isoform, KLF6 SV1, which antagonizes full-length KLF6-mediated growth suppression [111], presumably by promoting KLF6 degradation [112]. The ratio between SV1 and wild-type KLF6 was increased in HCV-associated HCC, which was related to higher tumor aggressiveness. Remarkably, elevated ratio SV1 and wild-type KLF6 promoted liver tumorigenesis in mice [112]. In contrast, the SV2 KLF6 variant is down-regulated in human HCC samples compared to adjacent non-tumor liver tissue. It has anti-tumorigenic activity, inducing cell cycle arrest by p21 activation and increasing pro-apoptotic factors, such as BAX [113].

4.1.5. FGFR2

The tyrosine kinase receptor fibroblast growth factor receptor 2 (FGFR2) undergoes AS generating two SVs (FGFR2-IIIb and FGFR2-IIIc) by incorporating mutually exclusive exons. EMT is associated with AS of FGFR2 switching from the epithelial- (FGFR2-IIIb) to the mesenchymal-type (FGFR2-IIIc) SV, which further promotes invasion and metastasis [114]. Although several proteins can affect AS of FGFR2, ESRP1 is the primary regulator, which promotes FGFR2-IIIb expression and hence favors epithelial phenotype [115,116]. In HCC, FGFR2-IIIb down-regulation has been associated with higher vascular invasion and more advanced tumor stage. Moreover, FGFR2-IIIb re-expression in HCC cells in vitro reduced proliferation and migratory potential [117].

4.1.6. FGFR3

Another tyrosine kinase receptor abnormally up-regulated in HCC is the fibroblast growth factor receptor 3 (FGFR3) [118], which also undergoes AS, plays an important role in cell proliferation, differentiation, and angiogenesis. The IgG-like-III domain of FGFR3, which is required for binding to its ligands (FGFs), is encoded by two separate exons, i.e., exon 8 and exon 9. The alternative selection of one of these exons results in FGFR3-IIIb and FGFR3-IIIc SVs, respectively. The expression levels of these SVs, both separately or in combination, are higher in 53% of all HCCs than in adjacent non-tumor liver tissue, especially in cases with early tumor infiltration, metastasis, and recurrence at the time of surgery [118]. Interestingly, FGFR3-IIIb overexpression in HCC cell lines using lentiviral constructs leads to enhanced proliferation. Moreover, both SVs could inhibit apoptosis and improve the tumorigenic activity of these HCC cells in vivo. Besides, blockade of FGFR3-IIIb/c dramatically reduced tumor growth [118], which suggests that these SVs act as oncogenes in HCC and are potential therapeutic targets. Moreover, aberrant AS of FGFR3 could also lead to the mutant FGFR3Δ7–9 variant, which has carcinogenic activity and has been associated with enhanced invasiveness in vitro and metastasis in vivo [119]. This isoform links exon 6 to exon 10 directly, resulting in a complete deletion of the IG-like-III domain. Nevertheless, the FGFR3Δ7–9 variant can be self-activated, independently of the FGF ligand binding. FGFR3Δ7–9 activates the AKT pathway and indirectly decreases the expression of the tumor suppressor PTEN [120]. In this way, FGFR3Δ7–9 enhances HCC cell proliferation and tumor growth in vivo [119]. Moreover, FGFR3Δ7–9 expression is accompanied by lower E-cadherin levels and a higher expression of stem/mesenchymal markers such as SNAIL and MMP-9 [119].

4.1.7. DNMT3b3

Some HCC-associated SVs have already been detected in preneoplastic stages, suggesting their possible early contribution to liver malignancy. For example, up-regulation of DNA methyltransferase 3b3 (DNMT3b3) has been described in the liver of patients with hepatitis and chronic cirrhosis, as well as in HCC. This was partly due to an elevated expression of DNMT3b4, a splice variant of DNMT3b lacking conserved methyltransferase motifs IX and X, which has been correlated with the degree of DNA hypomethylation in pericentromeric satellite regions both under precancerous conditions and in HCC [121].

4.1.8. PDSS2

Another example concerns the AS of prenyl diphosphate synthase subunit 2 (PDSS2), an essential enzyme in coenzyme Q10 synthesis involved in hepatocarcinogenesis [122]. PDSS2 down-regulation has been associated with a poor prognosis in HCC. Interestingly, an SV of PDSS2 lacking exon 2 (PDSS2Δ2) has been detected in HCC, and its overexpression was associated with shorter OS [123]. Apart from its pro-carcinogenic activity [122], PDSS2∆2 favors cell migration in vitro and metastasis in vivo, presumably, through Wnt/β-catenin and NF-Kβ signaling pathway [123].

4.1.9. AXL

The AXL tyrosine kinase receptor has attracted much attention as a potential therapeutical target. Skipping of exon 10 in AXL pre-mRNA gives rise to the AXL short (AXL-S) SV, which is up-regulated in HCC cells by hnRNPI (PTBP1) resulting in the promotion of cell migration [124]. Consistently, AXL-S silencing in HCC cells significantly inhibited pulmonary metastasis in a xenografted nude mouse model [124].

4.1.10. NF2

The tumor suppressor moesin-ezrin-radixin-like protein (Merlin) encoded by the NF2 gene is down-regulated in HCC [125]. Interestingly, an SV lacking exons 2, 3, and 4 (MerlinΔ2–4), identified in HCC, has the opposite effect on liver tumor progression to that of the wild-type isoform [125]. Thus, knocking-down Merlin or overexpressing MerlinΔ2–4 promotes cell migration and invasion through Twist1 up-regulation. Furthermore, MerlinΔ2–4 also stimulates the activation of β-catenin and the expression of stemness-related genes. Remarkably, using a pulmonary metastatic mouse model, the expression of Merlin reduced metastasis, whereas that of MerlinΔ2–4 induced distant metastasis [125].

4.1.11. XBP1

The SV of the X-box binding protein 1 (XBP1s) is a well-known activator of UPR (unfolded protein response), an adaptive response and defense mechanism through which tumor cells can survive under adverse conditions of enhanced ER stress. XBP1s expression is markedly high in HCC, where its up-regulation correlates with the existence of distant metastasis and poor prognosis. This could be due to the ability of XBP1 to induce EMT and migration in HCC cells [126].

4.1.12. OPN

Three SVs of osteopontin (OPN), a glycoprotein of the extracellular matrix involved in proliferation, angiogenesis, and invasion, have been described. OPN-a contains all exons, whereas OPN-b and OPN-c lack exons 5 and 4, respectively. Overexpression of OPN-a and OPN-b has been associated with greater invasiveness in HCC [127].

4.1.13. CD44

CD44 is a membrane receptor glycoprotein considered a cancer-stem cell (CSC) marker that binds to proteins and other ligands to activate signal transduction the control essential functions in proliferation, migration, and tumor invasion [128]. AS of CD44 can generate up to twelve SVs with different biological activities [129]. The switch between variants has been associated with EMT regulation [130]. In HCC, CD44v3 promotes metastasis, and its up-regulation is favored by the overexpression of nicotinamide N-methyltransferase (NNMT) [131]. Interestingly, studies in non-liver tumor tissue suggest that the generation of different CD44 variants may be affected by overexpression of ESRP1 and hnRNPM, two critical splicing factors involved in EMT regulation [132,133,134].

4.1.14. Genome Instability

Moreover, isoforms derived from aberrant splicing can also promote genome instability in HCC [135]. This is the case of the HCC-induced isoform SRSF3-Iso2, which has been implicated in the incorporation of introns 1 and 2 into CDCA5 mRNA, resulting in reduced levels of sororin, a protein essential to maintain sister chromatid cohesion [68].
Gene transcription and alterations in the expression of RBPs are important sources of replication stress and genome instability related to the pathogenesis of aging and neoplastic diseases, constituting one of the hallmarks of cancer [68]. Moreover, isoforms derived from aberrant splicing can also promote genome instability in HCC [135]. In this context, the splicing regulator SLU7 safeguards genomic integrity, preventing the formation of R-loops and the induction of DNA damage and ensuring proper chromosome segregation during mitosis [68]. Its expression is significantly down-regulated in HCC [66], which results in enhanced oxidative stress, DNA damage, mitotic aberrations, and genome instability, leading to an enhanced risk of carcinogenesis. Moreover, SLU7 is essential to prevent an increased expression of SRSF3 truncated proteins by dysregulation of SRSF3 splicing. The HCC-induced isoform SRSF3-Iso2 is involved in the incorporation of introns 1 and 2 into CDCA5 mRNA resulting in reduced levels of sororin, a protein essential to maintain sister chromatid cohesion [68]. The preservation of normal hepatic SRSF3 function appears to be critical to prevent liver carcinogenesis, since specific deletion of this gene in hepatocytes results in the spontaneous development of liver injury and HCC [86].
Other AS events related to chromosome instability occurring in HCC include exon 4 deletion (MAD1beta) in the mitotic arrest deficient 1 (MAD1) gene [136] and exon 6 deletion (AURKB-Sv2) in the serine/threonine kinase Aurora B (AURKB) [137].
Table 2. Alternative splicing of genes involved in hepatocellular carcinoma development and metastasis.
Table 2. Alternative splicing of genes involved in hepatocellular carcinoma development and metastasis.
GeneVariantSplicing EventConsequencesRefs.
AXLAXL-SExon 10 skippingEnhanced migration; Distant metastasis[124]
CD44CD44v3Lacks multiple coding exons compared to variant 1Promotes metastasis[131]
CDH17ΔEX7CDH17Exon 7 skippingDecreased OS; High tumor recurrence[109]
DNMT3b3DNMT3b4Loss of methyltransferase motifs IX and XInduced DNA hypomethylation and carcinogenesis[121]
FGFR2FGFR2-IIIb/IIIcMutually exclusive exonsPromoted invasion and metastasis[114]
FGFR3FGFR3-IIIbAlternative exon 8 inclusionIncreased proliferation in vitro; Increased tumor growth in vivo; Apoptosis inhibition[118]
FGFR3FGFR3-IIIcAlternative exon 9 inclusionIncreased tumor growth in vivo; Apoptosis inhibition[118]
FGFR3FGFR3Δ7–9Exons 7 to 9 skippingActivation of AKT and decreased expression of PTEN; Enhanced cell proliferation and tumor growth; Enhanced cell motility; Activation of EMT; Distant metastasis in vivo[119]
KLF6KLF6-SV(1,2)Exons 2 (partial) and 3 skippingEnhanced tumorigenesis and aggressiveness[112,113]
NF2Merlin Δ2–4Exons 2–4 skippingEnhanced migration and invasion; Activation of stemness; Distant metastasis.[125]
OPNOPN-bExon 5 skippingEnhanced invasiveness[127]
PDSS2PDSS2Δ2Exon 2 skippingEnhanced migration; Distant metastasis; Decreased OS[123]
TP53Short SVs
Δ40p53α
Several exons skippingVarious effects on cell proliferation[101,103,104,106]
TP73ΔEx2/3p73Exon 2 and 3 skippingTumor development in vivo[108]
XPB1XBP1sIntron skipping and frameshiftDistant metastasis; Poor prognosis[112]
EMT, epithelial-mesenchymal transition; OS, overall survival; SV, splicing variant.

4.2. Role of SVs in the Onset, Progression, and Metastasis of iCCA

Although more than six hundred genes have been reported to undergo AS in iCCA [38], there is scarce information on their role in iCCA initiation and metastasis. Several SVs associated with iCCA development are commented below and summarized in Table 3.

4.2.1. FOXP3

Recently, a prognosis signature of genomic instability-related AS events has been identified for CCA data available at TCGA, without discriminating the anatomical location of the tumor [7]. Moreover, this signature could be associated with the presence of infiltrating immune cells and the expression of PD-L1 [7], having, therefore therapeutic impact. Thus, skipping exons 3 and 4 in FOXP3 mRNA leads to a frameshift in the open reading frame (ORF) and hence in the amino acid sequence of FOXP3 results in a different protein involved in the development of CD25+ regulatory T cells [138]. This SV has also been detected in CCA cell lines [139]. Although the implication of this SV in iCCA onset is not well known, in other tumors, such as melanoma, this variant has been associated with immune suppression [138].

4.2.2. TFF2

Despite aberrant splicing is usually related to cancer promotion, some cases show the opposite association. For instance, this is the case of Trefoil factor 2 (TFF2), a secreted protein of the gastrointestinal mucosa that is up-regulated in many cancer types, including iCCA [140]. The SV generated by exon 2-skipping (ΔEX2TFF2) of TFF2 pre-mRNA, which results in the formation of a premature stop codon, has been identified in iCCA [141]. This non-functional TFF2 SV, in which most of the protein structure is lost, cannot mediate cell proliferation like the canonical wtTFF2 isoform does, which results in tumor progression inhibition. Interestingly, ΔEX2TFF2 expression correlated with more prolonged survival of iCCA patients. In contrast, the abundance of wtTTF2 in tumor tissue was associated with lower OS of these patients, suggesting that ΔEX2TFF2 could counteract the cancer-promoting activity of wtTFF2 in iCCA [141].

4.2.3. BAP1

Germline mutations in the tumor suppressor gene BAP1, encoding the BRCA1 associated protein-1 (BAP1), cause the so-called BAP1 tumor predisposition syndrome (BAP1-TPDS), characterized by an increased propensity to develop several types of cancer. In a human mesothelioma cell line, mutations in BAP1 pre-mRNA have been shown to contribute to tumorigenesis by disrupting normal splicing [142]. The mutations found in BAP1-TPDS patients affect ubiquitin carboxyl hydrolase (UCH) and nuclear localization signal (NLS) domains and create a non-functional truncated protein, disrupting BAP1 deubiquitinating activity and nuclear localization, both required for BAP1 tumor suppressor function [143]. Evidence for the relationship between germline mutations in 3′SS or 5′SS of exon 4 of BAP1 pre-mRNA and the development of iCCA has been reported [144].

4.2.4. AGR2

Regarding metastasis, several connections between AS and iCCA have been reported. Anterior gradient protein 2 (AGR2) is a member of the protein disulfide isomerases family, which is involved in protein modification and folding. AGR2 is up-regulated in a wide variety of solid tumors, where it is involved in their progression and metastasis [145]. AGR2 has several exon-skipping SVs, from AGR2vA to AGR2vH. The SV lacking exons 1 and 4–7 (AGR2vH) has been found in tumor samples derived from iCCA patients [146]. When overexpressed in CCA cell lines, this SV promotes not only survival and apoptosis evasion, thereby favoring cancer progression, but also metastasis-associated phenotypes [147], by increasing cell motility and up-regulating the expression of the mesenchymal marker vimentin [148]. On the contrary, down-regulation of AGR2vH has been found to reduce cell migration capacity [148].

4.2.5. WISP1

Wnt-inducible secreted protein 1 (WISP1, also known as CNN4) belongs to the cysteine-rich CCN family of proteins, which are involved in cell-matrix interaction and mediate cell adhesion, migration, proliferation, and apoptosis [149]. The SV WISP1v, lacking exon 3, has been found in approximately 50% of iCCA samples examined, whereas it was absent in the adjacent non-tumor liver tissue [150]. In the same study, WISP1v mRNA abundance was associated with perineural and lymphatic invasion and reduced OS after surgery. Moreover, WISP1v expression in HuCCT1 cells induced rapid growth and migration through p38 MAPK signaling activation [150].

4.2.6. PKM

PKM2 is an SV of pyruvate kinase (PKM), a rate-limiting enzyme in glycolysis. Immunohistochemical analysis of iCCA and hilar CCA revealed that PKM2 was up-regulated in tumor tissue compared with adjacent non-tumor liver tissue [151,152]. Higher PKM2 has been associated with enhanced proliferation and invasiveness of CCA cells, whereas patients presented shorter OS [151,152], which was related to lymph node invasion and distant metastasis [151]. Consistently, wound healing experiments showed that CCA cells transfected with shRNAs against PKM2 had lower motility than those with endogenous enzyme levels. Finally, in an orthotopic xenografted-mouse model of CCA, silencing of PKM2 expression remarkably inhibited both tumor growth and metastasis [152].

4.2.7. PTGER3

The AS of pre-mRNA of prostaglandin E2 receptor 3 (EP3, PTGER3) can generate eleven SVs. The SV EP3–4, which has been detected in several human tissues, such as gastric mucosa, mammary artery, and pulmonary vessels. This shorter mRNA includes exon 1, 2a, 5, and 10, which results in a truncated protein [153]. The overexpression of this SV in iCCA cells (CCLP1 and HuCCT1) enhances the expression of c-Myc and SNAIL, which supports the role of EP3–4 in iCCA progression and metastasis [154].

4.2.8. CD44

Among CD44 SVs, CD44v6 promotes EMT and activates the TGF-β pathway. CD44v6 expression was specifically found in proliferating cholangiocytes and CCA samples but not in healthy bile ducts [155]. In addition, patients with iCCA expressing CD44v6 and CD44v8–10 had shorter relapse-free survival (RFS) and OS than those without expression of these CD44 SVs. Moreover, the levels CD44v6 and CD44v8–10 proteins in serum were significantly increased in the recurrence group for early-stage iCCA, suggesting that soluble forms of CD44v6 and CD44v8–10 could serve as markers of post-operative iCCA recurrence [156].
Table 3. Alternative splicing of genes involved in intrahepatic cholangiocarcinoma development and metastasis.
Table 3. Alternative splicing of genes involved in intrahepatic cholangiocarcinoma development and metastasis.
GeneVariantSplicing EventConsequenceRefs.
AGR2AGR2vA to HSeveral combinations of exons 2 to 7 skippingAffect cancer cell survival and migration[148]
BAP1BAP1 p.E685VMultiple SV lacking exons 14–17Promote tumorigenesis[142]
CD44CD44v6; CD44v8–10Exon 6 skipping; exon 8–10 skippingDecreased OS and RFS; Tumor recurrence[156]
FOXP3FOXP3Exons 2–4 skippingImmune suppression[138,139]
PKMPKM2Mutual exclusive exons: exon 9 skipping/exon 10 retentionDecreased OS; Enhanced risk of metastasis[151,152]
PTGER3EP3–4Contains exon 1, 2a, 5, and 10Enhanced cell proliferation, migration, and invasion[153,154]
TFF2ΔEX2TFF2Exon 2 skippingIncreased OS[141]
WISP1WISP1vExon 3 skippingDecreased OS; Induction of invasion[150]
OS, overall survival; RFS, relapse-free survival; SV, splicing variant.

5. Altered Splicing of Genes Involved in the Resistance of Liver Cancer to Anticancer Drugs

The lack of response of liver cancer to drug treatment is one of the fundamental reasons for the high mortality caused by these tumors when they are diagnosed in an advanced non-operable stage, which unfortunately is a frequent situation in everyday clinical practice nowadays. The impact of AS on drug resistance mechanisms of cancer cells has recently been reviewed [157]. Regarding HCC and iCCA, the available data are still scarce. Hindering the access of anticancer drugs to their intracellular targets notably reduces their efficacy. This mechanism of pharmacoresistance (MPR), classified as MPR-1a [158,159], plays a crucial role in the lack of response. The uptake by HCC cells of sorafenib, a first-line drug against HCC, is mainly carried out through the organic cation transporter 1 (OCT1, SLC22A1) [160]. AS has a marked impact on this process because several OCT1 SVs are generated, most of them by exon skipping (exon 10, exon 9, or exon 9 + 10 skipping) and are translated into shorter OCT1 isoforms without transport activity. These non-functional SVs are abundantly expressed in HCC and iCCA [161]. Apart from sorafenib, a recent study has demonstrated that the failure of therapies targeting the androgen receptor (AR) in HCC could be associated with the expression of truncated AR-SVs [162] capable of driving AR signaling in the presence of antiandrogen therapy.
Drug resistance has also been associated with metabolic reprogramming in HCC, and more specifically, with the switch from oxidative phosphorylation to aerobic glycolysis [163,164]. As previously mentioned, PKM2 is a rate-limiting enzyme in the glycolytic pathway. This PKM isoform is produced by an AS mechanism that can be triggered, for instance, by down-regulation of the splicing factor SLU7 [69] or up-regulation of the RNA helicase MTR4 both events described in HCC [79].
Immune checkpoint inhibitors (ICI)-based immunotherapy is emerging as a promising strategy for treating advanced liver tumors [165,166]. A recent study comprehensively evaluated the differential AS events between HCC or iCCA tissues and the corresponding normal tissues focusing on immune features [167]. Interestingly, this work identified AS-related events that were not only associated with the prognosis of these patients but also with the formation of the differential immune microenvironment of these tumors. Unraveling the involvement of AS alterations in shaping the immune landscape of liver cancers may contribute to understanding their pathogenesis and improving the efficacy of ICI-based therapies.

6. Conclusions and Perspectives

The variability present in our proteome, thanks to the intervention of the splicing machinery during pre-mRNA maturation, constitutes a notable physiological advantage conferring the required flexibility to many cellular functions. Nevertheless, on the other side of the coin, AS may also provide tumor cells with dynamic and adaptative phenotypic traits that permit them to escape the host’s immune system and resist pharmacological treatment. The expression of SVs can lead to the loss of cell viability but also the development of a more malignant phenotype, owing to enhanced proliferative activity, reduced activation of pro-apoptotic pathways, enhanced capacity for local invasion, long-distance metastasis, and drug resistance. Several hundred splicing events have been described to occur differentially in liver cancer compared to normal liver tissue. Some of them have also been detected in other cancers. However, the complexity of the pre-mRNA maturation process, the differences in the transcriptome among tissues, and the heterogeneity concerning each type of cancer prevent us from reaching any general conclusion.
The study of the impact of AS on protein domain boundaries could offer the rationale to interpret the relevance of SV in liver cancer biology. Although, as commented above, there is some available information in this respect, for instance, regarding FGFR3 [119,120] and BAP1 [143] variants, this analysis requires individual gene investigation, which in many cases is still missing.
Identifying the impact of AS on cancer cell biology is crucial for the development of novel strategies targeting these mechanisms. Interfering with pro-carcinogenic AS events could reduce the malignancy of HCC and iCCA and eventually improve the outcomes of these patients. In this respect, there are already ongoing attempts to develop small molecules able to interact with splicing factors to modulate AS. However, there is still a long way for these drugs to reach clinical practice because they lack cancer cell specificity, and most of their molecular targets are constitutive elements of the spliceosome, whose pharmacological manipulation can have an impact on the correct expression of many different genes resulting in noxious effects for healthy cells [168,169].

Author Contributions

J.J.G.M., M.R., M.S., E.L., M.A., S.O.-R., C.B., M.A.A. and E.H. have contributed substantially to the elaboration of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the CIBERehd (EHD15PI05/2016) and “Fondo de Investigaciones Sanitarias, Instituto de Salud Carlos III”, Spain (PI16/00598 and PI19/00819, co-funded by European Regional Development Fund/European Social Fund, “Investing in your future”); Spanish Ministry of Economy, Industry and Competitiveness (SAF2016-75197-R,); “Junta de Castilla y Leon” (SA063P17); AECC Scientific Foundation (2017/2020), Spain; “Proyectos de Investigación. Modalidad C2”, University of Salamanca (18.K137/463AC01 and 18.K140/463AC01); “Centro Internacional sobre el Envejecimiento” (OLD-HEPAMARKER, 0348_CIE_6_E), Spain and Fundación University of Salamanca, Spain (PC-TCUE18-20_051); Fundació Marato TV3 (Ref. 201916-31). FEDER/Ministerio de Ciencia, Innovación y Universidades-Agencia Estatal de Investigación (PID2019-104878RB-100 and PI02019-104265RB-100 to MAA and CB, respectively). M.R. was supported by a predoctoral scholarship (FPU) funded by the Ministry of Science, Innovation and Universities, Spain. S.O.R. was supported by a postdoctoral contract (FPU) funded by the “Junta de Castilla y Leon” (SA074P20).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

3′SS, 3′ acceptor site; 5′SS, 5′ donor site; AR, androgen receptor; AREG, amphiregulin; AS, alternative splicing; AURKB, aurora kinase B; AXL, AXL tyrosine kinase receptor; AXL-S, AXL short; BAP1, BRCA1 associated protein-1; BAP1-TPDS, BAP1 tumor predisposition syndrome; BPS, branch point site; CCA, cholangiocarcinoma; CDH17, cadherin 17; CES, Carboxylesterase; CSC, cancer stem cell; DNMT, methyltransferase; EGFR, epidermal growth factor receptor; EMT, epithelial-mesenchymal transition; ENT1, Equilibrative nucleoside transporter 1; ER, estrogen receptor; ESE, exonic splicing enhancers; ESS, exonic splicing silencers; FGFR2, fibroblast growth factor receptor 2; HCC, hepatocellular carcinoma; HCCDB, Integrative Molecular Database of HCC; hnRNP, heterogeneous nuclear ribonucleoprotein; HR, homologous recombination; iCCA, intrahepatic cholangiocarcinoma; ICI, immune checkpoint inhibitors; ISE, intronic splicing enhancers; ISS, intronic splicing silencers; KLF6, Krüppel-like factor 6; lncRNA, long noncoding RNA; Merlin, moesin-ezrin-radixin-like protein; MPRs, mechanisms of pharmacoresistance; NLS, nuclear localization signal; NNMT, nicotinamide NN-methyltransferase; NSCLC, non-small cell lung cancer; OCT1, Organic cation transporter 1; OPN, osteopontin; ORF, open reading frame; OS, overall survival; PARPi, PARP inhibitors; PDSS2, prenyl diphosphate synthase subunit 2; PKM, pyruvate kinase gene; PPT, polypyrimidine tract; PRPF, pre-mRNA processing factors; PTBP, polypyrimidine tract binding proteins; PTGER3, Prostaglandin E2 receptor EP3; RBP, RNA-binding protein; RFS, relapse-free survival; Sf1, splicing factor 1; SLU7, step II splicing regulator SLU7; SNP, single nucleotide polymorphism; snRNP, small nuclear ribonucleoprotein; snRNA, small nuclear RNA; SV, splicing variant; TFF2, Trefoil factor 2; TKIs, tyrosine kinase inhibitors; TP, Thymidine; UCH, ubiquitin carboxyl hydrolase; UPR, unfolded protein response; WISP1, Wnt-inducible secreted protein 1; XBP1s, X-box binding protein 1.

References

  1. Yosudjai, J.; Wongkham, S.; Jirawatnotai, S.; Kaewkong, W. Aberrant mRNA splicing generates oncogenic RNA isoforms and contributes to the development and progression of cholangiocarcinoma. Biomed. Rep. 2019, 10, 147–155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Tremblay, M.P.; Armero, V.E.; Allaire, A.; Boudreault, S.; Martenon-Brodeur, C.; Durand, M.; Lapointe, E.; Thibault, P.; Tremblay-Letourneau, M.; Perreault, J.P.; et al. Global profiling of alternative RNA splicing events provides insights into molecular differences between various types of hepatocellular carcinoma. BMC Genom. 2016, 17, 683. [Google Scholar] [CrossRef] [Green Version]
  3. Li, S.; Hu, Z.; Zhao, Y.; Huang, S.; He, X. Transcriptome-Wide Analysis Reveals the Landscape of Aberrant Alternative Splicing Events in Liver Cancer. Hepatology 2019, 69, 359–375. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Xiong, Y.; Yang, G.; Wang, K.; Riaz, M.; Xu, J.; Lv, Z.; Zhou, H.; Li, Q.; Li, W.; Sun, J.; et al. Genome-Wide Transcriptional Analysis Reveals Alternative Splicing Event Profiles in Hepatocellular Carcinoma and Their Prognostic Significance. Front. Genet. 2020, 11, 879. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, D.; Duan, Y.; Wang, Z.; Lin, J. Systematic profiling of a novel prognostic alternative splicing signature in hepatocellular carcinoma. Oncol. Rep. 2019, 42, 2450–2472. [Google Scholar] [CrossRef]
  6. Wu, H.Y.; Wei, Y.; Liu, L.M.; Chen, Z.B.; Hu, Q.P.; Pan, S.L. Construction of a model to predict the prognosis of patients with cholangiocarcinoma using alternative splicing events. Oncol. Lett. 2019, 18, 4677–4690. [Google Scholar] [CrossRef] [PubMed]
  7. Lin, Z.; Gong, J.; Zhong, G.; Hu, J.; Cai, D.; Zhao, L.; Zhao, Z. Identification of Mutator-Derived Alternative Splicing Signatures of Genomic Instability for Improving the Clinical Outcome of Cholangiocarcinoma. Front. Oncol. 2021, 11, 666847. [Google Scholar] [CrossRef]
  8. Nilsen, T.W.; Graveley, B.R. Expansion of the eukaryotic proteome by alternative splicing. Nature 2010, 463, 457–463. [Google Scholar] [CrossRef] [Green Version]
  9. van den Hoogenhof, M.M.; Pinto, Y.M.; Creemers, E.E. RNA Splicing: Regulation and Dysregulation in the Heart. Circ. Res. 2016, 118, 454–468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Wang, E.T.; Sandberg, R.; Luo, S.; Khrebtukova, I.; Zhang, L.; Mayr, C.; Kingsmore, S.F.; Schroth, G.P.; Burge, C.B. Alternative isoform regulation in human tissue transcriptomes. Nature 2008, 456, 470–476. [Google Scholar] [CrossRef] [Green Version]
  11. Mthembu, N.N.; Mbita, Z.; Hull, R.; Dlamini, Z. Abnormalities in alternative splicing of angiogenesis-related genes and their role in HIV-related cancers. HIV AIDS 2017, 9, 77–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Kameyama, T.; Suzuki, H.; Mayeda, A. Re-splicing of mature mRNA in cancer cells promotes activation of distant weak alternative splice sites. Nucleic Acids Res. 2012, 40, 7896–7906. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Chua, H.H.; Kameyama, T.; Mayeda, A.; Yeh, T.H. Cancer-Specifically Re-Spliced TSG101 mRNA Promotes Invasion and Metastasis of Nasopharyngeal Carcinoma. Int. J. Mol. Sci. 2019, 20, 773. [Google Scholar] [CrossRef] [Green Version]
  14. Hall, S.L.; Padgett, R.A. Requirement of U12 snRNA for in vivo splicing of a minor class of eukaryotic nuclear pre-mRNA introns. Science 1996, 271, 1716–1718. [Google Scholar] [CrossRef]
  15. Chen, M.; Manley, J.L. Mechanisms of alternative splicing regulation: Insights from molecular and genomics approaches. Nat. Rev. Mol. Cell Biol. 2009, 10, 741–754. [Google Scholar] [CrossRef]
  16. Tarn, W.Y.; Steitz, J.A. Pre-mRNA splicing: The discovery of a new spliceosome doubles the challenge. Trends Biochem. Sci. 1997, 22, 132–137. [Google Scholar] [CrossRef]
  17. Patel, A.A.; Steitz, J.A. Splicing double: Insights from the second spliceosome. Nat. Rev. Mol. Cell Biol. 2003, 4, 960–970. [Google Scholar] [CrossRef]
  18. Tacke, R.; Manley, J.L. Determinants of SR protein specificity. Curr. Opin. Cell Biol. 1999, 11, 358–362. [Google Scholar] [CrossRef]
  19. Zhang, Y.; Madl, T.; Bagdiul, I.; Kern, T.; Kang, H.S.; Zou, P.; Mausbacher, N.; Sieber, S.A.; Kramer, A.; Sattler, M. Structure, phosphorylation and U2AF65 binding of the N-terminal domain of splicing factor 1 during 3’-splice site recognition. Nucleic Acids Res. 2013, 41, 1343–1354. [Google Scholar] [CrossRef]
  20. Xu, Y.Z.; Newnham, C.M.; Kameoka, S.; Huang, T.; Konarska, M.M.; Query, C.C. Prp5 bridges U1 and U2 snRNPs and enables stable U2 snRNP association with intron RNA. EMBO J. 2004, 23, 376–385. [Google Scholar] [CrossRef] [Green Version]
  21. Fredericks, A.M.; Cygan, K.J.; Brown, B.A.; Fairbrother, W.G. RNA-Binding Proteins: Splicing Factors and Disease. Biomolecules 2015, 5, 893–909. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Charenton, C.; Wilkinson, M.E.; Nagai, K. Mechanism of 5′ splice site transfer for human spliceosome activation. Science 2019, 364, 362–367. [Google Scholar] [CrossRef] [PubMed]
  23. Sidarovich, A.; Will, C.L.; Anokhina, M.M.; Ceballos, J.; Sievers, S.; Agafonov, D.E.; Samatov, T.; Bao, P.; Kastner, B.; Urlaub, H.; et al. Identification of a small molecule inhibitor that stalls splicing at an early step of spliceosome activation. Elife 2017, 6, e23533. [Google Scholar] [CrossRef] [PubMed]
  24. Fica, S.M.; Tuttle, N.; Novak, T.; Li, N.S.; Lu, J.; Koodathingal, P.; Dai, Q.; Staley, J.P.; Piccirilli, J.A. RNA catalyses nuclear pre-mRNA splicing. Nature 2013, 503, 229–234. [Google Scholar] [CrossRef] [Green Version]
  25. Jacquier, A. Self-splicing group II and nuclear pre-mRNA introns: How similar are they? Trends Biochem. Sci. 1990, 15, 351–354. [Google Scholar] [CrossRef]
  26. Will, C.L.; Luhrmann, R. Spliceosome structure and function. Cold Spring Harb. Perspect. Biol. 2011, 3, a003707. [Google Scholar] [CrossRef] [Green Version]
  27. Makarova, O.V.; Makarov, E.M.; Liu, S.; Vornlocher, H.P.; Luhrmann, R. Protein 61K, encoded by a gene (PRPF31) linked to autosomal dominant retinitis pigmentosa, is required for U4/U6*U5 tri-snRNP formation and pre-mRNA splicing. EMBO J. 2002, 21, 1148–1157. [Google Scholar] [CrossRef] [Green Version]
  28. Graveley, B.R. Sorting out the complexity of SR protein functions. RNA 2000, 6, 1197–1211. [Google Scholar] [CrossRef] [Green Version]
  29. Long, J.C.; Caceres, J.F. The SR protein family of splicing factors: Master regulators of gene expression. Biochem. J. 2009, 417, 15–27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Blencowe, B.J. Exonic splicing enhancers: Mechanism of action, diversity and role in human genetic diseases. Trends Biochem. Sci. 2000, 25, 106–110. [Google Scholar] [CrossRef]
  31. Martinez-Contreras, R.; Cloutier, P.; Shkreta, L.; Fisette, J.F.; Revil, T.; Chabot, B. hnRNP proteins and splicing control. Adv. Exp. Med. Biol. 2007, 623, 123–147. [Google Scholar] [CrossRef] [PubMed]
  32. Naro, C.; Sette, C. Phosphorylation-mediated regulation of alternative splicing in cancer. Int. J. Cell Biol. 2013, 2013, 151839. [Google Scholar] [CrossRef] [PubMed]
  33. Padgett, R.A. New connections between splicing and human disease. Trends Genet. 2012, 28, 147–154. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. He, C.; Zhou, F.; Zuo, Z.; Cheng, H.; Zhou, R. A global view of cancer-specific transcript variants by subtractive transcriptome-wide analysis. PLoS ONE 2009, 4, e4732. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Sebestyen, E.; Zawisza, M.; Eyras, E. Detection of recurrent alternative splicing switches in tumor samples reveals novel signatures of cancer. Nucleic Acids Res. 2015, 43, 1345–1356. [Google Scholar] [CrossRef] [Green Version]
  36. Danan-Gotthold, M.; Golan-Gerstl, R.; Eisenberg, E.; Meir, K.; Karni, R.; Levanon, E.Y. Identification of recurrent regulated alternative splicing events across human solid tumors. Nucleic Acids Res. 2015, 43, 5130–5144. [Google Scholar] [CrossRef] [Green Version]
  37. Luo, C.; Cheng, Y.; Liu, Y.; Chen, L.; Liu, L.; Wei, N.; Xie, Z.; Wu, W.; Feng, Y. SRSF2 Regulates Alternative Splicing to Drive Hepatocellular Carcinoma Development. Cancer Res. 2017, 77, 1168–1178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Chen, Y.; Liu, D.; Liu, P.; Chen, Y.; Yu, H.; Zhang, Q. Identification of biomarkers of intrahepatic cholangiocarcinoma via integrated analysis of mRNA and miRNA microarray data. Mol. Med. Rep. 2017, 15, 1051–1056. [Google Scholar] [CrossRef] [Green Version]
  39. Kim, E.; Goren, A.; Ast, G. Insights into the connection between cancer and alternative splicing. Trends Genet. 2008, 24, 7–10. [Google Scholar] [CrossRef]
  40. Srebrow, A.; Kornblihtt, A.R. The connection between splicing and cancer. J. Cell Sci. 2006, 119, 2635–2641. [Google Scholar] [CrossRef] [Green Version]
  41. Berasain, C.; Goni, S.; Castillo, J.; Latasa, M.U.; Prieto, J.; Avila, M.A. Impairment of pre-mRNA splicing in liver disease: Mechanisms and consequences. World J. Gastroenterol. 2010, 16, 3091–3102. [Google Scholar] [CrossRef] [PubMed]
  42. Blaustein, M.; Pelisch, F.; Srebrow, A. Signals, pathways and splicing regulation. Int. J. Biochem. Cell Biol. 2007, 39, 2031–2048. [Google Scholar] [CrossRef]
  43. Soto, M.; Reviejo, M.; Al-Abdulla, R.; Romero, M.R.; Macias, R.I.R.; Boix, L.; Bruix, J.; Serrano, M.A.; Marin, J.J.G. Relationship between changes in the exon-recognition machinery and SLC22A1 alternative splicing in hepatocellular carcinoma. Biochim. Biophys. Acta Mol. Basis Dis. 2020, 1866, 165687. [Google Scholar] [CrossRef] [PubMed]
  44. Phoomak, C.; Park, D.; Silsirivanit, A.; Sawanyawisuth, K.; Vaeteewoottacharn, K.; Detarya, M.; Wongkham, C.; Lebrilla, C.B.; Wongkham, S. O-GlcNAc-induced nuclear translocation of hnRNP-K is associated with progression and metastasis of cholangiocarcinoma. Mol. Oncol. 2019, 13, 338–357. [Google Scholar] [CrossRef]
  45. Pereira, B.; Billaud, M.; Almeida, R. RNA-Binding Proteins in Cancer: Old Players and New Actors. Trends Cancer 2017, 3, 506–528. [Google Scholar] [CrossRef] [PubMed]
  46. Xu, W.; Huang, H.; Yu, L.; Cao, L. Meta-analysis of gene expression profiles indicates genes in spliceosome pathway are up-regulated in hepatocellular carcinoma (HCC). Med. Oncol. 2015, 32, 96. [Google Scholar] [CrossRef]
  47. Lopez-Canovas, J.L.; Del Rio-Moreno, M.; Garcia-Fernandez, H.; Jimenez-Vacas, J.M.; Moreno-Montilla, M.T.; Sanchez-Frias, M.E.; Amado, V.; Fernando, L.; Fondevila, M.F.; Ciria, R.; et al. Splicing factor SF3B1 is overexpressed and implicated in the aggressiveness and survival of hepatocellular carcinoma. Cancer Lett. 2021, 496, 72–83. [Google Scholar] [CrossRef]
  48. Sun, C. The SF3b complex: Splicing and beyond. Cell Mol. Life Sci. 2020, 77, 3583–3595. [Google Scholar] [CrossRef] [Green Version]
  49. Kfir, N.; Lev-Maor, G.; Glaich, O.; Alajem, A.; Datta, A.; Sze, S.K.; Meshorer, E.; Ast, G. SF3B1 association with chromatin determines splicing outcomes. Cell Rep. 2015, 11, 618–629. [Google Scholar] [CrossRef] [Green Version]
  50. Zhang, Y.; Yuan, Z.; Jiang, Y.; Shen, R.; Gu, M.; Xu, W.; Gu, X. Inhibition of Splicing Factor 3b Subunit 1 (SF3B1) Reduced Cell Proliferation, Induced Apoptosis and Resulted in Cell Cycle Arrest by Regulating Homeobox A10 (HOXA10) Splicing in AGS and MKN28 Human Gastric Cancer Cells. Med. Sci. Monit. 2020, 26, e919460. [Google Scholar] [CrossRef]
  51. Hwang, H.M.; Heo, C.K.; Lee, H.J.; Kwak, S.S.; Lim, W.H.; Yoo, J.S.; Yu, D.Y.; Lim, K.J.; Kim, J.Y.; Cho, E.W. Identification of anti-SF3B1 autoantibody as a diagnostic marker in patients with hepatocellular carcinoma. J. Transl. Med. 2018, 16, 177. [Google Scholar] [CrossRef]
  52. Biankin, A.V.; Waddell, N.; Kassahn, K.S.; Gingras, M.C.; Muthuswamy, L.B.; Johns, A.L.; Miller, D.K.; Wilson, P.J.; Patch, A.M.; Wu, J.; et al. Pancreatic cancer genomes reveal aberrations in axon guidance pathway genes. Nature 2012, 491, 399–405. [Google Scholar] [CrossRef] [PubMed]
  53. Maguire, S.L.; Leonidou, A.; Wai, P.; Marchio, C.; Ng, C.K.; Sapino, A.; Salomon, A.V.; Reis-Filho, J.S.; Weigelt, B.; Natrajan, R.C. SF3B1 mutations constitute a novel therapeutic target in breast cancer. J. Pathol. 2015, 235, 571–580. [Google Scholar] [CrossRef] [Green Version]
  54. Jimenez-Vacas, J.M.; Herrero-Aguayo, V.; Gomez-Gomez, E.; Leon-Gonzalez, A.J.; Saez-Martinez, P.; Alors-Perez, E.; Fuentes-Fayos, A.C.; Martinez-Lopez, A.; Sanchez-Sanchez, R.; Gonzalez-Serrano, T.; et al. Spliceosome component SF3B1 as novel prognostic biomarker and therapeutic target for prostate cancer. Transl. Res. 2019, 212, 89–103. [Google Scholar] [CrossRef] [PubMed]
  55. Inoue, D.; Chew, G.L.; Liu, B.; Michel, B.C.; Pangallo, J.; D’Avino, A.R.; Hitchman, T.; North, K.; Lee, S.C.; Bitner, L.; et al. Spliceosomal disruption of the non-canonical BAF complex in cancer. Nature 2019, 574, 432–436. [Google Scholar] [CrossRef] [PubMed]
  56. te Raa, G.D.; Derks, I.A.; Navrkalova, V.; Skowronska, A.; Moerland, P.D.; van Laar, J.; Oldreive, C.; Monsuur, H.; Trbusek, M.; Malcikova, J.; et al. The impact of SF3B1 mutations in CLL on the DNA-damage response. Leukemia 2015, 29, 1133–1142. [Google Scholar] [CrossRef]
  57. The Cancer Genome Atlas Research Network. Comprehensive and Integrative Genomic Characterization of Hepatocellular Carcinoma. Cell 2017, 169, 1327–1341.e1323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Nault, J.C.; Martin, Y.; Caruso, S.; Hirsch, T.Z.; Bayard, Q.; Calderaro, J.; Charpy, C.; Copie-Bergman, C.; Ziol, M.; Bioulac-Sage, P.; et al. Clinical Impact of Genomic Diversity From Early to Advanced Hepatocellular Carcinoma. Hepatology 2020, 71, 164–182. [Google Scholar] [CrossRef]
  59. Shen, Q.; Nam, S.W. SF3B4 as an early-stage diagnostic marker and driver of hepatocellular carcinoma. BMB Rep. 2018, 51, 57–58. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, Z.; Li, W.; Pang, Y.; Zhou, Z.; Liu, S.; Cheng, K.; Qin, Q.; Jia, Y.; Liu, S. SF3B4 is regulated by microRNA-133b and promotes cell proliferation and metastasis in hepatocellular carcinoma. EBioMedicine 2018, 38, 57–68. [Google Scholar] [CrossRef] [Green Version]
  61. Iguchi, T.; Komatsu, H.; Masuda, T.; Nambara, S.; Kidogami, S.; Ogawa, Y.; Hu, Q.; Saito, T.; Hirata, H.; Sakimura, S.; et al. Increased Copy Number of the Gene Encoding SF3B4 Indicates Poor Prognosis in Hepatocellular Carcinoma. Anticancer Res. 2016, 36, 2139–2144. [Google Scholar] [PubMed]
  62. Peng, N.; Li, J.; He, J.; Shi, X.; Huang, H.; Mo, Y.; Ye, H.; Wu, G.; Wu, F.; Xiang, B.; et al. c-Myc-mediated SNRPB upregulation functions as an oncogene in hepatocellular carcinoma. Cell Biol. Int. 2020, 44, 1103–1111. [Google Scholar] [CrossRef] [PubMed]
  63. Bielli, P.; Pagliarini, V.; Pieraccioli, M.; Caggiano, C.; Sette, C. Splicing Dysregulation as Oncogenic Driver and Passenger Factor in Brain Tumors. Cells 2019, 9, 10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Correa, B.R.; de Araujo, P.R.; Qiao, M.; Burns, S.C.; Chen, C.; Schlegel, R.; Agarwal, S.; Galante, P.A.; Penalva, L.O. Functional genomics analyses of RNA-binding proteins reveal the splicing regulator SNRPB as an oncogenic candidate in glioblastoma. Genome Biol. 2016, 17, 125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Liu, N.; Wu, Z.; Chen, A.; Wang, Y.; Cai, D.; Zheng, J.; Liu, Y.; Zhang, L. SNRPB promotes the tumorigenic potential of NSCLC in part by regulating RAB26. Cell Death Dis. 2019, 10, 667. [Google Scholar] [CrossRef]
  66. Castillo, J.; Goni, S.; Latasa, M.U.; Perugorria, M.J.; Calvo, A.; Muntane, J.; Bioulac-Sage, P.; Balabaud, C.; Prieto, J.; Avila, M.A.; et al. Amphiregulin induces the alternative splicing of p73 into its oncogenic isoform DeltaEx2p73 in human hepatocellular tumors. Gastroenterology 2009, 137, 1805–1815. [Google Scholar] [CrossRef]
  67. Garate-Rascon, M.; Recalde, M.; Jimenez, M.; Elizalde, M.; Azkona, M.; Uriarte, I.; Latasa, M.U.; Urtasun, R.; Bilbao, I.; Sangro, B.; et al. Splicing Factor SLU7 Prevents Oxidative Stress-Mediated Hepatocyte Nuclear Factor 4alpha Degradation, Preserving Hepatic Differentiation and Protecting From Liver Damage. Hepatology 2021, 74, 2791–2807. [Google Scholar] [CrossRef] [PubMed]
  68. Jimenez, M.; Urtasun, R.; Elizalde, M.; Azkona, M.; Latasa, M.U.; Uriarte, I.; Arechederra, M.; Alignani, D.; Barcena-Varela, M.; Alvarez-Sola, G.; et al. Splicing events in the control of genome integrity: Role of SLU7 and truncated SRSF3 proteins. Nucleic Acids Res. 2019, 47, 3450–3466. [Google Scholar] [CrossRef] [Green Version]
  69. Elizalde, M.; Urtasun, R.; Azkona, M.; Latasa, M.U.; Goni, S.; Garcia-Irigoyen, O.; Uriarte, I.; Segura, V.; Collantes, M.; Di Scala, M.; et al. Splicing regulator SLU7 is essential for maintaining liver homeostasis. J. Clin. Invest. 2014, 124, 2909–2920. [Google Scholar] [CrossRef] [Green Version]
  70. Ji, F.J.; Wu, Y.Y.; An, Z.; Liu, X.S.; Jiang, J.N.; Chen, F.F.; Fang, X.D. Expression of both poly r(C) binding protein 1 (PCBP1) and miRNA-3978 is suppressed in peritoneal gastric cancer metastasis. Sci. Rep. 2017, 7, 15488. [Google Scholar] [CrossRef] [Green Version]
  71. Zhang, M.; Wang, X.; Tan, J.; Zhao, M.; Lian, L.; Zhang, W. Poly r(C) binding protein (PCBP) 1 is a negative regulator of thyroid carcinoma. Am. J. Transl. Res. 2016, 8, 3567–3573. [Google Scholar]
  72. Luo, K.; Zhuang, K. High expression of PCBP2 is associated with progression and poor prognosis in patients with glioblastoma. Biomed. Pharmacother. 2017, 94, 659–665. [Google Scholar] [CrossRef]
  73. Chen, C.; Lei, J.; Zheng, Q.; Tan, S.; Ding, K.; Yu, C. Poly(rC) binding protein 2 (PCBP2) promotes the viability of human gastric cancer cells by regulating CDK2. FEBS Open Bio. 2018, 8, 764–773. [Google Scholar] [CrossRef]
  74. Zhang, X.; Hua, L.; Yan, D.; Zhao, F.; Liu, J.; Zhou, H.; Liu, J.; Wu, M.; Zhang, C.; Chen, Y.; et al. Overexpression of PCBP2 contributes to poor prognosis and enhanced cell growth in human hepatocellular carcinoma. Oncol. Rep. 2016, 36, 3456–3464. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Qin, G.; Tu, X.; Li, H.; Cao, P.; Chen, X.; Song, J.; Han, H.; Li, Y.; Guo, B.; Yang, L.; et al. Long Noncoding RNA p53-Stabilizing and Activating RNA Promotes p53 Signaling by Inhibiting Heterogeneous Nuclear Ribonucleoprotein K deSUMOylation and Suppresses Hepatocellular Carcinoma. Hepatology 2020, 71, 112–129. [Google Scholar] [CrossRef] [PubMed]
  76. Shu, H.; Hu, J.; Deng, H. miR-1249-3p accelerates the malignancy phenotype of hepatocellular carcinoma by directly targeting HNRNPK. Mol. Genet. Genomic. Med. 2019, 7, e00867. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Zhou, Z.J.; Dai, Z.; Zhou, S.L.; Hu, Z.Q.; Chen, Q.; Zhao, Y.M.; Shi, Y.H.; Gao, Q.; Wu, W.Z.; Qiu, S.J.; et al. HNRNPAB induces epithelial-mesenchymal transition and promotes metastasis of hepatocellular carcinoma by transcriptionally activating SNAIL. Cancer Res. 2014, 74, 2750–2762. [Google Scholar] [CrossRef] [Green Version]
  78. Yang, X.; Qu, S.; Wang, L.; Zhang, H.; Yang, Z.; Wang, J.; Dai, B.; Tao, K.; Shang, R.; Liu, Z.; et al. PTBP3 splicing factor promotes hepatocellular carcinoma by destroying the splicing balance of NEAT1 and pre-miR-612. Oncogene 2018, 37, 6399–6413. [Google Scholar] [CrossRef]
  79. Yu, L.; Kim, J.; Jiang, L.; Feng, B.; Ying, Y.; Ji, K.Y.; Tang, Q.; Chen, W.; Mai, T.; Dou, W.; et al. MTR4 drives liver tumorigenesis by promoting cancer metabolic switch through alternative splicing. Nat. Commun. 2020, 11, 708. [Google Scholar] [CrossRef]
  80. Karni, R.; de Stanchina, E.; Lowe, S.W.; Sinha, R.; Mu, D.; Krainer, A.R. The gene encoding the splicing factor SF2/ASF is a proto-oncogene. Nat. Struct. Mol. Biol. 2007, 14, 185–193. [Google Scholar] [CrossRef] [Green Version]
  81. Liu, J.; Huang, B.; Xiao, Y.; Xiong, H.M.; Li, J.; Feng, D.Q.; Chen, X.M.; Zhang, H.B.; Wang, X.Z. Aberrant expression of splicing factors in newly diagnosed acute myeloid leukemia. Oncol. Res. Treat. 2012, 35, 335–340. [Google Scholar] [CrossRef]
  82. Yang, S.; Jia, R.; Bian, Z. SRSF5 functions as a novel oncogenic splicing factor and is upregulated by oncogene SRSF3 in oral squamous cell carcinoma. Biochim. Biophys. Acta Mol. Cell Res. 2018, 1865, 1161–1172. [Google Scholar] [CrossRef] [PubMed]
  83. Liu, F.; Dai, M.; Xu, Q.; Zhu, X.; Zhou, Y.; Jiang, S.; Wang, Y.; Ai, Z.; Ma, L.; Zhang, Y.; et al. SRSF10-mediated IL1RAP alternative splicing regulates cervical cancer oncogenesis via mIL1RAP-NF-kappaB-CD47 axis. Oncogene 2018, 37, 2394–2409. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Zhou, X.; Wang, R.; Li, X.; Yu, L.; Hua, D.; Sun, C.; Shi, C.; Luo, W.; Rao, C.; Jiang, Z.; et al. Splicing factor SRSF1 promotes gliomagenesis via oncogenic splice-switching of MYO1B. J. Clin. Invest. 2019, 129, 676–693. [Google Scholar] [CrossRef] [PubMed]
  85. Sen, S.; Jumaa, H.; Webster, N.J. Splicing factor SRSF3 is crucial for hepatocyte differentiation and metabolic function. Nat. Commun. 2013, 4, 1336. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Sen, S.; Langiewicz, M.; Jumaa, H.; Webster, N.J. Deletion of serine/arginine-rich splicing factor 3 in hepatocytes predisposes to hepatocellular carcinoma in mice. Hepatology 2015, 61, 171–183. [Google Scholar] [CrossRef]
  87. Chen, D.; Zhao, Z.; Chen, L.; Li, Q.; Zou, J.; Liu, S. PPM1G promotes the progression of hepatocellular carcinoma via phosphorylation regulation of alternative splicing protein SRSF3. Cell Death Dis. 2021, 12, 722. [Google Scholar] [CrossRef]
  88. Zhang, C.; Shen, L.; Yuan, W.; Liu, Y.; Guo, R.; Luo, Y.; Zhan, Z.; Xie, Z.; Wu, G.; Wu, W.; et al. Loss of SRSF2 triggers hepatic progenitor cell activation and tumor development in mice. Commun. Biol. 2020, 3, 210. [Google Scholar] [CrossRef]
  89. Lowdon, R.F.; Wang, T. Epigenomic annotation of noncoding mutations identifies mutated pathways in primary liver cancer. PLoS ONE 2017, 12, e0174032. [Google Scholar] [CrossRef] [Green Version]
  90. Xu, Y.; Gao, X.D.; Lee, J.H.; Huang, H.; Tan, H.; Ahn, J.; Reinke, L.M.; Peter, M.E.; Feng, Y.; Gius, D.; et al. Cell type-restricted activity of hnRNPM promotes breast cancer metastasis via regulating alternative splicing. Genes Dev. 2014, 28, 1191–1203. [Google Scholar] [CrossRef] [Green Version]
  91. Tian, S.; Liu, J.; Sun, K.; Liu, Y.; Yu, J.; Ma, S.; Zhang, M.; Jia, G.; Zhou, X.; Shang, Y.; et al. Systematic Construction and Validation of an RNA-Binding Protein-Associated Model for Prognosis Prediction in Hepatocellular Carcinoma. Front. Oncol. 2020, 10, 597996. [Google Scholar] [CrossRef]
  92. Li, H.; Liu, J.; Shen, S.; Dai, D.; Cheng, S.; Dong, X.; Sun, L.; Guo, X. Pan-cancer analysis of alternative splicing regulator heterogeneous nuclear ribonucleoproteins (hnRNPs) family and their prognostic potential. J. Cell Mol. Med. 2020, 24, 11111–11119. [Google Scholar] [CrossRef]
  93. Lee, Y.; Rio, D.C. Mechanisms and Regulation of Alternative Pre-mRNA Splicing. Annu. Rev. Biochem. 2015, 84, 291–323. [Google Scholar] [CrossRef] [Green Version]
  94. Pensabene, M.; Spagnoletti, I.; Capuano, I.; Condello, C.; Pepe, S.; Contegiacomo, A.; Lombardi, G.; Bevilacqua, G.; Caligo, M.A. Two mutations of BRCA2 gene at exon and splicing site in a woman who underwent oncogenetic counseling. Ann. Oncol. 2009, 20, 874–878. [Google Scholar] [CrossRef]
  95. Matsukuma, K.E.; Mullins, F.M.; Dietz, L.; Zehnder, J.L.; Ford, J.M.; Chun, N.M.; Schrijver, I. Hereditary diffuse gastric cancer due to a previously undescribed CDH1 splice site mutation. Hum. Pathol. 2010, 41, 1200–1203. [Google Scholar] [CrossRef] [PubMed]
  96. Hansford, S.; Kaurah, P.; Li-Chang, H.; Woo, M.; Senz, J.; Pinheiro, H.; Schrader, K.A.; Schaeffer, D.F.; Shumansky, K.; Zogopoulos, G.; et al. Hereditary Diffuse Gastric Cancer Syndrome: CDH1 Mutations and Beyond. JAMA Oncol. 2015, 1, 23–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Bartram, M.P.; Mishra, T.; Reintjes, N.; Fabretti, F.; Gharbi, H.; Adam, A.C.; Gobel, H.; Franke, M.; Schermer, B.; Haneder, S.; et al. Characterization of a splice-site mutation in the tumor suppressor gene FLCN associated with renal cancer. BMC Med. Genet. 2017, 18, 53. [Google Scholar] [CrossRef]
  98. Diederichs, S.; Bartsch, L.; Berkmann, J.C.; Frose, K.; Heitmann, J.; Hoppe, C.; Iggena, D.; Jazmati, D.; Karschnia, P.; Linsenmeier, M.; et al. The dark matter of the cancer genome: Aberrations in regulatory elements, untranslated regions, splice sites, non-coding RNA and synonymous mutations. EMBO Mol. Med. 2016, 8, 442–457. [Google Scholar] [CrossRef]
  99. Petasny, M.; Bentata, M.; Pawellek, A.; Baker, M.; Kay, G.; Salton, M. Splicing to Keep Cycling: The Importance of Pre-mRNA Splicing during the Cell Cycle. Trends Genet. 2021, 37, 266–278. [Google Scholar] [CrossRef] [PubMed]
  100. Murthy, T.; Bluemn, T.; Gupta, A.K.; Reimer, M., Jr.; Rao, S.; Pillai, M.M.; Minella, A.C. Cyclin-dependent kinase 1 (CDK1) and CDK2 have opposing roles in regulating interactions of splicing factor 3B1 with chromatin. J. Biol. Chem. 2018, 293, 10220–10234. [Google Scholar] [CrossRef] [Green Version]
  101. Shiraishi, Y.; Kataoka, K.; Chiba, K.; Okada, A.; Kogure, Y.; Tanaka, H.; Ogawa, S.; Miyano, S. A comprehensive characterization of cis-acting splicing-associated variants in human cancer. Genome Res. 2018, 28, 1111–1125. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Lai, M.Y.; Chang, H.C.; Li, H.P.; Ku, C.K.; Chen, P.J.; Sheu, J.C.; Huang, G.T.; Lee, P.H.; Chen, D.S. Splicing mutations of the p53 gene in human hepatocellular carcinoma. Cancer Res. 1993, 53, 1653–1656. [Google Scholar]
  103. Braakhuis, B.J.; Rietbergen, M.M.; Buijze, M.; Snijders, P.J.; Bloemena, E.; Brakenhoff, R.H.; Leemans, C.R. TP53 mutation and human papilloma virus status of oral squamous cell carcinomas in young adult patients. Oral Dis. 2014, 20, 602–608. [Google Scholar] [CrossRef]
  104. Supek, F.; Minana, B.; Valcarcel, J.; Gabaldon, T.; Lehner, B. Synonymous mutations frequently act as driver mutations in human cancers. Cell 2014, 156, 1324–1335. [Google Scholar] [CrossRef] [Green Version]
  105. Marcel, V.; Dichtel-Danjoy, M.L.; Sagne, C.; Hafsi, H.; Ma, D.; Ortiz-Cuaran, S.; Olivier, M.; Hall, J.; Mollereau, B.; Hainaut, P.; et al. Biological functions of p53 isoforms through evolution: Lessons from animal and cellular models. Cell Death Differ. 2011, 18, 1815–1824. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Ota, A.; Nakao, H.; Sawada, Y.; Karnan, S.; Wahiduzzaman, M.; Inoue, T.; Kobayashi, Y.; Yamamoto, T.; Ishii, N.; Ohashi, T.; et al. Delta40p53alpha suppresses tumor cell proliferation and induces cellular senescence in hepatocellular carcinoma cells. J. Cell Sci. 2017, 130, 614–625. [Google Scholar] [CrossRef] [Green Version]
  107. Stiewe, T.; Tuve, S.; Peter, M.; Tannapfel, A.; Elmaagacli, A.H.; Putzer, B.M. Quantitative TP73 transcript analysis in hepatocellular carcinomas. Clin. Cancer Res. 2004, 10, 626–633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  108. Stiewe, T.; Zimmermann, S.; Frilling, A.; Esche, H.; Putzer, B.M. Transactivation-deficient DeltaTA-p73 acts as an oncogene. Cancer Res. 2002, 62, 3598–3602. [Google Scholar]
  109. Wang, X.Q.; Luk, J.M.; Leung, P.P.; Wong, B.W.; Stanbridge, E.J.; Fan, S.T. Alternative mRNA splicing of liver intestine-cadherin in hepatocellular carcinoma. Clin. Cancer Res. 2005, 11, 483–489. [Google Scholar]
  110. Wang, X.Q.; Luk, J.M.; Garcia-Barcelo, M.; Miao, X.; Leung, P.P.; Ho, D.W.; Cheung, S.T.; Lam, B.Y.; Cheung, C.K.; Wong, A.S.; et al. Liver intestine-cadherin (CDH17) haplotype is associated with increased risk of hepatocellular carcinoma. Clin. Cancer Res. 2006, 12, 5248–5252. [Google Scholar] [CrossRef] [Green Version]
  111. Narla, G.; Difeo, A.; Reeves, H.L.; Schaid, D.J.; Hirshfeld, J.; Hod, E.; Katz, A.; Isaacs, W.B.; Hebbring, S.; Komiya, A.; et al. A germline DNA polymorphism enhances alternative splicing of the KLF6 tumor suppressor gene and is associated with increased prostate cancer risk. Cancer Res. 2005, 65, 1213–1222. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Vetter, D.; Cohen-Naftaly, M.; Villanueva, A.; Lee, Y.A.; Kocabayoglu, P.; Hannivoort, R.; Narla, G.; Llovet, J.M.; Thung, S.N.; Friedman, S.L. Enhanced hepatocarcinogenesis in mouse models and human hepatocellular carcinoma by coordinate KLF6 depletion and increased messenger RNA splicing. Hepatology 2012, 56, 1361–1370. [Google Scholar] [CrossRef]
  113. Hanoun, N.; Bureau, C.; Diab, T.; Gayet, O.; Dusetti, N.; Selves, J.; Vinel, J.P.; Buscail, L.; Cordelier, P.; Torrisani, J. The SV2 variant of KLF6 is down-regulated in hepatocellular carcinoma and displays anti-proliferative and pro-apoptotic functions. J. Hepatol. 2010, 53, 880–888. [Google Scholar] [CrossRef] [PubMed]
  114. Acevedo, V.D.; Ittmann, M.; Spencer, D.M. Paths of FGFR-driven tumorigenesis. Cell Cycle 2009, 8, 580–588. [Google Scholar] [CrossRef] [PubMed]
  115. Venables, J.P.; Brosseau, J.P.; Gadea, G.; Klinck, R.; Prinos, P.; Beaulieu, J.F.; Lapointe, E.; Durand, M.; Thibault, P.; Tremblay, K.; et al. RBFOX2 is an important regulator of mesenchymal tissue-specific splicing in both normal and cancer tissues. Mol. Cell Biol. 2013, 33, 396–405. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Pradella, D.; Naro, C.; Sette, C.; Ghigna, C. EMT and stemness: Flexible processes tuned by alternative splicing in development and cancer progression. Mol. Cancer 2017, 16, 8. [Google Scholar] [CrossRef] [Green Version]
  117. Amann, T.; Bataille, F.; Spruss, T.; Dettmer, K.; Wild, P.; Liedtke, C.; Muhlbauer, M.; Kiefer, P.; Oefner, P.J.; Trautwein, C.; et al. Reduced expression of fibroblast growth factor receptor 2IIIb in hepatocellular carcinoma induces a more aggressive growth. Am. J. Pathol. 2010, 176, 1433–1442. [Google Scholar] [CrossRef] [Green Version]
  118. Paur, J.; Nika, L.; Maier, C.; Moscu-Gregor, A.; Kostka, J.; Huber, D.; Mohr, T.; Heffeter, P.; Schrottmaier, W.C.; Kappel, S.; et al. Fibroblast growth factor receptor 3 isoforms: Novel therapeutic targets for hepatocellular carcinoma? Hepatology 2015, 62, 1767–1778. [Google Scholar] [CrossRef] [Green Version]
  119. Li, K.; Shen, B.; Cheng, X.; Ma, D.; Jing, X.; Liu, X.; Yang, W.; Peng, C.; Qiu, W. Phenotypic and Signaling Consequences of a Novel Aberrantly Spliced Transcript FGF Receptor-3 in Hepatocellular Carcinoma. Cancer Res. 2016, 76, 4205–4215. [Google Scholar] [CrossRef] [Green Version]
  120. Jin, Z.; Feng, H.; Liang, J.; Jing, X.; Zhao, Q.; Zhan, L.; Shen, B.; Cheng, X.; Su, L.; Qiu, W. FGFR3 big up tri, open7-9 promotes tumor progression via the phosphorylation and destabilization of ten-eleven translocation-2 in human hepatocellular carcinoma. Cell Death Dis. 2020, 11, 903. [Google Scholar] [CrossRef]
  121. Saito, Y.; Kanai, Y.; Sakamoto, M.; Saito, H.; Ishii, H.; Hirohashi, S. Overexpression of a splice variant of DNA methyltransferase 3b, DNMT3b4, associated with DNA hypomethylation on pericentromeric satellite regions during human hepatocarcinogenesis. Proc. Natl. Acad. Sci. USA 2002, 99, 10060–10065. [Google Scholar] [CrossRef] [Green Version]
  122. Li, Y.; Lin, S.; Li, L.; Tang, Z.; Hu, Y.; Ban, X.; Zeng, T.; Zhou, Y.; Zhu, Y.; Gao, S.; et al. PDSS2 Deficiency Induces Hepatocarcinogenesis by Decreasing Mitochondrial Respiration and Reprogramming Glucose Metabolism. Cancer Res. 2018, 78, 4471–4481. [Google Scholar] [CrossRef] [Green Version]
  123. Zeng, T.; Tang, Z.; Liang, L.; Suo, D.; Li, L.; Li, J.; Yuan, Y.; Guan, X.Y.; Li, Y. PDSS2-Del2, a new variant of PDSS2, promotes tumor cell metastasis and angiogenesis in hepatocellular carcinoma via activating NF-kappaB. Mol. Oncol. 2020, 14, 3184–3197. [Google Scholar] [CrossRef] [PubMed]
  124. Shen, L.; Lei, S.; Zhang, B.; Li, S.; Huang, L.; Czachor, A.; Breitzig, M.; Gao, Y.; Huang, M.; Mo, X.; et al. Skipping of exon 10 in Axl pre-mRNA regulated by PTBP1 mediates invasion and metastasis process of liver cancer cells. Theranostics 2020, 10, 5719–5735. [Google Scholar] [CrossRef] [PubMed]
  125. Luo, Z.L.; Cheng, S.Q.; Shi, J.; Zhang, H.L.; Zhang, C.Z.; Chen, H.Y.; Qiu, B.J.; Tang, L.; Hu, C.L.; Wang, H.Y.; et al. A splicing variant of Merlin promotes metastasis in hepatocellular carcinoma. Nat. Commun. 2015, 6, 8457. [Google Scholar] [CrossRef]
  126. Wu, S.; Du, R.; Gao, C.; Kang, J.; Wen, J.; Sun, T. The role of XBP1s in the metastasis and prognosis of hepatocellular carcinoma. Biochem. Biophys. Res. Commun. 2018, 500, 530–537. [Google Scholar] [CrossRef]
  127. Chae, S.; Jun, H.O.; Lee, E.G.; Yang, S.J.; Lee, D.C.; Jung, J.K.; Park, K.C.; Yeom, Y.I.; Kim, K.W. Osteopontin splice variants differentially modulate the migratory activity of hepatocellular carcinoma cell lines. Int. J. Oncol. 2009, 35, 1409–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Nam, K.; Oh, S.; Lee, K.M.; Yoo, S.A.; Shin, I. CD44 regulates cell proliferation, migration, and invasion via modulation of c-Src transcription in human breast cancer cells. Cell Signal 2015, 27, 1882–1894. [Google Scholar] [CrossRef]
  129. Ponta, H.; Sherman, L.; Herrlich, P.A. CD44: From adhesion molecules to signalling regulators. Nat. Rev. Mol. Cell Biol. 2003, 4, 33–45. [Google Scholar] [CrossRef]
  130. Brown, R.L.; Reinke, L.M.; Damerow, M.S.; Perez, D.; Chodosh, L.A.; Yang, J.; Cheng, C. CD44 splice isoform switching in human and mouse epithelium is essential for epithelial-mesenchymal transition and breast cancer progression. J. Clin. Investig. 2011, 121, 1064–1074. [Google Scholar] [CrossRef] [Green Version]
  131. Li, J.; You, S.; Zhang, S.; Hu, Q.; Wang, F.; Chi, X.; Zhao, W.; Xie, C.; Zhang, C.; Yu, Y.; et al. Elevated N-methyltransferase expression induced by hepatic stellate cells contributes to the metastasis of hepatocellular carcinoma via regulation of the CD44v3 isoform. Mol. Oncol. 2019, 13, 1993–2009. [Google Scholar] [CrossRef] [Green Version]
  132. Marzese, D.M.; Liu, M.; Huynh, J.L.; Hirose, H.; Donovan, N.C.; Huynh, K.T.; Kiyohara, E.; Chong, K.; Cheng, D.; Tanaka, R.; et al. Brain metastasis is predetermined in early stages of cutaneous melanoma by CD44v6 expression through epigenetic regulation of the spliceosome. Pigment Cell Melanoma Res. 2015, 28, 82–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Chen, L.; Yao, Y.; Sun, L.; Zhou, J.; Miao, M.; Luo, S.; Deng, G.; Li, J.; Wang, J.; Tang, J. Snail Driving Alternative Splicing of CD44 by ESRP1 Enhances Invasion and Migration in Epithelial Ovarian Cancer. Cell Physiol. Biochem. 2017, 43, 2489–2504. [Google Scholar] [CrossRef] [PubMed]
  134. Zhang, F.L.; Cao, J.L.; Xie, H.Y.; Sun, R.; Yang, L.F.; Shao, Z.M.; Li, D.Q. Cancer-Associated MORC2-Mutant M276I Regulates an hnRNPM-Mediated CD44 Splicing Switch to Promote Invasion and Metastasis in Triple-Negative Breast Cancer. Cancer Res. 2018, 78, 5780–5792. [Google Scholar] [CrossRef] [Green Version]
  135. Jimenez, M.; Arechederra, M.; Avila, M.A.; Berasain, C. Splicing alterations contributing to cancer hallmarks in the liver: Central role of dedifferentiation and genome instability. Transl. Gastroenterol. Hepatol. 2018, 3, 84. [Google Scholar] [CrossRef]
  136. Sze, K.M.; Ching, Y.P.; Jin, D.Y.; Ng, I.O. Role of a novel splice variant of mitotic arrest deficient 1 (MAD1), MAD1beta, in mitotic checkpoint control in liver cancer. Cancer Res. 2008, 68, 9194–9201. [Google Scholar] [CrossRef] [Green Version]
  137. Yasen, M.; Mizushima, H.; Mogushi, K.; Obulhasim, G.; Miyaguchi, K.; Inoue, K.; Nakahara, I.; Ohta, T.; Aihara, A.; Tanaka, S.; et al. Expression of Aurora B and alternative variant forms in hepatocellular carcinoma and adjacent tissue. Cancer Sci. 2009, 100, 472–480. [Google Scholar] [CrossRef] [PubMed]
  138. Ebert, L.M.; Tan, B.S.; Browning, J.; Svobodova, S.; Russell, S.E.; Kirkpatrick, N.; Gedye, C.; Moss, D.; Ng, S.P.; MacGregor, D.; et al. The regulatory T cell-associated transcription factor FoxP3 is expressed by tumor cells. Cancer Res. 2008, 68, 3001–3009. [Google Scholar] [CrossRef] [Green Version]
  139. Harada, K.; Shimoda, S.; Kimura, Y.; Sato, Y.; Ikeda, H.; Igarashi, S.; Ren, X.S.; Sato, H.; Nakanuma, Y. Significance of immunoglobulin G4 (IgG4)-positive cells in extrahepatic cholangiocarcinoma: Molecular mechanism of IgG4 reaction in cancer tissue. Hepatology 2012, 56, 157–164. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Kosriwong, K.; Menheniott, T.R.; Giraud, A.S.; Jearanaikoon, P.; Sripa, B.; Limpaiboon, T. Trefoil factors: Tumor progression markers and mitogens via EGFR/MAPK activation in cholangiocarcinoma. World J. Gastroenterol. 2011, 17, 1631–1641. [Google Scholar] [CrossRef]
  141. Kamlua, S.; Patrakitkomjorn, S.; Jearanaikoon, P.; Menheniott, T.R.; Giraud, A.S.; Limpaiboon, T. A novel TFF2 splice variant (EX2TFF2) correlates with longer overall survival time in cholangiocarcinoma. Oncol. Rep. 2012, 27, 1207–1212. [Google Scholar] [CrossRef] [Green Version]
  142. Morrison, A.; Chekaluk, Y.; Bacares, R.; Ladanyi, M.; Zhang, L. BAP1 missense mutation c.2054 A>T (p.E685V) completely disrupts normal splicing through creation of a novel 5′ splice site in a human mesothelioma cell line. PLoS ONE 2015, 10, e0119224. [Google Scholar] [CrossRef] [PubMed]
  143. Ventii, K.H.; Devi, N.S.; Friedrich, K.L.; Chernova, T.A.; Tighiouart, M.; Van Meir, E.G.; Wilkinson, K.D. BRCA1-associated protein-1 is a tumor suppressor that requires deubiquitinating activity and nuclear localization. Cancer Res. 2008, 68, 6953–6962. [Google Scholar] [CrossRef] [Green Version]
  144. Brandi, G.; Deserti, M.; Palloni, A.; Turchetti, D.; Zuntini, R.; Pedica, F.; Frega, G.; De Lorenzo, S.; Abbati, F.; Rizzo, A.; et al. Intrahepatic cholangiocarcinoma development in a patient with a novel BAP1 germline mutation and low exposure to asbestos. Cancer Genet. 2020, 248–249, 57–62. [Google Scholar] [CrossRef]
  145. Obacz, J.; Takacova, M.; Brychtova, V.; Dobes, P.; Pastorekova, S.; Vojtesek, B.; Hrstka, R. The role of AGR2 and AGR3 in cancer: Similar but not identical. Eur. J. Cell Biol. 2015, 94, 139–147. [Google Scholar] [CrossRef] [PubMed]
  146. Yosudjai, J.; Inpad, C.; Pothipan, P.; Saisomboon, S.; Surangkul, D.; Phimsen, S.; Hongsrichan, N.; Wongkham, S.; Jirawatnotai, S.; Roytrakul, S.; et al. Overexpression of AGR2vH, an oncogenic AGR2 spliced transcript, potentiates tumorigenicity and proteomic alterations in cholangiocarcinoma cell. Biosci. Biotechnol. Biochem. 2021, 85, 2263–2273. [Google Scholar] [CrossRef]
  147. Suwanmanee, G.; Yosudjai, J.; Phimsen, S.; Wongkham, S.; Jirawatnotai, S.; Kaewkong, W. Upregulation of AGR2vH facilitates cholangiocarcinoma cell survival under endoplasmic reticulum stress via the activation of the unfolded protein response pathway. Int. J. Mol. Med. 2020, 45, 669–677. [Google Scholar] [CrossRef] [PubMed]
  148. Yosudjai, J.; Inpad, C.; Chomwong, S.; Dana, P.; Sawanyawisuth, K.; Phimsen, S.; Wongkham, S.; Jirawatnotai, S.; Kaewkong, W. An aberrantly spliced isoform of anterior gradient-2, AGR2vH promotes migration and invasion of cholangiocarcinoma cell. Biomed. Pharmacother. 2018, 107, 109–116. [Google Scholar] [CrossRef]
  149. Weiskirchen, R. CCN proteins in normal and injured liver. Front. Biosci. 2011, 16, 1939–1961. [Google Scholar] [CrossRef]
  150. Tanaka, S.; Sugimachi, K.; Kameyama, T.; Maehara, S.; Shirabe, K.; Shimada, M.; Wands, J.R.; Maehara, Y. Human WISP1v, a member of the CCN family, is associated with invasive cholangiocarcinoma. Hepatology 2003, 37, 1122–1129. [Google Scholar] [CrossRef]
  151. Qian, Z.; Hu, W.; Lv, Z.; Liu, H.; Chen, D.; Wang, Y.; Wu, J.; Zheng, S. PKM2 upregulation promotes malignancy and indicates poor prognosis for intrahepatic cholangiocarcinoma. Clin. Res. Hepatol. Gastroenterol. 2020, 44, 162–173. [Google Scholar] [CrossRef]
  152. Yu, G.; Yu, W.; Jin, G.; Xu, D.; Chen, Y.; Xia, T.; Yu, A.; Fang, W.; Zhang, X.; Li, Z.; et al. PKM2 regulates neural invasion of and predicts poor prognosis for human hilar cholangiocarcinoma. Mol. Cancer 2015, 14, 193. [Google Scholar] [CrossRef] [Green Version]
  153. Kotelevets, L.; Foudi, N.; Louedec, L.; Couvelard, A.; Chastre, E.; Norel, X. A new mRNA splice variant coding for the human EP3-I receptor isoform. Prostaglandins Leukot. Essent. Fat. Acids 2007, 77, 195–201. [Google Scholar] [CrossRef]
  154. Du, M.; Shi, F.; Zhang, H.; Xia, S.; Zhang, M.; Ma, J.; Bai, X.; Zhang, L.; Wang, Y.; Cheng, S.; et al. Prostaglandin E2 promotes human cholangiocarcinoma cell proliferation, migration and invasion through the upregulation of beta-catenin expression via EP3-4 receptor. Oncol. Rep. 2015, 34, 715–726. [Google Scholar] [CrossRef] [Green Version]
  155. Yun, K.J.; Yoon, K.H.; Han, W.C. Immunohistochemical Study for CD44v6 in Hepatocellular Carcinoma and Cholangiocarcinoma. Cancer Res. Treat. 2002, 34, 170–174. [Google Scholar] [CrossRef]
  156. Padthaisong, S.; Thanee, M.; Namwat, N.; Phetcharaburanin, J.; Klanrit, P.; Khuntikeo, N.; Titapun, A.; Sungkhamanon, S.; Saya, H.; Loilome, W. Overexpression of a panel of cancer stem cell markers enhances the predictive capability of the progression and recurrence in the early stage cholangiocarcinoma. J. Transl. Med. 2020, 18, 64. [Google Scholar] [CrossRef] [Green Version]
  157. Reviejo, M.; Soto, M.; Lozano, E.; Asensio, M.; Martínez-Augustin, O.; de Medina, F.S.; Marin, J.J.G. Impact of alternative splicing on mechanisms of resistance to anticancer drugs. Biochem. Pharmacol. 2021, 193, 114810. [Google Scholar] [CrossRef]
  158. Marin, J.J.G.; Macias, R.I.R.; Monte, M.J.; Romero, M.R.; Asensio, M.; Sanchez-Martin, A.; Cives-Losada, C.; Temprano, A.G.; Espinosa-Escudero, R.; Reviejo, M.; et al. Molecular Bases of Drug Resistance in Hepatocellular Carcinoma. Cancers 2020, 12, 1663. [Google Scholar] [CrossRef]
  159. Marin, J.J.; Macias, R.I. Understanding drug resistance mechanisms in cholangiocarcinoma: Assisting the clinical development of investigational drugs. Expert. Opin. Investig. Drugs 2021, 30, 675–679. [Google Scholar] [CrossRef]
  160. Lozano, E.; Herraez, E.; Briz, O.; Robledo, V.S.; Hernandez-Iglesias, J.; Gonzalez-Hernandez, A.; Marin, J.J. Role of the plasma membrane transporter of organic cations OCT1 and its genetic variants in modern liver pharmacology. Biomed. Res. Int. 2013, 2013, 692071. [Google Scholar] [CrossRef] [Green Version]
  161. Herraez, E.; Lozano, E.; Macias, R.I.; Vaquero, J.; Bujanda, L.; Banales, J.M.; Marin, J.J.; Briz, O. Expression of SLC22A1 variants may affect the response of hepatocellular carcinoma and cholangiocarcinoma to sorafenib. Hepatology 2013, 58, 1065–1073. [Google Scholar] [CrossRef] [PubMed]
  162. Dauki, A.M.; Blachly, J.S.; Kautto, E.A.; Ezzat, S.; Abdel-Rahman, M.H.; Coss, C.C. Transcriptionally Active Androgen Receptor Splice Variants Promote Hepatocellular Carcinoma Progression. Cancer Res. 2020, 80, 561–575. [Google Scholar] [CrossRef]
  163. Bao, M.H.; Wong, C.C. Hypoxia, Metabolic Reprogramming, and Drug Resistance in Liver Cancer. Cells 2021, 10, 1715. [Google Scholar] [CrossRef] [PubMed]
  164. Li, X.; Ramadori, P.; Pfister, D.; Seehawer, M.; Zender, L.; Heikenwalder, M. The immunological and metabolic landscape in primary and metastatic liver cancer. Nat. Rev. Cancer 2021, 21, 541–557. [Google Scholar] [CrossRef]
  165. Sperandio, R.C.; Pestana, R.C.; Miyamura, B.V.; Kaseb, A.O. Hepatocellular Carcinoma Immunotherapy. Annu. Rev. Med. 2022, 73. [Google Scholar] [CrossRef]
  166. Charalampakis, N.; Papageorgiou, G.; Tsakatikas, S.; Fioretzaki, R.; Kole, C.; Kykalos, S.; Tolia, M.; Schizas, D. Immunotherapy for cholangiocarcinoma: A 2021 update. Immunotherapy 2021, 13, 1113–1134. [Google Scholar] [CrossRef]
  167. Luo, D.; Zhao, D.; Zhang, M.; Hu, C.; Li, H.; Zhang, S.; Chen, X.; Huttad, L.; Li, B.; Jin, C.; et al. Alternative Splicing-Based Differences between Hepatocellular Carcinoma and Intrahepatic Cholangiocarcinoma: Genes, Immune Microenvironment, and Survival Prognosis. Front. Oncol. 2021, 11, 731993. [Google Scholar] [CrossRef]
  168. Salton, M.; Misteli, T. Small Molecule Modulators of Pre-mRNA Splicing in Cancer Therapy. Trends Mol. Med. 2016, 22, 28–37. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Bonnal, S.C.; Lopez-Oreja, I.; Valcarcel, J. Roles and mechanisms of alternative splicing in cancer—Implications for care. Nat. Rev. Clin. Oncol. 2020, 17, 457–474. [Google Scholar] [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Marin, J.J.G.; Reviejo, M.; Soto, M.; Lozano, E.; Asensio, M.; Ortiz-Rivero, S.; Berasain, C.; Avila, M.A.; Herraez, E. Impact of Alternative Splicing Variants on Liver Cancer Biology. Cancers 2022, 14, 18. https://doi.org/10.3390/cancers14010018

AMA Style

Marin JJG, Reviejo M, Soto M, Lozano E, Asensio M, Ortiz-Rivero S, Berasain C, Avila MA, Herraez E. Impact of Alternative Splicing Variants on Liver Cancer Biology. Cancers. 2022; 14(1):18. https://doi.org/10.3390/cancers14010018

Chicago/Turabian Style

Marin, Jose J. G., Maria Reviejo, Meraris Soto, Elisa Lozano, Maitane Asensio, Sara Ortiz-Rivero, Carmen Berasain, Matias A. Avila, and Elisa Herraez. 2022. "Impact of Alternative Splicing Variants on Liver Cancer Biology" Cancers 14, no. 1: 18. https://doi.org/10.3390/cancers14010018

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

Article Metrics

Back to TopTop