Abstract
RNA splicing alterations are widespread and play critical roles in cancer pathogenesis and therapy. Lung cancer is highly heterogeneous and causes the most cancer-related deaths worldwide. Large-scale multi-omics studies have not only characterized the mutational landscapes but also discovered a plethora of transcriptional and post-transcriptional changes in lung cancer. Such resources have greatly facilitated the development of new diagnostic markers and therapeutic options over the past two decades. Intriguingly, altered RNA splicing has emerged as an important molecular feature and therapeutic target of lung cancer. In this review, we provide a brief overview of splicing dysregulation in lung cancer and summarize the recent progress on key splicing events and splicing factors that contribute to lung cancer pathogenesis. Moreover, we describe the general strategies targeting splicing alterations in lung cancer and highlight the potential of combining splicing modulation with currently approved therapies to combat this deadly disease. This review provides new mechanistic and therapeutic insights into splicing dysregulation in cancer.
Keywords: RNA splicing, Splicing factors, Splicing alterations, Lung cancer
Graphical abstract

Highlights
-
•
Splicing dysregulation is an emerging molecular feature of lung cancer.
-
•
Splicing alterations are critical for lung cancer pathogenesis.
-
•
Targeting dysregulated splicing holds great potential for lung cancer treatment.
Introduction
Lung cancer is the most prevalent cancer type and the leading cause of cancer-related death worldwide, with an estimated 2 million new cases and 1.8 million deaths every year.1 Histologically, it is classified as non-small cell lung cancer (NSCLC) and small cell lung cancer (SCLC), and NSCLC mainly comprises lung adenocarcinoma (LUAD) and lung squamous cell carcinoma (LUSC).2 With continuous advances in detection and treatment approaches, clinical outcomes have been markedly improved for patients with lung cancer. However, lung cancer management is still very challenging because of its tremendous complexity and heterogeneity in terms of clinicopathological and molecular features.3, 4, 5, 6 Over the past decade, genomic and functional studies have identified oncogenic molecular changes that greatly facilitate the development of targeted therapies for lung cancer. Intriguingly, the dysregulation of ribonucleic acid (RNA) splicing has been proven to be widespread and plays a critical role in NSCLC pathogenesis and treatment.7, 8, 9, 10
RNA splicing is the process of excising introns and ligating exons in precursor messenger RNA (pre-mRNA) to generate mature mRNA in eukaryotic cells. The spliceosome, a huge RNA-protein complex, catalyzes the two-step transesterification chemical reactions involved in RNA splicing. RNA splicing is highly dynamic and extensively regulated,11,12 and this has been described well in previous reviews.9,13, 14, 15 The interactions between cis-regulatory elements in pre-mRNA and various trans-acting factors (i.e., splicing factors) largely determine splicing outcomes through the modulation of basal spliceosome activity [Figure 1A]. Alternative splicing (AS), the selection and usage of different splice sites, frequently occurs due to splicing regulation [Figure 1B]. Almost all multi-exon genes in humans are regulated by AS, greatly expanding the complexity and diversity of the transcriptome and proteome.16 According to the pattern of splice-site usage, AS can be divided into five major basic types as follows: skipped exon, alternative 3′ or 5′ splice site, mutually exclusive exons, and intron retention [Figure 1B].
Figure 1.
Dysregulation of RNA splicing in cancer. (A) General mechanisms of alternative splicing regulation. (B) Categories of alternative splicing events. (C) Mechanisms by which splicing is dysregulated in cancer. Splicing alterations mainly arise from mutations in splicing regulatory cis-elements, trans-acting splicing factors, and snRNAs, as well as the altered expression of splicing factors in cancer. ESE/S: Exonic splicing enhancer/silencer; hnRNP: Heterogeneous nuclear ribonucleoprotein; ISE/S: Intronic splicing enhancer/silencer; snRNP: Small nuclear ribonucleoprotein; SR: Serine/arginine-rich protein.
Splicing dysregulation is a major cause of human diseases and has been proposed as an emerging molecular hallmark of cancer. Splicing alterations, arising from mutations in splicing cis-regulatory elements and mutations in or changes in the expression of splicing factors, have crucial functions in cancer development, progression, and therapy [Figure 1C; more details are provided in previous excellent reviews17, 18, 19, 20, 21]. To date, the role of aberrant splicing has been best characterized in hematopoietic malignancies, but this has also been increasingly investigated in solid tumors. Exciting progress has been made in understanding splicing alterations in lung cancer, making it one of the most representative examples of the role of this dysregulated process in solid tumors. Nevertheless, an updated summary of this research direction is currently lacking. In this review, we highlight key oncogenic splicing events and frequently mutated splicing factors in lung cancer, summarize the strategies used to target such splicing alterations, and discuss several potential challenges in the mechanistic understanding and clinical translation of splicing dysregulation in lung cancer.
Oncogenic splicing events in lung cancer
Based on the systematic identification and functional characterization, it is currently well-recognized that splicing alterations exert oncogenic effects and can serve as potential prognostic markers and therapeutic targets in lung cancer.22,23 As a representative example of global identification, a recent study systematically identified AS changes and investigated their biological implications via multi-omics analyses in NSCLC. Splicing changes in cancer-related genes, such as epidermal growth factor receptor (EGFR), fibroblast growth factor receptor 2 (FGFR2), and cluster of differentiation 44 (CD44), were found to be associated with prognosis in both LUAD and LUSC.22 In addition to global identification, a rapidly increasing number of aberrant AS events has been characterized in lung cancer. Several key examples are emphasized as follows.
Mesenchymal–epithelial transition (MET) exon 14 skipping
MET gene encodes a receptor tyrosine kinase whose over-activation promotes lung cancer by activating downstream oncogenic signaling pathways, encompassing mitogen-activated protein kinase (MAPK), phosphatidylinositol-4,5-bisphosphate 3-kinase (PI3K)–AKT serine/threonine kinase (AKT), and signal transducer and activator of transcription (STAT).24 Genomic alterations in MET, including amplification and mutations, lead to constitutively active MET signaling.25, 26, 27, 28, 29, 30 Notably, MET exon 14 skipping (MET-ΔE14) has been identified as one of the paradigmatic aberrant splicing events, with an oncogenic function and clear clinical significance, in LUAD [Figure 2A].6,31,32
Figure 2.
Key oncogenic splicing events in lung cancer. (A) Exon 14 skipping in MET is caused by mutations/deletions that disrupt splice donor and acceptor sites, leading to an in-frame deletion of 47 amino acids in the juxtamembrane domain of MET (MET-ΔE14). This deletion inhibits MET degradation and internalization, resulting in increased downstream signaling. (B) The 4th exon of KRAS is alternatively spliced, generating two splicing isoforms designated as KRAS4A and KRAS4B. KRAS4A and KRAS4B share common GTP/GDP-binding domains but differ in hypervariable regions (red). (C) Exon 3 skipping in PD-1 produces a soluble isoform, PD-1ΔEx3, which lacks the transmembrane domain. PD-1ΔEx3 might enhance anti-tumor immunity by interfering with the PD-1/PD-L1 signaling axis. Ex: Exon; FL: Full-length; IgV: Immunoglobulin variable domain; IPT: Immunoglobulins-plexins-transcription factors; ITIM: Immunoreceptor tyrosine-based inhibitory motif; ITSM: Immunoreceptor tyrosine-based switch motif; LUAD: Lung adenocarcinoma; MHC: Major histocompatibility complex; PSI: Plexins–semaphorins-integrins; SEMA: Semaphorin domain; TCR: T cell receptor.
MET-ΔE14 occurs in approximately 4% of NSCLC patients, based on the genomics data from multiple large-scale cohorts (The Cancer Genome Atlas Program (TCGA), Memorial Sloan Kettering Cancer Center (MSKCC), and Singapore Oncology Data Portal (OncoSG)). The genomic mutations in MET causing this change are heterogeneous, encompassing deletions and substitutions, and these mostly reside in or across from the splice donor and acceptor sites of exon 14 [Figure 2A]. Exon 14 skipping results in an in-frame deletion of 47 amino acids in the juxtamembrane domain of MET, which inhibits its degradation and internalization [Figure 2A], and thus, this is regarded as a gain-of-function alteration.33 The juxtamembrane domain is the key negative regulatory region of MET. This domain contains a caspase-cleavage sequence (ESVD1002) and a tyrosine-binding site (Y1003) for the E3 ubiquitin ligase Casitas B lineage lymphoma (c-CBL), which mediates the ubiquitination and degradation of MET.34 Compared to the wild type, MET-ΔE14 markedly slows down ligand-induced ubiquitination but has no significant effect on the phosphorylation of MET.35 In addition, exon 14 skipping causes a much more prominent association between MET and the p85 subunit of PI3K, which enhances the activation of downstream oncogenic signaling.36 Accordingly, mouse NIH3T3 fibroblasts were demonstrated to be tumorigenic in vivo when expressing MET-ΔE14.35
Multi-institutional studies have identified MET-ΔE14 as an independent oncogenic driver and a biomarker that is significantly associated with poorer survival in NSCLC.35 Clinical evidence has proven that patients harboring MET-ΔE14 could benefit from MET tyrosine kinase inhibitors (TKIs), including crizotinib, tepotinib, and capmatinib.37, 38, 39 Prior to the approval of new-generation MET inhibitors, crizotinib was recommended for NSCLC patients with MET-ΔE14, according to the National Comprehensive Cancer Network (NCCN) guidelines. The objective response rate (ORR) of crizotinib treatment in patients with MET-ΔE14 was determined to be 32.3%, based on the PROFILE 1001 clinical trial. However, the poor blood–brain barrier permeability of crizotinib has resulted in limited therapeutic efficacy in lung cancer patients with brain metastases.40 Recently, capmatinib, a highly potent, selective type 1b inhibitor of MET, was approved in the United States for the treatment of patients with advanced NSCLC.41, 42, 43 Moreover, multiple pre-clinical studies have provided strong evidence that it is more potent than previous MET TKIs (crizotinib and tepotinib) for the treatment of NSCLC with MET-ΔE14.41,42,44 In addition to the pre-clinical data, a phase I clinical trial supported the safety of the clinical application of capmatinib.45 Furthermore, a multi-institutional phase II clinical trial showed that the ORR of capmatinib was 68% and 41% in treatment-naive and pre-treated advanced NSCLC patients with MET-ΔE14, respectively.45, 46, 47
Kirsten rat sarcoma viral oncogene (KRAS) 4A and 4B
KRAS gene is the most frequently mutated oncogene in cancer, including LUAD.48,49 In Caucasian populations, KRAS mutations are found in approximately 30% of LUAD patients.6 Currently, novel compounds targeting the KRAS Gly12Cys mutation, such as sotorasib and adagrasib, have been developed and approved for clinical applications.50, 51, 52 Unfortunately, both are only effective against the KRAS G12C mutation, with limited effects on other KRAS driver mutations, such as G12D and G12V. Moreover, resistance to these inhibitors often occurs rapidly.51 Therefore, therapy targeting KRAS mutations remains a challenge in lung cancer.50,53,54
The 4th exon of KRAS is alternatively spliced, generating two splice variants designated as KRAS4A and KRAS4B [Figure 2B].55,56 Since the oncogenic mutation in KRAS is predominantly located in the 2nd and 3rd exons and leads to constitutively active oncoproteins, both isoforms were demonstrated to promote tumorigenesis in LUAD57 [Figure 2B]. Although KRAS4A was identified from the Kirsten rat sarcoma virus by Shimizu, early in 1983,48 over the past decades, the vast majority of studies have focused on KRAS4B largely owing to its higher expression (than KRAS4A) in lung cancer. KRAS4A and KRAS4B share the first 165 amino acids encoding G domains but differ substantially in their hypervariable regions that mediate membrane association and subcellular trafficking [Figure 2B]. The distinct landscape of interactomes for KRAS4A and KRAS4B was identified via affinity-purification mass spectrometry.58 For example, the v-ATPase A2 was shown to specifically interact with KRAS4B, but not KRAS4A, whereas the RAF-1 proto-oncogene serine/threonine kinase (RAF1) preferentially interacts with KRAS4A.58 In addition, KRAS4A was shown to profoundly enhance glycolysis by directly associating with hexokinase 1 on the outer mitochondrial membrane in cancer cells.59
KRAS4A is widely expressed in various types of cancer, especially in lung cancer and colon cancer.60,61 Recently, a multi-institutional study examined the expression of KRAS4A and KRAS4B in advanced-stage NSCLC patients and found that KRAS4A expression was elevated in most patients.56 Another study, based on the genomic and transcriptomic data of TCGA LUAD cohort, revealed that KRAS4A expression is positively correlated with genomic alterations in KRAS and significantly worse survival in LUAD.61 Further, an in vivo study consistently showed that KRAS4A alone could induce metastasis in LUAD in the absence of KRAS4B.62 Together, these findings indicate that KRAS4A plays critical roles, likely distinct from those of KRAS4B, in the development and progression of LUAD. Intriguingly, targeting KRAS4A splicing through degradation of the RNA-binding protein RBM39 was shown to inhibit cell stemness in lung cancer,63 providing a potential strategy to modulate KRAS splicing in cancer therapy. However, the regulatory mechanisms of KRAS splicing in cancer remain poorly understood and require further investigation.
Splicing alterations of programmed cell death protein 1/programmed cell death ligand 1 (PD-1/PD-L1)
Targeting the immune checkpoint molecules PD-1/PD-L1, known as immune checkpoint blockade (ICB), has resulted in remarkable clinical responses in various types of cancer, such as melanoma, NSCLC, and colon cancer.64, 65, 66 However, only a fraction of patients respond to ICB, calling for a deeper understanding of immune-escape mechanisms.67,68 Emerging evidence demonstrates that the AS of specific immune checkpoint molecules has significant effects on ICB. For example, exon 3 skipping in the PD-1-encoding gene PDCD1 produces a soluble form of the protein (designated as PD-1ΔEx3), which was shown to suppress the PD-1/PD-L1 signaling axis, thereby enhancing anti-tumor immunity [Figure 2C].69 Moreover, a clinical study in Denmark revealed that the upregulation of PD-1ΔEx3 expression correlates with improved survival in patients with EGFR-mutant NSCLC treated with TKIs.70 In addition, the expression of PD-1ΔEx3 was reported to enhance the response rate to immunotherapies, such as anti-PD-1 (a-PD-1) and anti-CTLA4 therapy, in NSCLC.71 Antisense oligonucleotides (ASOs) can shift PDCD1 splicing toward PD-1ΔEx3, providing an alternative approach when targeting PD-1 in lung cancer. Recently, a novel PD-L1 splice variant lacking the transmembrane domain has been identified.72 Compared with the canonical isoform expressed on the surfaces of cancer cells, this PD-L1 isoform appears to be secreted into the tumor immune microenvironment, conferring resistance to anti-PD-L1 immunotherapy in NSCLC.72,73
Dysregulation of splicing factors in lung cancer
Splicing factors are frequently mutated in hematologic malignancies, as well as in solid tumors.18,74 In addition to mutations, the abnormal expression or activity of splicing factors is also commonly observed in cancer.19,75 Accordingly, an increasing number of studies have demonstrated the oncogenic and tumor-suppressive functions of various splicing factors. Mechanistically, the dysregulation of splicing factors generally causes splicing alterations in target genes, consequently affecting cancer development, progression, and drug resistance. The modulation of dysregulated splicing factors and/or their target genes has started to show great potential for cancer therapy. However, despite this exciting progress, a large proportion of splicing factors dysregulated in cancer remain to be investigated. The spectrum of splicing factors exhibiting recurring mutations in lung cancer is shown in Figure 3A, and their functional roles and clinical implications are highlighted in the following sections.
Figure 3.
Splicing factors frequently mutated in lung cancer. (A) Spectrum of splicing factors frequently mutated in LUAD. Data are from the cBioPortal TCGA pan-cancer LUAD cohort. (B) Distribution of RBM10 mutations in LUAD along the RBM10 protein sequence. These mutations are primarily frameshift, nonsense, and splice-site mutations, leading to RBM10 loss-of-function alterations. Data are from the cBioPortal TCGA pan-cancer LUAD cohort. (C) RBM10 deficiency promotes lung cancer development, progression, and TKI resistance by regulating the alternative splicing of key target genes, such as NUMB, EIF4H, and BCLX. LUAD: Lung adenocarcinoma; RRM: RNA recognition motif; TKI: Tyrosine kinase inhibitor; ZnF: Zinc finger.
RNA binding motif protein 10 (RBM10)
RBM10 encodes an RNA-binding protein that has been identified as a component of U2 snRNP.76,77 We and other researchers revealed that RBM10 enhances exon skipping by binding to flanking intronic regions near splice sites78,79 or to exonic regions.80 RBM10 exhibits high mutation rates in multiple cancers, such as LUAD, colorectal carcinoma, pancreatic ductal adenocarcinoma, and bladder cancer.81, 82, 83 Moreover, RBM10 is the most frequently mutated splicing factor in lung cancers (e.g., 8.9% in a cohort of Chinese LUAD patients, 7.3% in the TCGA LUAD patient cohort,84 and even more frequently in early-stage or non-smoking LUAD patients85,86). RBM10 mutations are primarily loss-of-function [Figure 3B] and co-occur with known driver mutations, mostly EGFR and KRAS mutations, in lung cancer.84,87,88
Functional studies have demonstrated the tumor-suppressor functions of RBM10 in LUAD. Specifically, it was shown to repress Notch signaling via the AS-mediated regulation of NUMB exon 979 [Figure 3C]. We also found that RBM10 represses lung cancer progression by controlling the AS of eukaryotic translation initiation factor 4H (EIF4H) exon 5 [Figure 3C]. In particular, RBM10 loss was shown to enhance the inclusion of exon 5 in EIF4H. Importantly, expression of the long isoform of EIF4H containing exon 5 (EIF4H-L) is specifically upregulated in LUAD, is critical for LUAD cell proliferation and survival, and correlates with unfavorable prognosis, which makes it a promising therapeutic target.84 In addition, RBM10 was reported to suppress LUAD progression by inhibiting the Wnt/β-catenin and RAP1/AKT/CREB signaling pathways89,90 and inhibiting the invasion and metastasis of NSCLC cells by recruiting methyltransferase-like 3 (METTL3) to modulate the m6A methylation of its target long non-coding RNA metastasis-associated lung adenocarcinoma transcript 1 (MALAT1).91 Moreover, RBM10 deficiency in LUAD was demonstrated to confer high sensitivity to spliceosome inhibition,88 while compromising the efficacy of EGFR TKI therapy partially by regulating AS of the anti-apoptotic gene Bcl-x [Figure 3C].92 Notably, RBM10-deficient LUADs were linked to higher expression of human leukocyte antigen (HLA) and immune checkpoint molecules and increased immune cell infiltration compared to RBM10-wild-type LUADs.93,94 Additionally, RBM10 overexpression was found to significantly decrease the protein expression of PD-L1, whereas RBM10 silencing was determined to increase it.95 Although most studies support the tumor-suppressive functions of RBM10 in lung cancer,96 controversial oncogenic activities of RBM10 have been reported.97,98 Further in-depth studies, particularly those using in vivo mouse models combined with clinical samples, are needed to corroborate the functions and therapeutic value of RBM10 in lung cancer.
In addition to that in lung cancer, it was reported that RBM10 physically interacts with p53 in colon cancer and that its overexpression disrupts the mouse double minute 2 homolog (MDM2)-p53 interaction, subsequently repressing p53 ubiquitination.99 Further, RBM10 loss enhances sensitivity to BCL2 inhibitors partially through the mis-splicing of X-linked inhibitor of apoptosis (XIAP) in acute myeloid leukemia.100 Given the mutations and altered expression of RBM10 in multiple cancers, it will be important to elucidate its roles in different cancer types.
U2 small nuclear RNA auxiliary factor 1 (U2AF1)
U2AF1, an essential protein component of the splicing machinery, forms a heterodimer with U2AF2. U2AF1 interacts with the AG dinucleotide of the 3′ splice site through its RNA recognition motif and interacts with serine- and arginine-rich proteins, such as serine/arginine-rich splicing factor 2 (SRSF2), through its arginine–serine-rich domain.101 U2AF1 is frequently mutated in myelodysplastic syndromes (MDSs) and chronic myelomonocytic leukemia, as well as in several solid tumors, including LUAD.74,102 The S34F mutation, located in the first zinc finger domain of U2AF1, is the most pervasive hotspot in lung cancer and predicts worse survival.103 In general, the S34F mutation influences splicing by affecting the U2AF1-binding preference to the 3′ splice site, and it has been characterized as a change-of-function alteration.104 Intriguingly, the U2AF1 S34F mutation was shown to perturb mRNA translation by directly binding the mRNA 5′ untranslated region in the cytoplasm to promote cancer progression, implying a non-canonical role of splicing factors in cancer.105 Specifically, the overexpression of U2AF1S34F was found to lead to the elevated translation of genes associated with the epithelial–mesenchymal transition in lung cancer.105
Splicing factor 3b subunit 1 (SF3B1)
SF3B1 is a core component of U2 snRNP that is essential for the recognition and selection of the branch-point sequence. SF3B1 mutations have been intensively investigated in hematologic malignancies and also explored in several solid tumors, including uveal melanoma and LUAD.106,107 The hotspot missense mutations of SF3B1 K700 occur within the C-terminal HEAT repeat domains, and these result in the usage of cryptic 3′ splice sites and aberrant AS.108 Recent TCGA data analysis suggests that splicing changes induced by SF3B1 mutations share a similar pattern with that caused by SURP and G-patch domain containing 1 (SUGP1) deficiency in lung cancer.109 Moreover, the SF3B1 K700E mutation or a SUGP1 mutation disrupts the interaction between SUGP1 and SF3B1, leading to common splicing changes.110
Far upstream element (FUSE) binding protein 1 (FUBP1)
FUBP1 is involved in the regulation of transcription, splicing, and mRNA stabilization by binding to a single strand of deoxyribonucleic acid (DNA) or RNA.111,112 FUBP1 expression was found to be upregulated and correlated with poor prognosis in several cancers, including hepatocellular carcinoma, glioma, gastric cancer, ovarian cancer, and nasopharyngeal carcinoma. As such, it was regarded as an oncogene.113, 114, 115, 116, 117 In support of this notion, FUBP1 was shown to be required for efficient splicing of the oncogene MDM2 in MCF7 breast cancer cells.118 Conversely, loss-of-function mutations in FUBP1 have been identified in neuroblastoma, indicating a tumor-suppressive role.119 Interestingly, the FUBP1 S11Lfs∗43 mutation frequently occurs in LUAD (data from cBioPortal), but its functional significance remains to be determined. In vitro experiments showed that FUBP1 knockdown inhibits the proliferation and migration of lung cancer cells, suggesting its oncogenic functions in lung cancer.120 Mechanistically, FUBP1 was shown to be recruited by a novel long non-coding RNA, lung cancer-associated transcript 3 (LCAT3), and then bind the FUSE sequence to activate MYC transcription and promote cell proliferation.120 Another recent study indicated that FUBP1 knockdown decreases the expression of PD-L1 and inhibits LUSC tumor growth in vivo.73 Further investigations are required to reconcile the complex functions and underlying mechanisms of FUBP1 in lung cancer and other cancers.
Abnormal expression of splicing factors
The best-known example of splicing factors with altered expression in cancer is the proto-oncogene SRSF1. The expression of this protein was found to be upregulated and promote tumorigenesis in several cancer types, including lung cancer.121, 122, 123 Another representative example is SRSF2, whose P95 mutations occur frequently in MDSs but not in lung cancer.124,125 Previous studies showed that the levels of SRSF2 and phospho-SRSF2 proteins are overexpressed in LUAD, LUSC, and neuroendocrine lung tumors.126,127 Moreover, SRSF2 was shown to interact with E2F transcription factor 1 (E2F1) and positively regulate transcription to control the expression of cell cycle genes in neuroendocrine lung cancer.127 A recent study further revealed that SRSF2 expression is transcriptionally upregulated by SRY-box transcription factor 2 (SOX2), leading to a splicing change in vascular endothelial growth factor receptor (VEGFR) in LUSC.128 In addition, SRSF6 expression was found to be upregulated and induce the transformation of epithelial cells in lung cancer.129 Various splicing factors have also been found to be deregulated in lung cancer,19,75 but they still need to be functionally characterized.
Strategies to target splicing alterations in lung cancer
Owing to the fundamental roles of splicing alterations in cancer and technological advancements in the manipulation of splicing, targeting aberrant splicing has been considered an attractive cancer therapeutic approach. Currently, general strategies to modulate splicing mainly encompass targeting the spliceosome, splicing factors, and splicing isoforms using small-molecule inhibitors and ASOs, among others [Figure 4A and B]. Accumulating evidence has demonstrated the potent effects of targeting splicing in various cancers.7,9,10,17,130,131 Recently, pre-clinical and clinical studies have provided promising results regarding this strategy in lung cancer. Additionally, targeting splicing in combination with current standard treatment options for lung cancer can produce exciting results.
Figure 4.
Strategies to target RNA splicing alterations in lung cancer. (A) Approaches targeting spliceosomes and splicing factors mainly include small-molecule inhibitors and the PROTAC system. (B) Approaches to target aberrant splicing events in lung cancer, mainly including splice-switching ASOs and small molecules. (C) Schematic of targeting splicing in combination with standard treatments for lung cancer. ASO: Antisense oligonucleotide; ICB: Immune checkpoint blockade; PROTAC: Proteolysis-targeting chimeric molecule; TKI: Tyrosine kinase inhibitor.
Targeting the spliceosome and splicing factors
Various small-molecule inhibitors have been designed to inhibit the spliceosome and splicing factors, which have been tested in cancer. The classical small-molecule inhibitors of the spliceosome, namely the natural compound pladienolide B and its derivatives, E7107 and H3B-8800, were designed to target the SF3B1 complex.132, 133, 134 These small-molecule inhibitors bind to the branch-point-binding pocket of the SF3B complex, thereby preventing splicing [Figure 4A]. They also exhibited potent anti-tumor effects on lung cancer in pre-clinical studies;135,136 however, clinical evidence supporting their effectiveness is lacking. Two phase I clinical trials reported ocular complications caused by E7107 with unclear mechanisms, which hindered its further clinical application for advanced solid tumors.137,138
In addition to spliceosome inhibitors, small-molecule inhibitors targeting splicing factors, such as protein arginine methyltransferase 5 (PRMT5) and RBM39 have been developed, and their anti-tumor effects on lung cancer were tested in vitro and in vivo [Figure 4A].63,139, 140, 141, 142 For example, indisulam, a sulfonamide agent, can bind and bridge the splicing factor RBM39 with the CUL4–DDB1–DDA1–DCAF15 E3 ubiquitin ligase complex, leading to the polyubiquitination and proteasomal degradation of RBM39 and the inhibition of tumorigenesis in lung cancer [Figure 4A].63,143 In addition, proteolysis-targeting chimeric molecules, heterobifunctional compounds that utilize the ubiquitin-proteasome system to achieve specific protein degradation, can be applied to degrade abnormal splicing factors and have great potential for cancer therapy.142
Targeting aberrant splicing events
ASOs are short, artificially synthesized single-stranded nucleic acids with different modifications, which can directly bind to the splicing regulatory element in precursor mRNA to regulate splicing or pair with target mRNA to induce its degradation or repress translation. Further, they comprise an effective means to directly interfere with splicing abnormalities.18,144,145 The efficacy of ASOs targeting splicing alterations has been demonstrated in clinical trials for various cancers but not yet for lung cancer.146 Nonetheless, pre-clinical studies showed the significant anti-tumor effects of several ASOs targeting specific AS events in cancer-associated genes, such as EIF4H, BCLX, and MAPK interacting serine/threonine kinase 2 (MNK2), in lung cancer [Figure 4B].92,147,148 In addition to the splicing switch, it is worth noting that ASO-based RNA therapy has broad prospects in lung cancer. AZD9150, an ASO targeting STAT3, was shown to directly decrease the expression of STAT3 and exert anti-tumor effects on lymphoma and lung cancer in a phase I clinical trial.149 Further, AZD4785, a high-affinity ASO targeting KRAS, was found to exert prominent anti-tumor effects on KRAS-mutant NSCLC patient-derived xenografts by inhibiting KRAS expression.150 These studies provide foundations for the application of ASOs targeting aberrant splicing events in lung cancer.
Small-molecule inhibitors have also been pursued to modulate specific splicing isoforms. The most representative example is the MET TKIs mentioned previously herein. Capmatinib was tested in multiple clinical trials and resulted in optimistic outcomes for NSCLC patients with MET exon 14 skipping [Figure 4B].33,46,47 Salazosulfapyridine is a small-molecule compound that directly inhibits the splicing of isoforms of CD44 and prolongs progression-free survival in lung cancer.151 Another study showed that vorinostat could effectively target the oncogenic BIM splicing isoforms resulting from a deletion polymorphism.152,153
Targeting splicing in combination with standard treatment
According to NCCN guidelines, the standard treatment for NSCLC mainly consists of chemotherapy, radiotherapy, targeted therapy, and immunotherapy, beyond surgical approaches. However, therapy resistance is a major challenge encountered with these treatment options. Aberrant RNA splicing has been linked to therapy resistance in lung cancer. For example, the AS of the gene encoding caspase 9 was found to cause chemotherapy resistance in NSCLC.154 Further, SGOL1-B, a splice variant of shugoshin-like 1 (SGOL1), induces aberrant mitosis and resistance to taxane in LUAD.155 In addition, dysregulation of the splicing factor small nuclear ribonucleoprotein polypeptides B and B1 (SNRPB) was found to lead to platinum-based chemotherapy resistance in NSCLC.156
Currently, targeted therapies based on driver mutations, particularly TKIs, have brought about great benefits for patients with NSCLC, but resistance is almost inevitable. The mechanisms of resistance to TKIs include genomic alterations and other molecular and cellular changes.157 As such, the aberrant splicing of cancer-associated genes, such as HER2, BIM, and ATG16, was found to contribute to TKI resistance.158, 159, 160 In addition, deficiency of the splicing factor RBM10 was recently reported to limit the response to osimertinib in EGFR-mutant LUAD partially due to a splicing alteration in Bcl-x.92 Importantly, the combination of a Bcl-x inhibitor with osimertinib was found to synergistically inhibit LUAD [Figure 4C].92 Moreover, a phase I clinical trial led to the approval of the combination of vorinostat and gefitinib in BIM-deletion polymorphism/EGFR mutation-double positive LUAD [Figure 4C].152,153 Hence, targeting aberrant splicing in combination with conventional treatment options could be a very promising strategy to improve therapy efficacy and overcome resistance in lung cancer [Figure 4C].
The development of immunotherapy has revolutionized the treatment of lung cancer. The a-PD1 antibodies, represented by nivolumab and pembrolizumab, are a standard treatment strategy for advanced-stage NSCLC, especially those lacking driver mutations for targeted therapy. However, the ORR of a-PD1 was found to be approximately 20% in NSCLC due to primary or acquired resistance. Aberrant splicing has also been shown to affect the tumor's immune microenvironment.161, 162, 163 On one hand, the aberrant splicing of genes encoding immune checkpoint molecules could interfere with their normal functions, which in turn confers resistance or sensitivity to ICBs.69,72 Accordingly, ASOs designed to target those aberrant splicing events should be able to enhance the effects of ICBs in lung cancer. On the other hand, AS can generate neoantigens that reprogram the tumor immune microenvironment, similar to the tumor mutation burden, which correlates with the ORR of a-PD1 immunotherapy.164,165 The modulation of exon skipping and intron retention was predicted to generate numerous aberrant peptides, four times more than mutation-derived neoantigens,166,167 highlighting their important roles in anti-tumor immune responses. Interestingly, the splicing factor inhibitors indisulam and MS-023, targeting RBM39 and PRMTs, respectively, were found to significantly enhance sensitivity to ICBs in a pre-clinical study [Figure 4C].166 These studies provide direct evidence for combining the targeted modulation of splicing with ICBs as a promising therapeutic option for lung cancer [Figure 4C].
Conclusion and perspective
In this review, we summarized recent progress on the key splicing events and splicing factors that are altered in lung cancer. We also described the general strategies used to target splicing alterations in lung cancer and proposed a combination of splicing modulation with currently existing therapeutics as a promising direction to improve treatment outcomes. This review highlights the critical roles of RNA splicing alterations in the pathogenesis and treatment of lung cancer, providing new insights into cancer-related splicing dysregulation.
Despite encouraging advancements, there are pressing challenges that need to be addressed. First, many splicing events and splicing factors that are altered in lung cancer have not been functionally elucidated. Since splicing factors often regulate many RNA splicing events, it is difficult to determine whether a few key splicing alterations or many changes in combination are responsible for splicing factor dysregulation in cancer. Hence, efficient functional screening methods are important for elucidating splicing aberrations in lung cancer. Currently, several high-throughput screening libraries based on clustered regularly interspaced short palindromic repeats (CRISPR)-associated (Cas) systems have been developed and used to systematically interrogate cancer-related splicing factors and events.168, 169, 170 Such screening strategies have been applied to identify functional splicing alterations in lung cancer.171, 172, 173, 174 Second, splicing changes can be used as invaluable diagnostic biomarkers and therapeutic targets for lung cancer patients, yet these have not been translated to the clinic. Therefore, it is urgent to implement rationally designed clinical trials to test the efficacy of various splicing-modulating drugs, including large-scale, multi-institutional trials combining splicing modulation with targeted or immune therapy. It is also critical to develop sensitive, specific, and low-cost technologies to detect splicing changes in clinical samples, improve the delivery efficiency of ASOs to tumor sites, and limit the potential toxicity of spliceosome inhibitors. Such efforts will provide a foundation for the clinical application of splicing modulators in lung cancer treatment. Third, in stark contrast to knowledge on NSCLC, few studies have focused on splicing alterations in SCLC, for which more investigations are needed.
Funding
This work was supported by the National Natural Science Foundation of China (Nos. 81871878, 31371299), the Shanghai Municipal Natural Science Fund (No. 20ZR1406500), and the Innovation Research Team of High-level Local Universities in Shanghai.
Author contribution
Yongbo Wang,Yueren Yan, Yunpeng Ren, and Yufang Bao Bao conceived the review and wrote the manuscript. Yufang Bao, and Yueren Yan created the figures. Yongbo Wang edited the manuscript.
Ethics statement
None.
Data availability statement
The datasets used in the current study are available from the corresponding author on reasonable request.
Conflict of interest
None.
Acknowledgments
The figures are original works created by YF Bao and YR Yan with BioRender.com.
Managing Editor: Peng Lyu
References
- 1.Siegel R.L., Miller K.D., Fuchs H.E., Jemal A. Cancer statistics, 2022. CA Cancer J Clin. 2022;72:7–33. doi: 10.3322/caac.21708. [DOI] [PubMed] [Google Scholar]
- 2.Thai A.A., Solomon B.J., Sequist L.V., Gainor J.F., Heist R.S. Lung cancer. Lancet. 2021;398:535–554. doi: 10.1016/S0140-6736(21)00312-3. [DOI] [PubMed] [Google Scholar]
- 3.Cancer Genome Atlas Research Network Comprehensive genomic characterization of squamous cell lung cancers. Nature. 2012;489:519–525. doi: 10.1038/nature11404. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.George J., Lim J.S., Jang S.J., et al. Comprehensive genomic profiles of small cell lung cancer. Nature. 2015;524:47–53. doi: 10.1038/nature14664. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Hu J., Wang Y., Zhang Y., et al. Comprehensive genomic profiling of small cell lung cancer in Chinese patients and the implications for therapeutic potential. Cancer Med. 2019;8:4338–4347. doi: 10.1002/cam4.2199. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Collisson E.A., Campbell J.D., Brooks A.N., et al. Comprehensive molecular profiling of lung adenocarcinoma. Nature. 2014;511:543–550. doi: 10.1038/nature13385. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Venables J.P. Aberrant and alternative splicing in cancer. Cancer Res. 2004;64:7647–7654. doi: 10.1158/0008-5472.CAN-04-1910. [DOI] [PubMed] [Google Scholar]
- 8.Wang B.D., Lee N.H. Aberrant RNA splicing in cancer and drug resistance. Cancers. 2018;10:458. doi: 10.3390/cancers10110458. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Zhang Y., Qian J., Gu C., Yang Y. Alternative splicing and cancer: a systematic review. Signal Transduct Target Ther. 2021;6:78. doi: 10.1038/s41392-021-00486-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Song X., Zeng Z., Wei H., Wang Z. Alternative splicing in cancers: from aberrant regulation to new therapeutics. Semin Cell Dev Biol. 2018;75:13–22. doi: 10.1016/j.semcdb.2017.09.018. [DOI] [PubMed] [Google Scholar]
- 11.Wang Y.B., Bao Y.F., Zhang S.R., Wang Z.F. Splicing dysregulation in cancer: from mechanistic understanding to a new class of therapeutic targets. Science China Life Sci. 2020;63:469–484. doi: 10.1007/s11427-019-1605-0. [DOI] [PubMed] [Google Scholar]
- 12.Wahl M.C., Will C.L., Luhrmann R. The spliceosome: design principles of a dynamic RNP machine. Cell. 2009;136:701–718. doi: 10.1016/j.cell.2009.02.009. [DOI] [PubMed] [Google Scholar]
- 13.Jaganathan K., Panagiotopoulou S.K., McRae J.F., et al. Predicting splicing from primary sequence with deep learning. Cell. 2019;176:535–548. doi: 10.1016/j.cell.2018.12.015. [DOI] [PubMed] [Google Scholar]
- 14.Bao S.Y., Moakley D.F., Zhang C.L. The splicing code goes deep. Cell. 2019;176:414–416. doi: 10.1016/j.cell.2019.01.013. [DOI] [PubMed] [Google Scholar]
- 15.Braunschweig U., Gueroussov S., Plocik A.M., Graveley B.R., Blencowe B.J. Dynamic integration of splicing within gene regulatory pathways. Cell. 2013;152:1252–1269. doi: 10.1016/j.cell.2013.02.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Pan Q., Shai O., Lee L.J., Frey B.J., Blencowe B.J. Deep surveying of alternative splicing complexity in the human transcriptome by high-throughput sequencing. Nat Genet. 2008;40:1413–1415. doi: 10.1038/ng.259. [DOI] [PubMed] [Google Scholar]
- 17.Wang E., Aifantis I. RNA splicing and cancer. Trends Cancer. 2020;6:631–644. doi: 10.1016/j.trecan.2020.04.011. [DOI] [PubMed] [Google Scholar]
- 18.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. doi: 10.1038/s41571-020-0350-x. [DOI] [PubMed] [Google Scholar]
- 19.Anczukow O., Krainer A.R. Splicing-factor alterations in cancers. RNA. 2016;22:1285–1301. doi: 10.1261/rna.057919.116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Lee S.C.W., Abdel-Wahab O. Therapeutic targeting of splicing in cancer. Nat Med. 2016;22:976–986. doi: 10.1038/nm.4165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Bradley R.K., Anczuków O. RNA splicing dysregulation and the hallmarks of cancer. Nat Rev Cancer. 2023;23:135–155. doi: 10.1038/s41568-022-00541-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Li Y., Sun N., Lu Z., et al. Prognostic alternative mRNA splicing signature in non-small cell lung cancer. Cancer Lett. 2017;393:40–51. doi: 10.1016/j.canlet.2017.02.016. [DOI] [PubMed] [Google Scholar]
- 23.Wu Q., Feng L., Wang Y., et al. Multi-omics analysis reveals RNA splicing alterations and their biological and clinical implications in lung adenocarcinoma. Signal Transduct Target Ther. 2022;7:270. doi: 10.1038/s41392-022-01098-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 24.Nakamura T., Sakai K., Nakamura T., Matsumoto K. Hepatocyte growth factor twenty years on: much more than a growth factor. J Gastroenterol Hepatol. 2011;26(Suppl 1):188–202. doi: 10.1111/j.1440-1746.2010.06549.x. [DOI] [PubMed] [Google Scholar]
- 25.Vuong H.G., Ho A.T.N., Altibi A.M.A., Nakazawa T., Katoh R., Kondo T. Clinicopathological implications of MET exon 14 mutations in non-small cell lung cancer – a systematic review and meta-analysis. Lung Cancer. 2018;123:76–82. doi: 10.1016/j.lungcan.2018.07.006. [DOI] [PubMed] [Google Scholar]
- 26.Schrock A.B., Frampton G.M., Suh J., et al. Characterization of 298 patients with lung cancer harboring MET exon 14 skipping alterations. J Thorac Oncol. 2016;11:1493–1502. doi: 10.1016/j.jtho.2016.06.004. [DOI] [PubMed] [Google Scholar]
- 27.Frampton G.M., Ali S.M., Rosenzweig M., et al. Activation of MET via diverse exon 14 splicing alterations occurs in multiple tumor types and confers clinical sensitivity to MET inhibitors. Cancer Discov. 2015;5:850–859. doi: 10.1158/2159-8290.CD-15-0285. [DOI] [PubMed] [Google Scholar]
- 28.Tong J.H., Yeung S.F., Chan A.W., et al. MET amplification and exon 14 splice site mutation define unique molecular subgroups of non-small cell lung carcinoma with poor prognosis. Clin Cancer Res. 2016;22:3048–3056. doi: 10.1158/1078-0432.CCR-15-2061. [DOI] [PubMed] [Google Scholar]
- 29.Onozato R., Kosaka T., Kuwano H., Sekido Y., Yatabe Y., Mitsudomi T. Activation of MET by gene amplification or by splice mutations deleting the juxtamembrane domain in primary resected lung cancers. J Thorac Oncol. 2009;4:5–11. doi: 10.1097/JTO.0b013e3181913e0e. [DOI] [PubMed] [Google Scholar]
- 30.Awad M.M., Oxnard G.R., Jackman D.M., et al. MET exon 14 mutations in non-small-cell lung cancer area associated with advanced age and stage-dependent MET genomic amplification and c-Met overexpression. J Clin Oncol. 2016;34:721–730. doi: 10.1200/JCO.2015.63.4600. [DOI] [PubMed] [Google Scholar]
- 31.Van Der Steen N., Giovannetti E., Pauwels P., et al. cMET exon 14 skipping: from the structure to the clinic. J Thorac Oncol. 2016;11:1423–1432. doi: 10.1016/j.jtho.2016.05.005. [DOI] [PubMed] [Google Scholar]
- 32.Pilotto S., Gkountakos A., Carbognin L., Scarpa A., Tortora G., Bria E. MET exon 14 juxtamembrane splicing mutations: clinical and therapeutical perspectives for cancer therapy. Ann Transl Med. 2017;5:2. doi: 10.21037/atm.2016.12.33. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Wu Y.L., Smit E.F., Bauer T.M. Capmatinib for patients with non-small cell lung cancer with MET exon 14 skipping mutations: a review of preclinical and clinical studies. Cancer Treat Rev. 2021;95:102173. doi: 10.1016/j.ctrv.2021.102173. [DOI] [PubMed] [Google Scholar]
- 34.Schiering N., Knapp S., Marconi M., et al. Crystal structure of the tyrosine kinase domain of the hepatocyte growth factor receptor c-Met and its complex with the microbial alkaloid K-252a. Proc Natl Acad Sci U S A. 2003;100:12654–12659. doi: 10.1073/pnas.1734128100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Lee J.H., Gao C.F., Lee C.C., Kim M.D., Woude G.F.V. An alternatively spliced form of Met receptor is tumorigenic. Exp Mol Med. 2006;38:565–573. doi: 10.1038/emm.2006.66. [DOI] [PubMed] [Google Scholar]
- 36.Lee C.C., Yamada K.M. Alternatively spliced juxtamembrane domain of a tyrosine kinase receptor is a multifunctional regulatory site - deletion alters cellular tyrosine phosphorylation pattern and facilitates binding of phosphatidylinositol-3-oh kinase to the hepatocyte growth factor receptor. J Biol Chem. 1995;270:507–510. doi: 10.1074/jbc.270.2.507. [DOI] [PubMed] [Google Scholar]
- 37.Shalata W., Yakobson A., Weissmann S., et al. Crizotinib in MET exon 14-mutated or MET-amplified in advanced disease non-small cell lung cancer: a retrospective, single institution experience. Oncology. 2022;100:467–474. doi: 10.1159/000525188. [DOI] [PubMed] [Google Scholar]
- 38.Choi W., Park S.Y., Lee Y., et al. The clinical impact of capmatinib in the treatment of advanced non-small cell lung cancer with MET exon 14 skipping mutation or gene amplification. Cancer Res Treat. 2021;53:1024–1032. doi: 10.4143/crt.2020.1331. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Wu Z.X., Li J., Dong S., Lin L., Zou C., Chen Z.S. Tepotinib hydrochloride for the treatment of non-small cell lung cancer. Drugs Today. 2021;57:265–275. doi: 10.1358/dot.2021.57.4.3238323. [DOI] [PubMed] [Google Scholar]
- 40.Drilon A., Clark J.W., Weiss J., et al. Antitumor activity of crizotinib in lung cancers harboring a MET exon 14 alteration. Nat Med. 2020;26:47–51. doi: 10.1038/s41591-019-0716-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Baltschukat S., Engstler B.S., Huang A., et al. Capmatinib (INC280) is active against models of non-small cell lung cancer and other cancer types with defined mechanisms of MET activation. Clin Cancer Res. 2019;25:3164–3175. doi: 10.1158/1078-0432.CCR-18-2814. [DOI] [PubMed] [Google Scholar]
- 42.Liu X.D., Wang Q., Yang G.J., et al. A novel kinase inhibitor, INCB28060, blocks c-MET-dependent signaling, neoplastic activities, and cross-talk with EGFR and HER-3. Clin Cancer Res. 2011;17:7127–7138. doi: 10.1158/1078-0432.CCR-11-1157. [DOI] [PubMed] [Google Scholar]
- 43.Reungwetwattana T., Liang Y., Zhu V., Ou S.I. The race to target MET exon 14 skipping alterations in non-small cell lung cancer: the why, the how, the who, the unknown, and the inevitable. Lung Cancer. 2017;103:27–37. doi: 10.1016/j.lungcan.2016.11.011. [DOI] [PubMed] [Google Scholar]
- 44.Fujino T., Kobayashi Y., Suda K., et al. Sensitivity and resistance of MET exon 14 mutations in lung cancer to eight MET tyrosine kinase inhibitors in vitro. J Thorac Oncol. 2019;14:1753–1765. doi: 10.1016/j.jtho.2019.06.023. [DOI] [PubMed] [Google Scholar]
- 45.Schuler M., Berardi R., Lim W.T., et al. Molecular correlates of response to capmatinib in advanced non-small-cell lung cancer: clinical and biomarker results from a phase I trial. Ann Oncol. 2020;31:789–797. doi: 10.1016/j.annonc.2020.03.293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Wolf J., Seto T., Han J.Y., et al. Capmatinib in MET exon 14-mutated or MET-amplified non-small-cell lung cancer. N Engl J Med. 2020;383:944–957. doi: 10.1056/NEJMoa2002787. [DOI] [PubMed] [Google Scholar]
- 47.Groen H.J.M., Akerley W., Souquet P.J., et al. Capmatinib in patients with METex14-mutated or high-level MET-amplified advanced non-small- cell lung cancer (NSCLC): results from cohort 6 of the phase 2 GEOMETRY mono-1 study. J Clin Oncol. 2020;38(suppl 15):9520. doi: 10.1200/JCO.2020.38.15_suppl.9520. [DOI] [Google Scholar]
- 48.Shimizu K., Goldfarb M., Suard Y., et al. Three human transforming genes are related to the viral ras oncogenes. Proc Natl Acad Sci U S A. 1983;80:2112–2116. doi: 10.1073/pnas.80.8.2112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 49.Prior I.A., Lewis P.D., Mattos C. A comprehensive survey of Ras mutations in cancer. Cancer Res. 2012;72:2457–2467. doi: 10.1158/0008-5472.CAN-11-2612. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Hong D.S., Fakih M.G., Strickler J.H., et al. KRAS(G12C) Inhibition with sotorasib in advanced solid tumors. N Engl J Med. 2020;383:1207–1217. doi: 10.1056/NEJMoa1917239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Canon J., Rex K., Saiki A.Y., et al. The clinical KRAS(G12C) inhibitor AMG 510 drives anti-tumour immunity. Nature. 2019;575:217–223. doi: 10.1038/s41586-019-1694-1. [DOI] [PubMed] [Google Scholar]
- 52.Hallin J., Engstrom L.D., Hargis L., et al. The KRAS(G12C) inhibitor MRTX849 provides insight toward therapeutic susceptibility of KRAS-mutant cancers in mouse models and patients. Cancer Discov. 2020;10:54–71. doi: 10.1158/2159-8290.CD-19-1167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Ganguly A., Yoo E. Sotorasib: a KRAS(G12C) inhibitor for non-small cell lung cancer. Trends Pharmacol Sci. 2022;43:536–537. doi: 10.1016/j.tips.2022.03.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Li M., Liu L., Liu Z., et al. The status of KRAS mutations in patients with non-small cell lung cancers from mainland China. Oncol Rep. 2009;22:1013–1020. doi: 10.3892/or_00000529. [DOI] [PubMed] [Google Scholar]
- 55.Nuevo-Tapioles C., Philips M.R. The role of KRAS splice variants in cancer biology. Front Cell Dev Biol. 2022;10:1033348. doi: 10.3389/fcell.2022.1033348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Aran V., Masson Domingues P., Carvalho de Macedo F., et al. A cross-sectional study examining the expression of splice variants K-RAS4A and K-RAS4B in advanced non-small-cell lung cancer patients. Lung Cancer. 2018;116:7–14. doi: 10.1016/j.lungcan.2017.12.005. [DOI] [PubMed] [Google Scholar]
- 57.Patek C.E., Arends M.J., Wallace W.A., et al. Mutationally activated K-ras 4A and 4B both mediate lung carcinogenesis. Exp Cell Res. 2008;314:1105–1114. doi: 10.1016/j.yexcr.2007.11.004. [DOI] [PubMed] [Google Scholar]
- 58.Zhang X., Cao J., Miller S.P., Jing H., Lin H. Comparative nucleotide-dependent interactome analysis reveals shared and differential properties of KRas4a and KRas4b. ACS Cent Sci. 2018;4:71–80. doi: 10.1021/acscentsci.7b00440. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Amendola C.R., Mahaffey J.P., Parker S.J., et al. KRAS4A directly regulates hexokinase 1. Nature. 2019;576:482–486. doi: 10.1038/s41586-019-1832-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Tsai F.D., Lopes M.S., Zhou M., et al. K-Ras4A splice variant is widely expressed in cancer and uses a hybrid membrane-targeting motif. Proc Natl Acad Sci U S A. 2015;112:779–784. doi: 10.1073/pnas.1412811112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Yang I.S., Kim S. Isoform specific gene expression analysis of KRAS in the prognosis of lung adenocarcinoma patients. BMC Bioinform. 2018;19(Suppl 1):40. doi: 10.1186/s12859-018-2011-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Salmon M., Paniagua G., Lechuga C.G., et al. KRAS4A induces metastatic lung adenocarcinomas in vivo in the absence of the KRAS4B isoform. Proc Natl Acad Sci U S A. 2021:118. doi: 10.1073/pnas.2023112118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Chen W.C., To M.D., Westcott P.M.K., et al. Targeting KRAS4A splicing through the RBM39/DCAF15 pathway inhibits cancer stem cells. Nat Commun. 2021;12:4288. doi: 10.1038/s41467-021-24498-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Hu Z. The future of immune checkpoint blockade immunotherapy: towards personalized therapy or towards combination therapy. J Thorac Dis. 2017;9:4226–4229. doi: 10.21037/jtd.2017.10.31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Sharma P., Allison J.P. The future of immune checkpoint therapy. Science. 2015;348:56–61. doi: 10.1126/science.aaa8172. [DOI] [PubMed] [Google Scholar]
- 66.Smyth M.J., Ngiow S.F., Ribas A., Teng M.W. Combination cancer immunotherapies tailored to the tumour microenvironment. Nat Rev Clin Oncol. 2016;13:143–158. doi: 10.1038/nrclinonc.2015.209. [DOI] [PubMed] [Google Scholar]
- 67.Pitt J.M., Vetizou M., Daillere R., et al. Resistance mechanisms to immune-checkpoint blockade in cancer: tumor-intrinsic and -extrinsic factors. Immunity. 2016;44:1255–1269. doi: 10.1016/j.immuni.2016.06.001. [DOI] [PubMed] [Google Scholar]
- 68.Diesendruck Y., Benhar I. Novel immune check point inhibiting antibodies in cancer therapy-opportunities and challenges. Drug Resist Updates. 2017;30:39–47. doi: 10.1016/j.drup.2017.02.001. [DOI] [PubMed] [Google Scholar]
- 69.Sun J.J., Bai J.L., Jiang T., Gao Y., Hua Y.M. Modulation of PDCD1 exon 3 splicing. RNA Biol. 2019;16:1794–1805. doi: 10.1080/15476286.2019.1659080. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Sorensen S.F., Demuth C., Weber B., Sorensen B.S., Meldgaard P. Increase in soluble PD-1 is associated with prolonged survival in patients with advanced EGFR-mutated non-small cell lung cancer treated with erlotinib. Lung Cancer. 2016;100:77–84. doi: 10.1016/j.lungcan.2016.08.001. [DOI] [PubMed] [Google Scholar]
- 71.Khan M., Zhao Z., Arooj S., Fu Y., Liao G. Soluble PD-1: predictive, prognostic, and therapeutic value for cancer immunotherapy. Front Immunol. 2020;11 doi: 10.3389/fimmu.2020.587460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Gong B., Kiyotani K., Sakata S., et al. Secreted PD-L1 variants mediate resistance to PD-L1 blockade therapy in non-small cell lung cancer. J Exp Med. 2019;216:982–1000. doi: 10.1084/jem.20180870. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Yu J., Peng W., Xue Y.B., Li Y., Yang L., Geng Y. FUBP1 promotes the proliferation of lung squamous carcinoma cells and regulates tumor immunity through PD-L1. Allergol Immunopathol. 2022;50:68–74. doi: 10.15586/aei.v50i5.659. [DOI] [PubMed] [Google Scholar]
- 74.Seiler M., Peng S., Agrawal A.A., et al. Somatic mutational landscape of splicing factor genes and their functional consequences across 33 cancer types. Cell Rep. 2018;23:282–296.e4. doi: 10.1016/j.celrep.2018.01.088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Sveen A., Kilpinen S., Ruusulehto A., Lothe R.A., Skotheim R.I. Aberrant RNA splicing in cancer; expression changes and driver mutations of splicing factor genes. Oncogene. 2016;35:2413–2427. doi: 10.1038/onc.2015.318. [DOI] [PubMed] [Google Scholar]
- 76.Makarov E.M., Owen N., Bottrill A., Makarova O.V. Functional mammalian spliceosomal complex E contains SMN complex proteins in addition to U1 and U2 snRNPs. Nucleic Acids Res. 2012;40:2639–2652. doi: 10.1093/nar/gkr1056. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Inoue A. RBM10: structure, functions, and associated diseases. Gene. 2021;783 doi: 10.1016/j.gene.2021.145463. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Sun Y., Bao Y.F., Han W.J., et al. Autoregulation of RBM10 and cross-regulation of RBM10/RBM5 via alternative splicing-coupled nonsense-mediated decay. Nucleic Acids Res. 2017;45:8524–8540. doi: 10.1093/nar/gkx508. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Bechara Elias G., Sebestyén E., Bernardis I., Eyras E., Valcárcel J. RBM5, 6, and 10 differentially regulate NUMB alternative splicing to control cancer cell proliferation. Mol Cell. 2013;52:720–733. doi: 10.1016/j.molcel.2013.11.010. [DOI] [PubMed] [Google Scholar]
- 80.Collins K.M., Kainov Y.A., Christodolou E., et al. An RRM-ZnF RNA recognition module targets RBM10 to exonic sequences to promote exon exclusion. Nucleic Acids Res. 2017;45:6761–6774. doi: 10.1093/nar/gkx225. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Giannakis M., Mu X.J., Shukla S.A., et al. Genomic correlates of immune-cell infiltrates in colorectal carcinoma. Cell Rep. 2016;15:857–865. doi: 10.1016/j.celrep.2016.03.075. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Bailey P., Chang D.K., Nones K., et al. Genomic analyses identify molecular subtypes of pancreatic cancer. Nature. 2016;531:47–52. doi: 10.1038/nature16965. [DOI] [PubMed] [Google Scholar]
- 83.Nordentoft I., Lamy P., Birkenkamp-Demtroder K., et al. Mutational context and diverse clonal development in early and late bladder cancer. Cell Rep. 2014;7:1649–1663. doi: 10.1016/j.celrep.2014.04.038. [DOI] [PubMed] [Google Scholar]
- 84.Zhang S., Bao Y., Shen X., et al. RNA binding motif protein 10 suppresses lung cancer progression by controlling alternative splicing of eukaryotic translation initiation factor 4H. EBioMedicine. 2020;61 doi: 10.1016/j.ebiom.2020.103067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Li H., Sun Z., Xiao R., et al. Stepwise evolutionary genomics of early-stage lung adenocarcinoma manifesting as pure, heterogeneous and part-solid ground-glass nodules. Br J Cancer. 2022;127:747–756. doi: 10.1038/s41416-022-01821-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Chen Y.J., Roumeliotis T.I., Chang Y.H., et al. Proteogenomics of non-smoking lung cancer in East Asia delineates molecular signatures of pathogenesis and progression. Cell. 2020;182:226–244. doi: 10.1016/j.cell.2020.06.012. [DOI] [PubMed] [Google Scholar]
- 87.Zhao J., Sun Y., Huang Y., et al. Functional analysis reveals that RBM10 mutations contribute to lung adenocarcinoma pathogenesis by deregulating splicing. Sci Rep. 2017;7 doi: 10.1038/srep40488. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Bao Y., Zhang S., Zhang X., et al. RBM10 loss promotes EGFR-driven lung cancer and confers sensitivity to spliceosome inhibition. Cancer Res. 2023;22:1549. doi: 10.1158/0008-5472.CAN-22-1549. [DOI] [PubMed] [Google Scholar]
- 89.Cao Y., Geng J., Wang X., et al. RNA-binding motif protein 10 represses tumor progression through the Wnt/beta-catenin pathway in lung adenocarcinoma. Int J Biol Sci. 2022;18:124–139. doi: 10.7150/ijbs.63598. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.Jin X., Di X., Wang R., et al. RBM10 inhibits cell proliferation of lung adenocarcinoma via RAP1/AKT/CREB signalling pathway. J Cell Mol Med. 2019;23:3897–3904. doi: 10.1111/jcmm.14263. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Cao Y., Di X., Cong S., et al. RBM10 recruits METTL3 to induce N6-methyladenosine-MALAT1-dependent modification, inhibiting the invasion and migration of NSCLC. Life Sci. 2023;315 doi: 10.1016/j.lfs.2022.121359. [DOI] [PubMed] [Google Scholar]
- 92.Nanjo S., Wu W., Karachaliou N., et al. Deficiency of the splicing factor RBM10 limits EGFR inhibitor response in EGFR-mutant lung cancer. J Clin Investig. 2022;132 doi: 10.1172/JCI145099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Liu B., Wang Y.Q., Wang H., et al. RBM10 deficiency is associated with increased immune activity in lung adenocarcinoma. Front Oncol. 2021;11 doi: 10.3389/fonc.2021.677826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Hu C., Zhao L., Liu W., et al. Genomic profiles and their associations with TMB, PD-L1 expression, and immune cell infiltration landscapes in synchronous multiple primary lung cancers. J Immunother Cancer. 2021;9 doi: 10.1136/jitc-2021-003773. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Cao Y., Pang L., Jin S. RBM10 is a biomarker associated with pan-cancer prognosis and immune infiltration: system analysis combined with in vitro and vivo experiments. Oxid Med Cell Longev. 2022;2022 doi: 10.1155/2022/7654937.96. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Cao Y., Di X., Zhang Q., Li R., Wang K. RBM10 regulates tumor apoptosis, proliferation, and metastasis. Front Oncol. 2021;11 doi: 10.3389/fonc.2021.603932. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Sun X.N., Jia M.Q., Sun W., Feng L., Gu C.D., Wu T.H. Functional role of RBM10 in lung adenocarcinoma proliferation. Int J Oncol. 2019;54:467–478. doi: 10.3892/ijo.2018.4643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Loiselle J.J., Sutherland L.C. RBM10: harmful or helpful-many factors to consider. J Cell Biochem. 2018;119:3809–3818. doi: 10.1002/jcb.26644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Jung J.H., Lee H., Cao B., Liao P., Zeng S.X., Lu H. RNA-binding motif protein 10 induces apoptosis and suppresses proliferation by activating p53. Oncogene. 2020;39:1031–1040. doi: 10.1038/s41388-019-1034-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Wang E., Pineda J.M.B., Kim W.J., et al. Modulation of RNA splicing enhances response to BCL2 inhibition in leukemia. Cancer Cell. 2023;41:164–180.e8. doi: 10.1016/j.ccell.2022.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Yan C., Wan R., Shi Y. Molecular mechanisms of pre-mRNA splicing through structural biology of the spliceosome. Cold Spring Harb Perspect Biol. 2019;11:a032409. doi: 10.1101/cshperspect.a032409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Zhao Y.J., Cai W.L., Hua Y., Yang X.C., Zhou J.D. The biological and clinical consequences of RNA splicing factor U2AF1 mutation in myeloid malignancies. Cancers. 2022;14:4406. doi: 10.3390/cancers14184406. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Imielinski M., Berger A.H., Hammerman P.S., et al. Mapping the hallmarks of lung adenocarcinoma with massively parallel sequencing. Cell. 2012;150:1107–1120. doi: 10.1016/j.cell.2012.08.029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Esfahani M.S., Lee L.J., Jeon Y.J., et al. Functional significance of U2AF1 S34F mutations in lung adenocarcinomas. Nat Commun. 2019;10:5712. doi: 10.1038/s41467-019-13392-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Palangat M., Anastasakis D.G., Fei D.L., et al. The splicing factor U2AF1 contributes to cancer progression through a noncanonical role in translation regulation. Genes Dev. 2019;33:482–497. doi: 10.1101/gad.319590.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Foy A., McMullin M.F. Somatic SF3B1 mutations in myelodysplastic syndrome with ring sideroblasts and chronic lymphocytic leukaemia. J Clin Pathol. 2019;72:778–782. doi: 10.1136/jclinpath-2019-205895. [DOI] [PubMed] [Google Scholar]
- 107.Seiler M., Peng S.Y., Agrawal A.A., et al. Somatic mutational landscape of splicing factor genes and their functional consequences across 33 cancer types. Cell Rep. 2018;23:282–296.e4. doi: 10.1016/j.celrep.2018.01.088. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Alsafadi S., Houy A., Battistella A., et al. Cancer-associated SF3B1 mutations affect alternative splicing by promoting alternative branchpoint usage. Nat Commun. 2016;7 doi: 10.1038/ncomms10615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Alsafadi S., Dayot S., Tarin M., et al. Genetic alterations of SUGP1 mimic mutant-SF3B1splice pattern in lung adenocarcinoma and other cancers. Oncogene. 2021;40:85–96. doi: 10.1038/s41388-020-01507-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Zhang J., Ali A.M., Lieu Y.K., et al. Disease-causing mutations in SF3B1 alter splicing by disrupting interaction with SUGP1. Mol Cell. 2019;76:82–95.e7. doi: 10.1016/j.molcel.2019.07.017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Debaize L., Troadec M.B. The master regulator FUBP1: its emerging role in normal cell function and malignant development. Cell Mol Life Sci. 2019;76:259–281. doi: 10.1007/s00018-018-2933-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Zhang J., Chen Q.M. Far upstream element binding protein 1: a commander of transcription, translation and beyond. Oncogene. 2013;32:2907–2916. doi: 10.1038/onc.2012.350. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Malz M., Bovet M., Samarin J., et al. Overexpression of far upstream element (FUSE) binding protein (FBP)-interacting Repressor (FIR) supports growth of hepatocellular carcinoma. Hepatology. 2014;60:1241–1250. doi: 10.1002/hep.27218. [DOI] [PubMed] [Google Scholar]
- 114.Ding Z.M., Liu X.C., Liu Y.H., et al. Expression of far upstream element (FUSE) binding protein 1 in human glioma is correlated with c-Myc and cell proliferation. Mol Carcinog. 2015;54:405–415. doi: 10.1002/mc.22114. [DOI] [PubMed] [Google Scholar]
- 115.Venturutti L., Russo R.I.C., Rivas M.A., et al. MiR-16 mediates trastuzumab and lapatinib response in ErbB-2-positive breast and gastric cancer via its novel targets CCNJ and FUBP1. Oncogene. 2016;35:6189–6202. doi: 10.1038/onc.2016.151. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Liu Z.H., Hu J.L., Liang J.Z., et al. Far upstream element-binding protein 1 is a prognostic biomarker and promotes nasopharyngeal carcinoma progression. Cell Death Dis. 2015;6 doi: 10.1038/cddis.2015.258. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Penzvalto Z., Lanczky A., Lenart J., et al. MEK1 is associated with carboplatin resistance and is a prognostic biomarker in epithelial ovarian cancer. BMC Cancer. 2014;14:837. doi: 10.1186/1471-2407-14-837. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Jacob A.G., Singh R.K., Mohammad F., Bebee T.W., Chandler D.S. The splicing factor FUBP1 is required for the efficient splicing of oncogene MDM2 pre-mRNA. J Biol Chem. 2014;289:17350–17364. doi: 10.1074/jbc.M114.554717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Bailey M.H., Tokheim C., Porta-Pardo E., et al. Comprehensive characterization of cancer driver genes and mutations. Cell. 2018;173:371–385.e18. doi: 10.1016/j.cell.2018.02.060. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Qian X.Y., Yang J.Z., Qiu Q.Z., et al. LCAT3, a novel m6A-regulated long non-coding RNA, plays an oncogenic role in lung cancer via binding with FUBP1 to activate c-MYC. J Hematol Oncol. 2021;14:112. doi: 10.1186/s13045-021-01123-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Das T., Lee E.Y., You H.J., Kim E.E., Song E.J. USP15 and USP4 facilitate lung cancer cell proliferation by regulating the alternative splicing of SRSF1. Cell Death Discov. 2022;8:24. doi: 10.1038/s41420-022-00820-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Ye Y., Yu F., Li Z., Xie Y., Yu X. RNA binding protein serine/arginine splicing factor 1 promotes the proliferation, migration and invasion of hepatocellular carcinoma by interacting with RecQ protein-like 4 mRNA. Bioengineered. 2021;12:6144–6154. doi: 10.1080/21655979.2021.1972785. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Anczukow O., Akerman M., Clery A., et al. SRSF1-regulated alternative splicing in breast cancer. Mol Cell. 2015;60:105–117. doi: 10.1016/j.molcel.2015.09.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Haferlach T., Nagata Y., Grossmann V., et al. Landscape of genetic lesions in 944 patients with myelodysplastic syndromes. Leukemia. 2014;28:241–247. doi: 10.1038/leu.2013.336. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Thol F., Kade S., Schlarmann C., et al. Frequency and prognostic impact of mutations in SRSF2, U2AF1, and ZRSR2 in patients with myelodysplastic syndromes. Blood. 2012;119:3578–3584. doi: 10.1182/blood-2011-12-399337. [DOI] [PubMed] [Google Scholar]
- 126.Gout S., Brambilla E., Boudria A., et al. Abnormal expression of the pre-mRNA splicing regulators SRSF1, SRSF2, SRPK1 and SRPK2 in non small cell lung carcinoma. PLoS One. 2012;7 doi: 10.1371/journal.pone.0046539. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Edmond V., Merdzhanova G., Gout S., Brambilla E., Gazzeri S., Eymin B. A new function of the splicing factor SRSF2 in the control of E2F1-mediated cell cycle progression in neuroendocrine lung tumors. Cell Cycle. 2013;12:1267–1278. doi: 10.4161/cc.24363. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Abou Faycal C., Gazzeri S., Eymin B. A VEGF-A/SOX2/SRSF2 network controls VEGFR1 pre-mRNA alternative splicing in lung carcinoma cells. Sci Rep. 2019;9:336. doi: 10.1038/s41598-018-36728-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Cohen-Eliav M., Golan-Gerstl R., Siegfried Z., et al. The splicing factor SRSF6 is amplified and is an oncoprotein in lung and colon cancers. J Pathol. 2013;229:630–639. doi: 10.1002/path.4129. [DOI] [PubMed] [Google Scholar]
- 130.Grosso A.R., Martins S., Carmo-Fonseca M. The emerging role of splicing factors in cancer. EMBO Rep. 2008;9:1087–1093. doi: 10.1038/embor.2008.189. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Frankiw L., Baltimore D., Li G.D. Alternative mRNA splicing in cancer immunotherapy. Nat Rev Immunol. 2019;19:675–687. doi: 10.1038/s41577-019-0195-7. [DOI] [PubMed] [Google Scholar]
- 132.Folco E.G., Coil K.E., Reed R. The anti-tumor drug E7107 reveals an essential role for SF3b in remodeling U2 snRNP to expose the branch point-binding region. Genes Dev. 2011;25:440–444. doi: 10.1101/gad.2009411. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Finci L.I., Zhang X.F., Huang X.L., et al. The cryo-EM structure of the SF3b spliceosome complex bound to a splicing modulator reveals a pre-mRNA substrate competitive mechanism of action. Genes Dev. 2018;32:309–320. doi: 10.1101/gad.311043.117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Seiler M., Yoshimi A., Darman R., et al. H3B-8800, an orally available small-molecule splicing modulator, induces lethality in spliceosome-mutant cancers. Nat Med. 2018;24:497–504. doi: 10.1038/nm.4493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Bates D.O., Morris J.C., Oltean S., Donaldson L.F. Pharmacology of modulators of alternative splicing. Pharmacol Rev. 2017;69:63–79. doi: 10.1124/pr.115.011239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Aird D., Teng T., Huang C.L., et al. Sensitivity to splicing modulation of BCL2 family genes defines cancer therapeutic strategies for splicing modulators. Nat Commun. 2019;10:137. doi: 10.1038/s41467-018-08150-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 137.Eskens F.A., Ramos F.J., Burger H., et al. Phase I pharmacokinetic and pharmacodynamic study of the first-in-class spliceosome inhibitor E7107 in patients with advanced solid tumors. Clin Cancer Res. 2013;19:6296–6304. doi: 10.1158/1078-0432.CCR-13-0485. [DOI] [PubMed] [Google Scholar]
- 138.Hong D.S., Kurzrock R., Naing A., et al. A phase I, open-label, single-arm, dose-escalation study of E7107, a precursor messenger ribonucleic acid (pre-mRNA) splicesome inhibitor administered intravenously on days 1 and 8 every 21 days to patients with solid tumors. Invest New Drugs. 2014;32:436–444. doi: 10.1007/s10637-013-0046-5. [DOI] [PubMed] [Google Scholar]
- 139.Wu T.F., Millar H., Gaffney D., et al. JNJ-64619178, a selective and pseudo-irreversible PRMT5 inhibitor with potent in vitro and in vivo activity, demonstrated in several lung cancer models. Cancer Res. 2018;78:4859. doi: 10.1158/1538-7445.AM2018-4859. [DOI] [Google Scholar]
- 140.Villar M.V., Spreafico A., Moreno V., et al. First-in-human study of JNJ-64619178, a protein arginine methyltransferase 5 (PRMT5) inhibitor, in patients with advanced cancers. Ann Oncol. 2020;31:S470. doi: 10.1016/j.annonc.2020.08.651. [DOI] [Google Scholar]
- 141.Brehmer D., Beke L., Wu T.F., et al. Discovery and pharmacological characterization of JNJ-64619178, a novel small-molecule inhibitor of PRMT5 with potent antitumor activity. Mol Cancer Ther. 2021;20:2317–2328. doi: 10.1158/1535-7163.MCT-21-0367. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Uehara T., Minoshima Y., Sagane K., et al. Selective degradation of splicing factor CAPER alpha by anticancer sulfonamides. Nat Chem Biol. 2017;13:675–680. doi: 10.1038/nchembio.2363. [DOI] [PubMed] [Google Scholar]
- 143.Han T., Goralski M., Gaskill N., et al. Anticancer sulfonamides target splicing by inducing RBM39 degradation via recruitment to DCAF15. Science. 2017;356:aal3755. doi: 10.1126/science.aal3755. [DOI] [PubMed] [Google Scholar]
- 144.Kole R., Krainer A.R., Altman S. RNA therapeutics: beyond RNA interference and antisense oligonucleotides. Nat Rev Drug Discov. 2012;11:125–140. doi: 10.1038/nrd3625. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 145.Moreno P.M.D., Pego A.P. Therapeutic antisense oligonucleotides against cancer: hurdling to the clinic. Front Chem. 2014;2:87. doi: 10.3389/fchem.2014.00087. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Komaki H., Nagata T., Saito T., et al. Systemic administration of the antisense oligonucleotide NS-065/NCNP-01 for skipping of exon 53 in patients with Duchenne muscular dystrophy. Sci Transl Med. 2018;10 doi: 10.1126/scitranslmed.aan0713. [DOI] [PubMed] [Google Scholar]
- 147.Zhang S.R., Bao Y.F., Shen X.F., et al. RNA binding motif protein 10 suppresses lung cancer progression by controlling alternative splicing of eukaryotic translation initiation factor 4H. EBioMedicine. 2020;61 doi: 10.1016/j.ebiom.2020.103067. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 148.Anczukow O., Rosenberg A.Z., Akerman M., et al. The splicing factor SRSF1 regulates apoptosis and proliferation to promote mammary epithelial cell transformation. Nat Struct Mol Biol. 2012;19:220–228. doi: 10.1038/nsmb.2207. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.Hong D., Kim Y., Younes A., et al. Preclinical pharmacology and clinical efficacy of AZD9150 (ISIS-STAT3Rx), a potent next-generation antisense oligonucleotide inhibitor of STAT3. Cancer Res. 2014;74 doi: 10.1158/1538-7445.am2014-lb-227. LB-227-LB-227. [DOI] [Google Scholar]
- 150.Ross S.J., Revenko A.S., Hanson L.L., et al. Targeting KRAS-dependent tumors with AZD4785, a high-affinity therapeutic antisense oligonucleotide inhibitor of KRAS. Sci Transl Med. 2017;9 doi: 10.1126/scitranslmed.aal5253. [DOI] [PubMed] [Google Scholar]
- 151.Otsubo K., Nosaki K., Imamura C.K., et al. Phase I study of salazosulfapyridine in combination with cisplatin and pemetrexed for advanced non-small-cell lung cancer. Cancer Sci. 2017;108:1843–1849. doi: 10.1111/cas.13309. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 152.Tanimoto A., Takeuchi S., Arai S., et al. Histone deacetylase 3 inhibition overcomes BIM deletion polymorphism-mediated osimertinib resistance in EGFR-mutant lung cancer. Clin Cancer Res. 2017;23:3139–3149. doi: 10.1158/1078-0432.CCR-16-2271. [DOI] [PubMed] [Google Scholar]
- 153.Takeuchi S., Hase T., Shimizu S., et al. Phase I study of vorinostat with gefitinib in BIM deletion polymorphism/epidermal growth factor receptor mutation double-positive lung cancer. Cancer Sci. 2020;111:561–570. doi: 10.1111/cas.14260. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Shultz J.C., Goehe R.W., Murudkar C.S., et al. SRSF1 regulates the alternatives splicing of caspase 9 via a novel intronic splicing enhancer affecting the chemotherapeutic sensitivity of non-small cell lung cancer cells. Mol Cancer Res. 2011;9:889–900. doi: 10.1158/1541-7786.MCR-11-0061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 155.Matsuura S., Kahyo T., Shinmura K., et al. SGOL1 variant B induces abnormal mitosis and resistance to taxane in non-small cell lung cancers. Sci Rep. 2013;3:3012. doi: 10.1038/srep03012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Liu N.L., Chen A.X., Feng N., Liu X.C., Zhang L.Z. SNRPB is a mediator for cellular response to cisplatin in non-small-cell lung cancer. Med Oncol. 2021;38:57. doi: 10.1007/s12032-021-01502-0. [DOI] [PubMed] [Google Scholar]
- 157.Passaro A., Jänne P.A., Mok T., Peters S. Overcoming therapy resistance in EGFR-mutant lung cancer. Nat Cancer. 2021;2:377–391. doi: 10.1038/s43018-021-00195-8. [DOI] [PubMed] [Google Scholar]
- 158.Hsu C.C., Liao B.C., Liao W.Y., et al. Exon 16-skipping HER2 as a novel mechanism of osimertinib resistance in EGFR L858R/T790M-positive non-small cell lung cancer. J Thorac Oncol. 2020;15:50–61. doi: 10.1016/j.jtho.2019.09.006. [DOI] [PubMed] [Google Scholar]
- 159.Ng K.P., Hillmer A.M., Chuah C.T.H., et al. A common BIM deletion polymorphism mediates intrinsic resistance and inferior responses to tyrosine kinase inhibitors in cancer. Nat Med. 2012;18:521–528. doi: 10.1038/nm.2713. [DOI] [PubMed] [Google Scholar]
- 160.Hatat A.S., Benoit-Pilven C., Pucciarelli A., et al. Altered splicing of ATG16-L1 mediates acquired resistance to tyrosine kinase inhibitors of EGFR by blocking autophagy in non-small cell lung cancer. Mol Oncol. 2022;16:3490–3508. doi: 10.1002/1878-0261.13229. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 161.Frankiw L., Baltimore D., Li G. Alternative mRNA splicing in cancer immunotherapy. Nat Rev Immunol. 2019;19:675–687. doi: 10.1038/s41577-019-0195-7. [DOI] [PubMed] [Google Scholar]
- 162.Pan Y., Kadash-Edmondson K.E., Wang R., et al. RNA dysregulation: an expanding source of cancer immunotherapy targets. Trends Pharmacol Sci. 2021;42:268–282. doi: 10.1016/j.tips.2021.01.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 163.Peng Q., Zhou Y., Oyang L., et al. Impacts and mechanisms of alternative mRNA splicing in cancer metabolism, immune response, and therapeutics. Mole Ther. 2022;30:1018–1035. doi: 10.1016/j.ymthe.2021.11.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 164.Carbone D.P., Reck M., Paz-Ares L., et al. First-line nivolumab in stage IV or recurrent non-small-cell lung cancer. N Engl J Med. 2017;376:2415–2426. doi: 10.1056/NEJMoa1613493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Hellmann M.D., Ciuleanu T.E., Pluzanski A., et al. Nivolumab plus ipilimumab in lung cancer with a high tumor mutational burden. N Engl J Med. 2018;378:2093–2104. doi: 10.1056/NEJMoa1801946. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 166.Lu S.X., De Neef E., Thomas J.D., et al. Pharmacologic modulation of RNA splicing enhances anti-tumor immunity. Cell. 2021;184:4032–4047. doi: 10.1016/j.cell.2021.05.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Smart A.C., Margolis C.A., Pimentel H., et al. Intron retention is a source of neoepitopes in cancer. Nat Biotechnol. 2018;36:1056–1058. doi: 10.1038/nbt.4239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Thomas J.D., Polaski J.T., Feng Q., et al. RNA isoform screens uncover the essentiality and tumor-suppressor activity of ultraconserved poison exons. Nat Genet. 2020;52:84–94. doi: 10.1038/s41588-019-0555-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Xu F., Li C.H., Wong C.H., et al. Genome-wide screening and functional analysis identifies tumor suppressor long noncoding RNAs epigenetically silenced in hepatocellular carcinoma. Cancer Res. 2019;79:1305–1317. doi: 10.1158/0008-5472.CAN-18-1659. [DOI] [PubMed] [Google Scholar]
- 170.Liu Y., Cao Z., Wang Y., et al. Genome-wide screening for functional long noncoding RNAs in human cells by Cas9 targeting of splice sites. Nat Biotechnol. 2018;36:1203–1210. doi: 10.1038/nbt.4283. [DOI] [PubMed] [Google Scholar]
- 171.Li F., Huang Q., Luster T.A., et al. In vivo epigenetic CRISPR screen identifies Asf1a as an immunotherapeutic target in Kras-mutant lung adenocarcinoma. Cancer Discov. 2020;10:270–287. doi: 10.1158/2159-8290.CD-19-0780. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Li F., Ng W.L., Luster T.A., et al. Epigenetic CRISPR screens identify Npm1 as a therapeutic vulnerability in non-small cell lung cancer. Cancer Res. 2020;80:3556–3567. doi: 10.1158/0008-5472.CAN-19-3782. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Pickar-Oliver A., Gersbach C.A. The next generation of CRISPR-Cas technologies and applications. Nat Rev Mol Cell Biol. 2019;20:490–507. doi: 10.1038/s41580-019-0131-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 174.Wang G., Chow R.D., Zhu L., et al. CRISPR-GEMM pooled mutagenic screening identifies KMT2D as a major modulator of immune checkpoint blockade. Cancer Discov. 2020;10:1912–1933. doi: 10.1158/2159-8290.CD-19-1448. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Data Availability Statement
The datasets used in the current study are available from the corresponding author on reasonable request.




