Next Article in Journal
Inflammatory Pathways to Carcinogenesis: Deciphering the Rheumatoid Arthritis–Lung Cancer Connection
Previous Article in Journal
Combined TaTME with SP Robot for Low Anterior Resection in Rectal Cancer: rSPa TaTME
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Alternative Splicing in Lung Adenocarcinoma: From Bench to Bedside

Department of Surgery, Prince of Wales Hospital, The Chinese University of Hong Kong, Hong Kong SAR, China
*
Authors to whom correspondence should be addressed.
Cancers 2025, 17(), 1329; https://doi.org/10.3390/cancers17081329
Submission received: 25 January 2025 / Revised: 7 April 2025 / Accepted: 14 April 2025 / Published: 15 April 2025
(This article belongs to the Section Molecular Cancer Biology)

Simple Summary

Lung adenocarcinoma is the most prevalent pathological type of lung cancer. Alternative splicing generates multiple protein products from a single gene, significantly contributing to proteome diversity. Dysregulated splicing influences the development of lung adenocarcinoma and creates novel neoantigens for current immunotherapies, such as Chimeric Antigen Receptor T cell therapy. This review discusses alternative splicing events and therapeutic strategies, highlighting the potential to enhance anti-cancer treatments for lung adenocarcinoma.

Abstract

Lung adenocarcinoma (LUAD) is a highly heterogeneous tumor and the most prevalent pathological type of lung cancer. The alternative splicing (AS) of mRNA enables the generation of multiple protein products from a single gene. This is a tightly regulated process that significantly contributes to the proteome diversity in eukaryotes. Recent multi-omics studies have delineated the splicing profiles that underline LUAD tumorigenesis from initiation to metastasis. Such progress holds robust promise to facilitate the development of screening strategies and individualized therapies. Perturbed AS fosters the emergence of novel neoantigen resources and disturbances in the immune microenvironment, which allow new investigations into modulatory targets for LUAD immunotherapy. This review presents an update on the landscape of dysregulated splicing events in LUAD and the associated mechanisms and theranostic perspectives with unique insights into AS-based immunotherapy, such as Chimeric Antigen Receptor T cell therapy. These AS variants can be used in conjunction with current therapeutic modules in LUAD, allowing bench to bedside translation to combat this highly malignant cancer.

1. Introduction

With an estimated 2.20 million new cases and 1.79 million deaths annually, lung cancer is one of the most common and lethal cancer types worldwide [1]. Two primary forms of lung cancer are non-small-cell lung cancer (NSCLC) and small-cell lung cancer, although NSCLC comprises approximately 85% of cases. Lung adenocarcinoma (LUAD) is the most prevalent pathological subtype of NSCLC and is characterized by high genomic heterogeneity [2,3,4].
Significant achievements in the molecular profiling for LUAD, the utilization of predictive biomarkers, and the advancement of therapies have led to considerable progress and improved outcomes for many patients [2,5,6]. Recently, the development of high-quality genome assemblies and recent innovations in long-read sequencing technology have elucidated the intracellular mechanisms of alternative splicing (AS) and facilitated comparisons of splicing profiles among closely related organisms, offering insights into the genetic complexity and oncogenicity of LUAD [7,8,9,10].
AS generates multiple polypeptides from a single gene that significantly enhances proteomic complexity and contributes towards diverse phenotypic traits [11,12]. Splicing aberrations have been implicated in the initiation and progression of various cancers, with advanced therapeutics targeting AS events emerging as novel anti-tumor strategies, including those for LUAD [13,14,15,16]. Notably, recent studies have illustrated the contribution of AS to the antigen repertoire of cancer cells and have demonstrated the potential for expanding immunotherapies that recognize novel AS-derived epitopes [17,18]. The putative neoantigens generated through AS serve as a substantial source for tumor-specific therapeutic targets [19,20,21].
In this review, we highlight aberrant splicing events as major drivers of LUAD and summarize therapeutic strategies targeting splicing alterations. Additionally, we explore the potential and perspectives of AS-based anti-tumor strategies, such as Chimeric Antigen Receptor (CAR) T cell therapy.

2. The Mechanism of AS in LUAD

AS is a complex mechanism underscoring eukaryotic gene expression that allows a single gene to produce multiple proteins. It is a process, by which different combinations of exons are joined together, or “spliced”, from a pre-mRNA transcript during RNA processing, contributing to the diversity and adaptability in protein structures and functions without increasing genome size [22]. For example, the epidermal growth factor receptor (EGFR) is a vital membrane protein that regulates cell signaling and growth, mediating proliferation and differentiation in LUAD [23,24]. According to the NCBI Gene database, EGFR has nine isoforms, a through i, each derived from distinct splicing mechanisms and exhibiting unique biological functions. All EGFR-AS isoforms present unique crystal protein structure predicted by AlphaFold 3.0 [25], as illustrated in Figure 1. Alterations of diverse functional domains lead to different amino acids (a.a.s), featuring diverse protein sizes and putative subcellular locations, as shown in Table 1.
According to the splice site usage, AS can be categorized into seven elementary types as follows: exon skipping (ES), intron retention, mutually exclusive exons (MX), alternative 5′ splice site and alternative 3′ splice site selection, and alternative first exon and alternative last exon splicing, as illustrated in Figure 2 [11]. We reviewed the splicing data profiled in TCGA SpliceSeq by the MD Anderson Cancer Center and summarized the AS events in LUAD [26], as shown in Figure 3. Among the various AS events investigated, ES emerged as the most predominant splicing pattern, whereas MX was identified as the least prevalent (Figure 3A). Notably, our distribution analysis quantified the number of genes harboring varying frequencies of AS events and revealed that 6677 out of a total of 10,314 genes (64.7%) exhibit multiple splicing events. An increasing number of AS events per gene results in a smaller number of genes (Figure 3B). These results underscore the intrinsic mechanisms of AS that enable LUAD cells to generate a wide diversity of protein isoforms from a limited number of genes.
Increasingly, studies and prognostic splicing models have been developed by unraveling the mechanism of AS and correlating splicing aberrations with the tumor microenvironment and immune heterogeneity in LUAD [27,28]. For example, a recent study constructed a novel AS-based prognosis signature that is associated with regulatory B cells and malignant pleural effusion in LUAD [29]. Furthermore, a multi-center study led by a research team from Switzerland delineates the pan-cancer AS landscape through a comprehensive analysis across tumors from 8705 patients. Among 32 types of cancer, LUAD is recognized as one of the most prominent tumors, characterized by elevated levels of aberrant AS events, particularly ES, which demonstrate an increase of over 30% in tumor samples compared to matched normal tissues. Regarding therapeutic implications, the study elucidates that LUAD exhibits a significant prevalence of tumor-specific exon-exon junctions, indicating an increased complexity of splicing mechanisms and the potential neoantigen generation [20]. The development of advanced strategies necessitates a deeper understanding of the natural splicing properties in LUAD.

3. Emerging Technologies to Assess AS in LUAD

In the past decade, RNA sequencing (RNA-seq) has emerged as an essential methodological advancement for the comprehensive transcriptome-wide analyses of differential gene expression and mRNA splicing [30,31]. A pivotal advantage of this technology is single-cell RNA-seq (scRNA-seq), which facilitates the evaluation of gene expression at the high resolution of individual cells, thereby enabling the identification of distinct cellular populations within LUAD [32,33,34]. Advanced scRNA-seq, integrated with quantitative profiling techniques of both short-read and long-read sequencings, enables researchers to illuminate the intricacies of transcript diversity and AS events that may exhibit dramatic variability among heterogeneous tumor cell subpopulations [35,36]. In recent studies, the application of single-cell sequencing has been instrumental in elucidating the metastatic dynamics and aberrant AS, yielding novel therapeutic targets for various malignancies, including acute myeloid leukemia [37], ovarian cancer [38], and LUAD [39,40].
Furthermore, emerging advancements in spatial transcriptomics extend the genomic analysis by preserving the spatial arrangement of LUAD architecture [34,41]. Intriguingly, recent studies have introduced spatial isoform transcriptomics (SITs), an explorative methodology designed to characterize spatial AS isoforms and sequence heterogeneity utilizing long-read sequencing. The innovative SITs enable the in situ capture of full-length transcripts across different tissue areas, simultaneously providing a comprehensive perspective on gene expression dynamics and AS landscapes [42,43]. Notably, spatial sequencing combined with RNA-seq has unveiled critical findings in LUAD, revealing that hMENA/hMENAΔv6 splicing isoforms exhibit elevated levels in cancer-associated fibroblasts within tumor tissues, alongside the high expression of stromal FN1 and tertiary lymphoid structures (TLSs) that are preferentially localized in the peritumoral region. The dense expression of FN1 may serve as a barrier to intertumoral TLS localization, in agreement with significant anti-tumor immune responses, and may affect the efficacy of immunotherapy through interactions with ILT3 [44,45].
Researchers are endeavoring to address the limitations inherent in individual technologies while enhancing their capacity to elucidate the biological functions of AS isoforms. ScRNA-seq, employing short-read assembly, offers robust sensitivity for transcript detection at the individual cell level, whereas spatial transcriptomics enables the contextual interrogation of gene expression patterns within tissues. Collectively, the incorporation of long-read sequencing augments analytical depth by delivering intricate details regarding AS events. Effectively, these sophisticated technologies facilitate a holistic approach to examining gene expression and the spatial dynamics of LUAD, ultimately leading to profound insights into the mechanisms of AS within the tumor microenvironment.

4. The AS Events Involving Driver Genes of LUAD

Emerging evidence suggests that LUAD primarily develops from a specialized subset of alveolar type II (AT2) cells [46,47,48]. The intrinsic disorder of AT2 cell self-renewal is mediated by signals transduced through the dysregulation of significant drivers, such as the EGFR–Kirsten rat sarcoma oncogene (KRAS) pathway, which is frequently hijacked during oncogenesis through aberrant AS [49,50]. A crucial factor in this process is the generation of splicing isoforms from major driver genes, which influence the initiation and progression of LUAD by altering normal cellular functions and pathways. A summary of LUAD-related AS isoforms is given in Table 2.

4.1. KRAS

As one of the most frequently altered oncogenes in LUAD, KRAS encodes a small GTPase that mediates the connection between growth factor signaling and several downstream signaling pathways [67,68,69], including the MAPK pathway [70]. The KRAS locus produces two variants, KRAS4A and KRAS4B, which result from the AS of the fourth exon, leading to distinct membrane-targeting sequences. Common activating mutations found in exons 1 or 2 confer oncogenic properties to both splice variants [71,72,73]. Additionally, the expression levels of KRAS4A and KRAS4B show significant variability across different tissues and developmental stages. A recent multi-institutional study reported that KRAS4A expression was significantly elevated in advanced-stage LUAD patients [71,74,75]. Previous research has demonstrated that the stable expression of KRAS4A in in vivo LUAD models significantly promoted proximal metastasis [51].
Intriguingly, the disruption of KRAS splicing switching to preferentially produce KRAS4A, mediated by the regulation of the DCAF15/RBM39 pathway, has been shown to inhibit cellular stemness during lung tumorigenesis, presenting a potential therapeutic strategy for modulating KRAS splicing [76,77]. Furthermore, an exploratory analysis revealed that PD-L1 levels were significantly higher in patients with elevated KRAS isoforms, particularly those with high levels of KRAS4A [78]. Given these findings, KRAS4A merits further investigation as a potential biomarker for identifying patients who may benefit from immune checkpoint inhibitor therapy, thus enhancing personalized cancer immunotherapy.

4.2. EGFR

As a member of the receptor tyrosine kinase family, EGFR plays a crucial role in regulating cell growth and differentiation in response to its natural ligands, such as EGF and TGF-α [79,80,81]. Upon ligand binding, EGFR undergoes dimerization and autophosphorylation, which activates downstream signaling pathways [82,83,84], including the RAS-RAF-MAPK [85] and PI3K-AKT pathways [83,86]. Recently, mutations in EGFR, particularly within the tyrosine kinase domain, have been identified as vital drivers of LUAD tumorigenesis and have been associated with predicted sensitivity to tyrosine kinase inhibitors (TKIs) [87,88,89]. The most prevalent mutations, such as exon 19 deletion and the L858R point mutation in exon 21, lead to the constitutive activation of the receptor, which enhances tumor proliferation, survival, metastasis, and other cancer-associated properties [90,91].
Additionally, the AS of the EGFR gene can generate various isoforms that may significantly alter the receptor’s function and therapeutic responses, particularly to TKIs. The most common splicing variant, EGFRvIII, is caused by the truncation of exons 2–7 and leads to the deletion of a portion of the extracellular domain [92,93,94]. This variant exhibits unique constitutive activity independent of ligand binding, thereby contributing to TKI resistance in lung tumorigenesis [53]. Despite these insights, the regulatory mechanisms governing EGFR splicing in LUAD remain poorly understood and necessitate further investigation to elucidate their impact on tumor biology and therapeutic outcomes.

4.3. C-Met

The activation of the receptor tyrosine kinase, c-Met, has been implicated in the promotion of various cancers via the stimulation of multiple downstream oncogenic signaling pathways, including the MAPK, PI3K/AKT, and STAT pathways [95,96,97]. Aberrant c-Met signaling can facilitate lung tumorigenesis through a range of upstream mechanisms, such as c-Met gene amplification, mutations, rearrangements, and overexpression [98,99,100]. Recently, the exclusion of c-Met exon 14 (designated as MET-ΔEx14) [55] has been identified as a prototypical aberrant splicing event that exhibits oncogenic properties and holds considerable clinical significance in LUAD [101,102,103]. The exclusion of exon 14 leads to an in-frame deletion of 47 amino acids within the juxtamembrane domain of c-Met, which results in inhibited degradation and the internalization of the receptor [104]. Consequently, this alteration is classified as a gain-of-function event. The juxtamembrane domain functions as a crucial negative regulatory region of c-Met, containing a caspase-cleavage sequence and a tyrosine-binding site, which mediates the ubiquitination and subsequent degradation of c-Met [105,106].
Although MET-ΔEx14 occurs in approximately 4% of patients with LUAD [6,107,108], multiple studies have reported on MET-ΔEx14 as an independent oncogenic driver and a biomarker significantly associated with reduced survival rates in these patients [56,109,110]. Furthermore, growing clinical evidence indicates that patients harboring MET-ΔEx14 could benefit from MET TKIs, including crizotinib [111], tepotinib [112,113], and capmatinib [114]. Intriguingly, despite a substantial proportion of LUAD patients with MET-ΔEx14 alteration expressing PD-L1 [115], they often demonstrate a lower overall response rate to immune checkpoint blockade (ICB) therapies and experience shorter median progression-free survival [116].

4.4. CD44

CD44 is a glycoprotein that plays a critical role in cell–cell and cell–matrix interactions. According to the UniProt database, CD44 naturally exists as a diverse array of at least 20 different transmembrane isoforms [117]. Such structural heterogeneity is primarily attributed to AS [118]. The CD44 protein is encoded by a single gene consisting of 20 exons, 10 of which are subject to AS. Exons 1 to 5 and exons 16 to 20 are consistently expressed across various cell types. In contrast, exons 6 to 15 (designated as v1–v10) undergo AS, leading to the generation of numerous variant isoforms.
The protein variability of CD44 contributes to the functional heterogeneity in various biological processes [119,120,121], including cancer cell stemness and migration [122,123]. In LUAD, specific CD44 splicing patterns have been associated with tumor grade, stage, and patient survival [124,125,126]. Notably, the reduced expressions of CD44-v3 and -v6 have been linked to LUAD invasion [127]. CD44 AS variants could also influence the efficacy of anti-tumor immunotherapies, such as CAR T [58] and CAR NK therapy [128].

4.5. PD-1/PD-L1

Targeting the immune checkpoint molecules PD-1 and PD-L1 through ICB has shown significant clinical responses across various cancer types, including lung cancer [129,130], leukemia [131], and lymphoma [132,133]. Anti-PD-1/PD-L1 immunotherapy disrupts the PD-1/PD-L1 signaling pathway [134], which restores T cell activity, enhances anti-tumor immunity, and ultimately contributes to the elimination of cancer cells [135,136].
Despite these advancements, only a subset of patients respond to ICB, underscoring the need for a deeper understanding of the mechanism driving immune escape [59,137]. Research has revealed that exon 3 skipping in PD-1 leads to the generation of a soluble isoform, sPD-1, which lacks a transmembrane domain [138,139]. This soluble variant may antagonize PD-1, potentially enhancing anti-tumor immunity by interfering with the PD-1/PD-L1 signaling axis [140]. A clinical investigation in Denmark indicated that elevated sPD-1 expression is associated with improved outcomes for patients with EGFR-mutant NSCLC receiving the EGFR TKI erlotinib [141]. Moreover, high levels of sPD-1 expression have been linked to increased response rates to immunotherapies, such as anti-PD-1 and anti-CTLA4 therapies in NSCLC, including LUAD [142].
Additionally, a novel spliced variant of PD-L1, referred to as sPD-L1, which also lacks the transmembrane domain, has been identified in NSCLC [143]. The significance of pretreatment with ICB monotherapy for NSCLC patients harboring both sPD-1 and sPD-L1 has been emphasized [144]. Intriguingly, findings suggest that AS can generate a long non-coding RNA from the PD-L1 gene, which promotes the progression of LUAD by enhancing c-Myc activity [145,146]. Collectively, recent evidence highlights the intricate role of AS of PD-1 and PD-L1 in immune regulation and tumor progression, suggesting potential avenues for improving the efficacy of immunotherapy in LUAD.

5. Therapeutic Strategies Targeting AS in LUAD

Aberrant AS has emerged as a significant regulatory mechanism that diversifies the functional repertoire of proteins involved in lung tumorigenesis, thereby contributing to the genetic complexity and heterogeneity in LUAD. To enhance the therapeutic efficacy, recent research underscores the potential in targeting specific splicing events or the splicing machinery in LUAD. This positions AS as an essential component in future personalized cancer treatment, paving the way for innovative approaches to therapy.

5.1. Small-Molecule Inhibitors

The therapeutic strategy of small-molecule inhibitors acting on splicing machinery and regulatory elements is believed to effectively correct abnormal splicing events. By redirecting oncogenic spliced variants to non-oncogenic ones, there exists the potential to restore normal cellular functions and counteract malignancy in LUAD [147,148].
Recent studies have identified nonsynonymous mutations in genes encoding members of the spliceosome complex, such as SF3b1 [149], U2AF1 [150], and SRSF2 [151] in the development of LUAD. H3B-8800 is a small-molecule inhibitor that inducing preferential lethality in spliceosome-mutant epithelial tumor cells, such as LUAD [152]. Mechanistically, H3B-8800 specifically interacts with the SF3b complex, a critical component of the spliceosome complex within the branchpoint binding that occurs during splicing [153]. Intriguingly, the competitive binding of H3B-8800 to mutant SF3b1, which is a critical member of SF3b complexes in its canonical form and induced aberrant splicing events, leads to dose-independent inhibition through the dissociation of the mutant spliceosome [152,153,154]. Moreover, a recent study combined molecular dynamics simulation trajectories with metadynamics simulations to identify the existence of a putative druggable SF3b pocket in the vicinity of the SF3b1-K700E mutation, aiming to develop future therapeutic strategies that specifically target the mutant spliceosome while sparing the wild-type spliceosome [154]. Additionally, H3B-8800 has shown promising results in preclinical studies, demonstrating anti-tumor activity across various spliceosome-mutant cancer models, including LUAD [152]. A phase I clinical trial identified a subset of patients with acute myeloid leukemia who could benefit from H3B-8800 treatment, indicating an acceptable adverse effect profile [155]. Currently, clinical evaluations of H3B-8800 in LUAD remain underdeveloped and resistance to its inhibition has also been observed, underscoring the complexity of targeting splicing factor mutations [152,156,157].
Another spliceosome inhibitor is E7107, which specifically targets the U2 snRNP complex in pre-mRNA splicing [158,159]. Similarly to H3B-8800, E7107 exerts its effects through competitive binding to the mutant SF3b complex [160], effectively blocking spliceosome assembly and thereby activating multiple pro-apoptotic signaling pathways, such as the p53 pathway [161] and the regulation of BCL2 family [162]. Recent clinical trials have underscored its anti-tumor efficacy across a range of solid tumors, including LUAD [163,164]. Certain spliceosome inhibitors, including Spliceostatin A [165] and Pladienolide [166], modulate the fidelity of branch point selection by binding competitively to the SF3b complex, underscoring their potential as therapeutic agents for innovative anti-splicing strategies in LUAD. In conclusion, the potent anti-tumor effects of H3B-8800 and E7107, these two derivatives of pladienolide B [167], demonstrate the therapeutic potential of spliceosome inhibitors in addressing oncogenic splicing dysregulation in LUAD. The functional mechanisms of spliceosome inhibitors are illustrated in Figure 4.
Beyond the spliceosome, small-molecule inhibitors targeting specific splicing regulatory proteins have also demonstrated significant therapeutic potential. These compounds aim to disrupt the function of key splicing factors, thereby altering splicing patterns that drive tumorigenesis.
Protein arginine methyltransferase 5 (PRMT5) is a splicing regulator that plays a critical role in pre-mRNA splicing by methylating spliceosomal proteins. JNJ-64619178, inhibitor of PRMT5, leads to the inhibition of symmetric arginine dimethylation of SMD1/3 proteins, core components of the spliceosome in the tumor and alters aberrant splicing by suppressing PRMT5 enzymatic activity [168,169]. Xenograft models of LUAD have demonstrated that JNJ-64619178 effectively inhibits tumor growth and promotes regression, making it a promising candidate for splicing-targeted therapies [170].
Another key splicing regulator, RBM39, has been targeted by an aryl sulfonamide drug, indisulam [171]. Indisulam functions by recruiting DCAF15, which bridges RBM39 with the E3 ubiquitin ligase complex [172,173]. This action leads to the polyubiquitination and subsequent proteasomal degradation of RBM39, thus disrupting RBM39-dependent pre-mRNA splicing [172]. Intriguingly, a separate preclinical study indicates that targeting the DCAF15/RBM39 axis with Indisulam modulates the differential splicing of KRAS 4A/4B, which is essential for the development of KRAS mutant LUAD. Indisulam significantly inhibits initiation of LUAD by suppressing stem cell-like properties, positioning it as a cutting-edge therapeutic strategy against this malignancy [76].
Furthermore, small-molecule inhibitors have been actively pursued to target specific splicing isoforms, with one of the most representative examples being MET-ΔEx14 [98]. Small molecule MET TKIs such as crizotinib and cabozantinib, demonstrate potent activity against MET amplification thorough lower nanomolar potency in MET-ΔEx14 mutated cell lines [174]. A study conducted by Memorial Sloan Kettering Cancer Center enrolled Stage IV LUAD patients harboring exon 14 alterations, who exhibited significant radiographic responses to treatment with crizotinib and cabozantinib [175]. These results encourage prospective identification of MET exon 14 splicing alteration in LUAD patients.
In conclusion, small-molecule inhibitors directly disrupt the complex splicing machinery or isoforms within cancer cells, causing the mis-splicing of numerous RNA transcripts. This alteration can lead to selective lethality for tumor cells that rely on aberrant AS for their survival, underscoring the potential of splicing modulation in LUAD therapy.

5.2. RNA-Targeted Therapies

RNA-targeted therapies, including RNA interference (RNAi) and antisense oligonucleotides (ASOs) [176], aim to silence critical targets, such as KRAS and EGFR, by redirecting AS in LUAD metastasis [177,178].
A complementary RNAi strategy targeting KRAS has identified TBK1-activated NF-κB anti-apoptotic signals that involve c-Rel and the splicing of Bcl-xL. Both c-Rel and spliced Bcl-xL are essential for LUAD proliferation and thus provide mechanistic insights into this synthetic lethal interaction [179]. Furthermore, AZD4785, a high-affinity ASO targeting KRAS, can potently and selectively deplete cellular KRAS mRNA and protein resultant in the inhibition of downstream effector pathways and antiproliferative effects, especially in KRAS-mutant LUAD cells [52].
ASOs targeting EGFR are designed to induce exon skipping within the transmembrane or tyrosine kinase domains that can modulate the production of specific splice variants associated with LUAD proliferation and migration [180,181]. Wang et al. synthesized and introduced the ASO VF/S/A@CaP, which sensitizes osimertinib-resistant LUAD cells, characterized by a high mesenchymal state, to ferroptosis [182]. Recent findings indicate that EGFR-targeting ASO exhibits a more potent anti-cancer effect than traditional TKIs in NSCLC with EGFR mutations, including L858R and T790M, and can effectively suppress TKI-resistant patient-derived tumors [54]. By influencing EGFR splicing, these ASOs aim to alter receptor expression or function in the cell membrane, thus potentially reducing tumor growth and improving responses to existing therapies.
Recent studies showed that the strategy of splice-switching antisense oligonucleotides (SSOs) has high therapeutic efficiency [183,184]. SSOs hybridize pre-mRNA sequences and modulate AS by exposing or masking the exon–intron boundaries. The administration of nanoparticles containing Bcl-x SSOs, which interfere with the splicing pattern of Bcl-x pre-mRNA, exhibit in vivo anti-tumor activity in LUAD models [185]. A schematic presentation of RNA-targeted therapies is illustrated in Figure 5.

5.3. Emerging Gene Therapies

RNA targeted therapies have emerged as a formidable strategy for silencing specific oncogenes in LUAD. Despite its promise, the clinical applicability of them is hindered by the transient nature of its effects and the obstacles associated with effective delivery mechanisms [186,187]. To overcome these limitations, researchers have introduced innovative gene therapies utilizing engineered oncolytic viruses (OVs) that preferentially enter cancer cells by exploiting receptors that are overexpressed in tumors, while sparing healthy tissue and leveraging their natural properties to proliferate [188]. These OVs not only induce direct lysis of tumor cells but also provoke systemic anti-tumor immune responses through the release of tumor-associated antigens [189]. Notable OVs include Coxsackievirus B3 [190], Reolysin [191], and NDV-rL-RVG [192], each demonstrating promising results in LUAD treatment.
Recently, Yun et al. introduced an oncolytic adenovirus (OAs)-based short hairpin RNA (shRNA) expression system, specifically targeting vascular endothelial growth factor (VEGF) through the Ad-ΔE1-shVEGF construct. Among the seven isoforms of VEGF, the VEGF-121 sequence is conserved across all isoforms, allowing VEGF-121-specific shRNA to effectively degrade all seven VEGF variants. To achieve efficient and sustained VEGF silencing, the study incorporated the shVEGF construct under the regulation of a U6 promoter within the E3 region of a double mutant OAs characterized by E1A and E1B deletions. The modified OAs exhibited substantial tumor-suppressive effects in LUAD A549 cell lines in vitro [193]. Furthermore, research suggests that the inhibition of the specific isoform VEGF-165, a prominent splicing isoform of VEGF, can elicit long-term anti-tumor responses when integrated into standard oncolytic virus therapy regimens [194]. Such targeted inhibition tailors the complexities inherent in AS, which generates multiple VEGF isoforms with distinct biological functions. This underscores a pivotal intersection between oncolytic virotherapy and splicing modulation, offering a novel avenue for enhancing therapeutic efficacy against LUAD.
Furthermore, CRISPR/Cas9 gene editing provides a complementary strategy for gene therapies through its capacity for precise genomic alterations [195]. This technology allows for the disruption of oncogenes or the restoration of tumor suppressor genes in LUAD [196]. Recent studies utilized CRISPR/Cas9 to knock out the m6A-modified SNRPA, which further induced exon 8 skipping of the ERCC1 gene, thereby reversing the SNRPA-mediated cisplatin resistance observed in LUAD [197]. Additionally, researchers have developed pgFARM (paired guide RNAs for alternative exon removal), a CRISPR/Cas9-based methodology aimed at manipulating isoforms without necessitating gene inactivation using LUAD xenografts [198]. This approach employs a pooled screening technique to assess the functional significance of “poison” cassette exons, which disrupt the reading frames of their host genes while demonstrating a high degree of evolutionary conservation.
The emergence of gene therapy technologies, encompassing OVs therapy and CRISPR/Cas9 gene editing, offers promising new strategies for the treatment of LUAD. Continued exploration and research in this area are essential for unlocking the full potential of these strategies by investigating the intricate mechanisms of AS in LUAD.

5.4. Combining Splicing-Targeted Strategies with Standard Treatments

In recent decades, treatment of LUAD has experienced significant advancements in targeted therapies, particularly through the use of TKIs to target specific driver mutations such as EGFR, ALK, and ROS1 [199,200,201]. However, resistance to these therapies due to a combination of genetic alterations, cellular modifications, and microenvironmental factors remains a prevalent challenge [202,203]. Intriguingly, recent research has revealed that the dysregulated splicing of cancer-associated genes, such as HER2 [204,205] and BIM [206,207], can be implicated in mediating acquired TKI resistance in LUAD.
Vorinostat is an FDA-approved, orally taken, histone deacetylase inhibitor drug for primary T-cell lymphoma [208]. Required resistance to apoptosis induced by TKI osimertinib in LUAD can be overcome by the combined use of vorinostat, which affects the AS of BIM mRNA and increases the expression of active BIM protein [209]. This offers an alternative off-label use of vorinostat. Moreover, a clinical trial has indicated that the combination of vorinostat and gefitinib can be effective for patients with BIM-deletion polymorphisms and EGFR mutations [210].
Recent studies have shown that the co-occurring deficiency of the splicing factor RBM10 in EGFR-mutant LUAD can contribute to EGFR inhibitor osimertinib resistance [211,212]. This resistance is associated with a decreased apoptotic response resulting from splicing alterations in Bcl-xL, which is intrinsically modulated by RBM10 [213]. Importantly, researchers have found that the combination of a Bcl-xL-specific ASO and osimertinib could effectively overcome this resistance, leading to the enhanced killing of LUAD [62].
Targeting aberrant splicing in conjunction with conventional treatment options, such as TKIs, represents a promising strategy to combat LUAD. By addressing the AS-derived molecular vulnerabilities, clinicians aim to enhance therapeutic efficacy to further improve LUAD patient outcomes.

5.5. Upcoming AS-Based Immunotherapies

Immunotherapy has marked a significant advancement in cancer treatment by harnessing the immune system to recognize and combat cancer cells, including LUAD [214,215]. A noteworthy strategy in this field is adoptive cell therapy using CAR T cells, which is one of the antibody-based anti-tumor therapies and considered the most promising personalized immunotherapeutic treatment [216]. A notable success in this area is the CD19-targeted CAR T cell therapy, which has been approved for the treatment of relapsed or refractory acute lymphoblastic leukemia [217,218]. Although CAR T cell therapy has achieved remarkable success in treating hematological malignancies, efforts are underway to extend its application to solid tumors. Research is rapidly advancing to assess the feasibility and effectiveness of CAR T cell therapy for LUAD [219]. The identification of optimal tumor antigens is crucial for developing targeted CAR T immunotherapies [18,216].
Alternative mRNA processing in LUAD has been shown to alter proteomic diversity, providing an extensive but largely unexplored repertoire of novel immunotherapy targets [220,221]. Recent studies highlight the utility of a novel system called isoform peptides from RNA splicing for immunotherapy target screening (IRIS) [17]. IRIS utilizes a computational platform that is capable of identifying AS-derived tumor antigens, thereby broadening the scope of cancer immunotherapies. Among those upcoming tumor-specific targets, the CD44 AS isoform CD44v6 has been implicated in LUAD tumorigenesis, invasion, and metastasis [57,222]. Anti-CD44v6 CAR T cells have demonstrated functional activation in vitro, with the ability to infiltrate tumors, proliferate, inhibit LUAD growth in vivo, and improve overall survival in mouse models [58]. The interplay between LUAD cells and anti-CD44v6 CAR-engineered T cells is illustrated in Figure 6.
A substantial class of splicing-derived tumor-specific neojunctions highlights the potential to utilize AS-based proteins for diagnosis, prognosis, and therapeutic interventions [21]. As research advances, integrating these insights into immunotherapy strategies offers promises for developing more effective and personalized anti-LUAD treatments.

6. Conclusions

This review presents the multifaceted role of AS in the pathogenesis and treatment of LUAD. High-throughput screening methodologies, coupled with advanced bioinformatics analysis, are crucial for the comprehensive identification of novel AS events and associated biomarkers. This integration enables more precise patient stratification and the selection of appropriate targeted therapies, ultimately leading to improved treatment outcomes. Simultaneously, a deeper understanding of the complex regulatory mechanisms governing AS in LUAD is paramount, which necessitates the exploration of the intricate interplay between splicing machinery and the LUAD microenvironment. Integrating emerging technologies such as scRNA-seq and spatial transcriptomics enhances the elucidation of AS expression dynamics and spatial interactions.
Emerging therapeutic strategies aim to directly target AS machineries or specific aberrant splicing events. Small-molecule inhibitors and RNA-targeted therapies offer novel approaches to correct aberrant splicing and potentially overcome therapeutic resistance. The advent of gene therapy technologies, including engineered OVs and CRISPR/Cas9 gene editing, enriches the landscape of AS-modulating therapies for LUAD. Combinatory therapies, including the refinement and optimization of existing technologies, such as integrating AS-targeted ASOs with TKIs for patients harboring EGFR mutation, are able to enhance specificity and efficacy in LUAD while mitigating off-target effects. Furthermore, rigorous preclinical evaluations would prioritize the efficacy and safety of new drugs, and the addition of robust AS-derived biomarkers would further enable identification of LUAD patients who would most likely benefit from AS-targeted therapies.
In sum, the dynamic interaction between AS and the immune microenvironment plays an essential role in LUAD. The concerted effort to identify and characterize novel AS-derived neoantigens represents an unprecedented opportunity to advance cancer immunotherapies, such as CAR T cell therapy. The transition from basic research findings to clinical treatments allows us to bridge the gap from bench to bedside, and this hinges on continued AS-based investigations into LUAD.

Author Contributions

Conceptualization, C.S.-H.N., N.W. and W.L.; methodology, W.L.; software, W.L.; validation, M.X. and W.L.; formal analysis, W.L.; investigation, M.X. and W.L.; resources, W.L.; data curation, M.X. and W.L.; writing—original draft preparation, W.L.; writing—review and editing, C.S.-H.N., N.W. and M.X.; visualization, W.L.; supervision, C.S.-H.N. and N.W.; project administration, C.S.-H.N. and N.W.; funding acquisition, N.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors wish to acknowledge the support from the Hong Kong Research Grants Council Area Excellence Scheme (Ref. AoE/M-401/20) and the Research Impact Fund (Ref. R4017-18).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Siegel, R.L.; Giaquinto, A.N.; Jemal, A. Cancer statistics, 2024. CA Cancer J. Clin. 2024, 74, 12–49. [Google Scholar] [CrossRef]
  2. Herbst, R.S.; Heymach, J.V.; Lippman, S.M. Lung cancer. N. Engl. J. Med. 2008, 359, 1367–1380. [Google Scholar] [CrossRef]
  3. Chen, Z.; Fillmore, C.M.; Hammerman, P.S.; Kim, C.F.; Wong, K.K. Non-small-cell lung cancers: A heterogeneous set of diseases. Nat. Rev. Cancer 2014, 14, 535–546. [Google Scholar] [CrossRef] [PubMed]
  4. Han, G.; Sinjab, A.; Rahal, Z.; Lynch, A.M.; Treekitkarnmongkol, W.; Liu, Y.; Serrano, A.G.; Feng, J.; Liang, K.; Khan, K.; et al. An atlas of epithelial cell states and plasticity in lung adenocarcinoma. Nature 2024, 627, 656–663. [Google Scholar] [CrossRef] [PubMed]
  5. Howlader, N.; Forjaz, G.; Mooradian, M.J.; Meza, R.; Kong, C.Y.; Cronin, K.A.; Mariotto, A.B.; Lowy, D.R.; Feuer, E.J. The Effect of Advances in Lung-Cancer Treatment on Population Mortality. N. Engl. J. Med. 2020, 383, 640–649. [Google Scholar] [CrossRef] [PubMed]
  6. Collisson, E.A.; Campbell, J.D.; Brooks, A.N.; Berger, A.H.; Lee, W.; Chmielecki, J.; Beer, D.G.; Cope, L.; Creighton, C.J.; Danilova, L.; et al. Comprehensive molecular profiling of lung adenocarcinoma. Nature 2014, 511, 543–550. [Google Scholar] [CrossRef]
  7. Chen, H.; Gao, F.; He, M.; Ding, X.F.; Wong, A.M.; Sze, S.C.; Yu, A.C.; Sun, T.; Chan, A.W.; Wang, X.; et al. Long-Read RNA Sequencing Identifies Alternative Splice Variants in Hepatocellular Carcinoma and Tumor-Specific Isoforms. Hepatology 2019, 70, 1011–1025. [Google Scholar] [CrossRef]
  8. Whibley, A.; Kelley, J.L.; Narum, S.R. The changing face of genome assemblies: Guidance on achieving high-quality reference genomes. Mol. Ecol. Resour. 2021, 21, 641–652. [Google Scholar] [CrossRef]
  9. Sakamoto, Y.; Miyake, S.; Oka, M.; Kanai, A.; Kawai, Y.; Nagasawa, S.; Shiraishi, Y.; Tokunaga, K.; Kohno, T.; Seki, M.; et al. Phasing analysis of lung cancer genomes using a long read sequencer. Nat. Commun. 2022, 13, 3464. [Google Scholar] [CrossRef]
  10. Logsdon, G.A.; Vollger, M.R.; Eichler, E.E. Long-read human genome sequencing and its applications. Nat. Rev. Genet. 2020, 21, 597–614. [Google Scholar] [CrossRef]
  11. Wright, C.J.; Smith, C.W.J.; Jiggins, C.D. Alternative splicing as a source of phenotypic diversity. Nat. Rev. Genet. 2022, 23, 697–710. [Google Scholar] [CrossRef] [PubMed]
  12. Marasco, L.E.; Kornblihtt, A.R. The physiology of alternative splicing. Nat. Rev. Mol. Cell. Biol. 2023, 24, 242–254. [Google Scholar] [CrossRef]
  13. Wang, B.D.; Lee, N.H. Aberrant RNA Splicing in Cancer and Drug Resistance. Cancers 2018, 10, 458. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, Y.; Qian, J.; Gu, C.; Yang, Y. Alternative splicing and cancer: A systematic review. Signal Transduct. Target. Ther. 2021, 6, 78. [Google Scholar] [CrossRef] [PubMed]
  15. Wu, Q.; Feng, L.; Wang, Y.; Mao, Y.; Di, X.; Zhang, K.; Cheng, S.; Xiao, T. Multi-omics analysis reveals RNA splicing alterations and their biological and clinical implications in lung adenocarcinoma. Signal Transduct. Target. Ther. 2022, 7, 270. [Google Scholar] [CrossRef]
  16. Robinson, T.J.; Freedman, J.A.; Al Abo, M.; Deveaux, A.E.; LaCroix, B.; Patierno, B.M.; George, D.J.; Patierno, S.R. Alternative RNA Splicing as a Potential Major Source of Untapped Molecular Targets in Precision Oncology and Cancer Disparities. Clin. Cancer Res. 2019, 25, 2963–2968. [Google Scholar] [CrossRef]
  17. Pan, Y.; Phillips, J.W.; Zhang, B.D.; Noguchi, M.; Kutschera, E.; McLaughlin, J.; Nesterenko, P.A.; Mao, Z.; Bangayan, N.J.; Wang, R.; et al. IRIS: Discovery of cancer immunotherapy targets arising from pre-mRNA alternative splicing. Proc. Natl. Acad. Sci. USA 2023, 120, e2221116120. [Google Scholar] [CrossRef]
  18. Pan, Y.; Kadash-Edmondson, K.E.; Wang, R.; Phillips, J.; Liu, S.; Ribas, A.; Aplenc, R.; Witte, O.N.; Xing, Y. RNA Dysregulation: An Expanding Source of Cancer Immunotherapy Targets. Trends Pharmacol. Sci. 2021, 42, 268–282. [Google Scholar] [CrossRef]
  19. Wahl, M.C.; Will, C.L.; Lührmann, R. The spliceosome: Design principles of a dynamic RNP machine. Cell 2009, 136, 701–718. [Google Scholar] [CrossRef]
  20. Kahles, A.; Lehmann, K.V.; Toussaint, N.C.; Hüser, M.; Stark, S.G.; Sachsenberg, T.; Stegle, O.; Kohlbacher, O.; Sander, C.; Rätsch, G. Comprehensive Analysis of Alternative Splicing Across Tumors from 8,705 Patients. Cancer Cell 2018, 34, 211–224.e216. [Google Scholar] [CrossRef]
  21. Hoyos, L.E.; Abdel-Wahab, O. Cancer-Specific Splicing Changes and the Potential for Splicing-Derived Neoantigens. Cancer Cell 2018, 34, 181–183. [Google Scholar] [CrossRef] [PubMed]
  22. Ule, J.; Blencowe, B.J. Alternative Splicing Regulatory Networks: Functions, Mechanisms, and Evolution. Mol. Cell 2019, 76, 329–345. [Google Scholar] [CrossRef] [PubMed]
  23. Voldborg, B.R.; Damstrup, L.; Spang-Thomsen, M.; Poulsen, H.S. Epidermal growth factor receptor (EGFR) and EGFR mutations, function and possible role in clinical trials. Ann. Oncol. 1997, 8, 1197–1206. [Google Scholar] [CrossRef] [PubMed]
  24. El Zarif, T.; Meador, C.B.; Qiu, X.; Seo, J.H.; Davidsohn, M.P.; Savignano, H.; Lakshminarayanan, G.; McClure, H.M.; Canniff, J.; Fortunato, B.; et al. Detecting Small Cell Transformation in Patients with Advanced EGFR Mutant Lung Adenocarcinoma through Epigenomic cfDNA Profiling. Clin. Cancer Res. 2024, 30, 3798–3811. [Google Scholar] [CrossRef]
  25. Abramson, J.; Adler, J.; Dunger, J.; Evans, R.; Green, T.; Pritzel, A.; Ronneberger, O.; Willmore, L.; Ballard, A.J.; Bambrick, J.; et al. Accurate structure prediction of biomolecular interactions with AlphaFold 3. Nature 2024, 630, 493–500. [Google Scholar] [CrossRef]
  26. Ryan, M.; Wong, W.C.; Brown, R.; Akbani, R.; Su, X.; Broom, B.; Melott, J.; Weinstein, J. TCGASpliceSeq a compendium of alternative mRNA splicing in cancer. Nucleic Acids Res. 2016, 44, D1018–D1022. [Google Scholar] [CrossRef]
  27. Wang, Y.; Guo, E.; Zou, M.; Lv, C.; Cui, Y.; Zhai, S.; Sang, S.; Xiong, K.; Yang, X.; Zhuang, S.; et al. Unraveling immune heterogeneity across pan-cancer and deep insights in lung adenocarcinoma based on alternative splicing. J. Leukoc. Biol. 2025, 117, qiae104. [Google Scholar] [CrossRef]
  28. Liu, J.; Wang, Y.; Zhao, X.; Wang, K.; Wang, C.; Du, J. Prognostic alternative splicing events related splicing factors define the tumor microenvironment and pharmacogenomic landscape in lung adenocarcinoma. Aging 2022, 14, 6689–6715. [Google Scholar] [CrossRef]
  29. Shao, M.M.; Zhai, K.; Huang, Z.Y.; Yi, F.S.; Zheng, S.C.; Liu, Y.L.; Qiao, X.; Chen, Q.Y.; Wang, Z.; Shi, H.Z. Characterization of the alternative splicing landscape in lung adenocarcinoma reveals novel prognosis signature associated with B cells. PLoS ONE 2023, 18, e0279018. [Google Scholar] [CrossRef]
  30. Stark, R.; Grzelak, M.; Hadfield, J. RNA sequencing: The teenage years. Nat. Rev. Genet. 2019, 20, 631–656. [Google Scholar] [CrossRef]
  31. Trapnell, C.; Williams, B.A.; Pertea, G.; Mortazavi, A.; Kwan, G.; van Baren, M.J.; Salzberg, S.L.; Wold, B.J.; Pachter, L. Transcript assembly and quantification by RNA-Seq reveals unannotated transcripts and isoform switching during cell differentiation. Nat. Biotechnol. 2010, 28, 511–515. [Google Scholar] [CrossRef] [PubMed]
  32. Salmen, F.; De Jonghe, J.; Kaminski, T.S.; Alemany, A.; Parada, G.E.; Verity-Legg, J.; Yanagida, A.; Kohler, T.N.; Battich, N.; van den Brekel, F.; et al. High-throughput total RNA sequencing in single cells using VASA-seq. Nat. Biotechnol. 2022, 40, 1780–1793. [Google Scholar] [CrossRef]
  33. Salcher, S.; Sturm, G.; Horvath, L.; Untergasser, G.; Kuempers, C.; Fotakis, G.; Panizzolo, E.; Martowicz, A.; Trebo, M.; Pall, G.; et al. High-resolution single-cell atlas reveals diversity and plasticity of tissue-resident neutrophils in non-small cell lung cancer. Cancer Cell 2022, 40, 1503–1520.e1508. [Google Scholar] [CrossRef]
  34. Marjanovic, N.D.; Hofree, M.; Chan, J.E.; Canner, D.; Wu, K.; Trakala, M.; Hartmann, G.G.; Smith, O.C.; Kim, J.Y.; Evans, K.V.; et al. Emergence of a High-Plasticity Cell State during Lung Cancer Evolution. Cancer Cell 2020, 38, 229–246.e213. [Google Scholar] [CrossRef]
  35. Wenger, A.M.; Peluso, P.; Rowell, W.J.; Chang, P.C.; Hall, R.J.; Concepcion, G.T.; Ebler, J.; Fungtammasan, A.; Kolesnikov, A.; Olson, N.D.; et al. Accurate circular consensus long-read sequencing improves variant detection and assembly of a human genome. Nat. Biotechnol. 2019, 37, 1155–1162. [Google Scholar] [CrossRef] [PubMed]
  36. Hagemann-Jensen, M.; Ziegenhain, C.; Chen, P.; Ramsköld, D.; Hendriks, G.J.; Larsson, A.J.M.; Faridani, O.R.; Sandberg, R. Single-cell RNA counting at allele and isoform resolution using Smart-seq3. Nat. Biotechnol. 2020, 38, 708–714. [Google Scholar] [CrossRef]
  37. Penter, L.; Borji, M.; Nagler, A.; Lyu, H.; Lu, W.S.; Cieri, N.; Maurer, K.; Oliveira, G.; Al’Khafaji, A.M.; Garimella, K.V.; et al. Integrative genotyping of cancer and immune phenotypes by long-read sequencing. Nat. Commun. 2024, 15, 32. [Google Scholar] [CrossRef] [PubMed]
  38. Dondi, A.; Lischetti, U.; Jacob, F.; Singer, F.; Borgsmüller, N.; Coelho, R.; Heinzelmann-Schwarz, V.; Beisel, C.; Beerenwinkel, N. Detection of isoforms and genomic alterations by high-throughput full-length single-cell RNA sequencing in ovarian cancer. Nat. Commun. 2023, 14, 7780. [Google Scholar] [CrossRef]
  39. Wang, Y.; Ding, Y.; Liu, S.; Wang, C.; Zhang, E.; Chen, C.; Zhu, M.; Zhang, J.; Zhu, C.; Ji, M.; et al. Integrative splicing-quantitative-trait-locus analysis reveals risk loci for non-small-cell lung cancer. Am. J. Hum. Genet. 2023, 110, 1574–1589. [Google Scholar] [CrossRef]
  40. Kim, N.; Kim, H.K.; Lee, K.; Hong, Y.; Cho, J.H.; Choi, J.W.; Lee, J.I.; Suh, Y.L.; Ku, B.M.; Eum, H.H.; et al. Single-cell RNA sequencing demonstrates the molecular and cellular reprogramming of metastatic lung adenocarcinoma. Nat. Commun. 2020, 11, 2285. [Google Scholar] [CrossRef]
  41. De Zuani, M.; Xue, H.; Park, J.S.; Dentro, S.C.; Seferbekova, Z.; Tessier, J.; Curras-Alonso, S.; Hadjipanayis, A.; Athanasiadis, E.I.; Gerstung, M.; et al. Single-cell and spatial transcriptomics analysis of non-small cell lung cancer. Nat. Commun. 2024, 15, 4388. [Google Scholar] [CrossRef] [PubMed]
  42. Belchikov, N.; Hsu, J.; Li, X.J.; Jarroux, J.; Hu, W.; Joglekar, A.; Tilgner, H.U. Understanding isoform expression by pairing long-read sequencing with single-cell and spatial transcriptomics. Genome Res. 2024, 34, 1735–1746. [Google Scholar] [CrossRef] [PubMed]
  43. Lebrigand, K.; Bergenstråhle, J.; Thrane, K.; Mollbrink, A.; Meletis, K.; Barbry, P.; Waldmann, R.; Lundeberg, J. The spatial landscape of gene expression isoforms in tissue sections. Nucleic Acids Res. 2023, 51, e47. [Google Scholar] [CrossRef]
  44. Paavola, K.J.; Roda, J.M.; Lin, V.Y.; Chen, P.; O’Hollaren, K.P.; Ventura, R.; Crawley, S.C.; Li, B.; Chen, H.H.; Malmersjö, S.; et al. The Fibronectin-ILT3 Interaction Functions as a Stromal Checkpoint that Suppresses Myeloid Cells. Cancer Immunol. Res. 2021, 9, 1283–1297. [Google Scholar] [CrossRef]
  45. Di Modugno, F.; Di Carlo, A.; Spada, S.; Palermo, B.; D’Ambrosio, L.; D’Andrea, D.; Morello, G.; Belmonte, B.; Sperduti, I.; Balzano, V.; et al. Tumoral and stromal hMENA isoforms impact tertiary lymphoid structure localization in lung cancer and predict immune checkpoint blockade response in patients with cancer. EBioMedicine 2024, 101, 105003. [Google Scholar] [CrossRef]
  46. Desai, T.J.; Brownfield, D.G.; Krasnow, M.A. Alveolar progenitor and stem cells in lung development, renewal and cancer. Nature 2014, 507, 190–194. [Google Scholar] [CrossRef]
  47. Zacharias, W.J.; Frank, D.B.; Zepp, J.A.; Morley, M.P.; Alkhaleel, F.A.; Kong, J.; Zhou, S.; Cantu, E.; Morrisey, E.E. Regeneration of the lung alveolus by an evolutionarily conserved epithelial progenitor. Nature 2018, 555, 251–255. [Google Scholar] [CrossRef]
  48. Treutlein, B.; Brownfield, D.G.; Wu, A.R.; Neff, N.F.; Mantalas, G.L.; Espinoza, F.H.; Desai, T.J.; Krasnow, M.A.; Quake, S.R. Reconstructing lineage hierarchies of the distal lung epithelium using single-cell RNA-seq. Nature 2014, 509, 371–375. [Google Scholar] [CrossRef] [PubMed]
  49. Sutherland, K.D.; Song, J.Y.; Kwon, M.C.; Proost, N.; Zevenhoven, J.; Berns, A. Multiple cells-of-origin of mutant K-Ras-induced mouse lung adenocarcinoma. Proc. Natl. Acad. Sci. USA 2014, 111, 4952–4957. [Google Scholar] [CrossRef]
  50. Unni, A.M.; Lockwood, W.W.; Zejnullahu, K.; Lee-Lin, S.Q.; Varmus, H. Evidence that synthetic lethality underlies the mutual exclusivity of oncogenic KRAS and EGFR mutations in lung adenocarcinoma. eLife 2015, 4, e06907. [Google Scholar] [CrossRef]
  51. Salmón, M.; Paniagua, G.; Lechuga, C.G.; Fernández-García, F.; Zarzuela, E.; Álvarez-Díaz, R.; Musteanu, M.; Guerra, C.; Caleiras, E.; Muñoz, J.; et al. KRAS4A induces metastatic lung adenocarcinomas in vivo in the absence of the KRAS4B isoform. Proc. Natl. Acad. Sci. USA 2021, 118, e2023112118. [Google Scholar] [CrossRef] [PubMed]
  52. Ross, S.J.; Revenko, A.S.; Hanson, L.L.; Ellston, R.; Staniszewska, A.; Whalley, N.; Pandey, S.K.; Revill, M.; Rooney, C.; Buckett, L.K.; et al. Targeting KRAS-dependent tumors with AZD4785, a high-affinity therapeutic antisense oligonucleotide inhibitor of KRAS. Sci. Transl. Med. 2017, 9, eaal5253. [Google Scholar] [CrossRef] [PubMed]
  53. Ji, H.; Zhao, X.; Yuza, Y.; Shimamura, T.; Li, D.; Protopopov, A.; Jung, B.L.; McNamara, K.; Xia, H.; Glatt, K.A.; et al. Epidermal growth factor receptor variant III mutations in lung tumorigenesis and sensitivity to tyrosine kinase inhibitors. Proc. Natl. Acad. Sci. USA 2006, 103, 7817–7822. [Google Scholar] [CrossRef]
  54. Tran, T.T.T.; Phung, C.D.; Yeo, B.Z.J.; Prajogo, R.C.; Jayasinghe, M.K.; Yuan, J.; Tan, D.S.W.; Yeo, E.Y.M.; Goh, B.C.; Tam, W.L.; et al. Customised design of antisense oligonucleotides targeting EGFR driver mutants for personalised treatment of non-small cell lung cancer. EBioMedicine 2024, 108, 105356. [Google Scholar] [CrossRef] [PubMed]
  55. Van Der Steen, N.; Giovannetti, E.; Pauwels, P.; Peters, G.J.; Hong, D.S.; Cappuzzo, F.; Hirsch, F.R.; Rolfo, C. cMET Exon 14 Skipping: From the Structure to the Clinic. J. Thorac. Oncol. 2016, 11, 1423–1432. [Google Scholar] [CrossRef]
  56. Frampton, G.M.; Ali, S.M.; Rosenzweig, M.; Chmielecki, J.; Lu, X.; Bauer, T.M.; Akimov, M.; Bufill, J.A.; Lee, C.; Jentz, D.; 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. [Google Scholar] [CrossRef]
  57. Xie, Z.; Gao, Y.; Ho, C.; Li, L.; Jin, C.; Wang, X.; Zou, C.; Mao, Y.; Wang, X.; Li, Q.; et al. Exosome-delivered CD44v6/C1QBP complex drives pancreatic cancer liver metastasis by promoting fibrotic liver microenvironment. Gut 2022, 71, 568–579. [Google Scholar] [CrossRef] [PubMed]
  58. Porcellini, S.; Asperti, C.; Corna, S.; Cicoria, E.; Valtolina, V.; Stornaiuolo, A.; Valentinis, B.; Bordignon, C.; Traversari, C. CAR T Cells Redirected to CD44v6 Control Tumor Growth in Lung and Ovary Adenocarcinoma Bearing Mice. Front. Immunol. 2020, 11, 99. [Google Scholar] [CrossRef]
  59. Lin, X.; Kang, K.; Chen, P.; Zeng, Z.; Li, G.; Xiong, W.; Yi, M.; Xiang, B. Regulatory mechanisms of PD-1/PD-L1 in cancers. Mol. Cancer 2024, 23, 108. [Google Scholar] [CrossRef]
  60. Sun, J.; Bai, J.; Jiang, T.; Gao, Y.; Hua, Y. Modulation of PDCD1 exon 3 splicing. RNA Biol. 2019, 16, 1794–1805. [Google Scholar] [CrossRef]
  61. Bodac, A.; Mayet, A.; Rana, S.; Pascual, J.; Bowler, A.D.; Roh, V.; Fournier, N.; Craciun, L.; Demetter, P.; Radtke, F.; et al. Bcl-xL targeting eliminates ageing tumor-promoting neutrophils and inhibits lung tumor growth. EMBO Mol. Med. 2024, 16, 158–184. [Google Scholar] [CrossRef] [PubMed]
  62. Bertino, E.M.; Gentzler, R.D.; Clifford, S.; Kolesar, J.; Muzikansky, A.; Haura, E.B.; Piotrowska, Z.; Camidge, D.R.; Stinchcombe, T.E.; Hann, C.; et al. Phase IB Study of Osimertinib in Combination with Navitoclax in EGFR-mutant NSCLC Following Resistance to Initial EGFR Therapy (ETCTN 9903). Clin. Cancer Res. 2021, 27, 1604–1611. [Google Scholar] [CrossRef]
  63. Xu, N.; Ren, Y.; Bao, Y.; Shen, X.; Kang, J.; Wang, N.; Wang, Z.; Han, X.; Li, Z.; Zuo, J.; et al. PUF60 promotes cell cycle and lung cancer progression by regulating alternative splicing of CDC25C. Cell Rep. 2023, 42, 113041. [Google Scholar] [CrossRef] [PubMed]
  64. Qu, A.; Han, B.; Hua, M.; Wang, C.; Li, T. SF3B4 downregulation restrains lung adenocarcinoma tumorigenesis via 5′ alternative splicing of KAT2A. Sci. Rep. 2024, 14, 30. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, S.; Bao, Y.; Shen, X.; Pan, Y.; Sun, Y.; Xiao, M.; Chen, K.; Wei, H.; Zuo, J.; Saffen, D.; et al. RNA binding motif protein 10 suppresses lung cancer progression by controlling alternative splicing of eukaryotic translation initiation factor 4H. EBioMedicine 2020, 61, 103067. [Google Scholar] [CrossRef]
  66. Chen, Q.W.; Cai, Q.Q.; Yang, Y.; Dong, S.; Liu, Y.Y.; Chen, Z.Y.; Kang, C.L.; Qi, B.; Dong, Y.W.; Wu, W.; et al. LncRNA BC promotes lung adenocarcinoma progression by modulating IMPAD1 alternative splicing. Clin. Transl. Med. 2023, 13, e1129. [Google Scholar] [CrossRef]
  67. Klomp, J.A.; Klomp, J.E.; Stalnecker, C.A.; Bryant, K.L.; Edwards, A.C.; Drizyte-Miller, K.; Hibshman, P.S.; Diehl, J.N.; Lee, Y.S.; Morales, A.J.; et al. Defining the KRAS- and ERK-dependent transcriptome in KRAS-mutant cancers. Science 2024, 384, eadk0775. [Google Scholar] [CrossRef]
  68. Suda, K.; Tomizawa, K.; Mitsudomi, T. Biological and clinical significance of KRAS mutations in lung cancer: An oncogenic driver that contrasts with EGFR mutation. Cancer Metastasis Rev. 2010, 29, 49–60. [Google Scholar] [CrossRef]
  69. Prior, I.A.; Lewis, P.D.; Mattos, C. A comprehensive survey of Ras mutations in cancer. Cancer Res. 2012, 72, 2457–2467. [Google Scholar] [CrossRef]
  70. Collisson, E.A.; Trejo, C.L.; Silva, J.M.; Gu, S.; Korkola, J.E.; Heiser, L.M.; Charles, R.P.; Rabinovich, B.A.; Hann, B.; Dankort, D.; et al. A central role for RAF→MEK→ERK signaling in the genesis of pancreatic ductal adenocarcinoma. Cancer Discov. 2012, 2, 685–693. [Google Scholar] [CrossRef]
  71. Tsai, F.D.; Lopes, M.S.; Zhou, M.; Court, H.; Ponce, O.; Fiordalisi, J.J.; Gierut, J.J.; Cox, A.D.; Haigis, K.M.; Philips, M.R. K-Ras4A splice variant is widely expressed in cancer and uses a hybrid membrane-targeting motif. Proc. Natl. Acad. Sci. USA 2015, 112, 779–784. [Google Scholar] [CrossRef] [PubMed]
  72. Aran, V.; Masson Domingues, P.; Carvalho de Macedo, F.; Moreira de Sousa, C.A.; Caldas Montella, T.; de Souza Accioly, M.T.; Ferreira, C.G. 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. [Google Scholar] [CrossRef]
  73. Yang, I.S.; Kim, S. Isoform specific gene expression analysis of KRAS in the prognosis of lung adenocarcinoma patients. BMC Bioinformatics 2018, 19, 40. [Google Scholar] [CrossRef]
  74. Newlaczyl, A.U.; Coulson, J.M.; Prior, I.A. Quantification of spatiotemporal patterns of Ras isoform expression during development. Sci. Rep. 2017, 7, 41297. [Google Scholar] [CrossRef] [PubMed]
  75. Omerovic, J.; Hammond, D.E.; Clague, M.J.; Prior, I.A. Ras isoform abundance and signalling in human cancer cell lines. Oncogene 2008, 27, 2754–2762. [Google Scholar] [CrossRef]
  76. Chen, W.C.; To, M.D.; Westcott, P.M.K.; Delrosario, R.; Kim, I.J.; Philips, M.; Tran, Q.; Bollam, S.R.; Goodarzi, H.; Bayani, N.; et al. Targeting KRAS4A splicing through the RBM39/DCAF15 pathway inhibits cancer stem cells. Nat. Commun. 2021, 12, 4288. [Google Scholar] [CrossRef] [PubMed]
  77. Wang, H.; Lv, Q.; Xu, Y.; Cai, Z.; Zheng, J.; Cheng, X.; Dai, Y.; Jänne, P.A.; Ambrogio, C.; Köhler, J. An integrative pharmacogenomics analysis identifies therapeutic targets in KRAS-mutant lung cancer. EBioMedicine 2019, 49, 106–117. [Google Scholar] [CrossRef]
  78. Chiari, R.; Palladino, S.; Emili, R.; De Lisa, M.; Sarti, D.; Catalano, V.; Magnani, M.; Graziano, F.; Ruzzo, A. KRAS4A and KRAS4B in liquid biopsy of metastatic lung adenocarcinoma patients treated with Pembrolizumab or chemotherapy plus Pembrolizumab. Sci. Rep. 2023, 13, 21036. [Google Scholar] [CrossRef]
  79. Tomas, A.; Futter, C.E.; Eden, E.R. EGF receptor trafficking: Consequences for signaling and cancer. Trends Cell Biol. 2014, 24, 26–34. [Google Scholar] [CrossRef]
  80. Iamaroon, A.; Tait, B.; Diewert, V.M. Cell proliferation and expression of EGF, TGF-alpha, and EGF receptor in the developing primary palate. J. Dent. Res. 1996, 75, 1534–1539. [Google Scholar] [CrossRef]
  81. Ovrevik, J.; Refsnes, M.; Totlandsdal, A.I.; Holme, J.A.; Schwarze, P.E.; Låg, M. TACE/TGF-α/EGFR regulates CXCL8 in bronchial epithelial cells exposed to particulate matter components. Eur. Respir. J. 2011, 38, 1189–1199. [Google Scholar] [CrossRef]
  82. Wee, P.; Wang, Z. Epidermal Growth Factor Receptor Cell Proliferation Signaling Pathways. Cancers 2017, 9, 52. [Google Scholar] [CrossRef] [PubMed]
  83. Wang, R.Y.; Chen, L.; Chen, H.Y.; Hu, L.; Li, L.; Sun, H.Y.; Jiang, F.; Zhao, J.; Liu, G.M.; Tang, J.; et al. MUC15 inhibits dimerization of EGFR and PI3K-AKT signaling and is associated with aggressive hepatocellular carcinomas in patients. Gastroenterology 2013, 145, 1436–1448.e1–12. [Google Scholar] [CrossRef] [PubMed]
  84. Gala, K.; Chandarlapaty, S. Molecular pathways: HER3 targeted therapy. Clin. Cancer Res. 2014, 20, 1410–1416. [Google Scholar] [CrossRef] [PubMed]
  85. Nakakuki, T.; Yumoto, N.; Naka, T.; Shirouzu, M.; Yokoyama, S.; Hatakeyama, M. Topological analysis of MAPK cascade for kinetic ErbB signaling. PLoS ONE 2008, 3, e1782. [Google Scholar] [CrossRef]
  86. Zhang, R.; Fruhwirth, G.O.; Coban, O.; Barrett, J.E.; Burgoyne, T.; Lee, S.H.; Simonson, P.D.; Baday, M.; Kholodenko, B.N.; Futter, C.E.; et al. Probing the Heterogeneity of Protein Kinase Activation in Cells by Super-resolution Microscopy. ACS Nano 2017, 11, 249–257. [Google Scholar] [CrossRef]
  87. Ohtsuka, K.; Ohnishi, H.; Furuyashiki, G.; Nogami, H.; Koshiishi, Y.; Ooide, A.; Matsushima, S.; Watanabe, T.; Goya, T. Clinico-pathological and biological significance of tyrosine kinase domain gene mutations and overexpression of epidermal growth factor receptor for lung adenocarcinoma. J. Thorac. Oncol. 2006, 1, 787–795. [Google Scholar] [CrossRef]
  88. Castellanos, E.; Feld, E.; Horn, L. Driven by Mutations: The Predictive Value of Mutation Subtype in EGFR-Mutated Non-Small Cell Lung Cancer. J. Thorac. Oncol. 2017, 12, 612–623. [Google Scholar] [CrossRef]
  89. Scagliotti, G.V.; Selvaggi, G.; Novello, S.; Hirsch, F.R. The biology of epidermal growth factor receptor in lung cancer. Clin. Cancer Res. 2004, 10, 4227s–4232s. [Google Scholar] [CrossRef]
  90. Lee, V.H.; Tin, V.P.; Choy, T.S.; Lam, K.O.; Choi, C.W.; Chung, L.P.; Tsang, J.W.; Ho, P.P.; Leung, D.K.; Ma, E.S.; et al. Association of exon 19 and 21 EGFR mutation patterns with treatment outcome after first-line tyrosine kinase inhibitor in metastatic non-small-cell lung cancer. J. Thorac. Oncol. 2013, 8, 1148–1155. [Google Scholar] [CrossRef]
  91. Chung, K.P.; Wu, S.G.; Wu, J.Y.; Yang, J.C.; Yu, C.J.; Wei, P.F.; Shih, J.Y.; Yang, P.C. Clinical outcomes in non-small cell lung cancers harboring different exon 19 deletions in EGFR. Clin. Cancer Res. 2012, 18, 3470–3477. [Google Scholar] [CrossRef] [PubMed]
  92. Sasaki, H.; Kawano, O.; Endo, K.; Yukiue, H.; Yano, M.; Fujii, Y. EGFRvIII mutation in lung cancer correlates with increased EGFR copy number. Oncol. Rep. 2007, 17, 319–323. [Google Scholar] [CrossRef]
  93. Okamoto, I.; Kenyon, L.C.; Emlet, D.R.; Mori, T.; Sasaki, J.; Hirosako, S.; Ichikawa, Y.; Kishi, H.; Godwin, A.K.; Yoshioka, M.; et al. Expression of constitutively activated EGFRvIII in non-small cell lung cancer. Cancer Sci. 2003, 94, 50–56. [Google Scholar] [CrossRef]
  94. Piccione, E.C.; Lieu, T.J.; Gentile, C.F.; Williams, T.R.; Connolly, A.J.; Godwin, A.K.; Koong, A.C.; Wong, A.J. A novel epidermal growth factor receptor variant lacking multiple domains directly activates transcription and is overexpressed in tumors. Oncogene 2012, 31, 2953–2967. [Google Scholar] [CrossRef] [PubMed]
  95. Guo, R.; Luo, J.; Chang, J.; Rekhtman, N.; Arcila, M.; Drilon, A. MET-dependent solid tumours—Molecular diagnosis and targeted therapy. Nat. Rev. Clin. Oncol. 2020, 17, 569–587. [Google Scholar] [CrossRef]
  96. 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. S1), 188–202. [Google Scholar] [CrossRef] [PubMed]
  97. Recondo, G.; Che, J.; Jänne, P.A.; Awad, M.M. Targeting MET Dysregulation in Cancer. Cancer Discov. 2020, 10, 922–934. [Google Scholar] [CrossRef]
  98. Remon, J.; Hendriks, L.E.L.; Mountzios, G.; García-Campelo, R.; Saw, S.P.L.; Uprety, D.; Recondo, G.; Villacampa, G.; Reck, M. MET alterations in NSCLC-Current Perspectives and Future Challenges. J. Thorac. Oncol. 2023, 18, 419–435. [Google Scholar] [CrossRef]
  99. Schrock, A.B.; Frampton, G.M.; Suh, J.; Chalmers, Z.R.; Rosenzweig, M.; Erlich, R.L.; Halmos, B.; Goldman, J.; Forde, P.; Leuenberger, K.; et al. Characterization of 298 Patients with Lung Cancer Harboring MET Exon 14 Skipping Alterations. J. Thorac. Oncol. 2016, 11, 1493–1502. [Google Scholar] [CrossRef]
  100. Digumarthy, S.R.; Mendoza, D.P.; Zhang, E.W.; Lennerz, J.K.; Heist, R.S. Clinicopathologic and Imaging Features of Non-Small-Cell Lung Cancer with MET Exon 14 Skipping Mutations. Cancers 2019, 11, 2033. [Google Scholar] [CrossRef]
  101. Tan, A.C.; Tan, D.S.W. Targeted Therapies for Lung Cancer Patients With Oncogenic Driver Molecular Alterations. J. Clin. Oncol. 2022, 40, 611–625. [Google Scholar] [CrossRef] [PubMed]
  102. Awad, M.M.; Oxnard, G.R.; Jackman, D.M.; Savukoski, D.O.; Hall, D.; Shivdasani, P.; Heng, J.C.; Dahlberg, S.E.; Jänne, P.A.; Verma, S.; et al. MET Exon 14 Mutations in Non-Small-Cell Lung Cancer Are Associated With Advanced Age and Stage-Dependent MET Genomic Amplification and c-Met Overexpression. J. Clin. Oncol. 2016, 34, 721–730. [Google Scholar] [CrossRef] [PubMed]
  103. Lu, X.; Peled, N.; Greer, J.; Wu, W.; Choi, P.; Berger, A.H.; Wong, S.; Jen, K.Y.; Seo, Y.; Hann, B.; et al. MET Exon 14 Mutation Encodes an Actionable Therapeutic Target in Lung Adenocarcinoma. Cancer Res. 2017, 77, 4498–4505. [Google Scholar] [CrossRef]
  104. 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. [Google Scholar] [CrossRef] [PubMed]
  105. Schiering, N.; Knapp, S.; Marconi, M.; Flocco, M.M.; Cui, J.; Perego, R.; Rusconi, L.; Cristiani, C. 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. USA 2003, 100, 12654–12659. [Google Scholar] [CrossRef]
  106. Abella, J.V.; Peschard, P.; Naujokas, M.A.; Lin, T.; Saucier, C.; Urbé, S.; Park, M. Met/Hepatocyte growth factor receptor ubiquitination suppresses transformation and is required for Hrs phosphorylation. Mol. Cell. Biol. 2005, 25, 9632–9645. [Google Scholar] [CrossRef]
  107. Lee, G.D.; Lee, S.E.; Oh, D.Y.; Yu, D.B.; Jeong, H.M.; Kim, J.; Hong, S.; Jung, H.S.; Oh, E.; Song, J.Y.; et al. MET Exon 14 Skipping Mutations in Lung Adenocarcinoma: Clinicopathologic Implications and Prognostic Values. J. Thorac. Oncol. 2017, 12, 1233–1246. [Google Scholar] [CrossRef]
  108. Seo, J.S.; Ju, Y.S.; Lee, W.C.; Shin, J.Y.; Lee, J.K.; Bleazard, T.; Lee, J.; Jung, Y.J.; Kim, J.O.; Shin, J.Y.; et al. The transcriptional landscape and mutational profile of lung adenocarcinoma. Genome Res. 2012, 22, 2109–2119. [Google Scholar] [CrossRef]
  109. Lee, J.H.; Gao, C.F.; Lee, C.C.; Kim, M.D.; Vande Woude, G.F. An alternatively spliced form of Met receptor is tumorigenic. Exp. Mol. Med. 2006, 38, 565–573. [Google Scholar] [CrossRef]
  110. Park, S.; Koh, J.; Kim, D.W.; Kim, M.; Keam, B.; Kim, T.M.; Jeon, Y.K.; Chung, D.H.; Heo, D.S. MET amplification, protein expression, and mutations in pulmonary adenocarcinoma. Lung Cancer 2015, 90, 381–387. [Google Scholar] [CrossRef]
  111. Jenkins, R.W.; Oxnard, G.R.; Elkin, S.; Sullivan, E.K.; Carter, J.L.; Barbie, D.A. Response to Crizotinib in a Patient With Lung Adenocarcinoma Harboring a MET Splice Site Mutation. Clin. Lung Cancer 2015, 16, e101–e104. [Google Scholar] [CrossRef] [PubMed]
  112. Roth, K.G.; Mambetsariev, I.; Salgia, R. Prolonged survival and response to tepotinib in a non-small-cell lung cancer patient with brain metastases harboring MET exon 14 mutation: A research report. Cold Spring Harb. Mol. Case Stud. 2020, 6, a005785. [Google Scholar] [CrossRef] [PubMed]
  113. Li, R.; Liu, X.; Xu, Y.; Zhao, J.; Zhong, W.; Gao, X.; Chen, M.; Wang, M. Remarkable pathological response to neoadjuvant tepotinib in lung adenocarcinoma with MET exon 14 skipping mutation: A case report. Thorac. Cancer 2024, 15, 2339–2343. [Google Scholar] [CrossRef]
  114. Wolf, J.; Seto, T.; Han, J.Y.; Reguart, N.; Garon, E.B.; Groen, H.J.M.; Tan, D.S.W.; Hida, T.; de Jonge, M.; Orlov, S.V.; et al. Capmatinib in MET Exon 14-Mutated or MET-Amplified Non-Small-Cell Lung Cancer. N. Engl. J. Med. 2020, 383, 944–957. [Google Scholar] [CrossRef]
  115. Sun, X.; Li, C.W.; Wang, W.J.; Chen, M.K.; Li, H.; Lai, Y.J.; Hsu, J.L.; Koller, P.B.; Chan, L.C.; Lee, P.C.; et al. Inhibition of c-MET upregulates PD-L1 expression in lung adenocarcinoma. Am. J. Cancer Res. 2020, 10, 564–571. [Google Scholar] [PubMed]
  116. Sabari, J.K.; Leonardi, G.C.; Shu, C.A.; Umeton, R.; Montecalvo, J.; Ni, A.; Chen, R.; Dienstag, J.; Mrad, C.; Bergagnini, I.; et al. PD-L1 expression, tumor mutational burden, and response to immunotherapy in patients with MET exon 14 altered lung cancers. Ann. Oncol. 2018, 29, 2085–2091. [Google Scholar] [CrossRef]
  117. Bateman, A.; Martin, M.J.; Orchard, S.; Magrane, M.; Adesina, A.; Ahmad, S.; Bowler-Barnett, E.H.; Bye-A-Jee, H.; Carpentier, D.; Denny, P.; et al. UniProt: The Universal Protein Knowledgebase in 2025. Nucleic Acids Res. 2024, 52, D609–D617. [Google Scholar] [CrossRef]
  118. 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]
  119. Hassn Mesrati, M.; Syafruddin, S.E.; Mohtar, M.A.; Syahir, A. CD44: A Multifunctional Mediator of Cancer Progression. Biomolecules 2021, 11, 1850. [Google Scholar] [CrossRef]
  120. Weng, X.; Maxwell-Warburton, S.; Hasib, A.; Ma, L.; Kang, L. The membrane receptor CD44: Novel insights into metabolism. Trends Endocrinol. Metab. 2022, 33, 318–332. [Google Scholar] [CrossRef]
  121. Chen, C.; Zhao, S.; Karnad, A.; Freeman, J.W. The biology and role of CD44 in cancer progression: Therapeutic implications. J. Hematol. Oncol. 2018, 11, 64. [Google Scholar] [CrossRef] [PubMed]
  122. Zeilstra, J.; Joosten, S.P.; van Andel, H.; Tolg, C.; Berns, A.; Snoek, M.; van de Wetering, M.; Spaargaren, M.; Clevers, H.; Pals, S.T. Stem cell CD44v isoforms promote intestinal cancer formation in Apc(min) mice downstream of Wnt signaling. Oncogene 2014, 33, 665–670. [Google Scholar] [CrossRef]
  123. Li, Y.; Sun, N.; Lu, Z.; Sun, S.; Huang, J.; Chen, Z.; He, J. Prognostic alternative mRNA splicing signature in non-small cell lung cancer. Cancer Lett. 2017, 393, 40–51. [Google Scholar] [CrossRef]
  124. Clarke, M.R.; Landreneau, R.J.; Resnick, N.M.; Crowley, R.; Dougherty, G.J.; Cooper, D.L.; Yousem, S.A. Prognostic significance of CD44 expression in adenocarcinoma of the lung. Clin. Mol. Pathol. 1995, 48, M200–M204. [Google Scholar] [CrossRef]
  125. Ramasami, S.; Kerr, K.M.; Chapman, A.D.; King, G.; Cockburn, J.S.; Jeffrey, R.R. Expression of CD44v6 but not E-cadherin or beta-catenin influences prognosis in primary pulmonary adenocarcinoma. J. Pathol. 2000, 192, 427–432. [Google Scholar] [CrossRef]
  126. Lee, L.N.; Kuo, S.H.; Lee, Y.C.; Chang, Y.L.; Chang, H.C.; Jan, I.S.; Yang, P.C. CD44 splicing pattern is associated with disease progression in pulmonary adenocarcinoma. J. Formos. Med. Assoc. 2005, 104, 541–548. [Google Scholar] [PubMed]
  127. Suzuki, H.; Yamashiro, K. Reduced expression of CD44 v3 and v6 is related to invasion in lung adenocarcinoma. Lung Cancer 2002, 38, 137–141. [Google Scholar] [CrossRef]
  128. Ciulean, I.S.; Fischer, J.; Quaiser, A.; Bach, C.; Abken, H.; Tretbar, U.S.; Fricke, S.; Koehl, U.; Schmiedel, D.; Grunwald, T. CD44v6 specific CAR-NK cells for targeted immunotherapy of head and neck squamous cell carcinoma. Front. Immunol. 2023, 14, 1290488. [Google Scholar] [CrossRef] [PubMed]
  129. Reck, M.; Remon, J.; Hellmann, M.D. First-Line Immunotherapy for Non-Small-Cell Lung Cancer. J. Clin. Oncol. 2022, 40, 586–597. [Google Scholar] [CrossRef]
  130. Horn, L.; Mansfield, A.S.; Szczęsna, A.; Havel, L.; Krzakowski, M.; Hochmair, M.J.; Huemer, F.; Losonczy, G.; Johnson, M.L.; Nishio, M.; et al. First-Line Atezolizumab plus Chemotherapy in Extensive-Stage Small-Cell Lung Cancer. N. Engl. J. Med. 2018, 379, 2220–2229. [Google Scholar] [CrossRef]
  131. Xu, X.; Zhang, W.; Xuan, L.; Yu, Y.; Zheng, W.; Tao, F.; Nemechek, J.; He, C.; Ma, W.; Han, X.; et al. PD-1 signalling defines and protects leukaemic stem cells from T cell receptor-induced cell death in T cell acute lymphoblastic leukaemia. Nat. Cell Biol. 2023, 25, 170–182. [Google Scholar] [CrossRef] [PubMed]
  132. Wartewig, T.; Daniels, J.; Schulz, M.; Hameister, E.; Joshi, A.; Park, J.; Morrish, E.; Venkatasubramani, A.V.; Cernilogar, F.M.; van Heijster, F.H.A.; et al. PD-1 instructs a tumor-suppressive metabolic program that restricts glycolysis and restrains AP-1 activity in T cell lymphoma. Nat. Cancer 2023, 4, 1508–1525. [Google Scholar] [CrossRef] [PubMed]
  133. Wang, Y.; Nowakowski, G.S.; Wang, M.L.; Ansell, S.M. Advances in CD30- and PD-1-targeted therapies for classical Hodgkin lymphoma. J. Hematol. Oncol. 2018, 11, 57. [Google Scholar] [CrossRef] [PubMed]
  134. Huang, C.; Ren, S.; Chen, Y.; Liu, A.; Wu, Q.; Jiang, T.; Lv, P.; Song, D.; Hu, F.; Lan, J.; et al. PD-L1 methylation restricts PD-L1/PD-1 interactions to control cancer immune surveillance. Sci. Adv. 2023, 9, eade4186. [Google Scholar] [CrossRef]
  135. Jenkins, R.W.; Barbie, D.A.; Flaherty, K.T. Mechanisms of resistance to immune checkpoint inhibitors. Br. J. Cancer 2018, 118, 9–16. [Google Scholar] [CrossRef]
  136. Ribas, A.; Wolchok, J.D. Cancer immunotherapy using checkpoint blockade. Science 2018, 359, 1350–1355. [Google Scholar] [CrossRef]
  137. Pitt, J.M.; Vétizou, M.; Daillère, R.; Roberti, M.P.; Yamazaki, T.; Routy, B.; Lepage, P.; Boneca, I.G.; Chamaillard, M.; Kroemer, G.; et al. Resistance Mechanisms to Immune-Checkpoint Blockade in Cancer: Tumor-Intrinsic and -Extrinsic Factors. Immunity 2016, 44, 1255–1269. [Google Scholar] [CrossRef]
  138. Abu Hejleh, T.; Furqan, M.; Ballas, Z.; Clamon, G. The clinical significance of soluble PD-1 and PD-L1 in lung cancer. Crit. Rev. Oncol. Hematol. 2019, 143, 148–152. [Google Scholar] [CrossRef]
  139. Kruger, S.; Legenstein, M.L.; Rösgen, V.; Haas, M.; Modest, D.P.; Westphalen, C.B.; Ormanns, S.; Kirchner, T.; Heinemann, V.; Holdenrieder, S.; et al. Serum levels of soluble programmed death protein 1 (sPD-1) and soluble programmed death ligand 1 (sPD-L1) in advanced pancreatic cancer. Oncoimmunology 2017, 6, e1310358. [Google Scholar] [CrossRef]
  140. Niu, M.; Liu, Y.; Yi, M.; Jiao, D.; Wu, K. Biological Characteristics and Clinical Significance of Soluble PD-1/PD-L1 and Exosomal PD-L1 in Cancer. Front. Immunol. 2022, 13, 827921. [Google Scholar] [CrossRef]
  141. 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. [Google Scholar] [CrossRef] [PubMed]
  142. 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, 587460. [Google Scholar] [CrossRef] [PubMed]
  143. Széles, Á.; Fazekas, T.; Váncsa, S.; Váradi, M.; Kovács, P.T.; Krafft, U.; Grünwald, V.; Hadaschik, B.; Csizmarik, A.; Hegyi, P.; et al. Pre-treatment soluble PD-L1 as a predictor of overall survival for immune checkpoint inhibitor therapy: A systematic review and meta-analysis. Cancer Immunol. Immunother. 2023, 72, 1061–1073. [Google Scholar] [CrossRef]
  144. Himuro, H.; Nakahara, Y.; Igarashi, Y.; Kouro, T.; Higashijima, N.; Matsuo, N.; Murakami, S.; Wei, F.; Horaguchi, S.; Tsuji, K.; et al. Clinical roles of soluble PD-1 and PD-L1 in plasma of NSCLC patients treated with immune checkpoint inhibitors. Cancer Immunol. Immunother. 2023, 72, 2829–2840. [Google Scholar] [CrossRef] [PubMed]
  145. Qu, S.; Jiao, Z.; Lu, G.; Yao, B.; Wang, T.; Rong, W.; Xu, J.; Fan, T.; Sun, X.; Yang, R.; et al. PD-L1 lncRNA splice isoform promotes lung adenocarcinoma progression via enhancing c-Myc activity. Genome Biol. 2021, 22, 104. [Google Scholar] [CrossRef]
  146. Nielsen, C.; Ohm-Laursen, L.; Barington, T.; Husby, S.; Lillevang, S.T. Alternative splice variants of the human PD-1 gene. Cell Immunol. 2005, 235, 109–116. [Google Scholar] [CrossRef]
  147. Salton, M.; Misteli, T. Small Molecule Modulators of Pre-mRNA Splicing in Cancer Therapy. Trends Mol. Med. 2016, 22, 28–37. [Google Scholar] [CrossRef]
  148. Silvestri, I.; Manigrasso, J.; Andreani, A.; Brindani, N.; Mas, C.; Reiser, J.B.; Vidossich, P.; Martino, G.; McCarthy, A.A.; De Vivo, M.; et al. Targeting the conserved active site of splicing machines with specific and selective small molecule modulators. Nat. Commun. 2024, 15, 4980. [Google Scholar] [CrossRef]
  149. Alsafadi, S.; Dayot, S.; Tarin, M.; Houy, A.; Bellanger, D.; Cornella, M.; Wassef, M.; Waterfall, J.J.; Lehnert, E.; Roman-Roman, S.; et al. Genetic alterations of SUGP1 mimic mutant-SF3B1 splice pattern in lung adenocarcinoma and other cancers. Oncogene 2021, 40, 85–96. [Google Scholar] [CrossRef]
  150. Esfahani, M.S.; Lee, L.J.; Jeon, Y.J.; Flynn, R.A.; Stehr, H.; Hui, A.B.; Ishisoko, N.; Kildebeck, E.; Newman, A.M.; Bratman, S.V.; et al. Functional significance of U2AF1 S34F mutations in lung adenocarcinomas. Nat. Commun. 2019, 10, 5712. [Google Scholar] [CrossRef]
  151. Imielinski, M.; Berger, A.H.; Hammerman, P.S.; Hernandez, B.; Pugh, T.J.; Hodis, E.; Cho, J.; Suh, J.; Capelletti, M.; Sivachenko, A.; et al. Mapping the hallmarks of lung adenocarcinoma with massively parallel sequencing. Cell 2012, 150, 1107–1120. [Google Scholar] [CrossRef] [PubMed]
  152. Seiler, M.; Yoshimi, A.; Darman, R.; Chan, B.; Keaney, G.; Thomas, M.; Agrawal, A.A.; Caleb, B.; Csibi, A.; Sean, E.; et al. H3B-8800, an orally available small-molecule splicing modulator, induces lethality in spliceosome-mutant cancers. Nat. Med. 2018, 24, 497–504. [Google Scholar] [CrossRef]
  153. Cretu, C.; Schmitzová, J.; Ponce-Salvatierra, A.; Dybkov, O.; De Laurentiis, E.I.; Sharma, K.; Will, C.L.; Urlaub, H.; Lührmann, R.; Pena, V. Molecular Architecture of SF3b and Structural Consequences of Its Cancer-Related Mutations. Mol. Cell 2016, 64, 307–319. [Google Scholar] [CrossRef]
  154. Spinello, A.; Borišek, J.; Malcovati, L.; Magistrato, A. Investigating the Molecular Mechanism of H3B-8800: A Splicing Modulator Inducing Preferential Lethality in Spliceosome-Mutant Cancers. Int. J. Mol. Sci. 2021, 22, 11222. [Google Scholar] [CrossRef]
  155. Steensma, D.P.; Wermke, M.; Klimek, V.M.; Greenberg, P.L.; Font, P.; Komrokji, R.S.; Yang, J.; Brunner, A.M.; Carraway, H.E.; Ades, L.; et al. Phase I First-in-Human Dose Escalation Study of the oral SF3B1 modulator H3B-8800 in myeloid neoplasms. Leukemia 2021, 35, 3542–3550. [Google Scholar] [CrossRef] [PubMed]
  156. DeNicola, A.B.; Tang, Y. Therapeutic approaches to treat human spliceosomal diseases. Curr. Opin. Biotechnol. 2019, 60, 72–81. [Google Scholar] [CrossRef] [PubMed]
  157. Teng, T.; Tsai, J.H.; Puyang, X.; Seiler, M.; Peng, S.; Prajapati, S.; Aird, D.; Buonamici, S.; Caleb, B.; Chan, B.; et al. Splicing modulators act at the branch point adenosine binding pocket defined by the PHF5A-SF3b complex. Nat. Commun. 2017, 8, 15522. [Google Scholar] [CrossRef]
  158. 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. [Google Scholar] [CrossRef]
  159. Finci, L.I.; Zhang, X.; Huang, X.; Zhou, Q.; Tsai, J.; Teng, T.; Agrawal, A.; Chan, B.; Irwin, S.; Karr, C.; 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. [Google Scholar] [CrossRef]
  160. Kotake, Y.; Sagane, K.; Owa, T.; Mimori-Kiyosue, Y.; Shimizu, H.; Uesugi, M.; Ishihama, Y.; Iwata, M.; Mizui, Y. Splicing factor SF3b as a target of the antitumor natural product pladienolide. Nat. Chem. Biol. 2007, 3, 570–575. [Google Scholar] [CrossRef]
  161. Aptullahoglu, E.; Ciardullo, C.; Wallis, J.P.; Marr, H.; Marshall, S.; Bown, N.; Willmore, E.; Lunec, J. Splicing Modulation Results in Aberrant Isoforms and Protein Products of p53 Pathway Genes and the Sensitization of B Cells to Non-Genotoxic MDM2 Inhibition. Int. J. Mol. Sci. 2023, 24, 2410. [Google Scholar] [CrossRef] [PubMed]
  162. Aird, D.; Teng, T.; Huang, C.L.; Pazolli, E.; Banka, D.; Cheung-Ong, K.; Eifert, C.; Furman, C.; Wu, Z.J.; Seiler, M.; et al. Sensitivity to splicing modulation of BCL2 family genes defines cancer therapeutic strategies for splicing modulators. Nat. Commun. 2019, 10, 137. [Google Scholar] [CrossRef] [PubMed]
  163. Iwata, M. E7107, a new 7-urethan derivative of pladienplide D, displays curative effect against several human tumor xenografts. Proc. Am. Assoc. Cancer Res. 2004, 45, 691-a. [Google Scholar]
  164. Eskens, F.A.; Ramos, F.J.; Burger, H.; O’Brien, J.P.; Piera, A.; de Jonge, M.J.; Mizui, Y.; Wiemer, E.A.; Carreras, M.J.; Baselga, J.; et al. Phase I pharmacokinetic and pharmacodynamic study of the first-in-class spliceosome inhibitor E7107 in patients with advanced solid tumors. Clin. Cancer Res. 2013, 19, 6296–6304. [Google Scholar] [CrossRef]
  165. Corrionero, A.; Miñana, B.; Valcárcel, J. Reduced fidelity of branch point recognition and alternative splicing induced by the anti-tumor drug spliceostatin A. Genes Dev. 2011, 25, 445–459. [Google Scholar] [CrossRef]
  166. Sciarrillo, R.; Wojtuszkiewicz, A.; El Hassouni, B.; Funel, N.; Gandellini, P.; Lagerweij, T.; Buonamici, S.; Blijlevens, M.; Zeeuw van der Laan, E.A.; Zaffaroni, N.; et al. Splicing modulation as novel therapeutic strategy against diffuse malignant peritoneal mesothelioma. EBioMedicine 2019, 39, 215–225. [Google Scholar] [CrossRef]
  167. Sette, C.; Paronetto, M.P. Somatic Mutations in Core Spliceosome Components Promote Tumorigenesis and Generate an Exploitable Vulnerability in Human Cancer. Cancers 2022, 14, 1827. [Google Scholar] [CrossRef]
  168. Wu, T.; Millar, H.; Gaffney, D.; Beke, L.; Mannens, G.; Vinken, P.; Sommers, I.; Thuring, J.-W.; Sun, W.; Moy, C.; et al. Abstract 4859: 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. [Google Scholar] [CrossRef]
  169. Vieito, M.; Moreno, V.; Spreafico, A.; Brana, I.; Wang, J.S.; Preis, M.; Hernández, T.; Genta, S.; Hansen, A.R.; Doger, B.; et al. Phase 1 Study of JNJ-64619178, a Protein Arginine Methyltransferase 5 Inhibitor, in Advanced Solid Tumors. Clin. Cancer Res. 2023, 29, 3592–3602. [Google Scholar] [CrossRef]
  170. Brehmer, D.; Beke, L.; Wu, T.; Millar, H.J.; Moy, C.; Sun, W.; Mannens, G.; Pande, V.; Boeckx, A.; van Heerde, E.; 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. [Google Scholar] [CrossRef]
  171. Yamanaka, S.; Horiuchi, Y.; Matsuoka, S.; Kido, K.; Nishino, K.; Maeno, M.; Shibata, N.; Kosako, H.; Sawasaki, T. A proximity biotinylation-based approach to identify protein-E3 ligase interactions induced by PROTACs and molecular glues. Nat. Commun. 2022, 13, 183. [Google Scholar] [CrossRef] [PubMed]
  172. Han, T.; Goralski, M.; Gaskill, N.; Capota, E.; Kim, J.; Ting, T.C.; Xie, Y.; Williams, N.S.; Nijhawan, D. Anticancer sulfonamides target splicing by inducing RBM39 degradation via recruitment to DCAF15. Science 2017, 356, eaal3755. [Google Scholar] [CrossRef] [PubMed]
  173. Nijhuis, A.; Sikka, A.; Yogev, O.; Herendi, L.; Balcells, C.; Ma, Y.; Poon, E.; Eckold, C.; Valbuena, G.N.; Xu, Y.; et al. Indisulam targets RNA splicing and metabolism to serve as a therapeutic strategy for high-risk neuroblastoma. Nat. Commun. 2022, 13, 1380. [Google Scholar] [CrossRef]
  174. Drilon, A.; Clark, J.W.; Weiss, J.; Ou, S.I.; Camidge, D.R.; Solomon, B.J.; Otterson, G.A.; Villaruz, L.C.; Riely, G.J.; Heist, R.S.; et al. Antitumor activity of crizotinib in lung cancers harboring a MET exon 14 alteration. Nat. Med. 2020, 26, 47–51. [Google Scholar] [CrossRef] [PubMed]
  175. Paik, P.K.; Drilon, A.; Fan, P.D.; Yu, H.; Rekhtman, N.; Ginsberg, M.S.; Borsu, L.; Schultz, N.; Berger, M.F.; Rudin, C.M.; et al. Response to MET inhibitors in patients with stage IV lung adenocarcinomas harboring MET mutations causing exon 14 skipping. Cancer Discov. 2015, 5, 842–849. [Google Scholar] [CrossRef]
  176. Kole, R.; Krainer, A.R.; Altman, S. RNA therapeutics: Beyond RNA interference and antisense oligonucleotides. Nat. Rev. Drug Discov. 2012, 11, 125–140. [Google Scholar] [CrossRef]
  177. Fu, Z.; Zhang, X.; Zhou, X.; Ur-Rehman, U.; Yu, M.; Liang, H.; Guo, H.; Guo, X.; Kong, Y.; Su, Y.; et al. In vivo self-assembled small RNAs as a new generation of RNAi therapeutics. Cell Res. 2021, 31, 631–648. [Google Scholar] [CrossRef]
  178. Zhou, Y.; Zhou, G.; Tian, C.; Jiang, W.; Jin, L.; Zhang, C.; Chen, X. Exosome-mediated small RNA delivery for gene therapy. Wiley Interdiscip. Rev. RNA 2016, 7, 758–771. [Google Scholar] [CrossRef]
  179. Barbie, D.A.; Tamayo, P.; Boehm, J.S.; Kim, S.Y.; Moody, S.E.; Dunn, I.F.; Schinzel, A.C.; Sandy, P.; Meylan, E.; Scholl, C.; et al. Systematic RNA interference reveals that oncogenic KRAS-driven cancers require TBK1. Nature 2009, 462, 108–112. [Google Scholar] [CrossRef]
  180. Madanayake, T.W.; Welsh, E.A.; Darville, L.N.F.; Koomen, J.M.; Chalfant, C.E.; Haura, E.B.; Robinson, T.J. Inhibition of Epidermal Growth Factor Receptor Signaling by Antisense Oligonucleotides as a Novel Approach to Epidermal Growth Factor Receptor Inhibition. Nucleic Acid Ther. 2022, 32, 391–400. [Google Scholar] [CrossRef]
  181. Shi, B.; An, K.; Wang, Y.; Fei, Y.; Guo, C.; Cliff Zhang, Q.; Yang, Y.G.; Tian, X.; Kan, Q. RNA Structural Dynamics Modulate EGFR-TKI Resistance Through Controlling YRDC Translation in NSCLC Cells. Genom. Proteom. Bioinform. 2023, 21, 850–865. [Google Scholar] [CrossRef]
  182. Wang, L.; Fu, H.; Song, L.; Wu, Z.; Yu, J.; Guo, Q.; Chen, C.; Yang, X.; Zhang, J.; Wang, Q.; et al. Overcoming AZD9291 Resistance and Metastasis of NSCLC via Ferroptosis and Multitarget Interference by Nanocatalytic Sensitizer Plus AHP-DRI-12. Small 2023, 19, e2204133. [Google Scholar] [CrossRef] [PubMed]
  183. Kim, J.; Woo, S.; de Gusmao, C.M.; Zhao, B.; Chin, D.H.; DiDonato, R.L.; Nguyen, M.A.; Nakayama, T.; Hu, C.A.; Soucy, A.; et al. A framework for individualized splice-switching oligonucleotide therapy. Nature 2023, 619, 828–836. [Google Scholar] [CrossRef] [PubMed]
  184. Wan, L.; Kral, A.J.; Voss, D.; Schäfer, B.; Sudheendran, K.; Danielsen, M.; Caruthers, M.H.; Krainer, A.R. Screening Splice-Switching Antisense Oligonucleotides in Pancreas-Cancer Organoids. Nucleic Acid Ther. 2024, 34, 188–198. [Google Scholar] [CrossRef]
  185. Bauman, J.A.; Li, S.D.; Yang, A.; Huang, L.; Kole, R. Anti-tumor activity of splice-switching oligonucleotides. Nucleic Acids Res. 2010, 38, 8348–8356. [Google Scholar] [CrossRef] [PubMed]
  186. Chen, X.; Mangala, L.S.; Rodriguez-Aguayo, C.; Kong, X.; Lopez-Berestein, G.; Sood, A.K. RNA interference-based therapy and its delivery systems. Cancer Metastasis Rev. 2018, 37, 107–124. [Google Scholar] [CrossRef]
  187. Gagliardi, M.; Ashizawa, A.T. The Challenges and Strategies of Antisense Oligonucleotide Drug Delivery. Biomedicines 2021, 9, 433. [Google Scholar] [CrossRef]
  188. Choi, J.W.; Lee, J.S.; Kim, S.W.; Yun, C.O. Evolution of oncolytic adenovirus for cancer treatment. Adv. Drug Deliv. Rev. 2012, 64, 720–729. [Google Scholar] [CrossRef]
  189. Hemminki, O.; Dos Santos, J.M.; Hemminki, A. Oncolytic viruses for cancer immunotherapy. J. Hematol. Oncol. 2020, 13, 84. [Google Scholar] [CrossRef]
  190. Deng, H.; Liu, H.; de Silva, T.; Xue, Y.; Mohamud, Y.; Ng, C.S.; Qu, J.; Zhang, J.; Jia, W.W.G.; Lockwood, W.W.; et al. Coxsackievirus Type B3 Is a Potent Oncolytic Virus against KRAS-Mutant Lung Adenocarcinoma. Mol. Ther. Oncolytics 2019, 14, 266–278. [Google Scholar] [CrossRef]
  191. Villalona-Calero, M.A.; Lam, E.; Otterson, G.A.; Zhao, W.; Timmons, M.; Subramaniam, D.; Hade, E.M.; Gill, G.M.; Coffey, M.; Selvaggi, G.; et al. Oncolytic reovirus in combination with chemotherapy in metastatic or recurrent non-small cell lung cancer patients with KRAS-activated tumors. Cancer 2016, 122, 875–883. [Google Scholar] [CrossRef]
  192. Yan, Y.; Su, C.; Hang, M.; Huang, H.; Zhao, Y.; Shao, X.; Bu, X. Recombinant Newcastle disease virus rL-RVG enhances the apoptosis and inhibits the migration of A549 lung adenocarcinoma cells via regulating alpha 7 nicotinic acetylcholine receptors in vitro. Virol. J. 2017, 14, 190. [Google Scholar] [CrossRef]
  193. Yoo, J.Y.; Kim, J.H.; Kwon, Y.G.; Kim, E.C.; Kim, N.K.; Choi, H.J.; Yun, C.O. VEGF-specific short hairpin RNA-expressing oncolytic adenovirus elicits potent inhibition of angiogenesis and tumor growth. Mol. Ther. 2007, 15, 295–302. [Google Scholar] [CrossRef] [PubMed]
  194. Kottke, T.; Hall, G.; Pulido, J.; Diaz, R.M.; Thompson, J.; Chong, H.; Selby, P.; Coffey, M.; Pandha, H.; Chester, J.; et al. Antiangiogenic cancer therapy combined with oncolytic virotherapy leads to regression of established tumors in mice. J. Clin. Investig. 2010, 120, 1551–1560. [Google Scholar] [CrossRef] [PubMed]
  195. Guo, N.; Liu, J.B.; Li, W.; Ma, Y.S.; Fu, D. The power and the promise of CRISPR/Cas9 genome editing for clinical application with gene therapy. J. Adv. Res. 2022, 40, 135–152. [Google Scholar] [CrossRef] [PubMed]
  196. Platt, R.J.; Chen, S.; Zhou, Y.; Yim, M.J.; Swiech, L.; Kempton, H.R.; Dahlman, J.E.; Parnas, O.; Eisenhaure, T.M.; Jovanovic, M.; et al. CRISPR-Cas9 knockin mice for genome editing and cancer modeling. Cell 2014, 159, 440–455. [Google Scholar] [CrossRef]
  197. Fan, W.; Huang, J.; Tian, F.; Hong, X.; Zhu, K.; Zhan, Y.; Li, X.; Wang, X.; Wang, X.; Cai, L.; et al. m(6)A-Modified SNRPA Controls Alternative Splicing of ERCC1 Exon 8 to Induce Cisplatin Resistance in Lung Adenocarcinoma. Adv. Sci. 2024, 11, e2404609. [Google Scholar] [CrossRef]
  198. Thomas, J.D.; Polaski, J.T.; Feng, Q.; De Neef, E.J.; Hoppe, E.R.; McSharry, M.V.; Pangallo, J.; Gabel, A.M.; Belleville, A.E.; Watson, J.; et al. RNA isoform screens uncover the essentiality and tumor-suppressor activity of ultraconserved poison exons. Nat. Genet. 2020, 52, 84–94. [Google Scholar] [CrossRef]
  199. Li, K.; Peng, Z.Y.; Wang, R.; Li, X.; Du, N.; Liu, D.P.; Zhang, J.; Zhang, Y.F.; Ma, L.; Sun, Y.; et al. Enhancement of TKI sensitivity in lung adenocarcinoma through m6A-dependent translational repression of Wnt signaling by circ-FBXW7. Mol. Cancer 2023, 22, 103. [Google Scholar] [CrossRef]
  200. Huang, L.; Jiang, S.; Shi, Y. Tyrosine kinase inhibitors for solid tumors in the past 20 years (2001–2020). J. Hematol. Oncol. 2020, 13, 143. [Google Scholar] [CrossRef]
  201. Wu, Y.L.; Tsuboi, M.; He, J.; John, T.; Grohe, C.; Majem, M.; Goldman, J.W.; Laktionov, K.; Kim, S.W.; Kato, T.; et al. Osimertinib in Resected EGFR-Mutated Non-Small-Cell Lung Cancer. N. Engl. J. Med. 2020, 383, 1711–1723. [Google Scholar] [CrossRef] [PubMed]
  202. Han, H.; Li, S.; Chen, T.; Fitzgerald, M.; Liu, S.; Peng, C.; Tang, K.H.; Cao, S.; Chouitar, J.; Wu, J.; et al. Targeting HER2 Exon 20 Insertion-Mutant Lung Adenocarcinoma with a Novel Tyrosine Kinase Inhibitor Mobocertinib. Cancer Res. 2021, 81, 5311–5324. [Google Scholar] [CrossRef]
  203. Passaro, A.; Mok, T.; Peters, S.; Popat, S.; Ahn, M.J.; de Marinis, F. Recent Advances on the Role of EGFR Tyrosine Kinase Inhibitors in the Management of NSCLC With Uncommon, Non Exon 20 Insertions, EGFR Mutations. J. Thorac. Oncol. 2021, 16, 764–773. [Google Scholar] [CrossRef] [PubMed]
  204. Hsu, C.C.; Liao, B.C.; Liao, W.Y.; Markovets, A.; Stetson, D.; Thress, K.; Yang, J.C. 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. [Google Scholar] [CrossRef]
  205. Yamaoka, T.; Tsurutani, J.; Sagara, H.; Ohmori, T. HER2-D16 oncogenic driver mutation confers osimertinib resistance in EGFR mutation-positive non-small cell lung cancer. Transl. Lung Cancer Res. 2020, 9, 2178–2183. [Google Scholar] [CrossRef] [PubMed]
  206. Ng, K.P.; Hillmer, A.M.; Chuah, C.T.; Juan, W.C.; Ko, T.K.; Teo, A.S.; Ariyaratne, P.N.; Takahashi, N.; Sawada, K.; Fei, Y.; 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. [Google Scholar] [CrossRef]
  207. Yu, M.; Nah, G.S.S.; Krishnan, V.; Sulaimi, F.N.B.; Ng, K.P.; Wang, C.; Bhatt, S.; Chuah, C.; Bergstrom, D.E.; Ong, S.T. The BIM deletion polymorphism potentiates the survival of leukemia stem and progenitor cells and impairs response to targeted therapies. Leukemia 2024, 39, 134–143. [Google Scholar] [CrossRef]
  208. Mann, B.S.; Johnson, J.R.; Cohen, M.H.; Justice, R.; Pazdur, R. FDA approval summary: Vorinostat for treatment of advanced primary cutaneous T-cell lymphoma. Oncologist 2007, 12, 1247–1252. [Google Scholar] [CrossRef]
  209. Tanimoto, A.; Takeuchi, S.; Arai, S.; Fukuda, K.; Yamada, T.; Roca, X.; Ong, S.T.; Yano, S. Histone Deacetylase 3 Inhibition Overcomes BIM Deletion Polymorphism-Mediated Osimertinib Resistance in EGFR-Mutant Lung Cancer. Clin. Cancer Res. 2017, 23, 3139–3149. [Google Scholar] [CrossRef]
  210. Takeuchi, S.; Hase, T.; Shimizu, S.; Ando, M.; Hata, A.; Murakami, H.; Kawakami, T.; Nagase, K.; Yoshimura, K.; Fujiwara, T.; 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. [Google Scholar] [CrossRef]
  211. Bao, Y.; Zhang, S.; Zhang, X.; Pan, Y.; Yan, Y.; Wang, N.; Ren, Y.; Zuo, J.; Zong, W.X.; Wang, Z.; et al. RBM10 Loss Promotes EGFR-Driven Lung Cancer and Confers Sensitivity to Spliceosome Inhibition. Cancer Res. 2023, 83, 1490–1502. [Google Scholar] [CrossRef]
  212. Blakely, C.M.; Urisman, A.; Gubens, M.A.; Mulvey, C.K.; Allen, G.M.; Shiboski, S.C.; Rotow, J.K.; Chakrabarti, T.; Kerr, D.L.; Aredo, J.V.; et al. Neoadjuvant Osimertinib for the Treatment of Stage I-IIIA Epidermal Growth Factor Receptor-Mutated Non-Small Cell Lung Cancer: A Phase II Multicenter Study. J. Clin. Oncol. 2024, 42, 3105–3114. [Google Scholar] [CrossRef]
  213. Nanjo, S.; Wu, W.; Karachaliou, N.; Blakely, C.M.; Suzuki, J.; Chou, Y.T.; Ali, S.M.; Kerr, D.L.; Olivas, V.R.; Shue, J.; et al. Deficiency of the splicing factor RBM10 limits EGFR inhibitor response in EGFR-mutant lung cancer. J. Clin. Investig. 2022, 132, e145099. [Google Scholar] [CrossRef]
  214. Yang, K.; Halima, A.; Chan, T.A. Antigen presentation in cancer—Mechanisms and clinical implications for immunotherapy. Nat. Rev. Clin. Oncol. 2023, 20, 604–623. [Google Scholar] [CrossRef]
  215. Shepherd, F.A.; Douillard, J.Y.; Blumenschein, G.R., Jr. Immunotherapy for non-small cell lung cancer: Novel approaches to improve patient outcome. J. Thorac. Oncol. 2011, 6, 1763–1773. [Google Scholar] [CrossRef]
  216. Hong, M.; Clubb, J.D.; Chen, Y.Y. Engineering CAR-T Cells for Next-Generation Cancer Therapy. Cancer Cell 2020, 38, 473–488. [Google Scholar] [CrossRef]
  217. Neelapu, S.S.; Locke, F.L.; Bartlett, N.L.; Lekakis, L.J.; Miklos, D.B.; Jacobson, C.A.; Braunschweig, I.; Oluwole, O.O.; Siddiqi, T.; Lin, Y.; et al. Axicabtagene Ciloleucel CAR T-Cell Therapy in Refractory Large B-Cell Lymphoma. N. Engl. J. Med. 2017, 377, 2531–2544. [Google Scholar] [CrossRef]
  218. Maude, S.L.; Laetsch, T.W.; Buechner, J.; Rives, S.; Boyer, M.; Bittencourt, H.; Bader, P.; Verneris, M.R.; Stefanski, H.E.; Myers, G.D.; et al. Tisagenlecleucel in Children and Young Adults with B-Cell Lymphoblastic Leukemia. N. Engl. J. Med. 2018, 378, 439–448. [Google Scholar] [CrossRef]
  219. Maalej, K.M.; Merhi, M.; Inchakalody, V.P.; Mestiri, S.; Alam, M.; Maccalli, C.; Cherif, H.; Uddin, S.; Steinhoff, M.; Marincola, F.M.; et al. CAR-cell therapy in the era of solid tumor treatment: Current challenges and emerging therapeutic advances. Mol. Cancer 2023, 22, 20. [Google Scholar] [CrossRef]
  220. Frankiw, L.; Baltimore, D.; Li, G. Alternative mRNA splicing in cancer immunotherapy. Nat. Rev. Immunol. 2019, 19, 675–687. [Google Scholar] [CrossRef]
  221. Peng, Q.; Zhou, Y.; Oyang, L.; Wu, N.; Tang, Y.; Su, M.; Luo, X.; Wang, Y.; Sheng, X.; Ma, J.; et al. Impacts and mechanisms of alternative mRNA splicing in cancer metabolism, immune response, and therapeutics. Mol. Ther. 2022, 30, 1018–1035. [Google Scholar] [CrossRef] [PubMed]
  222. Yoshida, C.; Kadota, K.; Yamada, K.; Fujimoto, S.; Ibuki, E.; Ishikawa, R.; Haba, R.; Yajima, T. CD44v6 downregulation as a prognostic factor for distant recurrence in resected stage I lung adenocarcinomas. Clin. Exp. Med. 2023, 23, 5191–5200. [Google Scholar] [CrossRef]
Figure 1. Predicted protein structures of EGFR isoforms using PyMOL 2.6 and AlphaFold 3.0.
Figure 1. Predicted protein structures of EGFR isoforms using PyMOL 2.6 and AlphaFold 3.0.
Cancers 17 01329 g001
Figure 2. Constitutive splicing (top) and seven elementary types of alternative splicing (created with Microsoft PowerPoint 2024).
Figure 2. Constitutive splicing (top) and seven elementary types of alternative splicing (created with Microsoft PowerPoint 2024).
Cancers 17 01329 g002
Figure 3. Summary of AS events in LUAD derived from TCGA SpliceSeq. (A) Proportion of different types of AS events in LUAD. (B) Distribution analysis of a total of 10,314 genes characterized by the number of AS events per gene. A percentage of samples with Percent Spliced In (PSI) values greater than 75% was included in analysis. (Created with GraphPad Prism 8.0 and Microsoft PowerPoint 2024).
Figure 3. Summary of AS events in LUAD derived from TCGA SpliceSeq. (A) Proportion of different types of AS events in LUAD. (B) Distribution analysis of a total of 10,314 genes characterized by the number of AS events per gene. A percentage of samples with Percent Spliced In (PSI) values greater than 75% was included in analysis. (Created with GraphPad Prism 8.0 and Microsoft PowerPoint 2024).
Cancers 17 01329 g003
Figure 4. Small-molecule inhibitors, such as H3B-8800 and E7107, inhibit the spliceosome through direct interaction with the SF3b complex (created with Microsoft PowerPoint 2024 and Adobe Illustrator 2025).
Figure 4. Small-molecule inhibitors, such as H3B-8800 and E7107, inhibit the spliceosome through direct interaction with the SF3b complex (created with Microsoft PowerPoint 2024 and Adobe Illustrator 2025).
Cancers 17 01329 g004
Figure 5. Overview of RNA-targeted therapies, including RNAi, ASOs, and SSOs (created with Microsoft PowerPoint 2024 and Adobe Illustrator 2025).
Figure 5. Overview of RNA-targeted therapies, including RNAi, ASOs, and SSOs (created with Microsoft PowerPoint 2024 and Adobe Illustrator 2025).
Cancers 17 01329 g005
Figure 6. Tumor-specific CD44v6 derived from aberrant AS represents a promising neoantigen for CAR T cell immunotherapy in LUAD (created with Microsoft PowerPoint 2024 and Adobe Illustrator 2025).
Figure 6. Tumor-specific CD44v6 derived from aberrant AS represents a promising neoantigen for CAR T cell immunotherapy in LUAD (created with Microsoft PowerPoint 2024 and Adobe Illustrator 2025).
Cancers 17 01329 g006
Table 1. Detailed characteristics of EGFR AS isoforms in the NCBI Gene database.
Table 1. Detailed characteristics of EGFR AS isoforms in the NCBI Gene database.
EGFR AS IsoformAS TypeNumber of a.a.sProtein Size (kDa)Predicted Subcellular Location
Isoform aCanonical isoform1210134.3Cell surface protein
Isoform bAlternative last exon62869.2Secreted protein
Isoform cAlternative last exon40544.7Secreted protein
Isoform dAlternative last exon70577.3Secreted protein
Isoform eExon skipping and alternative last exon1091120.7Cell surface protein
Isoform fAlternative last exon1136125.8Cell surface protein
Isoform gExon skipping1165129.2Cell surface protein
Isoform hAlternative first exon1157128.7Cell surface protein
Isoform I (EGFRvIII)Alternative first exon943104.3Cell surface protein
Table 2. LUAD-related AS isoforms.
Table 2. LUAD-related AS isoforms.
GeneAS EventBiological or Clinical ImplicationsAS-Related TherapyReference(s)
KRASInclusion of either exon 4A (KRAS4A) or exon 4B (KRAS4B)KRAS4A induces metastatic LUAD in vivoASO[51,52]
EGFRDeletion of exons 2–7 (EGFRvIII)EGFRvIII contributes to resistance against TKIsASO[53,54]
c-MetDeletion of exon 14 (MET-ΔEx14)MET-ΔEx14 confers clinical sensitivity to MET inhibitorsSmall-molecule inhibitors[55,56]
CD44Inclusion of variant exon 6 (CD44v6)CD44v6 is implicated in the process of lung tumorigenesisCAR T cell therapy[57,58]
PD-1Deletion of exon 3 (sPD-1)sPD-1 enhances anti-tumor immunityAnti-PD-1/PD-L1 therapy[59,60]
Bcl-xAlternative last exon (Bcl-xL)Bcl-xL inhibits tumor cell apoptosisASO[61,62]
CDC25CExon 3 skipping (CDC25C-ΔEx3)CDC25C-ΔEx3 inhibits cell proliferation/[63]
KAT2AAlternative 5′ splicingThe splicing variant promotes LUAD progression/[64]
EIF4HExon 5 skipping (EIF4H-ΔEx5)EIF4H-ΔEx5 promotes LUAD progression/[65]
IMPAD1Alternative last exonThe splicing variant promotes LUAD proliferation/[66]
C-Met, c-mesenchymal–epithelial transition factor; PD-1, programmed cell death 1; PD-L1, programmed cell death ligand 1; Bcl-x(L), B-cell lymphoma extra-large; CDC25C, cell division cycle 25C; KAT2A, lysine acetyltransferase 2A; EIF4H, eukaryotic translation initiation factor 4H; IMPAD1, inositol monophosphatase domain-containing protein 1.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Luo, W.; Xu, M.; Wong, N.; Ng, C.S.-H. Alternative Splicing in Lung Adenocarcinoma: From Bench to Bedside. Cancers 2025, 17, 1329. https://doi.org/10.3390/cancers17081329

AMA Style

Luo W, Xu M, Wong N, Ng CS-H. Alternative Splicing in Lung Adenocarcinoma: From Bench to Bedside. Cancers. 2025; 17():1329. https://doi.org/10.3390/cancers17081329

Chicago/Turabian Style

Luo, Wenjie, Mingjing Xu, Nathalie Wong, and Calvin Sze-Hang Ng. 2025. "Alternative Splicing in Lung Adenocarcinoma: From Bench to Bedside" Cancers 17, no. : 1329. https://doi.org/10.3390/cancers17081329

APA Style

Luo, W., Xu, M., Wong, N., & Ng, C. S.-H. (2025). Alternative Splicing in Lung Adenocarcinoma: From Bench to Bedside. Cancers, 17(), 1329. https://doi.org/10.3390/cancers17081329

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