Next Article in Journal
Role of Single-Nucleotide Polymorphisms in Genes Implicated in Capecitabine Pharmacodynamics on the Effectiveness of Adjuvant Therapy in Colorectal Cancer
Next Article in Special Issue
Separation of Lipoproteins for Quantitative Analysis of 14C-Labeled Lipid-Soluble Compounds by Accelerator Mass Spectrometry
Previous Article in Journal
Substances Secreted by Lactobacillus spp. from the Urinary Tract Microbiota Play a Protective Role against Proteus mirabilis Infections and Their Complications
Previous Article in Special Issue
MiR-221-3p/222-3p Cluster Expression in Human Adipose Tissue Is Related to Obesity and Type 2 Diabetes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Actionable Driver Events in Small Cell Lung Cancer

by
Mirian Gutiérrez
1,
Irene Zamora
1,
Michael R. Freeman
2,3,
Ignacio J. Encío
1,4,* and
Mirja Rotinen
1,4,*
1
Department of Health Sciences, Public University of Navarre, 31008 Pamplona, Spain
2
Departments of Urology and Biomedical Sciences, Cedars-Sinai Medical Center, Los Angeles, CA 90048, USA
3
Department of Medicine, University of California Los Angeles, Los Angeles, CA 90095, USA
4
IdiSNA, Navarre Institute for Health Research, 31006 Pamplona, Spain
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(1), 105; https://doi.org/10.3390/ijms25010105
Submission received: 20 November 2023 / Revised: 15 December 2023 / Accepted: 18 December 2023 / Published: 20 December 2023

Abstract

:
Small cell lung cancer (SCLC) stands out as the most aggressive form of lung cancer, characterized by an extremely high proliferation rate and a very poor prognosis, with a 5-year survival rate that falls below 7%. Approximately two-thirds of patients receive their diagnosis when the disease has already reached a metastatic or extensive stage, leaving chemotherapy as the remaining first-line treatment option. Other than the recent advances in immunotherapy, which have shown moderate results, SCLC patients cannot yet benefit from any approved targeted therapy, meaning that this cancer remains treated as a uniform entity, disregarding intra- or inter-tumoral heterogeneity. Continuous efforts and technological improvements have enabled the identification of new potential targets that could be used to implement novel therapeutic strategies. In this review, we provide an overview of the most recent approaches for SCLC treatment, providing an extensive compilation of the targeted therapies that are currently under clinical evaluation and inhibitor molecules with promising results in vitro and in vivo.

1. Introduction

Lung cancer is the leading cause of cancer death, with 1.8 million estimated deaths and over 2 million cases diagnosed worldwide in 2020 [1,2]. The disease is subdivided into two fundamental histological groups: non-small cell lung cancer (NSCLC) and small cell lung cancer (SCLC). This classification is based on the microscopic characteristics of the tumor cells and is crucial for determining the appropriate therapeutic approach. SCLC makes up about 15% of lung cancer cases and is the most aggressive subtype, characterized by an exceptionally high proliferation rate and significant metastatic potential, leading to a very poor prognosis. Despite the initial responsiveness to chemotherapy, the majority of patients with SCLC will eventually relapse, with a dismal 5-year survival of less than 7% and most patients dying within 12 months [3]. The etymology of SCLC is strongly associated with tobacco, with smokers representing 98% of patients [4]. The average patient is a male over 70 years of age with significant current previous tobacco exposure and pre-existing pulmonary, cardiovascular, and metabolic conditions [5].
At present, chemotherapy is the primary approach to treatment of SCLC. According to the ESMO guidelines, surgery is considered for patients diagnosed at stages I or II, and is typically combined with chemoradiotherapy. For cases with complete resection and limited or no nodal involvement, cisplatin-etoposide is recommended. However, two-thirds of patients are diagnosed at the metastatic or extensive stage disease [3]. In these situations where resection is not possible, chemotherapy takes precedence as the first-line treatment. Irrespective of the disease stage at diagnosis, the guidelines advise prophylactic brain irradiation to mitigate the risk of brain metastasis [6].
Despite the initial favorable responses to chemotherapy, relapses are frequent after the first year of treatment [7]. Over the past few decades, second-line treatments such as cyclophosphamide/doxorubicin/vincristine (CAV), topotecan, or lurbinectedin have been approved. Immunotherapy is gaining attention due to its potential to suppress immunologically active tumors, and may be particularly applicable in SCLC due to its high immunogenicity, which promotes expression of neoantigens and their subsequent recognition by the immune system [6,8,9]. Currently, immunotherapy is used as monotherapy as a third-line strategy, and many trials are exploring the possibility of its application as a first-line therapy in combination with chemotherapy [6]. The recent IMpower133 and CASPIAN clinical trials have provided compelling support for the immune checkpoint inhibitors (ICIs) atezolizumab and durvalumab in combination with chemotherapy. This approach has demonstrated significant enhancements in both overall survival (OS) and progression-free survival (PFS) when compared with chemotherapy alone in the context of extensive stage disease [10,11]. However, in SCLC the effectiveness of immunotherapy is lower than in other solid tumors such as NSCLC due to the variety of mechanisms employed by these tumors to evade the immune system, and only a small percentage of patients demonstrate lasting responses [3].
In an era where recent advances in personalized medicine have enabled the development of targeted therapies tailored to the molecular alteration characteristic of each condition, SCLC continues to be managed as a single entity. However, there are many ongoing efforts for the discovery of new targets and their therapeutic application. This review compiles the most recent studies that have successfully translated findings into clinical trials or may do so in the near future.

1.1. Mutational Profile and Molecular Landscape of SCLC

SCLC is a genetically intricate condition characterized by a high mutation rate of approximately nine non-synonymous mutations per mega-base. Its mutational profile is highly associated with tobacco carcinogens, and most patients present dual loss of the tumor suppressors p53 and RB [12]. Their functional paralogs TP73 and RBL2 (RB Transcriptional Corepressor Like 2), respectively, are mutated in a significant portion of SCLC patients as well [13].
Among other common events in this type of cancer are mutations in genes that encode enzymes modifying histones, such as the histone acetyltransferases CREBBP (CREB Binding Protein) and EP300 (E1A Binding Protein P300), as well as the histone methyltransferases MLL (Mixed Lineage Leukemia). Genes encoding chromatin modifying complexes (e.g., BAF-SWI/SNF and PBAF-SWI/SNF complexes) have been identified as being frequently mutated in SCLC, including the ARID1A/B (AT-rich Interaction Domain 1A and 1B), BRG1 (or SMARCA4), PBRM1 (Polybromo-1), and CHD7 (Chromodomain Helicase DNA Binding Protein 7) genes [14]. SCLC tumors show recurrent somatic alterations in proteins related to cytoskeleton formation or cell–cell and cell–matrix interactions, such as SLIT2 (Slit Guidance Ligand 2) and mutagenic alterations in kinase receptors, like the inactivation of EPHA7 (EPH Receptor A7) and/or the amplification of FGFR (Fibroblast Growth Factor Receptor) and MET [15]. The PI3K (Phosphatidylinositol 3-kinase)/AKT/mTOR (Mammalian Target of Rapamycin) signaling pathway, linked to enhanced proliferation and migration of SCLC cells, is another frequently altered pathway, with activating mutations found in PIK3CA (Phosphatidylinositol-4,5-Bisphosphate 3-Kinase Catalytic Subunit Alpha), AKT, and PTEN (Phosphatase and Tensin Homolog) genes [16].
Inactivating mutations in NOTCH family members, especially NOTCH1, are present in 25% of SCLC tumors [12]. In the context of RB1 and p53 loss, loss of Notch drives SCLC cells into a stem-like state, thereby contributing to cancer progression. Nevertheless, Notch is active in a subset of human SCLC, and this pathway has been described as playing a key role in directing tumor transdifferentiation towards a non-neuroendocrine state, thereby promoting drug resistance mechanisms [17]. The MYC family of transcription factors (C-MYC, L-MYC and N-MYC) exhibit mutually exclusive amplifications in 20% of SCLC tumors [15]. C-MYC expression and activity is associated with increased tumor progression and metastatic potential, and a large body of evidence suggests that it is involved in the reverse neuroendocrine transdifferentiation (NET) process [17]. The SOX family of transcription factors, particularly SOX2 (SRY-Box Transcription Factor 2), is frequently amplified and overexpressed in SCLC tumors and cell lines. Upregulation of SOX2 is regarded as a pivotal event in SCLC, where the protein functions as a lineage-survival oncogene that contributes to the transcriptional deregulation observed in the disease and promotes the classical neuroendocrine phenotype [13,18]. The whole set of molecular alterations described above is depicted in Figure 1.

1.2. Tumor Heterogeneity and Plasticity in SCLC

SCLC is a high-grade malignant epithelial tumor. The pulmonary epithelium contains various populations of stem and progenitor cells that give rise to the variety of differentiated cell types in the lung. The fact that SCLC cells express neuroendocrine markers has led to the assumption that they originate from neuroendocrine cells in the pulmonary epithelium or their neuroendocrine progenitors [19]. Supporting this theory, these cells express neuroendocrine transcription factors such as ASCL1 (Achaete-Scute Homolog 1), which plays a crucial role in the proper development of pulmonary neuroendocrine cells and has been shown to be essential for the survival of SCLC cells [20]. However, it has been demonstrated that deleting p53 and RB1 in distinct lung cell types can give rise to different SCLC subtypes, implying that the cell of origin can potentially define the trajectory of tumor progression [21,22]. In addition to a de novo origin, SCLC can arise as a result of histological transformation in response to adverse conditions, such as hypoxia or selective pressure to escape targeted therapy against an oncogenic driver in lung adenocarcinoma. This plasticity, referred to as NET, has been identified in more than 14% of patients with EGFR (Epidermal Growth Factor Receptor)-mutated NSCLC in response to resistance to EGFR inhibitor therapy [23,24].
While the majority of SCLC tumors and cell lines display a morphology in concordance with the ‘classic’ WHO description of SCLC, and express a comprehensive range of neuroendocrine markers, there is a subgroup of variants characterized by reduced or absent expression of these markers coupled with changes in morphology and growth potential [25,26]. The transcriptomic analysis of both these cell lines and SCLC tumors has enabled the categorization of SCLC into four molecular subtypes, determined by the expression of a range of transcription factors that specify cell lineage: ASCL1, NEUROD1 (Neuronal Differentiation 1), POU2F3 (POU Class 2 Homeobox 3), or YAP1 (Yes1 Associated Transcriptional Regulator). The distinctions among the different transcriptional profiles of these subtypes involve different levels of neuroendocrine differentiation and disparities in their metabolic traits [3].
The ASCL1 subtype (SCLC-A), which comprises about 70% of SCLC cases, is characterized by predominant expression of ASCL1, a transcription factor critical for the normal development of pulmonary neuroendocrine cells (PNECs) [17]. ASCL1 also plays a vital role in the growth and survival of in vitro cell lines and the development of high-grade neuroendocrine lung tumors in murine models [27]. ASCL1 triggers the expression of oncogenes such as L-MYC, RET, SOX2, and NFIB (Nuclear Factor I B), in addition to other anti-apoptotic genes such as BCL2 (B-cell Lymphoma 2) and genes responsible for the neuroendocrine phenotype [28]. Importantly, the Notch signaling pathway negatively regulates ASCL1 expression, establishing an antagonistic relationship between the activation of the Notch signaling pathway and ASCL1 expression [29].
NEUROD1 also plays a crucial role in development of PNECs, and is the main driving transcription factor for 17% of SCLC tumors [28]. Note that there is a subset of tumors with varying levels of dual expression of ASCL1 and NEUROD1, usually with a predominance of ASCL1 expression and activity. The NEUROD1 subtype (SCLC-N) shows high expression of neuroendocrine markers, although to a lesser extent than SCLC-A [28]. NEUROD1 significantly contributes to the promotion of malignant traits, enhancing the survival and migratory abilities of tumor cells, and activates oncogenes such as C-MYC [27].
The YAP1 subtype (SCLC-Y), accounting for approximately 7% of SCLC tumors, is characterized by predominant expression of YAP1, a critical mediator in the activation of the Hippo signaling pathway. The existence of this subtype has sparked controversy, as several groups have failed to identify tumors exclusively expressing YAP1 [30,31,32]. Its association with an inflammatory phenotype has led to the proposal of the SCLC-I (inflamed) subtype to replace SCLC-Y. This subtype exhibits the highest levels of immune cell infiltration into the tumor, and is identified by a mesenchymal phenotype and a non-neuroendocrine profile with low or absent expression of ASCL1 and NEUROD1 [33,34].
The POU2F3 subtype (SCLC-P), representing 7% of SCLC tumors, is characterized by the absence of expression of ASCL1, NEUROD1, and neuroendocrine markers [28]. POU2F3 is a transcription factor essential for the differentiation of tuft cells, a rare type of chemosensory cells found in the pulmonary epithelium. These cells, similar to neuroendocrine cells, respond to external stimuli by releasing bioactive substances that regulate the functions of local immune and epithelial cells [35]. This subtype is rarely found in tumors originating from PNECs [28]. Similar to the SCLC-Y subtype, it exhibits a non-neuroendocrine phenotype and has been associated with an inflammatory phenotype [30].
Recently, it has been demonstrated that these subpopulations can coexist within the same tumor as a result of lineage plasticity, with the ASCL1 subtype evolving into the NEUROD1 subtype and the latter into the YAP1 subtype [17,36], implying that these phenotypes are different phases within the same evolutionary process rather than distinct subtypes. This transformation from a neuroendocrine to a non-neuroendocrine state appears to be orchestrated by a set of critical genes and cellular programs such as the previously stated C-MYC (which activates Notch), Hippo, and TGFβ (Transforming Growth Factor β) signaling pathways as well as genes involved in the epithelial-to-mesenchymal transition (EMT) [17,26,37] (Figure 2). This continuum of plastic cell states illustrates the capacity of SCLC to trade off cancer hallmark-related tasks, enabling the cells to adapt to different microenvironments. This transforming capacity has been hypothesized to derive from the dysregulation of the innate plasticity of normal PNECs, which have been shown to differentiate in response to lung injury [38,39]. Other factors that may induce cell plasticity include tumor treatment and changes in microenvironment [38].
Plasticity should be considered an important target for SCLC treatment. The presence of multiple cellular subpopulations and their interactions in SCLC could explain the ability of these cells to acquire chemoresistance and high metastatic potential. The high degree of heterogeneity in SCLC highlights the need to develop new precision therapies specific to each subtype, as they present different degrees of neuroendocrine differentiation and distinctive alterations that imply different therapeutic vulnerabilities [40,41].

2. Actionable Drivers in SCLC

2.1. Targeting Transcription Factors and Epigenomic Regulators

LSD1 (also known as KDM1A) and KDM5A are lysine-specific histone demethylases involved in the regulation of the Notch signaling pathway. LSD1 is overexpressed in SCLC [42]. Iadademstat (ORY-1001) is an LSD1 inhibitor in phase 2 clinical trials in leukemia (NCT05546580) that has been shown to reactivate Notch signaling, thereby suppressing ASCL1 activity and reducing tumor growth in SCLC mouse models [43]. Another LSD1 targeted drug, GSK2879552, has been tested in SCLC in combination with a PD-1 inhibitor in a syngeneic model of SCLC, resulting in strong tumor growth inhibition [44]. However, the clinical trial conducted with GSK2879552 was terminated, as the risk–benefit assessment did not favor continuation [45]. KDM5A represses Notch signaling to sustain neuroendocrine differentiation and promote tumorigenesis. Concomitant inhibition of KDM5A and LSD1 with KDM5-C70 and ORY-1001 molecules, respectively, has been shown to synergistically repress ASCL1 and suppress proliferation of SCLC cell lines [46].
Another epigenetic factor that has emerged as a significant target in SCLC is EZH2 (Enhancer of Zeste Homolog 2), the enzymatic subunit of the Polycomb Repressive Complex 2 (PRC2). This histone methyltransferase is frequently overexpressed in this cancer type, and promotes SCLC progression by suppressing the TGFβ-Smad-ASCL1 pathway [47]. Various EZH2 inhibitors are under investigation for SCLC and other tumor types. A number of these inhibitors have progressed to clinical trials in SCLC, such as DS-3201b (NCT03879798), PF-06821497 (NCT03460977), and XNW5004 (NCT06022757). These molecules may be relevant in SCLC, as evidence suggests that they enhance standard cytotoxic therapy and prevent the development of resistance both in vitro and in vivo [48].
As previously mentioned, C-MYC activates Notch, and MYC family genes are amplified and overexpressed in SCLC [46]. The lack of an accessible enzymatic active site in this protein has entailed a challenge in the search for MYC targeting strategies [49]. A dominant-negative form that sequesters MYC away from DNA has been developed, exhibiting anti-tumoral efficacy in SCLC in vitro and in NSCLC models in vivo [50,51]. Other C-MYC repressors, such as the BET (Bromodomain and Extra-Terminal Domain) inhibitor JQ1, have shown good results in different cancer types, and could be tested in SCLC patients that express high C-MYC levels [52,53]. C-MYC has not only been studied as a target, but also as a biomarker. C-MYC-driven variants are vulnerable to Aurora kinase inhibition [54] and Chk1 (Checkpoint Kinase 1) inhibitors [55]. Considering that C-MYC amplification is characteristic of NEUROD1-positive tumors, therapies targeting C-MYC would be of most benefit in these patients [56].
The transcription factor CREB imparts neuroendocrine features and has been found to be elevated in SCLC tumors. CREB activity is related to cell proliferation, and its blockade with PKA inhibitors abolished the development of SCLC in an in vivo model [57]. As mentioned above, the acetyltransferase CREBBP is a frequently mutated gene in SCLC and its inactivation is associated with the SCLC-A subtype. Jia et al. demonstrated that knocking out CREBBP reduces transcription of cellular adhesion genes and drives tumorigenesis [58]. These effects could be restored with HDAC inhibitors such as pracinostat or tinostamustine, breaking new ground in the treatment of ASCL1-positive tumors [59].
ELF3 (E74-Like Factor 3) is an ETS transcription factor that exhibits context-dependent functionality, serving both as a lineage-dependent oncogene and as a tumor-suppressive gene [60,61]. In SCLC, ELF3 acts as an oncogenic regulator in ASCL1-positive cells. Because AURKA (Aurora Kinase A) is an ELF3 target gene, the use of AURKA inhibitors has been proposed for patients with high ELF3 expression [33]. Upstream inhibition of this factor using auranofin (a PKC inhibitor) leads to cell death in NSCLC [62]; thus, its use could be a potential therapy for patients with high levels of ELF3 in SCLC. Recently, the transcription factors ETV4 (ETS Variant Transcription Factor 4) and ETV5 (ETS Variant Transcription Factor 5) have been identified as key mediators of SCLC progression. Considering that both are downstream transcriptional effectors of FGFR (Fibroblast Growth Factor Receptor) signaling, the pan-FGFR inhibitor LY2874455 has emerged as a potential therapeutic candidate for SCLC treatment. LY2874455 not only decreases ETV4 and ETV5 protein expression, it blocks the downstream MAPK and PI3K-AKT signaling pathways. In pre-clinical studies, LY2874455 has demonstrated efficacy as an anti-tumor agent when used alone or in combination with cisplatin-etoposide. These findings highlight the potential of LY2874455 as a promising therapeutic approach for SCLC [63].
The signal transducer and activator of transcription 3 (STAT3) is often activated in SCLC; this is partly attributed to tumor-associated macrophages (TAMs). Consequently, the crosstalk between SCLC cells and TAMs is being considered as a potential target for SCLC treatment. Onionin A, a natural compound isolated from onion, has shown the ability to counteract the influence of macrophage-derived soluble factors, including IL-6, resulting in effective suppression of the SCLC cell proliferation both in vitro and in vivo [64].

2.2. Targeting DNA Damage Response Proteins

Lurbinectedin, aforementioned as one of the second-line therapies, is a synthetic marine-derived anticancer agent that acts as a selective inhibitor of oncogenic transcription [65]. This drug binds covalently to GC-rich regions and blocks the DNA repair mechanism, resulting in RNA Pol II elongation arrest and degradation [66]. Lurbinectedin selectively targets the CpG islands located downstream of the transcriptional start site of ASCL1 and NEUROD1, leading to suppression of these genes and their targets, including BCL2, INSM1 (Insulinoma-associated 1), C-MYC, and AURKA [67]. This drug was approved by the FDA for treatment of patients with metastatic SCLC after showing favorable results in a phase 2 basket trial of relapsed SCLC patients [68,69]. Although the phase 3 ATLANTIS trial demonstrated that a combination of lurbinectedin and doxorubicin did not improve OS compared to control therapy, this combination shows a safer hematological profile [65].
The high mutational burden of SCLC makes the DNA damage repair (DDR) pathway an attractive target for tumor control. The DNA repair protein PARP1 (Poly (ADP Ribose) Polymerase 1) was identified as a target in SCLC in a proteomic analysis [70]. This seminal study showed that PARP expression is higher in both SCLC cells and patients compared to NSCLC and demonstrated the high sensitivity of many SCLC cell lines to treatment with PARP inhibitors (PARPi), although sensitivity to these drugs was not universal for all cell lines tested [70,71]. The PARPi veliparib has been shown to be mostly ineffective when used as single-agent therapy; however, it potentiates standard chemotherapy in SCLC patients [72]. Veliparib combined with temozolomide, an oral alkylating agent that induces O6-alkyl-guanine lesions on DNA, has shown significant improvement in these patients, especially in those with SLFN11 (Schlafen 11)-expressing tumors [73]. SLFN11 is a putative DNA/RNA helicase that induces irreversible replication blockade. The ongoing NCT04334941 clinical trial aims to test the efficacy of the PARPi talazoparib plus immunotherapy based on SLFN11 status. SLFN11 serves as a predictive biomarker in a wide range of cancer therapies that cause DNA damage, including platinum salts, topoisomerase I/II inhibitors, and lurbinectedin [74,75]. The epigenetic inactivation of SLFN11-mediated resistance to DNA-targeted agents in different cancers has been bypassed with class I HDAC (Histone DeACetylases) inhibitors [76]. Other PARPis, such as olaparib and talazoparib (BMN-673), have been evaluated in SCLC as well [77,78].
Another potential target is the primary activator of the replication stress response, ATR (Ataxia Telangiectasia and Rad3-related). Its inhibitor, M6620 (berzosertib), has been shown to be effective in pre-clinical studies when used in combination with other SCLC therapies [79]. In combination with topotecan, it has a cytotoxic effect and activates the STING (STimulator of INterferon Genes) pathway, implying that ATR-TOP1 inhibition may improve response to immune checkpoint therapy in STING-low SCLC tumors [80]. In addition, synergy of M6620 with lurbinectedin has been demonstrated in multiple SCLC cell lines, organoids, and in vivo models [81], and their combined effectiveness is currently being assessed in a clinical trial (NCT04802174). Finally, other M6620 combinations are under investigation in SCLC, such as a combination with the Trop2 inhibitor molecule sacituzumab-govitecan (NCT04826341), a transmembrane glycoprotein that has emerged as a promising molecular target for lung cancer treatment [82].

2.3. Kinase Inhibitors

Continuing with the DDR pathway, certain kinases have emerged as compelling targets for SCLC. Indeed, one of the most studied family of molecules for SCLC treatment is the Aurora family of serine/threonine kinases (AURK). These enzymes are frequently found to be overexpressed in cancer, and play a key role in the regulation of mitosis and chromosome alignment [83]. It has been shown that MYC family-driven cell lines, in particular the NEUROD1 high subtype with C-MYC amplification or overexpression, are sensitive to Aurora K inhibitors [54,84]. Barasertib (AZD1152), an AURKB inhibitor, has shown efficacy suppressing growth of a subset of SCLC cell lines both in vitro and in vivo [85]. An extensive safety evaluation of this molecule and its derivatives has been conducted in clinical trials for advanced solid tumors, including patients with SCLC [86]. Additionally, Alisertib (MLN8237), a selective inhibitor of AURKA, has demonstrated improvement in PFS when combined with paclitaxel compared to paclitaxel alone in patients with SCLC tumors positive for C-MYC [87]. As a caveat, Aurora K inhibitor effectiveness and safety must be carefully validated in phase 3 clinical trials, as many patients, despite showing an initial good response, have had to discontinue treatment due to tumor progression or adverse effects, particularly those linked to blood malignancies [88].
As mentioned before, after DNA damage and in the absence of p53 suppressor activity, the ATR kinase activates the cell-cycle checkpoint kinase Chk1, acting as a principal mediator of cell cycle arrest in the G2/M phase [89]. Chk1 is highly expressed in SCLC lines compared to NSCLC [55]. Another study has shown that Chk1 expression is directly correlated with poor OS among SCLC patients [90]. In pre-clinical SCLC models, the Chk1 inhibitors prexasertib (LY2606368) and AZD7762 sensitize cells resistant to cisplatin [91]. Sen et al. demonstrated that the effectiveness of Chk1 inhibition using prexasertib was particularly pronounced in cells with C-MYC amplification or overexpression [55]. Although prexasertib displays strong efficacy in pre-clinical studies as a single agent, in clinical trials it did not achieve sufficient activity and exhibited limitations in terms of bioavailability [92]. As a result, novel Chk1 inhibitors such as SRA-737 are currently being tested in patients with advanced solid tumors, including SCLC [93]. Of note, it has been observed that targeting Chk1 in combination with chemotherapy enhances the effect of PD-L1 blockade by modulating the immune microenvironment, providing compelling rationale for combining Chk1 inhibitors with chemotherapy and ICIs [94].
Another regulator of the G2/M checkpoint, preventing the initiation of mitosis in the presence of DNA damage, is the evolutionarily highly conserved kinase WEE1. This kinase has become a target of interest in SCLC, and the first-in-class WEE1 inhibitor adavosertib (AZD1775) has demonstrated a good safety and tolerability profile in a phase 1 study conducted in patients with advanced solid tumors including SCLC [95]. WEE1 expression levels are directly correlated with drug resistance in SCLC cell lines. Thus, WEE1 inhibitors have the potential for synergistic interactions with various other treatment agents, underscoring their significance in combinatorial therapies for SCLC. For example, upregulation of WEE1 represents a mechanism of acquired resistance to Chk1 inhibitors in SCLC [89]. Adavosertib in combination with the PARPi olaparib enhances the efficacy of olaparib in SCLC circulating tumor cell patient-derived xenografts (CDX) [96], and a clinical trial with this combination in SCLC has been conducted under the identifier NCT02511795. Finally, WEE1 inhibition activates the STING-TBK1-IRF3 and STAT1 pathways, resulting in elevated expression of IFN-γ and PD-L1. Thus, the combination of adavosertib and PD-L1 blockade induces tumor regression, activation of type I and II interferon pathways, and infiltration of cytotoxic T cells in SCLC mouse models, representing a promising immunotherapeutic approach in SCLC [97].
Cyclin-dependent kinase 7 (CDK7) plays a pivotal role in controlling both cell cycle and gene transcription. Zhang et al. demonstrated that the selective CDK7 inhibitor YKL-5-124 induces DNA replication stress response and genome instability. Simultaneously, it activates immune response signaling, triggering a T cell-mediated surveillance system. Consistent with this, YKL-5-124 has shown improved survival in several mouse models of SCLC when combined with anti-PD-1 therapy [98]. Sun et al. showed a potential synergy between the CDK7 inhibitor THZ1 and topotecan. Inhibition of CDK7 induces RNA polymerase II proteosomal degradation. Pol II depletion prevents transcription-coupled ubiquitin-proteasome-mediated repair of topotecan-induced DNA lesions, leading to enhanced cell killing [99].
Src (c-Src) is a non-receptor tyrosine kinase that is frequently overexpressed in SCLC and plays a significant role in motility, cell adhesion, and angiogenesis. The Src inhibitor dasatinib was tested in clinical trials, showing poor results in SCLC (NCT00470054). However, the compound KC-180-2, with a dual mechanism that reduces Src phosphorylation and inhibits tubulin polymerization, was recently shown to reduce SCLC cell viability both in vitro and in vivo [100]. Additionally, YES1, a non-receptor tyrosine kinase that belongs to the Src family of kinases, has been proposed as a druggable oncogenic target in SCLC. The YES1 inhibitor CH6953755 has shown anti-tumor activity in YES1-positive organoid models as well as in cell line- and patient-derived xenografts (PDX) [101].

2.4. Targeting Metabolic Pathways

Heightened glucose uptake is a well-known characteristic of cancer cells, including SCLC. Consequently, research has been dedicated to investigating and targeting pathways related to glucose metabolism [102]. SCLC is distinguished by increased glutamine anabolism, which supports the synthesis of nucleic acids, thereby facilitating tumor cell proliferation. Pre-clinical studies have demonstrated that SCLC exhibits high sensitivity to the simultaneous inhibition of both salvage and de novo pathways of purine nucleotide synthesis using 6-mercaptopurine (6-MP) and methotrexate (MTX). Currently employed in clinical practice for specific types of leukemia, repurposing of this combinatory treatment for SCLC appears to be a viable prospect in the future [103]. On the other hand, SCLC tumors exhibit increased sensitivity to disruption of the pyrimidine biosynthesis pathway. Pharmacological inhibition of dihydroorotate dehydrogenase (DHODH), a crucial enzyme in this pathway, through the use of brequinar markedly reduced the viability of SCLC cells in vitro and suppressed tumor growth in vivo. This promising approach could be employed as a stand-alone agent or in combination with standard treatments for SCLC [104].
Additionally, it has been demonstrated that the MEK5/ERK5 axis regulates lipid metabolism, including the mevalonate pathway, which is responsible for cholesterol synthesis in SCLC. Depletion of MEK5/ERK5 sensitizes SCLC cells to pharmacological inhibition of the mevalonate pathway by statins, suggesting that these two kinases may be therapeutic targets in SCLC [105]. Several pre-clinical studies are in progress to investigate inhibitory molecules for these targets, such as BIX02189, which has already been studied in NSCLC but remains unexplored in SCLC [106].
In addition, Chalishazar et al. have demonstrated that different subtypes of SCLC exhibit unique metabolic vulnerabilities. The depletion of arginine using pegylated arginine deiminase (ADI-PEG20) significantly suppresses tumor growth and enhances the survival of mice, specifically in MYC-driven tumors. These findings suggest that arginine deprivation represents a subtype-specific therapeutic vulnerability [107]. Although a phase 2 clinical trial (NCT01266018) was terminated due to lack of efficacy when ADI-PEG20 was used as a single agent, an ongoing trial is currently evaluating the effects of combined treatment with gemcitabine and docetaxel in SCLC patients (NCT05616624).
In addition to its function regulating the intrinsic pathway of apoptosis, BCL2 regulates mitochondrial bioenergetics [108]. BCL2 is frequently upregulated in SCLC [109]. BCL2 inhibitors such as venetoclax [110], navitoclax (NCT03366103) [111], oblimersen sodium (G31399) [112], and palcitoclax (NCT03366103) show modest results when combined with other therapies [113], and enhance sensitivity to AURKB inhibition in pre-clinical SCLC models [114].

2.5. Other Targetable Proteins and Lipids Involved in SCLC

A number of cell surface proteins have been shown to play important roles in SCLC tumorigenesis. One such protein is Somatostatin receptor 2 (SSTR2), which promotes tumor growth and survival [115]. Targeting SSTR2 with the miniaturized drug conjugate PEN-221 leads to cell-cycle arrest and apoptosis, resulting in tumor regression in multiple SCLC SSTR2-expressing human xenograft models [116]. CD47 is another cell-surface molecule that acts as an anti-phagocytic signal. CD47 blockade has been shown to enhance the anti-SCLC effects of radiotherapy in murine and human pre-clinical models and to stimulate abscopal effects, inhibiting the growth of distant non-irradiated tumors [117]. There are other ADC (Antibody–Drug Conjugate) surface targets being explored in SCLC research, such as NRXN1 [118] and SEZ6 [119]. Notably, the ADC against SEZ6 ABBV-011 has advanced to a phase 1 clinical trial for relapsed SCLC (NCT03639194).
Karyopherin β1 (KPNB1) has been identified as a cytoplasmic-to-nuclear transport receptor for ASCL1 and NEUROD1 in SCLC. Inhibition of KPNB1 results in cytoplasmic accumulation and impaired transcriptional activity of both transcription factors. Pharmacologic targeting of KPNB1 using the small molecule INI-43 preferentially disrupted the growth of SCLC-A and SCLC-N cells in in vitro studies and effectively suppressed the growth of SCLC-A PDX tumors in vivo. These findings highlight the therapeutic potential of targeting KPNB1 in particular subtypes of SCLC [120].
The monosialoganglioside fucosyl-GM1 is a biomarker for SCLC [121]. BMS-986012 is a monoclonal antibody against fucosyl-GM1 that is currently being tested in a phase 1/2 SCLC clinical trial (NCT04702880). Previous studies with this antibody have shown a good tolerability profile and promising results in combination with the ICI nivolumab [122].

2.6. Targeting Pathways and Biological Processes

Developmental pathways such as Hedgehog, Notch, and Wnt are closely associated with SCLC pathogenesis. Although pre-clinical studies have demonstrated that Hedgehog inhibition can suppress tumor growth in vivo [123], a phase 2 trial conducted with vismodegib (GDC-0449) in combination with chemotherapy did not reach the OS and PFS endpoints in patients with SCLC [87].
Tarextumab (OMP-59R5) is a Notch2/Notch3 antagonist; in combination with platinum-etoposide, it did not improve PFS in SCLC patients (NCT01859741). Other Notch targeting strategies include inhibition of the cell-surface Notch ligand DLL3. Recently, the use of modified CAR-T cells to recognize this ligand has emerged as a personalized therapy for SCLC patients: Tarlatamab (AMG 757) has demonstrated manageable safety and response durability in patients with relapsed/refractory SCLC [124]; HPN328 and BI764532 are currently being evaluated (NCT04471727, NCT05963867).
Wnt signaling activation has been described as a mechanism of chemoresistance in relapsed SCLC [125]. Wnt5A, a ligand of β-catenin-independent noncanonical Wnt pathways, has been identified as a promoter of SCLC neoplastic transformation and cell proliferation through activation of RHOA (Ras Homolog Family Member A). Rhosin, a selective inhibitor of GTPase activity among the RHOA subfamily, has demonstrated dose-dependent effects on SCLC viability [126]. This small molecule has been tested in various cancer animal models, exhibiting optimal safety profiles and significant efficacy [127].
Blockade of the PI3K/AKT/mTOR pathway has shown encouraging pre-clinical results, indicating a potential strategy for SCLC treatment through a combination of clinically approved PI3K and mTOR inhibitors [128]. The activation of this pathway has been described as a potential mechanism underlying therapeutic resistance [129]; thus, this approach holds promise for reversing resistance to standard therapy in SCLC. Interestingly, pharmacological inhibition of the PI3K/AKT pathway with the AKT inhibitor samotolisib delayed tumor growth and neuroendocrine transformation in an EGFR-mutant lung adenocarcinoma PDX model [130].
The MET/HGF (Hepatocyte Growth Factor) axis plays an important role in regulating cell motility in SCLC. Consequently, various MET inhibitors have been tested to target this pathway. Pre-clinical data indicate that inhibiting the MET pathway can reverse chemoresistance and impede tumor growth [131]. However, combination of MET inhibitors with chemotherapy has yielded unsatisfactory results in clinical trials [132,133]. Intriguingly, modulation of MET appears to enhance the efficacy of standard checkpoint inhibitors. Nevertheless, comprehensive pre-clinical and clinical data evaluating the combination of immunotherapy and MET inhibitors in SCLC is still missing [134].
c-Kit expression and activity are upregulated in various cancers, including SCLC [135,136]. Pre-clinical studies have demonstrated that imatinib, which inhibits c-Kit kinase activity, enhanced chemotherapy induced growth inhibition in SCLC models [137]. However, efficacy could not be demonstrated in clinical trials [138,139]. Consequently, novel strategies, such as coupling the ADC-targeting c-Kit with the microtubule inhibitor DM1, have been developed. This antibody has demonstrated significant anti-tumor activity both in vitro and in vivo and exhibits a synergistic effect when combined with carboplatin-etoposide [136].
The NFIB/CARM1 pathway plays a pivotal oncogenic role in SCLC, facilitating metastasis and regulating chromatin accessibility. Unlike NFIB, CARM1 activity can be effectively suppressed by several small molecule inhibitors that have recently become available. The CARM1 inhibitor TP-064 has been tested in SCLC PDX models, yielding the most significant reduction in tumor volume when combined with cisplatin-etoposide chemotherapy [140].
Targeting the ubiquitin-protease system (UPS) has emerged as a novel therapeutic strategy in various tumors, including SCLC. UBA1 (Ubiquitin-Like Modifier Activating Enzyme 1) is highly expressed in microcytic tumors and can be inhibited by TAK-243. SCLC cell lines and PDX models are sensitive to this agent, and it exhibits a synergistic effect when combined with chemotherapy, olaparib, and radiotherapy [141].

2.7. Angiogenesis Related Therapies

Angiogenesis is a mechanism that tumors often employ to sustain uncontrolled growth and avoid immune surveillance. Many tumor types benefit from therapies based on inhibition of the vasculogenesis regulator VEGF (Vascular Endothelial Growth Factor) and/or its receptor VEGFR. The use of the VEGFR inhibitor anlotinib has been approved as a third-line treatment in China, where a clinical trial demonstrated improved PFS and OS compared to placebo and a favorable safety profile [142]. Several phase 3 and phase 4 clinical trials have recently explored or are currently exploring the use of anlotinib in combination with other therapies such as chemotherapy (NCT03890055) (NCT04073550) and immunotherapy (NCT04192682). In addition, the combination of anlotinib with both chemotherapy and immunotherapy has been tested (NCT04234607). Other VEGFR inhibitors such as sunitinib and aflibercept (NCT00828139) have shown good and modest results, respectively. Pazopanib, an oral angiogenesis inhibitor targeting VEGFR, PDGFR, and c-Kit, has shown significant efficacy, prolonging PFS in extensive disease as maintenance therapy following chemotherapy [143]. Finally, the combination of the VEGFR2 inhibitor apatinib with the immunotherapeutic agent camrelizumab has demonstrated anti-tumor activity in patients who previously failed platinum-based chemotherapy [144]. Similar results for extensive-stage SCLC were achieved with the combination of the monoclonal antibody against VEGF bevacizumab and dual immunotherapy against PD-1 and CTLA-4 [145].

3. Conclusions and Future Directions

For decades, SCLC has been considered an undruggable cancer. However, the identification of distinct molecular subtypes, each with unique vulnerabilities, has opened a range of possibilities for repurposing or developing drugs against specific targets. Many of these agents have already been tested, and although to date there are still no approved targeted drugs for SCLC treatment, several have shown improved outcomes, especially when used in combination with chemo- or immunotherapies. In this review, we have compiled the most recent and promising pre-clinical studies (Table 1) and clinical trials (Table 2) in the field. Hopefully, with these efforts, precision medicine will revolutionize treatment for this universally lethal disease in the near future.

Author Contributions

Conceptualization, I.J.E. and M.R.; writing—original draft preparation, M.G., I.Z. and M.R.; writing—review and editing, M.R.F., I.J.E. and M.R.; supervision, I.J.E. and M.R.; funding acquisition, M.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Spanish Ministry of Science, Innovation, and Universities, grant number PID2019-109577RA-I00/AEI/10.13039/501100011033; the Spanish Ministry of Science, Innovation, and Universities (Ramón y Cajal Program), grant number RYC-2018-023874-I (M.R.); 2021/22 Predoctoral Fellowship UPNA-RYC program (M.G.); the Department of Health, Government of Navarre, grant number 026/2022.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sung, H.; Ferlay, J.; Siegel, R.L.; Laversanne, M.; Soerjomataram, I.; Jemal, A.; Bray, F. Global Cancer Statistics 2020: GLOBOCAN Estimates of Incidence and Mortality Worldwide for 36 Cancers in 185 Countries. CA Cancer J. Clin. 2021, 71, 209–249. [Google Scholar] [CrossRef] [PubMed]
  2. Ferlay, J.; Colombet, M.; Soerjomataram, I.; Parkin, D.M.; Piñeros, M.; Znaor, A.; Bray, F. Cancer Statistics for the Year 2020: An Overview. Int. J. Cancer 2021, 149, 778–789. [Google Scholar] [CrossRef] [PubMed]
  3. Rudin, C.M.; Brambilla, E.; Faivre-Finn, C.; Sage, J. Small-Cell Lung Cancer. Nat. Rev. Dis. Primers 2021, 7, 3. [Google Scholar] [CrossRef] [PubMed]
  4. Varghese, A.M.; Zakowski, M.F.; Yu, H.A.; Won, H.H.; Riely, G.J.; Krug, L.M.; Kris, M.G.; Rekhtman, N.; Ladanyi, M.; Wang, L.; et al. Small-Cell Lung Cancers in Patients Who Never Smoked Cigarettes. J. Thorac. Oncol. 2014, 9, 892–896. [Google Scholar] [CrossRef] [PubMed]
  5. Basumallik, N.; Agarwal, M. Small Cell Lung Cancer. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2023. [Google Scholar]
  6. Dingemans, A.-M.C.; Früh, M.; Ardizzoni, A.; Besse, B.; Faivre-Finn, C.; Hendriks, L.E.; Lantuejoul, S.; Peters, S.; Reguart, N.; Rudin, C.M.; et al. Small-Cell Lung Cancer: ESMO Clinical Practice Guidelines for Diagnosis, Treatment and Follow-Up. Ann. Oncol. 2021, 32, 839–853. [Google Scholar] [CrossRef]
  7. Kalemkerian, G.P.; Akerley, W.; Bogner, P.; Borghaei, H.; Chow, L.Q.; Downey, R.J.; Gandhi, L.; Ganti, A.K.P.; Govindan, R.; Grecula, J.C.; et al. Small Cell Lung Cancer. J. Natl. Compr. Cancer Netw. 2013, 11, 78–98. [Google Scholar] [CrossRef]
  8. Lim, J.U.; Kang, H.S. A Narrative Review of Current and Potential Prognostic Biomarkers for Immunotherapy in Small-Cell Lung Cancer. Ann. Transl. Med. 2021, 9, 809. [Google Scholar] [CrossRef]
  9. Hellmann, M.D.; Callahan, M.K.; Awad, M.M.; Calvo, E.; Ascierto, P.A.; Atmaca, A.; Rizvi, N.A.; Hirsch, F.R.; Selvaggi, G.; Szustakowski, J.D.; et al. Tumor Mutational Burden and Efficacy of Nivolumab Monotherapy and in Combination with Ipilimumab in Small-Cell Lung Cancer. Cancer Cell 2018, 33, 853–861.e4. [Google Scholar] [CrossRef]
  10. 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]
  11. Goldman, J.W.; Dvorkin, M.; Chen, Y.; Reinmuth, N.; Hotta, K.; Trukhin, D.; Statsenko, G.; Hochmair, M.J.; Özgüroğlu, M.; Ji, J.H.; et al. Durvalumab, with or without Tremelimumab, plus Platinum–Etoposide versus Platinum–Etoposide Alone in First-Line Treatment of Extensive-Stage Small-Cell Lung Cancer (CASPIAN): Updated Results from a Randomised, Controlled, Open-Label, Phase 3 Trial. Lancet Oncol. 2021, 22, 51–65. [Google Scholar] [CrossRef]
  12. George, J.; Lim, J.S.; Jang, S.J.; Cun, Y.; Ozretić, L.; Kong, G.; Leenders, F.; Lu, X.; Fernández-Cuesta, L.; Bosco, G.; et al. Comprehensive Genomic Profiles of Small Cell Lung Cancer. Nature 2015, 524, 47–53. [Google Scholar] [CrossRef] [PubMed]
  13. Kim, K.-B.; Dunn, C.T.; Park, K.-S. Recent Progress in Mapping the Emerging Landscape of the Small-Cell Lung Cancer Genome. Exp. Mol. Med. 2019, 51, 1–13. [Google Scholar] [CrossRef] [PubMed]
  14. Kim, D.-W.; Kim, K.-C.; Kim, K.-B.; Dunn, C.T.; Park, K.-S. Transcriptional Deregulation Underlying the Pathogenesis of Small Cell Lung Cancer. Transl. Lung Cancer Res. 2018, 7, 4–20. [Google Scholar] [CrossRef] [PubMed]
  15. Peifer, M.; Fernández-Cuesta, L.; Sos, M.L.; George, J.; Seidel, D.; Kasper, L.H.; Plenker, D.; Leenders, F.; Sun, R.; Zander, T.; et al. Integrative Genome Analyses Identify Key Somatic Driver Mutations of Small-Cell Lung Cancer. Nat. Genet. 2012, 44, 1104–1110. [Google Scholar] [CrossRef] [PubMed]
  16. Umemura, S.; Mimaki, S.; Makinoshima, H.; Tada, S.; Ishii, G.; Ohmatsu, H.; Niho, S.; Yoh, K.; Matsumoto, S.; Takahashi, A.; et al. Therapeutic Priority of the PI3K/AKT/MTOR Pathway in Small Cell Lung Cancers as Revealed by a Comprehensive Genomic Analysis. J. Thorac. Oncol. 2014, 9, 1324–1331. [Google Scholar] [CrossRef] [PubMed]
  17. Ireland, A.S.; Micinski, A.M.; Kastner, D.W.; Guo, B.; Wait, S.J.; Spainhower, K.B.; Conley, C.C.; Chen, O.S.; Guthrie, M.R.; Soltero, D.; et al. MYC Drives Temporal Evolution of Small Cell Lung Cancer Subtypes by Reprogramming Neuroendocrine Fate. Cancer Cell 2020, 38, 60–78.e12. [Google Scholar] [CrossRef] [PubMed]
  18. Voigt, E.; Wallenburg, M.; Wollenzien, H.; Thompson, E.; Kumar, K.; Feiner, J.; McNally, M.; Friesen, H.; Mukherjee, M.; Afeworki, Y.; et al. Sox2 Is an Oncogenic Driver of Small-Cell Lung Cancer and Promotes the Classic Neuroendocrine Subtype. Mol. Cancer Res. 2021, 19, 2015–2025. [Google Scholar] [CrossRef]
  19. Park, K.-S.; Liang, M.-C.; Raiser, D.M.; Zamponi, R.; Roach, R.R.; Curtis, S.J.; Walton, Z.; Schaffer, B.E.; Roake, C.M.; Zmoos, A.-F.; et al. Characterization of the Cell of Origin for Small Cell Lung Cancer. Cell Cycle 2011, 10, 2806–2815. [Google Scholar] [CrossRef]
  20. Augustyn, A.; Borromeo, M.; Wang, T.; Fujimoto, J.; Shao, C.; Dospoy, P.D.; Lee, V.; Tan, C.; Sullivan, J.P.; Larsen, J.E.; et al. ASCL1 Is a Lineage Oncogene Providing Therapeutic Targets for High-Grade Neuroendocrine Lung Cancers. Proc. Natl. Acad. Sci. USA 2014, 111, 14788–14793. [Google Scholar] [CrossRef]
  21. Yang, D.; Denny, S.K.; Greenside, P.G.; Chaikovsky, A.C.; Brady, J.J.; Ouadah, Y.; Granja, J.M.; Jahchan, N.S.; Lim, J.S.; Kwok, S.; et al. Intertumoral Heterogeneity in SCLC Is Influenced by the Cell Type of Origin. Cancer Discov. 2018, 8, 1316–1331. [Google Scholar] [CrossRef]
  22. Ferone, G.; Lee, M.C.; Sage, J.; Berns, A. Cells of Origin of Lung Cancers: Lessons from Mouse Studies. Genes Dev. 2020, 34, 1017–1032. [Google Scholar] [CrossRef] [PubMed]
  23. Offin, M.; Guo, R.; Wu, S.L.; Sabari, J.; Land, J.D.; Ni, A.; Montecalvo, J.; Halpenny, D.F.; Buie, L.W.; Pak, T.; et al. Immunophenotype and Response to Immunotherapy of RET-Rearranged Lung Cancers. JCO Precis. Oncol. 2019, 3, 1–8. [Google Scholar] [CrossRef] [PubMed]
  24. Pizzutilo, E.G.; Pedrani, M.; Amatu, A.; Ruggieri, L.; Lauricella, C.; Veronese, S.M.; Signorelli, D.; Cerea, G.; Giannetta, L.; Siena, S.; et al. Liquid Biopsy for Small Cell Lung Cancer Either De Novo or Transformed: Systematic Review of Different Applications and Meta-Analysis. Cancers 2021, 13, 2265. [Google Scholar] [CrossRef] [PubMed]
  25. Scarpa, S.; Morstyn, G.; Carney, D.N.; Modesti, A.; Triche, T.J. Small Cell Lung Cancer Cell Lines: Pure and Variant Types Can Be Distinguished by Their Extracellular Matrix Synthesis. Eur. Respir. J. 1988, 1, 639–644. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, W.; Girard, L.; Zhang, Y.-A.; Haruki, T.; Papari-Zareei, M.; Stastny, V.; Ghayee, H.K.; Pacak, K.; Oliver, T.G.; Minna, J.D.; et al. Small Cell Lung Cancer Tumors and Preclinical Models Display Heterogeneity of Neuroendocrine Phenotypes. Transl. Lung Cancer Res. 2018, 7, 32–49. [Google Scholar] [CrossRef] [PubMed]
  27. Borromeo, M.D.; Savage, T.K.; Kollipara, R.K.; He, M.; Augustyn, A.; Osborne, J.K.; Girard, L.; Minna, J.D.; Gazdar, A.F.; Cobb, M.H.; et al. ASCL1 and NEUROD1 Reveal Heterogeneity in Pulmonary Neuroendocrine Tumors and Regulate Distinct Genetic Programs. Cell Rep. 2016, 16, 1259–1272. [Google Scholar] [CrossRef] [PubMed]
  28. Baine, M.K.; Hsieh, M.-S.; Lai, W.V.; Egger, J.V.; Jungbluth, A.A.; Daneshbod, Y.; Beras, A.; Spencer, R.; Lopardo, J.; Bodd, F.; et al. SCLC Subtypes Defined by ASCL1, NEUROD1, POU2F3, and YAP1: A Comprehensive Immunohistochemical and Histopathologic Characterization. J. Thorac. Oncol. 2020, 15, 1823–1835. [Google Scholar] [CrossRef] [PubMed]
  29. Sriuranpong, V.; Borges, M.W.; Strock, C.L.; Nakakura, E.K.; Watkins, D.N.; Blaumueller, C.M.; Nelkin, B.D.; Ball, D.W. Notch Signaling Induces Rapid Degradation of Achaete-Scute Homolog 1. Mol. Cell. Biol. 2002, 22, 3129–3139. [Google Scholar] [CrossRef]
  30. Qu, S.; Fetsch, P.; Thomas, A.; Pommier, Y.; Schrump, D.S.; Miettinen, M.M.; Chen, H. Molecular Subtypes of Primary SCLC Tumors and Their Associations With Neuroendocrine and Therapeutic Markers. J. Thorac. Oncol. 2022, 17, 141–153. [Google Scholar] [CrossRef]
  31. Poirier, J.T.; George, J.; Owonikoko, T.K.; Berns, A.; Brambilla, E.; Byers, L.A.; Carbone, D.; Chen, H.J.; Christensen, C.L.; Dive, C.; et al. New Approaches to SCLC Therapy: From the Laboratory to the Clinic. J. Thorac. Oncol. 2020, 15, 520–540. [Google Scholar] [CrossRef]
  32. Rudin, C.M.; Poirier, J.T.; Byers, L.A.; Dive, C.; Dowlati, A.; George, J.; Heymach, J.V.; Johnson, J.E.; Lehman, J.M.; MacPherson, D.; et al. Molecular Subtypes of Small Cell Lung Cancer: A Synthesis of Human and Mouse Model Data. Nat. Rev. Cancer 2019, 19, 289–297. [Google Scholar] [CrossRef]
  33. Horie, M.; Tanaka, H.; Suzuki, M.; Sato, Y.; Takata, S.; Takai, E.; Miyashita, N.; Saito, A.; Nakatani, Y.; Yachida, S. An Integrative Epigenomic Approach Identifies ELF3 as an Oncogenic Regulator in ASCL1 -positive Neuroendocrine Carcinoma. Cancer Sci. 2023, 114, 2596–2608. [Google Scholar] [CrossRef] [PubMed]
  34. Owonikoko, T.K.; Dwivedi, B.; Chen, Z.; Zhang, C.; Barwick, B.; Ernani, V.; Zhang, G.; Gilbert-Ross, M.; Carlisle, J.; Khuri, F.R.; et al. YAP1 Expression in SCLC Defines a Distinct Subtype With T-Cell–Inflamed Phenotype. J. Thorac. Oncol. 2021, 16, 464–476. [Google Scholar] [CrossRef] [PubMed]
  35. Huang, Y.-H.; Klingbeil, O.; He, X.-Y.; Wu, X.S.; Arun, G.; Lu, B.; Somerville, T.D.D.; Milazzo, J.P.; Wilkinson, J.E.; Demerdash, O.E.; et al. POU2F3 Is a Master Regulator of a Tuft Cell-like Variant of Small Cell Lung Cancer. Genes Dev. 2018, 32, 915–928. [Google Scholar] [CrossRef] [PubMed]
  36. Shue, Y.T.; Lim, J.S.; Sage, J. Tumor Heterogeneity in Small Cell Lung Cancer Defined and Investigated in Pre-Clinical Mouse Models. Transl. Lung Cancer Res. 2018, 7, 21–31. [Google Scholar] [CrossRef] [PubMed]
  37. Horie, M.; Saito, A.; Ohshima, M.; Suzuki, H.I.; Nagase, T. YAP and TAZ Modulate Cell Phenotype in a Subset of Small Cell Lung Cancer. Cancer Sci. 2016, 107, 1755–1766. [Google Scholar] [CrossRef] [PubMed]
  38. Groves, S.M.; Ildefonso, G.V.; McAtee, C.O.; Ozawa, P.M.M.; Ireland, A.S.; Stauffer, P.E.; Wasdin, P.T.; Huang, X.; Qiao, Y.; Lim, J.S.; et al. Archetype Tasks Link Intratumoral Heterogeneity to Plasticity and Cancer Hallmarks in Small Cell Lung Cancer. Cell Systems 2022, 13, 690–710.e17. [Google Scholar] [CrossRef] [PubMed]
  39. Ouadah, Y.; Rojas, E.R.; Riordan, D.P.; Capostagno, S.; Kuo, C.S.; Krasnow, M.A. Rare Pulmonary Neuroendocrine Cells Are Stem Cells Regulated by Rb, P53, and Notch. Cell 2019, 179, 403–416.e23. [Google Scholar] [CrossRef]
  40. Frese, K.K.; Simpson, K.L.; Dive, C. Small Cell Lung Cancer Enters the Era of Precision Medicine. Cancer Cell 2021, 39, 297–299. [Google Scholar] [CrossRef]
  41. Gay, C.M.; Stewart, C.A.; Park, E.M.; Diao, L.; Groves, S.M.; Heeke, S.; Nabet, B.Y.; Fujimoto, J.; Solis, L.M.; Lu, W.; et al. Patterns of Transcription Factor Programs and Immune Pathway Activation Define Four Major Subtypes of SCLC with Distinct Therapeutic Vulnerabilities. Cancer Cell 2021, 39, 346–360.e7. [Google Scholar] [CrossRef]
  42. Yan, W.; Chung, C.; Xie, T.; Ozeck, M.; Nichols, T.C.; Frey, J.; Udyavar, A.R.; Sharma, S.; Paul, T.A. Intrinsic and Acquired Drug Resistance to LSD1 Inhibitors in Small Cell Lung Cancer Occurs through a TEAD4-driven Transcriptional State. Mol. Oncol. 2022, 16, 1309–1328. [Google Scholar] [CrossRef] [PubMed]
  43. Augert, A.; Eastwood, E.; Ibrahim, A.H.; Wu, N.; Grunblatt, E.; Basom, R.; Liggitt, D.; Eaton, K.D.; Martins, R.; Poirier, J.T.; et al. Targeting NOTCH Activation in Small Cell Lung Cancer through LSD1 Inhibition. Sci. Signal. 2019, 12, eaau2922. [Google Scholar] [CrossRef] [PubMed]
  44. Hiatt, J.B.; Sandborg, H.; Garrison, S.M.; Arnold, H.U.; Liao, S.-Y.; Norton, J.P.; Friesen, T.J.; Wu, F.; Sutherland, K.D.; Rienhoff, H.Y.; et al. Inhibition of LSD1 with Bomedemstat Sensitizes Small Cell Lung Cancer to Immune Checkpoint Blockade and T-Cell Killing. Clin. Cancer Res. 2022, 28, 4551–4564. [Google Scholar] [CrossRef] [PubMed]
  45. Bauer, T.M.; Besse, B.; Martinez-Marti, A.; Trigo, J.M.; Moreno, V.; Garrido, P.; Ferron-Brady, G.; Wu, Y.; Park, J.; Collingwood, T.; et al. Phase I, Open-Label, Dose-Escalation Study of the Safety, Pharmacokinetics, Pharmacodynamics, and Efficacy of GSK2879552 in Relapsed/Refractory SCLC. J. Thorac. Oncol. 2019, 14, 1828–1838. [Google Scholar] [CrossRef] [PubMed]
  46. Oser, M.G.; Sabet, A.H.; Gao, W.; Chakraborty, A.A.; Schinzel, A.C.; Jennings, R.B.; Fonseca, R.; Bonal, D.M.; Booker, M.A.; Flaifel, A.; et al. The KDM5A/RBP2 Histone Demethylase Represses NOTCH Signaling to Sustain Neuroendocrine Differentiation and Promote Small Cell Lung Cancer Tumorigenesis. Genes Dev. 2019, 33, 1718–1738. [Google Scholar] [CrossRef] [PubMed]
  47. Murai, F.; Koinuma, D.; Shinozaki-Ushiku, A.; Fukayama, M.; Miyaozono, K.; Ehata, S. EZH2 Promotes Progression of Small Cell Lung Cancer by Suppressing the TGF-β-Smad-ASCL1 Pathway. Cell Discov. 2015, 1, 15026. [Google Scholar] [CrossRef] [PubMed]
  48. Gardner, E.E.; Lok, B.H.; Schneeberger, V.E.; Desmeules, P.; Miles, L.A.; Arnold, P.K.; Ni, A.; Khodos, I.; de Stanchina, E.; Nguyen, T.; et al. Chemosensitive Relapse in Small Cell Lung Cancer Proceeds through an EZH2-SLFN11 Axis. Cancer Cell 2017, 31, 286–299. [Google Scholar] [CrossRef] [PubMed]
  49. Beaulieu, M.-E.; Jauset, T.; Massó-Vallés, D.; Martínez-Martín, S.; Rahl, P.; Maltais, L.; Zacarias-Fluck, M.F.; Casacuberta-Serra, S.; Serrano del Pozo, E.; Fiore, C.; et al. Intrinsic Cell-Penetrating Activity Propels Omomyc from Proof of Concept to Viable Anti-MYC Therapy. Sci. Transl. Med. 2019, 11, eaar5012. [Google Scholar] [CrossRef]
  50. Fiorentino, F.P.; Tokgün, E.; Solé-Sánchez, S.; Giampaolo, S.; Tokgün, O.; Jauset, T.; Kohno, T.; Perucho, M.; Soucek, L.; Yokota, J. Growth Suppression by MYC Inhibition in Small Cell Lung Cancer Cells with TP53 and RB1 Inactivation. Oncotarget 2016, 7, 31014–31028. [Google Scholar] [CrossRef]
  51. Massó-Vallés, D.; Soucek, L. Blocking Myc to Treat Cancer: Reflecting on Two Decades of Omomyc. Cells 2020, 9, 883. [Google Scholar] [CrossRef]
  52. Chen, H.; Liu, H.; Qing, G. Targeting Oncogenic Myc as a Strategy for Cancer Treatment. Signal Transduct. Target. Ther. 2018, 3, 5. [Google Scholar] [CrossRef] [PubMed]
  53. Delmore, J.E.; Issa, G.C.; Lemieux, M.E.; Rahl, P.B.; Shi, J.; Jacobs, H.M.; Kastritis, E.; Gilpatrick, T.; Paranal, R.M.; Qi, J.; et al. BET Bromodomain Inhibition as a Therapeutic Strategy to Target C-Myc. Cell 2011, 146, 904–917. [Google Scholar] [CrossRef] [PubMed]
  54. Mollaoglu, G.; Guthrie, M.R.; Böhm, S.; Brägelmann, J.; Can, I.; Ballieu, P.M.; Marx, A.; George, J.; Heinen, C.; Chalishazar, M.D.; et al. MYC Drives Progression of Small Cell Lung Cancer to a Variant Neuroendocrine Subtype with Vulnerability to Aurora Kinase Inhibition. Cancer Cell 2017, 31, 270–285. [Google Scholar] [CrossRef] [PubMed]
  55. Sen, T.; Tong, P.; Stewart, C.A.; Cristea, S.; Valliani, A.; Shames, D.S.; Redwood, A.B.; Fan, Y.H.; Li, L.; Glisson, B.S.; et al. CHK1 Inhibition in Small-Cell Lung Cancer Produces Single-Agent Activity in Biomarker-Defined Disease Subsets and Combination Activity with Cisplatin or Olaparib. Cancer Res. 2017, 77, 3870–3884. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, H.; Gesumaria, L.; Park, Y.-K.; Oliver, T.G.; Singer, D.S.; Ge, K.; Schrump, D.S. BET Inhibitors Target the SCLC-N Subtype of Small-Cell Lung Cancer by Blocking NEUROD1 Transactivation. Mol. Cancer Res. 2023, 21, 91–101. [Google Scholar] [CrossRef] [PubMed]
  57. Xia, Y.; Zhan, C.; Feng, M.; Leblanc, M.; Ke, E.; Yeddula, N.; Verma, I.M. Targeting CREB Pathway Suppresses Small Cell Lung Cancer. Mol. Cancer Res. 2018, 16, 825–832. [Google Scholar] [CrossRef] [PubMed]
  58. Jia, D.; Augert, A.; Kim, D.-W.; Eastwood, E.; Wu, N.; Ibrahim, A.H.; Kim, K.-B.; Dunn, C.T.; Pillai, S.P.S.; Gazdar, A.F.; et al. Crebbp Loss Drives Small Cell Lung Cancer and Increases Sensitivity to HDAC Inhibition. Cancer Discov. 2018, 8, 1422–1437. [Google Scholar] [CrossRef]
  59. Megyesfalvi, Z.; Gay, C.M.; Popper, H.; Pirker, R.; Ostoros, G.; Heeke, S.; Lang, C.; Hoetzenecker, K.; Schwendenwein, A.; Boettiger, K.; et al. Clinical Insights into Small Cell Lung Cancer: Tumor Heterogeneity, Diagnosis, Therapy, and Future Directions. CA Cancer J. Clin. 2023, 73, 620–652. [Google Scholar] [CrossRef]
  60. Suzuki, M.; Saito-Adachi, M.; Arai, Y.; Fujiwara, Y.; Takai, E.; Shibata, S.; Seki, M.; Rokutan, H.; Maeda, D.; Horie, M.; et al. E74-like Factor 3 Is a Key Regulator of Epithelial Integrity and Immune Response Genes in Biliary Tract Cancer. Cancer Res. 2021, 81, 489–500. [Google Scholar] [CrossRef]
  61. Enfield, K.S.S.; Marshall, E.A.; Anderson, C.; Ng, K.W.; Rahmati, S.; Xu, Z.; Fuller, M.; Milne, K.; Lu, D.; Shi, R.; et al. Epithelial Tumor Suppressor ELF3 Is a Lineage-Specific Amplified Oncogene in Lung Adenocarcinoma. Nat. Commun. 2019, 10, 5438. [Google Scholar] [CrossRef]
  62. Lee, J.-S.; Choi, Y.E.; Kim, S.; Han, J.-Y.; Goh, S.-H. ELF3 Is a Target That Promotes Therapeutic Efficiency in EGFR Tyrosine Kinase Inhibitor-Resistant Non-Small Cell Lung Cancer Cells via Inhibiting PKCί. Int. J. Mol. Sci. 2021, 22, 12287. [Google Scholar] [CrossRef] [PubMed]
  63. Shia, D.W.; Choi, W.; Vijayaraj, P.; Vuong, V.; Sandlin, J.M.; Lu, M.M.; Aziz, A.; Marin, C.; Aros, C.J.; Sen, C.; et al. Targeting PEA3 Transcription Factors to Mitigate Small Cell Lung Cancer Progression. Oncogene 2023, 42, 434–448. [Google Scholar] [CrossRef] [PubMed]
  64. Mito, R.; Iriki, T.; Fujiwara, Y.; Pan, C.; Ikeda, T.; Nohara, T.; Suzuki, M.; Sakagami, T.; Komohara, Y. Onionin A Inhibits Small-Cell Lung Cancer Proliferation through Suppressing STAT3 Activation Induced by Macrophages-Derived IL-6 and Cell–Cell Interaction with Tumor-Associated Macrophage. Human Cell 2023, 36, 1068–1080. [Google Scholar] [CrossRef]
  65. Aix, S.P.; Ciuleanu, T.E.; Navarro, A.; Cousin, S.; Bonanno, L.; Smit, E.F.; Chiappori, A.; Olmedo, M.E.; Horvath, I.; Grohé, C.; et al. Combination Lurbinectedin and Doxorubicin versus Physician’s Choice of Chemotherapy in Patients with Relapsed Small-Cell Lung Cancer (ATLANTIS): A Multicentre, Randomised, Open-Label, Phase 3 Trial. Lancet Respir. Med. 2023, 11, 74–86. [Google Scholar] [CrossRef] [PubMed]
  66. Santamaría Nuñez, G.; Robles, C.M.G.; Giraudon, C.; Martínez-Leal, J.F.; Compe, E.; Coin, F.; Aviles, P.; Galmarini, C.M.; Egly, J.-M. Lurbinectedin Specifically Triggers the Degradation of Phosphorylated RNA Polymerase II and the Formation of DNA Breaks in Cancer Cells. Mol. Cancer Ther. 2016, 15, 2399–2412. [Google Scholar] [CrossRef] [PubMed]
  67. Costanzo, F.; Martínez Diez, M.; Santamaría Nuñez, G.; Díaz-Hernandéz, J.I.; Genes Robles, C.M.; Díez Pérez, J.; Compe, E.; Ricci, R.; Li, T.; Coin, F.; et al. Promoters of ASCL1- and NEUROD1-dependent Genes Are Specific Targets of Lurbinectedin in SCLC Cells. EMBO Mol. Med. 2022, 14, e14841. [Google Scholar] [CrossRef] [PubMed]
  68. Singh, S.; Jaigirdar, A.A.; Mulkey, F.; Cheng, J.; Hamed, S.S.; Li, Y.; Liu, J.; Zhao, H.; Goheer, A.; Helms, W.S.; et al. FDA Approval Summary: Lurbinectedin for the Treatment of Metastatic Small Cell Lung Cancer. Clin. Cancer Res. 2021, 27, 2378–2382. [Google Scholar] [CrossRef] [PubMed]
  69. Trigo, J.; Subbiah, V.; Besse, B.; Moreno, V.; López, R.; Sala, M.A.; Peters, S.; Ponce, S.; Fernández, C.; Alfaro, V.; et al. Lurbinectedin as Second-Line Treatment for Patients with Small-Cell Lung Cancer: A Single-Arm, Open-Label, Phase 2 Basket Trial. Lancet Oncol. 2020, 21, 645–654. [Google Scholar] [CrossRef]
  70. Byers, L.A.; Wang, J.; Nilsson, M.B.; Fujimoto, J.; Saintigny, P.; Yordy, J.; Giri, U.; Peyton, M.; Fan, Y.H.; Diao, L.; et al. Proteomic Profiling Identifies Dysregulated Pathways in Small Cell Lung Cancer and Novel Therapeutic Targets Including PARP1. Cancer Discov. 2012, 2, 798–811. [Google Scholar] [CrossRef]
  71. Barayan, R.; Ran, X.; Lok, B.H. PARP Inhibitors for Small Cell Lung Cancer and Their Potential for Integration into Current Treatment Approaches. J. Thorac. Dis. 2020, 12, 6240–6252. [Google Scholar] [CrossRef]
  72. Owonikoko, T.K.; Dahlberg, S.E.; Sica, G.L.; Wagner, L.I.; Wade, J.L.; Srkalovic, G.; Lash, B.W.; Leach, J.W.; Leal, T.B.; Aggarwal, C.; et al. Randomized Phase II Trial of Cisplatin and Etoposide in Combination With Veliparib or Placebo for Extensive-Stage Small-Cell Lung Cancer: ECOG-ACRIN 2511 Study. J. Clin. Oncol. 2019, 37, 222–229. [Google Scholar] [CrossRef] [PubMed]
  73. Pietanza, M.C.; Waqar, S.N.; Krug, L.M.; Dowlati, A.; Hann, C.L.; Chiappori, A.; Owonikoko, T.K.; Woo, K.M.; Cardnell, R.J.; Fujimoto, J.; et al. Randomized, Double-Blind, Phase II Study of Temozolomide in Combination With Either Veliparib or Placebo in Patients With Relapsed-Sensitive or Refractory Small-Cell Lung Cancer. J. Clin. Oncol. 2018, 36, 2386–2394. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, B.; Ramkumar, K.; Cardnell, R.J.; Gay, C.M.; Stewart, C.A.; Wang, W.-L.; Fujimoto, J.; Wistuba, I.I.; Byers, L.A. A Wake-up Call for Cancer DNA Damage: The Role of Schlafen 11 (SLFN11) across Multiple Cancers. Br. J. Cancer 2021, 125, 1333–1340. [Google Scholar] [CrossRef] [PubMed]
  75. Kundu, K.; Cardnell, R.J.; Zhang, B.; Shen, L.; Stewart, C.A.; Ramkumar, K.; Cargill, K.R.; Wang, J.; Gay, C.M.; Byers, L.A. SLFN11 Biomarker Status Predicts Response to Lurbinectedin as a Single Agent and in Combination with ATR Inhibition in Small Cell Lung Cancer. Transl. Lung Cancer Res. 2021, 10, 4095–4105. [Google Scholar] [CrossRef] [PubMed]
  76. Tang, S.-W.; Thomas, A.; Murai, J.; Trepel, J.B.; Bates, S.E.; Rajapakse, V.N.; Pommier, Y. Overcoming Resistance to DNA-Targeted Agents by Epigenetic Activation of Schlafen 11 ( SLFN11) Expression with Class I Histone Deacetylase Inhibitors. Clin. Cancer Res. 2018, 24, 1944–1953. [Google Scholar] [CrossRef] [PubMed]
  77. Krebs, M.G.; Delord, J.-P.; Jeffry Evans, T.R.; De Jonge, M.; Kim, S.-W.; Meurer, M.; Postel-Vinay, S.; Lee, J.-S.; Angell, H.K.; Rocher-Ros, V.; et al. Olaparib and Durvalumab in Patients with Relapsed Small Cell Lung Cancer (MEDIOLA): An Open-Label, Multicenter, Phase 1/2, Basket Study. Lung Cancer 2023, 180, 107216. [Google Scholar] [CrossRef] [PubMed]
  78. Laird, J.H.; Lok, B.H.; Ma, J.; Bell, A.; de Stanchina, E.; Poirier, J.T.; Rudin, C.M. Talazoparib Is a Potent Radiosensitizer in Small Cell Lung Cancer Cell Lines and Xenografts. Clin. Cancer Res. 2018, 24, 5143–5152. [Google Scholar] [CrossRef]
  79. Thomas, A.; Takahashi, N.; Rajapakse, V.N.; Zhang, X.; Sun, Y.; Ceribelli, M.; Wilson, K.M.; Zhang, Y.; Beck, E.; Sciuto, L.; et al. Therapeutic Targeting of ATR Yields Durable Regressions in Small Cell Lung Cancers with High Replication Stress. Cancer Cell 2021, 39, 566–579.e7. [Google Scholar] [CrossRef]
  80. Li, X.; Li, Y.; Zhao, Z.; Miao, N.; Liu, G.; Deng, L.; Wei, S.; Hou, J. Immunogenicity of Small-cell Lung Cancer Associates with STING Pathway Activation and Is Enhanced by ATR and TOP1 Inhibition. Cancer Med. 2023, 12, 4864–4881. [Google Scholar] [CrossRef]
  81. Schultz, C.W.; Zhang, Y.; Elmeskini, R.; Zimmermann, A.; Fu, H.; Murai, Y.; Wangsa, D.; Kumar, S.; Takahashi, N.; Atkinson, D.; et al. ATR Inhibition Augments the Efficacy of Lurbinectedin in Small-cell Lung Cancer. EMBO Mol. Med. 2023, 15, e17313. [Google Scholar] [CrossRef]
  82. Inamura, K.; Yokouchi, Y.; Kobayashi, M.; Ninomiya, H.; Sakakibara, R.; Subat, S.; Nagano, H.; Nomura, K.; Okumura, S.; Shibutani, T.; et al. Association of Tumor TROP2 Expression with Prognosis Varies among Lung Cancer Subtypes. Oncotarget 2017, 8, 28725–28735. [Google Scholar] [CrossRef] [PubMed]
  83. Gautschi, O.; Heighway, J.; Mack, P.C.; Purnell, P.R.; Lara, P.N.; Gandara, D.R. Aurora Kinases as Anticancer Drug Targets. Clin. Cancer Res. 2008, 14, 1639–1648. [Google Scholar] [CrossRef] [PubMed]
  84. Hook, K.E.; Garza, S.J.; Lira, M.E.; Ching, K.A.; Lee, N.V.; Cao, J.; Yuan, J.; Ye, J.; Ozeck, M.; Shi, S.T.; et al. An Integrated Genomic Approach to Identify Predictive Biomarkers of Response to the Aurora Kinase Inhibitor PF-03814735. Mol. Cancer Ther. 2012, 11, 710–719. [Google Scholar] [CrossRef] [PubMed]
  85. Helfrich, B.A.; Kim, J.; Gao, D.; Chan, D.C.; Zhang, Z.; Tan, A.-C.; Bunn, P.A. Barasertib (AZD1152), a Small Molecule Aurora B Inhibitor, Inhibits the Growth of SCLC Cell Lines In Vitro and In Vivo. Mol. Cancer Ther. 2016, 15, 2314–2322. [Google Scholar] [CrossRef] [PubMed]
  86. Johnson, M.L.; Wang, J.S.; Falchook, G.; Greenlees, C.; Jones, S.; Strickland, D.; Fabbri, G.; Kennedy, C.; Elizabeth Pease, J.; Sainsbury, L.; et al. Safety, Tolerability, and Pharmacokinetics of Aurora Kinase B Inhibitor AZD2811: A Phase 1 Dose-Finding Study in Patients with Advanced Solid Tumours. Br. J. Cancer 2023, 128, 1906–1915. [Google Scholar] [CrossRef] [PubMed]
  87. Owonikoko, T.K.; Niu, H.; Nackaerts, K.; Csoszi, T.; Ostoros, G.; Mark, Z.; Baik, C.; Joy, A.A.; Chouaid, C.; Jaime, J.C.; et al. Randomized Phase II Study of Paclitaxel plus Alisertib versus Paclitaxel plus Placebo as Second-Line Therapy for SCLC: Primary and Correlative Biomarker Analyses. J. Thorac. Oncol. 2020, 15, 274–287. [Google Scholar] [CrossRef] [PubMed]
  88. Stefani, A.; Piro, G.; Schietroma, F.; Strusi, A.; Vita, E.; Fiorani, S.; Barone, D.; Monaca, F.; Sparagna, I.; Valente, G.; et al. Unweaving the Mitotic Spindle: A Focus on Aurora Kinase Inhibitors in Lung Cancer. Front. Oncol. 2022, 12, 1026020. [Google Scholar] [CrossRef]
  89. Patil, M.; Pabla, N.; Dong, Z. Checkpoint Kinase 1 in DNA Damage Response and Cell Cycle Regulation. Cell. Mol. Life Sci. 2013, 70, 4009–4021. [Google Scholar] [CrossRef]
  90. Hsu, W.-H.; Zhao, X.; Zhu, J.; Kim, I.-K.; Rao, G.; McCutcheon, J.; Hsu, S.-T.; Teicher, B.; Kallakury, B.; Dowlati, A.; et al. Checkpoint Kinase 1 Inhibition Enhances Cisplatin Cytotoxicity and Overcomes Cisplatin Resistance in SCLC by Promoting Mitotic Cell Death. J. Thorac. Oncol. 2019, 14, 1032–1045. [Google Scholar] [CrossRef]
  91. Zhao, X.; Kim, I.; Kallakury, B.; Chahine, J.J.; Iwama, E.; Pierobon, M.; Petricoin, E.; McCutcheon, J.N.; Zhang, Y.; Umemura, S.; et al. Acquired Small Cell Lung Cancer Resistance to Chk1 Inhibitors Involves Wee1 Up-regulation. Mol. Oncol. 2021, 15, 1130–1145. [Google Scholar] [CrossRef]
  92. Byers, L.A.; Navarro, A.; Schaefer, E.; Johnson, M.; Özgüroğlu, M.; Han, J.-Y.; Bondarenko, I.; Cicin, I.; Dragnev, K.H.; Abel, A.; et al. A Phase II Trial of Prexasertib (LY2606368) in Patients With Extensive-Stage Small-Cell Lung Cancer. Clin. Lung Cancer 2021, 22, 531–540. [Google Scholar] [CrossRef] [PubMed]
  93. Jones, R.; Plummer, R.; Moreno, V.; Carter, L.; Roda, D.; Garralda, E.; Kristeleit, R.; Sarker, D.; Arkenau, T.; Roxburgh, P.; et al. A Phase I/II Trial of Oral SRA737 (a Chk1 Inhibitor) Given in Combination with Low-Dose Gemcitabine in Patients with Advanced Cancer. Clin. Cancer Res. 2023, 29, 331–340. [Google Scholar] [CrossRef] [PubMed]
  94. Sen, T.; Della Corte, C.M.; Milutinovic, S.; Cardnell, R.J.; Diao, L.; Ramkumar, K.; Gay, C.M.; Stewart, C.A.; Fan, Y.; Shen, L.; et al. Combination Treatment of the Oral CHK1 Inhibitor, SRA737, and Low-Dose Gemcitabine Enhances the Effect of Programmed Death Ligand 1 Blockade by Modulating the Immune Microenvironment in SCLC. J. Thorac. Oncol. 2019, 14, 2152–2163. [Google Scholar] [CrossRef] [PubMed]
  95. Bauer, T.M.; Moore, K.N.; Rader, J.S.; Simpkins, F.; Mita, A.C.; Beck, J.T.; Hart, L.; Chu, Q.; Oza, A.; Tinker, A.V.; et al. A Phase Ib Study Assessing the Safety, Tolerability, and Efficacy of the First-in-Class Wee1 Inhibitor Adavosertib (AZD1775) as Monotherapy in Patients with Advanced Solid Tumors. Target. Oncol. 2023, 18, 517–530. [Google Scholar] [CrossRef] [PubMed]
  96. Lallo, A.; Frese, K.K.; Morrow, C.J.; Sloane, R.; Gulati, S.; Schenk, M.W.; Trapani, F.; Simms, N.; Galvin, M.; Brown, S.; et al. The Combination of the PARP Inhibitor Olaparib and the WEE1 Inhibitor AZD1775 as a New Therapeutic Option for Small Cell Lung Cancer. Clin. Cancer Res. 2018, 24, 5153–5164. [Google Scholar] [CrossRef] [PubMed]
  97. Taniguchi, H.; Caeser, R.; Chavan, S.S.; Zhan, Y.A.; Chow, A.; Manoj, P.; Uddin, F.; Kitai, H.; Qu, R.; Hayatt, O.; et al. WEE1 Inhibition Enhances the Antitumor Immune Response to PD-L1 Blockade by the Concomitant Activation of STING and STAT1 Pathways in SCLC. Cell Rep. 2022, 39, 110814. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, H.; Christensen, C.L.; Dries, R.; Oser, M.G.; Deng, J.; Diskin, B.; Li, F.; Pan, Y.; Zhang, X.; Yin, Y.; et al. CDK7 Inhibition Potentiates Genome Instability Triggering Anti-Tumor Immunity in Small Cell Lung Cancer. Cancer Cell 2020, 37, 37–54.e9. [Google Scholar] [CrossRef] [PubMed]
  99. Sun, Y.; Zhang, Y.; Schultz, C.W.; Pommier, Y.; Thomas, A. CDK7 Inhibition Synergizes with Topoisomerase I Inhibition in Small Cell Lung Cancer Cells by Inducing Ubiquitin-Mediated Proteolysis of RNA Polymerase II. Mol. Cancer Ther. 2022, 21, 1430–1438. [Google Scholar] [CrossRef]
  100. Peng, J.; Zeng, Y.; Hu, X.; Huang, S.; Gao, X.; Tian, D.; Tian, S.; Qiu, L.; Liu, J.; Dong, R.; et al. KC-180-2 Exerts Anti-SCLC Effects via Dual Inhibition of Tubulin Polymerization and Src Signaling. ACS Omega 2022, 7, 32164–32175. [Google Scholar] [CrossRef]
  101. Redin, E.; Garrido-Martin, E.M.; Valencia, K.; Redrado, M.; Solorzano, J.L.; Carias, R.; Echepare, M.; Exposito, F.; Serrano, D.; Ferrer, I.; et al. YES1 Is a Druggable Oncogenic Target in SCLC. J. Thorac. Oncol. 2022, 17, 1387–1403. [Google Scholar] [CrossRef]
  102. Nomura, M.; Morita, M.; Tanuma, N. A Metabolic Vulnerability of Small-Cell Lung Cancer. Oncotarget 2018, 9, 32278–32279. [Google Scholar] [CrossRef] [PubMed]
  103. Kodama, M.; Toyokawa, G.; Sugahara, O.; Sugiyama, S.; Haratake, N.; Yamada, Y.; Wada, R.; Takamori, S.; Shimokawa, M.; Takenaka, T.; et al. Modulation of Host Glutamine Anabolism Enhances the Sensitivity of Small Cell Lung Cancer to Chemotherapy. Cell Rep. 2023, 42, 112899. [Google Scholar] [CrossRef] [PubMed]
  104. Li, L.; Ng, S.R.; Colón, C.I.; Drapkin, B.J.; Hsu, P.P.; Li, Z.; Nabel, C.S.; Lewis, C.A.; Romero, R.; Mercer, K.L.; et al. Identification of DHODH as a Therapeutic Target in Small Cell Lung Cancer. Sci. Transl. Med. 2019, 11, eaaw7852. [Google Scholar] [CrossRef] [PubMed]
  105. Cristea, S.; Coles, G.L.; Hornburg, D.; Gershkovitz, M.; Arand, J.; Cao, S.; Sen, T.; Williamson, S.C.; Kim, J.W.; Drainas, A.P.; et al. The MEK5–ERK5 Kinase Axis Controls Lipid Metabolism in Small-Cell Lung Cancer. Cancer Res. 2020, 80, 1293–1303. [Google Scholar] [CrossRef] [PubMed]
  106. Sánchez-Fdez, A.; Re-Louhau, M.F.; Rodríguez-Núñez, P.; Ludeña, D.; Matilla-Almazán, S.; Pandiella, A.; Esparís-Ogando, A. Clinical, Genetic and Pharmacological Data Support Targeting the MEK5/ERK5 Module in Lung Cancer. NPJ Precis. Oncol. 2021, 5, 78. [Google Scholar] [CrossRef] [PubMed]
  107. Chalishazar, M.D.; Wait, S.J.; Huang, F.; Ireland, A.S.; Mukhopadhyay, A.; Lee, Y.; Schuman, S.S.; Guthrie, M.R.; Berrett, K.C.; Vahrenkamp, J.M.; et al. MYC-Driven Small-Cell Lung Cancer Is Metabolically Distinct and Vulnerable to Arginine Depletion. Clin. Cancer Res. 2019, 25, 5107–5121. [Google Scholar] [CrossRef] [PubMed]
  108. Lucantoni, F.; Salvucci, M.; Düssmann, H.; Lindner, A.U.; Lambrechts, D.; Prehn, J.H.M. BCL(X)L and BCL2 Increase the Metabolic Fitness of Breast Cancer Cells: A Single-Cell Imaging Study. Cell Death Differ. 2021, 28, 1512–1531. [Google Scholar] [CrossRef] [PubMed]
  109. Ben-Ezra, J.M.; Kornstein, M.J.; Grimes, M.M.; Krystal, G. Small Cell Carcinomas of the Lung Express the Bcl-2 Protein. Am. J. Pathol. 1994, 145, 1036–1040. [Google Scholar]
  110. Lochmann, T.L.; Floros, K.V.; Naseri, M.; Powell, K.M.; Cook, W.; March, R.J.; Stein, G.T.; Greninger, P.; Maves, Y.K.; Saunders, L.R.; et al. Venetoclax Is Effective in Small-Cell Lung Cancers with High BCL-2 Expression. Clin. Cancer Res. 2018, 24, 360–369. [Google Scholar] [CrossRef]
  111. Rudin, C.M.; Hann, C.L.; Garon, E.B.; Ribeiro de Oliveira, M.; Bonomi, P.D.; Camidge, D.R.; Chu, Q.; Giaccone, G.; Khaira, D.; Ramalingam, S.S.; et al. Phase II Study of Single-Agent Navitoclax (ABT-263) and Biomarker Correlates in Patients with Relapsed Small Cell Lung Cancer. Clin. Cancer Res. 2012, 18, 3163–3169. [Google Scholar] [CrossRef]
  112. Rudin, C.M.; Salgia, R.; Wang, X.; Hodgson, L.D.; Masters, G.A.; Green, M.; Vokes, E.E. Randomized Phase II Study of Carboplatin and Etoposide with or without the Bcl-2 Antisense Oligonucleotide Oblimersen for Extensive-Stage Small-Cell Lung Cancer: CALGB 30103. J. Clin. Oncol. 2008, 26, 870–876. [Google Scholar] [CrossRef] [PubMed]
  113. Ploumaki, I.; Triantafyllou, E.; Koumprentziotis, I.-A.; Karampinos, K.; Drougkas, K.; Karavolias, I.; Trontzas, I.; Kotteas, E.A. Bcl-2 Pathway Inhibition in Solid Tumors: A Review of Clinical Trials. Clin. Transl. Oncol. 2023, 25, 1554–1578. [Google Scholar] [CrossRef] [PubMed]
  114. Ramkumar, K.; Tanimoto, A.; Della Corte, C.M.; Stewart, C.A.; Wang, Q.; Shen, L.; Cardnell, R.J.; Wang, J.; Polanska, U.M.; Andersen, C.; et al. Targeting BCL2 Overcomes Resistance and Augments Response to Aurora Kinase B Inhibition by AZD2811 in Small Cell Lung Cancer. Clin. Cancer Res. 2023, 29, 3237–3249. [Google Scholar] [CrossRef] [PubMed]
  115. Lehman, J.M.; Hoeksema, M.D.; Staub, J.; Qian, J.; Harris, B.; Callison, J.C.; Miao, J.; Shi, C.; Eisenberg, R.; Chen, H.; et al. Somatostatin Receptor 2 Signaling Promotes Growth and Tumor Survival in Small-cell Lung Cancer. Int. J. Cancer 2019, 144, 1104–1114. [Google Scholar] [CrossRef] [PubMed]
  116. Whalen, K.A.; White, B.H.; Quinn, J.M.; Kriksciukaite, K.; Alargova, R.; Au Yeung, T.P.; Bazinet, P.; Brockman, A.; DuPont, M.M.; Oller, H.; et al. Targeting the Somatostatin Receptor 2 with the Miniaturized Drug Conjugate, PEN-221: A Potent and Novel Therapeutic for the Treatment of Small Cell Lung Cancer. Mol. Cancer Ther. 2019, 18, 1926–1936. [Google Scholar] [CrossRef] [PubMed]
  117. Nishiga, Y.; Drainas, A.P.; Baron, M.; Bhattacharya, D.; Barkal, A.A.; Ahrari, Y.; Mancusi, R.; Ross, J.B.; Takahashi, N.; Thomas, A.; et al. Radiotherapy in Combination with CD47 Blockade Elicits a Macrophage-Mediated Abscopal Effect. Nat. Cancer 2022, 3, 1351–1366. [Google Scholar] [CrossRef] [PubMed]
  118. Yotsumoto, T.; Maemura, K.; Watanabe, K.; Amano, Y.; Matsumoto, Y.; Zokumasu, K.; Ando, T.; Kawakami, M.; Kage, H.; Nakajima, J.; et al. NRXN1 as a Novel Potential Target of Antibody-Drug Conjugates for Small Cell Lung Cancer. Oncotarget 2020, 11, 3590–3600. [Google Scholar] [CrossRef] [PubMed]
  119. Wiedemeyer, W.R.; Gavrilyuk, J.; Schammel, A.; Zhao, X.; Sarvaiya, H.; Pysz, M.; Gu, C.; You, M.; Isse, K.; Sullivan, T.; et al. ABBV-011, A Novel, Calicheamicin-Based Antibody–Drug Conjugate, Targets SEZ6 to Eradicate Small Cell Lung Cancer Tumors. Mol. Cancer Ther. 2022, 21, 986–998. [Google Scholar] [CrossRef]
  120. Kelenis, D.P.; Rodarte, K.E.; Kollipara, R.K.; Pozo, K.; Choudhuri, S.P.; Spainhower, K.B.; Wait, S.J.; Stastny, V.; Oliver, T.G.; Johnson, J.E. Inhibition of Karyopherin Β1-Mediated Nuclear Import Disrupts Oncogenic Lineage-Defining Transcription Factor Activity in Small Cell Lung Cancer. Cancer Res. 2022, 82, 3058–3073. [Google Scholar] [CrossRef]
  121. Groux-Degroote, S.; Delannoy, P. Cancer-Associated Glycosphingolipids as Tumor Markers and Targets for Cancer Immunotherapy. Int. J. Mol. Sci. 2021, 22, 6145. [Google Scholar] [CrossRef]
  122. Chu, Q.; Leighl, N.B.; Surmont, V.; van Herpen, C.; Sibille, A.; Markman, B.; Clarke, S.; Juergens, R.A.; Rivera, M.A.; Andelkovic, V.; et al. BMS-986012, an Anti–Fucosyl-GM1 Monoclonal Antibody as Monotherapy or in Combination With Nivolumab in Relapsed/Refractory SCLC: Results From a First-in-Human Phase 1/2 Study. JTO Clin. Res. Rep. 2022, 3, 100400. [Google Scholar] [CrossRef] [PubMed]
  123. Park, K.-S.; Martelotto, L.G.; Peifer, M.; Sos, M.L.; Karnezis, A.N.; Mahjoub, M.R.; Bernard, K.; Conklin, J.F.; Szczepny, A.; Yuan, J.; et al. A Crucial Requirement for Hedgehog Signaling in Small Cell Lung Cancer. Nat. Med. 2011, 17, 1504–1508. [Google Scholar] [CrossRef] [PubMed]
  124. Paz-Ares, L.; Champiat, S.; Lai, W.V.; Izumi, H.; Govindan, R.; Boyer, M.; Hummel, H.-D.; Borghaei, H.; Johnson, M.L.; Steeghs, N.; et al. Tarlatamab, a First-in-Class DLL3-Targeted Bispecific T-Cell Engager, in Recurrent Small-Cell Lung Cancer: An Open-Label, Phase I Study. J. Clin. Oncol. 2023, 41, 2893–2903. [Google Scholar] [CrossRef] [PubMed]
  125. Wagner, A.H.; Devarakonda, S.; Skidmore, Z.L.; Krysiak, K.; Ramu, A.; Trani, L.; Kunisaki, J.; Masood, A.; Waqar, S.N.; Spies, N.C.; et al. Recurrent WNT Pathway Alterations Are Frequent in Relapsed Small Cell Lung Cancer. Nat. Commun. 2018, 9, 3787. [Google Scholar] [CrossRef] [PubMed]
  126. Kim, K.-B.; Kim, D.-W.; Kim, Y.; Tang, J.; Kirk, N.; Gan, Y.; Kim, B.; Fang, B.; Park, J.; Zheng, Y.; et al. WNT5A–RHOA Signaling Is a Driver of Tumorigenesis and Represents a Therapeutically Actionable Vulnerability in Small Cell Lung Cancer. Cancer Res. 2022, 82, 4219–4233. [Google Scholar] [CrossRef] [PubMed]
  127. Tsubaki, M.; Genno, S.; Takeda, T.; Matsuda, T.; Kimura, N.; Yamashita, Y.; Morii, Y.; Shimomura, K.; Nishida, S. Rhosin Suppressed Tumor Cell Metastasis through Inhibition of Rho/YAP Pathway and Expression of RHAMM and CXCR4 in Melanoma and Breast Cancer Cells. Biomedicines 2021, 9, 35. [Google Scholar] [CrossRef] [PubMed]
  128. Hung, M.-C.; Wang, W.-P.; Chi, Y.-H. AKT Phosphorylation as a Predictive Biomarker for PI3K/MTOR Dual Inhibition-Induced Proteolytic Cleavage of MTOR Companion Proteins in Small Cell Lung Cancer. Cell Biosci. 2022, 12, 122. [Google Scholar] [CrossRef] [PubMed]
  129. Jin, Y.; Chen, Y.; Tang, H.; Hu, X.; Hubert, S.M.; Li, Q.; Su, D.; Xu, H.; Fan, Y.; Yu, X.; et al. Activation of PI3K/AKT Pathway Is a Potential Mechanism of Treatment Resistance in Small Cell Lung Cancer. Clin. Cancer Res. 2022, 28, 526–539. [Google Scholar] [CrossRef]
  130. Quintanal-Villalonga, A.; Taniguchi, H.; Zhan, Y.A.; Hasan, M.M.; Chavan, S.S.; Meng, F.; Uddin, F.; Manoj, P.; Donoghue, M.T.A.; Won, H.H.; et al. Multiomic Analysis of Lung Tumors Defines Pathways Activated in Neuroendocrine Transformation. Cancer Discov. 2021, 11, 3028–3047. [Google Scholar] [CrossRef]
  131. Taniguchi, H.; Yamada, T.; Takeuchi, S.; Arai, S.; Fukuda, K.; Sakamoto, S.; Kawada, M.; Yamaguchi, H.; Mukae, H.; Yano, S. Impact of MET Inhibition on Small-cell Lung Cancer Cells Showing Aberrant Activation of the Hepatocyte Growth Factor/ MET Pathway. Cancer Sci. 2017, 108, 1378–1385. [Google Scholar] [CrossRef]
  132. Byers, L.A.; Horn, L.; Ghandi, J.; Kloecker, G.; Owonikoko, T.; Waqar, S.N.; Krzakowski, M.; Cardnell, R.J.; Fujimoto, J.; Taverna, P.; et al. A Phase 2, Open-Label, Multi-Center Study of Amuvatinib in Combination with Platinum Etoposide Chemotherapy in Platinum-Refractory Small Cell Lung Cancer Patients. Oncotarget 2017, 8, 81441–81454. [Google Scholar] [CrossRef] [PubMed]
  133. Glisson, B.; Besse, B.; Dols, M.C.; Dubey, S.; Schupp, M.; Jain, R.; Jiang, Y.; Menon, H.; Nackaerts, K.; Orlov, S.; et al. A Randomized, Placebo-Controlled, Phase 1b/2 Study of Rilotumumab or Ganitumab in Combination With Platinum-Based Chemotherapy as First-Line Treatment for Extensive-Stage Small-Cell Lung Cancer. Clin. Lung Cancer 2017, 18, 615–625.e8. [Google Scholar] [CrossRef] [PubMed]
  134. Hardy-Werbin, M.; del Rey-Vergara, R.; Galindo-Campos, M.A.; Moliner, L.; Arriola, E. MET Inhibitors in Small Cell Lung Cancer: From the Bench to the Bedside. Cancers 2019, 11, 1404. [Google Scholar] [CrossRef] [PubMed]
  135. Sekido, Y.; Obata, Y.; Ueda, R.; Hida, T.; Suyama, M.; Shimokata, K.; Ariyoshi, Y.; Takahashi, T. Preferential Expression of C-Kit Protooncogene Transcripts in Small Cell Lung Cancer. Cancer Res. 1991, 51, 2416–2419. [Google Scholar] [PubMed]
  136. Kim, K.-H.; Kim, J.-O.; Park, J.-Y.; Seo, M.-D.; Park, S.G. Antibody-Drug Conjugate Targeting c-Kit for the Treatment of Small Cell Lung Cancer. Int. J. Mol. Sci. 2022, 23, 2264. [Google Scholar] [CrossRef] [PubMed]
  137. Decaudin, D.; de Cremoux, P.; Sastre, X.; Judde, J.-G.; Nemati, F.; Tran-Perennou, C.; Fréneaux, P.; Livartowski, A.; Pouillart, P.; Poupon, M.-F. In Vivo Efficacy of STI571 in Xenografted Human Small Cell Lung Cancer Alone or Combined with Chemotherapy. Int. J. Cancer 2005, 113, 849–856. [Google Scholar] [CrossRef] [PubMed]
  138. Johnson, B.E.; Fischer, T.; Fischer, B.; Dunlop, D.; Rischin, D.; Silberman, S.; Kowalski, M.O.; Sayles, D.; Dimitrijevic, S.; Fletcher, C.; et al. Phase II Study of Imatinib in Patients with Small Cell Lung Cancer. Clin. Cancer Res. 2003, 9, 5880–5887. [Google Scholar] [PubMed]
  139. Dy, G.K.; Miller, A.A.; Mandrekar, S.J.; Aubry, M.-C.; Langdon, R.M.; Morton, R.F.; Schild, S.E.; Jett, J.R.; Adjei, A.A. A Phase II Trial of Imatinib (ST1571) in Patients with c-Kit Expressing Relapsed Small-Cell Lung Cancer: A CALGB and NCCTG Study. Ann. Oncol. 2005, 16, 1811–1816. [Google Scholar] [CrossRef]
  140. Gao, G.; Hausmann, S.; Flores, N.M.; Benitez, A.M.; Shen, J.; Yang, X.; Person, M.D.; Gayatri, S.; Cheng, D.; Lu, Y.; et al. The NFIB/CARM1 Partnership Is a Driver in Preclinical Models of Small Cell Lung Cancer. Nat. Commun. 2023, 14, 363. [Google Scholar] [CrossRef]
  141. Majeed, S.; Aparnathi, M.K.; Nixon, K.C.J.; Venkatasubramanian, V.; Rahman, F.; Song, L.; Weiss, J.; Barayan, R.; Sugumar, V.; Barghout, S.H.; et al. Targeting the Ubiquitin–Proteasome System Using the UBA1 Inhibitor TAK-243 Is a Potential Therapeutic Strategy for Small-Cell Lung Cancer. Clin. Cancer Res. 2022, 28, 1966–1978. [Google Scholar] [CrossRef]
  142. Cheng, Y.; Wang, Q.; Li, K.; Shi, J.; Liu, Y.; Wu, L.; Han, B.; Chen, G.; He, J.; Wang, J.; et al. Anlotinib vs Placebo as Third- or Further-Line Treatment for Patients with Small Cell Lung Cancer: A Randomised, Double-Blind, Placebo-Controlled Phase 2 Study. Br. J. Cancer 2021, 125, 366–371. [Google Scholar] [CrossRef]
  143. Sun, J.-M.; Lee, K.H.; Kim, B.-S.; Kim, H.-G.; Min, Y.J.; Yi, S.Y.; Yun, H.J.; Jung, S.-H.; Lee, S.-H.; Ahn, J.S.; et al. Pazopanib Maintenance after First-Line Etoposide and Platinum Chemotherapy in Patients with Extensive Disease Small-Cell Lung Cancer: A Multicentre, Randomised, Placebo-Controlled Phase II Study (KCSG-LU12-07). Br. J. Cancer 2018, 118, 648–653. [Google Scholar] [CrossRef]
  144. Fan, Y.; Zhao, J.; Wang, Q.; Huang, D.; Li, X.; Chen, J.; Fang, Y.; Duan, J.; Zhou, C.; Hu, Y.; et al. Camrelizumab Plus Apatinib in Extensive-Stage SCLC (PASSION): A Multicenter, Two-Stage, Phase 2 Trial. J. Thorac. Oncol. 2021, 16, 299–309. [Google Scholar] [CrossRef]
  145. Song, L.; Zhou, R.; Li, X.; Pan, D. Combination of Bevacizumab and Dual Immunotherapy for Extensive-Disease Small-Cell Lung Cancer: A Case Report. Immunotherapy 2021, 13, 1309–1315. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Representation of molecular alterations that play a role in SCLC progression and recurrence.
Figure 1. Representation of molecular alterations that play a role in SCLC progression and recurrence.
Ijms 25 00105 g001
Figure 2. Scheme of dynamic evolution across the different subtypes of SCLC, showing the cell plasticity drivers and frequency of occurrence for each molecular subtype or cell state. EMT: epithelial-to-mesenchymal transition.
Figure 2. Scheme of dynamic evolution across the different subtypes of SCLC, showing the cell plasticity drivers and frequency of occurrence for each molecular subtype or cell state. EMT: epithelial-to-mesenchymal transition.
Ijms 25 00105 g002
Table 1. Targetable molecules in SCLC with positive pre-clinical results. This table does not include targets and inhibitors that have already reached clinical trials.
Table 1. Targetable molecules in SCLC with positive pre-clinical results. This table does not include targets and inhibitors that have already reached clinical trials.
TargetTargeting MoleculeResultsRef.
AURKBBarasertibBarasertib has growth-inhibitory effects in some SCLC lines and suppresses tumor growth xenografts. C-MYC amplification or high gene expression or C-MYC gene signature is a useful predictive biomarker.[85]
c-SrcKC-180-2KC-180-2 suppresses SCLC lines proliferation and inhibits the growth of SCLC xenograft tumors.[100]
YES1CH6953755CH6953755 presents anti-tumor activity in organoids and in cell- and patient-derived xenografts.[101]
HPRT16-MP6-MP in combination with MTX attenuates the growth of mouse SCLC xenograft models. The glutamine synthetase inhibitor methionine sulfoximine (MSO) enhances this effect.[103]
DHODHBrequinarBrequinar reduces SCLC cells viability in vitro and suppresses tumor growth in PDX and mouse models.[104]
SSTR2PEN-221PEN-221 inhibits in vitro cellular proliferation and suppresses tumor growth in SSTR2-positive SCLC xenografts.[116]
CD47CD47-blocking antibodyCD47 blockade plus irradiation inhibits tumor growth in SCLC xenograft models, promoting abscopal effects.[117]
NRXN1ADCThe combination of a primary anti-NRXN1 monoclonal antibody and a secondary ADC exhibits anti-tumor activity in SCLC cell lines in vitro.[118]
KPNB1INI-43INI-43 disrupts SCLC-A and SCLC-N cells proliferation in vitro and suppresses the growth of SCLC-A PDX tumors.[120]
WNT5A/RHOARhosinRhosin inhibits SCLC cell proliferation in vitro.[126]
c-Kit 4C9-DM14C9-DM1 suppresses SCLC proliferation in vitro and tumor growth in a xenograft mouse model, with synergistic effects when combined with carboplatin-etoposide therapy.[136]
NFIB/CARM1TP-064TP-064 alone cannot cause cancer regression; however, in combination with cisplatin-etoposide chemotherapy, it exhibits anti-tumor activity in a SCLC xenograft model.[140]
UBA1TAK-243The combination of TAK-243 and genotoxic therapies (e.g., olaparib) exhibits a synergistic effect in cell lines and PDX models.[141]
KDM5A/RBP2KDM5-C70KDM5-C70 decreases ASCL1 levels and inhibits cellular proliferation of a subset of SCLC lines. Combining KDM5-C70 and the LSD1 inhibitor ORY-1001 synergistically suppresses ASCL1 levels.[46]
FGFR/PEA3LY2874455LY2874455 presents anti-tumor activity when used alone or in combination with cisplatin-etoposide in vitro and in SCLC xenografts.[63]
STAT3 activationOnionin ABy suppressing STAT3 activation, onionin A inhibits SCLC cells proliferation and suppresses tumor growth in a murine model.[64]
Table 2. Compilation of clinical trials conducted in SCLC patients. OS: Overall survival, PSF: Progression-free survival.
Table 2. Compilation of clinical trials conducted in SCLC patients. OS: Overall survival, PSF: Progression-free survival.
TargetTargeting MoleculeStageStudyResults
Oncogenic transcriptionLurbinectedinApprovedNCT04291937Used as a second line treatment.
Phase 3NCT05153239Ongoing; as a single agent or combined with irinotecan.
Phase 3NCT05091567Ongoing; combined with atezolizumab.
Phase 3NCT02566993Did not improve OS vs. topotecan.
Phase 2NCT05578326Ongoing; combined with trilaciclib.
Phase 2NCT05572476Ongoing; combined with durvalumab in pre-treated patients.
Phase 2NCT04607954Ongoing; combined with durvalumab.
Phase 1/2NCT04358237Ongoing; combined with pembrolizumab.
PARPVeliparibPhase 2NCT02289690Potentiates chemotherapy.
Phase 2NCT01638546Improved PFS and OS in SLFN11-positive patients.
Phase 1/2NCT01642251Demonstrated efficacy in patients with extensive stage SCLC when combined with chemotherapy.
TalazoparibPhase 2NCT04334941Ongoing; in SLFN11-positive patients.
Phase 2NCT03672773Ongoing; combined with temozolomide in low doses.
OlaparibPhase 3NCT04624204Ongoing; in combination with pembrolizumab in pre-treated patients with chemotherapy.
Phase 2NCT03009682In patients with gene mutations in the homologous recombination pathway; no results posted yet.
Phase 2NCT04538378Ongoing; in combination with durvalumab in SCLC transformed from EGFR-mutated adenocarcinomas.
Phase 2NCT05623319Ongoing; in combination with pembrolizumab.
Phase 2NCT04939662Ongoing; in combination with bevacizumab.
Phase 2NCT02498613Ongoing; in combination with cediranib in metastatic tumors.
Phase 2NCT05245994Ongoing; in combination with durvalumab.
Phase 2NCT03428607In combination with AZD6738; no results posted yet.
Phase 2NCT02937818In combination with AZD6738 in refractory patients.
Phase 1/2NCT02446704Ongoing; in recurrent patients.
Phase 1/2NCT05975944Ongoing; in combination with selinexor.
Phase 1/2NCT02769962Ongoing; in combination with EP0057.
Phase 1/2NCT02734004Ongoing; in combination with MEDI4736.
Phase 1/2NCT04728230Ongoing; in combination with carboplatin, etoposide and/or radiation.
ATRBerzosertibPhase 2NCT04768296No results posted yet.
Phase 1/2NCT04802174Ongoing; in combination with lurbinectedin.
Phase 1/2NCT04826341Ongoing; in combination with sacituzumab-govitecan.
AlisertibPhase 2NCT06095505Ongoing; in patients with extensive stage.
Phase 2NCT01045421An objective response was noted in 21% of the patients.
Chk1LY2606368Phase 2NCT02735980Availability limitations. Another compound has been developed and is currently being tested in other neoplasms.
SRA-737Phase 1/2NCT02797977Used in combination with gemcitabine and cisplatin; no results posted yet.
WEE1AZD1775Phase 2NCT02688907Ongoing; in patients with MYC amplification or CDKN2A + TP53 mutation.
Phase 2NCT02593019Used as a single agent in relapsed patients; no results posted yet.
Phase 2NCT02688907Terminated, regimen change.
Phase 2NCT02937818Ongoing; in combination carboplatin and olaparib.
ArginineADI-PEG 20Phase 2NCT01266018Terminated for lack of efficacy.
Phase 1/2NCT05616624Ongoing; in combination with gemcitabine and docetaxel.
BCR-ABL, PDGFR, c-KitImatinibPhase 2NCT00156286No results posted.
Phase 2NCT00248482No results posted.
Phase 2NCT00052949No results posted.
Phase 2NCT00193349No results posted.
EZH2DS-3201bPhase 1/2NCT03879798In combination with irinotecan; no results posted.
XNW5004Phase 1/2NCT06022757Ongoing; in combination with pembrolizumab.
VEGFR, PDGFR and c-KitPazopanibPhase 2NCT01253369Notable rate of stable disease and slightly improve of PFS compared to second-line therapies.
BCL2Oblimersen SodiumPhase 2NCT00042978No results posted yet.
GM1BMS-986012Phase 1/2NCT04702880Ongoing; in combination with carboplatin, etoposide and nivolumab.
Phase 1/2NCT02247349No results posted yet.
LSD1IadademstatPhase 2NCT05420636Ongoing; in combination with paclitaxel in relapsed patients.
DLL3AMG 757Phase 3NCT05740566Ongoing; in combination with chemotherapy.
Phase 2NCT05060016Ongoing; in relapsed patients.
HPN328Phase 1/2NCT04471727Ongoing; in patients with DDL3 expression.
Phase 1/2NCT05879978Ongoing; in combination with ezabenlimab in patients with DDL3 expression.
BI764532Phase 2NCT05882058Ongoing; different doses of the compound tested.
CREBBPTinostamustinePhase 1/2NCT03345485No results posted yet.
VEFGRSunitinibPhase 2NCT00695292No OS and PFS results posted.
Phase 2NCT00620347Approved as a second-line treatment in China.
Phase 2NCT00616109No improvements over control.
Phase 2NCT01306045Ongoing; tested among several drugs.
Phase 1/2NCT00453154Safe and improved PFS.
AnlotinibPhase 4NCT03890055In combination with chemotherapy in relapsed patients.
Phase 3NCT04073550In combination with topotecan.
Phase 3NCT04234607In combination with benmelstobart, etoposide and carboplatin. No results posted yet.
Phase 3NCT04192682In combination with sintilimab.
AfliberceptPhase 2NCT00828139In combination with topotecan.
Improved PFS but increased toxicity.
VEGFBevacizumabPhase 2/3NCT00930891In combination with chemotherapy, did not improve outcome in extensive SCLC patients.
Phase 2NCT00403403In combination with cisplatin or carboplatin plus etoposide for treatment of extensive-stage SCLC, improved PFS with an acceptable toxicity profile. No improvement in OS was observed.
Phase 2NCT04730999Ongoing; in patients with extensive disease.
Phase 2NCT04939662Ongoing; in combination with olaparib.
Phase 2NCT00698516After treatment with topotecan-bevacizumab, improvement in PFS.
Phase 2NCT05588388Ongoing; in combination with chemo-immunotherapy and atezolizumab.
Phase 2NCT001933752-Year PFS: 22% of participants. Overall response rate: 80%.
Phase 2NCT00726986Terminated due to extreme toxicity when combined with cisplatin and etoposide.
VEGFR2ApatinibPhase 3NCT03651219No results posted.
Phase 3NCT04490421No results posted.
Phase 3NCT02875457No results posted.
Phase 3NCT03100955No results posted.
Phase 2NCT03547804Successful, already in Phase 3.
Phase 2NCT04901754In combination with camrelizumab; no results posted yet.
Phase 2NCT04128800In combination with S-1; no results posted yet.
Phase 2NCT04683198Ongoing; in combination with cambralizumab, carboplatin and etoposide.
Phase 2NCT02945852No results posted.
Phase 2NCT02995187No results posted.
Phase 2NCT03135977No results posted.
Phase 2NCT03389087No results posted.
Phase 2NCT02980809No results posted.
Phase 2NCT03129698No results posted.
Phase 2NCT03417895No results posted.
Phase 2NCT04659785No results posted.
Phase 2NCT04453930No results posted.
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

Gutiérrez, M.; Zamora, I.; Freeman, M.R.; Encío, I.J.; Rotinen, M. Actionable Driver Events in Small Cell Lung Cancer. Int. J. Mol. Sci. 2024, 25, 105. https://doi.org/10.3390/ijms25010105

AMA Style

Gutiérrez M, Zamora I, Freeman MR, Encío IJ, Rotinen M. Actionable Driver Events in Small Cell Lung Cancer. International Journal of Molecular Sciences. 2024; 25(1):105. https://doi.org/10.3390/ijms25010105

Chicago/Turabian Style

Gutiérrez, Mirian, Irene Zamora, Michael R. Freeman, Ignacio J. Encío, and Mirja Rotinen. 2024. "Actionable Driver Events in Small Cell Lung Cancer" International Journal of Molecular Sciences 25, no. 1: 105. https://doi.org/10.3390/ijms25010105

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