Skip to main content
Signal Transduction and Targeted Therapy logoLink to Signal Transduction and Targeted Therapy
. 2025 Feb 10;10:37. doi: 10.1038/s41392-024-02121-7

AXL signaling in cancer: from molecular insights to targeted therapies

Monika Yadav 1,2, Akansha Sharma 3, Ketki Patne 4, Saba Tabasum 5,6, Jyoti Suryavanshi 7, Laxminarayan Rawat 6,8, Marc Machaalani 5,6, Marc Eid 5,6, Rana P Singh 1, Toni K Choueiri 5,6, Soumitro Pal 6,8,, Akash Sabarwal 6,8,
PMCID: PMC11808115  PMID: 39924521

Abstract

AXL, a member of the TAM receptor family, has emerged as a potential target for advanced-stage human malignancies. It is frequently overexpressed in different cancers and plays a significant role in various tumor-promoting pathways, including cancer cell proliferation, invasion, metastasis, epithelial–mesenchymal transition (EMT), angiogenesis, stemness, DNA damage response, acquired therapeutic resistance, immunosuppression, and inflammatory responses. Beyond oncology, AXL also facilitates viral infections, including SARS-CoV-2 and Zika highlighting its importance in both cancer and virology. In preclinical models, small-molecule kinase inhibitors targeting AXL have shown promising anti-tumorigenic potential. This review primarily focuses on the induction, regulation and biological functions of AXL in mediating these tumor-promoting pathways. We discuss a range of therapeutic strategies, including recently developed small-molecule tyrosine kinase inhibitors (TKIs), monoclonal antibodies, and antibody–drug conjugates (ADCs), anti-AXL-CAR, and combination therapies. These interventions are being examined in both preclinical and clinical studies, offering the potential for improved drug sensitivity and therapeutic efficacy. We further discuss the mechanisms of acquired therapeutic resistance, particularly the crosstalk between AXL and other critical receptor tyrosine kinases (RTKs) such as c-MET, EGFR, HER2/HER3, VEGFR, PDGFR, and FLT3. Finally, we highlight key research areas that require further exploration to enhance AXL-mediated therapeutic approaches for improved clinical outcomes.

Subject terms: Cancer therapy, Oncogenes

Introduction

Cancer progression is significantly influenced by aberrant signaling through receptor tyrosine kinases (RTKs), including c-MET, EGFR, VEGFR, PDGFR, HER2/HER3, and FLT3. Anticancer therapies utilizing small-molecule inhibitors against RTKs are important treatment strategies for advanced-stage human malignancies.14 However, acquired drug resistance remains a serious concern, limiting the long-term efficacy of these inhibitors. Thus, urgent measures are needed to identify novel therapeutic targets to prevent resistance and limit tumor progression. AXL, a member of the TAM (TYRO3, AXL, and MERTK) family of RTKs, has emerged as a promising target. The name AXL is derived from the Greek word anexelekto, meaning uncontrolled, reflecting its involvement in cancer progression. TAM RTKs are recognized for their role in various biological functions, including promoting apoptotic cell phagocytosis, maintaining vascular integrity and modulating immune responses.58 In addition, they serve as phosphatidylserine-mediated virus entry-enhancing receptors (PVEERs), facilitating the infiltration of enveloped viruses into human cells.911 Structurally, all three TAM share common features, where ligand binding induces homodimerization, leading to autophosphorylation and kinase activation. This activation triggers several downstream signaling pathways involved in cancer onset and progression.1214 Unlike many RTKs, TAM receptors are rarely mutated. Instead, their dysregulation is typically due to overexpression caused by gene amplification or stimulation by the tumor microenvironment. This makes TAM receptors, particularly AXL, attractive therapeutic targets with a lower risk of developing drug resistance through mutations.15 AXL has gained attention in cancer therapy due to its frequent overexpression in various malignancies including renal,16,17 lung,18 breast,19 prostate,20 colon,21,22 esophageal,23 and head and neck cancers.24,25 Its role in therapeutic resistance has spurred several clinical trials.26 Notably, the selective AXL inhibitor bemcentinib (BGB324/R428) recently received fast-track designation by the US FDA in combination with a PD-1/PD-L1 agent, highlighting its growing importance in cancer research. AXL plays critical role in cancer cell proliferation, stemness, migration, invasion, angiogenesis, therapeutic resistance and poor prognosis.27,28 For instance, in prostate cancer, AXL regulates cell proliferation through the Akt/NF-kβ pathway.20 In breast cancer, it maintains acquired stemness by co-opting epithelial plasticity programs within mammary stem cells.29 Beyond cancer cells, AXL promotes angiogenesis by supporting endothelial cell function, facilitating metastasis, as seen in hepatocellular carcinoma (HCC) through activation of the PI3K/Akt/SOX2/DKK-1 network.30 In head and neck cancer, AXL derives resistance to cetuximab by upregulating the HER3 ligand neuregulin1.31 Similarly, in non-small-cell lung cancer (NSCLC), AXL enhances MYC transcription, which disrupts purine metabolism and increases the likelihood of drug-resistant mutations.32 In addition, patients with pancreatic ductal adenocarcinoma (PDA),33 clear cell renal cell carcinoma (ccRCC),34 and cholangiocarcinoma (CCA)35 often face poor prognosis due to dysregulation of the growth arrest-specific protein 6 (GAS6)/AXL signaling pathway. These findings collectively suggest that AXL could offer significant potential for treating various cancers and preventing therapeutic resistance.

In this review, we explore recent studies on structural and functional characteristics of AXL, its regulatory mechanisms, and the crosstalk with other RTKs. We also examine preclinical and clinical evidences of AXL’s role in therapeutic resistance. Finally, we discuss the translational relevance of AXL-targeted therapies in enhancing the patient outcomes by reducing resistance and promoting tumor regression.

Structure, signaling, regulation, and functions of AXL

Structure of AXL receptor

The AXL gene is located at locus 19q13.2, near the BCL3 proto-oncogene. It consists of 20 exons encoding an 894 amino acid multidomain protein.26 AXL is a 140 kDa single-pass transmembrane receptor, comprising extracellular, transmembrane and intracellular domains. The extracellular domain of AXL contains two immunoglobulin (Ig)-like repeats and two fibronectin type III (FNIII)-like repeats, structurally resembling neural cell adhesion molecules (NCAMs).36 AXL binds to its ligand GAS6, primarily through the Ig motifs, with regulation modulated by the FNIII.37 The intracellular domain is crucial for autophosphorylation and subsequent kinase activity.38 Upon GAS6 binding, AXL undergoes dimerization and is activated through phosphorylation at specific tyrosine residues. This phosphorylation event is essential for the recruitment of key adaptor and effector molecules, initiating downstream signaling pathways from AXL.39

AXL ligands

Initially believed to be orphan receptors, the TAM family of receptors are now known to have several ligands,40 including GAS6, protein S (Pros1), Tubby, Tubby-like protein 1 (Tulp-1), and galectin-3.41 GAS6 and protein S are vitamin K-dependent secretory proteins that bind to the extracellular domains of AXL, TYRO3, and MER, triggering dimerization and phosphorylation within the cytoplasmic domain, leading to receptor activation.42,43 Among these, ligands, GAS6 and Tulp-1 are recognized as universal ligands for all the TAM receptors, binding to each member of the TAM family.44 The full activation of these receptors, particularly in both healthy tissues and the cancer microenvironment, depends on their interactions with ligands and phosphatidylserine (PtdSer). GAS6 binds to PtdSer present on the cell membrane, forming an extracellular lipid-protein complex. This complex is critical for initiating TAM receptor dimerization and activating of the AXL signaling pathway, crucial for cancer cell survival and proliferation (Fig. 1).45

Fig. 1.

Fig. 1

Structure and activation mechanisms of the TYRO3, AXL, and MERTK (TAM) receptors. a The TAM receptors and ligands share closely related structures. The structure of growth arrest-specific protein 6 (GAS6) and vitamin K-dependent protein S (PROS1) comprises a γ-carboxyglutamic acid (GLA) domain, a loop region, four epidermal growth factor-like (EGF-like) repeats, and two C-terminal globular laminin G-like (LG) domains. TAM receptors consist of two immunoglobulin-like (Ig) domains, two fibronectin III (FNIII) repeat domains, and a kinase domain. GAS6 binds to all three receptors, while PROS1 binds to MERTK and TYRO3, but not to AXL. The decreasing width of the arrows indicates a reduction in receptor activation strength by the ligands, reflecting their relative affinities. b Ligand-Dependent Activation: Binding of GAS6 to AXL leads to receptor activation. Optimal activation requires PtdSer (phosphatidylserine) on the opposed cell membrane, which binds to the GLA domain of the ligand. c Ligand-Independent Activation: Various factors in the tumor microenvironment can activate AXL in a ligand-independent manner. d Homophilic Interaction: Activation can occur through interaction between AXL monomers on neighboring cells. e Heterophilic Activation: AXL can also be activated heterophilically with non-TAM receptors. Activation of AXL through these mechanisms results in autophosphorylation of intracellular tyrosine residues, initiating signaling pathways that promote tumor proliferation, invasion, migration, angiogenesis, drug resistance, and immune evasion

This interaction between GAS6 and AXL, mediated by Ig motifs and FNIII, is unique. It prevents direct AXL/AXL or GAS6/GAS6 interactions, influencing processes such as epithelial-to-mesenchymal transition (EMT), a key event in cancer progression.46

Activation and regulation of AXL signaling

The activation of the AXL receptor occurs through several mechanisms: ligand-dependent (GAS6-driven dimerization), ligand-independent (atypical) activation, monomer interactions with neighboring cells, and heteromeric dimerization with non-TAM receptors.42,47,48

Mechanisms of AXL activation

GAS6-dependent AXL activation

AXL binds to GAS6 in a 1:1 ratio, forming a complex that undergoes dimerization, activating downstream signaling pathways such as phosphatidylinositol 3-kinase/Akt kinase (PI3K/Akt) and mitogen-activated protein kinase/extracellular signal-regulated kinase (MAPK/ERK). These pathways are involved in promoting cancer cell growth and metastasis.14,37,49,50

GAS6-independent AXL activation

Beyond the canonical GAS6-driven mechanism, AXL can be activated in the absence of GAS6 through crosstalk with other receptors like c-MET or EGFR,51 self-dimerization,52 or as a response to oxidative stress.53 These alternative mechanisms further drive cancer progression.

AXL regulation

AXL signaling is tightly regulated through a range of complex mechanisms, including epigenetic alterations, transcriptional, translational, and posttranslational modifications.26 Several studies have demonstrated that AXL gene expression can be modulated by promoter methylation,54 as well as by various transcription factors (HIF-��, AP-1, YAP1/TEAD, CREB, MZF1, and SP-1)5559. In addition, interactions with miRNAs also play a role in the regulation of AXL expression.60

Posttranslational modifications, including glycosylation, phosphorylation, proteolytic cleavage, and ubiquitination, further influence the functionality of AXL in cancer progression.6163 Beyond these mechanisms, studies have identified the role of long non-coding RNA (LncRNAs) (DANCR, and XIST), which act as endogenous RNAs that compete with miRNAs (miR-33a-5p, and miR-93-5p) to regulate AXL signaling at the post-transcriptional level in various malignancies.6466 In addition to these regulatory mechanisms, transcriptomic sequencing of primary lung adenocarcinomas revealed the AXL-MBIP (MAP3K12 binding inhibitory protein 1) fusion gene as another potential mechanism for AXL activation.67 These findings expand the understanding of AXL’s regulatory landscape and could provide new avenues for the development of targeted cancer therapies.

Biological functions of AXL and its role in cancer progression

AXL overexpression has been linked to several key biological processes that drive cancer development and metastasis. These functions, outlined below, highlight the crucial role of AXL in promoting tumor progression (Fig. 2).

Fig. 2.

Fig. 2

AXL regulates multiple biological processes in cancer. AXL is pivotal in regulating several key biological processes in cancer. Upon binding to its ligand, GAS6, AXL becomes activated, triggering downstream signaling pathways that promote cancer cell proliferation, survival, migration, invasion, epithelial-to-mesenchymal transition (EMT), angiogenesis, stem cell maintenance, and immune suppression or evasion

AXL in cell proliferation and survival

AXL signaling plays a pivotal role in promoting cell proliferation through various downstream pathways, including PI3K/Akt, JAK/STAT, NF-κβ, and RAS/RAF/MEK/ERK.68,69 In addition, AXL enhances cell survival by regulating NF-κβ nuclear transport, which leads to increased expression of anti-apoptotic markers such as survivin and BCL-2, while simultaneously reducing the expression of pro-apoptotic markers like caspase 3 and BAD.70 AXL inhibition has been shown to lower the levels of anti-apoptotic proteins, such as MCL-1 and BCL-2, while upregulating apoptosis-related proteins like Bim, particularly in chronic lymphocytic leukemia (CLL).71 Blocking the GAS6/AXL axis is critical for suppressing tumor growth in various cancers, including renal cell carcinoma (RCC), prostate cancer, lung adenocarcinoma, and colorectal cancer.72 For example, a study by Yu et al. identified AXL as a key survival factor for RCC cells, where it is overexpressed in human RCC tissues and cell lines. Targeting AXL using the monoclonal antibody hMAb173 induces apoptosis in both ex vivo and in vivo models.73 These findings suggest that AXL plays a significant role in cancer cell proliferation and survival, making it a promising therapeutic target to inhibit cancer growth and enhance apoptosis in a range of malignancies.

AXL in cell migration and invasion

AXL overexpression is strongly linked to increased cell migration and invasion across various cancer types. AXL activity has been associated with the upregulation of GTP-binding proteins, such as Rho and Rac, which are key facilitators of cell migration.74 Additionally, AXL activation promotes the expression of important factors such as Akt and MMP9, primarily through NF-κβ and Brg-1 signaling pathways.70,75 In cancers, such as oral squamous cell carcinoma (OSCC) and ovarian tumors, AXL activation stimulates the PI3K/Akt pathway that mediates cancer cell invasion through increased expression of proteolytic enzymes like MMP2 and MMP9. In hepatocellular carcinoma, studies have shown that YAP1 knockdown results in reduced AXL activity, leading to decreased cell invasion. This emphasizes the significant role of AXL in invasion and migration processes.76 Furthermore, AXL regulates genes associated with metastasis (e.g., somatostatin receptor 2 (SSTR2), fms-related receptor tyrosine kinase 4 (FLT4), matrix metalloproteinase 10 (MMP10), kisspeptin-1 (KISS1), collagen type IV alpha 2 chain (COL4A2), RAR-related orphan receptor B (RORB)) and is linked to the expression of stem cell markers.77 It also facilitates the migration and invasion of breast cancer stem cells. Notably, AXL is involved in the production of cathepsin B, a facilitator of invasion, and the peripheral distribution of lysosomes in esophageal cancer cell lines.78 These findings collectively suggest that AXL could be a promising target to inhibit cancer invasion and migration, thereby disrupting tumor progression.

AXL in epithelial-to-mesenchymal transition (EMT)

Recent studies have demonstrated that AXL plays a role in promoting EMT. Activation of AXL initiates a reversible transition from an epithelial to a mesenchymal phenotype, enhancing cell invasiveness and metastatic potential.79 Several growth factors, including TGF-β, PDGF, and HGF as well as transcription factors such as Slug, Snail, ZEB1, and TWIST1, mediate EMT in various cancers.77,80 AXL has been identified as a phenotypical marker in Non-small-cell lung cancer (NSCLC) cell lines and breast cancer stem cells.79,81 Elevated levels of AXL are crucial for the downregulation of certain pro-epithelial factors such as E-cadherin, while simultaneously upregulating mesenchymal markers like N-cadherin, Slug, Snail, ��-catenin, α-SMA, and vimentin.82,83

In addition to downregulating pro-epithelial factors and inducing mesenchymal markers, AXL enhances the transcriptional expression of EMT regulators Slug and N-cadherin by activating PI3K and NF-κβ pathways.77,84 Overexpression of AXL in primary breast cancers is associated with reduced patient survival. Experimental breast cancer models have shown that knocking down AXL inhibits the metastasis of highly invasive breast carcinoma cells to lymph nodes and other organs, leading to improved overall survival.81 Similarly, TGF-β promotes AXL expression in HER2-positive breast cancer tissues, contributing to their invasive and metastatic properties.85 In summary, AXL plays a dual role in promoting EMT across various cancers. It functions as an upstream signaling molecule that drives EMT and maintains cells in a mesenchymal state independently of GAS6. This mesenchymal state can be reversed by downregulating AXL, restoring an epithelial-like morphology.

EMT is also involved in the formation of circulating tumor cells (CTCs), which drive metastatic relapse and are key components of liquid biopsies.86 A study by de Miguel-Pérez et al. suggested that assessing both tissue AXL expression and CTCs could improve predictions of patient outcome and tailor post-surgical therapy for resected lung adenocarcinomas.87 Currently, CellSearch is the only FDA-cleared platform for CTC enumeration, demonstrating clinical value across various malignancies.8890 However, the limited expression of epithelial cell adhesion molecule (EPCAM) in CTCs has restricted their utility in advanced tumors.9193 Microcavity array technology, utilizing mesenchymal markers, has proven effective in identifying AXL-expressing CTCs in advanced NSCLC patients.94 Furthermore, Bardol et al. provided the first evidence of CellSearch® detecting AXL+ CTCs in metastatic breast cancer, paving the way for future studies to explore anti-AXL treatments.95

AXL in angiogenesis

TAM receptors, including AXL, are crucial not only for cancer cell proliferation and survival but also for vessel integrity and angiogenesis. They modulate various growth factors such as VEGF, FGF, and PDGF, which are essential for these processes.96,97 AXL is expressed in endothelial and vascular smooth muscle cells, where it supports the stabilization and survival of endothelial cells and contributes to tissue remodeling, wound healing, and vessel integrity.98 During neoplasm formation, both AXL and its ligand GAS6 are often overexpressed. They regulate the expression of matrix proteins, promoting endothelial cell proliferation, migration, and survival.99 AXL has been implicated in stimulating growth and tube formation in human umbilical vein endothelial cells (HUVEC), through the VEGF-A-mediated activation of the PI3K/Akt pathway, which is crucial for tumor angiogenesis.100,101 Knockdown studies of AXL and GAS6 have demonstrated reduced endothelial tube formation, attributed to decreased expression of pro-angiogenic proteins such as Dickkopf-related protein 3 (DKK3) and angiopoietin-2 (Ang-2),27 which are key mediators of tumor angiogenesis. Inhibitors of AXL, when used in conjunction with antiangiogenic agents, have shown a significant reduction in vessel or tube formation in renal carcinoma xenografts.102 In addition, the GAS6/AXL/S100A10 axis has been found to promote plasmin generation, endothelial cell migration, and angiogenesis in clear cell renal cell carcinoma (ccRCC) cells, with AXL expression being associated with resistance to antiangiogenic therapies.102

AXL in stemness

AXL is a key driver of stemness in cancer cells, facilitating self-renewal and contributing to the long-term maintenance of tumors. Its expression corelates with various cancer stem cell markers, including cell division cycle 2a (Cdc2a), bone gamma-carboxyglutamate protein 1 (Bglap1), aldehyde dehydrogenase 1 (ALDH1), enhancer of zeste homolog 2 (EZH2) and cluster of differentiation 44 (CD44).70,103,104 Increased levels of these stemness markers enhance resistance to chemotherapy in squamous cell carcinoma and promote self-renewal activities in human mammary epithelial cells.105 AXL maintains breast cancer stem cells (BCSCs) and epithelial plasticity through a positive feedback loop. Inhibiting AXL with inhibitors such as MP-470 or amuvatinib can effectively reverse the EMT process in BCSCs by downregulating the NF-κβ pathway. AXL expression is linked to the regulation of stem cell genes, metastatic genes, and enhanced tumor progression through increased cellular invasion and migration.77 Specifically, the AXL-PYK2 (proline-rich tyrosine kinase 2)-PKCα (protein kinase C alpha) signaling axis has been shown to induce stemness in triple-negative breast cancer (TNBC). In TNBC, the expression of AXL, PYK2, and PKCα correlates with stemness in breast cancer patients. These molecules activate signal transducer and activator of transcription (STAT3), SMAD family member 3 (SMAD3), transcriptional coactivator with PDZ-binding motif (TAZ), and Fos-related antigen-1 (FRA1), along with pluripotent stem cell transcription factors Nanog and Oct4.106

In cutaneous squamous cell carcinoma (SCC), AXL contributes to cell invasiveness and stemness by modulating Wnt and TGF-β signaling pathways and disrupting cell-to-cell adhesion.103 Furthermore, suppressing AXL through inhibitors or shRNA extends the survival of chronic myeloid leukemia (CML) mice and reduces the growth of leukemia stem cells (LSCs). The GAS6/AXL complex has been shown to stabilize β-catenin in human CML CD4+ cells, presenting a potential target for eliminating CML LSCs.107 Therefore, targeting AXL or its signaling pathways may enhance clinical response to anticancer therapies by reducing stemness, recurrence and metastasis.

AXL in immunosuppression

In recent years, significant progress has been made in cancer immunotherapy; however, acquired resistance remains a major clinical challenge. AXL has emerged as a key player in the development of tumor immune tolerance.26 AXL induces immune evasion by upregulating BCL-2 and Twist, suppressing inflammatory signaling, and reducing the production of pro-inflammatory cytokines and chemokines such as interleukin 1 alpha (IL-1 α), interleukin 6 (IL-6), tumor necrosis factor-alpha (TNF-α), C-X-C motif chemokine ligand 1, 2 and 5 (CXCL1, CXCL2, CXCL5), colony-stimulating factor 1-3 (CSF1-3) and C-C motif chemokine ligand 2-5 (CCL2-5).108 It also contributes to radio-resistant and checkpoint immune checkpoint resistance by blocking MHC-I antigen presentation and promoting the production of myeloid-supporting chemokines and cytokines, which trigger an adverse immune response.109,110 AXL confers therapeutic resistance by activating the PI3K/Akt pathway and increasing the transcription of programmed cell death ligand 1 (PD-L1) in head and neck cancer cells.111,112 Overexpression of AXL has been linked to both intrinsic and acquired resistance to immune checkpoint blockade (ICB) in real cell carcinoma (RCC). A strong correlation exists between AXL and PD-L1 expression, with high levels of AXL in tumor cells associated with lower response rates and shorter progression-free survival following anti-PD-1 treatment. Furthermore, patients with tumors that display both high PD-L1 and AXL expression have the poorest overall survival and a reduced antitumor immune response.113

Knockout studies of AXL have demonstrated reduced tumor growth and increased sensitivity to immunotherapy due to the infiltration of CD8 + T cells in a murine model of breast cancer.114,115 This indicates that AXL plays a multifaceted role in immunosuppression by regulating the secretion of immunosuppressive cytokines and mediators.116 AXL-mediated remodeling of the tumor microenvironment (TME) leads to the downregulation of immune effector cells, such as natural killer (NK) cells, T cells, and dendritic cells, while promoting the recruitment of immunosuppressive cells such as regulatory T cells (Tregs), tumor-related macrophages, and myeloid-derived suppressor cells (MDSCs).117

In addition, AXL signaling results in a reduction of cytotoxic T cells at effector sites and induces the transformation of macrophages from M1 to M2 polarization, ultimately facilitating tumor evasion and a poor prognosis.118 Therefore, inhibiting AXL can remodel the TME, enhance immune response and improve the therapeutic potential of immune checkpoint blockade inhibitors.

AXL in viral infections and inflammatory responses

Beyond its role in immune regulation, AXL has been implicated in the progression and severity of various viral infections, including Ebola virus (EBOV), severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2), and Zika virus (ZIKV).119121 Its interactions with these viruses highlight its dual role in facilitating viral entry and modifying inflammatory responses, establishing a connection between infectious diseases and cancer progression.

Using a series of AXL mutants, Shimojima et al. demonstrated that EBOV particles activate AXL through cell surface adhesion and interaction with its immunoglobulin domains, enabling viral entry by endocytosis.120 Similarly, AXL has been identified as a potential receptor for SARS-CoV-2, interacting directly with the spike protein’s N-terminal to facilitate viral attachment and entry into pulmonary epithelial cells.122 During Zika virus infection, AXL enhances viral entry in glial cells through GAS6 bridging and impacts innate immune responses.119 Studies involving Zika virus strain MR-766 in human glioblastoma cell lines revealed that AXL is essential for viral entry, though the exact mechanisms remain unclear and needs further investigations.123 Moreover, AXL was recently identified as a co-receptor for parvovirus B19 in human erythroid progenitors, further emphasizing its role in viral pathogenesis.124 These findings suggest that targeting AXL could be a promising strategy for future antiviral treatments.

In addition to facilitating viral entry, AXL plays a crucial role in regulating inflammatory responses. Activation of AXL by its ligand, GAS6, triggers a signaling cascade that suppresses innate immune defenses, allowing viruses to evade detection. This occurs through the upregulation of suppressors of cytokine signaling (SOCS), which in turn inhibits toll-like receptor (TLR) pathways essential for a robust immune response.125 While this immunomodulatory function can reduce tissue damage during infections, it also creates a vulnerability to malignancy. Tumors often exploit these inflammatory pathways to escape immune surveillance, and elevated AXL expression can impair the function of natural killer (NK) cells and macrophages, weakening the antitumor immune response. For instance, the human papillomavirus type 16E6 (HPV16E6) oncoprotein has been shown to upregulate AXL expression in cervical cancer via the MAGI-2/PTEN/Akt/MZF1 network, increasing tumor invasiveness and promoting immune evasion by reducing responsiveness to NK cells-mediated lysis.126

Thus, AXL serves as a critical link between viral infections, inflammatory responses, and cancer progression.127 By facilitating viral entry and modulating the host’s immune response, AXL contributes to the pathogenesis of both infectious diseases and cancer metastasis. These insights position AXL as a potential therapeutic target in both virology and oncology, where its inhibition could enhance antitumor immunity and improve clinical outcomes.

However, studies suggest that AXL inhibition may also create a tumor-promoting environment by enhancing pro-inflammatory cytokine production and impairing the clearance of apoptotic neutrophils.128 This is particularly relevant in comorbid conditions like inflammatory arthritis, where AXL, regulated by miR-34a, is essential for controlling dendritic cell (CD1c+ DC) activation and the generation of pro-inflammatory cytokines such as TNF, IL-6, IL-1β, and IL-23.129 Moreover, AXL deficiency has been observed to exacerbate pulmonary arterial hypertension (PAH) via the bone morphogenetic protein receptor 2 (BMPR2) pathway, suggesting a protective role for the GAS6/AXL axis in PAH.130 Therefore, AXL’s paradoxical roles as both a proto-oncogene and an immunological protector highlight its complexity in cancer and related comorbidities. Blocking AXL could interfere with its beneficial immunoregulatory effects, potentially exacerbating inflammation or immune dysregulation in comorbid diseases. Thus, a thorough investigation of AXL’s context-dependent effects is necessary to develop effective therapies.

Future research should focus on clarifying AXL’s roles in viral pathogenesis, tumor biology, and related comorbidities, paving the way for innovative treatment strategies. Nonetheless, the unpredictable mutations in viral genomes present a significant challenge for therapeutic development, necessitating further investigation.

Targeted therapy toward AXL

The studies discussed above confirm the pleiotropic effects of AXL on cancer growth, metastasis, and drug resistance, establishing it as a promising target for anticancer therapies. AXL inhibitors have shown significant benefits both in vitro and in vivo by promoting apoptosis and reducing cell proliferation and migration.131 Targeting AXL can also enhance the therapeutic efficacy of chemotherapy132, and various small-molecule inhibitors that target EGFR, HER2, PARP, PI3K, and VEGF-A.133137 Currently, different types of AXL inhibitors are in various stages of development including small-molecule tyrosine kinase inhibitors (TKIs) that block AXL kinase activity, anti-AXL antibodies against AXL receptors, soluble AXL decoy receptors for neutralizing GAS6, nucleotide aptamers that bind to and inhibit AXL, and natural compounds as AXL inhibitors.70,138 In this review, we primarily focus on small-molecule AXL inhibitors and anti-AXL antibodies that currently being investigated in preclinical and clinical studies for the treatment of various human malignancies (Table 1).

Table 1.

Targeted therapies against AXL

Small-molecule AXL TKI Target(s) Mechanism(s) Cancer type Monotherapy/combination Clinical trial no./phasea Statusa Adverse events
Selective AXL Inhibitors
 Bemcentinib/BGB324/R428 AXL Inhibition of Akt and MAPK pathways, upregulate PUMA and downregulate Bcl-2 NSCLC + Docetaxel NCT02922777- I Completed Neutropenia, diarrhea, nausea, vomiting, QTc prolongation, transaminase increase, asthenia, and fatigue
AML or MDS Monotherapy in AML or MDS/+ cytarabine or decitabine in AML NCT02488408- Ib/II Unknown status
NSCLC + Erlotinib NCT02424617- I/II Completed
TNBC or TN-IBC + Pembrolizumab NCT03184558- II Terminated
MDS patients failing standard-of-care therapy Monotherapy NCT03824080- II Completed
Advanced NSCLC + Pembrolizumab NCT03184571- II Completed
Malignant mesothelioma + Pembrolizumab NCT03654833- II Active, not recruiting
Metastatic melanoma + Pembrolizumab or Dabrafenib/Trametinib NCT02872259- Ib/II Completed
Recurrent glioblastoma undergoing surgery Monotherapy NCT03965494- I Terminated
Metastatic pancreatic cancer + Chemotherapy (nab-paclitaxel/gemcitabine) NCT03649321- I/II Terminated
Untreated non-squamous NSCLC + Pembrolizumab/carboplatin/pemetrexed NCT05469178- Ib/IIa Recruiting
 TP-0903/Dubermatinib AXL Inhibit Aurora B activation to induce G2/M arrest; Induction of apoptosis by suppressing Mcl-1, Bcl-2 and XIAP; upregulation of BIM Advanced solid tumors Monotherapy NCT02729298- I Completed Thrombocytopenia, anemia, nausea, syncope, and vomiting
FLT3 mutated AML TP-0903 alone/+azacitidine NCT04518345- Ib/II Completed
AML Monotherapy NCT03013998- Ib/II Recruiting
Previously treated CLL Monotherapy/+ ibrutinib NCT03572634-I/II Terminated
 DS-1205 AXL Inhibit GAS6-mediated cell migration Metastatic or unresectable EGFR-mutant NSCLC + Gefitinib NCT03599518-I Terminated Combination with gefitinib may induce elevation of liver enzymes, vomiting, diarrhea, and fatigue
Metastatic or unresectable EGFR-mutant NSCLC + Osimertinib NCT03255083-I Terminated
 AVB-S6-500/Batiraxcept AXL decoy receptor Inhibit GAS6-induced AXL or Src phosphorylation Platinum-resistant recurrent ovarian cancer + Pegylated liposomal doxorubicin or paclitaxel NCT03639246- I Completed Avelumab in combination with AVB-S6-500 may result in UTI, hyponatremia, elevated creatinine, anemia, thrombocytopenia, hematuria, anorexia, and sepsis
Platinum-resistant recurrent ovarian, fallopian tube or primary peritoneal cancer + Durvalumab (MEDI4736) NCT04019288- I/II Active, not recruiting
Platinum-resistant recurrent ovarian cancer + Paclitaxel vs placebo and paclitaxel NCT04729608- III Terminated
Advanced or metastatic ccRCC AVB-S6-500 alone/+ cabozantinib/+ cabozantinib and nivolumab NCT04300140- Ib/II Terminated
Advanced pancreatic adenocarcinoma + Nab-paclitaxel and gemcitabine NCT04983407- Ib/II Terminated
Advanced urothelial carcinoma + Avelumab NCT04004442- I Active, not recruiting
Multitargeted AXL inhibitors
 BMS777607

AXL,

RON,

c-MET, TYRO3, MER

Inhibition of HGF-induced c-MET-phosphorylation and downregulate MEK-ERK and PI3K/Akt pathways Advanced or metastatic solid tumors Monotherapy NCT01721148-I Completed Anemia, nausea, Constipation
Advanced or metastatic solid tumors Monotherapy NCT00605618- I/II Completed
 Cabozantinib/BMS-907351 AXL, VEGFR2, c-MET, RET, TIE2, c-KIT, Flt3, ROS1 Decrease AXL phosphorylation and TGF-β-induced E-cadherin expression NSCLC With or without erlotinib NCT00596648- Ib/II Completed Fatigue, diarrhea, hypertension, palmar-plantar erythrodysesthesia syndrome, fistula, abdominal and pelvic abscess
Previously treated metastatic NSCLC + Erlotinib NCT01866410- II Completed
Stage IV NSCLC Monotherapy/+ erlotinib NCT01708954- II Active, not recruiting
KRAS wild-type metastatic colorectal cancer Monotherapy in MET amplified metastatic colorectal cancer/+ panitumumab NCT02008383- I Completed
Locally advanced or metastatic solid tumors Cabozantinib alone / + atezolizumab NCT03170960- I/II Active, not recruiting
Metastatic castration-resistant prostate cancer + Atezolizumab vs second novel hormonal therapy NCT04446117- III Active, not recruiting
Advanced HCC with no previous systemic anticancer therapy + Atezolizumab vs sorafenib NCT03755791- III Active, not recruiting
Metastatic NSCLC previously treated with anti-PD-L1/ PD-1 Antibody and platinum- chemotherapy + Atezolizumab vs docetaxel monotherapy NCT04471428- III Active, not recruiting
Advanced RCC after immune checkpoint inhibitor treatment + Atezolizumab vs cabozantinib alone NCT04338269- III Active, not recruiting
Previously untreated advanced or metastatic RCC + Nivolumab vs Sunitinib NCT03141177- III Active, not recruiting
Non-clear cell RCC + Nivolumab NCT03635892- II Active, not recruiting
Previously untreated advanced or metastatic RCC + Nivolumab and ipilimumab vs nivolumab and ipilimumab with matched placebo NCT03937219- III Active, not recruiting
NSCLC With or without erlotinib NCT00596648- Ib/II Completed
Advanced kidney cancer Ipilimumab and nivolumab followed by nivolumab alone vs nivolumab with cabozantinib NCT03793166- III Active, not recruiting

Recurrent stage

IV NSCLC

+ Nivolumab or + nivolumab and ipilimumab vs nivolumab alone NCT03468985- II Active, not recruiting
Advanced liver cancer Nivolumab alone/ nivolumab and sorafenib/ nivolumab and ipilimumab/ + nivolumab/ + nivolumab and ipilimumab NCT01658878-I/II Active, not recruiting
Radioiodine-refractory differentiated thyroid cancer progressed after prior VEGFR -targeted therapy Cabozantinib vs placebo NCT03690388-III Active, not recruiting
Advanced pancreatic neuroendocrine and carcinoid tumors Cabozantinib vs Placebo NCT03375320-III Active, not recruiting
Hepatocellular carcinoma patients received prior sorafenib Cabozantinib vs Placebo NCT01908426- III Completed
Metastatic RCC Cabozantinib vs Everolimus NCT01865747- III Completed
 Bosutinib/SKI-606 AXL, Src kinase, BCR-ABL Inhibits autophosphorylation of AXL with downregulation of slug and inhibits PI3K/Akt/mTOR pathways Recurrent glioblastoma Monotherapy NCT01331291- II Completed Diarrhea, rash, liver enzyme elevations
Advanced breast cancer Monotherapy NCT00319254- II Completed
Advanced malignant solid tumors Monotherapy NCT01001936- I Completed
Philadelphia chromosome +ve CML previously treated with one or more TKIs Monotherapy NCT02228382- IV Terminated
CML-CP previously treated with 2 or more TKIs ABL001 vs Bosutinib NCT03106779- III Active, not recruiting
Selected metastatic solid tumors + Pemetrexed NCT03023319- I Completed
HR + HER2- advanced breast cancer refractory to CDK4/ 6 inhibitor + Palbocicilib and fulvestrant NCT03854903-I Active, not recruiting
 Crizotinib/PF-02341066 AXL, ALK, c-MET, RON Inhibits phosphorylation of target receptors ALK+ advanced NSCLC Brigatinib vs Crizotinib NCT02737501- III Completed Abdominal pain, headache, pain in extremity, nausea, pyrexia
High-risk uveal melanoma following definitive therapy Monotherapy NCT02223819- II Completed
c-MET +ve gastric adenocarcinoma Monotherapy NCT02435108- II Completed
mCRPC + Enzalutamide NCT02207504- I Completed
 Sunitinib/SU11248 AXL, c-KIT, FLT3, PDGFR, VEGFR2 Inhibits ligand-mediated VEGFR2, PDGFR-β phosphorylation Metastatic RCC Monotherapy NCT00706706- IV Completed Diarrhea, fatigue, hypertension, hematologic adverse events and palmar-plantar erythrodysesthesia syndrome
Refractory solid tumors Monotherapy NCT02691793- IV Completed
High-risk non-muscle invasive lower urinary tract urothelial carcinoma Monotherapy NCT00794950- II Completed
Advanced cholangiocarcinoma Monotherapy NCT01718327- II Completed
Advanced well-differentiated pancreatic neuroendocrine tumors Monotherapy NCT01525550- IV Completed
Imatinib resistant or intolerant malignant GIST Monotherapy NCT00793871- IV Completed
Locally advanced or recurrent soft tissue sarcoma Preoperative sunitinib and radiation NCT01498835- I Completed
Recurrent ovarian clear cell carcinoma Monotherapy NCT01824615- II Completed
NSCLC patients with brain metastases Monotherapy NCT00372775-II Completed
Metastatic or recurrent thymic carcinoma Monotherapy NCT02623127- II Completed
 Sitravatinib/MGCD516

AXL,

c-KIT, VEGFR3 PDGFR,

c-MET, RET, MER, FLT3

Immune-modulation of tumors by reducing immunosuppressive myeloid cells and increase CD4+ & CD+8 T cells Metastatic, pre-treated TNBC Monotherapy NCT04123704- II Terminated Diarrhea, fatigue, rash, nausea
ccRCC + Nivolumab NCT03680521- II Completed
Advanced solid tumor malignancies Monotherapy NCT02219711- I/Ib Completed
Advanced or metastatic kidney cancer progressed on prior VEGF-targeted therapy + Nivolumab NCT03015740- I/II Completed
Locally recurrent or metastatic TNBC + Tislelizumab/ + tislelizumab and nab-paclitaxel NCT04734262- II Active, not recruiting
Advanced solid tumors + Tislelizumab NCT03666143- Ib Completed
NSCLC + Nivolumab NCT02954991- II Terminated
Advanced or metastatic ccRCC or other solid malignancies + Nivolumab and ipilimumab NCT04518046- I/Ib Completed
Advanced or metastatic urothelial carcinoma + PD-(L)1 checkpoint inhibitors regimens (nivolumab/ pembrolizumab and enfortumab vedotin) NCT03606174- II Terminated
Advanced NSCLC + Nivolumab vs Docetaxel NCT03906071- III Active, not recruiting
Unresectable locally advanced or metastatic HCC or GC/GEJC Monotherapy/ + tislelizumab NCT03941873- I/II Completed
Locally advanced or metastatic NSCLC + Tislelizumab vs Docetaxel NCT04921358- III Terminated
Advanced biliary tract cancer + Tislelizumab NCT04727996- II Active, not recruiting
Advanced treatment-naïve PD-L1+ non-squamous NSCLC + Pembrolizumab NCT04925986- II Terminated
Advanced or metastatic malignancies Monotherapy/ + other anticancer therapies (nivolumab, pembrolizumab, enfortumab vedotin, ipilimumab) NCT04887870- II/III Active, not recruiting
S49076 AXL/MET, c-MET, FGFR1/2/3 Inhibits Aurora b kinase inhibits phosphorylation of target receptors Advanced solid tumors Monotherapy ISRCTN00759419- I Completed Peripheral edema and hypalbuminaemia
Recurrent glioblastoma multiforme + Bevacizumab ISRCTN11619481- I/II Completed
 ONO-7475

AXL,

MER

Acute leukemia or MDS Monotherapy/ + venetoclax NCT03176277- I/II Terminated
 PF-07265807

AXL,

MERTK

Enhance dendritic cell activity to prime CD8 + T cells Metastatic solid tumors Monotherapy/ + sasanlimab/ + sasanlimab and axitinib NCT04458259- I Active, not recruiting
 Ningetinib c-MET, AXL and VEGFR2 Stage IIIB or IV NSCLC patients with EGFR mutation and T790M-negative + Gefitinib NCT03758287- I/II Unknown status Myocardial enzyme elevation, transaminase elevation, skin rash, diarrhea, hypertension, coagulation abnormalities, and albuminuria
 CB469

AXL,

c-MET

NSCLC

+ Erlotinib,

gefitinib,

osimertinib

Preclinical
 Foretinib/XL880/GSK1363089 AXL, c-MET, MERTK, VEGFR1/2/3, TIE2, TYRO3, RON, FLT3, PDGFR-α/ β, c-KIT, ROS1 Inhibition of inter-receptor tyrosine kinase phosphorylation Metastatic breast cancer + Lapatinib NCT01138384-I/II Completed Fatigue, alopecia, diarrhea
Recurrent/metastatic Triple-negative breast cancer Monotherapy NCT01147484-II Completed
Papillary RCC Monotherapy NCT00726323-II Completed
Metastatic gastric cancer Monotherapy NCT00725712-II Completed
 Merestinib/LY2801653 MERTK, DDR1/2, AXL, MKNK1/2, FLT3, RON, ROS1, TYRO3, c-MET, PDGFR-α, TEK Advanced or metastatic biliary tract cancer + Cisplatin and gemcitabine NCT02711553-II Active, not recruiting Anemia, thrombocytopenia, leukopenia, neutropenia, nausea, constipation
NSCLC harboring MET Exon 14 mutations and solid tumors with NTRK rearrangements Monotherapy NCT02920996-II Terminated
Relapsed or refractory AML + LY2874455 NCT03125239-I Completed
Advanced or metastatic cancer Monotherapy/ + cisplatin and gemcitabine NCT03027284-I Completed
Advanced cancer Monotherapy/ + cetuximab in HNSCC/ + cisplatin in cholangiocarcinoma/ + gemcitabine and cisplatin in cholangiocarcinoma/ + ramucirumab in gastric cancer NCT01285037-I Completed
 Gilteritinib/ASP2215 FLT3, AXL, ALK, RET, MERTK FLT3/ITD AML in first complete remission Gilteritinib vs placebo NCT02927262-II Completed Febrile Neutropenia, anemia, thrombocytopenia, sepsis, pneumonia, diarrhea, fatigue
Advanced solid tumors Gilteritinib and 14C-labeled gilteritinib NCT02456883-I Completed
EGFRm+ advanced NSCLC with acquired resistance to an EGFR-TKI + Erlotinib NCT02495233-I/II Terminated
 Amuvatinib/MP-470 AXL-c-KIT, FLT3, c-MET, c-RET, PDGFRs Suppresses Rad51, which can disrupt DNA repair and make tumor cells more sensitive to radiation Solid tumors + paclitaxel and carboplatin/ + carboplatin and etoposide/ + topotecan/ + docetaxel/ + erlotinib NCT00881166-I Completed Fatigue, alopecia, diarrhea, nausea, anorexia, neutropenia, anemia, thrombocytopenia, leukopenia
Solid malignancies Monotherapy NCT00894894-I Completed
Small-cell lung cancer + Platinum-etoposide chemotherapy NCT01357395-II Completed
Unresectable or metastatic solid tumor or lymphoma Monotherapy NCT00504205-I Terminated
 MRX-2843 MERTK, FLT3, AXL relapsed/refractory AML, ALL, or MPAL Monotherapy NCT04872478-I Recruiting
Anti-AXL monoclonal antibodies (mAb) and antibody–drug conjugate (ADC)
 Tilvestamab/ BGB149 AXL mAb Directly binds with AXL and prevents GAS6-dependent AXL activation Relapsed, platinum-resistant, high-grade serous ovarian cancer Monotherapy NCT04893551- Ib Terminated
 YW327.6S2 AXL mAb Binds to AXL and blocks GAS6 from binding to AXL In (A549) NSCLC and (MDA-MB-231) breast cancer model + Erlotinib Preclinical
 CDX-0168 AXL mAb Block GAS6-dependent AXL phosphorylation, stimulate ADCC, release pro-inflammatory cytokines, enhance T-cell activation Human cancer lines and primary human myeloid cells Monotherapy Preclinical
 Enapotamab Vedotin/ AXL-107-MMAE/ HuMax-AXL-ADC AXL-ADC Hinder microtubules and tubulin polymerization to block mitotic spindle assembly and induces cell-cycle arrest Solid tumors Monotherapy NCT02988817- I/II Completed Febrile neutropenia, nausea, vomiting, constipation, diarrhea and a rise in γ-glutamyl transferase
 Mipasetamab Uzoptirine/ ADCT-601 AXL-ADC DNA interstrand cross-linking by pyrrolobenzodiazepine dimer Solid tumors Monotherapy/ + gemcitabine NCT05389462-I Recruiting Palmar-plantar erythrodysesthesia syndrome, anemia, rash maculopapular, and cheilitis and constipation

 Mecbotamab Vedotin

/BA3011/ CAB-AXL-ADC

AXL-ADC Binding to AXL only under tumor-specific conditions, such as low pH, allows the delivery of a cytotoxic agent directly to the tumor cells Sarcoma Mecbotamab vedotin alone / + nivolumab NCT03425279-I/II Recruiting
NSCLC CAB-AXL-ADC alone/ + PD-1 inhibitor NCT04681131-II Recruiting
Platinum-resistant high-grade serous ovarian cancer + Durvalumab NCT04918186-II Recruiting
Anti-AXL-CAR therapy
 CCT301-38 AXL Patient’s T cells are engineered to recognize and attack cancer cells with high-AXL expression Recurrent or refractory stage IV RCC Monotherapy NCT03393936-I/II Unknown status
Relapsed or refractory AXL +ve sarcomas Monotherapy NCT05128786-I Recruiting
 AXL-CAR-T AXL TNBC Monotherapy Preclinical
 AXL-CAR.C7R AXL TNBC Monotherapy Preclinical
 AXL-CAR-T AXL Advanced solid tumors Monotherapy NCT04842812-I Recruiting
 AXL-CAR-T HER2, Mesothelin, Lewis-Y, PSCA, MUC1, GPC3, AXL, EGFR, or B7-H3 Advanced lung or other cancers Monotherapy NCT03198052-I Recruiting
 AXL-CAR-NK AXL Advanced solid tumors Monotherapy/ IL7/CCL19 secreting CAR-NK cell therapy/ PD-1/PD-L1/CTLA4-scfv secreting CAR-NK cell therapy/ + Cannabidiol/ + Nicotinamide adenine dinucleotide NCT05410717-I/II Recruiting

AML acute myeloid leukemia, ALL acute lymphoblastic leukemia, ALK anaplastic lymphoma kinase, BIM Bcl-2 interacting mediator of cell death, CLL chronic lymphocytic leukemia, RCC renal cell carcinoma, ccRCC clear cell renal cell carcinoma, CML chronic myelogenous leukemia, CML-CP chronic myelogenous leukemia in chronic phase, DDR discoidin domain receptor, EGFR epidermal growth factor receptor, ESCC esophageal squamous cell carcinoma, FLT3 Fms-like tyrosine kinase 3, GC/GEJC gastric/gastroesophageal junction cancer, HCC hepatocellular carcinoma, HGF hepatocyte growth factor, HNSCC head and neck squamous cell carcinoma, MDS myelodysplastic syndromes, MPAL mixed phenotype acute leukemia, c-MET mesenchymal–epithelial transition factor, NSCLC non-small cell lung cancer, PDGFR platelet-derived growth factor receptor, ROS1 c-ros oncogene 1 receptor tyrosine kinase, TIE2 tyrosine kinase with immunoglobulin-like and EGF-like domains 2, TNBC triple-negative breast cancer, TN-IBC triple-negative inflammatory breast cancer, TRK tropomyosin receptor kinase, TKI tyrosine kinase inhibitor, VEGFR vascular endothelial growth factor receptor, XIAP X-linked inhibitor of apoptosis protein

aThe clinical trial and related information can be found on the public clinical trial registry website (https://www.clinicaltrials.gov/). Here, we have listed a partial list of some of the clinically relevant clinical trials

Small-molecule AXL tyrosine kinase inhibitors (AXL TKIs)

In recent years, there has been a significant progress in the development of small-molecule kinase inhibitors that target AXL. These inhibitors work either by specifically targeting AXL or as multitargeted agents in preclinical and clinical studies.139 Typically, AXL has not been the primary focus of small-molecule inhibitors; instead, they target AXL due to its kinase domain’s similarities with other RTKs like c-MET. Consequently, inhibitors like cabozantinib140 and bosutinib,141 often exhibit lower potencies against AXL compared to their intended targets. Interestingly, the c-MET inhibitor BMS777607 has demonstrated greater potency against AXL than c-MET in a laboratory setting.70 In addition to this, several highly selective AXL-specific inhibitors have been developed, including BGB324, NA80xl, TP-0903, and SGI-7079. These belong to the category of ATP-competitive inhibitors containing an adenine-mimicking heterocyclic moiety. They bind to hinge region of ATP-binding site, either in active conformation of AXL (Aspartate-phenylalanine-glycine or D–F-G motif oriented toward the active site, DFG-in), known as ‘Type I inhibitors’142 or in an extended conformation by interacting with D–F-G motif through allosteric position, i.e., DFG-out conformation are known as ‘Type II inhibitors’.143 NPS-1034, BMS777607, cabozantinib, and sitravatinib are type II multitargeted inhibitors that are intrinsically less selective compared to type I inhibitors. These small molecules are currently at various stages of clinical trial phases and hold the potential to improve clinical outcomes for patients.144,145

Anti-AXL monoclonal antibodies (mAb) and antibody–drug conjugates (ADC)

Many AXL inhibitors currently under investigation either exhibit cytotoxicity toward normal cells, lack specificity, or show only moderate effectiveness. To address these limitations, several monoclonal antibodies targeting AXL (such as Tilvestamab, YW327.6S2, CDX-0168, 20G7-D9, and MAb173) are being developed to enhance specificity and efficacy against cancer cells.73,114,115,146,147 Furthermore, antibody–drug conjugates (ADC) like Enapotamab Vedotin, DAXL-88, Mipasetamab Uzoptirine, and Mecbotamab Vedotin have been constructed. These ADCs deliver cytotoxic drugs via monoclonal antibodies, resulting in improved therapeutic efficacy (Table 1).148151 Recently, AXL-specific single-domain antibodies (sdAbs) have demonstrated both diagnostic potential and anticancer efficacy in acute myeloid leukemia (AML).152 The fusion of sdAbs to an Fc domain also grants them new therapeutic capabilities, potentially enabling a multidrug strategy for treating cancer patients who test positive for AXL.

Anti-AXL chimeric antigen receptor (CAR)-therapy

A novel cancer immunotherapy, anti-AXL chimeric antigen receptor (CAR)-T-cell therapy, has shown promise as a potential treatment option. For instance, an AXL-CAR and AXL synNotch receptors engineered using humanized single-chain variable fragment (scFv) have demonstrated the ability to target and kill tumor cells. When AXL-CAR is incorporated into human primary T cells, it enables targeted tumor cell destruction, while the AXL synNotch receptor can generate IL-10 in an antigen-specific manner.153 Similarly, AXL-CAR-T cells have shown antigen-specific cytotoxicity and cytokine release capacity against triple-negative breast cancer (TNBC).154 This therapy is precisely tailored to target AXL-positive osteomyeloid leukemic and TNBC cells,153,154 and is currently under clinical investigation for refractory stage IV renal cell carcinoma (RCC) (NCT03393936) and AXL-positive sarcoma (NCT05128786). However, due to the adverse side effects associated with CAR-T-cell treatment, its effectiveness in combating in battling solid tumors has been limited.155

To address this, AXL-CAR-T cells with constitutively activated IL7 receptor (C7R) have shown enhanced antitumor activity against TNBC cells both in vitro and in vivo, along with extended CAR-T-cell survival.156 In addition, the combination of AXL-CAR-T cells and microwave ablation (MWA) has exhibited improved antitumor activity in AXL-positive non-small-cell lung cancer (NSCLC) patient-derived xenograft (PDX) tumors by remodeling the tumor microenvironment (TME).157 AXL-CAR-T-cell immunotherapy for advanced AXL-positive lung cancer patients is currently being evaluated in phase I clinical trials to further assess its safety and efficacy (NCT03198052). Interestingly, GAS6-CAR-T cells have demonstrated effective targeting of TAM-positive pancreatic tumor cells and suppression of in vivo xenograft tumor growth with minimal toxicity by targeting both tumor cells and tumor-associated macrophages.158 In a recent study, Sakemura et al. found that inhibiting AXL with TP-0903 can modulate the function of CART19, enhancing its anticancer effect towards CD19+ cells through the suppression of Th2 cells and M2 macrophages. This finding suggests a novel impact of AXL inhibition on adoptive T-cell therapy’s immunological modulation.159 Furthermore, AXL-ADC and anti-AXL-CAR-T cells additionally showed potent and targeted cytotoxicity, improved primary tumor TME, and reduced lung metastasis in breast and lung cancer models, rendering it a plausible cancer prevention strategy.160 Consequently, extensive clinical research is ongoing to modify the CAR construct and explore various immune cell types as platforms to enhance the safety and effectiveness of CAR-based treatments. For instance, a phase I clinical trial involving 40 patients with AXL-positive advanced solid malignancies is underway to determine the safety and feasibility of targeting AXL with CAR-NK cells. In patients with advanced ovarian cancer or other AXL-expressing tumors, engineered CARs that target AXL are introduced into NK cells isolated from the patients’ peripheral blood.161 Some CAR-NK are further genetically modified to produce and release IL7/CCL19 and/or scFvs against PD-1/CTLA4/Lag3, while others are combined with cannabidiol or nicotinamide adenine dinucleotide to enhance their cytotoxic potential (NCT05410717).

AXL-mediated acquired therapeutic resistance

Acquired resistance to current cancer therapies frequently results in unfavorable clinical outcomes, leading to tumor relapse and metastasis. Molecular studies on recurrent tumors have revealed AXL’s multifaceted role in conferring resistance to a range of treatments, including conventional and targeted therapies.70,138

In ovarian cancer, AXL has been identified as a key mediator of resistance to cisplatin.162 Similarly, acute myeloid leukemia (AML) cells from patients who developed chemotherapy resistance exhibit significantly elevated AXL expression.108 Additionally, AXL has been implicated in mediating resistance to cetuximab in NSCLC and head and neck squamous cell carcinoma (HNSCC),163 highlighting its role across various treatment modalities.164166

The acquisition of secondary mutations that activate kinase signaling pathways is thought to be a primary driver of acquired therapeutic resistance, even in the presence of TKIs. These mutations can directly interfere with the effective interaction of TKIs and their target kinases by altering kinase conformation or ATP-binding affinity. While genetic alterations in the AXL gene, such as mutations (1.69%), fusions (0.04%), amplification (0.21%), or gene loss (0.09%), are relatively uncommon.28,167 AXL overexpression is frequently observed in multiple cancer types and is often associated with acquired therapeutic resistance.168

Although the precise molecular mechanisms underlying AXL-driven therapeutic resistance in the absence of genetic changes remain incompletely understood, they are thought to involve AXL-independent pathways. These may include the modulation of DNA damage and repair processes, phenotypic changes, activation of alternative signaling pathways, alterations in glucose metabolism, and interactions with other RTKs, as discussed in the following sections.

Modulating DNA damage and DNA damage response processes

Numerous studies have highlighted the role of AXL signaling in orchestrating DNA damage and DNA damage response (DDR) in drug-resistant cancers. The DDR involves complex mechanisms, including repair pathways, cell-cycle checkpoints, metabolic alterations, and apoptosis induction, all of which are critical for maintaining genomic stability. When these repair pathways are dysregulated, genomic instability arises, a hallmark feature of cancer closely associated with therapy resistance (Fig. 3).169

Fig. 3.

Fig. 3

AXL induces acquired therapeutic resistance by modulating DNA damage and repair processes. AXL induces acquired therapeutic resistance by modulating DNA damage and DNA damage response processes. AXL plays a vital role in maintaining this equilibrium by inhibiting DNA damage and regulating various DNA damage response mechanisms leading to error-prone DNA replication, genomic instability and inhibition of cell death mechanisms. (HR- Homologous Recombination, NHEJ- Non-Homologous End Joining)

Targeting AXL signaling presents a promising approach to enhance cell death by promoting DNA damage-induced activation of DDR pathways. In esophageal adenocarcinoma, AXL has been found to block apoptosis and foster drug resistance by suppressing c-ABL/p73 signaling in response to cisplatin.170 In addition, increased levels of AXL have been shown to stimulate ERK/mTOR signaling, triggering simultaneous DNA damage and repair signals. This promotes cell survival through CHK1 activation and confers resistance to WEE1 inhibition, a kinase involved in replication stress response.171

However, some studies indicate that AXL inhibition can sensitize cancer cells to other treatments. For example, study by Ramkumar et al. demonstrated that AXL inhibition activates ATR/CHK1 signaling, making NSCLC cells more susceptible to ATR inhibitors.172 Similarly, AXL inhibition has been observed to impede cell proliferation and induce G2/M arrest in response to DNA damage. This is evidenced by increased levels of γH2AX and other DDR markers, such as ATM/ATR/CHK2/CHK1, thereby sensitizing ovarian cancer cells to ATR inhibitors.173

Combining DDR and AXL signaling inhibitors has been found to have a synergistic effect in enhancing DNA damage, leading to increased cell death and improved therapy efficacy. For instance, simultaneous treatment with AXL and ATR inhibitors induces replication fork collapse and double-strand breaks (DSBs), as indicated by increased levels of phospho-RPA32 and γH2Ax proteins, ultimately leading to mitotic catastrophe.172 In ovarian cancer cells, combining chemotherapy with a GAS6/AXL inhibitor (AVB-500) enhances γH2AX and 53BP1 foci formation, suppresses RAD51 foci, and impedes replication fork progression compared to chemotherapy alone, thereby impairing homologous recombination repair (HRR). Conversely, the combination of a PARP inhibitor and AVB-500 accelerates replication fork progression, exacerbating DNA damage and genomic instability.174

Furthermore, AXL contributes to drug resistance by promoting neddylation and activation of RAD18, leading to the recruitment of low-fidelity DNA polymerases for error-prone DNA replication. This process accelerates the emergence of the T790M mutation in NSCLC cells.30 Additionally, AXL upregulates and activates MYC, resulting in a purine metabolic imbalance and increased adaptive mutations.32 In HER2+ breast cancer, replication stress response, therapy resistance, and metachronous metastasis have been linked to the nuclear AXL/WRNIP1 axis, indicating that this pathway could be therapeutically exploited.175 Overall, these findings highlight the complex interplay between AXL and DDR, suggesting that the simultaneous targeting of both pathways could offer significant therapeutic benefits.

Phenotypic changes

As discussed earlier, AXL is intricately involved in phenotypic changes such as epithelial-to-mesenchymal transition (EMT), a process closely associated with the development of therapeutic resistance in cancer cells (Fig. 4).176,177 Increased AXL expression has been shown to upregulate vimentin while concurrently downregulating E-cadherin and the epithelial-associated miR-34a, leading to EMT and subsequent resistance to erlotinib in head and neck cancers.178 In addition, the GAS6/AXL axis has been reported to enhance aggressiveness and resistance to doxorubicin in breast cancer cells through the Akt/GSK-3β/β-catenin/ZEB1 pathway.179 Similarly, in esophageal cancer, epirubicin-induced resistance is associated with AXL-mediated transcriptional upregulation of c-MYC expression via the Akt/β-catenin signaling pathways.180

Fig. 4.

Fig. 4

AXL-mediated phenotypic changes in cancer cells induce therapeutic resistance. AXL promotes epithelial-to-mesenchymal transition (EMT) and therapeutic resistance by inducing several EMT transcription factors, such as Snail, Slug, Smad-3, β-catenin, and TWIST1, as well as markers like MMP9, ZEB1, and Vimentin. The TGF-β/Smad-4/YAP pathway is reported to upregulate AXL expression, facilitating these phenotypic changes in cancer cells

Furthermore, the interaction between AXL and 14-3-3ζ in hepatocellular carcinoma (HCC) leads to the phosphorylation of SMAD3 at Ser213, upregulating TGF-β target genes such as Snail, PAI1, and MMP9, and promoting autocrine TGF-β secretion.181 Studies on HCC further show that AXL overexpression drives metastatic progression by activating mesenchymal markers like Slug and Vimentin, contributing to sorafenib resistance.182 In colorectal cancer, increased AXL expression induces EMT and enhances resistance to polo-like kinase (PLK1) inhibitors via TWIST1 overexpression.183

Notably, the EMT state can rewire AXL signaling, further inducing drug resistance. For instance, in ovarian cancer, AXL signaling is modulated by the EMT state, with mesenchymal subtypes augmenting GAS6/AXL signaling.184 Similarly, TGF-β/SMAD4 activation facilitates YAP-dependent AXL expression, contributing to acquired chemoresistance in mesenchymal lung cancer cells.185 Further investigations into these molecular mechanisms could reveal additional targetable pathways for AXL-EMT regulation.

AXL silencing has shown potential in overcoming drug resistance. For example, using a covalent siRNA-gelatin-antibody nanoconjugate to silence AXL inhibits mTOR/EMT signaling and activates p53, thereby sensitizing NSCLC cells to TKIs.186 In addition, miR-625-3p-mediated targeting of AXL reverses TGF-β/SMAD driven-EMT, enhancing gefitinib sensitivity in NSCLC.187 Recent findings by Chen et al. demonstrated that AXL-siRNA loaded paclitaxel-poly-L-lysine prodrugs, delivered via activated T-cell-derived exosomes, can overcome AXL-mediated paclitaxel resistance and immune evasion in triple-negative breast cancer (TNBC) cells by reversing EMT and stemness processes.188

Given AXL’s role in inducing and maintaining the EMT phenotype, stratifying patients based on their AXL dependence is crucial. This approach could potentially reverse EMT and sensitize cancer cells to other TKIs or chemotherapeutic drugs.

Activation of bypass signals

AXL has been implicated in resistance to various therapeutic agents, particularly those targeting receptor tyrosine kinases (TKIs). It maintains pathway functionality by employing alternative effectors, activating alternate signaling networks, or disrupting feedback signals (Fig. 5).189 In EGFR-mutant NSCLC, upregulation of AXL serves as a bypass mechanism, contributing to acquired resistance to first-generation EGFR TKIs. This resistance is partly through the activation of MAPK/ERK and PI3K/Akt pathways.190,191 Furthermore, Saab et al. proposed a novel perspective on how YAP and AXL contribute to EGFR-TKI resistance, amplifying signaling through a feed-forward system.192

Fig. 5.

Fig. 5

AXL-directed therapeutic resistance via activation of bypass signaling. a Under basal conditions, AXL is shed through the action of ADAM10/17. b In bypass signaling, MAPK inhibition decreases sheddase activity by enhancing the association between TIMP1 and ADAM10/17. This inhibition boosts mitogenic signaling through bypass kinases like JNK, leading to sustained secondary tolerance to EGFR TKIs

Resistance to osimertinib, a third-generation EGFR-TKI, has been associated with the reactivation of multiple signaling pathways, including the AXL/CDCP1/SRC/Akt and STC2/JUN/AXL/ERK network in lung cancer.1,193 In addition, overexpression of AXL in NSCLC cells has been shown to induce drug-resistant persister cell phenotypes through the activation of autophagy.194 Another mechanism contributing to enhanced bypass signals and cancer drug resistance is extracellular proteolytic rewiring caused by blocking AXL receptor shedding.195 Complementing this, AXL activation and delayed degradation have been identified as critical pathways through which chemotherapeutic agents confer preserved secondary tolerance to EGFR TKIs.196

Given the insights, a potentially viable treatment approach for overcoming and delaying acquired resistance in cancer may involve simultaneously targeting AXL and other RTKs that share similar downstream signaling networks through combination therapy.197,198

Altering glucose metabolism

Recent studies have highlighted the role of AXL in modulating glucose metabolism in cancer cells. The AXL/TNS2/IRS-1 crosstalk has been shown to influence glucose metabolism in pancreatic cancer cells by upregulating glycolytic enzymes such as Glut4 and PDK1 (Fig. 6).199 In the context of drug resistance, Tian et al. demonstrated that inhibiting AXL enhanced the chemosensitivity of ovarian cancer cells to cisplatin by reducing glycolysis. This effect is attributed to decreased phosphorylation of pyruvate kinase isoform M2 (PKM2) at Y105.132 Similarly, GAS6-AXL inhibition was found to impair the glycolytic capacity of endometrial cancer cells by reducing PI3K/Akt activation, thereby increasing sensitivity to paclitaxel.200

Fig. 6.

Fig. 6

AXL alters glucose metabolism to develop acquired resistance in cancer cells. AXL-GAS6 signaling promotes glycolysis through multiple pathways, reducing sensitivity to cytotoxic stress and fostering chemoresistance. AXL-mediated PI3K/Akt signaling activates glycolysis. After AXL phosphorylates TNS2 at Y483, TNS2 becomes more stable and disengages from IRS-1. The overexpression of GLUT4 and PDK1, resulting from AXL/TNS2/IRS-1 crosstalk, enhances related metabolism. Additionally, AXL phosphorylates PKM2 at Y105, reducing its ability to bind phosphoenolpyruvate and further promoting glycolysis

Although these studies provide a foundation, further research across different cancer types is needed to determine the precise molecular mechanisms through which AXL alters glucose metabolism to induce drug resistance.

Crosstalk of AXL with other RTKs

AXL engages in significant crosstalk with other RTKS in both the presence or absence of GAS6, amplifying pro-tumorigenic signaling. In mesenchymal ovarian cancer cells, GAS6-activated AXL forms clusters with RTKs like c-MET, HER2, and EGFR, driving motility through ERK activation and the transcription factors FRA1 and Slug.184 Similarly, the co-expression of AXL and PDGFR has been linked to increased invasiveness and metastasis in thyroid, esophagus, and bladder carcinomas.201203

Recent studies have identified various mechanisms through which AXL interacts with other RTKs, to promote cancer cell survival, proliferation, invasiveness, and drug resistance. These mechanisms include heterodimerization,204 transphosphorylation,204 activation through intermediate molecules,101 receptor stabilization,85 compensation for the loss of one RTK by another,205 and co-expression of RTKs in specific cancers (Fig. 7).206 Some of them are discussed in the following sections.

Fig. 7.

Fig. 7

AXL crosstalk with different receptor tyrosine kinases promotes drug resistance. AXL interacts with several oncogenic receptor tyrosine kinases (RTKs), such as TYRO3 and MERTK, and heterodimerizes with several non-TAM receptors like EGFR, MET, PDGFR, and FLT3. It also forms dimers with HER2/3. These interactions activate downstream signaling pathways, leading to cancer cell proliferation, metastasis, angiogenesis, and drug resistance. Multiple tyrosine kinase inhibitors (TKIs), monoclonal antibodies (mAbs), and antibody–drug conjugates (ADCs) shown have potential in overcoming therapeutic resistance in various cancers

AXL-TYRO3

TYRO3, a member of the TAM RTK family, shares the ligands GAS6 and Protein S with AXL. Although the signal transduction pathways of TYRO3 are not fully explored, its overexpression and activation have been linked to enhanced cell survival, proliferation, metastasis, and therapy resistance in various cancers.207

A study by Brown et al. identified a functional connection between TYRO3 and AXL, demonstrating that overexpressing kinase-active TYRO3 in Rat2 cells leads to increased GAS6-mediated AXL phosphorylation, independent of AXL expression levels. Moreover, the co-expression of TYRO3 and AXL in these cells significantly enhances proliferation, up to fivefold, upon GAS6 treatment. AXL overexpression also promotes transphosphorylation of both kinase-active and kinase-dead forms of TYRO3. This cross-phosphorylation modulates downstream of MAPK and PI3K signaling pathways, highlighting how RTK crosstalk alters cellular signaling.202

Both AXL and TYRO3 are overexpressed in thyroid cancer cells and leiomyosarcoma cells, contributing to cell growth and resistance to apoptosis.208,209 Investigating the interplay between AXL and TYRO3 in cancer progression could offer insights into the varying responses of different cancer cells to AXL inhibitors.

AXL-MERTK

MERTK, another member of the TAM family of RTKs, plays a physiological role in regulating innate immunity, platelet aggregation, tissue homeostasis and repair. However, its aberrant expression in various cancers has been implicated in promoting neoplasia processes such as growth factor independence, increased proliferation, resistance to apoptosis, and metastasis.210

While AXL and MERTK do not directly interact, their downstream signaling pathways significantly overlap. MERTK is primarily expressed in cells of the hematopoietic lineage, including monocytes, macrophages, dendritic cells, NK cells, and platelets, whereas AXL expression is mainly confined to epithelial and endothelial tissue and is linked to EMT.99,189 Consequently, MERTK is more closely associated with the tumor microenvironment, while AXL is linked to both tumor cells and the tumor microenvironment. This cooperation between AXL and MERTK has been shown to accelerate the growth of breast cancer cells by combining oncogenic signaling and evading host antitumor immunity.211

AXL and MERTK are frequently overexpressed in multiple cancers including head and neck, lung, colon, lung, skin, sarcoma, and glioblastoma multiforme.212216 McDaniel et al. identified MERTK upregulation as an inherent and adaptive feedback mechanism that leads to resistance against anti-AXL agents. However, co-targeting both MERTK and AXL significantly inhibited tumor growth by affecting several signaling pathways, including Akt, STAT6, P70S6K, S6RP, and C-RAF.215 Interestingly, the dual inhibitor INCB081776 demonstrated immunomodulatory and antitumor effects primarily by inhibiting MERTK activation in macrophages. This effect partially counteracted M2 macrophage-mediated inhibition of T-cell proliferation and suppression of phospho-Akt.216

These findings suggest that co-targeting the TAM RTKs could confer dual anticancer effects. First in the tumor cells, where AXL inhibition would disrupt developmental, proliferative, and migratory pathways, and second, in the tumor microenvironment, where MERTK suppression would modulate the innate immune response.

AXL-c-MET

The c-MET RTK is a proto-oncogene known to derive the invasion and metastasis of various cancers.215,217 In clear cell renal cell carcinoma (ccRCC), GAS6/AXL signaling enhances cellular invasion through lateral stimulation of c-MET, mediated by another proto-oncogene, SRC.57 AXL and c-MET crosstalk has been observed in triple-negative breast cancer (TNBC), where exposure to hepatocyte growth factor (HGF) leads to phosphorylation of both c-MET and AXL. Interestingly, AXL knockdown does not affect the HGF-mediated phosphorylation of downstream proteins, indicating a complex interplay between c-MET and AXL. Conversely, GAS6 exposure results only in AXL phosphorylation, c-MET knockdown alters the timing and intensity of GAS6-mediated signaling. Despite these complexities, Reverse Phase Protein Arrays (RPPAs or RPLAs) analyses suggest that AXL and c-MET receptors exhibit crosstalk and may physically interact in breast cancer cells.218

In addition, Li et al. identified that the clustering of c-MET, AXL, ELMO2, and DOCK180 proteins on the plasma membrane constitutes a crucial mechanism through which HGF-dependent signaling initiates RAC1-mediated migration and invasion of glioblastoma cells.219 Notably, in the HGF/c-MET pathway, AXL (in its higher-molecular-weight form, p140) undergoes transphosphorylation at Tyr-779, potentially playing a significant role in tumorigenesis. This contrasts with the phosphorylation of the lower-molecular-weight AXL (p120) by GAS6 at Y702.219

Overall, these findings suggest that using inhibitors targeting both AXL and c-MET may be a promising strategy for preventing the progression and metastasis of cancers driven by AXL/c-MET signaling.

AXL-EGFR

Epidermal growth factor receptor (EGFR) is primarily involved in maintaining epithelial tissue. However, aberrant EGFR signaling can lead to epithelial transformation, making it one of the most studied tyrosine kinase receptors in cancer.220 Despite the clinical efficacy demonstrated by various EGFR-targeted therapies, acquired resistance often limits long-term effectiveness in cancer treatment.

Interestingly, AXL has been found to influence acquired resistance to EGFR-targeted therapies in both breast and lung cancers. In triple-negative breast cancer (TNBC), EGFR-mediated transactivation of AXL reduces responsiveness to EGFR-targeted therapies by diversifying EGFR signaling, including its interactions with c-MET and PDGFR, but not with IGF1R (insulin-like growth factor-1 receptor) or INSR (insulin receptor).221 In addition, co-expression of AXL and EGFR is frequently observed in lung adenocarcinomas, regardless of EGFR mutation status.206,222 Furthermore, crosstalk between AXL and EGFR allows squamous cell carcinomas to bypass the anti-proliferative effects of PI3Kα inhibitors by activating the PLCγ-PKC/mTOR signaling pathway.133 In glioblastoma multiforme, this hetero-interaction between AXL and EGFR facilitates EGFR-mediated pro-invasive signaling.51

Examining the co-expression patterns of EGFR and AXL in various cancers could provide valuable insights into potential biomarkers. Additionally, the concurrent targeting EGFR and AXL may enhance the therapeutic efficacy of EGFR-targeted therapies and help mitigate therapeutic resistance.

AXL-HER2

Another mechanism of RTK cross-communication involves the stabilization of one receptor by another, as seen with AXL and human epidermal growth factor receptor 2 (HER2). HER2, a member of the epidermal growth factor receptor family, plays a significant role in the etiology of various human cancer by regulating cellular proliferation and differentiation. Approximately 20% of breast cancer cases exhibit HER2 gene amplification and overexpression.223

In HER2-positive breast cancer cells, AXL physically interact with HER2, facilitating AXL’s stability and recruitment to the cell surface. This interaction leads to AXL transphosphorylation by HER2, promoting metastatic processes such as intravasation, growth of metastatic cells, and extravasation.85 In addition, AXL stabilizes HER2 and regulate Hypoxia-inducible factor-1 alpha (HIF-1α) levels under hypoxic conditions, potentially through activation of the PI3K/Akt pathway, which drives metastasis and modulates the tumor microenvironment in HER2-positive breast cancer cells.224 The crosstalk between AXL and HER2 also activates the PI3K/Akt and ERK pathways, enabling HER2-positive breast cancer cells to develop resistance against lapatinib, an inhibitor of HER2/neu and EGFR pathways.225

Interestingly, AXL promotes EMT and therapeutic resistance in breast cancer by heterodimerizing with HER2, thereby inducing PI3K/Akt and MAPK signaling independent of its ligands. This suggests that targeting HER2-AXL heterodimerization could be crucial for preventing HER2 inhibitor-mediated therapeutic resistance, especially in breast cancer.137

AXL-HER3

Human epidermal growth factor receptor 3 (HER3) is a member of the HER-RTK family, with neuregulins 1 (NRG1) and 2 (NRG2) as its primary ligands. Although HER3 has low intrinsic tyrosine kinase activity, it can interact with other RTK or non-RTK members to increase transphosphorylation and activate downstream pathways, ultimately promoting metastasis and drug resistance.226

AXL and HER3 collaborate to counteract the effects of AXL inhibitors (such as BMS777607 or R428) in MDA-MB231 and Ovcar8 cells. when AXL activity is lost, HER3 transcription is upregulated, leading to NRG1-dependent phosphorylation (Y1289) as a compensatory mechanism. Notably, the activation of phospho-Akt (S473) at low basal levels is typically required for HER3 activation upon AXL inhibition.227

In addition, crosstalk between AXL and HER3 promotes invasion and metastatic progression in melanoma cells by regulating invadopodia formation.205 In NSCLC, AXL interacts with HER3 with EGFR within a negative feedback system involving SPRY4, resulting in resistance to the EGFR inhibitor osimertinib.222,228 In head and neck cancer, AXL mediates cetuximab resistance through two mechanisms: activation of c-ABL kinase via AXL’s Y821 phosphorylation229 and HER3 activation due to upregulated NRG1.31

This predictive relevance of the AXL and HER3 association highlights the importance of developing dual-targeted inhibitors or combination therapies to enhance treatment efficacy against acquired resistance.

AXL-VEGFR

Vascular endothelial growth factor receptor (VEGFR), a member of the RTK family, plays a pivotal role in angiogenesis, with its three primary subtypes identified as VEGFR1, VEGFR2, and VEGFR3. RTKs can communicate indirectly through various signaling intermediates. One such example is how VEGFR2 activates Src family kinases via TSAd (T-cell-specific adaptor protein), promoting ligand-independent autophosphorylation of AXL at YXXM motifs. This activation facilitates interaction with PI3K, stimulating Akt signaling. Consequently, activated AXL contributes to VEGF-A/VEGFR2-dependent cancer cell migration, tube formation, vascular permeability, and corneal neovascularization.101

Conversely, GAS6 impedes ligand-dependent VEGFR2 activation and angiogenic signaling in vascular endothelial cells by inducing the phosphorylation of SHP-2 (Src homology region 2-containing protein tyrosine phosphatase 2) and facilitating its interaction with AXL.230 This intricate regulatory mechanism underscores the complexity of crosstalk between AXL and VEGFR in modulating angiogenic programs.

In addition, the localization and expression of AXL and VEGFR exhibit significant variation between monolayer and spheroid ovarian cancer models. For instance, OVCAR8 spheroids show a tenfold higher concentration of VEGFR1 on the plasma membrane and greater heterogeneity compared to monolayers, characterized by a bimodal distribution of low and high-AXL subpopulations. This variation, including a 100-fold difference in plasma membrane AXL concentration between chemo-sensitive and chemoresistance cells,231 highlights the importance of selecting appropriate cancer models for therapeutic screening.

In conclusion, comprehensive studies across various cancers are needed to unravel the precise molecular mechanisms underlying GAS6/AXL crosstalk with VEGFR and its regulation of angiogenic programs.

AXL-FLT3

The Fms-related receptor tyrosine kinase 3 (FLT3) gene, a class III RTK, plays a crucial role in hematopoiesis regulation. Upon binding to its ligand, FLT3 undergoes homodimerization on the plasma membrane, leading to autophosphorylation and initiation of downstream signaling pathways involved in cell proliferation, apoptosis, and differentiation in bone marrow. Mutations that result in constitutive activation of FLT3 are notably implicated in myeloid and lymphoblastic leukemia.232

AXL significantly influences the progression of FLT3 internal tandem duplication (ITD)-positive acute myeloid leukemia (AML) by promoting constitutive FLT3 phosphorylation.233 In FLT3-ITD positive AML cells, elevated AXL expression promotes cell growth, increases therapy resistance, and suppresses apoptosis and differentiation through regulation of downstream molecules such as Akt, STAT5, and ERK.151,234 Notably, inhibition of AXL activity has shown a marked reduction in the development of resistance in AML against FLT3 inhibitors, including PKC412 and AC220.235 Furthermore, dual inhibition of FLT3 and AXL using Gilteritinib in FLT3-ITD AML has demonstrated effectiveness in overcoming resistance mechanisms driven by the hematopoietic niche.236

The potential of FLT3-ITD inhibitors to induce apoptosis is influenced by the GSK-3β/Mcl-1 axis, which is regulated by both FLT3-ITD and AXL signaling pathways.237 Overall, combined inhibition of FLT3 and AXL presents a promising therapeutic strategy for leukemia, offering new alternatives to overcome resistance and improve patient outcomes.

AXL-PDGFR

Platelet-derived growth factors (PDGFs) and their receptors (PDGFRs) are expressed in various cancers, where PDGF/PDGFR signaling activation influences cell proliferation, migration, invasion, and angiogenesis through the PI3K-PKB/Akt and MAPK/ERK pathways.238 AXL and PDGFR demonstrate crosstalk, as observed in gastrointestinal stromal tumors (GISTs) treated with imatinib mesylate. In this context, inhibition of α-PDGFR (KIT) by imatinib mesylate leads to AXL overexpression, resulting in resistance to therapy.239

In human bladder cancer, c-MET has been shown to promote disease progression by transactivating the expression of AXL and PDGFR-α via MEK/ERK1/2 signaling, independent of Ras and Src.202 Similarly, in triple-negative breast cancer cells, AXL interacts with other ErbB receptor family members, c-MET, and PDGFR contributing to resistance against EGFR TKIs.221

Further exploration of the mechanisms underlying receptor switching from PDGFR to AXL in different cancers could provide valuable insights into the signaling networks and help overcome acquired resistance.

Overall, the crosstalk between AXL with other RTKs contributes to resistance against both conventional and targeted therapies across diverse cancers. Therefore, co-targeting AXL along with other RTKs presents a promising strategy to address the challenges associated with RTK signaling rewiring and improve cancer management.

Combination therapies can potentially prevent therapeutic resistance

Given that AXL plays a crucial role in cancer progression and therapeutic resistance, combining AXL inhibitors with other receptor tyrosine kinase inhibitors offers a promising strategy for overcoming acquired resistance and enhancing drug sensitivity. Consequently, ongoing research is exploring the effectiveness of various combinations of AXL inhibitors with chemotherapies, targeted therapies, and immunotherapies. These combinations can be broadly classified into selective or multitargeted AXL inhibitor approaches.

Selective AXL inhibitor combinations

Bemcentinib combinations

Bemcentinib, the first selective small-molecule AXL inhibitor to receive US FDA fast-track designation, has shown promising results in combination studies across various cancer types. Its potent inhibition of both ligand-dependent and ligand-independent AXL signaling makes it an attractive candidate for combination therapies aimed at overcoming resistance and improving treatment outcomes.

In uterine serous carcinoma, bemcentinib enhanced sensitivity to paclitaxel chemotherapy by inhibiting AXL activation.240 Similarly, in diffuse intrinsic pontine glioma, and aggressive colorectal adenocarcinoma, combining bemcentinib and agents like panobinostat (an HDAC inhibitor) and galunisertib (a TGF-β inhibitor) showed synergistic effects in reversing mesenchymal transition and reducing the mesenchymal phenotype, respectively.241,242

In NSCLC, bemcentinib combined with VX-970 (an ATR inhibitor) enhanced DNA damage and replication stress, increasing susceptibility to ATR inhibitors.172 In addition, in metastatic melanoma, bemcentinib improved the therapeutic efficacy of the BRAF inhibitor vemurafenib by triggering apoptosis, accelerating ferroptosis, and reducing autophagy.243 In older AML patients who cannot tolerate intense chemotherapy, bemcentinib showed an anti-leukemic effect when combined with low-dose cytarabine or decitabine (NCT02488408). Parallel studies have shown that combining bemcentinib with erlotinib (NCT02424617), docetaxel (NCT02922777), and pembrolizumab (a humanized antibody that targets PD-1; NCT03184571) is viable and tolerable in advanced NSCLC.

A phase Ib/II trial is currently studying the combination of bemcentinib with pembrolizumab or dabrafenib/trametinib in advanced non-resectable or metastatic myeloma (NCT02872259). Similarly, the combination of bemcentinib and pembrolizumab is being evaluated in a phase II study for mesothelioma compared to other targeted treatments (NCT03654833).

While bemcentinib/pembrolizumab has demonstrated a favorable clinical response in several trials, its investigation in refractory TNBC was halted due to the lack of complete or partial response in patients (NCT03184558). The encouraging results from these combination studies collectively support further exploration of bemcentinib in conjunction with other small-molecule inhibitors or immunotherapeutic agents for the effective treatment of various human malignancies.

TP-0903 combinations

In small-cell lung cancer, combining WEE1 inhibitors (such as AZD1775) and AXL inhibitors (like TP-0903, an oral selective AXL inhibitor) or with mTOR inhibitors (such as RAD001) can overcome resistance to WEE1 inhibition driven by downstream activation of mTOR and CHK1 which are involved in DNA damage repair pathways.171 Additionally, the combination of AXL (TP-0903) and JAK1 (ruxolitinib) inhibition has been identified as a novel therapeutic strategy based on single-cell proteomic profiling in lung cancer.244 Our lab has demonstrated that silencing AXL (either through gene knockdown or using TP-0903), effectively inhibits therapeutic resistance against cabozantinib in renal cell carcinoma both in vitro and in vivo.245247 Moreover, targeting the heterodimerization of AXL-HER2 with TP-0903 and trastuzumab (a HER2 inhibitor) in breast cancer has been observed to overcome resistance to HER2 blockade.137

The preliminary clinical activity of TP-0903 in combination with decitabine has also been reported in the phase 1b/2 trial (NCT03013998) involving patients aged ≥60 years with prognostically poor TP53 mutant/complex karyotype AML.248 Together, these studies support the use of TP-0903 in combinations with various therapeutics to overcome the emergence of therapeutic resistance.

Antibody–drug conjugate combinations (CAB-AXL-ADC & Enapotamab vedotin)

Several clinical trials are investigating the safety and efficacy of antibody–drug conjugates (ADCs), such as CAB-AXL-ADC. These trials are examining its effectiveness in combination with PD-(L)1 inhibitor against various cancers, including sarcomas (phase I, NCT03425279), NSCLC (phase II, NCT04681131) and ovarian cancers (phase II, NCT04918186).

Recently, Boshuizen et al. also demonstrated the role of another AXL-targeting antibody–drug conjugate (Enapotamab vedotin) in potentiating anti-PD-1 (pembrolizumab) therapy in melanoma and lung cancer models resistant to immunotherapy. This combination results in a significant benefit from de novo immune checkpoint inhibition.249

Multitargeted AXL inhibitor combinations

The development of multitargeted AXL TKI combinations has gained significant attention in recent years. These combinations aim to disrupt multiple independent significant pathways, reducing the likelihood of resistance development. For instance, TAM receptors, which include AXL, are part of complex signaling networks that being explored for combination therapies. These approaches seek to mitigate resistance to various chemotherapeutic agents and enhance treatment efficacy.189

ONO-7475 combinations

ONO-7475, a dual inhibitor targeting AXL and MERTK, has shown potential in combination therapies for overcoming resistance in various cancers. In FLT3-ITD-dependent AML, ONO-7475 can synergize with sorafenib (a multikinase inhibitor) and venetoclax (a BCL-2 inhibitor) to address resistance to FLT3 and BCL-2 inhibitors, respectively.250,251 In addition, combining ONO-7475 with Osimertinib (an EGFR inhibitor) has demonstrated efficacy in suppressing resistance in EGFR-mutated NSCLC cells that overexpress AXL.252

Despite these promising results, some clinical trials have faced challenges. The phase I/II trial (NCT03176277) exploring ONO-7475 alone or with venetoclax in AML was terminated due to protocol-defined futility criteria. Similarly, the dose-escalation study of ONO-7475 monotherapy, and in combination with anti-PD-1 therapy (ONO-4538/Nivolumab) in advanced solid tumors (NCT03730337) was also suspended following temporary completion of enrollment.

PF-07265807 combinations

PF-07265807, an inhibitor of AXL and MERTK, has demonstrated promising antitumor activity in preclinical models, particularly when combined with PD-1 inhibitors.253 A clinical trial (NCT04458259) is currently assessing the safety, tolerability, pharmacokinetics, and early antitumor efficacy of PF-07265807 as a single agent and in combination with sasanlimab (an anti-PD-1 antibody), with or without axitinib (a VEGFR inhibitor), in patients with advanced or metastatic renal cell carcinoma.254

In addition, Q702, a selective inhibitor targeting AXL, MERTK, and CSF1R, has shown favorable pharmacologic activity and an acceptable safety profile in patients with advanced solid tumors.255 This inhibitor is being evaluated in a clinical trial (NCT04648254).

Recently studies have also explored the use of AXL, TAM, and non-TAM RTK inhibitors in combination therapies to reverse acquired resistance to EGFR TKIs and improve the outcomes of immunotherapy. These combinations often show greater efficacy compared to TAM receptor-specific inhibitors alone.

Cabozantinib combinations

Cabozantinib, a multikinase inhibitor targeting VEGFR2, c-MET, RET, and AXL, has shown efficacy in overcoming resistance to other therapies across various cancers. In renal cancer, prolonged sunitinib treatment increases c-MET and AXL expression, leading to acquired resistance against sunitinib, which can be countered with cabozantinib.256 Similarly, cabozantinib combined with gefitinib has been effective in countering resistance to crizotinib and ROS1-TKIs in lung cancers with ROS1 fusions, where active HB-EGF/EGFR and AXL signaling contribute to resistance.257

Cabozantinib combined with osimertinib has also shown potential in overcoming AXL-mediated resistance in EGFR-mutated NSCLC.168 Multiple trials (NCT00596648, NCT01866410, and NCT01708954) are exploring the combination of cabozantinib with erlotinib to address EGFR-TKI resistance in NSCLC, while cabozantinib combined with panitumumab has demonstrated a favorable safety profile and encouraging activity in RAS wild-type colorectal cancer (NCT02008383).258

Cabozantinib is also being investigated in combination with immunotherapies. The COSMIC-021 phase 1b trial is evaluating cabozantinib alone or with atezolizumab in various solid tumors. In metastatic castration-resistant prostate cancer, cabozantinib and atezolizumab have shown a tolerable safety profile and clinically relevant activity (NCT03170960).259 The CONTACT-2 phase III trial reported that the combination significantly improved progression-free survival compared to abiraterone, enzalutamide, and prednisone (NCT04446117).260

In renal cell carcinoma, the combination of cabozantinib and nivolumab has improved overall survival compared to sunitinib and showed durable responses in advanced cases (NCT03141177).261 The COSMIC-313 trial showed adding cabozantinib to nivolumab and ipilimumab significantly prolonged progression-free survival in advanced renal cell carcinoma (NCT03937219).262 Ongoing trials, including the PDIGREE study (NCT03793166) and a phase I/II trial (NCT01658878), continue to explore these combinations, highlighting Cabozantinib’s versatility in treating various cancers and improving outcomes.

Sitravatinib combinations

Sitravatinib, a receptor kinase inhibitor targeting TYRO3, AXL, MERTK, and VEGFR2, is being studied in combination with nivolumab for various cancers. Ongoing clinical trials focus on its efficacy in locally advanced clear cell renal cell carcinoma (ccRCC) (NCT03015740; NCT04904302; NCT03680521), advanced NSCLC (NCT02954991; NCT03906071) and oral cancer (NCT03575598). A phase I study is also exploring the combination of sitravatinib with nivolumab and ipilimumab in ccRCC and other solid malignancies (NCT04518046). Sitravatinib combined with the anti-PD-1 antibody tislelizumab has demonstrated tolerance and early anticancer activity in patients with unresectable advanced hepatocellular carcinoma and gastric cancer (NCT03941873) and promising results in TNBC263 (NCT04734262), advanced melanoma264 and NSCLC265 (NCT03666143). A phase III trial is testing sitravatinib and nivolumab in advanced NSCLC (NCT04921358), while a phase II trial investigates this combination in advanced biliary tract cancer (NCT04727996). However, a phase II study in urothelial carcinoma found a tolerable safety profile but no significant objective response (NCT03606174), with 51.2% of patients experienced grade 3 treatment-related adverse events (TRAEs) 3.3% encountered grade 4 TRAEs, and one patient died from cardiac failure.266 The phase III SAPPHIRE trial also failed to show improved survival for sitravatinib plus nivolumab compared to docetaxel in advanced NSCLC (NCT03906071).267 Additional multitargeted AXL inhibitors are in development. NPS-1034, a dual inhibitor of AXL/MET, enhances gefitinib or erlotinib sensitivity in EGFR-mutant NSCLC.268 Ningetinib, an inhibitor of c-MET, AXL and VEGFR2 showed promise when combined with gefitinib in EGFR-TKI-resistant T790M-negative NSCLC (NCT03758287). CB469, another AXL/MET inhibitor, demonstrated efficacy with EGFR TKIs in NSCLC with acquired resistance from AXL and c-MET activation.269

Anlotinib combined with osimertinib has also been found to sensitize osimertinib-resistant NSCLC by targeting c-MET/MYC/AXL signaling.270

In addition, BMS777607, a selective MET kinase inhibitor, combined with lapatinib, overcame HER3-induced resistance to AXL TKIs.227 S49076, an inhibitor of MET, AXL, and FGFR, enhanced radiotherapy efficacy in MET-dependent and independent models.271

These studies support the ongoing exploration of AXL inhibitors in combination therapies to improve safety and overcome resistance in various cancers.

Future prospects and conclusions

Considering AXL’s pivotal role in cancer initiation, progression, metastasis and therapeutic resistance, future research holds promise in uncovering the translational potential of AXL as a biomarker for predicting treatment responses. The numerous ongoing clinical trials involving various AXL inhibitors will help assess the clinical benefits of targeting AXL. However, several key questions about AXL remain that necessitate further investigations.

Since DNA changes are infrequent and may not reliably indicate sensitivity to AXL inhibitors, unlike targeted therapies such as EGFR inhibitors, it is crucial to explore whether the upregulated expression of AXL, often associated with poor prognosis, can serve as a reliable biomarker for identifying patients who would benefit most from AXL-targeted therapies. In addition, more research is needed to understand the epigenetic alterations governing AXL expression.

The clinical utility of AXL TKIs is limited by inherent drawbacks, such as off-target side effects resulting from the inhibition of multiple kinases. These side effects can lead to unforeseen toxicities, especially with multitargeted AXL inhibitors.272,273 Furthermore, the inevitability of relapse in patients with AXL-driven cancers highlights the need for high throughput screening strategies and in-depth understanding of resistance mechanisms. Such efforts are crucial to develop new treatment approaches that can mitigate potential toxicities and suppress drug resistance, resulting in more sustained therapeutic responses.

AXL’s involvement in immunological evasion and resistance to immune checkpoint blockade, mediated through regulation of the tumor microenvironment (TME), presents another significant challenge, as the survival advantage conferred by these therapies is limited to a small percentage of patients.117 Therefore, it is essential to delineate the mechanistic role of AXL signaling within different populations of stromal cells to better understand TME-induced antitumor defenses.

Future research should also focus on various variables affecting TME metabolism, including nutritional deprivation, pH dysregulation, hypoxia, and oxidative stress, as these factors may influence AXL expression and the selection of cells that express it.

Beyond its role in immune suppression, AXL is also implicated in EMT and cancer stemness, suggesting that inhibiting AXL could provide insights into these processes in metastatic disease and their impact on chemosensitivity and resistance. Trials investigating the potential of AXL inhibition to prevent resistance to conventional or targeted therapies are currently underway. Furthermore, non-invasive diagnostics using AXL-expressing CTC detection in future clinical studies could significantly improve personalized cancer therapy decision-making. Beyond cancer, AXL’s role in viral entry and replication (e.g., Ebola and SARS-CoV-2) and its impact on immune response modulation in comorbid conditions requires further investigation. AXL contributes to tissue homeostasis by regulating cytokine secretion and facilitating apoptotic cell clearance. Loss of AXL can lead to heightened inflammation and delayed recovery from immune-mediated damage, suggesting its potential involvement in cancer-related comorbid conditions.

Exploring the crosstalk and cooperation between AXL and other RTKs is vital for understanding the precise downstream signaling mechanisms underlying cancer cell survival. This understanding is necessary to design rational single-agent, anti-AXL-CAR or combination therapies that enhance antitumor defense and personalized treatment strategies.

In conclusion, while numerous cancer studies have demonstrated the potential of AXL as a biomarker and therapeutic target, further research is warranted to ascertain the clinical utility of AXL-based diagnostic and treatment tools.

Acknowledgements

M.Y. is supported by MK Bhan- Young Researcher Fellowship Programme (YRFP), Department of Biotechnology (DBT), Ministry of Science & Technology, Government of India. S.T. is supported by MacNaught Fellowship at DFCI. T. K. Choueiri is supported in part by the Dana-Farber/Harvard Cancer Center Kidney SPORE (2P50CA101942-16) and Program 5P30CA006516-56, the Kohlberg Chair at Harvard Medical School and the Trust Family, Michael Brigham, Pan Mass Challenge, Hinda and Arthur Marcus Fund and Loker Pinard Funds for Kidney Cancer Research at DFCI. S.P. acknowledges National Institutes of Health Grants (RO1 CA193675 and RO1 CA222355). A.Sa. acknowledges the Dana-Farber/Harvard Cancer Centre (DF/HCC), Kidney Cancer SPORE, Career Enhancement Award (CEP) 5P50CA101942-18 subaward. The figures are created with Biorender.com.

Author contributions

M.Y., T.K.C., S.P., and A.Sa. conceptualized the study. R.P.S., T.K.C., S.P., and A.Sa. supervised the study. Funding acquisition was done by M.Y., S.T., T.K.C., S.P., and A.Sa. Investigation—M.Y., A.Sh., K.P., J.S., and S.T. Methodology—M.Y., S.P., S.T., T.K.C., and A.Sa. Project administration: M.Y., S.P., and A.Sa. Resources: R.P.S., T.K.C., S.P., and A.Sa. Validation—M.M., M.E., L.R., S.T., and R.P.S. Visualization—M.M., M.E., L.R., and R.P.S. Writing—original draft—M.Y., A.Sh., K.P., S.T., and A.Sa. Writing—review and editing—T.K.C., S.P., and A.Sa. M.Y., A.Sh., J.S., R.P.S., L.R., and S.T. have created the figures. Figures were further discussed, reviewed, and edited by M.Y., S.T., S.P., and A.Sa. S.P. and A.Sa. reviewed and prepared the final manuscript. All authors have read and approved the article.

Competing interests

Dr. Choueiri reports institutional and/or personal, paid and/or unpaid support for research, advisory boards, consultancy, and/or honoraria past 5 years, ongoing or not, from: Alkermes, Arcus Bio, AstraZeneca, Aravive, Aveo, Bayer, Bristol Myers-Squibb, Calithera, Circle Pharma, Deciphera Pharmaceuticals, Eisai, EMD Serono, Exelixis, GlaxoSmithKline, Gilead, HiberCell, IQVA, Infinity, Ipsen, Jansen, Kanaph, Lilly, Merck, Nikang, Neomorph, Nuscan/PrecedeBio, Novartis, Oncohost, Pfizer, Roche, Sanofi/Aventis, Scholar Rock, Surface Oncology, Takeda, Tempest, Up-To-Date, CME events (Peerview, OncLive, MJH, CCO and others), outside the submitted work; Institutional patents filed on molecular alterations and immunotherapy response/toxicity, and ctDNA; Equity: Tempest, Pionyr, Osel, PrecedeBio, CureResponse, InnDura Therapeutics, Primium; Committees: NCCN, GU Steering Committee, ASCO (BOD 6-2024-, ESMO, ACCRU, KidneyCan; Medical writing and editorial assistance support may have been funded by Communications companies in part; No speaker’s bureau; Mentored several non-US citizens on research projects with potential funding (in part) from non-US sources/Foreign Components; The institution (Dana-Farber Cancer Institute) may have received additional independent funding of drug companies or/and royalties potentially involved in research around the subject matter.

Contributor Information

Soumitro Pal, Email: soumitro.pal@childrens.harvard.edu.

Akash Sabarwal, Email: akash.sabarwal@childrens.harvard.edu.

References

  • 1.Murakami, Y. et al. AXL/CDCP1/SRC axis confers acquired resistance to osimertinib in lung cancer. Sci. Rep.12, 8983 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 2.Patel, S. A. et al. Molecular mechanisms and future implications of VEGF/VEGFR in cancer therapy. Clin. Cancer Res.29, 30–39 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Kumar, R. et al. Recent developments in receptor tyrosine kinase inhibitors: a promising mainstay in targeted cancer therapy. Med. Drug Discov.23, 100195 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Rawat, L. et al. A novel combination therapy with Cabozantinib and Honokiol effectively inhibits c-Met-Nrf2-induced renal tumor growth through increased oxidative stress. Redox Biol.68, 102945 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Lu, Q. et al. Tyro-3 family receptors are essential regulators of mammalian spermatogenesis. Nature398, 723–728 (1999). [DOI] [PubMed] [Google Scholar]
  • 6.Scott, R. S. et al. Phagocytosis and clearance of apoptotic cells is mediated by MER. Nature411, 207–211 (2001). [DOI] [PubMed] [Google Scholar]
  • 7.Angelillo-Scherrer, A. et al. Role of Gas6 in erythropoiesis and anemia in mice. J. Clin. Investig.118, 583–596 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Angelillo-Scherrer, A. et al. Role of Gas6 receptors in platelet signaling during thrombus stabilization and implications for antithrombotic therapy. J. Clin. Investig.115, 237–246 (2005). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Moller-Tank, S. & Maury, W. Phosphatidylserine receptors: enhancers of enveloped virus entry and infection. Virology468-470, 565–580 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Morizono, K. et al. The soluble serum protein Gas6 bridges virion envelope phosphatidylserine to the TAM receptor tyrosine kinase Axl to mediate viral entry. Cell Host Microbe9, 286–298 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Miao, Y. R., Rankin, E. B. & Giaccia, A. J. Therapeutic targeting of the functionally elusive TAM receptor family. Nat. Rev. Drug Discov.23, 201–217 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Challier, C., Uphoff, C. C., Janssen, J. W. & Drexler, H. G. Differential expression of the ufo/axl oncogene in human leukemia-lymphoma cell lines. Leukemia10, 781–787 (1996). [PubMed] [Google Scholar]
  • 13.Graham, D. K. et al. Ectopic expression of the proto-oncogene Mer in pediatric T-cell acute lymphoblastic leukemia. Clin. Cancer Res.12, 2662–2669 (2006). [DOI] [PubMed] [Google Scholar]
  • 14.DeRyckere, D., Huelse, J. M., Earp, H. S. & Graham, D. K. TAM family kinases as therapeutic targets at the interface of cancer and immunity. Nat. Rev. Clin. Oncol.20, 755–779 (2023). [DOI] [PubMed] [Google Scholar]
  • 15.Ling, L., Templeton, D. & Kung, H. J. Identification of the major autophosphorylation sites of Nyk/Mer, an NCAM-related receptor tyrosine kinase. J. Biol. Chem.271, 18355–18362 (1996). [DOI] [PubMed] [Google Scholar]
  • 16.Zucca, L. E. et al. Expression of tyrosine kinase receptor AXL is associated with worse outcome of metastatic renal cell carcinomas treated with sunitinib. Urol. Oncol.36, 11.e13–11.e21 (2018). [DOI] [PubMed] [Google Scholar]
  • 17.Sabarwal, A. et al. A Combination therapy using an mTOR inhibitor and Honokiol effectively induces autophagy through the modulation of AXL and Rubicon in renal cancer cells and restricts renal tumor growth following organ transplantation. Carcinogenesis43, 360–370 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Sang, Y. B. et al. The development of AXL inhibitors in lung cancer: recent progress and challenges. Front. Oncol.12, 811247 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Ozyurt, R. & Ozpolat, B. Therapeutic landscape of AXL receptor kinase in triple-negative breast cancer. Mol. Cancer Ther.22, 818–832 (2023). [DOI] [PubMed] [Google Scholar]
  • 20.Paccez, J. D. et al. The receptor tyrosine kinase Axl is an essential regulator of prostate cancer proliferation and tumor growth and represents a new therapeutic target. Oncogene32, 689–698 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Ashouri, K. et al. Clinical and molecular characterization of AXL in colorectal cancer: CALGB (Alliance)/SWOG 80405 and real-world data. J. Clin. Oncol.42, 3145–3145 (2024). [Google Scholar]
  • 22.De Rosa, L., Di Stasi, R., Fusco, V. & D’Andrea, L. D. AXL receptor as an emerging molecular target in colorectal cancer. Drug Discov. Today29, 104005 (2024). [DOI] [PubMed] [Google Scholar]
  • 23.Hsieh, M. S., Yang, P. W., Wong, L. F. & Lee, J. M. The AXL receptor tyrosine kinase is associated with adverse prognosis and distant metastasis in esophageal squamous cell carcinoma. Oncotarget7, 36956–36970 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Brand, T. M. et al. AXL is a logical molecular target in head and neck squamous cell carcinoma. Clin. Cancer Res.21, 2601–2612 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Busch, C. J. et al. Tissue microarray analyses suggest Axl as a predictive biomarker in HPV-negative head and neck cancer. Cancers14, 1829 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Tang, Y., Zang, H., Wen, Q. & Fan, S. AXL in cancer: a modulator of drug resistance and therapeutic target. J. Exp. Clin. Cancer Res.42, 148 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Li, Y. et al. Axl as a potential therapeutic target in cancer: role of Axl in tumor growth, metastasis and angiogenesis. Oncogene28, 3442–3455 (2009). [DOI] [PubMed] [Google Scholar]
  • 28.Gay, C. M., Balaji, K. & Byers, L. A. Giving AXL the axe: targeting AXL in human malignancy. Br. J. Cancer116, 415–423 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Engelsen, A. S. T. et al. AXL is a driver of stemness in normal mammary gland and breast cancer. iScience23, 101649 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Chai, Z. T. et al. AXL overexpression in tumor-derived endothelial cells promotes vessel metastasis in patients with hepatocellular carcinoma. Front. Oncol.11, 650963 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Iida, M. et al. AXL regulates neuregulin1 expression leading to cetuximab resistance in head and neck cancer. BMC Cancer22, 447 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Noronha, A. et al. AXL and error-prone DNA replication confer drug resistance and offer strategies to treat EGFR-mutant lung cancer. Cancer Discov.12, 2666–2683 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Song, X. et al. Overexpression of receptor tyrosine kinase Axl promotes tumor cell invasion and survival in pancreatic ductal adenocarcinoma. Cancer117, 734–743 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Wang, Y. et al. Prognostic value and immunological role of AXL gene in clear cell renal cell carcinoma associated with identifying LncRNA/RBP/AXL mRNA networks. Cancer Cell Int.21, 625 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Khamko, R. et al. Combined OPCML and AXL expression as a prognostic marker and OPCML enhances AXL inhibitor in cholangiocarcinoma. Vivo36, 1168–1177 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Yamagata, M., Sanes, J. R. & Weiner, J. A. Synaptic adhesion molecules. Curr. Opin. Cell Biol.15, 621–632 (2003). [DOI] [PubMed] [Google Scholar]
  • 37.Linger, R. M., Keating, A. K., Earp, H. S. & Graham, D. K. TAM receptor tyrosine kinases: biologic functions, signaling, and potential therapeutic targeting in human cancer. Adv. Cancer Res.100, 35–83 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Kimani, S. G. et al. Small molecule inhibitors block Gas6-inducible TAM activation and tumorigenicity. Sci. Rep.7, 43908 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Wium, M., Ajayi-Smith, A. F., Paccez, J. D. & Zerbini, L. F. The role of the receptor tyrosine kinase Axl in carcinogenesis and development of therapeutic resistance: an overview of molecular mechanisms and future applications. Cancers13, 1521 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.O’Bryan, J. P. et al. axl, a transforming gene isolated from primary human myeloid leukemia cells, encodes a novel receptor tyrosine kinase. Mol. Cell Biol.11, 5016–5031 (1991). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Verma, A. et al. Targeting Axl and Mer kinases in cancer. Mol. Cancer Ther.10, 1763–1773 (2011). [DOI] [PubMed] [Google Scholar]
  • 42.Hafizi, S. & Dahlbäck, B. Signalling and functional diversity within the Axl subfamily of receptor tyrosine kinases. Cytokine Growth Factor Rev.17, 295–304 (2006). [DOI] [PubMed] [Google Scholar]
  • 43.Sadahiro, H. et al. Activation of the receptor tyrosine kinase AXL regulates the immune microenvironment in glioblastoma. Cancer Res.78, 3002–3013 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Caberoy, N. B., Zhou, Y. & Li, W. Tubby and tubby-like protein 1 are new MerTK ligands for phagocytosis. EMBO J.29, 3898–3910 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Auyez, A., Sayan, A. E., Kriajevska, M. & Tulchinsky, E. AXL receptor in cancer metastasis and drug resistance: when normal functions go askew. Cancers13, 4864 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Axelrod, H. & Pienta, K. J. Axl as a mediator of cellular growth and survival. Oncotarget5, 8818–8852, (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Korshunov, V. A. Axl-dependent signalling: a clinical update. Clin. Sci.122, 361–368 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Schmidt, T., Ben-Batalla, I., Schultze, A. & Loges, S. Macrophage-tumor crosstalk: role of TAMR tyrosine kinase receptors and of their ligands. Cell Mol. Life Sci.69, 1391–1414, (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Sasaki, T. et al. Structural basis for Gas6-Axl signalling. EMBO J.25, 80–87 (2006). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Brandao, L., Migdall-Wilson, J., Eisenman, K. & Graham, D. K. TAM receptors in leukemia: expression, signaling, and therapeutic implications. Crit. Rev. Oncog.16, 47–63 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Vouri, M. et al. Axl-EGFR receptor tyrosine kinase hetero-interaction provides EGFR with access to pro-invasive signalling in cancer cells. Oncogenesis5, e266 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Bellosta, P., Costa, M., Lin, D. A. & Basilico, C. The receptor tyrosine kinase ARK mediates cell aggregation by homophilic binding. Mol. Cell Biol.15, 614–625, (1995). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Huang, J. S. et al. Oxidative stress enhances Axl-mediated cell migration through an Akt1/Rac1-dependent mechanism. Free Radic. Biol. Med.65, 1246–1256 (2013). [DOI] [PubMed] [Google Scholar]
  • 54.Mudduluru, G. & Allgayer, H. The human receptor tyrosine kinase Axl gene-promoter characterization and regulation of constitutive expression by Sp1, Sp3 and CpG methylation. Biosci. Rep.28, 161–176, (2008). [DOI] [PubMed] [Google Scholar]
  • 55.Mudduluru, G., Leupold, J. H., Stroebel, P. & Allgayer, H. PMA up-regulates the transcription of Axl by AP-1 transcription factor binding to TRE sequences via the MAPK cascade in leukaemia cells. Biol. Cell103, 21–33 (2010). [DOI] [PubMed] [Google Scholar]
  • 56.Linger, R. M. et al. Mer or Axl receptor tyrosine kinase inhibition promotes apoptosis, blocks growth and enhances chemosensitivity of human non-small cell lung cancer. Oncogene32, 3420–3431 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Rankin, E. B. et al. Direct regulation of GAS6/AXL signaling by HIF promotes renal metastasis through SRC and MET. Proc. Natl. Acad. Sci. USA111, 13373–13378 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Lamhamedi-Cherradi, S. E. et al. Transcriptional activators YAP/TAZ and AXL orchestrate dedifferentiation, cell fate, and metastasis in human osteosarcoma. Cancer Gene Ther.28, 1325–1338 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Badarni, M. et al. Repression of AXL expression by AP-1/JNK blockage overcomes resistance to PI3Ka therapy. JCI Insight5, e125341 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Safaric Tepes, P. et al. An epigenetic switch regulates the ontogeny of AXL-positive/EGFR-TKi-resistant cells by modulating miR-335 expression. eLife10, e66109 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Li, J. et al. Axl glycosylation mediates tumor cell proliferation, invasion and lymphatic metastasis in murine hepatocellular carcinoma. World J. Gastroenterol.18, 5369–5376 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Lauter, M., Weber, A. & Torka, R. Targeting of the AXL receptor tyrosine kinase by small molecule inhibitor leads to AXL cell surface accumulation by impairing the ubiquitin-dependent receptor degradation. Cell Commun. Signal17, 59 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Wang, J. et al. Leukemogenic chromatin alterations promote AML leukemia stem cells via a KDM4C-ALKBH5-AXL signaling axis. Cell Stem Cell27, 81–97.e88 (2020). [DOI] [PubMed] [Google Scholar]
  • 64.De Martino, M., Esposito, F. & Pallante, P. Long non-coding RNAs regulating multiple proliferative pathways in cancer cell. Transl. Cancer Res.10, 3140–3157 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 65.Jiang, N. et al. lncRNA DANCR promotes tumor progression and cancer stemness features in osteosarcoma by upregulating AXL via miR-33a-5p inhibition. Cancer Lett.405, 46–55 (2017). [DOI] [PubMed] [Google Scholar]
  • 66.Yang, L. G. et al. LncRNA XIST modulates HIF-1A/AXL signaling pathway by inhibiting miR-93-5p in colorectal cancer. Mol. Genet Genom. Med.8, e1112 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Seo, J. S. et al. The transcriptional landscape and mutational profile of lung adenocarcinoma. Genome Res.22, 2109–2119 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Rothlin, C. V. et al. TAM receptors are pleiotropic inhibitors of the innate immune response. Cell131, 1124–1136 (2007). [DOI] [PubMed] [Google Scholar]
  • 69.Stenhoff, J., Dahlback, B. & Hafizi, S. Vitamin K-dependent Gas6 activates ERK kinase and stimulates growth of cardiac fibroblasts. Biochem Biophys. Res. Commun.319, 871–878, (2004). [DOI] [PubMed] [Google Scholar]
  • 70.Zhu, C., Wei, Y. & Wei, X. AXL receptor tyrosine kinase as a promising anti-cancer approach: functions, molecular mechanisms and clinical applications. Mol. Cancer18, 153 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Sinha, S. et al. Targeted Axl inhibition primes chronic lymphocytic leukemia B cells to apoptosis and shows synergistic/additive effects in combination with BTK inhibitors. Clin. Cancer Res.21, 2115–2126 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 72.Paccez, J. D., Vogelsang, M., Parker, M. I. & Zerbini, L. F. The receptor tyrosine kinase Axl in cancer: biological functions and therapeutic implications. Int. J. Cancer134, 1024–1033 (2014). [DOI] [PubMed] [Google Scholar]
  • 73.Yu, H. et al. Axl receptor tyrosine kinase is a potential therapeutic target in renal cell carcinoma. Br. J. Cancer113, 616–625 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Koorstra, J. B. et al. The Axl receptor tyrosine kinase confers an adverse prognostic influence in pancreatic cancer and represents a new therapeutic target. Cancer Biol. Ther.8, 618–626 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Rankin, E. B. et al. AXL is an essential factor and therapeutic target for metastatic ovarian cancer. Cancer Res.70, 7570–7579 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Xu, M. Z. et al. AXL receptor kinase is a mediator of YAP-dependent oncogenic functions in hepatocellular carcinoma. Oncogene30, 1229–1240 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Asiedu, M. K. et al. AXL induces epithelial-to-mesenchymal transition and regulates the function of breast cancer stem cells. Oncogene33, 1316–1324 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 78.Maacha, S. et al. AXL mediates esophageal adenocarcinoma cell invasion through regulation of extracellular acidification and lysosome trafficking. Neoplasia20, 1008–1022 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Wilson, C. et al. AXL inhibition sensitizes mesenchymal cancer cells to antimitotic drugs. Cancer Res.74, 5878–5890 (2014). [DOI] [PubMed] [Google Scholar]
  • 80.Storci, G. et al. The basal-like breast carcinoma phenotype is regulated by SLUG gene expression. J. Pathol.214, 25–37 (2008). [DOI] [PubMed] [Google Scholar]
  • 81.Gjerdrum, C. et al. Axl is an essential epithelial-to-mesenchymal transition-induced regulator of breast cancer metastasis and patient survival. Proc. Natl. Acad. Sci. USA107, 1124–1129 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Thiery, J. P., Acloque, H., Huang, R. Y. & Nieto, M. A. Epithelial-mesenchymal transitions in development and disease. Cell139, 871–890, (2009). [DOI] [PubMed] [Google Scholar]
  • 83.Nieto, M. A., Huang, R. Y., Jackson, R. A. & Thiery, J. P. Emt: 2016. Cell166, 21–45 (2016). [DOI] [PubMed] [Google Scholar]
  • 84.Li, Y. et al. Axl mediates tumor invasion and chemosensitivity through PI3K/Akt signaling pathway and is transcriptionally regulated by slug in breast carcinoma. IUBMB Life66, 507–518 (2014). [DOI] [PubMed] [Google Scholar]
  • 85.Goyette, M. A. et al. The receptor tyrosine kinase AXL is required at multiple steps of the metastatic cascade during HER2-positive breast cancer progression. Cell Rep.23, 1476–1490 (2018). [DOI] [PubMed] [Google Scholar]
  • 86.Genna, A. et al. EMT-associated heterogeneity in circulating tumor cells: sticky friends on the road to metastasis. Cancers12, 1632 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.de Miguel-Pérez, D. et al. Post-surgery circulating tumor cells and AXL overexpression as new poor prognostic biomarkers in resected lung adenocarcinoma. Cancers11, 1750 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Grover, P. K. et al. Circulating tumour cells: the evolving concept and the inadequacy of their enrichment by EpCAM-based methodology for basic and clinical cancer research. Ann. Oncol.25, 1506–1516 (2014). [DOI] [PubMed] [Google Scholar]
  • 89.Allard, W. J. et al. Tumor cells circulate in the peripheral blood of all major carcinomas but not in healthy subjects or patients with nonmalignant diseases. Clin. Cancer Res.10, 6897–6904 (2004). [DOI] [PubMed] [Google Scholar]
  • 90.Cristofanilli, M. et al. Circulating tumor cells, disease progression, and survival in metastatic breast cancer. New Engl. J. Med.351, 781–791 (2004). [DOI] [PubMed] [Google Scholar]
  • 91.Andree, K. C., van Dalum, G. & Terstappen, L. W. Challenges in circulating tumor cell detection by the CellSearch system. Mol. Oncol.10, 395–407 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Stoecklein, N. H., Fischer, J. C., Niederacher, D. & Terstappen, L. W. Challenges for CTC-based liquid biopsies: low CTC frequency and diagnostic leukapheresis as a potential solution. Expert Rev. Mol. Diagn.16, 147–164 (2016). [DOI] [PubMed] [Google Scholar]
  • 93.Hanssen, A. et al. Characterization of different CTC subpopulations in non-small cell lung cancer. Sci. Rep.6, 28010 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Ikeda, M. et al. Detection of AXL expression in circulating tumor cells of lung cancer patients using an automated microcavity array system. Cancer Med.9, 2122–2133 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Bardol, T. et al. First evidence of AXL expression on circulating tumor cells in metastatic breast cancer patients: a proof-of-concept study. Cancer Med.13, e6843 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 96.Rothlin, C. V. & Lemke, G. TAM receptor signaling and autoimmune disease. Curr. Opin. Immunol.22, 740–746, (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.Zagorska, A. et al. Diversification of TAM receptor tyrosine kinase function. Nat. Immunol.15, 920–928 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Ruan, G. X. & Kazlauskas, A. Lactate engages receptor tyrosine kinases Axl, Tie2, and vascular endothelial growth factor receptor 2 to activate phosphoinositide 3-kinase/Akt and promote angiogenesis. J. Biol. Chem.288, 21161–21172 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 99.Graham, D. K., DeRyckere, D., Davies, K. D. & Earp, H. S. The TAM family: phosphatidylserine sensing receptor tyrosine kinases gone awry in cancer. Nat. Rev. Cancer14, 769–785 (2014). [DOI] [PubMed] [Google Scholar]
  • 100.Holland, S. J. et al. Multiple roles for the receptor tyrosine kinase axl in tumor formation. Cancer Res.65, 9294–9303 (2005). [DOI] [PubMed] [Google Scholar]
  • 101.Ruan, G. X. & Kazlauskas, A. Axl is essential for VEGF-A-dependent activation of PI3K/Akt. EMBO J.31, 1692–1703 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Xiao, Y. et al. S100A10 is a critical mediator of GAS6/AXL-induced angiogenesis in renal cell carcinoma. Cancer Res.79, 5758–5768 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 103.Ott, M. et al. Promotion of glioblastoma cell motility by enhancer of zeste homolog 2 (EZH2) is mediated by AXL receptor kinase. PLoS ONE7, e47663 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 104.Rankin, E. B. & Giaccia, A. J. The receptor tyrosine kinase AXL in cancer progression. Cancers8, 103 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Cichon, M. A. et al. The receptor tyrosine kinase Axl regulates cell-cell adhesion and stemness in cutaneous squamous cell carcinoma. Oncogene33, 4185–4192 (2014). [DOI] [PubMed] [Google Scholar]
  • 106.Khera, L., Vinik, Y., Maina, F. & Lev, S. The AXL-PYK2-PKCα axis as a nexus of stemness circuits in TNBC. Life Sci. Alliance4, e202000985 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Jin, Y. et al. Gas6/AXL signaling regulates self-renewal of chronic myelogenous leukemia stem cells by stabilizing beta-catenin. Clin. Cancer Res.23, 2842–2855 (2017). [DOI] [PubMed] [Google Scholar]
  • 108.Hong, C. C. et al. Receptor tyrosine kinase AXL is induced by chemotherapy drugs and overexpression of AXL confers drug resistance in acute myeloid leukemia. Cancer Lett.268, 314–324 (2008). [DOI] [PubMed] [Google Scholar]
  • 109.Aguilera, T. A. et al. Reprogramming the immunological microenvironment through radiation and targeting Axl. Nat. Commun.7, 13898 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Guo, Z., Li, Y., Zhang, D. & Ma, J. Axl inhibition induces the antitumor immune response which can be further potentiated by PD-1 blockade in the mouse cancer models. Oncotarget8, 89761–89774 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Kasikara, C. et al. Phosphatidylserine sensing by TAM receptors regulates AKT-dependent chemoresistance and PD-L1 expression. Mol. Cancer Res.15, 753–764 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Skinner, H. D. et al. Integrative analysis identifies a novel AXL-PI3 kinase-PD-L1 signaling axis associated with radiation resistance in head and neck cancer. Clin. Cancer Res.23, 2713–2722 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 113.Terry, S. et al. Association of AXL and PD-L1 expression with clinical outcomes in patients with advanced renal cell carcinoma treated with PD-1 blockade. Clin. Cancer Res.27, 6749–6760 (2021). [DOI] [PubMed] [Google Scholar]
  • 114.Ye, X. et al. An anti-Axl monoclonal antibody attenuates xenograft tumor growth and enhances the effect of multiple anticancer therapies. Oncogene29, 5254–5264 (2010). [DOI] [PubMed] [Google Scholar]
  • 115.Leconet, W. et al. Therapeutic activity of anti-AXL antibody against triple-negative breast cancer patient-derived xenografts and metastasis. Clin. Cancer Res.23, 2806–2816 (2017). [DOI] [PubMed] [Google Scholar]
  • 116.Bejarano, L., Jordao, M. J. C. & Joyce, J. A. Therapeutic targeting of the tumor microenvironment. Cancer Discov.11, 933–959 (2021). [DOI] [PubMed] [Google Scholar]
  • 117.Engelsen, A. S. T. et al. Dissecting the role of AXL in cancer immune escape and resistance to immune checkpoint inhibition. Front. Immunol.13, 869676 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Myers, K. V., Amend, S. R. & Pienta, K. J. Targeting Tyro3, Axl and MerTK (TAM receptors): implications for macrophages in the tumor microenvironment. Mol. Cancer18, 94 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Meertens, L. et al. Axl mediates ZIKA virus entry in human glial cells and modulates innate immune responses. Cell Rep.18, 324–333 (2017). [DOI] [PubMed] [Google Scholar]
  • 120.Shimojima, M., Ikeda, Y. & Kawaoka, Y. The mechanism of Axl-mediated Ebola virus infection. J. Infect. Dis.196, S259–S263 (2007). [DOI] [PubMed] [Google Scholar]
  • 121.Zhang, Q. et al. Molecular mechanism of interaction between SARS-CoV-2 and host cells and interventional therapy. Signal Transduct. Target Ther.6, 233 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Wang, S. et al. AXL is a candidate receptor for SARS-CoV-2 that promotes infection of pulmonary and bronchial epithelial cells. Cell Res.31, 126–140 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Zwernik, S. D. et al. AXL receptor is required for Zika virus strain MR-766 infection in human glioblastoma cell lines. Mol. Ther. Oncolytics23, 447–457 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Ning, K. et al. Identification of AXL as a co-receptor for human parvovirus B19 infection of human erythroid progenitors. Sci. Adv.9, eade0869 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Tanaka, M. & Siemann, D. W. Gas6/Axl signaling pathway in the tumor immune microenvironment. Cancers12, 1850 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 126.Lee, E. H. et al. Blockade of Axl signaling ameliorates HPV16E6-mediated tumorigenecity of cervical cancer. Sci. Rep.7, 5759 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Zhang, W. N. et al. Comprehensive analysis of the novel omicron receptor AXL in cancers. Comput Struct. Biotechnol. J.20, 3304–3312 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 128.Bosurgi, L. et al. Paradoxical role of the proto-oncogene Axl and Mer receptor tyrosine kinases in colon cancer. Proc. Natl. Acad. Sci. USA110, 13091–13096 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.Kurowska-Stolarska, M. et al. MicroRNA-34a dependent regulation of AXL controls the activation of dendritic cells in inflammatory arthritis. Nat. Commun.8, 15877 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Novoyatleva, T. et al. Deficiency of Axl aggravates pulmonary arterial hypertension via BMPR2. Commun. Biol.4, 1002 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 131.Onken, J. et al. Inhibiting receptor tyrosine kinase AXL with small molecule inhibitor BMS-777607 reduces glioblastoma growth, migration, and invasion in vitro and in vivo. Oncotarget7, 9876–9889 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 132.Tian, M. et al. Inhibition of AXL enhances chemosensitivity of human ovarian cancer cells to cisplatin via decreasing glycolysis. Acta Pharm. Sin.42, 1180–1189 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 133.Elkabets, M. et al. AXL mediates resistance to PI3Kα inhibition by activating the EGFR/PKC/mTOR axis in head and neck and esophageal squamous cell carcinomas. Cancer Cell27, 533–546 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Balaji, K. et al. AXL inhibition suppresses the DNA damage response and sensitizes cells to PARP inhibition in multiple cancers. Mol. Cancer Res.15, 45–58 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 135.Wang, F. et al. Blockade of AXL activation overcomes acquired resistance to EGFR tyrosine kinase inhibition in non-small cell lung cancer. Transl. Cancer Res.8, 2425–2438 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 136.Toboni, M. D. et al. Inhibition of AXL and VEGF-A has improved therapeutic efficacy in uterine serous cancer. Cancers13, 5877 (2021). [DOI] [PMC free article] [PubMed]
  • 137.Adam-Artigues, A. et al. Targeting HER2-AXL heterodimerization to overcome resistance to HER2 blockade in breast cancer. Sci. Adv.8, eabk2746 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 138.Liu, Z. et al. Recent discovery and development of AXL inhibitors as antitumor agents. Eur. J. Med. Chem.272, 116475 (2024). [DOI] [PubMed] [Google Scholar]
  • 139.Shen, Y. et al. Axl inhibitors as novel cancer therapeutic agents. Life Sci.198, 99–111 (2018). [DOI] [PubMed] [Google Scholar]
  • 140.Hart, C. D. & De Boer, R. H. Profile of cabozantinib and its potential in the treatment of advanced medullary thyroid cancer. Onco Targets Ther.6, 1–7 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 141.Lee, H. J. et al. Gas6/Axl pathway promotes tumor invasion through the transcriptional activation of Slug in hepatocellular carcinoma. Carcinogenesis35, 769–775 (2014). [DOI] [PubMed] [Google Scholar]
  • 142.Davis, M. I. et al. Comprehensive analysis of kinase inhibitor selectivity. Nat. Biotechnol.29, 1046–1051 (2011). [DOI] [PubMed] [Google Scholar]
  • 143.Zhao, Z. et al. Exploration of type II binding mode: a privileged approach for kinase inhibitor focused drug discovery? ACS Chem. Biol.9, 1230–1241 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 144.Rodon, J. et al. First-in-human phase I study of oral S49076, a unique MET/AXL/FGFR inhibitor, in advanced solid tumours. Eur. J. Cancer81, 142–150 (2017). [DOI] [PubMed] [Google Scholar]
  • 145.Felip, E. et al. A phase II study of bemcentinib (BGB324), a first-in-class highly selective AXL inhibitor, with pembrolizumab in pts with advanced NSCLC: OS for stage I and preliminary stage II efficacy. J. Clin. Oncol.37, 9098–9098 (2019). [Google Scholar]
  • 146.Chen, T. J. et al. AXL targeting by a specific small molecule or monoclonal antibody inhibits renal cell carcinoma progression in an orthotopic mice model. Physiol. Rep.9, e15140 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 147.Alvarado, D. et al. 550 An Axl-targeting monoclonal antibody that inhibits Axl activity and potently stimulates the innate immune response. J. Immunother. Cancer8, A334–A334 (2020). [Google Scholar]
  • 148.Pei, J. P. et al. AXL antibody and AXL-ADC mediate antitumor efficacy via targeting AXL in tumor-intrinsic epithelial-mesenchymal transition and tumor-associated M2-like macrophage. Acta Pharm. Sin.44, 1290–1303 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 149.Van Renterghem, B. et al. Enapotamab vedotin, an AXL-specific antibody-drug conjugate, demonstrates antitumor efficacy in patient-derived xenograft models of soft tissue sarcoma. Int. J. Mol. Sci.23, 7493 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 150.Fuentes-Antras, J., Genta, S., Vijenthira, A. & Siu, L. L. Antibody-drug conjugates: in search of partners of choice. Trends Cancer9, 339–354 (2023). [DOI] [PubMed] [Google Scholar]
  • 151.Liu, Y. et al. Novel AXL-targeted agents overcome FLT3 inhibitor resistance in FLT3-ITD(+) acute myeloid leukemia cells. Oncol. Lett.21, 397 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 152.Vandewalle, N. et al. AXL-specific single domain antibodies show diagnostic potential and anti-tumor activity in acute myeloid leukemia. Theranostics14, 2656–2674 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 153.Cho, J. H. et al. Engineering Axl specific CAR and SynNotch receptor for cancer therapy. Sci. Rep.8, 3846 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 154.Wei, J. et al. A novel AXL chimeric antigen receptor endows T cells with anti-tumor effects against triple negative breast cancers. Cell Immunol.331, 49–58 (2018). [DOI] [PubMed] [Google Scholar]
  • 155.Włodarczyk, M. & Pyrzynska, B. CAR-NK as a rapidly developed and efficient immunotherapeutic strategy against cancer. Cancers15, 117 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 156.Zhao, Z., Li, Y., Liu, W. & Li, X. Engineered IL-7 receptor enhances the therapeutic effect of AXL-CAR-T cells on triple-negative breast cancer. Biomed. Res. Int.2020, 4795171 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 157.Cao, B. et al. Remodelling of tumour microenvironment by microwave ablation potentiates immunotherapy of AXL-specific CAR T cells against non-small cell lung cancer. Nat. Commun.13, 6203 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 158.Fan, J. et al. GAS6-based CAR-T cells exhibit potent antitumor activity against pancreatic cancer. J. Hematol. Oncol.16, 77 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 159.Sakemura, R. L. et al. AXL inhibition improves the antitumor activity of chimeric antigen receptor T cells. Cancer Immunol. Res.11, 1222–1236 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 160.Wu, Z. et al. Abstract LB342: potent anti-tumor activity demonstrated by anti-AXL CAR T cell therapy and AXL-ADC in lung cancer and breast cancer models. Cancer Res.84, LB342 (2024). [Google Scholar]
  • 161.Li, J. et al. NK-92MI cells engineered with anti-claudin-6 chimeric antigen receptors in immunotherapy for ovarian cancer. Int. J. Biol. Sci.20, 1578–1601 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 162.Macleod, K. et al. Altered ErbB receptor signaling and gene expression in cisplatin-resistant ovarian cancer. Cancer Res.65, 6789–6800 (2005). [DOI] [PubMed] [Google Scholar]
  • 163.Brand, T. M. et al. AXL mediates resistance to cetuximab therapy. Cancer Res.74, 5152–5164 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar] [Research Misconduct Found]
  • 164.Cardone, C. et al. AXL has a prognostic role in metastatic colorectal cancer (mCRC) and is a predictive biomarker of lack of efficacy of chemotherapy (CT) + cetuximab in RAS wild type (WT) patients (pts). Ann. Oncol.29, viii30 (2018). [Google Scholar]
  • 165.Quinn, J. M. et al. Therapeutic inhibition of the receptor tyrosine kinase AXL improves sensitivity to platinum and taxane in ovarian cancer. Mol. Cancer Ther.18, 389–398 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 166.Scherschinski, L. et al. Regulation of the receptor tyrosine kinase AXL in response to therapy and its role in therapy resistance in glioblastoma. Int. J. Mol. Sci.23, 982 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 167.AACR Project GENIE. Powering precision medicine through an International Consortium. Cancer Discov.7, 818–831 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 168.Namba, K. et al. Activation of AXL as a preclinical acquired resistance mechanism against osimertinib treatment in EGFR-mutant non–small cell lung cancer cells. Mol. Cancer Res.17, 499–507 (2019). [DOI] [PubMed] [Google Scholar]
  • 169.Li, L. Y. et al. DNA repair pathways in cancer therapy and resistance. Front. Pharm.11, 629266 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 170.Hong, J. et al. ABL regulation by AXL promotes cisplatin resistance in esophageal cancer. Cancer Res.73, 331–340 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 171.Sen, T. et al. Targeting AXL and mTOR pathway overcomes primary and acquired resistance to WEE1 inhibition in small-cell lung cancer. Clin. Cancer Res.23, 6239–6253 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 172.Ramkumar, K. et al. AXL inhibition induces DNA damage and replication stress in non-small cell lung cancer cells and promotes sensitivity to ATR inhibitors. Mol. Cancer Res.19, 485–497 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 173.Yeo, X. H. et al. The effect of inhibition of receptor tyrosine kinase AXL on DNA damage response in ovarian cancer. Commun. Biol.6, 660 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 174.Mullen, M. M. et al. GAS6/AXL inhibition enhances ovarian cancer sensitivity to chemotherapy and PARP inhibition through increased DNA damage and enhanced replication stress. Mol. Cancer Res.20, 265–279 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 175.Marquez-Palencia, M. et al. AXL/WRNIP1 mediates replication stress response and promotes therapy resistance and metachronous metastasis in HER2+ breast cancer. Cancer Res.84, 675–687 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 176.Xu, Z., Zhang, Y., Dai, H. & Han, B. Epithelial-mesenchymal transition-mediated tumor therapeutic resistance. Molecules27, 4750 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 177.Shao, H., Teramae, D. & Wells, A. Axl contributes to efficient migration and invasion of melanoma cells. PLoS ONE18, e0283749 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 178.Giles, K. M. et al. Axl mediates acquired resistance of head and neck cancer cells to the epidermal growth factor receptor inhibitor erlotinib. Mol. Cancer Ther.12, 2541–2558 (2013). [DOI] [PubMed] [Google Scholar]
  • 179.Wang, C. et al. Gas6/Axl axis contributes to chemoresistance and metastasis in breast cancer through Akt/GSK-3β/β-catenin signaling. Theranostics6, 1205–1219 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 180.Hong, J., Maacha, S. & Belkhiri, A. Transcriptional upregulation of c-MYC by AXL confers epirubicin resistance in esophageal adenocarcinoma. Mol. Oncol.12, 2191–2208 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 181.Reichl, P. et al. Axl activates autocrine transforming growth factor-beta signaling in hepatocellular carcinoma. Hepatology61, 930–941 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 182.Pinato, D. J. et al. Integrated analysis of multiple receptor tyrosine kinases identifies Axl as a therapeutic target and mediator of resistance to sorafenib in hepatocellular carcinoma. Br. J. Cancer120, 512–521 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 183.Solanes-Casado, S. et al. Overcoming PLK1 inhibitor resistance by targeting mevalonate pathway to impair AXL-TWIST axis in colorectal cancer. Biomed. Pharmacother.144, 112347 (2021). [DOI] [PubMed] [Google Scholar]
  • 184.Antony, J. et al. The GAS6-AXL signaling network is a mesenchymal (Mes) molecular subtype-specific therapeutic target for ovarian cancer. Sci. Signal9, ra97 (2016). [DOI] [PubMed] [Google Scholar]
  • 185.Choi, J. Y. et al. TGFbeta promotes YAP-dependent AXL induction in mesenchymal-type lung cancer cells. Mol. Oncol.15, 679–696 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 186.Suresh, D. et al. Silencing AXL by covalent siRNA-gelatin-antibody nanoconjugate inactivates mTOR/EMT pathway and stimulates p53 for TKI sensitization in NSCLC. Nanomedicine20, 102007 (2019). [DOI] [PubMed] [Google Scholar]
  • 187.Du, W. et al. The miR‑625‑3p/AXL axis induces non‑T790M acquired resistance to EGFR‑TKI via activation of the TGF‑beta/Smad pathway and EMT in EGFR‑mutant non‑small cell lung cancer. Oncol. Rep.44, 185–195 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 188.Chen, C. et al. Activated T cell-derived exosomes for targeted delivery of AXL-siRNA loaded paclitaxel-poly-L-lysine prodrug to overcome drug resistance in triple-negative breast cancer. Chem. Eng. J.468, 143454 (2023). [Google Scholar]
  • 189.Schoumacher, M. & Burbridge, M. Key roles of AXL and MER receptor tyrosine kinases in resistance to multiple anticancer therapies. Curr. Oncol. Rep.19, 19 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 190.Jimbo, T. et al. DS-1205b, a novel selective inhibitor of AXL kinase, blocks resistance to EGFR-tyrosine kinase inhibitors in a non-small cell lung cancer xenograft model. Oncotarget10, 5152–5167 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 191.Tian, X., Gu, T., Lee, M. H. & Dong, Z. Challenge and countermeasures for EGFR targeted therapy in non-small cell lung cancer. Biochim Biophys. Acta Rev. Cancer1877, 188645 (2022). [DOI] [PubMed] [Google Scholar]
  • 192.Saab, S. et al. The potential role of YAP in Axl-mediated resistance to EGFR tyrosine kinase inhibitors. Am. J. Cancer Res.9, 2719–2729 (2019). [PMC free article] [PubMed] [Google Scholar]
  • 193.Liu, Y. N. et al. Acquired resistance to EGFR tyrosine kinase inhibitors is mediated by the reactivation of STC2/JUN/AXL signaling in lung cancer. Int. J. Cancer145, 1609–1624 (2019). [DOI] [PubMed] [Google Scholar]
  • 194.Lotsberg, M. L. et al. AXL targeting abrogates autophagic flux and induces immunogenic cell death in drug-resistant cancer cells. J. Thorac. Oncol.15, 973–999 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 195.Miller, M. A. et al. Reduced proteolytic shedding of receptor tyrosine kinases is a post-translational mechanism of kinase inhibitor resistance. Cancer Discov.6, 382–399 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 196.Aldonza, M. B. D. et al. Chemotherapy confers a conserved secondary tolerance to EGFR inhibition via AXL-mediated signaling bypass. Sci. Rep.11, 8016 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 197.Han, S. Y. et al. Marsdenia tenacissima extract overcomes Axl- and Met-mediated erlotinib and gefitinib cross-resistance in non-small cell lung cancer cells. Oncotarget8, 56893–56905 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 198.Kim, D. et al. AXL degradation in combination with EGFR-TKI can delay and overcome acquired resistance in human non-small cell lung cancer cells. Cell Death Dis.10, 361 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 199.Cheng, L. C. et al. AXL phosphorylates and up-regulates TNS2 and its implications in IRS-1-associated metabolism in cancer cells. J. Biomed. Sci.25, 80 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 200.Bruce, S. F. et al. GAS6-AXL inhibition by AVB-500 overcomes resistance to paclitaxel in endometrial cancer by decreasing tumor cell glycolysis. Mol. Cancer Ther.21, 1348–1359 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 201.Tanaka, K. et al. Expression profile of receptor-type protein tyrosine kinase genes in the human thyroid. Endocrinology139, 852–858 (1998). [DOI] [PubMed] [Google Scholar]
  • 202.Yeh, C. Y. et al. Transcriptional activation of the Axl and PDGFR-α by c-Met through a ras- and Src-independent mechanism in human bladder cancer. BMC Cancer11, 139 (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 203.Kashyap, M. K. & Abdel-Rahman, O. Expression, regulation and targeting of receptor tyrosine kinases in esophageal squamous cell carcinoma. Mol. Cancer17, 54 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 204.Brown, J. E. et al. Cross-phosphorylation, signaling and proliferative functions of the Tyro3 and Axl receptors in Rat2 cells. PLoS ONE7, e36800 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 205.Revach, O. Y., Sandler, O., Samuels, Y. & Geiger, B. Cross-Talk between receptor tyrosine kinases AXL and ERBB3 regulates invadopodia formation in melanoma cells. Cancer Res.79, 2634–2648 (2019). [DOI] [PubMed] [Google Scholar]
  • 206.Wu, Z. et al. Coexpression of receptor tyrosine kinase AXL and EGFR in human primary lung adenocarcinomas. Hum. Pathol.46, 1935–1944 (2015). [DOI] [PubMed] [Google Scholar]
  • 207.Hsu, P. L., Jou, J. & Tsai, S. J. TYRO3: a potential therapeutic target in cancer. Exp. Biol. Med.244, 83–99 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 208.Avilla, E. et al. Activation of TYRO3/AXL tyrosine kinase receptors in thyroid cancer. Cancer Res.71, 1792–1804 (2011). [DOI] [PubMed] [Google Scholar]
  • 209.Dantas-Barbosa, C. et al. Expression and role of TYRO3 and AXL as potential therapeutical targets in leiomyosarcoma. Br. J. Cancer117, 1787–1797 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 210.Huelse, J. M. et al. MERTK in cancer therapy: targeting the receptor tyrosine kinase in tumor cells and the immune system. Pharm. Ther.213, 107577 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 211.Davra, V. et al. Axl and Mertk receptors cooperate to promote breast cancer progression by combined oncogenic signaling and evasion of host antitumor immunity. Cancer Res.81, 698–712 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 212.Linger, R. M., Keating, A. K., Earp, H. S. & Graham, D. K. Taking aim at Mer and Axl receptor tyrosine kinases as novel therapeutic targets in solid tumors. Expert Opin. Ther. Targets14, 1073–1090 (2010). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 213.Yoshizawa, T. et al. Abstract LB-218: development of Axl/Mer inhibitor, ONO-9330547: preclinical evidence supporting the combination with immunotherapeutics. Cancer Res.76, LB-218-LB-218 (2016). [Google Scholar]
  • 214.Vouri, M. & Hafizi, S. TAM receptor tyrosine kinases in cancer drug resistance. Cancer Res.77, 2775–2778 (2017). [DOI] [PubMed] [Google Scholar]
  • 215.McDaniel, N. K. et al. MERTK mediates intrinsic and adaptive resistance to AXL-targeting agents. Mol. Cancer Ther.17, 2297–2308 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 216.Rios-Doria, J. et al. A potent and selective dual inhibitor of AXL and MERTK possesses both immunomodulatory and tumor-targeted activity. Front. Oncol.10, 598477 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 217.Phan, L. M. et al. Hepatocyte growth factor/cMET pathway activation enhances cancer hallmarks in adrenocortical carcinoma. Cancer Res.75, 4131–4142 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 218.Gujral, T. S. et al. Profiling phospho-signaling networks in breast cancer using reverse-phase protein arrays. Oncogene32, 3470–3476 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 219.Li, W. et al. HGF-induced formation of the MET-AXL-ELMO2-DOCK180 complex promotes RAC1 activation, receptor clustering, and cancer cell migration and invasion. J. Biol. Chem.293, 15397–15418 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 220.Levantini, E., Maroni, G., Del Re, M. & Tenen, D. G. EGFR signaling pathway as therapeutic target in human cancers. Semin Cancer Biol.85, 253–275 (2022). [DOI] [PubMed] [Google Scholar]
  • 221.Meyer, A. S., Miller, M. A., Gertler, F. B. & Lauffenburger, D. A. The receptor AXL diversifies EGFR signaling and limits the response to EGFR-targeted inhibitors in triple-negative breast cancer cells. Sci. Signal6, ra66 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 222.Taniguchi, H. et al. AXL confers intrinsic resistance to osimertinib and advances the emergence of tolerant cells. Nat. Commun.10, 259 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 223.Martínez-Sáez, O. & Prat, A. Current and future management of HER2-positive metastatic breast cancer. JCO Oncol. Pr.17, 594–604 (2021). [DOI] [PubMed] [Google Scholar]
  • 224.Goyette, M. A. et al. Targeting Axl favors an antitumorigenic microenvironment that enhances immunotherapy responses by decreasing Hif-1α levels. Proc. Natl. Acad. Sci. USA. 118, e2023868118 (2021). [DOI] [PMC free article] [PubMed]
  • 225.Liu, L. et al. Novel mechanism of lapatinib resistance in HER2-positive breast tumor cells: activation of AXL. Cancer Res.69, 6871–6878 (2009). [DOI] [PubMed] [Google Scholar]
  • 226.Uliano, J., Corvaja, C., Curigliano, G. & Tarantino, P. Targeting HER3 for cancer treatment: a new horizon for an old target. ESMO Open8, 100790 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 227.Torka, R. et al. Activation of HER3 interferes with antitumor effects of Axl receptor tyrosine kinase inhibitors: suggestion of combination therapy. Neoplasia16, 301–318 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 228.Chhouri, H., Alexandre, D. & Grumolato, L. Mechanisms of acquired resistance and tolerance to EGFR targeted therapy in non-small cell lung cancer. Cancers15, 504 (2023). [DOI] [PMC free article] [PubMed]
  • 229.McDaniel, N. K. et al. AXL mediates cetuximab and radiation resistance through tyrosine 821 and the c-ABL kinase pathway in head and neck cancer. Clin. Cancer Res.26, 4349–4359 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 230.Gallicchio, M. et al. Inhibition of vascular endothelial growth factor receptor 2-mediated endothelial cell activation by Axl tyrosine kinase receptor. Blood105, 1970–1976 (2005). [DOI] [PubMed] [Google Scholar]
  • 231.Fang, Y. & Imoukhuede, P. I. Axl and vascular endothelial growth factor receptors exhibit variations in membrane localization and heterogeneity across monolayer and spheroid high-grade serous ovarian cancer models. GEN Biotechnol.2, 43–56 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 232.Kiyoi, H., Kawashima, N. & Ishikawa, Y. FLT3 mutations in acute myeloid leukemia: therapeutic paradigm beyond inhibitor development. Cancer Sci.111, 312–322 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 233.Park, I. K. et al. Inhibition of the receptor tyrosine kinase Axl impedes activation of the FLT3 internal tandem duplication in human acute myeloid leukemia: implications for Axl as a potential therapeutic target. Blood121, 2064–2073 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 234.Mori, M. et al. Gilteritinib, a FLT3/AXL inhibitor, shows antileukemic activity in mouse models of FLT3 mutated acute myeloid leukemia. Invest. N. Drugs35, 556–565 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 235.Park, I. K. et al. Receptor tyrosine kinase Axl is required for resistance of leukemic cells to FLT3-targeted therapy in acute myeloid leukemia. Leukemia29, 2382–2389 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 236.Dumas, P. Y. et al. Dual inhibition of FLT3 and AXL by gilteritinib overcomes hematopoietic niche-driven resistance mechanisms in FLT3-ITD acute myeloid leukemia. Clin. Cancer Res.27, 6012–6025 (2021). [DOI] [PubMed] [Google Scholar]
  • 237.Qiu, Y. et al. The GSK3β/Mcl-1 axis is regulated by both FLT3-ITD and Axl and determines the apoptosis induction abilities of FLT3-ITD inhibitors. Cell Death Discov.9, 44 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 238.Zou, X. et al. Targeting the PDGF/PDGFR signaling pathway for cancer therapy: a review. Int. J. Biol. Macromol.202, 539–557 (2022). [DOI] [PubMed] [Google Scholar]
  • 239.Mahadevan, D. et al. A novel tyrosine kinase switch is a mechanism of imatinib resistance in gastrointestinal stromal tumors. Oncogene26, 3909–3919 (2007). [DOI] [PubMed] [Google Scholar]
  • 240.Palisoul, M. L. et al. Inhibition of the receptor tyrosine kinase AXL restores paclitaxel chemosensitivity in uterine serous cancer. Mol. Cancer Ther.16, 2881–2891 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 241.Meel, M. H. et al. Combined therapy of AXL and HDAC inhibition reverses mesenchymal transition in diffuse intrinsic pontine glioma. Clin. Cancer Res.26, 3319–3332 (2020). [DOI] [PubMed] [Google Scholar]
  • 242.Ciardiello, D. et al. Dual inhibition of TGFβ and AXL as a novel therapy for human colorectal adenocarcinoma with mesenchymal phenotype. Med. Oncol.38, 24 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 243.Nyakas, M. et al. AXL inhibition improves BRAF-targeted treatment in melanoma. Sci. Rep.12, 5076 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 244.Taverna, J. A. et al. Single-cell proteomic profiling identifies combined AXL and JAK1 inhibition as a novel therapeutic strategy for lung cancer. Cancer Res.80, 1551–1563 (2020). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 245.Sabarwal, A. et al. Abstract 3866: targeting AXL can effectively inhibit c-Met-induced therapeutic resistance in renal cancer. Cancer Res.83, 3866–3866 (2023). [Google Scholar]
  • 246.Sabarwal, A. et al. Abstract 4555: c-Met/AXL crosstalk in mediating therapeutic resistance in renal cell carcinoma. Cancer Res.84, 4555–4555 (2024). [Google Scholar]
  • 247.Machaalani, M. et al. 49 Inhibition of AXL along with c-Met potentially halts resistance development in renal cell carcinoma. Oncologist29, S27–S28 (2024). [Google Scholar]
  • 248.Mims, A. S. et al. A phase 1b/2 study of TP-0903 and decitabine targeting mutant TP53 and/or complex karyotype in patients with untreated acute myeloid leukemia ≥ age 60 years: phase 1b interim results. J. Clin. Oncol.40, 7027–7027 (2022). [Google Scholar]
  • 249.Boshuizen, J. et al. Cooperative targeting of immunotherapy-resistant melanoma and lung cancer by an AXL-targeting antibody-drug conjugate and immune checkpoint blockade. Cancer Res.81, 1775–1787 (2021). [DOI] [PubMed] [Google Scholar]
  • 250.Ruvolo, P. P. et al. TAM Inhibitor ONO-7475 combination with sorafenib is effective using in vitro and in vivo models of FLT3 ITD acute myeloid leukemia. Blood132, 2642–2642 (2018). [Google Scholar]
  • 251.Post, S. M. et al. AXL/MERTK inhibitor ONO-7475 potently synergizes with venetoclax and overcomes venetoclax resistance to kill FLT3-ITD acute myeloid leukemia. Haematologica107, 1311–1322 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 252.Okura, N. et al. ONO-7475, a novel AXL inhibitor, suppresses the adaptive resistance to initial EGFR-TKI treatment in EGFR-mutated non-small cell lung cancer. Clin. Cancer Res.26, 2244–2256 (2020). [DOI] [PubMed] [Google Scholar]
  • 253.Wong, J. et al. Abstract 1735: The potent and selective MERTK/AXL inhibitor PF-5807/ARRY-067 activates dendritic cells to cross-prime CD8+ T cells for anti-tumor activity. Cancer Res.81, 1735 (2021). [Google Scholar]
  • 254.Hirohashi, T. et al. Trials in progress: TAM kinase inhibitor PF-07265807 and sasanlimab plus axitinib in patients with advanced or metastatic renal cell carcinoma—A phase 1, open-label, pharmacokinetic, safety and tolerability study. J. Clin. Oncol.41, TPS743 (2023). [Google Scholar]
  • 255.Choi, B. J. et al. Abstract 2255: patient pharmacodynamic biomarker and pk evaluation results from an ongoing phase I dose-escalation study of q702, an axl, mer and csf1r kinase inhibitor in patients with advanced solid tumors. Cancer Res.83, 2255 (2023). [Google Scholar]
  • 256.Zhou, L. et al. Targeting MET and AXL overcomes resistance to sunitinib therapy in renal cell carcinoma. Oncogene35, 2687–2697 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 257.Kato, Y. et al. Combined effect of cabozantinib and gefitinib in crizotinib-resistant lung tumors harboring ROS1 fusions. Cancer Sci.109, 3149–3158 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 258.Jia, J. et al. Cabozantinib with or without Panitumumab for RAS wild-type metastatic colorectal cancer: impact of MET amplification on clinical outcomes and circulating biomarkers. Cancer Chemother. Pharm.89, 413–422 (2022). [DOI] [PubMed] [Google Scholar]
  • 259.Agarwal, N. et al. Cabozantinib (C) in combination with atezolizumab (A) in patients (pts) with metastatic castration-resistant prostate cancer (mCRPC): results of Cohort 6 of the COSMIC-021 Study. J. Clin. Oncol.38, 139 (2020). [Google Scholar]
  • 260.Agarwal, N. et al. CONTACT-02: Phase 3 study of cabozantinib (C) plus atezolizumab (A) vs second novel hormonal therapy (NHT) in patients (pts) with metastatic castration-resistant prostate cancer (mCRPC). J. Clin. Oncol.42, 18 (2024). [Google Scholar]
  • 261.Powles, T. et al. Nivolumab plus cabozantinib versus sunitinib for first-line treatment of advanced renal cell carcinoma: extended follow-up from the phase III randomised CheckMate 9ER trial. ESMO Open9, 102994 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 262.Choueiri, T. K. et al. Cabozantinib plus nivolumab and ipilimumab in renal-cell carcinoma. N. Engl. J. Med388, 1767–1778 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 263.Fan, L. et al. The safety, tolerability, and preliminary antitumor activity of sitravatinib plus tislelizumab in patients with locally recurrent or metastatic triple-negative breast cancer. J. Clin. Oncol.40, 1070 (2022). [Google Scholar]
  • 264.Wang, X. et al. SAFFRON-103: a phase Ib study of sitravatinib plus tislelizumab in anti-PD-(L)1 refractory/resistant advanced melanoma. Immunotherapy16, 243–256 (2024). [DOI] [PubMed] [Google Scholar]
  • 265.Zhao, J. et al. SAFFRON-103: a phase 1b study of the safety and efficacy of sitravatinib combined with tislelizumab in patients with locally advanced or metastatic non-small cell lung cancer. J. Immunother. Cancer11, e006055 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 266.Msaouel, P. et al. A phase 2 study of sitravatinib in combination with nivolumab in patients with advanced or metastatic urothelial carcinoma. Eur. Urol. Oncol.7, 933–943 (2024). [DOI] [PubMed] [Google Scholar]
  • 267.Borghaei, H. et al. SAPPHIRE: phase III study of sitravatinib plus nivolumab versus docetaxel in advanced nonsquamous non-small-cell lung cancer. Ann. Oncol.35, 66–76 (2024). [DOI] [PubMed] [Google Scholar]
  • 268.Rho, J. K. et al. MET and AXL inhibitor NPS-1034 exerts efficacy against lung cancer cells resistant to EGFR kinase inhibitors because of MET or AXL activation. Cancer Res.74, 253–262 (2014). [DOI] [PubMed] [Google Scholar]
  • 269.Yang, Y. M. et al. AXL/MET dual inhibitor, CB469, has activity in non-small cell lung cancer with acquired resistance to EGFR TKI with AXL or MET activation. Lung Cancer146, 70–77 (2020). [DOI] [PubMed] [Google Scholar]
  • 270.Lei, T. et al. Anlotinib combined with osimertinib reverses acquired osimertinib resistance in NSCLC by targeting the c-MET/MYC/AXL axis. Pharm. Res.188, 106668 (2023). [DOI] [PubMed] [Google Scholar]
  • 271.Clemenson, C. et al. The MET/AXL/FGFR inhibitor S49076 impairs Aurora B activity and improves the antitumor efficacy of radiotherapy. Mol. Cancer Ther.16, 2107–2119 (2017). [DOI] [PubMed] [Google Scholar]
  • 272.Powles, T. et al. Outcomes based on prior therapy in the phase 3 METEOR trial of cabozantinib versus everolimus in advanced renal cell carcinoma. Br. J. Cancer119, 663–669 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 273.Hellerstedt, B. A. et al. Results of a phase II placebo-controlled randomized discontinuation trial of cabozantinib in patients with non-small-cell lung carcinoma. Clin. Lung Cancer20, 74–81.e71 (2019). [DOI] [PubMed] [Google Scholar]

Articles from Signal Transduction and Targeted Therapy are provided here courtesy of Nature Publishing Group

RESOURCES