Next Article in Journal
IL-1β and HMGB1 in Epileptogenesis: Recent Advances and Clinical Translation
Previous Article in Journal
Tumor-Derived Microvesicles Promote Kidney Regeneration and Cytoprotective Immunomodulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Targeting Ion Channels for Cancer Therapy: From Pathophysiological Mechanisms to Clinical Translation

Department of Anesthesiology, State Key Laboratory of Oncology in South China, Collaborative Innovation Centre for Cancer Medicine, Guangdong Provincial Clinical Research Center for Cancer, Sun Yat-sen University Cancer Center, Guangzhou 510060, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this study.
Pharmaceuticals 2025, 18(10), 1521; https://doi.org/10.3390/ph18101521
Submission received: 14 August 2025 / Revised: 23 September 2025 / Accepted: 26 September 2025 / Published: 10 October 2025
(This article belongs to the Special Issue Ion Channels as Pharmacological Targets in Cancer)

Abstract

Cancer remains a major global health burden, representing one of the leading causes of mortality among noncommunicable diseases worldwide. Although conventional treatment modalities, including surgical resection, chemotherapy, radiotherapy, targeted therapy, and immunotherapeutic interventions, have demonstrated clinical benefits, their therapeutic efficacy is often constrained by inherent limitations such as low specificity, systemic toxicity, or tumor heterogeneity. These challenges underscore the imperative for developing innovative treatment strategies. Emerging evidence has implicated ion channels as critical players in oncogenesis and cancer progression. These proteins modulate diverse oncogenic phenotypes, including uncontrolled proliferation, metastatic dissemination, and apoptotic resistance. Their frequent dysregulation in malignancies correlates with disease aggressiveness and clinical outcomes, positioning them as promising targets for precision oncology. Notably, pharmacological modulation of ion channels exerts multifaceted antitumor effects, with several channel-targeting agents advancing through clinical trials. This review explores recent advances in ion channel-targeted therapies, emphasizing their mechanisms, clinical applications, and challenges. Furthermore, we examine the pathophysiological contributions of ion channels to tumor biology and evaluate their emerging utility as predictive biomarkers, providing perspectives on addressing critical gaps in current oncologic management.

1. Introduction

Cancer remains one of the most formidable global public health challenges of the 21st century, with nearly 20 million new cases diagnosed annually and accounting for almost one-quarter of all deaths from noncommunicable diseases. Demographic projections indicate that the global cancer burden will escalate to approximately 35 million new cases per year by 2050 [1]. Despite significant progress in conventional antitumor strategies—including surgery, chemotherapy, radiotherapy, targeted therapy, and the rapidly evolving field of immunotherapy—used alone or in combination, a substantial proportion of patients continue to derive limited clinical benefit because the efficacy of these therapies is affected by factors such as low therapeutic specificity, severe toxicities, acquired resistance, limited applicable population, and dynamic tumor microenvironment [2,3,4,5]. As a result, clinical outcomes remain suboptimal for many patients, highlighting an urgent and unmet need for innovative therapeutic strategies.
The asymmetric distribution of ions across the cell membrane establishes electrochemical gradients that are fundamental to numerous cellular processes. Ion movement across the lipid bilayer is indispensable for maintaining cellular homeostasis and can be broadly classified into passive and active transport [6]. Active transport, exemplified by the sodium–potassium pump (Na+/K+-ATPase), requires direct ATP hydrolysis to move ions against their electrochemical gradient (from low to high concentration), thereby maintaining essential transmembrane gradients [7]. These gradients not only enable passive ion movement but also underpin a wide range of physiological functions. By contrast, passive transport relies on diffusion along the electrochemical gradient (from high to low concentration) without energy expenditure. It occurs either by simple diffusion across the lipid bilayer or via facilitated diffusion mediated by specific membrane transport proteins.
Ion channels are essential mediators of facilitated diffusion, enabling rapid and selective passive ion flux across cellular membranes in response to defined stimuli. This process is vital for maintaining the membrane potential (Vm) and regulating a broad spectrum of physiological functions, including electrophysiological signaling, molecular transport, and cellular energy metabolism [8,9]. Based on their gating mechanisms, ion channels are commonly classified into four major categories [10]: (a) voltage-gated channels, which respond to changes in membrane potential to control cellular excitability; (b) ligand-gated channels, which undergo conformational changes upon ligand binding; (c) mechanosensitive channels, which are activated by mechanical forces such as pressure and contribute to processes like vascular tone and sensory transduction; and (d) thermosensitive channels, which are modulated by temperature fluctuations. Additionally, ion channels are categorized according to ion selectivity, most notably for sodium (Na+), potassium (K+), calcium (Ca2+), and chloride (Cl) ions [11]. Dysregulation of these channels has been increasingly linked to tumorigenesis, influencing key processes such as cancer cell proliferation, migration, apoptosis, and resistance to therapy [12,13]. Aberrant expression patterns have also been associated with patient prognosis, highlighting ion channels as promising therapeutic targets in the context of precision oncology [14].
Targeting ion channels has emerged as a promising anticancer strategy, offering multifaceted therapeutic benefits, including the inhibition of tumor growth, enhancement of treatment sensitivity, and reversal of drug resistance (Table 1) [12]. Several ion channel-modulating agents have progressed into clinical evaluation, reflecting growing translational interest in this therapeutic avenue (Table 2) [15,16,17]. Given the breadth and complexity of research accumulated over the past three decades, this review does not aim to provide a comprehensive inventory of all ion channels implicated in tumorigenesis. Rather, it focuses on a selection of representative channels and their modulators, chosen for their established relevance to cancer hallmarks and the robustness of supporting preclinical and clinical evidence. Specifically, we examine their pathophysiological functions, current and potential clinical applications, and the principal challenges hindering their broader implementation. Due to constraints of length and scope, some ion channels, transporters, and cancer types of interest are not discussed.

2. The Role of Ion Channels in Cancer

2.1. Sodium Channel

Sodium ions (Na+) traverse the plasma membrane via multiple mechanisms, including the activity of the transmembrane Na+ gradient, Na+-dependent transporters, specific membrane channels, and energy-dependent processes [7]. Central to maintaining this critical gradient is the Na+/K+-ATPase, which hydrolyzes ATP to actively export Na+ and import K+. Sodium-selective channels—such as voltage-gated sodium channels (VGSCs), epithelial sodium channels (ENaCs), and acid-sensing ion channels (ASICs)—further mediate Na+ influx. This coordinated regulation results in an extracellular Na+ concentration approximately tenfold higher than that within the cytoplasm. Maintenance of Na+ homeostasis is indispensable for electrophysiological stability, and its disruption is a recognized pathogenic mechanism in acute inflammatory responses and ischemic injury [105,106]. Increasing evidence indicates that sodium imbalance also contributes to tumor progression, as malignant tissues often accumulate higher Na+ concentrations compared with normal tissues [107,108,109,110]. Such alterations may arise from aberrant Na+/K+-ATPase activity or dysregulated Na+ channel expression. Consequently, quantifying intratumoral Na+ concentration has been proposed to be a predictive biomarker for chemotherapy responsiveness [111].
VGSCs are integral membrane proteins that mediate Na+ flux across cellular membranes and are essential initiators of action potentials in excitable cells. They thereby govern electrical signaling and play fundamental roles in diverse physiological functions. Structurally, VGSCs are composed of a principal pore-forming α-subunit and one or more auxiliary β-subunits [112]. To date, nine mammalian α-subunit isoforms (NaV1.1–NaV1.9), encoded by SCN1A-SCN5A and SCN8A-SCN11A, have been identified. The β-subunits consist of five variants (β1a, β1b, β2, β3, and β4) encoded by SCN1B-SCN4B. Based on their sensitivity to tetrodotoxin (TTX), VGSCs are classified into two subtypes: TTX-resistant isoforms (NaV1.5, NaV1.8, and NaV1.9), which require micromolar TTX concentrations for inhibition, and TTX-sensitive isoforms, which are inhibited at nanomolar concentrations [113].
Aberrant VGSC expression and activity have been widely implicated in cancer biology. These channels promote tumor cell proliferation, invasion, and metastatic dissemination by modulating cellular excitability and downstream signaling. VGSCs are expressed across multiple malignancies, including colorectal, prostate, breast, lung, cervical, ovarian cancers, and melanoma. For example, NaV1.2 mRNA is upregulated in ovarian cancer [114], while NaV1.6 protein is markedly overexpressed in prostate and colorectal cancers [115,116]. In breast cancer, NaV1.5 is significantly overexpressed at both mRNA and protein levels compared with normal tissue, correlating strongly with disease recurrence and metastatic progression [117]. Mechanistic studies support these observations: siRNA-mediated silencing of NaV1.5 markedly inhibits proliferation and invasion of astrocytoma cells [118]. NaV1.6, upregulated in follicular thyroid carcinoma, promotes proliferation, epithelial–mesenchymal transition (EMT), and invasion through activation of the JAK2/STAT3 pathway [119]. Similarly, NaV1.7 drives gastric cancer progression via the MACC1-NHE1 axis [120]. In addition, the β1 subunit regulates cell–cell adhesion and migration in breast cancer [121].
Beyond VGSCs, ENaCs constitute another group of Na+ channels linked to oncogenesis. ENaCs, which mediate sodium reabsorption and water retention in renal collecting ducts, are fundamental to systemic fluid–electrolyte balance and blood pressure regulation. Their aberrant expression in malignancies such as hepatocellular carcinoma, melanoma, brain cancer, and breast cancer has been shown to facilitate tumor progression. ASICs, members of the ENaC/degenerin (ENaC/DEG) superfamily [122], also enhance oncogenic behaviors, including proliferation, EMT, and invasion, in colorectal and pancreatic cancers [123,124].
The tumor microenvironment (TME), comprising malignant cells, stromal elements, immune infiltrates, and the extracellular matrix (ECM), profoundly influences Na+ channel expression and activity. Hypoxia—a hallmark of solid tumors—induces Na+ channel upregulation, as demonstrated by increased NaV1.7 mRNA expression in prostate cancer cells [125]. In colorectal cancer, hypoxia-driven invasiveness depends on sustained Na+ current carried by the neonatal splice variant nNaV1.5 [126]. Moreover, extracellular acidosis in the TME modulates Na+ channel activity and exacerbates cancer progression. In breast cancer, a NaV1.5-dependent positive feedback loop enhances Na+ influx and glycolytic H+ export, thereby promoting extracellular acidification, ECM degradation, and tumor invasion [117].

2.2. Calcium Channel

As a vital second messenger, intracellular Ca2+ regulates fundamental biological processes, including proliferation, differentiation, migration, and programmed cell death [127,128]. Cells maintain an exceptionally low cytosolic Ca2+ concentration through active transport mechanisms, creating a steep 10- to 15,000-fold gradient relative to the extracellular milieu [128]. In response to diverse stimuli, transient elevations in cytosolic Ca2+ are initiated through release from the endoplasmic reticulum (ER) and Golgi apparatus (GA), efflux from mitochondria or lysosomes, and influx from the extracellular space. Termination of Ca2+ signaling occurs via sequestration into the ER, GA, or mitochondria, or by extrusion into the extracellular compartment.
Mounting evidence indicates that disruption of intracellular Ca2+ homeostasis accompanies malignant transformation, driving uncontrolled proliferation, enhanced survival, and metastatic potential. Ca2+ channels, which regulate cellular Ca2+ flux, are thus critically involved in tumor progression through alterations in their expression and function [129]. Plasma membrane Ca2+ channels are broadly divided into four classes: voltage-gated calcium channels (VGCCs), ligand-gated calcium channels (LGCCs), store-operated calcium channels (SOCCs), and mechanosensitive calcium channels (Figure 1). VGCCs open in response to membrane depolarization, and they share a conserved structure consisting of a pore-forming α1 subunit and auxiliary β subunits. Based on α1 subunit composition, they are classified into L-, P/Q-, N-, R-, and T-types, encoded by 10 distinct genes and grouped into three subfamilies (CaV1.x, CaV2.x, and CaV3.x) [130]. LGCCs are activated by specific ligands and typically couple to intracellular signaling cascades. SOCCs are triggered by depletion of intracellular Ca2+ stores: when ER Ca2+ falls, the ER-resident protein stromal interaction molecule 1 (STIM1) undergoes conformational change, enabling interaction with the plasma membrane channel ORAI1. This opens ORAI1 to facilitate extracellular Ca2+ influx, replenishing ER stores via sarco/endoplasmic reticulum Ca2+-ATPase (SERCA)—a process termed store-operated Ca2+ entry (SOCE) [131]. Mechanosensitive Ca2+ channels, including Piezo and transient receptor potential (TRP) families, respond to mechanical stimuli such as stretch, compression, or membrane tension, mediating transmembrane Ca2+ influx [132].
Ca2+ channels function as pivotal regulators of intracellular Ca2+ influx, essential for sustaining physiological functions within neuronal networks, the cardiovascular system (including cardiac and smooth muscle), and skeletal muscle. Substantial evidence indicates that dysregulation of Ca2+ signaling is pathologically implicated in neurodegenerative disorders such as Parkinson’s disease; neurological conditions, including epilepsy; pain syndromes; autism spectrum disorder; and various cardiovascular dysfunctions [133]. Clinically, L-type Ca2+ channels are established pharmacological targets for widely used antihypertensive and antiarrhythmic agents, such as nifedipine and verapamil. The N-type Ca2+ channel is targeted by ziconotide for pain management [134,135]. Furthermore, gabapentinoids, which act on the α2δ-1 subunit of VGCCs, are employed in the treatment of neuropathic pain [136].
Recent research has underscored the critical role of Ca2+ channels in malignant tumors. Accumulating evidence indicates that Ca2+ signaling remodels TME and regulates key cancer cell behaviors, including proliferation, apoptosis, migration, and invasion. Specifically, He et al. demonstrated that Ca2+ influx promotes the polarization of tumor-associated macrophages (TAMs) from M2 toward M1 phenotype, thereby enhancing systemic antitumor immune responses [137]. Aberrant expression of multiple Ca2+ channels is a common feature in tumor cells. Notably, the VGCC subunit CaV1.3 is frequently dysregulated in various cancers, with its encoding gene CACNA1D exhibiting elevated expression across diverse malignancies. In prostate cancer, CACNA1D mRNA and protein levels are significantly upregulated, and its overexpression correlates with an aggressive tumor phenotype [138]. Correspondingly, pharmacological blockade of Ca2+ channels has been associated with a reduced incidence of prostate cancer [139]. Similarly, CaV1.3 dysregulation extends to breast cancer. CaV1.3 is also highly expressed in breast cancer tissues. Downregulation of CACNA1D inhibits filopodia formation and suppresses the invasive capacity of breast cancer cells [140]. Beyond prostate and breast cancer, the α1D protein of CaV1.3 is overexpressed in colorectal cancer and endometrial carcinoma compared to adjacent normal tissues. Silencing the CACNA1D gene (encoding α1D) disrupts Ca2+ homeostasis and impairs the migration of colorectal cancer and endometrial carcinoma cells [141,142]. Furthermore, dysregulation extends to other Ca2+ channels. The gene CACNA1E, encoding the α1 subunit of R-type VGCCs, is elevated in non-small-cell lung cancer (NSCLC) tissues and promotes NSCLC cell proliferation [143]. Amplification and overexpression of CACNA1E are also associated with disease relapse in favorable histology Wilms’ tumors [144].
SOCCs are also critically important in tumor biology. Yang et al. demonstrated that STIM1/ORAI1-mediated SOCE is indispensable for breast cancer cell migration and metastasis, establishing that Ca2+ influx via SOCE is a prerequisite for these processes. Genetic silencing of ORAI1 or STIM1 using RNA interference, as well as pharmacological blockade of SOCCs, significantly reduced metastatic burden in vivo [145]. Motiani et al. further identified ORAI3, a mammalian homolog of ORAI1, as a key contributor to the pathogenesis of estrogen receptor-positive (ER+) breast cancer [146,147,148]. Subsequent work by Faouzi et al. confirmed that ORAI3 regulates proliferation and cell-cycle progression in breast cancer cells [149,150]. In colorectal cancer (CRC), STIM1 overexpression correlates with tumor size, lymph node metastasis, depth of invasion, and elevated carcinoembryonic antigen levels. Mechanistic studies demonstrated that STIM1 enhances CRC cell migration by upregulating COX-2 expression and PGE2 production [151].
Mechanosensitive Ca2+ channels also play essential roles in tumor progression. Hasegawa et al. showed that the transcription factor YAP regulates Piezo1 expression in oral squamous cell carcinoma (OSCC), and that Piezo1 upregulation promotes OSCC cell proliferation [152]. Moreover, Piezo1 is highly expressed in human glioma and serves as an independent prognostic marker for patient outcomes [153].
TRP channels represent another major class of mechanosensitive Ca2+ channels. TRPs are nonselective cation channels that respond to diverse environmental and intracellular stimuli, and they are grouped into seven subfamilies based on sequence homology: TRPA (ankyrin), TRPC (canonical), TRPM (melastatin), TRPML (mucolipin), TRPN (nompC), TRPP (polycystin), and TRPV (vanilloid) [154,155]. With the exception of TRPN1, all subfamilies are expressed in mammals and fulfil critical physiological functions. In addition to their canonical roles in sensing thermal, mechanical, and chemical cues, TRP channels are increasingly recognized as regulators of cancer cell fate, modulating proliferation, apoptosis, and metastatic potential. Intriguingly, the identification of several TRP channels is historically linked to cancer research. For example, TRPM1 was initially discovered as a prognostic marker in melanoma, leading to its designation as “melastatin” [156].
TRPA1, the only member of the ankyrin subfamily, exhibits context-dependent effects in cancer. Under hypoxic conditions, TRPA1 suppresses invasion in lung cancer cells in vitro, whereas in breast and lung cancer spheroids, it enhances resistance to reactive oxygen species (ROS) and activates Ca2+-dependent anti-apoptotic signaling pathways [157]. Although TRPV1 is primarily localized to the plasma membrane and TRPV2 to intracellular membranes, both contribute to pro-survival signaling. Elevated TRPV1 expression has been reported in lung and breast cancers, where its overactivation promotes proliferation and migration of esophageal squamous cells [158,159]. TRPV2 expression is markedly increased in gastric cancer, where it has been proposed to be a prognostic biomarker, and in prostate cancer, where it is associated with castration resistance and metastatic progression [160,161].
Beyond their roles in survival signaling, TRP channels also regulate tumor angiogenesis and metastatic dissemination [162]. TRPV3 promotes angiogenesis in non-small-cell lung cancer (NSCLC) through the HIF-1α/VEGF pathway [163], while TRPV4 modulates mechanosensitivity in tumor endothelial cells to regulate vascular maturation [164]. TRPV5 and TRPV6, both highly Ca2+-selective channels with structural and functional homology, have also been implicated in cancer. TRPV6 is upregulated in thyroid, prostate, colon, and breast cancers [165], and its silencing suppresses pancreatic cancer progression [166]. Similarly, TRPC1 drives proliferation in follicular thyroid cancer via a Ca2+-dependent mechanism involving S1P3 and VEGFR2 modulation [167].
Among the TRPM subfamily, TRPM1 was the first identified in melanoma and is associated with tumor progression and survival, potentially through CaMKII-mediated activation of AKT signaling, thereby enhancing colony formation and tumor growth [168,169]. TRPM7 is the most extensively studied TRPM member in oncology. It has been shown to be indispensable for metastasis formation in breast cancer mouse models and is associated with poor prognosis in breast cancer patients [170]. Similarly, TRPM7 promotes migration in nasopharyngeal carcinoma (NPC) cells and correlates with unfavorable clinical outcomes [171]. TRPM8, initially identified as a cold- and menthol-sensitive receptor in prostate tissue [172], also exerts context-dependent functions in cancer. In breast cancer, TRPM8 activation inhibits proliferation and migration via AMPK signaling and autophagy modulation [173], whereas, in bladder cancer, it induces cell death through mitochondrial regulation [174].
Specific Ca2+ channels have also been linked to therapeutic resistance. TRPC6 has been shown to promote paclitaxel resistance in breast cancer cells by modulating integrin α6 mRNA splicing and TAZ activation. Pharmacological inhibition of TRPC6 with BI-749327 attenuates cancer stem cell-like properties and restores paclitaxel sensitivity in resistant organoids [30]. In gastric cancer, high TRPV2 expression correlates with advanced stage and poor prognosis, and mechanistically, TRPV2 confers cisplatin resistance by modulating cisplatin-induced apoptosis [175]. VGCC subunit CaV1.3 contributes to resistance against androgen deprivation therapy in prostate cancer [176], while the α2δ-1 subunit encoded by CACNA2D1 has been shown to enhance radioresistance in NSCLC [177].
Collectively, these findings demonstrate that Ca2+ channels regulate not only cancer initiation and progression but also therapeutic sensitivity. They therefore hold promise as diagnostic biomarkers and therapeutic targets in precision oncology.

2.3. Potassium Channel

Potassium ions (K+) represent the predominant intracellular cation in human cells, and their concentration is critical for maintaining cellular electrophysiological stability. K+ channels are the principal transmembrane conduits for K+ flux, regulating electrochemical gradients essential for action potential generation and cellular excitability.
Among ion channels, K+ channels display the greatest structural diversity and functional complexity. The Nomenclature Committee of the International Union of Basic and Clinical Pharmacology (NC-IUPHAR) has established a standardized classification system based on gating mechanisms and structural features [10]. Structurally, K+ channels are grouped into three major categories according to their number of transmembrane domains (TMs): 2TM, 4TM, and 6TM. Functionally, they are divided into four classes: voltage-gated potassium channels (KV), inwardly rectifying potassium channels (Kir), two-pore domain potassium channels (K2P), and calcium-activated potassium channels (KCa) (Figure 1). KV channels open or close in response to changes in membrane potential and are further subdivided into 12 subfamilies (KV1-12). Kir channels exhibit strong inward rectification, characterized by significantly greater inward than outward K+ conductance. This property is crucial for maintaining the resting membrane potential and regulating cellular excitability. Based on structural and functional properties, Kir channels are classified into seven subfamilies (Kir1-7). K2P channels, also termed background or leak channels, provide basal K+ conductance and are divided into six subgroups: TWIK, TASK, TREK, TRESK, THIK, and TALK [178,179]. KCa channels are activated by intracellular Ca2+ and are subclassified according to their single-channel conductance: large conductance (BKCa/MaxiK), intermediate conductance (IKCa), and small conductance (SKCa) [180].
Aberrant expression and dysregulation of K+ channels are observed across multiple cancer types, where they contribute to tumor initiation, progression, and clinical outcomes. In breast cancer, several K+ channels display dynamic expression changes. BKCa channels, for instance, show specific expression patterns that significantly correlate with clinical prognosis [181]. The oncogenic role of K+ channels is exemplified by the voltage-gated channel EAG1 (KV10.1), which is aberrantly expressed in diverse tumors and has been shown to promote cellular transformation in vitro [182,183,184]. KCNK1, a K2P channel member, is highly expressed in breast cancer and promotes metabolic reprogramming through activation of lactate dehydrogenase A (LDHA) [185]. Similarly, elevated KCa3.1 expression in breast cancer confers radioresistance by enhancing activation in response to ionizing radiation [186]. In lung cancer, the calcium-activated potassium channel KCNN4 is significantly overexpressed in lung adenocarcinoma, where it regulates apoptosis and EMT via the PI3K/AKT and MEK/ERK pathways [187]. TASK-1, a two-pore domain acid-sensitive K+ channel, modulates proliferation and apoptosis in non-small-cell lung cancer (NSCLC) [188]. Inwardly rectifying K+ channel Kir2.1 is upregulated in small-cell lung cancer (SCLC), with expression levels correlating with disease stage [189]. Beyond lung cancer, KCa3.1 expression has prognostic significance in pancreatic ductal adenocarcinoma (PDAC) [190], while the voltage-gated K+ channel KV1.3 regulates apoptotic processes in melanoma [191].

2.4. Chloride Channel

As the most abundant anions in the human body, chloride ions (Cl) are indispensable for maintaining osmotic pressure, cellular volume, and acid–base balance, as well as facilitating carbon dioxide transport through the chloride shift mechanism [192]. The transmembrane Cl gradient is tightly regulated by the coordinated action of passive Cl channels and active transporters.
Electrophysiological studies classify Cl channels into several major groups: the chloride channel (ClC) family, the cystic fibrosis transmembrane conductance regulator (CFTR), calcium-activated chloride channels (CaCCs), volume-regulated anion channels (VRACs), and ligand-gated chloride channels (LGClCs) [193]. The ClC family plays a central role in modulating membrane potential and maintaining Cl homeostasis within intracellular organelles. It consists of nine members (ClC-1 to ClC-7, ClC-Ka, and ClC-Kb), which exhibit diverse functions despite conserved sequence motifs. Evolutionarily, the ClC family has diverged into two functional classes: voltage-gated chloride channels and Cl/H+ antiporters [192].
CFTR, a unique member of the ATP-binding cassette (ABC) transporter superfamily, is best known for the ΔF508 deletion mutation, which disrupts protein folding and trafficking, leading to cystic fibrosis (CF). Beyond its role in Cl conductance, CFTR also regulates the activity of other channels, including epithelial sodium channels (ENaCs) and CaCCs. CaCCs are predominantly formed by members of the TMEM16 family, with TMEM16A (ANO1) identified as the principal molecular determinant of CaCC function. VRACs facilitate regulatory volume decrease (RVD) by mediating Cl efflux in response to cell swelling [194]. These channels are heteromeric complexes that require LRRC8A as an essential structural component, which combines with other LRRC8 isoforms (LRRC8B-E) to form functional assemblies [195]. LGClCs, members of the Cys-loop ligand-gated ion channel superfamily, include γ-aminobutyric acid type A receptors (GABA_A-Rs) and glycine receptors (GlyRs), the predominant neuronal LGClCs mediating inhibitory neurotransmission in the central nervous system [196].
Dysregulation of specific Cl channels in cancer cells functionally contributes to survival and progression. As reviewed by Hong et al. [197], ClC-3 plays a particularly prominent role in cancer biology. In malignant glial cells, ClC-3 acts as a key regulator of the cell cycle [198]. In human glioma cells, ClC-3 is enriched at the plasma membrane and activated through Ca2+/calmodulin-dependent kinase II (CaMKII) signaling [199]. Genetic ablation of ClC-3 suppresses CaMKII-mediated chloride currents and reduces bradykinin-stimulated migration of glioma cells [200]. Similarly, silencing ClC-3 expression induces G0/G1 cell-cycle arrest and inhibits migration in nasopharyngeal carcinoma (NPC) cells [201,202]. In cervical carcinoma, both ClC-3 mRNA and protein are significantly upregulated compared with normal cervical tissues, and overexpression correlates with lymph node metastasis and poor prognosis [203].
While dysfunction of ClC channels has been implicated in diverse cancers, the role of CFTR in cancer risk is particularly notable in the context of cystic fibrosis (CF)—a multisystem genetic disorder caused by mutations in the CFTR gene. CF is associated with an increased risk of multiple malignancies through multifaceted mechanisms, including disruption of ion homeostasis, dysregulation of EMT and immune responses, and alterations in signaling pathways critical for proliferation and survival [204]. Clinical studies consistently support this association, particularly within the gastrointestinal tract. Initial reports in The New England Journal of Medicine identified a significantly increased risk of digestive tract cancers in CF patients [205], which was subsequently confirmed by long-term follow-up studies [206,207]. Notably, CF patients exhibit a 5-to-10-fold-higher risk of colorectal cancer (CRC), prompting recommendations for routine colonoscopic surveillance as the primary screening strategy for this population [208]. Elevated risks of pancreatic cancer, testicular cancer, and lymphoid leukemia have also been documented [207,209]. Nevertheless, the precise molecular mechanisms by which CFTR modulates tumor biology remain incompletely understood.
Beyond CFTR, other Clchannels such as TMEM16A/ANO1 have been strongly linked to oncogenesis. TMEM16A encodes a member of the CaCC family, which regulates diverse physiological processes, including glandular secretion, smooth muscle contraction, and sensory transduction. Identified as a CaCC in 2008, TMEM16A has also been implicated in vascular regulation, emerging as a potential therapeutic target for hypertension [210,211,212,213,214]. Dysregulated ANO1 expression promotes oncogenic progression in multiple malignancies, including breast cancer, gastric cancer, and head and neck squamous cell carcinoma (HNSCC) [77,215,216]. Furthermore, aberrant ANO1 serves as a prognostic biomarker for CRC and pancreatic cancer [217,218]. Mechanistically, ANO1 activates ERK1/2 signaling and modulates the subcellular localization of p27Kip1, thereby regulating cell viability and apoptosis in HNSCC [215,219]. More recently, ANO1-mediated PI3K-AKT activation has been shown to impair ferroptosis in gastrointestinal cancers while simultaneously remodeling the tumor microenvironment through recruitment of cancer-associated fibroblasts (CAFs) and suppression of CD8+ T-cell activity [220].
Similarly, VRACs and LGClCs have been implicated in cancer biology. LRRC8A, the essential structural component of VRACs [194], has emerged as a promising prognostic biomarker across several malignancies [221,222]. For example, Zhang et al. demonstrated that LRRC8A promotes colon cancer metastasis by regulating the PIP5K1B/PIP2 signaling pathway [223]. With respect to LGClCs, growing evidence supports their involvement in tumorigenesis and progression. In breast cancer, the α3 subunit-encoding gene GABRA3 is significantly upregulated in both clinical specimens and cell models. Interestingly, its RNA-edited isoform exerts tumor-suppressive effects by inhibiting cellular migration, invasion, and metastasis [224]. Conversely, the pi subunit-encoding gene GABRP promotes migration in basal-like breast cancer cells through ERK1/2 pathway activation [225].

2.5. Trace Metal Ion Channel

Unlike the major ions that typically diffuse passively through selective channels, trace elements predominantly rely on active transport mechanisms mediated by specialized transmembrane transporters. Distinct metalloprotein families exemplify this regulation: iron homeostasis is maintained by divalent metal transporter 1 (DMT1) and ferroportin (FPN); zinc balance is achieved through the coordinated actions of Zrt-/Irt-like protein (ZIP, SLC39A) importers and zinc transporter (ZnT) exporters; and copper transport is governed by copper transporter 1 (CTR1) and the P-type ATPases ATP7A/B. In addition to these dedicated transporters, increasing evidence highlights the contribution of nonselective cation channels, such as TRPM7 and TRPML1, in mediating the transmembrane flux of essential divalent cations (e.g., Zn2+, Mn2+, and Fe2+). These channels operate in concert with ATP-dependent transporters to ensure cellular metal homeostasis [226,227].
TRPM7, a unique bifunctional protein combining ion channel and kinase activities, exerts pleiotropic effects on tumor initiation, progression, and metastasis. Aberrant overexpression of TRPM7 has been reported across multiple cancers, where it regulates proliferation, migration, invasion, and EMT through both its ion-conducting and kinase functions [228]. For example, in pancreatic ductal adenocarcinoma PDAC, TRPM7 overexpression correlates with poor prognosis, and its kinase domain is required for interaction with PAK1 to maintain the mesenchymal phenotype of PDAC cells [229]. In breast cancer, TRPM7’s oncogenic role is partly mediated by Zn2+-dependent regulation of MDMX stability [230]. Beyond intrinsic tumor cell regulation, TRPM7 also shapes the TME: it mediates Mg2+ influx into macrophages, driving M2 polarization and thereby accelerating tumor progression [231].
Among the mucolipin TRP subfamily, TRPML1 (MCOLN1) is widely expressed and primarily localized to lysosomes. Lysosomal patch-clamp studies have confirmed TRPML1 as a nonselective cation channel facilitating efflux of divalent cations (Ca2+, Zn2+, Fe2+) from the lysosomal lumen. Recent findings implicate TRPML1 in regulating lysosomal functions that confer tumor resistance to ferroptosis and modulate autophagy [232,233]. Notably, TRPML1-mediated lysosomal exocytosis has been identified as a critical anti-ferroptotic mechanism. In AKT-hyperactivated cancers, TRPML1 is phosphorylated by AKT at Ser343, preventing ubiquitination-mediated degradation. Stabilized TRPML1 interacts with ARL8B to initiate lysosomal exocytosis, thereby reducing intracellular Fe2+ levels, enhancing membrane repair, and promoting resistance to ferroptosis. Importantly, a synthetic peptide targeting TRPML1 disrupts the TRPML1-ARL8B interaction, suppresses lysosomal exocytosis, amplifies ferroptosis, and enhances sensitivity of AKT-driven tumors to radiotherapy and immunotherapy in vivo [233]. Furthermore, TRPML1 has been reported to regulate oncogenic autophagy through Zn2+ influx, further underscoring its multifaceted role in cancer [234].

2.6. Ion Transporters

Similar to ion channels, ion transporters are essential membrane proteins that mediate ion movement across the lipid bilayer [235]. However, their operational mechanisms are fundamentally distinct: ion channels permit passive diffusion along electrochemical gradients, whereas ion transporters actively translocate ions against these gradients, typically through energy expenditure. Based on their functional mechanisms, ion transporters can be broadly divided into two categories: ion pumps and ion exchangers.
Ion pumps are active transporters that directly utilize ATP hydrolysis to move ions against their gradients, thereby establishing and maintaining transmembrane ion homeostasis. Classic examples include the sodium–potassium pump (Na+/K+-ATPase), which expels three Na+ and imports two K+ ions per ATP hydrolyzed to sustain resting membrane potential and cellular volume stability, and the calcium pump (Ca2+-ATPase), which extrudes Ca2+ to the extracellular space or sequesters it into the endoplasmic reticulum, maintaining low cytosolic Ca2+ essential for signaling. Proton pumps (H+-ATPases) are indispensable for gastric acid secretion and the acidification of intracellular organelles. By contrast, ion exchangers mediate secondary active transport, harnessing the energy stored in pre-established ion gradients rather than directly consuming ATP. For example, the Na+/Ca2+ exchanger (NCX) couples Ca2+ efflux to Na+ influx, while the Na+/H+ exchanger (NHE) regulates intracellular pH by extruding protons in exchange for Na+ uptake. Similarly, the Na+/K+/2Cl cotransporter (NKCC) facilitates the coupled transport of Na+, K+, and Cl, thereby contributing to cell volume regulation and intracellular chloride homeostasis.
Ion transporters play fundamental roles in intracellular ion balance, pH regulation, volume control, and signal transduction. Dysregulation of their expression or activity has been widely implicated in malignancies. For instance, Na+/K+-ATPase is upregulated in gastric and breast cancers, where enhanced pump activity promotes invasion, tumor progression, and poor prognosis through activation of oncogenic PI3K/AKT signaling [236,237]. Ca2+-ATPases comprise three subtypes based on subcellular localization: plasma membrane Ca2+-ATPases (PMCAs), ER/sarcoplasmic reticulum Ca2+-ATPases (SERCAs), and Golgi/secretory pathway Ca2+-ATPases (SPCAs). PMCA2 is markedly overexpressed in certain breast cancer cell lines [238], whereas downregulation of PMCA4 and PMCA1 has been reported in colon and oral squamous cell carcinoma, respectively, resulting in elevated cytosolic Ca2+ and enhanced proliferation [239,240]. Among SERCAs, isoform-specific dysregulation has been associated with cancer [241]: SERCA2 is overexpressed in colorectal cancer and promotes proliferation and migration [242], while SERCA3 expression declines during colon tumor progression despite its initial elevation during differentiation [243]. In breast cancer, SPCA1 is preferentially expressed in basal-like subtypes but reduced in luminal tumors [244]. SPCA2 is upregulated in breast cancers, where knockdown suppresses proliferation and xenograft tumor growth, while overexpression enhances proliferation and activates ERK1/2 signaling through constitutive Ca2+ entry [245].
Although studies on NCX in cancer remain limited, elevated NCX3 transcript levels have been observed in therapy-resistant ovarian carcinoma and medulloblastoma cells compared with sensitive counterparts [88,89,246]. NHE1, the predominant isoform mediating proton efflux, is consistently upregulated in diverse malignancies. Hyperactive NHE1 promotes malignant progression by maintaining intracellular alkalinization, generating an acidic tumor microenvironment, regulating cell volume, altering morphology, and enhancing migration. Elevated NHE1 expression has been documented in gliomas [247], and it contributes to 5-fluorouracil resistance in gastric cancer cells via JAK/STAT3 signaling [248]. NKCC1 is similarly dysregulated in cancer, with overexpression enhancing proliferation and invasion in HCC [249]. In gastric cancer, NKCC1 promotes malignant progression by activating the MAPK–JNK pathway and inducing EMT [250].

3. Ion Channels as Pharmacological Targets in Cancer

The recognition of ion channels as critical regulators of oncogenesis has stimulated extensive research into their therapeutic potential. Two key questions now dominate the field: Is targeting ion channels a viable strategy to mechanistically disrupt oncogenic processes? And do these targets hold genuine promise for successful clinical translation into novel therapeutics?

3.1. Na+ Channel Inhibitors

Several pharmacological classes—including local anesthetics, antiarrhythmic agents, and antiepileptic drugs—exert their therapeutic effects primarily through inhibition of Na+ channels. Preclinical studies have shown that agents such as lidocaine, ropivacaine, phenytoin, and carbamazepine significantly modulate cancer cell behavior. Lidocaine, one of the most widely used local anesthetics in clinical practice, primarily functions by blocking VGSCs [251]. Recent studies demonstrate that lidocaine suppresses ovarian cancer metastasis via inhibition of NaV1.5-mediated EMT and reduces pulmonary metastasis in a murine breast cancer surgical model [18,20]. Similarly, phenytoin has been shown to decrease migration and invasion in metastatic breast cancer by selectively targeting NaV1.5 [21]. Ropivacaine enhances sorafenib-induced apoptosis in HCC cells through suppression of the IL-6/STAT3 signaling pathway [22].
In addition to these conventional drugs, peptide toxins and monoclonal antibodies targeting specific VGSC isoforms represent innovative strategies in oncology. For example, the spider venom-derived toxin JZTX-14 selectively blocks NaV1.5, thereby inhibiting breast cancer cell proliferation and migration [23]. Likewise, the peptide toxin HNTX-III, which targets NaV1.7, regulates migration and invasion in rat prostate cancer cells [25]. Furthermore, a polyclonal antibody against the neonatal splice variant of NaV1.5 significantly reduces the invasive capacity of MDA-MB-231 breast cancer cells in a dose-dependent manner, underscoring its therapeutic potential in breast cancer [24].
Considerable attention has also been directed toward the synergistic potential of Na+ channel modulators with established anticancer therapies. For instance, Sui et al. reported that the NaV1.5 activator veratridine increased the chemosensitivity of colorectal cancer cells to 5-fluorouracil [252]. Lidocaine enhances the sensitivity of ovarian cancer cells to carboplatin and paclitaxel [18,253] and restores temozolomide (TMZ) responsiveness in glioblastoma by inhibiting the HGF/MET pathway [19]. In the context of radiotherapy, the use of antiepileptic drugs has been associated with improved prognosis in breast cancer patients with brain metastases undergoing whole-brain radiotherapy [254].
The impact of Na+ channel inhibitors on antitumor immunity remains incompletely understood; however, emerging evidence suggests that certain local anesthetics may exert immunomodulatory effects. Preclinical studies have demonstrated that lidocaine and ropivacaine possess significant immune-regulatory properties. In vitro experiments using clinically relevant concentrations of lidocaine revealed enhanced natural killer (NK) cell activity [255]. Additionally, ropivacaine has been shown to upregulate MHC-I expression on tumor cells, thereby facilitating recognition and elimination by cytotoxic T lymphocytes within the TME [256].
Clinical investigations evaluating Na+ channel modulators are currently underway in various malignancies. An open-label, multicenter, randomized trial reported that peritumoral lidocaine injection prior to surgery significantly improved both disease-free survival (DFS) and overall survival (OS) in patients with early breast cancer (NCT01916317) [95]. In colorectal cancer, a prospective study of 150 surgical patients demonstrated that those receiving sevoflurane anesthesia combined with a 48 h lidocaine infusion exhibited a significantly lower one-year recurrence rate compared with patients receiving sevoflurane alone (NCT02786329) [15]. Furthermore, riluzole, another VGSC inhibitor, demonstrated favorable tolerability in a phase I trial investigating its combination with mFOLFOX6 and bevacizumab in metastatic colorectal cancer patients (NCT04761614) [96].

3.2. Ca2+ Channel Inhibitors

Calcium channel blockers (CCBs), a class of agents that inhibit Ca2+ influx through voltage-gated or other calcium channels, are widely used in the management of cardiovascular dysfunction, psychiatric disorders, and neurodegenerative diseases. In recent years, their potential applications in oncology have attracted increasing attention.
Commercially available antitumor agents targeting Ca2+ channels have primarily focused on VGCCs. Several FDA-approved CCBs, including lacidipine, manidipine, lomerizine, and benidipine, have demonstrated efficacy in inhibiting stem cell-like properties and inducing apoptosis in ovarian CSCs [257]. Bepridil, commonly prescribed for arrhythmia and heart failure, suppresses EMT-associated gene expression in ovarian cancer and exhibits antitumor activity in chronic lymphocytic leukemia through attenuation of NOTCH1 activation [26,27]. Diltiazem has been shown to modulate EMT and inhibit metastasis in triple-negative breast cancer (TNBC) mouse models [28]. In addition, the T-type CCB KTt-45 exerts anticancer effects by inducing mitochondrial-dependent apoptosis in HeLa cells [29].
Blockade of TRP channels also demonstrates substantial therapeutic promise. For example, inhibition of TRPC6 by BI-749327 significantly reduces mammosphere formation in breast CSCs [30]. Both natural compounds and synthetic analogues targeting TRP channels have exhibited anticancer properties across diverse tumor types. Cannabinoids, acting as TRP channel blockers, induce tumor cell death, inhibit angiogenesis and invasion, and modulate antitumor immunity [258]. In orthotopic HCC xenograft models, cannabinoids inhibited tumor growth through AMPK activation and regulation of autophagy [31]. In lung cancer, cannabinoids reduced invasion by modulating ICAM-1 and TIMP-1 expression [32]. However, it is noteworthy that delta-9-tetrahydrocannabinol has been reported to promote breast cancer growth and metastasis by impairing antitumor immunity, mediated via CB2 receptor-driven Th1/Th2 imbalance [259]. Monoclonal antibodies targeting Ca2+ channels represent another emerging therapeutic approach. For instance, mAb82, directed against TRPV6, induces apoptosis and suppresses tumor growth in prostate cancer models [33].
Moreover, specific CCBs have demonstrated efficacy in enhancing sensitivity to, or reversing resistance against, anticancer therapies. A recent study reported that the L-type CCB lacidipine potentiates the effects of doxorubicin and cisplatin in TNBC by selectively targeting Ca2+ channels to inhibit the Pyk2-JAK1–calmodulin complex-mediated IDO1 immunosuppressive pathway [34]. Similarly, lercanidipine and amlodipine inhibit the YY1/ERK/TGF-β axis, significantly sensitizing gastric cancer cells to doxorubicin [35]. The T-type CCB mibefradil (MIB) synergizes with carboplatin, promoting apoptosis and restoring platinum sensitivity in resistant ovarian tumors in murine xenograft models [36]. In gliomas, MIB induces glioblastoma cell apoptosis while suppressing glioblastoma stem-like cell phenotypes, thereby enhancing chemosensitivity to temozolomide [37,260]. Pharmacological inhibition of SOCE with SKF96365 augments 5-FU chemosensitivity in HCC through autophagic mechanisms [38]. In small-cell lung cancer patient-derived xenograft (PDX) models, the α2δ1 subunit-specific monoclonal antibody 1B50-1 enhanced the efficacy of chemotherapy and delayed relapse, highlighting its therapeutic potential in both combination and sequential regimens [39]. TRPV1 has also been implicated in chemoresistance: it promotes cisplatin resistance in cervical cancer via the Ca2+–AMPK pathway, whereas pharmacological inhibition with the antagonist AMG9810 reverses cisplatin resistance [40]. Similarly, pharmacological inhibition of TRPM2 with compound D9 synergizes with osimertinib, a third-generation EGFR-TKI, effectively inducing apoptosis and reducing viability in osimertinib-resistant lung cancer cells [41]. In addition to chemosensitization, CCBs also augment antitumor immune responses. Wu et al. demonstrated that nifedipine suppresses colorectal cancer immune evasion via NFAT2-dependent mechanisms and further enhances the efficacy of PD-1 blockade in preclinical murine models [42].
Encouraged by robust preclinical evidence, several clinical trials have been initiated to evaluate CCBs as novel oncological agents. A phase I trial (NCT01480050) assessed the safety and tolerability of MIB administered sequentially with TMZ in patients with recurrent high-grade gliomas (rHGGs). At the maximum tolerated dose (MTD), this regimen was well tolerated, demonstrating a favorable safety profile [16]. Another phase I trial (NCT02202993) conducted at Yale University investigated the combination of mibefradil dihydrochloride with concurrent hypofractionated radiotherapy in progressive or recurrent glioblastoma, although results have not yet been published. An open-label, dose-escalation phase I trial (NCT01578564) evaluating the TRPV6 inhibitor SOR-C13 demonstrated acceptable safety and tolerability in patients with advanced epithelial-derived solid tumors [97]. The ARISTOCRAT trial (NCT05629702), a multicenter, phase II, double-blind, placebo-controlled randomized study, is currently assessing the efficacy of cannabinoids combined with TMZ in recurrent glioblastoma [98]. Furthermore, a monoclonal antibody targeting a variant of P2X7 (nfP2X7) completed phase I trials in basal cell carcinoma (NCT02587819), demonstrating promising efficacy, with 65% of participants exhibiting measurable reductions in lesion area following 28 days of daily administration [99].
Carboxyamidotriazole (CAI), an inhibitor of non-voltage-gated Ca2+ channels and Ca2+-mediated signaling pathways, has also demonstrated potent anticancer activity by suppressing angiogenesis, tumor growth, and metastasis. Although CAI promotes IFN-γ secretion by T cells, it paradoxically enhances immune evasion through the IDO1-Kyn-AhR axis. Notably, combining CAI with IDO1 or AhR antagonists reduces PD-1 expression, improves CD8+ T-cell infiltration and effector function, suppresses tumor growth, and prolongs survival in murine models [261]. Several clinical trials have evaluated CAI’s safety and efficacy [17,100,104]. A phase II trial in relapsed epithelial ovarian cancer indicated its potential for disease stabilization [104], while a phase III randomized controlled trial in advanced NSCLC (CTR20160395) demonstrated that CAI combined with platinum-based chemotherapy improved progression-free survival, supporting its therapeutic potential [101].

3.3. K+ Channel Inhibitors

K+ channels represent promising therapeutic targets in oncology. Several pharmacological agents targeting distinct K+ channel subfamilies, including voltage-gated (KV), calcium-activated (KCa), two-pore domain (K2P), and inwardly rectifying (Kir) channels, have demonstrated antitumor activity in both in vitro and in vivo models.
Among KV channels, KV1.3, KV10.1, and KV11.1 have emerged as prominent therapeutic candidates. FS48 from Xenopsylla cheopis salivary glands inhibits migration and invasion of human lung cancer cells by blocking KV1.3 [43]. KAaH1 from scorpion venom suppresses migration and adhesion of glioblastoma cells by targeting KV1.1 and KV1.3 [44]. Psoralen derivatives induce tumor cell death by inhibiting mitochondrial KV1.3 expression and inducing mitochondrial dysfunction [45]. Memantine triggers acute lymphoblastic leukemia (ALL) cell death via KV1.3 inhibition, mechanistically linked to suppression of AKT and ERK1/2 signaling [47]. Small-molecule inhibitors such as Psora-4, PAP-1, and clofazimine reduce proliferation and induce apoptosis via Kv1.3 channel blockade [262]. Two PAP-1 derivatives, PAPTP and PCARBTP, enhance the efficacy of cisplatin in murine melanoma models [45]. Other Kv1.3 inhibitors, including the plant-derived polycyclic compound genistein, exert antiproliferative and pro-apoptotic effects in cancer cells [48,49]. KV10.1 has also been extensively investigated as an oncogenic driver. Multiple agents, including Chloroquine, Astemizole, Procyanidin B1, Tetrandrine, and Nutlin-3, selectively target KV10.1 to suppress malignant proliferation and migration [50,51,52,53,54]. Similarly, agents such as celastrol, berberine, resveratrol, and clarithromycin modulate cancer cell fate via KV11.1-dependent mechanisms [55,56,57,58].
Targeting KCa channels is also an emerging anticancer strategy. The BKCa blocker Paxilline prevents hypoxia-induced migration in glioblastoma cells [59] and, together with Iberiotoxin, regulates cell cycle and proliferation in neuroblastoma cells [60]. Iberiotoxin impairs proliferation and migration in endometrial cancer cells [61] and inhibits remote colonization of HCC cells in xenograft models [62]. Tetraethylammonium (nonselective BKCa blocker) induces G2 phase arrest and reduces viability of HCC cells under hypoxic conditions, while Penitrem A (selective BKCa blocker) induces G1 phase arrest and exerts antiproliferative effects in breast cancer [62,63]. Trimebutine maleate downregulates stemness-related protein expression and slows the growth of ovarian CSCs [64]. Pharmacological or genetic blockade of IKCa and SKCa channels effectively curbs tumor progression. Temozolomide and Vigabatrin decrease IKCa current amplitude in glioma cells, potentially contributing to their antineoplastic effects [65,66]. Clotrimazole and Triarylmethane-34 inhibit proliferation of cervical cancer and intrahepatic cholangiocarcinoma cells, respectively [67,68]. Piperine antagonizes IKCa channels, triggering dose-dependent G0/G1 phase arrest, proliferation suppression, and apoptosis in prostate cancer cells [69]. Blocking KCa3.1 with TRAM-34 or Senicapoc slows glioma cell growth [263]. Similarly, Miconazole targeting SKCa inhibits melanoma cell growth in a dose-dependent manner [70].
The TREK-1 (K2P2.1) activator BL1249 inhibits proliferation and migration of PDAC cells [71]. For Kir channels, Withaferin A suppresses breast cancer cell proliferation by inhibiting TASK-3 (K2P9.1) channel function [72]. The ATP-sensitive potassium channel (KATP) is a Kir subfamily member modulated by intracellular ATP. Minoxidil triggers mitochondrial disruption and extensive DNA damage in ovarian cancer by activating Kir6.2 [73], and Glibenclamide acts as a blocker of KATP channels decreases proliferation and migration in endometrial adenocarcinoma cells [74].
Pharmacological modulation of K+ channels can enhance the efficacy of conventional anticancer agents by synergistically inducing cell death and suppressing therapy resistance. Memantine synergistically enhances cytarabine-induced cell death in acute leukemia models via KV1.3 inhibition [47]. Inhibition of mitochondrial KV1.3 channels using PCARBTP and PAPTP potently suppress cancer cell viability in vitro and synergize with Gemcitabine and Abraxane to amplify antitumor efficacy in vivo [46]. Combining BKCa blocker Penitrem A with HER-targeted agents induces synthetic lethality in breast cancer by reducing EGFR, HER2, and AKT/STAT3 activation [63]. The selective KCa1.1 blocker paxilline reverses multidrug resistance in 3D-cultured sarcoma and prostate cancer spheroid models [264,265].
K+ channels are also emerging as promising immunotherapeutic targets. Blockade of Kir2.1 with ML133 has been shown to repolarize TAMs toward an antitumor phenotype, thereby suppressing tumor growth in both murine and PDX models [266]. KCa3.1 (KCNN4) has been identified as a robust prognostic biomarker and predictor of immunotherapy response across multiple cancer types [267]. Mechanistically, KCa3.1 activators suppress IL-10 transcription in T-cell lymphoma cells and attenuate IL-10-mediated immune evasion by reducing TAM-derived IL-10 production [268,269]. Preclinical studies further demonstrate that pharmacological blockade of KCa3.1 with TRAM-34 enhances the cytotoxic activity of CAR T cells against target tumor cells [270].

3.4. Cl Channel Inhibitors

Recent advances in elucidating the roles of Cl channels across malignancies have catalyzed the evaluation of these channels as druggable targets for precision oncology. Tamoxifen and 5-Nitro-2-[3-phenylpropylamino]benzoic acid (NPPB), non-specific Cl channel inhibitors, suppress ClC-3 currents and cancer cell proliferation. However, their concurrent inhibition of channels such as TMEM16A/ANO1 precludes clinical utility as selective ClC-3 blockers [75,271]. Similarly, while 4,4′-diisothiocyanatostilbene-2,2•-disulfonic acid disodium salt hydrate (DIDS) enhances hyperthermia-mediated tumor suppression via Cl channel inhibition, its broad-spectrum activity raises concerns regarding off-target effects, necessitating further validation of its therapeutic applicability in oncology [272]. Chlorotoxin (ClTx), an animal venom-derived peptide initially characterized as a Cl channel blocker, specifically interacts with MMP-2 and exerts anti-invasive effect in glioma [76,273]. Qin et al. further demonstrated that ClTx-modified liposomes activate the ClC-3 channel via MMP-2 interaction, subsequently suppressing Cl currents and cell migration in gliomas [274]. Although ClC-3-specific antibodies have been proven to inhibit Cl currents [275], their antitumor efficacy requires further investigation. Conversely, ClC-3 activators like bufadienolides exert antitumor activity by activating ClC-3 channel, leading to suppression of the PI3K/Akt/mTOR signaling axis [276].
The ANO1 (TMEM16A) channel is inhibited by CaCCinh-A01, which impedes breast cancer progression [77]. Natural compounds such as schisandrathera D and vitekwangin B reduce cell viability and induce apoptosis in cancers, such as prostate cancer, by downregulating ANO1 protein expression [78,79,80]. TMEM16A-targeting natural products show broad antitumor potential in pulmonary malignancies [277], with matairesinoside and daidzein exhibiting potent inhibitory activity against lung cancer progression [81,82]. Phytochemicals like narirutin and allicin demonstrate synergistic effects with cisplatin, enhancing chemotherapeutic efficacy while reducing toxicity [278,279]. Significantly, zafirlukast, an FDA-approved asthma therapeutic, suppresses lung adenocarcinoma growth through targeted TMEM16A inhibition, revealing novel therapeutic applications for existing pharmacological agents [83]. Benzodiazepines exhibit context-dependent effects in oncology. The GABAA-R agonist lorazepam (LOR) enhances GABA-mediated signaling, promoting glioma proliferation and growth while shortening survival in diffuse midline glioma patient-derived models [280]. A retrospective cohort study in pancreatic cancer patients linked LOR to stimulated IL-6 secretion by cancer-associated fibroblasts (CAFs) and decreased survival [281]. Similarly, diazepam use in NSCLC patients correlates with poorer responses to chemoimmunotherapy compared to non-users [282]. Conversely, in a murine melanoma model, benzodiazepines synergized with radiotherapy and immune checkpoint inhibitors, enhancing overall antitumor efficacy [283]. This evidence indicates that benzodiazepines manifest dual regulatory roles in cancer. Subsequent studies should delineate their context-specific molecular mechanisms to refine potential clinical applications.

3.5. Trace Metal Ion Channel Inhibitors

Trace metal ion channels play essential roles in regulating cellular physiology by maintaining trace ion homeostasis, and their dysregulation has been implicated as a critical driver of tumor progression. Accordingly, targeting these channels has emerged as a promising therapeutic strategy. TRPM7 represents one of the most extensively studied candidates. Its unique bifunctional structure, encompassing both ion channel and kinase domains, provides multiple therapeutic entry points. Pharmacological inhibitors such as TG100-115 and waixenicin A, as well as gene-editing approaches, have demonstrated robust preclinical efficacy in suppressing tumor growth and metastasis [229,284,285,286]. Therapeutic targeting of TRPML1 requires consideration of tumor-specific signaling contexts. TRPML1 agonists, including ML-SA5 and MK6-83, selectively induce cancer cell death in breast, gastric, melanoma, and glioma models by triggering autophagic arrest, while sparing normal cells [234]. Conversely, TRPML1 inhibitors such as ML-SI1 promote ferroptosis in breast CSCs and enhance the chemosensitivity of breast cancer cells to doxorubicin [84].

3.6. Ion Transporter Inhibitors

Ion transporters play essential roles in maintaining ion balance and pH homeostasis, positioning them as attractive targets for pharmacological intervention. Cardiac glycosides such as ouabain and digoxin, which potently inhibit Na+/K+ ATPase, have demonstrated broad-spectrum antitumor activity. These agents suppress cancer cell proliferation, migration, and invasion, and can also induce tumor cell lysis [85,86,237,287]. Furthermore, their clinical use has been associated with reduced cancer risk and improved survival outcomes in certain patient populations [288,289].
Elevated cytoplasmic Ca2+ concentrations exert cytotoxic effects by activating cell death pathways. Inhibition of Ca2+-ATPases induces sustained increases in cytosolic Ca2+, ultimately leading to apoptosis or necrosis. For example, the selective PMCA inhibitor [Pt(O,O′-acac)(γ-acac)(DMS)] rapidly triggers apoptosis in breast cancer cells in association with cytosolic Ca2+ accumulation [87]. Thapsigargin (TG), a specific SERCA inhibitor, disrupts Ca2+ uptake into the endoplasmic reticulum, resulting in endoplasmic reticulum calcium store depletion [290]. Although TG has been extensively investigated as a chemotherapeutic candidate for prostate and other cancers, its nonselective toxicity has limited clinical translation. To improve tumor specificity, the prodrug G202 (Mipsagargin) was developed by conjugating a TG analogue to a peptide targeting prostate-specific membrane antigen (PSMA). This design restricts uptake to PSMA-expressing cells, thereby conferring tumor selectivity [291]. G202 has shown potent antitumor activity in models of prostate, breast, and bladder cancer with minimal host toxicity. A phase I clinical trial in patients with advanced solid tumors demonstrated its acceptable tolerability and favorable pharmacokinetic profile [102], while subsequent phase II studies suggested promising efficacy in hepatocellular carcinoma [103].
Research on NCX in relation to cancer remains limited. The NCX inhibitor KB-R7943 has been shown to enhance cisplatin-induced cell death in therapy-resistant ovarian carcinoma cells, without affecting sensitive counterparts, and to significantly increase radiation-induced apoptosis in medulloblastoma cells [88,89].
NHE also represent attractive therapeutic targets, with inhibitors exerting antitumor effects primarily through disruption of pH homeostasis. Representative agents include the non-specific inhibitor amiloride and the more selective derivative cariporide. Amiloride sensitizes prostate cancer cells to the tyrosine kinase inhibitor lapatinib by altering ERBB3 subcellular localization [90], and has also demonstrated antimyeloma activity [91]. Cariporide significantly enhances doxorubicin sensitivity in breast cancer cells both in vitro and in vivo [92]. Similarly, the Na+/K+/2Cl cotransporter 1 (NKCC1) has emerged as a therapeutic target. The NKCC1-specific inhibitor bumetanide exerts anti-angiogenic effects in preclinical colon cancer models and significantly suppresses the migration of metastatic glioma cells [93,94].

4. Limitations and Challenges

Ion channels represent a promising and rapidly expanding class of anticancer drug targets. Nevertheless, their therapeutic exploitation is still constrained by several important limitations and challenges.
First, achieving adequate drug selectivity and specificity remains a major obstacle. Ion channels are ubiquitously expressed throughout the human body, particularly in excitable tissues such as the heart and nervous system, where they mediate essential physiological functions. Nonselective inhibition therefore carries a substantial risk of off-target toxicity, including arrhythmia, hypotension, and neurotoxicity. Although efforts have increasingly focused on developing subtype-selective modulators, the high degree of structural conservation within ion channel families makes it difficult to selectively target tumor-associated subtypes without affecting their physiological counterparts.
Second, limitations inherent to current research models substantially hinder progress. Conventional in vitro systems, such as two-dimensional cell cultures, fail to recapitulate the complexity of the tumor microenvironment, including factors such as extracellular pH, hypoxia, and cell–cell interactions, all of which profoundly influence ion channel function. Consequently, compounds that appear effective in simplified models often fail in vivo. Furthermore, the absence of robust preclinical models that faithfully reproduce human ion channel characteristics and tumor–host interactions limits the predictive value of efficacy and toxicity assessments. Even in vivo, tumor regulation frequently involves multiple cooperating ion channels; thus, mono-channel inhibition may be offset by compensatory mechanisms. Adding to this complexity, certain ion channels display divergent, or even opposing, functions depending on tumor type, stage, or microenvironmental context. Spatiotemporal heterogeneity in ion concentrations (e.g., H+ and Ca2+), oxidative stress, and extracellular matrix composition within the TME further complicates drug response.
Third, the path toward clinical translation presents additional difficulties. Most ion channel-targeting drugs currently available are broad-spectrum, with limited capacity to distinguish between malignant and normal cells. This lack of specificity has led to the termination of many candidates in early-phase development due to unacceptable toxicity. Mechanisms of tumor resistance to ion channel inhibition—whether genetic, epigenetic, or adaptive—remain poorly understood, while the absence of reliable biomarkers impedes patient stratification and precision therapy. Moreover, the potential synergistic effects between ion channel inhibitors and existing treatment modalities (chemotherapy, targeted therapy, and immunotherapy) remain inadequately explored, and evidence supporting rational combination strategies is limited. Although several ion channel inhibitors have advanced to early-phase clinical trials, most studies are small, single-arm investigations with limited controls, thereby providing low-quality evidence. Finally, critical barriers related to drug distribution, metabolism, and tumor-specific delivery continue to restrict therapeutic translation.

5. Conclusions and Future Perspectives

Malignant tumors frequently exhibit dysregulated expression and functional alterations of ion channels. These aberrations directly contribute to oncogenic processes, including proliferation, migration, invasion, apoptosis, and therapeutic resistance, by modulating membrane potential, intracellular ion homeostasis, and downstream signaling cascades. Moreover, ion channels play critical roles in regulating immune cell activation and cytotoxic function, underscoring their importance in shaping antitumor immunity and mediating tumor–immune cell interactions. Collectively, these insights provide a strong scientific rationale for targeting ion channels as a therapeutic strategy in oncology.
Given their central role in tumor initiation and progression, ion channels have emerged as highly attractive candidates for anticancer drug development. Current strategies encompass several complementary approaches: (i) repurposing existing ion channel modulators (such as calcium channel blockers, local anesthetics, and selected natural compounds) for their antitumor properties; (ii) designing subtype-selective agents, including monoclonal antibodies and advanced drug delivery systems, to enhance specificity; and (iii) exploring rational combination therapies that integrate ion channel inhibitors with chemotherapy, targeted agents, or immunotherapies to achieve synergistic efficacy. These approaches collectively hold considerable promise for advancing precision oncology.
Addressing the current challenges in this field will require multidisciplinary collaboration and innovative methodologies. Structure-based drug design, informed by high-resolution datasets from resources such as UniProt and RCSB-PDB, will be critical for developing selective inhibitors against specific ion channel isoforms. Comparative structural analyses of conserved versus divergent domains, including pore regions, voltage-sensing modules, and ligand-binding sites, may help to discriminate tumor-associated subtypes and reduce off-target toxicity. Improved preclinical models, including tumor organoids and three-dimensional co-culture systems, can better recapitulate the tumor microenvironment and ion channel dynamics, thereby improving translational predictability. Additionally, artificial intelligence-driven tools offer new opportunities to integrate structural and functional data, predict binding affinities, and optimize subtype selectivity, thereby accelerating the discovery of safer and more effective ligands.
Future research should prioritize elucidating the structural determinants of ion channel function in cancer, developing subtype-specific modulators, and designing advanced delivery platforms to achieve tumor-selective targeting. By strengthening the interface between basic mechanistic research and clinical translation, ion channels may evolve from potential drug targets into clinically accessible therapeutic tools, ultimately expanding the landscape of precision oncology and providing novel treatment options for cancer patients.

Author Contributions

Conceptualization, S.Z., X.S., W.Z. and D.C.; writing—original draft preparation, S.Z., X.S. and D.C.; writing—review and editing, S.Z., X.S., W.Z. and D.C.; visualization, S.Z. and X.S.; supervision, D.C.; project administration, W.Z. and D.C.; funding acquisition, S.Z. and D.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by grants from the National Natural Science Foundation of China (No. 82203321) and the Basic and Applied Basic Research Foundation of Guangdong Province (2024A1515011643).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable.

Acknowledgments

Graphic Abstract was created in BioRender. Quan, t. (2025) https://BioRender.com/j4o0rtq. Figure 1 was created in BioRender. Quan, t. (2025) https://BioRender.com/i4fjbpw. The authors have reviewed and edited the output and take full responsibility for the content of this publication.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
ASICsAcid-sensing ion channels
BKCaLarge-conductance calcium-activated potassium channel
CaCCsCalcium-activated chloride channels
CAFCancer-associated fibroblast
CCBsCalcium channel blockers
CFCystic fibrosis
CFTRCystic fibrosis transmembrane conductance regulator
CSCCancer stem cell
ECMExtracellular matrix
EMTEpithelial-mesenchymal transition
ENaCsEpithelial Na+ channels
GABAA-RsGamma-aminobutyric acid type A receptors
GlyRsGlycine receptors
HCCHepatocellular carcinoma
HNSCCHead and neck squamous cell carcinoma
IKCaIntermediate-conductance calcium-activated potassium channel
K2PTwo-pore domain potassium channels
KCaCalcium-activated potassium channels
KirInwardly rectifying potassium channels
KVVoltage-gated potassium channels
LGCCsLigand-gated calcium channels
LGClCsLigand-gated chloride channels
NPCNasopharyngeal carcinoma
NSCLCNon-small-cell lung cancer
OSCCOral squamous cell carcinoma
PDACPancreatic ductal adenocarcinoma
SCLCSmall-cell lung cancer
SKCaSmall-conductance calcium-activated potassium channel
SOCCsStore-operated calcium channels
SOCEStore-operated calcium entry
TAMsTumor-associated macrophages
TMETumor microenvironment
TMsTransmembrane domains
TMZTemozolomide
TNBCTriple-negative breast cancer
TRPTransient receptor potential
TTXTetrodotoxin
VGCCsVoltage-gated calcium channels
VGSCVoltage-gated sodium channel
VRACsVolume-regulated anion channels

References

  1. Bray, F.; Laversanne, M.; Sung, H.; Ferlay, J.; Siegel, R.L.; Soerjomataram, I.; Jemal, A. Global cancer statistics 2022: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA Cancer J. Clin. 2024, 74, 229–263. [Google Scholar] [CrossRef] [PubMed]
  2. Liu, S.; Jiang, A.; Tang, F.; Duan, M.; Li, B. Drug-induced tolerant persisters in tumor: Mechanism, vulnerability and perspective implication for clinical treatment. Mol. Cancer 2025, 24, 150. [Google Scholar] [CrossRef] [PubMed]
  3. Zhang, M.; Liu, C.; Tu, J.; Tang, M.; Ashrafizadeh, M.; Nabavi, N.; Sethi, G.; Zhao, P.; Liu, S. Advances in cancer immunotherapy: Historical perspectives, current developments, and future directions. Mol. Cancer 2025, 24, 136. [Google Scholar] [CrossRef]
  4. Du, B.; Qin, J.; Lin, B.; Zhang, J.; Li, D.; Liu, M. CAR-T therapy in solid tumors. Cancer Cell 2025, 43, 665–679. [Google Scholar] [CrossRef]
  5. Pauken, K.E.; Alhalabi, O.; Goswami, S.; Sharma, P. Neoadjuvant immune checkpoint therapy: Enabling insights into fundamental human immunology and clinical benefit. Cancer Cell 2025, 43, 623–640. [Google Scholar] [CrossRef]
  6. Gadsby, D.C. Ion channels versus ion pumps: The principal difference, in principle. Nat. Rev. Mol. Cell Biol. 2009, 10, 344–352. [Google Scholar] [CrossRef]
  7. Leslie, T.K.; James, A.D.; Zaccagna, F.; Grist, J.T.; Deen, S.; Kennerley, A.; Riemer, F.; Kaggie, J.D.; Gallagher, F.A.; Gilbert, F.J.; et al. Sodium homeostasis in the tumour microenvironment. Biochim. Biophys. Acta (BBA) Rev. Cancer 2019, 1872, 188304. [Google Scholar] [CrossRef] [PubMed]
  8. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion Channels in Cancer: Are Cancer Hallmarks Oncochannelopathies? Physiol. Rev. 2018, 98, 559–621. [Google Scholar] [CrossRef] [PubMed]
  9. Sanghvi, S.; Sridharan, D.; Evans, P.; Dougherty, J.; Szteyn, K.; Gabrilovich, D.; Dyta, M.; Weist, J.; Pierre, S.V.; Gururaja Rao, S.; et al. Functional large-conductance calcium and voltage-gated potassium channels in extracellular vesicles act as gatekeepers of structural and functional integrity. Nat. Commun. 2025, 16, 42. [Google Scholar] [CrossRef] [PubMed]
  10. Alexander, S.P.H.; Mathie, A.A.; Peters, J.A.; Veale, E.L.; Striessnig, J.; Kelly, E.; Armstrong, J.F.; Faccenda, E.; Harding, S.D.; Davies, J.A.; et al. The Concise Guide to PHARMACOLOGY 2023/24: Ion channels. Br. J. Pharmacol. 2023, 180 (Suppl. 2), S145–S222. [Google Scholar] [CrossRef] [PubMed]
  11. Huang, J.; Pan, X.; Yan, N. Structural biology and molecular pharmacology of voltage-gated ion channels. Nat. Rev. Mol. Cell Biol. 2024, 25, 904–925. [Google Scholar] [CrossRef] [PubMed]
  12. Shi, Q.; Yang, Z.; Yang, H.; Xu, L.; Xia, J.; Gu, J.; Chen, M.; Wang, Y.; Zhao, X.; Liao, Z.; et al. Targeting ion channels: Innovative approaches to combat cancer drug resistance. Theranostics 2025, 15, 521–545. [Google Scholar] [CrossRef]
  13. Capatina, A.L.; Lagos, D.; Brackenbury, W.J. Targeting Ion Channels for Cancer Treatment: Current Progress and Future Challenges. Rev. Physiol. Biochem. Pharmacol. 2022, 183, 1–43. [Google Scholar] [PubMed]
  14. Ru, Q.; Li, Y.; Chen, L.; Wu, Y.; Min, J.; Wang, F. Iron homeostasis and ferroptosis in human diseases: Mechanisms and therapeutic prospects. Signal Transduct. Target. Ther. 2024, 9, 271. [Google Scholar] [CrossRef] [PubMed]
  15. Alexa, A.L.; Ciocan, A.; Zaharie, F.; Valean, D.; Sargarovschi, S.; Breazu, C.; Al Hajjar, N.; Ionescu, D. The Influence of Intravenous Lidocaine Infusion on Postoperative Outcome and Neutrophil-to-Lymphocyte Ratio in Colorectal Cancer Patients. A Pilot Study. J. Gastrointestin. Liver Dis. 2023, 32, 156–161. [Google Scholar] [CrossRef] [PubMed]
  16. Holdhoff, M.; Ye, X.; Supko, J.G.; Nabors, L.B.; Desai, A.S.; Walbert, T.; Lesser, G.J.; Read, W.L.; Lieberman, F.S.; Lodge, M.A.; et al. Timed sequential therapy of the selective T-type calcium channel blocker mibefradil and temozolomide in patients with recurrent high-grade gliomas. Neuro-Oncology 2017, 19, 845–852. [Google Scholar] [CrossRef]
  17. Kohn, E.C.; Figg, W.D.; Sarosy, G.A.; Bauer, K.S.; Davis, P.A.; Soltis, M.J.; Thompkins, A.; Liotta, L.A.; Reed, E. Phase I trial of micronized formulation carboxyamidotriazole in patients with refractory solid tumors: Pharmacokinetics, clinical outcome, and comparison of formulations. J. Clin. Oncol. 1997, 15, 1985–1993. [Google Scholar] [CrossRef] [PubMed]
  18. Liu, C.; Yu, M.; Li, Y.; Wang, H.; Xu, C.; Zhang, X.; Li, M.; Guo, H.; Ma, D.; Guo, X. Lidocaine inhibits the metastatic potential of ovarian cancer by blocking Nav1.5-mediated EMT and FAK/Paxillin signaling pathway. Cancer Med. 2021, 10, 337–349. [Google Scholar] [CrossRef]
  19. Chen, M.S.; Chong, Z.Y.; Huang, C.; Huang, H.C.; Su, P.H.; Chen, J.C. Lidocaine attenuates TMZ resistance and inhibits cell migration by modulating the MET pathway in glioblastoma cells. Oncol. Rep. 2024, 51, 72. [Google Scholar] [CrossRef]
  20. Freeman, J.; Crowley, P.D.; Foley, A.G.; Gallagher, H.C.; Iwasaki, M.; Ma, D.; Buggy, D.J. Effect of Perioperative Lidocaine, Propofol and Steroids on Pulmonary Metastasis in a Murine Model of Breast Cancer Surgery. Cancers 2019, 11, 613. [Google Scholar] [CrossRef] [PubMed]
  21. Yang, M.; Kozminski, D.J.; Wold, L.A.; Modak, R.; Calhoun, J.D.; Isom, L.L.; Brackenbury, W.J. Therapeutic potential for phenytoin: Targeting Nav1.5 sodium channels to reduce migration and invasion in metastatic breast cancer. Breast Cancer Res. Treat. 2012, 134, 603–615. [Google Scholar] [CrossRef]
  22. Wang, W.; Lin, H.; Liu, D.; Wang, T.; Zhu, Z.; Yu, P.; Zhang, J. Ropivacaine synergizes with sorafenib to induce apoptosis of hepatocellular carcinoma cells via the IL-6/STAT3 pathway. Cancer Sci. 2024, 115, 2923–2930. [Google Scholar] [CrossRef]
  23. Wu, W.; Yin, Y.; Feng, P.; Chen, G.; Pan, L.; Gu, P.; Zhou, S.; Lin, F.; Ji, S.; Zheng, C.; et al. Spider venom-derived peptide JZTX-14 prevents migration and invasion of breast cancer cells via inhibition of sodium channels. Front. Pharmacol. 2023, 14, 1067665. [Google Scholar] [CrossRef]
  24. Brackenbury, W.J.; Chioni, A.M.; Diss, J.K.; Djamgoz, M.B. The neonatal splice variant of Nav1.5 potentiates in vitro invasive behaviour of MDA-MB-231 human breast cancer cells. Breast Cancer Res. Treat. 2007, 101, 149–160. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, B.; Zhang, C.; Wang, Z.; Chen, Y.; Xie, H.; Li, S.; Liu, X.; Liu, Z.; Chen, P. Mechanistic insights into Nav1.7-dependent regulation of rat prostate cancer cell invasiveness revealed by toxin probes and proteomic analysis. FEBS J. 2019, 286, 2549–2561. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, S.; Kim, D.; Park, M.; Yin, J.H.; Park, J.; Chung, Y.J. Suppression of Metastatic Ovarian Cancer Cells by Bepridil, a Calcium Channel Blocker. Life 2023, 13, 1607. [Google Scholar] [CrossRef] [PubMed]
  27. Baldoni, S.; Del Papa, B.; Dorillo, E.; Aureli, P.; De Falco, F.; Rompietti, C.; Sorcini, D.; Varasano, E.; Cecchini, D.; Zei, T.; et al. Bepridil exhibits anti-leukemic activity associated with NOTCH1 pathway inhibition in chronic lymphocytic leukemia. Int. J. Cancer 2018, 143, 958–970. [Google Scholar] [CrossRef]
  28. Chen, Y.C.; Wu, C.T.; Chen, J.H.; Tsai, C.F.; Wu, C.Y.; Chang, P.C.; Yeh, W.L. Diltiazem inhibits breast cancer metastasis via mediating growth differentiation factor 15 and epithelial-mesenchymal transition. Oncogenesis 2022, 11, 48. [Google Scholar] [CrossRef] [PubMed]
  29. Du, N.H.; Ngoc, T.T.B.; Cang, H.Q.; Luyen, N.T.T.; Thuoc, T.L.; Le Quan, T.; Thao, D.T.P. KTt-45, a T-type calcium channel blocker, acts as an anticancer agent by inducing apoptosis on HeLa cervical cancer cell line. Sci. Rep. 2023, 13, 22092. [Google Scholar] [CrossRef] [PubMed]
  30. Mukhopadhyay, D.; Goel, H.L.; Xiong, C.; Goel, S.; Kumar, A.; Li, R.; Zhu, L.J.; Clark, J.L.; Brehm, M.A.; Mercurio, A.M. The calcium channel TRPC6 promotes chemotherapy-induced persistence by regulating integrin α6 mRNA splicing. Cell Rep. 2023, 42, 113347. [Google Scholar] [CrossRef]
  31. Vara, D.; Salazar, M.; Olea-Herrero, N.; Guzmán, M.; Velasco, G.; Díaz-Laviada, I. Anti-tumoral action of cannabinoids on hepatocellular carcinoma: Role of AMPK-dependent activation of autophagy. Cell Death Differ. 2011, 18, 1099–1111. [Google Scholar] [CrossRef]
  32. Ramer, R.; Bublitz, K.; Freimuth, N.; Merkord, J.; Rohde, H.; Haustein, M.; Borchert, P.; Schmuhl, E.; Linnebacher, M.; Hinz, B. Cannabidiol inhibits lung cancer cell invasion and metastasis via intercellular adhesion molecule-1. FASEB J. 2012, 26, 1535–1548. [Google Scholar] [CrossRef]
  33. Haustrate, A.; Cordier, C.; Shapovalov, G.; Mihalache, A.; Desruelles, E.; Soret, B.; Essonghé, N.C.; Spriet, C.; Yassine, M.; Barras, A.; et al. Trpv6 channel targeting using monoclonal antibody induces prostate cancer cell apoptosis and tumor regression. Cell Death Dis. 2024, 15, 419. [Google Scholar] [CrossRef] [PubMed]
  34. Sheng, Y.; Qiao, C.; Zhang, Z.; Shi, X.; Yang, L.; Xi, R.; Yu, J.; Liu, W.; Zhang, G.; Wang, F. Calcium Channel Blocker Lacidipine Promotes Antitumor Immunity by Reprogramming Tryptophan Metabolism. Adv. Sci. 2025, 12, e2409310. [Google Scholar] [CrossRef]
  35. Panneerpandian, P.; Rao, D.B.; Ganesan, K. Calcium channel blockers lercanidipine and amlodipine inhibit YY1/ERK/TGF-β mediated transcription and sensitize the gastric cancer cells to doxorubicin. Toxicol. In Vitro 2021, 74, 105152. [Google Scholar] [CrossRef] [PubMed]
  36. Dziegielewska, B.; Casarez, E.V.; Yang, W.Z.; Gray, L.S.; Dziegielewski, J.; Slack-Davis, J.K. T-Type Ca2+ Channel Inhibition Sensitizes Ovarian Cancer to Carboplatin. Mol. Cancer Ther. 2016, 15, 460–470. [Google Scholar] [CrossRef] [PubMed]
  37. Valerie, N.C.; Dziegielewska, B.; Hosing, A.S.; Augustin, E.; Gray, L.S.; Brautigan, D.L.; Larner, J.M.; Dziegielewski, J. Inhibition of T-type calcium channels disrupts Akt signaling and promotes apoptosis in glioblastoma cells. Biochem. Pharmacol. 2013, 85, 888–897. [Google Scholar] [CrossRef] [PubMed]
  38. Tang, B.D.; Xia, X.; Lv, X.F.; Yu, B.X.; Yuan, J.N.; Mai, X.Y.; Shang, J.Y.; Zhou, J.G.; Liang, S.J.; Pang, R.P. Inhibition of Orai1-mediated Ca2+ entry enhances chemosensitivity of HepG2 hepatocarcinoma cells to 5-fluorouracil. J. Cell Mol. Med. 2017, 21, 904–915. [Google Scholar] [CrossRef] [PubMed]
  39. Yu, J.; Wang, S.; Zhao, W.; Duan, J.; Wang, Z.; Chen, H.; Tian, Y.; Wang, D.; Zhao, J.; An, T.; et al. Mechanistic Exploration of Cancer Stem Cell Marker Voltage-Dependent Calcium Channel α2δ1 Subunit-mediated Chemotherapy Resistance in Small-Cell Lung Cancer. Clin. Cancer Res. 2018, 24, 2148–2158. [Google Scholar] [CrossRef]
  40. Oh, S.J.; Lim, J.Y.; Son, M.K.; Ahn, J.H.; Song, K.H.; Lee, H.J.; Kim, S.; Cho, E.H.; Chung, J.Y.; Cho, H.; et al. TRPV1 inhibition overcomes cisplatin resistance by blocking autophagy-mediated hyperactivation of EGFR signaling pathway. Nat. Commun. 2023, 14, 2691. [Google Scholar] [CrossRef]
  41. Chen, Z.; Vallega, K.A.; Boda, V.K.; Quan, Z.; Wang, D.; Fan, S.; Wang, Q.; Ramalingam, S.S.; Li, W.; Sun, S.Y. Targeting Transient Receptor Potential Melastatin-2 (TRPM2) Enhances Therapeutic Efficacy of Third Generation EGFR Inhibitors against EGFR Mutant Lung Cancer. Adv. Sci. 2024, 11, e2310126. [Google Scholar] [CrossRef]
  42. Wu, L.; Lin, W.; Liao, Q.; Wang, H.; Lin, C.; Tang, L.; Lian, W.; Chen, Z.; Li, K.; Xu, L.; et al. Calcium Channel Blocker Nifedipine Suppresses Colorectal Cancer Progression and Immune Escape by Preventing NFAT2 Nuclear Translocation. Cell Rep. 2020, 33, 108327. [Google Scholar] [CrossRef]
  43. Xiong, W.; Fan, H.; Zeng, Q.; Deng, Z.; Li, G.; Lu, W.; Zhang, B.; Lai, S.; Chen, X.; Xu, X. The in vitro anticancer effects of FS48 from salivary glands of Xenopsylla cheopis on NCI-H460 cells via its blockage of voltage-gated K+ channels. Acta Pharm. 2023, 73, 145–155. [Google Scholar] [CrossRef]
  44. Aissaoui, D.; Mlayah-Bellalouna, S.; Jebali, J.; Abdelkafi-Koubaa, Z.; Souid, S.; Moslah, W.; Othman, H.; Luis, J.; ElAyeb, M.; Marrakchi, N.; et al. Functional role of Kv1.1 and Kv1.3 channels in the neoplastic progression steps of three cancer cell lines, elucidated by scorpion peptides. Int. J. Biol. Macromol. 2018, 111, 1146–1155. [Google Scholar] [CrossRef] [PubMed]
  45. Leanza, L.; Romio, M.; Becker, K.A.; Azzolini, M.; Trentin, L.; Managò, A.; Venturini, E.; Zaccagnino, A.; Mattarei, A.; Carraretto, L.; et al. Direct Pharmacological Targeting of a Mitochondrial Ion Channel Selectively Kills Tumor Cells In Vivo. Cancer Cell 2017, 31, 516–531.e10. [Google Scholar] [CrossRef] [PubMed]
  46. Li, W.; Wilson, G.C.; Bachmann, M.; Wang, J.; Mattarei, A.; Paradisi, C.; Edwards, M.J.; Szabo, I.; Gulbins, E.; Ahmad, S.A.; et al. Inhibition of a Mitochondrial Potassium Channel in Combination with Gemcitabine and Abraxane Drastically Reduces Pancreatic Ductal Adenocarcinoma in an Immunocompetent Orthotopic Murine Model. Cancers 2022, 14, 2618. [Google Scholar] [CrossRef] [PubMed]
  47. Lowinus, T.; Heidel, F.H.; Bose, T.; Nimmagadda, S.C.; Schnöder, T.; Cammann, C.; Schmitz, I.; Seifert, U.; Fischer, T.; Schraven, B.; et al. Memantine potentiates cytarabine-induced cell death of acute leukemia correlating with inhibition of Kv1.3 potassium channels, AKT and ERK1/2 signaling. Cell Commun. Signal. 2019, 17, 5. [Google Scholar] [CrossRef] [PubMed]
  48. Shon, Y.H.; Park, S.D.; Nam, K.S. Effective chemopreventive activity of genistein against human breast cancer cells. J. Biochem. Mol. Biol. 2006, 39, 448–451. [Google Scholar] [CrossRef]
  49. Yu, Z.; Li, W.; Liu, F. Inhibition of proliferation and induction of apoptosis by genistein in colon cancer HT-29 cells. Cancer Lett. 2004, 215, 159–166. [Google Scholar] [CrossRef]
  50. Valdés-Abadía, B.; Morán-Zendejas, R.; Rangel-Flores, J.M.; Rodríguez-Menchaca, A.A. Chloroquine inhibits tumor-related Kv10.1 channel and decreases migration of MDA-MB-231 breast cancer cells in vitro. Eur. J. Pharmacol. 2019, 855, 262–266. [Google Scholar] [CrossRef] [PubMed]
  51. Hernández-Reséndiz, I.; Pacheu-Grau, D.; Sánchez, A.; Pardo, L.A. Inhibition of Kv10.1 Channels Sensitizes Mitochondria of Cancer Cells to Antimetabolic Agents. Cancers 2020, 12, 920. [Google Scholar] [CrossRef] [PubMed]
  52. Na, W.; Ma, B.; Shi, S.; Chen, Y.; Zhang, H.; Zhan, Y.; An, H. Procyanidin B1, a novel and specific inhibitor of Kv10.1 channel, suppresses the evolution of hepatoma. Biochem. Pharmacol. 2020, 178, 114089. [Google Scholar] [CrossRef]
  53. Wang, X.; Chen, Y.; Li, J.; Guo, S.; Lin, X.; Zhang, H.; Zhan, Y.; An, H. Tetrandrine, a novel inhibitor of ether-à-go-go-1 (Eag1), targeted to cervical cancer development. J. Cell. Physiol. 2019, 234, 7161–7173. [Google Scholar] [CrossRef] [PubMed]
  54. Wang, X.; Chen, Y.; Liu, H.; Guo, S.; Hu, Y.; Zhan, Y.; An, H. A novel anti-cancer mechanism of Nutlin-3 through downregulation of Eag1 channel and PI3K/AKT pathway. Biochem. Biophys. Res. Commun. 2019, 517, 445–451. [Google Scholar] [CrossRef]
  55. Ji, N.; Li, J.; Wei, Z.; Kong, F.; Jin, H.; Chen, X.; Li, Y.; Deng, Y. Effect of celastrol on growth inhibition of prostate cancer cells through the regulation of hERG channel in vitro. BioMed Res. Int. 2015, 2015, 308475. [Google Scholar] [CrossRef]
  56. Zhi, D.; Zhou, K.; Yu, D.; Fan, X.; Zhang, J.; Li, X.; Dong, M. hERG1 is involved in the pathophysiological process and inhibited by berberine in SKOV3 cells. Oncol. Lett. 2019, 17, 5653–5661. [Google Scholar] [CrossRef] [PubMed]
  57. Yan, C.; Li, F.; Zhang, Y.; Li, Y.; Li, M.; Wang, F.; Zhang, G.; Li, Y.; Li, B.; Zhao, X. Effects of As2O3 and Resveratrol on the Proliferation and Apoptosis of Colon Cancer Cells and the hERG-mediated Potential Mechanisms. Curr. Pharm. Des. 2019, 25, 1385–1391. [Google Scholar] [CrossRef]
  58. Petroni, G.; Bagni, G.; Iorio, J.; Duranti, C.; Lottini, T.; Stefanini, M.; Kragol, G.; Becchetti, A.; Arcangeli, A. Clarithromycin inhibits autophagy in colorectal cancer by regulating the hERG1 potassium channel interaction with PI3K. Cell Death Dis. 2020, 11, 161. [Google Scholar] [CrossRef]
  59. Rosa, P.; Catacuzzeno, L.; Sforna, L.; Mangino, G.; Carlomagno, S.; Mincione, G.; Petrozza, V.; Ragona, G.; Franciolini, F.; Calogero, A. BK channels blockage inhibits hypoxia-induced migration and chemoresistance to cisplatin in human glioblastoma cells. J. Cell. Physiol. 2018, 233, 6866–6877. [Google Scholar] [CrossRef] [PubMed]
  60. Maqoud, F.; Curci, A.; Scala, R.; Pannunzio, A.; Campanella, F.; Coluccia, M.; Passantino, G.; Zizzo, N.; Tricarico, D. Cell Cycle Regulation by Ca2+-Activated K+ (BK) Channels Modulators in SH-SY5Y Neuroblastoma Cells. Int. J. Mol. Sci. 2018, 19, 2442. [Google Scholar] [CrossRef] [PubMed]
  61. Li, N.; Liu, L.; Li, G.; Xia, M.; Du, C.; Zheng, Z. The role of BKCa in endometrial cancer HEC-1-B cell proliferation and migration. Gene 2018, 655, 42–47. [Google Scholar] [CrossRef]
  62. He, Y.; Lin, Y.; He, F.; Shao, L.; Ma, W.; He, F. Role for calcium-activated potassium channels (BK) in migration control of human hepatocellular carcinoma cells. J. Cell. Mol. Med. 2021, 25, 9685–9696. [Google Scholar] [CrossRef]
  63. Goda, A.A.; Siddique, A.B.; Mohyeldin, M.; Ayoub, N.M.; El Sayed, K.A. The Maxi-K (BK) Channel Antagonist Penitrem A as a Novel Breast Cancer-Targeted Therapeutic. Mar. Drugs 2018, 16, 157. [Google Scholar] [CrossRef]
  64. Lee, H.; Kwon, O.B.; Lee, J.E.; Jeon, Y.H.; Lee, D.S.; Min, S.H.; Kim, J.W. Repositioning Trimebutine Maleate as a Cancer Treatment Targeting Ovarian Cancer Stem Cells. Cells 2021, 10, 918. [Google Scholar] [CrossRef] [PubMed]
  65. Yeh, P.S.; Wu, S.J.; Hung, T.Y.; Huang, Y.M.; Hsu, C.W.; Sze, C.I.; Hsieh, Y.J.; Huang, C.W.; Wu, S.N. Evidence for the Inhibition by Temozolomide, an Imidazotetrazine Family Alkylator, of Intermediate-Conductance Ca2+-Activated K+ Channels in Glioma Cells. Cell. Physiol. Biochem. 2016, 38, 1727–1742. [Google Scholar] [CrossRef] [PubMed]
  66. Hung, T.Y.; Huang, H.I.; Wu, S.N.; Huang, C.W. Depressive effectiveness of vigabatrin (γ-vinyl-GABA), an antiepileptic drug, in intermediate-conductance calcium-activated potassium channels in human glioma cells. BMC Pharmacol. Toxicol. 2021, 22, 6. [Google Scholar] [CrossRef] [PubMed]
  67. Zhan, P.; Liu, L.; Nie, D.; Liu, Y.; Mao, X. Effect of inhibition of intermediate-conductance-Ca2+-activated K+ channels on HeLa cell proliferation. J. Cancer Res. Ther. 2018, 14, S41–S45. [Google Scholar] [CrossRef] [PubMed]
  68. Song, P.; Du, Y.; Song, W.; Chen, H.; Xuan, Z.; Zhao, L.; Chen, J.; Chen, J.; Guo, D.; Jin, C.; et al. KCa3.1 as an Effective Target for Inhibition of Growth and Progression of Intrahepatic Cholangiocarcinoma. J. Cancer 2017, 8, 1568–1578. [Google Scholar] [CrossRef] [PubMed]
  69. Ba, Y.; Malhotra, A. Potential of piperine in modulation of voltage-gated K+ current and its influences on cell cycle arrest and apoptosis in human prostate cancer cells. Eur. Rev. Med. Pharmacol. Sci. 2018, 22, 8999–9011. [Google Scholar]
  70. D’Arcangelo, D.; Scatozza, F.; Giampietri, C.; Marchetti, P.; Facchiano, F.; Facchiano, A. Ion Channel Expression in Human Melanoma Samples: In Silico Identification and Experimental Validation of Molecular Targets. Cancers 2019, 11, 446. [Google Scholar] [CrossRef] [PubMed]
  71. Sauter, D.R.; Sørensen, C.E.; Rapedius, M.; Brüggemann, A.; Novak, I. pH-sensitive K+ channel TREK-1 is a novel target in pancreatic cancer. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2016, 1862, 1994–2003. [Google Scholar] [CrossRef] [PubMed]
  72. Zúñiga, R.; Concha, G.; Cayo, A.; Cikutović-Molina, R.; Arevalo, B.; González, W.; Catalán, M.A.; Zúñiga, L. Withaferin A suppresses breast cancer cell proliferation by inhibition of the two-pore domain potassium (K2P9) channel TASK-3. Biomed. Pharmacother. 2020, 129, 110383. [Google Scholar] [CrossRef]
  73. Fukushiro-Lopes, D.; Hegel, A.D.; Russo, A.; Senyuk, V.; Liotta, M.; Beeson, G.C.; Beeson, C.C.; Burdette, J.; Potkul, R.K.; Gentile, S. Repurposing Kir6/SUR2 Channel Activator Minoxidil to Arrests Growth of Gynecologic Cancers. Front. Pharmacol. 2020, 11, 577. [Google Scholar] [CrossRef] [PubMed]
  74. Erdem Kış, E.; Tiftik, R.N.; Al Hennawi, K.; Ün, İ. The role of potassium channels in the proliferation and migration of endometrial adenocarcinoma HEC1-A cells. Mol. Biol. Rep. 2022, 49, 7447–7454. [Google Scholar] [CrossRef]
  75. Wang, L.; Ma, W.; Zhu, L.; Ye, D.; Li, Y.; Liu, S.; Li, H.; Zuo, W.; Li, B.; Ye, W.; et al. ClC-3 is a candidate of the channel proteins mediating acid-activated chloride currents in nasopharyngeal carcinoma cells. Am. J. Physiol. Cell Physiol 2012, 303, C14–C23. [Google Scholar] [CrossRef] [PubMed]
  76. Deshane, J.; Garner, C.C.; Sontheimer, H. Chlorotoxin inhibits glioma cell invasion via matrix metalloproteinase-2. J. Biol. Chem. 2003, 278, 4135–4144. [Google Scholar] [CrossRef]
  77. Britschgi, A.; Bill, A.; Brinkhaus, H.; Rothwell, C.; Clay, I.; Duss, S.; Rebhan, M.; Raman, P.; Guy, C.T.; Wetzel, K.; et al. Calcium-activated chloride channel ANO1 promotes breast cancer progression by activating EGFR and CAMK signaling. Proc. Natl. Acad. Sci. USA 2013, 110, E1026–E1034. [Google Scholar] [CrossRef]
  78. Park, S.; Das, R.; Nhiem, N.X.; Jeong, S.B.; Kim, M.; Kim, D.; Oh, H.I.; Cho, S.H.; Kwon, O.B.; Choi, J.H.; et al. ANO1-downregulation induced by schisandrathera D: A novel therapeutic target for the treatment of prostate and oral cancers. Front. Pharmacol. 2023, 14, 1163970. [Google Scholar] [CrossRef]
  79. Seo, Y.; Lee, S.; Kim, M.; Kim, D.; Jeong, S.B.; Das, R.; Sultana, A.; Park, S.; Nhiem, N.X.; Huong, P.T.T.; et al. Discovery of a novel natural compound, vitekwangin B, with ANO1 protein reduction properties and anticancer potential. Front. Pharmacol. 2024, 15, 1382787. [Google Scholar] [CrossRef]
  80. Jeon, D.; Jo, M.; Lee, Y.; Park, S.H.; Phan, H.T.L.; Nam, J.H.; Namkung, W. Inhibition of ANO1 by Cis- and Trans-Resveratrol and Their Anticancer Activity in Human Prostate Cancer PC-3 Cells. Int. J. Mol. Sci. 2023, 24, 1186. [Google Scholar] [CrossRef] [PubMed]
  81. Wang, Z.; Geng, R.; Chen, Y.; Qin, J.; Guo, S. Matairesinoside, a novel inhibitor of TMEM16A ion channel, loaded with functional hydrogel for lung cancer treatment. Int. J. Biol. Macromol. 2024, 264, 130618. [Google Scholar] [CrossRef]
  82. Wang, X.; Hao, A.; Song, G.; Elena, V.; Sun, Y.; Zhang, H.; Zhan, Y.; An, H.; Chen, Y. Inhibitory effect of daidzein on the calcium-activated chloride channel TMEM16A and its anti-lung adenocarcinoma activity. Int. J. Biol. Macromol. 2023, 253, 127261. [Google Scholar] [CrossRef]
  83. Shi, S.; Ma, B.; Sun, F.; Qu, C.; Li, G.; Shi, D.; Liu, W.; Zhang, H.; An, H. Zafirlukast inhibits the growth of lung adenocarcinoma via inhibiting TMEM16A channel activity. J. Biol. Chem. 2022, 298, 101731. [Google Scholar] [CrossRef] [PubMed]
  84. Fan, C.; Wu, H.; Du, X.; Li, C.; Zeng, W.; Qu, L.; Cang, C. Inhibition of lysosomal TRPML1 channel eliminates breast cancer stem cells by triggering ferroptosis. Cell Death Discov. 2024, 10, 256. [Google Scholar] [CrossRef]
  85. Kepp, O.; Menger, L.; Vacchelli, E.; Adjemian, S.; Martins, I.; Ma, Y.; Sukkurwala, A.Q.; Michaud, M.; Galluzzi, L.; Zitvogel, L.; et al. Anticancer activity of cardiac glycosides: At the frontier between cell-autonomous and immunological effects. Oncoimmunology 2012, 1, 1640–1642. [Google Scholar] [CrossRef] [PubMed]
  86. Gould, H.J., III; Norleans, J.; Ward, T.D.; Reid, C.; Paul, D. Selective lysis of breast carcinomas by simultaneous stimulation of sodium channels and blockade of sodium pumps. Oncotarget 2018, 9, 15606–15615. [Google Scholar] [CrossRef] [PubMed]
  87. Muscella, A.; Calabriso, N.; Vetrugno, C.; Fanizzi, F.P.; De Pascali, S.A.; Storelli, C.; Marsigliante, S. The platinum (II) complex [Pt(O,O′-acac)(γ-acac)(DMS)] alters the intracellular calcium homeostasis in MCF-7 breast cancer cells. Biochem. Pharmacol. 2011, 81, 91–103. [Google Scholar] [CrossRef] [PubMed]
  88. Pelzl, L.; Hosseinzadeh, Z.; Alzoubi, K.; Al-Maghout, T.; Schmidt, S.; Stournaras, C.; Lang, F. Impact of Na+/Ca2+ Exchangers on Therapy Resistance of Ovary Carcinoma Cells. Cell. Physiol. Biochem. 2015, 37, 1857–1868. [Google Scholar] [CrossRef] [PubMed]
  89. Pelzl, L.; Hosseinzadeh, Z.; Al-Maghout, T.; Singh, Y.; Sahu, I.; Bissinger, R.; Schmidt, S.; Alkahtani, S.; Stournaras, C.; Toulany, M.; et al. Role of Na+/Ca2+ Exchangers in Therapy Resistance of Medulloblastoma Cells. Cell. Physiol. Biochem. 2017, 42, 1240–1251. [Google Scholar] [CrossRef]
  90. Jathal, M.K.; Mudryj, M.; Dall’Era, M.A.; Ghosh, P.M. Amiloride sensitizes prostate cancer cells to the reversible tyrosine kinase inhibitor lapatinib by modulating Erbb3 subcellular localization. Cell. Mol. Life Sci. 2024, 82, 24. [Google Scholar] [CrossRef]
  91. Rojas, E.A.; Corchete, L.A.; San-Segundo, L.; Martínez-Blanch, J.F.; Codoñer, F.M.; Paíno, T.; Puig, N.; García-Sanz, R.; Mateos, M.V.; Ocio, E.M.; et al. Amiloride, An Old Diuretic Drug, Is a Potential Therapeutic Agent for Multiple Myeloma. Clin. Cancer Res. 2017, 23, 6602–6615. [Google Scholar] [CrossRef] [PubMed]
  92. Chen, Q.; Liu, Y.; Zhu, X.L.; Feng, F.; Yang, H.; Xu, W. Increased NHE1 expression is targeted by specific inhibitor cariporide to sensitize resistant breast cancer cells to doxorubicin in vitro and in vivo. BMC Cancer 2019, 19, 211. [Google Scholar] [CrossRef]
  93. Malamas, A.S.; Jin, E.; Zhang, Q.; Haaga, J.; Lu, Z.R. Anti-angiogenic Effects of Bumetanide Revealed by DCE-MRI with a Biodegradable Macromolecular Contrast Agent in a Colon Cancer Model. Pharm. Res. 2015, 32, 3029–3043. [Google Scholar] [CrossRef]
  94. Haas, B.R.; Sontheimer, H. Inhibition of the Sodium-Potassium-Chloride Cotransporter Isoform-1 reduces glioma invasion. Cancer Res. 2010, 70, 5597–5606. [Google Scholar] [CrossRef]
  95. Badwe, R.A.; Parmar, V.; Nair, N.; Joshi, S.; Hawaldar, R.; Pawar, S.; Kadayaprath, G.; Borthakur, B.B.; Rao Thammineedi, S.; Pandya, S.; et al. Effect of Peritumoral Infiltration of Local Anesthetic Before Surgery on Survival in Early Breast Cancer. J. Clin. Oncol. 2023, 41, 3318–3328. [Google Scholar] [CrossRef] [PubMed]
  96. Li, C.; Noonan, A.M.; Hays, J.; Roychowdhury, S.; Malalur, P.; Elkhatib, R.; Manne, A.; Mittra, A.; Rahman, S.; Yan, L.; et al. Riluzole in Combination with mFOLFOX6 and Bevacizumab in Treating Patients with Metastatic Colorectal Cancer: A Phase I Clinical Trial. Clin. Cancer Res. 2025, 31, 2115–2123. [Google Scholar] [CrossRef] [PubMed]
  97. Fu, S.; Hirte, H.; Welch, S.; Ilenchuk, T.T.; Lutes, T.; Rice, C.; Fields, N.; Nemet, A.; Dugourd, D.; Piha-Paul, S.; et al. First-in-human phase I study of SOR-C13, a TRPV6 calcium channel inhibitor, in patients with advanced solid tumors. Investig. New Drugs 2017, 35, 324–333. [Google Scholar] [CrossRef] [PubMed]
  98. Bhaskaran, D.; Savage, J.; Patel, A.; Collinson, F.; Mant, R.; Boele, F.; Brazil, L.; Meade, S.; Buckle, P.; Lax, S.; et al. A randomised phase II trial of temozolomide with or without cannabinoids in patients with recurrent glioblastoma (ARISTOCRAT): Protocol for a multi-centre, double-blind, placebo-controlled trial. BMC Cancer 2024, 24, 83. [Google Scholar] [CrossRef] [PubMed]
  99. Gilbert, S.M.; Gidley Baird, A.; Glazer, S.; Barden, J.A.; Glazer, A.; Teh, L.C.; King, J. A phase I clinical trial demonstrates that nfP2X(7) -targeted antibodies provide a novel, safe and tolerable topical therapy for basal cell carcinoma. Br. J. Dermatol. 2017, 177, 117–124. [Google Scholar] [CrossRef] [PubMed]
  100. Omuro, A.; Beal, K.; McNeill, K.; Young, R.J.; Thomas, A.; Lin, X.; Terziev, R.; Kaley, T.J.; DeAngelis, L.M.; Daras, M.; et al. Multicenter Phase IB Trial of Carboxyamidotriazole Orotate and Temozolomide for Recurrent and Newly Diagnosed Glioblastoma and Other Anaplastic Gliomas. J. Clin. Oncol. 2018, 36, 1702–1709. [Google Scholar] [CrossRef]
  101. Si, X.; Wang, J.; Cheng, Y.; Shi, J.; Cui, L.; Zhang, H.; Huang, Y.; Liu, W.; Chen, L.; Zhu, J.; et al. A phase III, randomized, double-blind, controlled trial of carboxyamidotriazole plus chemotherapy for the treatment of advanced non-small cell lung cancer. Ther. Adv. Med. Oncol. 2020, 12, 1758835920965849. [Google Scholar] [CrossRef] [PubMed]
  102. Mahalingam, D.; Wilding, G.; Denmeade, S.; Sarantopoulas, J.; Cosgrove, D.; Cetnar, J.; Azad, N.; Bruce, J.; Kurman, M.; Allgood, V.E.; et al. Mipsagargin, a novel thapsigargin-based PSMA-activated prodrug: Results of a first-in-man phase I clinical trial in patients with refractory, advanced or metastatic solid tumours. Br. J. Cancer 2016, 114, 986–994. [Google Scholar] [CrossRef]
  103. Mahalingam, D.; Peguero, J.; Cen, P.; Arora, S.P.; Sarantopoulos, J.; Rowe, J.; Allgood, V.; Tubb, B.; Campos, L. A Phase II, Multicenter, Single-Arm Study of Mipsagargin (G-202) as a Second-Line Therapy Following Sorafenib for Adult Patients with Progressive Advanced Hepatocellular Carcinoma. Cancers 2019, 11, 833. [Google Scholar] [CrossRef]
  104. Hussain, M.M.; Kotz, H.; Minasian, L.; Premkumar, A.; Sarosy, G.; Reed, E.; Zhai, S.; Steinberg, S.M.; Raggio, M.; Oliver, V.K.; et al. Phase II trial of carboxyamidotriazole in patients with relapsed epithelial ovarian cancer. J. Clin. Oncol. 2003, 21, 4356–4363. [Google Scholar] [CrossRef] [PubMed]
  105. Biller, A.; Pflugmann, I.; Badde, S.; Diem, R.; Wildemann, B.; Nagel, A.M.; Jordan, J.; Benkhedah, N.; Kleesiek, J. Sodium MRI in Multiple Sclerosis is Compatible with Intracellular Sodium Accumulation and Inflammation-Induced Hyper-Cellularity of Acute Brain Lesions. Sci. Rep. 2016, 6, 31269. [Google Scholar] [CrossRef] [PubMed]
  106. Murphy, E.; Eisner, D.A. Regulation of intracellular and mitochondrial sodium in health and disease. Circ. Res. 2009, 104, 292–303. [Google Scholar] [CrossRef]
  107. Cameron, I.L.; Smith, N.K.; Pool, T.B.; Sparks, R.L. Intracellular concentration of sodium and other elements as related to mitogenesis and oncogenesis in vivo. Cancer Res. 1980, 40, 1493–1500. [Google Scholar] [PubMed]
  108. Hürter, T.; Bröcker, W.; Bosma, H.J. Investigations on vasogenic and cytotoxic brain edema, comparing results from X-ray microanalysis and flame photometry. Microsc. Acta 1982, 85, 285–293. [Google Scholar]
  109. Barrett, T.; Riemer, F.; McLean, M.A.; Kaggie, J.; Robb, F.; Tropp, J.S.; Warren, A.; Bratt, O.; Shah, N.; Gnanapragasam, V.J.; et al. Quantification of Total and Intracellular Sodium Concentration in Primary Prostate Cancer and Adjacent Normal Prostate Tissue With Magnetic Resonance Imaging. Investig. Radiol. 2018, 53, 450–456. [Google Scholar] [CrossRef]
  110. Ouwerkerk, R.; Jacobs, M.A.; Macura, K.J.; Wolff, A.C.; Stearns, V.; Mezban, S.D.; Khouri, N.F.; Bluemke, D.A.; Bottomley, P.A. Elevated tissue sodium concentration in malignant breast lesions detected with non-invasive 23Na MRI. Breast Cancer Res. Treat. 2007, 106, 151–160. [Google Scholar] [CrossRef]
  111. Jacobs, M.A.; Ouwerkerk, R.; Wolff, A.C.; Gabrielson, E.; Warzecha, H.; Jeter, S.; Bluemke, D.A.; Wahl, R.; Stearns, V. Monitoring of neoadjuvant chemotherapy using multiparametric, 23Na sodium MR, and multimodality (PET/CT/MRI) imaging in locally advanced breast cancer. Breast Cancer Res. Treat. 2011, 128, 119–126. [Google Scholar] [CrossRef]
  112. Sanchez-Sandoval, A.L.; Hernández-Plata, E.; Gomora, J.C. Voltage-gated sodium channels: From roles and mechanisms in the metastatic cell behavior to clinical potential as therapeutic targets. Front. Pharmacol. 2023, 14, 1206136. [Google Scholar] [CrossRef] [PubMed]
  113. de Lera Ruiz, M.; Kraus, R.L. Voltage-Gated Sodium Channels: Structure, Function, Pharmacology, and Clinical Indications. J. Med. Chem. 2015, 58, 7093–7118. [Google Scholar] [CrossRef] [PubMed]
  114. Gao, R.; Shen, Y.; Cai, J.; Lei, M.; Wang, Z. Expression of voltage-gated sodium channel alpha subunit in human ovarian cancer. Oncol. Rep. 2010, 23, 1293–1299. [Google Scholar]
  115. Shan, B.; Dong, M.; Tang, H.; Wang, N.; Zhang, J.; Yan, C.; Jiao, X.; Zhang, H.; Wang, C. Voltage-gated sodium channels were differentially expressed in human normal prostate, benign prostatic hyperplasia and prostate cancer cells. Oncol. Lett. 2014, 8, 345–350. [Google Scholar] [CrossRef] [PubMed]
  116. Lin, S.; Lv, Y.; Xu, J.; Mao, X.; Chen, Z.; Lu, W. Over-expression of Nav1.6 channels is associated with lymph node metastases in colorectal cancer. World J. Surg. Oncol. 2019, 17, 175. [Google Scholar] [CrossRef] [PubMed]
  117. Leslie, T.K.; Tripp, A.; James, A.D.; Fraser, S.P.; Nelson, M.; Sajjaboontawee, N.; Capatina, A.L.; Toss, M.; Fadhil, W.; Salvage, S.C.; et al. A novel Nav1.5-dependent feedback mechanism driving glycolytic acidification in breast cancer metastasis. Oncogene 2024, 43, 2578–2594. [Google Scholar] [CrossRef]
  118. Xing, D.; Wang, J.; Ou, S.; Wang, Y.; Qiu, B.; Ding, D.; Guo, F.; Gao, Q. Expression of neonatal Nav1.5 in human brain astrocytoma and its effect on proliferation, invasion and apoptosis of astrocytoma cells. Oncol. Rep. 2014, 31, 2692–2700. [Google Scholar] [CrossRef]
  119. Li, H.; Liu, J.; Fan, N.; Wang, H.; Thomas, A.M.; Yan, Q.; Li, S.; Qin, H. Nav1.6 promotes the progression of human follicular thyroid carcinoma cells via JAK-STAT signaling pathway. Pathol. Res. Pract. 2022, 236, 153984. [Google Scholar] [CrossRef] [PubMed]
  120. Xia, J.; Huang, N.; Huang, H.; Sun, L.; Dong, S.; Su, J.; Zhang, J.; Wang, L.; Lin, L.; Shi, M.; et al. Voltage-gated sodium channel Nav 1.7 promotes gastric cancer progression through MACC1-mediated upregulation of NHE1. Int. J. Cancer 2016, 139, 2553–2569. [Google Scholar] [CrossRef]
  121. Chioni, A.M.; Brackenbury, W.J.; Calhoun, J.D.; Isom, L.L.; Djamgoz, M.B. A novel adhesion molecule in human breast cancer cells: Voltage-gated Na+ channel β1 subunit. Int. J. Biochem. Cell Biol. 2009, 41, 1216–1227. [Google Scholar] [CrossRef]
  122. Yoder, N.; Yoshioka, C.; Gouaux, E. Gating mechanisms of acid-sensing ion channels. Nature 2018, 555, 397–401. [Google Scholar] [CrossRef]
  123. Zhou, Z.H.; Song, J.W.; Li, W.; Liu, X.; Cao, L.; Wan, L.M.; Tan, Y.X.; Ji, S.P.; Liang, Y.M.; Gong, F. The acid-sensing ion channel, ASIC2, promotes invasion and metastasis of colorectal cancer under acidosis by activating the calcineurin/NFAT1 axis. J. Exp. Clin. Cancer Res. 2017, 36, 130. [Google Scholar] [CrossRef]
  124. Zhu, S.; Zhou, H.Y.; Deng, S.C.; Deng, S.J.; He, C.; Li, X.; Chen, J.Y.; Jin, Y.; Hu, Z.L.; Wang, F.; et al. ASIC1 and ASIC3 contribute to acidity-induced EMT of pancreatic cancer through activating Ca2+/RhoA pathway. Cell Death Dis. 2017, 8, e2806. [Google Scholar] [CrossRef]
  125. Rizaner, N.; Uzun, S.; Fraser, S.P.; Djamgoz, M.B.A.; Altun, S. Riluzole: Anti-invasive effects on rat prostate cancer cells under normoxic and hypoxic conditions. Basic Clin. Pharmacol. Toxicol. 2020, 127, 254–264. [Google Scholar] [CrossRef] [PubMed]
  126. Guzel, R.M.; Ogmen, K.; Ilieva, K.M.; Fraser, S.P.; Djamgoz, M.B.A. Colorectal cancer invasiveness in vitro: Predominant contribution of neonatal Nav1.5 under normoxia and hypoxia. J. Cell. Physiol. 2019, 234, 6582–6593. [Google Scholar] [CrossRef]
  127. Giorgi, C.; Danese, A.; Missiroli, S.; Patergnani, S.; Pinton, P. Calcium Dynamics as a Machine for Decoding Signals. Trends Cell Biol. 2018, 28, 258–273. [Google Scholar] [CrossRef]
  128. Marchi, S.; Giorgi, C.; Galluzzi, L.; Pinton, P. Ca2+ Fluxes and Cancer. Mol. Cell 2020, 78, 1055–1069. [Google Scholar] [CrossRef] [PubMed]
  129. Wu, L.; Lian, W.; Zhao, L. Calcium signaling in cancer progression and therapy. FEBS J. 2021, 288, 6187–6205. [Google Scholar] [CrossRef] [PubMed]
  130. Cui, C.; Merritt, R.; Fu, L.; Pan, Z. Targeting calcium signaling in cancer therapy. Acta Pharm. Sin. B 2017, 7, 3–17. [Google Scholar] [CrossRef] [PubMed]
  131. Prakriya, M.; Lewis, R.S. Store-Operated Calcium Channels. Physiol. Rev. 2015, 95, 1383–1436. [Google Scholar] [CrossRef] [PubMed]
  132. Karska, J.; Kowalski, S.; Saczko, J.; Moisescu, M.G.; Kulbacka, J. Mechanosensitive Ion Channels and Their Role in Cancer Cells. Membranes 2023, 13, 167. [Google Scholar] [CrossRef]
  133. Zamponi, G.W. Targeting voltage-gated calcium channels in neurological and psychiatric diseases. Nat. Rev. Drug Discov. 2016, 15, 19–34. [Google Scholar] [CrossRef]
  134. Schmidtko, A.; Lötsch, J.; Freynhagen, R.; Geisslinger, G. Ziconotide for treatment of severe chronic pain. Lancet 2010, 375, 1569–1577. [Google Scholar] [CrossRef]
  135. Gao, S.; Yao, X.; Yan, N. Structure of human Cav2.2 channel blocked by the painkiller ziconotide. Nature 2021, 596, 143–147. [Google Scholar] [CrossRef] [PubMed]
  136. Huang, Y.; Chen, S.R.; Pan, H.L. α2δ-1-Linked NMDA and AMPA Receptors in Neuropathic Pain and Gabapentinoid Action. J. Neurochem. 2025, 169, e70064. [Google Scholar] [CrossRef] [PubMed]
  137. He, Y.; Yang, X.; Yuan, M.; Zhang, X.; Tu, W.; Xue, W.; Wang, D.; Gao, D. Wireless discharge of piezoelectric nanogenerator opens voltage-gated ion channels for calcium overload-mediated tumor treatment. Biomaterials 2025, 321, 123311. [Google Scholar] [CrossRef]
  138. Zhu, G.; Liu, Z.; Epstein, J.I.; Davis, C.; Christudass, C.S.; Carter, H.B.; Landis, P.; Zhang, H.; Chung, J.Y.; Hewitt, S.M.; et al. A Novel Quantitative Multiplex Tissue Immunoblotting for Biomarkers Predicts a Prostate Cancer Aggressive Phenotype. Cancer Epidemiol. Biomark. Prev. 2015, 24, 1864–1872. [Google Scholar] [CrossRef] [PubMed]
  139. Debes, J.D.; Roberts, R.O.; Jacobson, D.J.; Girman, C.J.; Lieber, M.M.; Tindall, D.J.; Jacobsen, S.J. Inverse association between prostate cancer and the use of calcium channel blockers. Cancer Epidemiol. Biomark. Prev. 2004, 13, 255–259. [Google Scholar] [CrossRef] [PubMed]
  140. Jacquemet, G.; Baghirov, H.; Georgiadou, M.; Sihto, H.; Peuhu, E.; Cettour-Janet, P.; He, T.; Perälä, M.; Kronqvist, P.; Joensuu, H.; et al. L-type calcium channels regulate filopodia stability and cancer cell invasion downstream of integrin signalling. Nat. Commun. 2016, 7, 13297. [Google Scholar] [CrossRef]
  141. Fourbon, Y.; Guéguinou, M.; Félix, R.; Constantin, B.; Uguen, A.; Fromont, G.; Lajoie, L.; Magaud, C.; Lecomte, T.; Chamorey, E.; et al. Ca2+ protein alpha 1D of CaV1.3 regulates intracellular calcium concentration and migration of colon cancer cells through a non-canonical activity. Sci. Rep. 2017, 7, 14199. [Google Scholar] [CrossRef]
  142. Hao, J.; Bao, X.; Jin, B.; Wang, X.; Mao, Z.; Li, X.; Wei, L.; Shen, D.; Wang, J.L. Ca2+ channel subunit α 1D promotes proliferation and migration of endometrial cancer cells mediated by 17β-estradiol via the G protein-coupled estrogen receptor. FASEB J. 2015, 29, 2883–2893. [Google Scholar] [CrossRef] [PubMed]
  143. Gao, S.H.; Wang, G.Z.; Wang, L.P.; Feng, L.; Zhou, Y.C.; Yu, X.J.; Liang, F.; Yang, F.Y.; Wang, Z.; Sun, B.B.; et al. Mutations and clinical significance of calcium voltage-gated channel subunit alpha 1E (CACNA1E) in non-small cell lung cancer. Cell Calcium 2022, 102, 102527. [Google Scholar] [CrossRef] [PubMed]
  144. Natrajan, R.; Little, S.E.; Reis-Filho, J.S.; Hing, L.; Messahel, B.; Grundy, P.E.; Dome, J.S.; Schneider, T.; Vujanic, G.M.; Pritchard-Jones, K.; et al. Amplification and overexpression of CACNA1E correlates with relapse in favorable histology Wilms’ tumors. Clin. Cancer Res. 2006, 12, 7284–7293. [Google Scholar] [CrossRef] [PubMed]
  145. Yang, S.; Zhang, J.J.; Huang, X.Y. Orai1 and STIM1 are critical for breast tumor cell migration and metastasis. Cancer Cell 2009, 15, 124–134. [Google Scholar] [CrossRef]
  146. Motiani, R.K.; Abdullaev, I.F.; Trebak, M. A novel native store-operated calcium channel encoded by Orai3: Selective requirement of Orai3 versus Orai1 in estrogen receptor-positive versus estrogen receptor-negative breast cancer cells. J. Biol. Chem. 2010, 285, 19173–19183. [Google Scholar] [CrossRef]
  147. Motiani, R.K.; Stolwijk, J.A.; Newton, R.L.; Zhang, X.; Trebak, M. Emerging roles of Orai3 in pathophysiology. Channels 2013, 7, 392–401. [Google Scholar] [CrossRef]
  148. Motiani, R.K.; Zhang, X.; Harmon, K.E.; Keller, R.S.; Matrougui, K.; Bennett, J.A.; Trebak, M. Orai3 is an estrogen receptor α-regulated Ca2+ channel that promotes tumorigenesis. FASEB J. 2013, 27, 63–75. [Google Scholar] [CrossRef] [PubMed]
  149. Faouzi, M.; Hague, F.; Potier, M.; Ahidouch, A.; Sevestre, H.; Ouadid-Ahidouch, H. Down-regulation of Orai3 arrests cell-cycle progression and induces apoptosis in breast cancer cells but not in normal breast epithelial cells. J. Cell. Physiol. 2011, 226, 542–551. [Google Scholar] [CrossRef]
  150. Faouzi, M.; Kischel, P.; Hague, F.; Ahidouch, A.; Benzerdjeb, N.; Sevestre, H.; Penner, R.; Ouadid-Ahidouch, H. ORAI3 silencing alters cell proliferation and cell cycle progression via c-myc pathway in breast cancer cells. Biochim. Biophys. Acta (BBA) Mol. Cell Res 2013, 1833, 752–760. [Google Scholar] [CrossRef]
  151. Wang, J.Y.; Sun, J.; Huang, M.Y.; Wang, Y.S.; Hou, M.F.; Sun, Y.; He, H.; Krishna, N.; Chiu, S.J.; Lin, S.; et al. STIM1 overexpression promotes colorectal cancer progression, cell motility and COX-2 expression. Oncogene 2015, 34, 4358–4367. [Google Scholar] [CrossRef] [PubMed]
  152. Hasegawa, K.; Fujii, S.; Matsumoto, S.; Tajiri, Y.; Kikuchi, A.; Kiyoshima, T. YAP signaling induces PIEZO1 to promote oral squamous cell carcinoma cell proliferation. J. Pathol. 2021, 253, 80–93. [Google Scholar] [CrossRef] [PubMed]
  153. Qu, S.; Li, S.; Hu, Z. Upregulation of Piezo1 Is a Novel Prognostic Indicator in Glioma Patients. Cancer Manag. Res. 2020, 12, 3527–3536. [Google Scholar] [CrossRef] [PubMed]
  154. Clapham, D.E. TRP channels as cellular sensors. Nature 2003, 426, 517–524. [Google Scholar] [CrossRef]
  155. Zhang, M.; Ma, Y.; Ye, X.; Zhang, N.; Pan, L.; Wang, B. TRP (transient receptor potential) ion channel family: Structures, biological functions and therapeutic interventions for diseases. Signal Transduct. Target. Ther. 2023, 8, 261. [Google Scholar] [CrossRef]
  156. Duncan, L.M.; Deeds, J.; Hunter, J.; Shao, J.; Holmgren, L.M.; Woolf, E.A.; Tepper, R.I.; Shyjan, A.W. Down-regulation of the novel gene melastatin correlates with potential for melanoma metastasis. Cancer Res. 1998, 58, 1515–1520. [Google Scholar]
  157. Takahashi, N.; Chen, H.Y.; Harris, I.S.; Stover, D.G.; Selfors, L.M.; Bronson, R.T.; Deraedt, T.; Cichowski, K.; Welm, A.L.; Mori, Y.; et al. Cancer Cells Co-opt the Neuronal Redox-Sensing Channel TRPA1 to Promote Oxidative-Stress Tolerance. Cancer Cell 2018, 33, 985–1003.e7. [Google Scholar] [CrossRef] [PubMed]
  158. Prevarskaya, N.; Zhang, L.; Barritt, G. TRP channels in cancer. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2007, 1772, 937–946. [Google Scholar] [CrossRef]
  159. Huang, R.; Wang, F.; Yang, Y.; Ma, W.; Lin, Z.; Cheng, N.; Long, Y.; Deng, S.; Li, Z. Recurrent activations of transient receptor potential vanilloid-1 and vanilloid-4 promote cellular proliferation and migration in esophageal squamous cell carcinoma cells. FEBS Open Bio 2019, 9, 206–225. [Google Scholar] [CrossRef] [PubMed]
  160. Zoppoli, P.; Calice, G.; Laurino, S.; Ruggieri, V.; La Rocca, F.; La Torre, G.; Ciuffi, M.; Amendola, E.; De Vita, F.; Petrillo, A.; et al. TRPV2 Calcium Channel Gene Expression and Outcomes in Gastric Cancer Patients: A Clinically Relevant Association. J. Clin. Med. 2019, 8, 662. [Google Scholar] [CrossRef] [PubMed]
  161. Monet, M.; Lehen’kyi, V.; Gackiere, F.; Firlej, V.; Vandenberghe, M.; Roudbaraki, M.; Gkika, D.; Pourtier, A.; Bidaux, G.; Slomianny, C.; et al. Role of cationic channel TRPV2 in promoting prostate cancer migration and progression to androgen resistance. Cancer Res. 2010, 70, 1225–1235. [Google Scholar] [CrossRef]
  162. Santoni, G.; Farfariello, V.; Amantini, C. TRPV channels in tumor growth and progression. Adv. Exp. Med. Biol. 2011, 704, 947–967. [Google Scholar] [PubMed]
  163. Li, X.; Li, H.; Li, Z.; Wang, T.; Yu, D.; Jin, H.; Cao, Y. TRPV3 promotes the angiogenesis through HIF-1α-VEGF signaling pathway in A549 cells. Acta Histochem. 2022, 124, 151955. [Google Scholar] [CrossRef]
  164. Adapala, R.K.; Thoppil, R.J.; Ghosh, K.; Cappelli, H.C.; Dudley, A.C.; Paruchuri, S.; Keshamouni, V.; Klagsbrun, M.; Meszaros, J.G.; Chilian, W.M.; et al. Activation of mechanosensitive ion channel TRPV4 normalizes tumor vasculature and improves cancer therapy. Oncogene 2016, 35, 314–322. [Google Scholar] [CrossRef] [PubMed]
  165. Zhuang, L.; Peng, J.B.; Tou, L.; Takanaga, H.; Adam, R.M.; Hediger, M.A.; Freeman, M.R. Calcium-selective ion channel, CaT1, is apically localized in gastrointestinal tract epithelia and is aberrantly expressed in human malignancies. Lab. Investig. 2002, 82, 1755–1764. [Google Scholar] [CrossRef] [PubMed]
  166. Song, H.; Dong, M.; Zhou, J.; Sheng, W.; Li, X.; Gao, W. Expression and prognostic significance of TRPV6 in the development and progression of pancreatic cancer. Oncol. Rep. 2018, 39, 1432–1440. [Google Scholar] [CrossRef]
  167. Asghar, M.Y.; Magnusson, M.; Kemppainen, K.; Sukumaran, P.; Löf, C.; Pulli, I.; Kalhori, V.; Törnquist, K. Transient Receptor Potential Canonical 1 (TRPC1) Channels as Regulators of Sphingolipid and VEGF Receptor Expression: Implications for Thyroid Cancer Cell Migration and Proliferation. J. Biol. Chem. 2015, 290, 16116–16131. [Google Scholar] [CrossRef] [PubMed]
  168. Duncan, L.M.; Deeds, J.; Cronin, F.E.; Donovan, M.; Sober, A.J.; Kauffman, M.; McCarthy, J.J. Melastatin expression and prognosis in cutaneous malignant melanoma. J. Clin. Oncol. 2001, 19, 568–576. [Google Scholar] [CrossRef]
  169. Hsieh, C.C.; Su, Y.C.; Jiang, K.Y.; Ito, T.; Li, T.W.; Kaku-Ito, Y.; Cheng, S.T.; Chen, L.T.; Hwang, D.Y.; Shen, C.H. TRPM1 promotes tumor progression in acral melanoma by activating the Ca2+/CaMKIIδ/AKT pathway. J. Adv. Res. 2023, 43, 45–57. [Google Scholar] [CrossRef]
  170. Middelbeek, J.; Kuipers, A.J.; Henneman, L.; Visser, D.; Eidhof, I.; van Horssen, R.; Wieringa, B.; Canisius, S.V.; Zwart, W.; Wessels, L.F.; et al. TRPM7 is required for breast tumor cell metastasis. Cancer Res. 2012, 72, 4250–4261. [Google Scholar] [CrossRef]
  171. Chen, J.P.; Wang, J.; Luan, Y.; Wang, C.X.; Li, W.H.; Zhang, J.B.; Sha, D.; Shen, R.; Cui, Y.G.; Zhang, Z.; et al. TRPM7 promotes the metastatic process in human nasopharyngeal carcinoma. Cancer Lett. 2015, 356, 483–490. [Google Scholar] [CrossRef] [PubMed]
  172. Peier, A.M.; Moqrich, A.; Hergarden, A.C.; Reeve, A.J.; Andersson, D.A.; Story, G.M.; Earley, T.J.; Dragoni, I.; McIntyre, P.; Bevan, S.; et al. A TRP channel that senses cold stimuli and menthol. Cell 2002, 108, 705–715. [Google Scholar] [CrossRef]
  173. Huang, Y.; Li, S.; Jia, Z.; Zhao, W.; Zhou, C.; Zhang, R.; Ali, D.W.; Michalak, M.; Chen, X.Z.; Tang, J. Transient Receptor Potential Melastatin 8 (TRPM8) Channel Regulates Proliferation and Migration of Breast Cancer Cells by Activating the AMPK-ULK1 Pathway to Enhance Basal Autophagy. Front. Oncol. 2020, 10, 573127. [Google Scholar] [CrossRef] [PubMed]
  174. Li, Q.; Wang, X.; Yang, Z.; Wang, B.; Li, S. Menthol induces cell death via the TRPM8 channel in the human bladder cancer cell line T24. Oncology 2009, 77, 335–341. [Google Scholar] [CrossRef]
  175. Laurino, S.; Mazzone, P.; Ruggieri, V.; Zoppoli, P.; Calice, G.; Lapenta, A.; Ciuffi, M.; Ignomirelli, O.; Vita, G.; Sgambato, A.; et al. Cationic Channel TRPV2 Overexpression Promotes Resistance to Cisplatin-Induced Apoptosis in Gastric Cancer Cells. Front. Pharmacol. 2021, 12, 746628. [Google Scholar] [CrossRef]
  176. O’Reilly, D.; Downing, T.; Kouba, S.; Potier-Cartereau, M.; McKenna, D.J.; Vandier, C.; Buchanan, P.J. CaV1.3 enhanced store operated calcium promotes resistance to androgen deprivation in prostate cancer. Cell Calcium 2022, 103, 102554. [Google Scholar] [CrossRef] [PubMed]
  177. Sui, X.; Geng, J.H.; Li, Y.H.; Zhu, G.Y.; Wang, W.H. Calcium channel α2δ1 subunit (CACNA2D1) enhances radioresistance in cancer stem-like cells in non-small cell lung cancer cell lines. Cancer Manag. Res. 2018, 10, 5009–5018. [Google Scholar] [CrossRef] [PubMed]
  178. Khoubza, L.; Chatelain, F.C.; Feliciangeli, S.; Lesage, F.; Bichet, D. Physiological roles of heteromerization: Focus on the two-pore domain potassium channels. J. Physiol. 2021, 599, 1041–1055. [Google Scholar] [CrossRef] [PubMed]
  179. Goldstein, S.A.; Bockenhauer, D.; O’Kelly, I.; Zilberberg, N. Potassium leak channels and the KCNK family of two-P-domain subunits. Nat. Rev. Neurosci. 2001, 2, 175–184. [Google Scholar] [CrossRef]
  180. Guéguinou, M.; Chantôme, A.; Fromont, G.; Bougnoux, P.; Vandier, C.; Potier-Cartereau, M. KCa and Ca2+ channels: The complex thought. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2014, 1843, 2322–2333. [Google Scholar] [CrossRef]
  181. Mohr, C.J.; Schroth, W.; Mürdter, T.E.; Gross, D.; Maier, S.; Stegen, B.; Dragoi, A.; Steudel, F.A.; Stehling, S.; Hoppe, R.; et al. Subunits of BK channels promote breast cancer development and modulate responses to endocrine treatment in preclinical models. Br. J. Pharmacol. 2022, 179, 2906–2924. [Google Scholar] [CrossRef] [PubMed]
  182. Chen, J.; Xuan, Z.; Song, W.; Han, W.; Chen, H.; Du, Y.; Xie, H.; Zhao, Y.; Zheng, S.; Song, P. EAG1 enhances hepatocellular carcinoma proliferation by modulating SKP2 and metastasis through pseudopod formation. Oncogene 2021, 40, 163–176. [Google Scholar] [CrossRef] [PubMed]
  183. Gao, S.; Wang, W.; Ye, W.; Wang, K. The mechanism study of Eag1 potassium channel in gastric cancer. Transl. Cancer Res. 2022, 11, 3827–3840. [Google Scholar] [CrossRef]
  184. Wang, X.; Chen, Y.; Zhang, Y.; Guo, S.; Mo, L.; An, H.; Zhan, Y. Eag1 Voltage-Dependent Potassium Channels: Structure, Electrophysiological Characteristics, and Function in Cancer. J. Membr. Biol. 2017, 250, 123–132. [Google Scholar] [CrossRef]
  185. Hou, X.; Ouyang, J.; Tang, L.; Wu, P.; Deng, X.; Yan, Q.; Shi, L.; Fan, S.; Fan, C.; Guo, C.; et al. KCNK1 promotes proliferation and metastasis of breast cancer cells by activating lactate dehydrogenase A (LDHA) and up-regulating H3K18 lactylation. PLoS Biol. 2024, 22, e3002666. [Google Scholar] [CrossRef]
  186. Mohr, C.J.; Gross, D.; Sezgin, E.C.; Steudel, F.A.; Ruth, P.; Huber, S.M.; Lukowski, R. K(Ca)3.1 Channels Confer Radioresistance to Breast Cancer Cells. Cancers 2019, 11, 1285. [Google Scholar] [CrossRef] [PubMed]
  187. Xu, P.; Mo, X.; Xia, R.; Jiang, L.; Zhang, C.; Xu, H.; Sun, Q.; Zhou, G.; Zhang, Y.; Wang, Y.; et al. KCNN4 promotes the progression of lung adenocarcinoma by activating the AKT and ERK signaling pathways. Cancer Biomark. 2021, 31, 187–201. [Google Scholar] [CrossRef] [PubMed]
  188. Leithner, K.; Hirschmugl, B.; Li, Y.; Tang, B.; Papp, R.; Nagaraj, C.; Stacher, E.; Stiegler, P.; Lindenmann, J.; Olschewski, A.; et al. TASK-1 Regulates Apoptosis and Proliferation in a Subset of Non-Small Cell Lung Cancers. PLoS ONE 2016, 11, e0157453. [Google Scholar] [CrossRef] [PubMed]
  189. Liu, H.; Huang, J.; Peng, J.; Wu, X.; Zhang, Y.; Zhu, W.; Guo, L. Upregulation of the inwardly rectifying potassium channel Kir2.1 (KCNJ2) modulates multidrug resistance of small-cell lung cancer under the regulation of miR-7 and the Ras/MAPK pathway. Mol. Cancer 2015, 14, 59. [Google Scholar] [CrossRef]
  190. Bachmann, M.; Rossa, A.; Varanita, T.; Fioretti, B.; Biasutto, L.; Milenkovic, S.; Checchetto, V.; Peruzzo, R.; Ahmad, S.A.; Patel, S.H.; et al. Pharmacological targeting of the mitochondrial calcium-dependent potassium channel KCa3.1 triggers cell death and reduces tumor growth and metastasis in vivo. Cell Death Dis. 2022, 13, 1055. [Google Scholar] [CrossRef] [PubMed]
  191. Capera, J.; Pérez-Verdaguer, M.; Peruzzo, R.; Navarro-Pérez, M.; Martínez-Pinna, J.; Alberola-Die, A.; Morales, A.; Leanza, L.; Szabó, I.; Felipe, A. A novel mitochondrial Kv1.3-caveolin axis controls cell survival and apoptosis. eLife 2021, 10, e69099. [Google Scholar] [CrossRef]
  192. Jentsch, T.J.; Pusch, M. CLC Chloride Channels and Transporters: Structure, Function, Physiology, and Disease. Physiol. Rev. 2018, 98, 1493–1590. [Google Scholar] [CrossRef] [PubMed]
  193. Hu, Y.; Tuo, B. The function of chloride channels in digestive system disease (Review). Int. J. Mol. Med. 2025, 55, 99. [Google Scholar] [CrossRef] [PubMed]
  194. Qiu, Z.; Dubin, A.E.; Mathur, J.; Tu, B.; Reddy, K.; Miraglia, L.J.; Reinhardt, J.; Orth, A.P.; Patapoutian, A. SWELL1, a plasma membrane protein, is an essential component of volume-regulated anion channel. Cell 2014, 157, 447–458. [Google Scholar] [CrossRef] [PubMed]
  195. Voss, F.K.; Ullrich, F.; Münch, J.; Lazarow, K.; Lutter, D.; Mah, N.; Andrade-Navarro, M.A.; von Kries, J.P.; Stauber, T.; Jentsch, T.J. Identification of LRRC8 heteromers as an essential component of the volume-regulated anion channel VRAC. Science 2014, 344, 634–638. [Google Scholar] [CrossRef] [PubMed]
  196. Lynch, J.W. Molecular structure and function of the glycine receptor chloride channel. Physiol. Rev. 2004, 84, 1051–1095. [Google Scholar] [CrossRef]
  197. Hong, S.; Bi, M.; Wang, L.; Kang, Z.; Ling, L.; Zhao, C. CLC-3 channels in cancer (review). Oncol. Rep. 2015, 33, 507–514. [Google Scholar] [CrossRef]
  198. Habela, C.W.; Olsen, M.L.; Sontheimer, H. ClC3 is a critical regulator of the cell cycle in normal and malignant glial cells. J. Neurosci. 2008, 28, 9205–9217. [Google Scholar] [CrossRef]
  199. Cuddapah, V.A.; Sontheimer, H. Molecular interaction and functional regulation of ClC-3 by Ca2+/calmodulin-dependent protein kinase II (CaMKII) in human malignant glioma. J. Biol. Chem. 2010, 285, 11188–11196. [Google Scholar] [CrossRef]
  200. Cuddapah, V.A.; Turner, K.L.; Seifert, S.; Sontheimer, H. Bradykinin-induced chemotaxis of human gliomas requires the activation of KCa3.1 and ClC-3. J. Neurosci. 2013, 33, 1427–1440. [Google Scholar] [CrossRef] [PubMed]
  201. Ye, D.; Luo, H.; Lai, Z.; Zou, L.; Zhu, L.; Mao, J.; Jacob, T.; Ye, W.; Wang, L.; Chen, L. ClC-3 Chloride Channel Proteins Regulate the Cell Cycle by Up-regulating cyclin D1-CDK4/6 through Suppressing p21/p27 Expression in Nasopharyngeal Carcinoma Cells. Sci. Rep. 2016, 6, 30276. [Google Scholar] [CrossRef] [PubMed]
  202. Mao, J.; Chen, L.; Xu, B.; Wang, L.; Li, H.; Guo, J.; Li, W.; Nie, S.; Jacob, T.J.; Wang, L. Suppression of ClC-3 channel expression reduces migration of nasopharyngeal carcinoma cells. Biochem. Pharmacol. 2008, 75, 1706–1716. [Google Scholar] [CrossRef]
  203. Guan, Y.T.; Xie, Y.; Zhou, H.; Shi, H.Y.; Zhu, Y.Y.; Zhang, X.L.; Luan, Y.; Shen, X.M.; Chen, Y.P.; Xu, L.J.; et al. Overexpression of chloride channel-3 (ClC-3) is associated with human cervical carcinoma development and prognosis. Cancer Cell Int. 2019, 19, 8. [Google Scholar] [CrossRef]
  204. Parisi, G.F.; Papale, M.; Pecora, G.; Rotolo, N.; Manti, S.; Russo, G.; Leonardi, S. Cystic Fibrosis and Cancer: Unraveling the Complex Role of CFTR Gene in Cancer Susceptibility. Cancers 2023, 15, 4244. [Google Scholar] [CrossRef]
  205. Neglia, J.P.; FitzSimmons, S.C.; Maisonneuve, P.; Schöni, M.H.; Schöni-Affolter, F.; Corey, M.; Lowenfels, A.B. The risk of cancer among patients with cystic fibrosis. Cystic Fibrosis and Cancer Study Group. N. Engl. J. Med. 1995, 332, 494–499. [Google Scholar] [CrossRef] [PubMed]
  206. Maisonneuve, P.; FitzSimmons, S.C.; Neglia, J.P.; Campbell, P.W., III; Lowenfels, A.B. Cancer risk in nontransplanted and transplanted cystic fibrosis patients: A 10-year study. J. Natl. Cancer Inst. 2003, 95, 381–387. [Google Scholar] [CrossRef] [PubMed]
  207. Maisonneuve, P.; Marshall, B.C.; Knapp, E.A.; Lowenfels, A.B. Cancer risk in cystic fibrosis: A 20-year nationwide study from the United States. J. Natl. Cancer Inst. 2013, 105, 122–129. [Google Scholar] [CrossRef] [PubMed]
  208. Hadjiliadis, D.; Khoruts, A.; Zauber, A.G.; Hempstead, S.E.; Maisonneuve, P.; Lowenfels, A.B. Cystic Fibrosis Colorectal Cancer Screening Consensus Recommendations. Gastroenterology 2018, 154, 736–745.e14. [Google Scholar] [CrossRef]
  209. Maisonneuve, P.; Marshall, B.C.; Lowenfels, A.B. Risk of pancreatic cancer in patients with cystic fibrosis. Gut 2007, 56, 1327–1328. [Google Scholar] [CrossRef]
  210. Schroeder, B.C.; Cheng, T.; Jan, Y.N.; Jan, L.Y. Expression cloning of TMEM16A as a calcium-activated chloride channel subunit. Cell 2008, 134, 1019–1029. [Google Scholar] [CrossRef] [PubMed]
  211. Yang, Y.D.; Cho, H.; Koo, J.Y.; Tak, M.H.; Cho, Y.; Shim, W.S.; Park, S.P.; Lee, J.; Lee, B.; Kim, B.M.; et al. TMEM16A confers receptor-activated calcium-dependent chloride conductance. Nature 2008, 455, 1210–1215. [Google Scholar] [CrossRef] [PubMed]
  212. Caputo, A.; Caci, E.; Ferrera, L.; Pedemonte, N.; Barsanti, C.; Sondo, E.; Pfeffer, U.; Ravazzolo, R.; Zegarra-Moran, O.; Galietta, L.J. TMEM16A, a membrane protein associated with calcium-dependent chloride channel activity. Science 2008, 322, 590–594. [Google Scholar] [CrossRef] [PubMed]
  213. Hartzell, H.C.; Yu, K.; Xiao, Q.; Chien, L.T.; Qu, Z. Anoctamin/TMEM16 family members are Ca2+-activated Cl channels. J. Physiol. 2009, 587, 2127–2139. [Google Scholar] [CrossRef]
  214. Heinze, C.; Seniuk, A.; Sokolov, M.V.; Huebner, A.K.; Klementowicz, A.E.; Szijártó, I.A.; Schleifenbaum, J.; Vitzthum, H.; Gollasch, M.; Ehmke, H.; et al. Disruption of vascular Ca2+-activated chloride currents lowers blood pressure. J. Clin. Investig. 2014, 124, 675–686. [Google Scholar] [CrossRef] [PubMed]
  215. Filippou, A.; Pehkonen, H.; Karhemo, P.R.; Väänänen, J.; Nieminen, A.I.; Klefström, J.; Grénman, R.; Mäkitie, A.A.; Joensuu, H.; Monni, O. ANO1 Expression Orchestrates p27Kip1/MCL1-Mediated Signaling in Head and Neck Squamous Cell Carcinoma. Cancers 2021, 13, 1170. [Google Scholar] [CrossRef]
  216. Zeng, X.; Pan, D.; Wu, H.; Chen, H.; Yuan, W.; Zhou, J.; Shen, Z.; Chen, S. Transcriptional activation of ANO1 promotes gastric cancer progression. Biochem. Biophys. Res. Commun. 2019, 512, 131–136. [Google Scholar] [CrossRef] [PubMed]
  217. Chen, J.; Wang, H.; Peng, F.; Qiao, H.; Liu, L.; Wang, L.; Shang, B. Ano1 is a Prognostic Biomarker That is Correlated with Immune Infiltration in Colorectal Cancer. Int. J. Gen. Med. 2022, 15, 1547–1564. [Google Scholar] [CrossRef]
  218. Zhang, G.; Shu, Z.; Yu, J.; Li, J.; Yi, P.; Wu, B.; Deng, D.; Yan, S.; Li, Y.; Ren, D.; et al. High ANO1 expression is a prognostic factor and correlated with an immunosuppressive tumor microenvironment in pancreatic cancer. Front. Immunol. 2024, 15, 1341209. [Google Scholar] [CrossRef] [PubMed]
  219. Duvvuri, U.; Shiwarski, D.J.; Xiao, D.; Bertrand, C.; Huang, X.; Edinger, R.S.; Rock, J.R.; Harfe, B.D.; Henson, B.J.; Kunzelmann, K.; et al. TMEM16A induces MAPK and contributes directly to tumorigenesis and cancer progression. Cancer Res. 2012, 72, 3270–3281. [Google Scholar] [CrossRef] [PubMed]
  220. Jiang, F.; Jia, K.; Chen, Y.; Ji, C.; Chong, X.; Li, Z.; Zhao, F.; Bai, Y.; Ge, S.; Gao, J.; et al. ANO1-Mediated Inhibition of Cancer Ferroptosis Confers Immunotherapeutic Resistance through Recruiting Cancer-Associated Fibroblasts. Adv. Sci. 2023, 10, e2300881. [Google Scholar] [CrossRef]
  221. Carpanese, V.; Festa, M.; Prosdocimi, E.; Bachmann, M.; Sadeghi, S.; Bertelli, S.; Stein, F.; Velle, A.; Abdel-Salam, M.A.L.; Romualdi, C.; et al. Interactomic exploration of LRRC8A in volume-regulated anion channels. Cell Death Discov. 2024, 10, 299. [Google Scholar] [CrossRef] [PubMed]
  222. Xu, R.; Hu, Y.; Xie, Q.; Zhang, C.; Zhao, Y.; Zhang, H.; Shi, H.; Wang, X.; Shi, C. LRRC8A Is a Promising Prognostic Biomarker and Therapeutic Target for Pancreatic Adenocarcinoma. Cancers 2022, 14, 5526. [Google Scholar] [CrossRef] [PubMed]
  223. Zhang, H.; Liu, R.; Jing, Z.; Li, C.; Fan, W.; Li, H.; Li, H.; Ren, J.; Cui, S.; Zhao, W.; et al. LRRC8A as a central mediator promotes colon cancer metastasis by regulating PIP5K1B/PIP2 pathway. Biochim. Biophys. Acta (BBA) Mol. Basis Dis. 2024, 1870, 167066. [Google Scholar] [CrossRef]
  224. Gumireddy, K.; Li, A.; Kossenkov, A.V.; Sakurai, M.; Yan, J.; Li, Y.; Xu, H.; Wang, J.; Zhang, P.J.; Zhang, L.; et al. The mRNA-edited form of GABRA3 suppresses GABRA3-mediated Akt activation and breast cancer metastasis. Nat. Commun. 2016, 7, 10715. [Google Scholar] [CrossRef]
  225. Sizemore, G.M.; Sizemore, S.T.; Seachrist, D.D.; Keri, R.A. GABA(A) receptor pi (GABRP) stimulates basal-like breast cancer cell migration through activation of extracellular-regulated kinase 1/2 (ERK1/2). J. Biol. Chem. 2014, 289, 24102–24113. [Google Scholar] [CrossRef]
  226. Correia, L.; Shalygin, A.; Erbacher, A.; Zaisserer, J.; Gudermann, T.; Chubanov, V. TRPM7 underlies cadmium cytotoxicity in pulmonary cells. Arch. Toxicol. 2025, 99, 3269–3281. [Google Scholar] [CrossRef]
  227. Dong, X.P.; Cheng, X.; Mills, E.; Delling, M.; Wang, F.; Kurz, T.; Xu, H. The type IV mucolipidosis-associated protein TRPML1 is an endolysosomal iron release channel. Nature 2008, 455, 992–996. [Google Scholar] [CrossRef]
  228. Köles, L.; Ribiczey, P.; Szebeni, A.; Kádár, K.; Zelles, T.; Zsembery, Á. The Role of TRPM7 in Oncogenesis. Int. J. Mol. Sci. 2024, 25, 719. [Google Scholar] [CrossRef]
  229. Auwercx, J.; Neve, B.; Vanlaeys, A.; Fourgeaud, M.; Bourrin-Reynard, I.; Souidi, M.; Brassart-Pasco, S.; Hague, F.; Guenin, S.; Duchene, B.; et al. The kinase domain of TRPM7 interacts with PAK1 and regulates pancreatic cancer cell epithelial-to-mesenchymal transition. Cell Death Dis. 2025, 16, 335. [Google Scholar] [CrossRef] [PubMed]
  230. Wang, H.; Li, B.; Asha, K.; Pangilinan, R.L.; Thuraisamy, A.; Chopra, H.; Rokudai, S.; Yu, Y.; Prives, C.L.; Zhu, Y. The ion channel TRPM7 regulates zinc-depletion-induced MDMX degradation. J. Biol. Chem. 2021, 297, 101292. [Google Scholar] [CrossRef] [PubMed]
  231. Nascimento Da Conceicao, V.; Sun, Y.; Venkatesan, M.; De La Chapa, J.; Ramachandran, K.; Jasrotia, R.S.; Drel, V.; Chai, X.; Mishra, B.B.; Madesh, M.; et al. Naltriben promotes tumor growth by activating the TRPM7-mediated development of the anti-inflammatory M2 phenotype. NPJ Precis. Oncol. 2025, 9, 29. [Google Scholar] [CrossRef] [PubMed]
  232. Qi, J.; Li, Q.; Xin, T.; Lu, Q.; Lin, J.; Zhang, Y.; Luo, H.; Zhang, F.; Xing, Y.; Wang, W.; et al. MCOLN1/TRPML1 in the lysosome: A promising target for autophagy modulation in diverse diseases. Autophagy 2024, 20, 1712–1722. [Google Scholar] [CrossRef]
  233. Zhang, H.L.; Hu, B.X.; Ye, Z.P.; Li, Z.L.; Liu, S.; Zhong, W.Q.; Du, T.; Yang, D.; Mai, J.; Li, L.C.; et al. TRPML1 triggers ferroptosis defense and is a potential therapeutic target in AKT-hyperactivated cancer. Sci. Transl. Med. 2024, 16, eadk0330. [Google Scholar] [CrossRef] [PubMed]
  234. Qi, J.; Xing, Y.; Liu, Y.; Wang, M.M.; Wei, X.; Sui, Z.; Ding, L.; Zhang, Y.; Lu, C.; Fei, Y.H.; et al. MCOLN1/TRPML1 finely controls oncogenic autophagy in cancer by mediating zinc influx. Autophagy 2021, 17, 4401–4422. [Google Scholar] [CrossRef]
  235. Liu, S.; Liu, H.; Jiang, J.; Liu, G.; Liu, J. Molecular machines for transmembrane ion transport. Chem. Commun. 2025, 61, 14598–14610. [Google Scholar] [CrossRef]
  236. Li, L.; Feng, R.; Xu, Q.; Zhang, F.; Liu, T.; Cao, J.; Fei, S. Expression of the β3 subunit of Na+/K+-ATPase is increased in gastric cancer and regulates gastric cancer cell progression and prognosis via the PI3/AKT pathway. Oncotarget 2017, 8, 84285–84299. [Google Scholar] [CrossRef] [PubMed]
  237. Khajah, M.A.; Mathew, P.M.; Luqmani, Y.A. Na+/K+ ATPase activity promotes invasion of endocrine resistant breast cancer cells. PLoS ONE 2018, 13, e0193779. [Google Scholar] [CrossRef] [PubMed]
  238. Lee, W.J.; Roberts-Thomson, S.J.; Monteith, G.R. Plasma membrane calcium-ATPase 2 and 4 in human breast cancer cell lines. Biochem. Biophys. Res. Commun. 2005, 337, 779–783. [Google Scholar] [CrossRef]
  239. Aung, C.S.; Kruger, W.A.; Poronnik, P.; Roberts-Thomson, S.J.; Monteith, G.R. Plasma membrane Ca2+-ATPase expression during colon cancer cell line differentiation. Biochem. Biophys. Res. Commun. 2007, 355, 932–936. [Google Scholar] [CrossRef] [PubMed]
  240. Saito, K.; Uzawa, K.; Endo, Y.; Kato, Y.; Nakashima, D.; Ogawara, K.; Shiba, M.; Bukawa, H.; Yokoe, H.; Tanzawa, H. Plasma membrane Ca2+ ATPase isoform 1 down-regulated in human oral cancer. Oncol. Rep. 2006, 15, 49–55. [Google Scholar] [CrossRef]
  241. Dang, D.; Rao, R. Calcium-ATPases: Gene disorders and dysregulation in cancer. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2016, 1863, 1344–1350. [Google Scholar] [CrossRef] [PubMed]
  242. Fan, L.; Li, A.; Li, W.; Cai, P.; Yang, B.; Zhang, M.; Gu, Y.; Shu, Y.; Sun, Y.; Shen, Y.; et al. Novel role of Sarco/endoplasmic reticulum calcium ATPase 2 in development of colorectal cancer and its regulation by F36, a curcumin analog. Biomed. Pharmacother. 2014, 68, 1141–1148. [Google Scholar] [CrossRef]
  243. Brouland, J.P.; Gélébart, P.; Kovàcs, T.; Enouf, J.; Grossmann, J.; Papp, B. The loss of sarco/endoplasmic reticulum calcium transport ATPase 3 expression is an early event during the multistep process of colon carcinogenesis. Am. J. Pathol. 2005, 167, 233–242. [Google Scholar] [CrossRef]
  244. Grice, D.M.; Vetter, I.; Faddy, H.M.; Kenny, P.A.; Roberts-Thomson, S.J.; Monteith, G.R. Golgi calcium pump secretory pathway calcium ATPase 1 (SPCA1) is a key regulator of insulin-like growth factor receptor (IGF1R) processing in the basal-like breast cancer cell line MDA-MB-231. J. Biol. Chem. 2010, 285, 37458–37466. [Google Scholar] [CrossRef] [PubMed]
  245. Feng, M.; Grice, D.M.; Faddy, H.M.; Nguyen, N.; Leitch, S.; Wang, Y.; Muend, S.; Kenny, P.A.; Sukumar, S.; Roberts-Thomson, S.J.; et al. Store-independent activation of Orai1 by SPCA2 in mammary tumors. Cell 2010, 143, 84–98. [Google Scholar] [CrossRef]
  246. Rodrigues, T.; Estevez, G.N.N.; Tersariol, I. Na+/Ca2+ exchangers: Unexploited opportunities for cancer therapy? Biochem. Pharmacol. 2019, 163, 357–361. [Google Scholar] [CrossRef] [PubMed]
  247. McLean, L.A.; Roscoe, J.; Jorgensen, N.K.; Gorin, F.A.; Cala, P.M. Malignant gliomas display altered pH regulation by NHE1 compared with nontransformed astrocytes. Am. J. Physiol. Cell Physiol. 2000, 278, C676–C688. [Google Scholar] [CrossRef] [PubMed]
  248. Sun, Z.; Luan, S.; Yao, Y.; Qin, T.; Xu, X.; Shen, Z.; Yao, R.; Yue, L. NHE1 Mediates 5-Fu Resistance in Gastric Cancer via STAT3 Signaling Pathway. OncoTargets Ther. 2020, 13, 8521–8532. [Google Scholar] [CrossRef]
  249. Zhou, Y.; Sun, W.; Chen, N.; Xu, C.; Wang, X.; Dong, K.; Zhang, B.; Zhang, J.; Hao, N.; Sun, A.; et al. Discovery of NKCC1 as a potential therapeutic target to inhibit hepatocellular carcinoma cell growth and metastasis. Oncotarget 2017, 8, 66328–66342. [Google Scholar] [CrossRef]
  250. Wang, J.F.; Zhao, K.; Chen, Y.Y.; Qiu, Y.; Zhu, J.H.; Li, B.P.; Wang, Z.; Chen, J.Q. NKCC1 promotes proliferation, invasion and migration in human gastric cancer cells via activation of the MAPK-JNK/EMT signaling pathway. J. Cancer 2021, 12, 253–263. [Google Scholar] [CrossRef] [PubMed]
  251. Chida, K.; Kanazawa, H.; Kinoshita, H.; Roy, A.M.; Hakamada, K.; Takabe, K. The role of lidocaine in cancer progression and patient survival. Pharmacol. Ther. 2024, 259, 108654. [Google Scholar] [CrossRef]
  252. Sui, Q.; Peng, J.; Han, K.; Lin, J.; Zhang, R.; Ou, Q.; Qin, J.; Deng, Y.; Zhou, W.; Kong, L.; et al. Voltage-gated sodium channel Nav1.5 promotes tumor progression and enhances chemosensitivity to 5-fluorouracil in colorectal cancer. Cancer Lett. 2021, 500, 119–131. [Google Scholar] [CrossRef] [PubMed]
  253. Brummelhuis, I.S.; Fiascone, S.J.; Hasselblatt, K.T.; Frendl, G.; Elias, K.M. Voltage-Gated Sodium Channels as Potential Biomarkers and Therapeutic Targets for Epithelial Ovarian Cancer. Cancers 2021, 13, 5437. [Google Scholar] [CrossRef] [PubMed]
  254. Reddy, J.P.; Dawood, S.; Mitchell, M.; Debeb, B.G.; Bloom, E.; Gonzalez-Angulo, A.M.; Sulman, E.P.; Buchholz, T.A.; Woodward, W.A. Antiepileptic drug use improves overall survival in breast cancer patients with brain metastases in the setting of whole brain radiotherapy. Radiother. Oncol. 2015, 117, 308–314. [Google Scholar] [CrossRef] [PubMed]
  255. Ramirez, M.F.; Tran, P.; Cata, J.P. The effect of clinically therapeutic plasma concentrations of lidocaine on natural killer cell cytotoxicity. Reg. Anesth. Pain Med. 2015, 40, 43–48. [Google Scholar] [CrossRef] [PubMed]
  256. Qing, X.; Dou, R.; Wang, P.; Zhou, M.; Cao, C.; Zhang, H.; Qiu, G.; Yang, Z.; Zhang, J.; Liu, H.; et al. Ropivacaine-loaded hydrogels for prolonged relief of chemotherapy-induced peripheral neuropathic pain and potentiated chemotherapy. J. Nanobiotechnol. 2023, 21, 462. [Google Scholar] [CrossRef]
  257. Lee, H.; Kim, J.W.; Kim, D.K.; Choi, D.K.; Lee, S.; Yu, J.H.; Kwon, O.B.; Lee, J.; Lee, D.S.; Kim, J.H.; et al. Calcium Channels as Novel Therapeutic Targets for Ovarian Cancer Stem Cells. Int. J. Mol. Sci. 2020, 21, 2327. [Google Scholar] [CrossRef] [PubMed]
  258. Vecera, L.; Gabrhelik, T.; Prasil, P.; Stourac, P. The role of cannabinoids in the treatment of cancer. Bratisl. Lek. Listy 2020, 121, 79–95. [Google Scholar] [CrossRef]
  259. McKallip, R.J.; Nagarkatti, M.; Nagarkatti, P.S. Δ-9-Tetrahydrocannabinol enhances breast cancer growth and metastasis by suppression of the antitumor immune response. J. Immunol. 2005, 174, 3281–3289. [Google Scholar] [CrossRef]
  260. Zhang, Y.; Cruickshanks, N.; Yuan, F.; Wang, B.; Pahuski, M.; Wulfkuhle, J.; Gallagher, I.; Koeppel, A.F.; Hatef, S.; Papanicolas, C.; et al. Targetable T-type Calcium Channels Drive Glioblastoma. Cancer Res. 2017, 77, 3479–3490. [Google Scholar] [CrossRef]
  261. Shi, J.; Chen, C.; Ju, R.; Wang, Q.; Li, J.; Guo, L.; Ye, C.; Zhang, D. Carboxyamidotriazole combined with IDO1-Kyn-AhR pathway inhibitors profoundly enhances cancer immunotherapy. J. Immunother. Cancer 2019, 7, 246. [Google Scholar] [CrossRef] [PubMed]
  262. Teisseyre, A.; Palko-Labuz, A.; Sroda-Pomianek, K.; Michalak, K. Voltage-Gated Potassium Channel Kv1.3 as a Target in Therapy of Cancer. Front. Oncol. 2019, 9, 933. [Google Scholar] [CrossRef]
  263. Hausmann, D.; Hoffmann, D.C.; Venkataramani, V.; Jung, E.; Horschitz, S.; Tetzlaff, S.K.; Jabali, A.; Hai, L.; Kessler, T.; Azoŕin, D.D.; et al. Autonomous rhythmic activity in glioma networks drives brain tumour growth. Nature 2023, 613, 179–186. [Google Scholar] [CrossRef]
  264. Ohya, S.; Kajikuri, J.; Endo, K.; Kito, H.; Elboray, E.E.; Suzuki, T. Ca2+-activated K+ channel KCa1.1 as a therapeutic target to overcome chemoresistance in three-dimensional sarcoma spheroid models. Cancer Sci. 2021, 112, 3769–3783. [Google Scholar] [CrossRef] [PubMed]
  265. Ohya, S.; Kajikuri, J.; Endo, K.; Kito, H.; Matsui, M. KCa1.1 K+ Channel Inhibition Overcomes Resistance to Antiandrogens and Doxorubicin in a Human Prostate Cancer LNCaP Spheroid Model. Int. J. Mol. Sci. 2021, 22, 13553. [Google Scholar] [CrossRef]
  266. Chen, S.; Cui, W.; Chi, Z.; Xiao, Q.; Hu, T.; Ye, Q.; Zhu, K.; Yu, W.; Wang, Z.; Yu, C.; et al. Tumor-associated macrophages are shaped by intratumoral high potassium via Kir2.1. Cell Metab. 2022, 34, 1843–1859.e11. [Google Scholar] [CrossRef] [PubMed]
  267. Chen, S.; Su, X.; Mo, Z. KCNN4 is a Potential Biomarker for Predicting Cancer Prognosis and an Essential Molecule that Remodels Various Components in the Tumor Microenvironment: A Pan-Cancer Study. Front. Mol. Biosci 2022, 9, 812815. [Google Scholar] [CrossRef] [PubMed]
  268. Matsui, M.; Kajikuri, J.; Kito, H.; Endo, K.; Hasegawa, Y.; Murate, S.; Ohya, S. Inhibition of Interleukin 10 Transcription through the SMAD2/3 Signaling Pathway by Ca2+-Activated K+ Channel KCa3.1 Activation in Human T-Cell Lymphoma HuT-78 Cells. Mol. Pharmacol. 2019, 95, 294–302. [Google Scholar] [CrossRef] [PubMed]
  269. Ohya, S.; Matsui, M.; Kajikuri, J.; Kito, H.; Endo, K. Downregulation of IL-8 and IL-10 by the Activation of Ca2+-Activated K+ Channel KCa3.1 in THP-1-Derived M2 Macrophages. Int. J. Mol. Sci. 2022, 23, 8603. [Google Scholar] [CrossRef] [PubMed]
  270. Medyouni, G.; Vörös, O.; Jusztus, V.; Panyi, G.; Vereb, G.; Szöőr, Á.; Hajdu, P. Inhibition of K+ Channels Affects the Target Cell Killing Potential of CAR T Cells. Cancers 2024, 16, 3750. [Google Scholar] [CrossRef]
  271. Yang, L.; Ye, D.; Ye, W.; Jiao, C.; Zhu, L.; Mao, J.; Jacob, T.J.; Wang, L.; Chen, L. ClC-3 is a main component of background chloride channels activated under isotonic conditions by autocrine ATP in nasopharyngeal carcinoma cells. J. Cell. Physiol. 2011, 226, 2516–2526. [Google Scholar] [CrossRef] [PubMed]
  272. Lyons, J.C.; Ross, B.D.; Song, C.W. Enhancement of hyperthermia effect in vivo by amiloride and DIDS. Int. J. Radiat. Oncol. Biol. Phys. 1993, 25, 95–103. [Google Scholar] [CrossRef]
  273. DeBin, J.A.; Strichartz, G.R. Chloride channel inhibition by the venom of the scorpion Leiurus quinquestriatus. Toxicon 1991, 29, 1403–1408. [Google Scholar] [CrossRef]
  274. Qin, C.; He, B.; Dai, W.; Lin, Z.; Zhang, H.; Wang, X.; Wang, J.; Zhang, X.; Wang, G.; Yin, L.; et al. The impact of a chlorotoxin-modified liposome system on receptor MMP-2 and the receptor-associated protein ClC-3. Biomaterials 2014, 35, 5908–5920. [Google Scholar] [CrossRef]
  275. Wang, G.X.; Hatton, W.J.; Wang, G.L.; Zhong, J.; Yamboliev, I.; Duan, D.; Hume, J.R. Functional effects of novel anti-ClC-3 antibodies on native volume-sensitive osmolyte and anion channels in cardiac and smooth muscle cells. Am. J. Physiol. Heart Circ. Physiol. 2003, 285, H1453–H1463. [Google Scholar] [CrossRef]
  276. Liu, J.; Zhang, D.; Li, Y.; Chen, W.; Ruan, Z.; Deng, L.; Wang, L.; Tian, H.; Yiu, A.; Fan, C.; et al. Discovery of bufadienolides as a novel class of ClC-3 chloride channel activators with antitumor activities. J. Med. Chem. 2013, 56, 5734–5743. [Google Scholar] [CrossRef]
  277. Li, S.; Wang, Z.; Geng, R.; Zhang, W.; Wan, H.; Kang, X.; Guo, S. TMEM16A ion channel: A novel target for cancer treatment. Life Sci. 2023, 331, 122034. [Google Scholar] [CrossRef]
  278. Shi, S.; Bai, X.; Ji, Q.; Wan, H.; An, H.; Kang, X.; Guo, S. Molecular mechanism of ion channel protein TMEM16A regulated by natural product of narirutin for lung cancer adjuvant treatment. Int. J. Biol. Macromol. 2022, 223, 1145–1157. [Google Scholar] [CrossRef] [PubMed]
  279. Bai, X.; Cheng, Y.; Wan, H.; Li, S.; Kang, X.; Guo, S. Natural Compound Allicin Containing Thiosulfinate Moieties as Transmembrane Protein 16A (TMEM16A) Ion Channel Inhibitor for Food Adjuvant Therapy of Lung Cancer. J. Agric. Food Chem. 2023, 71, 535–545. [Google Scholar] [CrossRef] [PubMed]
  280. Barron, T.; Yalçın, B.; Su, M.; Byun, Y.G.; Gavish, A.; Shamardani, K.; Xu, H.; Ni, L.; Soni, N.; Mehta, V.; et al. GABAergic neuron-to-glioma synapses in diffuse midline gliomas. Nature 2025, 639, 1060–1068. [Google Scholar] [CrossRef] [PubMed]
  281. Cornwell, A.C.; Tisdale, A.A.; Venkat, S.; Maraszek, K.E.; Alahmari, A.A.; George, A.; Attwood, K.; George, M.; Rempinski, D.; Franco-Barraza, J.; et al. Lorazepam Stimulates IL6 Production and Is Associated with Poor Survival Outcomes in Pancreatic Cancer. Clin. Cancer Res. 2023, 29, 3793–3812. [Google Scholar] [CrossRef] [PubMed]
  282. Montégut, L.; Derosa, L.; Messaoudene, M.; Chen, H.; Lambertucci, F.; Routy, B.; Zitvogel, L.; Martins, I.; Kroemer, G. Benzodiazepines compromise the outcome of cancer immunotherapy. Oncoimmunology 2024, 13, 2413719. [Google Scholar] [CrossRef]
  283. Pomeranz Krummel, D.A.; Nasti, T.H.; Kaluzova, M.; Kallay, L.; Bhattacharya, D.; Melms, J.C.; Izar, B.; Xu, M.; Burnham, A.; Ahmed, T.; et al. Melanoma Cell Intrinsic GABA(A) Receptor Enhancement Potentiates Radiation and Immune Checkpoint Inhibitor Response by Promoting Direct and T Cell-Mediated Antitumor Activity. Int. J. Radiat. Oncol. Biol. Phys. 2021, 109, 1040–10535. [Google Scholar] [CrossRef]
  284. Song, C.; Choi, S.; Oh, K.B.; Sim, T. Suppression of TRPM7 enhances TRAIL-induced apoptosis in triple-negative breast cancer cells. J. Cell. Physiol. 2020, 235, 10037–10050. [Google Scholar] [CrossRef]
  285. Nadolni, W.; Immler, R.; Hoelting, K.; Fraticelli, M.; Ripphahn, M.; Rothmiller, S.; Matsushita, M.; Boekhoff, I.; Gudermann, T.; Sperandio, M.; et al. TRPM7 Kinase Is Essential for Neutrophil Recruitment and Function via Regulation of Akt/mTOR Signaling. Front. Immunol. 2020, 11, 606893. [Google Scholar] [CrossRef]
  286. Wong, R.; Gong, H.; Alanazi, R.; Bondoc, A.; Luck, A.; Sabha, N.; Horgen, F.D.; Fleig, A.; Rutka, J.T.; Feng, Z.P.; et al. Inhibition of TRPM7 with waixenicin A reduces glioblastoma cellular functions. Cell Calcium 2020, 92, 102307. [Google Scholar] [CrossRef]
  287. Zhang, H.; Qian, D.Z.; Tan, Y.S.; Lee, K.; Gao, P.; Ren, Y.R.; Rey, S.; Hammers, H.; Chang, D.; Pili, R.; et al. Digoxin and other cardiac glycosides inhibit HIF-1alpha synthesis and block tumor growth. Proc. Natl. Acad. Sci. USA 2008, 105, 19579–19586. [Google Scholar] [CrossRef] [PubMed]
  288. Haux, J.; Klepp, O.; Spigset, O.; Tretli, S. Digitoxin medication and cancer; case control and internal dose-response studies. BMC Cancer 2001, 1, 11. [Google Scholar] [CrossRef]
  289. Menger, L.; Vacchelli, E.; Adjemian, S.; Martins, I.; Ma, Y.; Shen, S.; Yamazaki, T.; Sukkurwala, A.Q.; Michaud, M.; Mignot, G.; et al. Cardiac glycosides exert anticancer effects by inducing immunogenic cell death. Sci. Transl. Med. 2012, 4, 143ra199. [Google Scholar] [CrossRef] [PubMed]
  290. O’Neill, J.P.; Velalar, C.N.; Lee, D.I.; Zhang, B.; Nakanishi, T.; Tang, Y.; Selaru, F.; Ross, D.; Meltzer, S.J.; Hussain, A. Thapsigargin resistance in human prostate cancer cells. Cancer 2006, 107, 649–659. [Google Scholar] [CrossRef] [PubMed]
  291. Denmeade, S.R.; Mhaka, A.M.; Rosen, D.M.; Brennen, W.N.; Dalrymple, S.; Dach, I.; Olesen, C.; Gurel, B.; Demarzo, A.M.; Wilding, G.; et al. Engineering a prostate-specific membrane antigen-activated tumor endothelial cell prodrug for cancer therapy. Sci. Transl. Med. 2012, 4, 140ra186. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Classification and major subtypes of ion channels. Ion channels enable passive ion movement along electrochemical gradients and are categorized by their ion selectivity and gating mechanisms. The intensity of an ion current through a channel depends on its conductance and the electrochemical driving force, the difference between the membrane potential and the ion’s Nernst equilibrium potential. At rest, the driving force is strongly negative for sodium and calcium ions, so channel opening causes influx of these ions into the cell. For potassium, the driving force is positive, leading to efflux of these ions upon channel opening. Chloride ion flow direction through chloride channels may be inward or outward, depending on its transmembrane concentration gradient. These ion movements are critical for electrical signaling and cellular function.
Figure 1. Classification and major subtypes of ion channels. Ion channels enable passive ion movement along electrochemical gradients and are categorized by their ion selectivity and gating mechanisms. The intensity of an ion current through a channel depends on its conductance and the electrochemical driving force, the difference between the membrane potential and the ion’s Nernst equilibrium potential. At rest, the driving force is strongly negative for sodium and calcium ions, so channel opening causes influx of these ions into the cell. For potassium, the driving force is positive, leading to efflux of these ions upon channel opening. Chloride ion flow direction through chloride channels may be inward or outward, depending on its transmembrane concentration gradient. These ion movements are critical for electrical signaling and cellular function.
Pharmaceuticals 18 01521 g001
Table 1. Ion channels as pharmacological targets in cancer.
Table 1. Ion channels as pharmacological targets in cancer.
CompoundsTargetCancer TypeBiological EffectMolecular MechanismRef.
LidocaineNaV1.5OvarianEMT ↓, Sensitivity to cisplatin ↑FAK/Paxillin pathway ↓[18]
LidocaineNaV1.5GlioblastomaMigration ↓, Sensitivity to TMZ ↑HGF/MET pathway ↓[19]
JZTX-14NaV1.5BreastMetastasis ↓-[20]
PhenytoinNaV1.5BreastMigration ↓, Invasion ↓-[21]
RopivacaineVGSCsHepatocellular carcinomaApoptosis ↑IL-6/STAT3 pathway ↓[22]
JZTX-14NaV1.5BreastMigration ↓, Invasion ↓E-cadherin ↑, N-cadherin ↓, vimentin ↓, MMP2 ↓[23]
VeratridineNaV1.5ColorectalSensitivity to 5-Fu ↑Ca2+/calmodulin-dependent Ras signaling ↑[24]
HNTX-IIINaV1.7ProstateInvasion ↓Rho signaling pathway ↓[25]
NESO-pAbnNaV1.5BreastInvasion ↓-[24]
BepridilVGCCsOvarianCell viability ↓, EMT ↓-[26]
BepridilVGCCsChronic lymphocytic leukemiaApoptosis ↑NOTCH1 pathway ↓[27]
DiltiazemVGCCsBreastEMT ↓GDF-15 ↑[28]
KTt-45VGCCsCervicalApoptosis ↑Caspase-3 and caspase-9 activation[29]
BI-749327TRPC6BreastStemness ↓Myc ↑[30]
CannabinoidsTRP channelHepatocellular carcinomaAutophagy ↑AMPK activation[31]
CannabinoidsTRP channelLungInvasion ↓ICAM-1 ↑, TIMP-1 ↑[32]
mAb82TRPV6ProstateApoptosis ↑-[33]
LacidipineCaV1.2/1.3BreastAntitumor immunity ↑IDO1 ↓[34]
Lercanidipine/amlodipineL-type Ca2+ channelGastricSensitize to doxorubicin ↑YY1/ERK/TGF-β ↓[35]
MibefradilCaV3.2OvarianApoptosis ↑p-Akt ↓, FOXM1 ↓, survivin ↓[36]
MibefradilCaV3.2GlioblastomaApoptosis ↑mTORC2/Akt pathway ↓[37]
SKF96365ORAI1HepatocarcinomaAutophagy ↑, Chemosensitivity to 5-Fu ↑PI3K/AKT/mTOR pathway ↓[38]
1B50-1α2δ1Small-cell lung cancerChemosensitivity ↑ERK ↓[39]
AMG9810TRPV1CervicalSensitivity to cisplatin ↑EGFR/AKT pathway ↓[40]
Compound D9TRPM2LungTherapeutic efficacy of osimertinib ↑EGFR pathway ↓[41]
NifedipineVGCCsColorectalProliferation ↓, Metastasis ↓, Immune escapePD-1/PD-L1 ↓[42]
FS48KV1.1/KV1.3LungMigration ↓, Invasion ↓MMP-9 ↓, TIMP-1 ↑[43]
KAaH1KV1.1/KV1.3GlioblastomaAdhesion ↓, Migration ↓-[44]
PAP-1/PAPTP/PCARBTPKV1.3MelanomaROS-induced apoptosis ↑, Sensitivity to cisplatin ↑-[45]
PAPTP/PCARBTPKV1.3Pancreas ductal adenocarcinomaCell viability ↓p38 phosphorylation ↑[46]
MemantineKV1.3Acute lymphoblastic leukemiaApoptosis ↑AKT and ERK1/2 signaling ↓[47]
GenisteinKV1.3BreastProliferation ↓EROD and ODC activity ↓[48]
GenisteinKV1.3ColonApoptosis ↑Bax ↑, p21WAF1 ↑[49]
Chloroquine KV10.1BreastMigration ↓-[50]
AstemizoleKV10.1CervicalAutophagy ↑-[51]
Procyanidin B1KV10.1LiverProliferation ↓, Migration ↓-[52]
TetrandrineKV10.1CervicalProliferation ↓-[53]
Nutlin-3KV10.1CervicalCell viability ↓, Migration ↓, Invasion ↓PI3K/AKT pathway ↓[54]
CelastrolKV11.1ProstateApoptosis ↑-[55]
BerberineKV11.1OvarianProliferation ↓, Migration ↓, Invasion ↓-[56]
ResveratrolKV11.1ColonProliferation ↓, Apoptosis ↑-[57]
ClarithromycinKV11.1ColorectalApoptosis ↑PI3K/AKT pathway ↓[58]
PaxillineBKCaGlioblastomaMigration ↓, Sensitivity to cisplatin ↑-[59]
IberiotoxinBKCaNeuroblastomaProliferation ↓AKT1pser473 dephosphorylation ↑[60]
IberiotoxinBKCaEndometrialProliferation ↓, Migration ↓-[61]
IberiotoxinBKCaHepatocellular carcinomaMigration ↓E-cadherin ↑, Vimentin ↓[62]
Penitrem ABKCaBreastCell-cycle arrestp-AKT ↓, p-STAT3 ↓[63]
Trimebutine maleateVGCC/BKCaOvarianStemness ↓Wnt/β-catenin pathway ↓[64]
TemozolomideIKCaGlioma--[65]
VigabatrinIKCaGlioma--[66]
ClotrimazoleIKCaCervicalProliferation ↓-[67]
Triarylmethane-34KCa3.1Intrahepatic cholangiocarcinomaProliferation ↓, Invasion ↓NF-κB activation ↓[68]
PiperineIKCaProstateCell-cycle arrest, Proliferation ↓, Apoptosis ↑-[69]
MiconazoleSKCaMelanomaProliferation ↓-[70]
BL1249K2P2.1PancreaticProliferation ↓, Migration ↓-[71]
Withaferin ATASK-3BreastProliferation ↓-[72]
MinoxidilKir6.2OvarianTumor growth ↓-[73]
GlibenclamideKATPEndometrial adenocarcinomaProliferation ↓, Migration ↓-[74]
NPPBClC-3Nasopharyngeal carcinoma--[75]
ChlorotoxinClC-3GliomaInvasion ↓MMP-2 ↓[76]
CaCCinh-A01ANO1BreastCell viability ↓EGFR and CAMK signaling ↓[77]
Schisandrathera DANO1ProstateApoptosis ↑Caspase-3 ↑[78]
Vitekwangin BANO1ProstateApoptosis ↑-[79]
Cis-resveratrolANO1ProstateProliferation ↓, Migration ↓Caspase-3 ↑[80]
MatairesinosideANO1LungMigration ↓, Invasion ↓, Apoptosis ↑-[81]
DaidzeinANO1LungCell viability ↓, Migration ↓, Cell-cycle arrest ↑-[82]
ZafirlukastANO1LungProliferation ↓, Migration ↓-[83]
ML-SI1TRPML1BreastFerroptosis ↑, Stemness ↓, Sensitivity to doxorubicin ↑-[84]
OuabainNa+/K+ ATPaseBreastCell lysis ↑-[85]
OuabainNa+/K+ ATPaseBreastProliferation ↑, Motility ↑, Invasion ↑p-Rac/cdc42 ↓, p-ERK1/2 ↓[86]
DMSPMCABreastApoptosis ↑-[87]
KB-R7943NCXOvarianSensitivity to cisplatin ↑ -[88]
KB-R7943NCXMedulloblastomaApoptosis ↑-[89]
AmilorideNHEProstateSensitivity to lapatinib ↑-[90]
AmilorideNHEMultiple myelomaApoptosis ↑p53 signaling ↑[91]
CariporideNHEBreastApoptosis ↑MDR1 ↑, Cleaved caspase 3/9 ↑[92]
BumetanideNKCC1ColonVascularity ↓CD31 ↓, VEGF ↓[93]
BumetanideNKCC1GliomaMigration ↓, Invasion ↓-[94]
Arrows indicate the following: ↑, upregulation/activation; ↓, downregulation/inhibition.
Table 2. Clinical trials of anticancer drugs targeting ion channels.
Table 2. Clinical trials of anticancer drugs targeting ion channels.
ClinicalTrials.gov ID.PhaseAgentsStatusCancer TypesEnd PointsRef.
NCT01916317IIILidocaineCompletedBreastDFS and OS[95]
NCT02786329ILidocaineCompletedColorectalRecurrence rate[15]
NCT04761614I RiluzoleCompletedColorectalSafety profile[96]
NCT01480050I MibefradilCompletedGliomasSafety and the maximum tolerated dose[16]
NCT01578564I SOR-C13CompletedAdvanced solid tumorsSafety/tolerability and pharmacokinetics[97]
NCT05629702IICannabinoidsRecruitingRecurrent glioblastomaOS[98]
NCT02587819I nfP2X7-targeted antibodiesCompletedBasal cell carcinomaSafety, tolerability and pharmacokinetics[99]
NCT01107522IBCarboxyamidotriazoleCompletedRecurrent and newly diagnosed glioblastomaSafety profile[100]
CTR20160395IIICarboxyamidotriazoleCompletedAdvanced non-small-cell lung cancer PFS[101]
NCT01056029IMipsagarginCompletedAdvanced solid tumorsSafety, the maximum tolerated dose and pharmacokinetics[102]
NCT01777594IIMipsagarginCompletedProgressive advanced hepatocellular carcinomaResponse rate, PFS, OS[103]
-ICarboxyamidotriazoleCompletedRefractory solid tumorsPharmacokinetic analysis[17]
-IICarboxyamidotriazoleCompletedRelapsed epithelial ovarian cancerSafety, tolerability and pharmacokinetics[104]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhou, S.; Song, X.; Zeng, W.; Chen, D. Targeting Ion Channels for Cancer Therapy: From Pathophysiological Mechanisms to Clinical Translation. Pharmaceuticals 2025, 18, 1521. https://doi.org/10.3390/ph18101521

AMA Style

Zhou S, Song X, Zeng W, Chen D. Targeting Ion Channels for Cancer Therapy: From Pathophysiological Mechanisms to Clinical Translation. Pharmaceuticals. 2025; 18(10):1521. https://doi.org/10.3390/ph18101521

Chicago/Turabian Style

Zhou, Sha, Xiong Song, Weian Zeng, and Dongtai Chen. 2025. "Targeting Ion Channels for Cancer Therapy: From Pathophysiological Mechanisms to Clinical Translation" Pharmaceuticals 18, no. 10: 1521. https://doi.org/10.3390/ph18101521

APA Style

Zhou, S., Song, X., Zeng, W., & Chen, D. (2025). Targeting Ion Channels for Cancer Therapy: From Pathophysiological Mechanisms to Clinical Translation. Pharmaceuticals, 18(10), 1521. https://doi.org/10.3390/ph18101521

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

Article Metrics

Back to TopTop