Next Article in Journal
Robotic-Assisted Colonoscopy Platform with a Magnetically-Actuated Soft-Tethered Capsule
Next Article in Special Issue
Cancer and pH Dynamics: Transcriptional Regulation, Proteostasis, and the Need for New Molecular Tools
Previous Article in Journal
Targeting Signaling Pathways in Inflammatory Breast Cancer
Previous Article in Special Issue
Lysosome as a Central Hub for Rewiring PH Homeostasis in Tumors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

pH-Channeling in Cancer: How pH-Dependence of Cation Channels Shapes Cancer Pathophysiology

1
Institute of Physiology II, University Münster, 48147 Münster, Germany
2
Department of Biosciences, Biotechnologies, and Biopharmaceutics, University of Bari, 90126 Bari, Italy
3
Department of Radiology and Nuclear Medicine, Erasmus Medical Center, 3035 GD Rotterdam, The Netherlands
4
Percuros B.V., 2333 CL Leiden, The Netherlands
5
Department of Oncology CHUV, UNIL and Ludwig Cancer Center, 1011 Lausanne, Switzerland
*
Author to whom correspondence should be addressed.
Cancers 2020, 12(9), 2484; https://doi.org/10.3390/cancers12092484
Submission received: 16 July 2020 / Revised: 25 August 2020 / Accepted: 26 August 2020 / Published: 2 September 2020
(This article belongs to the Special Issue New Insights into Tumour pH Regulation)

Abstract

:
Tissue acidosis plays a pivotal role in tumor progression: in particular, interstitial acidosis promotes tumor cell invasion, and is a major contributor to the dysregulation of tumor immunity and tumor stromal cells. The cell membrane and integral membrane proteins commonly act as important sensors and transducers of altered pH. Cell adhesion molecules and cation channels are prominent membrane proteins, the majority of which is regulated by protons. The pathophysiological consequences of proton-sensitive ion channel function in cancer, however, are scarcely considered in the literature. Thus, the main focus of this review is to highlight possible events in tumor progression and tumor immunity where the pH sensitivity of cation channels could be of great importance.

Graphical Abstract

1. Introduction

The acidification of the tumor microenvironment (TME) is a prominent environmental driver of cancer progression. One major cause of acidification is a metabolic shift of tumor cells towards glycolysis and lactate production, ultimately leading to accumulation of metabolic acids in the tissue [1,2]. The interstitial acidosis can reach pH values of <6.5 in highly fibrotic tumors such as pancreatic cancer [3]. Thus, the concentration of free protons can be ten times higher than at physiological pH values of ~7.4. Acidosis in the TME is linked to tumor progression via a large variety of factors including among others: extracellular matrix remodeling; the promotion of invasion and metastasis by affecting cell adhesion; and it instigates many immunosuppressive processes, which result in a significant dampening of the immune response. These phenomena have been widely discussed in comprehensive reviews in multiple types of cancer, e.g., breast cancer [4,5], melanoma [6,7,8] and pancreatic cancer [9]. In contrast, in well-vascularized regions of a tumor sufficient perfusion is ensured to nurture the cancer cells as well as to wash out the produced metabolic acids. This contrast between differentially perfused areas of a tumor gives rise to pH gradients facilitating invasion of the surrounding tissues and metastasis.
The intracellular pH (pHi) of cancer cells is usually slightly (0.1–0.2 pH units) more alkaline than the extracellular pH (pHo) [10]. It is important to point out that this is only a relative intracellular alkalinization; in the poorly vascularized, highly acidic regions of a tumor (e.g., pHo 6.5), the pHi would still be acidic in absolute terms (e.g., pHi 6.7). The importance of pH in cellular physiology is straightforward as many proteins function in a pH-dependent manner having a very small spectrum of optimal pH. Prolonged acidification of pHi severely affects cell fate, as it inhibits cell metabolism, migration, proliferation and induces cell death. Therefore, it is vital for the cell to have high-sensitivity pH sensors in order to respond according to pH alterations using different effector proteins. Transmembrane proteins, such as ion channels, are optimal cellular sensors as they can be affected by both pHo and pHi [11]. Ion channels have a key role in cancer progression affecting every hallmark of cancer, as pointed out in numerous comprehensive reviews, e.g., [12,13,14,15]. However, our critical review of the literature revealed that most of the studies investigating the role of ion channels in cancer do not consider the pH landscape of tumors. Therefore, the present review specifically emphasizes the complex interplay between pH and individual ion channels in cancer pathophysiology. We propose multiple mechanisms of how ion channel function and interaction with the cell adhesion machinery in cancer cells and in tumor immunity can be affected by pH. We discuss possible pathophysiological consequences of pH-dependent ion channel activity and highlight the need for further studies in this direction. For limiting the scope of this review, we will only focus on cation channels for which functional data in tumors is available.

2. pH Homeostasis Is Tightly Coupled to Changes in Membrane Potential

The membrane potential is a characteristic feature of every cell and one of the most fundamental signaling modules. In 1971, Cone postulated that there is a functional correlation between the baseline (resting) membrane potential of cells and their mitotic activity [16]. Notably, the resting membrane potential of cancer cells ranges from −5 to −55 mV, whereas noncancer cells usually have a more negative resting membrane potential of −35 to −95 mV [17,18]. The more positive membrane potential is indeed beneficial for tumor progression, since it enhances tumor growth along with invasive and metastatic capabilities [17]. This concept is endorsed by the discovery that membrane hyperpolarization of tumor foci induced by the light-gated nonselective cation channel ChR2D156A—that passes mostly H+ at physiological pH or by the proton pump Arch,—suppresses tumor progression and induces tumor regression in zebrafish [18].
Changes in membrane potential are closely related to pHi alterations. Generally, cellular alkalization leads to membrane hyperpolarization, whereas acidification leads to depolarization. However, there are exceptions depending on the expression pattern of channels and transporters in the respective cells (see below). Acidification produces more pronounced effects on membrane potential than alkalization: acidification from pHi 7.4 to 6.5 leads to a depolarization of up to +20–+60 mV, whereas alkalization changes the membrane potential by −5 to −20 mV per unit of pH. This phenomenon is frequent in eukaryotic cells (reviewed in [19]), and has also been described in multiple human cancer cell lines [20,21]. Changes of pHo have no immediate extracellular effects on membrane potential, only after prolonged exposure, presumably because of subsequent intracellular acidification [22]. The central role of K+ channels in this phenomenon is revealed by the fact that the pH-dependent depolarization can be abolished with K+ channel inhibitors [19,21,23,24]. However, it should be noted that there are exceptions to this rule [25] and that changes of the pH homeostasis can also affect the potential of intracellular membranes such as that of the inner mitochondrial membrane [26,27,28]. Considering that depolarization frequently accompanies pHi acidification and vice versa, it is remarkable that depolarized cancer cells can maintain their pHi elevated compared to environmental pH. To understand this phenomenon, we need to discuss how cells handle surrounding H+. Overall, cells are facing a constant passive load of H+ towards the intracellular side of the plasma membrane [29]. The main reasons for this are the high membrane H+ permeability, as well as the electrochemical driving force for H+ pointing to the intracellular direction. The driving force of protons depends on their equilibrium (Nernst-) potential, which is about EH = −12 mV at physiological pH in somatic cells (pHi = 7.2, pHo = 7.4) [29]. Assuming a membrane potential of −70 mV, there is a driving force for H+ influx of −58 mV. As outlined in the previous section, cancer cells have a slightly more alkaline pHi compared to pHo [3], so the EH could become positive, generating a slightly higher chemical driving force for H+. However, as the resting membrane potential of tumor cells is highly depolarized compared to somatic cells, they would still have a reduced driving force for H+ influx. Thus, the depolarized membrane potential of cancer cells will reduce the electrochemical driving force for the extrusion of metabolic acid and thus help the cells to cope with the intracellular and environmental acid load.
The membrane potential of cells is temporally highly dynamic, and its changes accompany fundamental processes such as proliferation, migration, volume control, secretion and contractility [30,31]. Membrane potential and pHi changes go hand in hand throughout the life of a cell and contribute to the fine-tuning of cell proliferation, migration and metabolism:
(I).
The plasma membrane is hyperpolarized during DNA synthesis, in the S phase, and depolarized in the G2/M phase [32,33]. Interestingly, pHi has an alkaline peak of ΔpHi ≈ 0.2 in lymphocyte populations in the S phase of the cell cycle [34,35], whereas an acidification in G2/M has not been shown to date. Since pH changes alter microtubule stability, it has been proposed that pHi may be a “clock” for mitosis [36].
(II).
Migrating cells have a pHi gradient with the leading edge of the cells having a higher pHi [37,38]. Such local pH-microenvironments could also affect the function of ion channels involved in cell migration.
(III).
Glycolysis acidifies the intracellular milieu, whereas oxidative phosphorylation induces an alkalinization [26,27]. Aerobic glycolysis is prominent in the lamellipodia of multiple cancer cell types [39], which can act as a local source of metabolically produced acid during migration and further depolarize the plasma membrane. Additionally, the mitochondrial membrane potential depolarizes during glycolysis, and hyperpolarizes during oxidative phosphorylation [40,41]. It is conceivable that the plasma membrane potential is altered concordantly, as the mitochondrial membrane potential is in the phase with the plasma membrane potential [28]. Therefore, the enhanced aerobic glycolysis of cancer cells can contribute to their depolarized membrane potential.
In summary, eukaryotic cells alter their membrane potential and pHi harmoniously, and changes in each one of them can affect cancer invasiveness. Ion channels are key players in linking pHi to the membrane potential and vice versa, as they are sensitive sensors for both. By altering the activity of pH-sensitive ion channels in cancer cells, it could be possible to dysregulate the membrane potential of the tumor cells and render them inactive. Additionally, in case of tumor immunity, it would be critical to enhance the antitumor effects of immune cells in the acidic milieu of poorly perfused tumor areas. Therefore, to improve our understanding of tumor pathophysiology, it is vital to evaluate the individual families of ion channels and how they respond to changes in pH or, conversely, how they, themselves, modify pHi.

3. Ion Channels Are Modulated Directly and Indirectly by Changes in pH

Acidification of both pHi and pHo can modulate channel function through multiple mechanisms—outlined in Section 3.1. Moreover, acidification can indirectly modulate ion channel function, among others, through alteration of cell–extracellular matrix interactions, as described in Section 3.2.

3.1. Direct Interaction between Protons and Ion Channels

Proteins can coordinate H+ to specific proton-binding sites that involve titratable side chains with pKa values close to the physiological pH. Thus, the pHo-sensitivity of the voltage-gated potassium channel KV1.5 is attributed to a histidine residue located in the linker between S5 and the pore domain [42]. However, other residues (e.g., arginine and lysine) can form interaction sites for protons as well.
Moreover, H+ competes with other cation-binding sites and can thereby alter channel function. Competition arises primarily between divalent cations and H+. For example, Ca2+ ions modulate the pH-dependent gating of ASIC channels (detailed in Section 4.4.3). The apparent pH sensitivity of mammalian ASIC1 and ASIC3 is reduced with an increase in the extracellular Ca2+ concentration [43]. Ca2+ and H+ compete for ASICs, with Ca2+ binding favoring the closed state and H+ binding leading to the open state [44]. A similar competition between Ca2+ and H+ has been postulated in the case of the KV11.1 (hERG1) channel (Section 4.1.1). However, protons render this channel towards a closed state [45]. Lastly, competition between Zn2+ and H+ is well established for the HV1 proton channel [46].
Protons, similarly to Ca2+ [47], from the extracellular side can also provide an electrical shielding biasing the closed-open equilibrium of voltage-sensitive channels. This is known as the surface charge screening effect [48]. Thus, it is crucial to differentiate the screening effect from the specific inhibitory effects of protons. The change in surface potential following pHo acidification shifts the activation threshold of ion channels towards more depolarized voltages [49,50,51]. This indicates that in the hypoxic–acidic regions of solid tumors, surface charge screening could be relevant in inhibiting voltage-gated ion channels.

3.2. pH-Dependent Interaction of Cell Adhesion Proteins with Ion Channels

A principal mechanism for cells to crosstalk with the TME is via integrins. Integrins are transmembrane proteins formed by the noncovalent association of α and β subunits. So far, 18 α and 18 β subunits have been identified in mammals which can assemble in multiple combinations to form at least 24 different functional heterodimers with distinct specificities [52]. The conformation of each integrin and its binding to the ECM is regulated by the interaction of the cytoplasmatic domain with intracellular signaling and cytoskeletal proteins [53].
Integrins are more than just cell adhesion molecules, since they are able to transmit signals bidirectionally across the plasma membrane through “inside-out” and “outside-in” signaling, respectively. In the resting (“bent”) state, integrins are usually in the low affinity conformation and they are activated (“extended”) by the “inside-out” signaling pathway. Two intracellular activators (talin and kindlin) are crucial in this process, and the binding of talin to the cytoplasmatic tail of the β subunit is believed to be the final step of integrin activation [54]. Talin contains specific binding domains for the cytoplasmic tail of integrin, for filamentous actin and for other regulators of focal adhesions such as focal adhesion kinase (FAK) [55]. The binding of the integrin receptor to its extracellular ligands leads to the “outside-in” signaling that exerts crucial regulation of key cellular processes such as cell mobility, proliferation and differentiation [56].
Several integrins, namely α2β1, α5β1 or αvβ3, are pH-dependent in melanoma and other cell types [57,58,59]. Importantly, both pHi and pHo modify “inside-out” and “outside-in” signaling of integrins. Talin binding to actin filaments is pH-sensitive, and this association is decreased at a pHi > 7.2, allowing faster focal adhesion turnover and thus increased migration [60]. Migration and focal adhesion remodeling is also regulated by pH-dependent FAK autophosphorylation which is conferred by the protonation of His58 [61]. Furthermore, the ability of integrin receptors to bind to their extracellular ligands is regulated by pHo [10,60,62,63]. Low extracellular pH leads to increased cell adhesiveness, which can be explained by several possible mechanisms. First, conformational changes in acidic pHo increase the avidity of the integrin headpieces to ECM proteins [59]. Another mechanism could be a pH-dependence of the mechanical stability of focal adhesions [64]. Lastly, protons themselves could compete with divalent cations such as Ca2+ bound to integrin and thereby influence integrin activation.
Increasing evidence shows that ion channels [52,65,66,67] are able to form signaling complexes with cell adhesion proteins including integrins. Following their association, integrins and channels regulate their function bidirectionally [65]. They co-assemble in the plasma membrane and form supramolecular complexes that can recruit further cytosolic signaling proteins to orchestrate downstream intracellular signals [66]. These cytoplasmic messengers include Ca2+ or protein kinases [52,68]. Moreover, Ca2+ mediates the transmission of mechanical forces at focal adhesion sites triggered by integrins via the activation of FAK and c-Src [69].
Integrin–ion channel complexes coordinate several cellular processes relevant in cancer, namely adhesiveness, migration and differentiation [70,71,72]. By forming complexes with integrins, K+ channels regulate several proteins implicated in cell movement such as FAK, cortactin and integrins themselves [73,74,75]. It is well recognized that multiple Ca2+-activated and voltage-dependent K+ channels are involved in integrin-directed cell migration [76,77]. In prostate cancer, a KCa/αvβ3 integrin complex recruits FAK and promotes its phosphorylation, resulting in increased cancer cell proliferation [78]. β1 integrins can also bind to other K+ channels, such as KV1.3 and KV11.1, in tumor cells but not in cardiac cells [79,80,81,82]. Additionally, the KV1.3 channel involved in the immune response of T lymphocytes can be modulated by β1 integrins, which could be crucial for tumor immunity [83]. In colorectal cancer cells, the ability of KV11.1 to form a complex with β1 integrin correlates with increased in vitro invasiveness [84]. Moreover, in acute myeloid leukemic cells, the KV11.1–β1 complex couples with FLT-1, a receptor for vascular endothelial growth factor. This interplay is crucial for cell migration, triggers chemoresistance or modulates angiogenesis and cancer progression [81,85,86].
The interplay between integrins and ion channels also regulates cell differentiation. In the PC12 pheochromocytoma cell line, the shift to a neuronal phenotype caused by the neural cell adhesion molecules and N-cadherins is facilitated by voltage-gated Ca2+ channels [87]. Moreover, integrins seem to be important in the regulation of erythroleukemic cell differentiation and neurite extension in certain neuroblastoma cells through the activation of K+ channels [70,71].
Given the pH sensitivity of integrins and many ion channels, it is plausible to assume that their interaction itself may be pH-dependent as well. However, the experimental proof of whether and how the interaction, its kinetics and the downstream activity of these two classes of transmembrane proteins is indeed pH-dependent still remains to be established. One could for example mutate pH-sensitive residues on one protein class (e.g., an integrin) and observe the changes in expression, phosphorylation state, trafficking, intracellular signaling systems, binding capacity to the partner protein and downstream function of the ”partner” protein (e.g., an ion channel) both in vitro and in vivo. A study following a similar logic [37] could be performed to understand the role of the intratumoral and intercellular pH gradients in determining the specificity of integrin subunit interactions and the resulting downstream effects.

4. pH-Dependent Regulation of Ion Channels in Cancer Cells

As summarized by Figure 1 and detailed in Table 1, protonation by an acidic environment often results in a loss-of-function of cation channels involved in cancer cells and thereby to an inhibition of ion fluxes, similarly to what Prevarskaya et al. proposed [13]. This general phenomenon, however, shows a huge variability depending on how strongly pH-sensitive a channel is, ranging from almost insensitive to markedly pH-dependent. What is also noticeable from Table 1 is beyond some primarily proton-sensory channels (e.g., HV1 and K2P channels), there is limited information to date on how pHi modifies ion channel function as opposed to pHo. The section below briefly summarizes our knowledge about the pH-dependence of cation channels involved in cancer cells and details the involvement of the pH-sensitive function of cation channels in cancer pathogenesis and progression.

4.1. Modulation of K+ Channels in Cancer by pH

The tumor interstitial fluid can reach extreme potassium concentrations of 40 mm or higher [92]. This high [K+]o depolarizes the tumor cell membrane and increases the open probability of voltage-gated ion channels. On the other hand, the outward electrochemical gradient for K+ is largely decreased under such conditions which will lead to a low single-channel conductance for outward rectifying K+ channels. Furthermore, as shown in Table 1 and Figure 1, most cancer-relevant K+ channels are inhibited by pHo and/or pHi acidification. This is aggravated by the fact that the voltage-dependence of ion channels is generally shifted towards more positive membrane potentials in the case of an acidic pHo due to surface charge screening (discussed in Section 3.1). Together, these factors could mean that in poorly perfused regions of a tumor, K+ channels are largely nonfunctional. Along these lines, overexpression of K+ channels in tumors can be seen as a compensatory mechanism to enable at least some K+ efflux even in the unfavorable K+ gradient (similarly to lymphocytes, proposed in [49]). Therefore, it is likely that the sensory function of K+ channels could come into play when cancer cells steer from hypoxic areas to well-perfused regions, where the K+ and pH gradient are more favorable. For a deeper understanding of how K+ channels contribute to cancer progression in a pH-dependent manner, this section discusses the individual families of K+ channels.

4.1.1. KV Channels

Acidification of pHo potently inhibits KV channels, as demonstrated exhaustively in the literature [45,49,89,90,202]. However, the exact molecular mechanisms are so heterogeneous and often controversially discussed that it would warrant a separate systematic review.
The pHi-dependence of KV channels is more confusing and is considerably less well studied. Intracellular alkalinization leads to the activation of KV1.3 in lymphocytes [49], whereas KV channels are inhibited by alkaline pHi in canine pulmonary arterial smooth muscle cells [203]. Moreover, the KV11.1 currents are not affected by pHi when overexpressed in HEK293 cells [109]. This indicates that different KV channels are differentially regulated by pHi so that it is not yet possible to extrapolate the pHi dependence of KV channels onto cancer cells.
Another distinguishing feature of KV channels is that their voltage-dependent gating largely depends on the presence of charged amino acids in their transmembrane domains, e.g., arginine residues of the S4 domain [204]). Mutation of particular arginine residues in this domain creates a proton conducting pore through the voltage-sensor domain of KV channels (giving rise to so-called ω-currents). Since cancer cells carry a multitude of somatic mutations, such mutations could also be possible in cancer cells giving rise to functionally aberrant KV channels. Indeed, a recent study shows an enrichment of mutations in the S4 domain of KV7.x (KCNQ) channels in gastrointestinal cancers [205].
As indicated by Table 1, several pH-sensitive KV channels have been associated with tumors. Their function could be drastically modified along the heterogeneous pH-landscape of a solid tumor. However, the exact mechanism on how the pH-dependence manifests itself in cancer has yet to be experimentally discovered. One study assessing KV channels and pH in cancer cells was performed in the T-84 colon cancer cell line [103]. Here, the authors showed that inhibition of KV10.1 by astemizole in T-84 cells inhibits pHi regeneration after an NH4+ prepulse, which points to an inhibition of the H+ extrusion pathways upon K+ channel inhibition. Their study concludes that KV channels (namely KV1.5, KV3.4 and KV10.1) control T-84 cell proliferation by maintenance of pH regulation and Ca2+ signaling through hyperpolarization of the membrane potential. Furthermore, two studies about the KV1.5 channel show increased expression of KV1.5 in cancer cells treated with dichloracetate (DCA), accompanied by an increased pHi, and the generation of H2O2 that induces apoptosis in A549 and Hela cells [27,95].

4.1.2. KCa Channels

There are three major groups of KCa channels: the large conductance KCa1.1 (BKCa), the small conductance KCa2.1–2.3 (SKCa) and the intermediate conductance KCa3.1 (IKCa) channels. KCa3.1 channels have been linked to several tumor entities [206]. Some studies also revealed that KCa channels are pH-sensitive (summarized in Table 1). Intracellular acidification in the range of pHi 6.4–5.4 reduces channel open probability in C6 glioma cells of the rat [113]. Similar results were observed in pancreatic β-cells [207], and in HEK-293 cells overexpressing KCa3.1 [208]. Here, the authors also identified that changes in pHo in the range of pH 6.0 to 8.2 do not affect KCa3.1 current. Two other studies support the idea that intracellular acidification inhibits the activity KCa channels. In smooth muscle cells of the rabbit trachea, nearly no channel activity was observed even at slight acidification to pHi 7.0 [209]. In the ventricular membrane of choroid plexus, it was found that lowering of the pHi from 7.4 to 6.4 also reduced the channel open probability of Ca2+-activated K+ channels [210].
Certainly, most of the existing studies are not directly connected to cancer cells, but they show very similar effects to the findings of Strupp et al. on C6 glioma cells [113]. In conclusion, the present studies give evidence that KCa channel activity depends on the regulation by intracellular pH. It remains to be determined how this channel property modulates its impact on tumor pathophysiology.

4.1.3. Kir Channels

Some inwardly rectifying potassium channels (Kir) such as Kir1.1, Kir4.1, Kir4.2 and the heteromeric Kir4.x/Kir5.1 are sensitive to changes of the intracellular pH [211]. Kir 4.1, Kir 5.1, Kir 1.1 and Kir 2.3 channels are inhibited by lowering the pHi [212]. Several studies showed that Kir channels are also expressed in cancer cells, as in glioma cells [213], as well as in lung and prostate cancer cells [214]. These channels act primarily as tumor suppressors by hyperpolarizing the membrane potential of cancer cells. Notably, there is a strong correlation between Kir4.1 expression and the pathologic grade of astrocytic tumors [215]. However, to the best of our knowledge, there are no additional studies showing a link between the pH-dependent function of Kir channels in cancer cells. Yet, the importance of local regulation of Kir4.2 channels for migration illustrates how channel regulation by local pH microdomains might affect tumor cell behavior [216].

4.1.4. K2P Channels

K2P channels conduct outwardly directed background K+ currents, which are crucial for setting the resting membrane potential [133]. Eight members of the K2P family are regulated by changes in pHo and/or pHi (detailed in Table 1). Among these, K2P2.1 (TREK-1), K2P3.1 (TASK-1), K2P5.1 (TASK-2) and K2P9 (TASK-3) have already been confirmed to be functionally involved in solid tumors; while for other members of the family, only differential expression patterns have been observed to date, as summarized by [217].
K2P2.1 (TREK-1) is only active in the narrow physiological pH range [128]. Its current is potently inhibited by an extracellular acidification to pHo 6.9 [126] and alkalization above pHo 7.4 [128]. On the other hand, a similar degree of intracellular acidification, especially in the presence of lactate [127,218], activates the channel through the protonation of a conserved glutamate in the C-terminal domain [219]. As pHo and pHi acidification have opposite consequences to channel function, it would be important to investigate how a chronic interstitial acidosis also involving pHi acidification would affect channel function. The exact role of K2P2.1 in cancer development is still unclear. However, it has been postulated that K2P2.1 expression might be linked to different stages of cancer, and its regulatory effect on the membrane potential might be strongly connected with the progression of the cell cycle during cancer proliferation [141]. K2P2.1 overexpression is associated with a poor prognosis for the patients with prostate cancer, but not with ovarian cancer [125]. In both prostate [125] and ovarian cancer [220], the inhibition or knockdown of K2P2.1 inhibits proliferation by inducing cell cycle arrest at the G1/S checkpoint. On the contrary, in the pancreatic cancer cell line BxPC-3, a small-molecule activator of K2P2.1, BL-1249 has an inhibitory effect on cell proliferation and migration, observed at both pHo 6.7 and 8.5 [128]. This argues for a tumor-specific role of K2P2.1 channels in cancer. It remains to be seen whether this is due to potentially different pH landscapes in these tumor entities. K2P9 (TASK-3) and K2P3.1(TASK-1) are inhibited by an extracellular acidification through the protonation of conserved histidines [221,222,223]. K2P9 is inhibited by more acidic pHo (6.5–6.7) [222], while K2P3.1 is already partially inhibited at physiological pHo (7.2–7.4) [130,224]. K2P3.1 and K2P9 are important for apoptosis. Experiments on rat cerebellar granule neurons show decreased apoptosis due to the inhibition of these channels through high extracellular K+ (25 mm) and extracellular acidification. Similarly, K2P9 has been linked to the death of human glioma cells [129]. K2P9 and K2P3.1 have oncogenic potential for their ability to respond to pH and hypoxia. It has been postulated that either K2P9 or K2P3.1 are pH-sensors in glioma. K2P9 has been also found to be relevant for glioma cell death [129]. Prolonged interstitial acidification prevents glioma cells from passing through the hyperpolarization-dependent G1-to-S phase cell cycle [21]. Overexpression of K2P9 increases the resistance to hypoxia and apoptosis, enhancing its tumorigenic potential in embryonic fibroblast cells. Similarly, as discussed in case of KV channels, a possible mechanism could be that K2P9 overexpression compensates for acidic pH-induced channel inhibition. K2P9 contributes to proliferation when a loss-of-function mutation of K2P9, K2P9G95E, is introduced in a partially transformed mouse embryonic fibroblast cell line, C8 [130]. K2P9G95E expressing cells show tumor necrosis factor-induced apoptosis, followed by a loss of oncogenic properties inhibiting the tumor growth.
K2P5.1 (TASK-2), in contrast to K2P9 or K2P3.1, is almost completely closed in acidic pHo, and activated by extracellular and intracellular alkalization above the physiological pH (detailed in [133,221]). This channel plays a role in the proliferation of estrogen α receptor positive breast cancer cell. Additionally, the gene enhancer of K2P5.1 can bind to estrogen α receptor through its estrogen-responsive elements [131]. While there is yet no functional data on K2P5.1 in other types of cancer, its exquisite regulation by pH suggests that it could be relevant considering the alkaline pHi in tumor cells and points to important future studies.
In summary, evidence highlights the role that K2P channels have in the apoptotic process [129,130,220,222]. It is likely that there is a functional interplay between pH and K2P channels in cell cycle and apoptosis, where differential pH-sensitivity of the K2P channels could either induce or prevent cell death in tumors. While there is promising evidence in this respect for K2P2.1, K2P3.1 and K2P9, the full scale of the pH-dependence of K2P channels in cancer remains to be investigated.

4.2. TRP Channel in Tumors: Multimodal pH-Depencence

Transient receptor potential (TRP) channels are a large family of proteins often involved in the cellular sensing of external cues. Several TRP channels are sensitive to changes in intra- and extracellular pH (see Table 1 and Figure 1) [225]. Protons can activate (TRPV1, TRPV4, TRPC4, TRPC5) or inhibit the channels (TRPC6, TRPM2, TRPM7) (reviewed in: [143,156,226]). Their presence and function was also shown in multiple malignancies [13,227]. Here, we focus specifically on TRP channels which are pH sensitive and play a role in cancer progression.
TRPV1 mediates acute as well as chronic pain perception, also elicited by the acidic environment of cancer [228]. Extracellular protons can potentiate the effects of other stimuli (pH = 6.3) [151] and also directly open the channel (pH ≤ 5.9) [229]. Proton-induced modulations depend on glutamic acid and phenylalanine residues where the negative charge of Glu appears to be crucial for proton sensing [230,231]. Moreover, TRPV1 may itself cause cytoplasmic acidification, since in low pHo, the channel becomes highly permeable for protons [232].
The prognostic value of TRPV1 channels in cancer often depends on the cancer type, stage and two-sided nature of [Ca2+]i elevation, which can either induce cancer cell apoptosis or enhance proliferation [233,234,235]. In breast cancer tissues and cell lines, TRPV1 mRNA is highly overexpressed. Immunocytochemical staining demonstrated TRPV1 protein expression in SUM149PT cells, a model of triple-negative (PR, ER, HER2) breast cancer. Treating these cells with capsaicin, a TRPV1 activator, induces [Ca2+]i increase, reduced cell proliferation and migration, and leads to cell death [150]. Expression of functional TRPV1 was also observed in human hyperplastic prostate tissue and prostate cancer cell lines [236]. In agreement with channel activity being beneficial for disease outcome, a case report of a prostate cancer patient regularly ingesting capsaicin suggested that TRPV1 activation slows prostate-specific antigen increase [237]. It is noteworthy that tumor acidification favoring channel opening occurs predominantly in aggressive breast cancers [238] and late-stage prostate cancer [239]. This suggests that in these cases, the pH dependence of TRPV1 could be addressed therapeutically. However, the heterogeneity of breast and prostate cancer cells and possibility of channel desensitization, casts doubts on TRPV1 targeting in malignancies [240]. In other types of cancers, i.e., gastric and colorectal cancer, astrocytoma and bladder carcinoma, altered TRPV1 expression is also observed, but putative therapeutical benefits of the channel modulation still needs in vivo validation [152,153,241,242]. Lastly, TRPV1 is already considered as a therapeutic target in cancers because of its involvement in pain sensation. For example, chronic pain induced by acidosis in bone cancer can be diminished by the inhibition of TRPV1 [228].
TRPV4 is activated by H+ in a dose-dependent manner. Accordingly, TRPV4 can mediate acid-sensing at a low pH (pH < 6) [156]. Concerning the disrupted ionic composition and the high pressure of solid tumors, TRPV4 osmo- and mechanosensitivity is an additional important modulatory aspect. Activation and/or overexpression of TRPV4 induces normalization of the tumor vasculature and decreases tumor endothelial cell migration. The restored vascularization led to diminished tumor growth in a murine model of Lewis lung carcinoma [155]. In contrast, in breast cancer, TRPV4 activity promoted cancer cell invasion, transendothelial migration and metastasis. Thus, its expression is associated with poor prognosis [157,243]. Pharmacological inhibition of TRPV4 in patient-derived hepatocellular carcinoma cells leads to increased apoptosis [158] and in gastric cancer, TRPV4 activity augments oncogenic potential [159]. Thus, inhibition of the TRPV4 seems to be predominantly (but not exclusively) favorable for cancer treatment.
The activity of TRPC4 (β isoform) and TRPC5 is potentiated by low extracellular pH (~6.5). They can assemble into homo- or heteromultimers (also with TRPC1). Channel modulation shows biphasic characteristics, since further lowering of the pHo (<6.5) inhibits the currents [135]. Involvement of these channels in cancer development is rather poorly investigated; however, they may play a role in tumor angiogenesis, ovarian carcinoma proliferation, multidrug resistance and induce cancer cell death. Findings that cytotoxicity of (−)-englerin A on renal cell carcinoma cells is mediated by TRPC4/5, triggered promising new investigations. However, in vivo application will require the development of metabolically stable (−)-englerin A derivatives [134,244].
Nondestructive low pHo inhibits several TRP channels. TRPC6 lacks glutamate residues common for TRPC4 and TRPC5, so that it is inhibited at pHo ~6.5 and lower [135]. TRPC6 involvement in cancer is based on its Ca2+ permeability. Channel activity predominantly promotes cell proliferation and migration. TRPC6 expression was shown in multiple malignancies; however, its prognostic value is not well described. High TRPC6 expression is present in glioblastoma, prostate cancer and in esophageal squamous cell carcinoma where it correlates with poor prognosis [245]. Inhibition of TRPC6 by low pH of the tumor environment could be an advantageous factor, considering that this channel is also highly expressed in cancer-infiltrating immune cells (see Section 5.1) and in stromal cells such as pancreatic stellate cells [140].
TRPM2 is a Na+- and Ca2+-permeable channel. The inhibition of TRPM2 by low pH may be due to ion competition [246] or induced by H+ binding causing conformational changes in a state-dependent manner [247,248]. An intricate feature of this channel is its sensitivity to reactive oxygen species (ROS). In the context of cancer, oxidative stress is an important aspect but challenging to fully comprehend. Cancer cells often produce high amounts of ROS and at the same time they overexpress cellular antioxidants [249]. ROS accumulation may introduce mutagenic DNA damage, but conversely, oxidative stress is cytotoxic against malignant cells. TRPM2 expression in cancer cells was mostly reported to correlate with poor prognosis, as shown in neuroblastoma, squamous cells carcinoma, T-cell leukemia, breast, gastric, lung, pancreatic and prostate cancer cell lines and in patient samples. In all these cases, inhibition of TRPM2 displays antitumorigenic effects (reviewed in [250]). However, this general conclusion needs to be seen with a considerable amount of caution because of the frequent use of highly unspecific TRPM2 inhibitors (e.g., 2-APB, ACA). A recent study indicated that TRPM2 in cancer cells mediates H2O2-induced neutrophil cytotoxicity and the authors proposed that activation of the channel initiates cancer cell apoptosis [141]. Concerning the inhibitory action of low pH, the expression of the TRPM2 channel in immune cells rather than cancer cells might play a more decisive role (see below).
TRPM7 is among the most thoroughly examined TRP channels in cancer. TRPM7 is pivotal for Mg2+ homeostasis. At physiological membrane potentials, TRPM7 shows small inward current, which is potentiated by extracellular protons when pH is greatly reduced (pH < 6.0). Negatively charged glutamic acid residues serve as molecular proton sensors, similarly to TRPV1 [251,252].
TRPM7 is permeable for Mg2+, Ca2+ and monovalent ions. Mg2+ permeability is a distinguishing feature of this channel [253]. Cancer cells often accumulate these ions through TRPM7 activity [254]. TRPM7 overexpression in cancer was shown mainly in breast cancer but also in prostate and pancreatic cancer. In nasopharyngeal, lung and pancreatic cancer cells, TRPM7 promotes cell migration. ([149]). TRPM7 channel inhibition in cancer cells is almost unequivocally beneficial for disease outcome. However, the ubiquity of TRPM7 in healthy tissues makes it a difficult therapeutic target [150]. An alternative approach, but still challenging, would be to target the acidic tumor environment. Acidic pHi blocks the activity of TRPM7 [146], as does acidic pHo. However, the inhibition of TRPM7 by pHo appears to be dependent on the absence of Ca2+ and Mg2+ [255].

4.3. Voltage-Gated Proton Channel HV1

The voltage-gated proton channel HV1 is ubiquitously expressed in the human body. It has a very similar structure to the voltage-sensing domain of voltage-gated ion channels [256]. In the context of this review, it is highly relevant that HV1 channels can gate according to the pH gradient between the intracellular and the extracellular space (detailed in [257]). Because of their outward rectification, HV1 channels essentially only mediate outwardly directed H+ flux. The rate of H+ flow through HV1 is ~6000 H+ per second under physiological circumstances [258], which is high enough to cause an intracellular H+ depletion in the vicinity of the channel. In this way, HV1 can create localized alkaline pH microdomains within the cell [259] which in turn may steer cell polarity and actin cytoskeleton remodeling in comparison with Ca2+ microdomains [260]. Additionally, the voltage-dependence of the channel could be easily exploited by cancer cells which already have a depolarized membrane potential. Lastly, as HV1 is sensitive to mechanical stress [261], the channel may be more likely to open even without strong depolarizing signals in solid tumors that give rise to a multitude of mechanical stressors [262]. Taken together, HV1 can be a potent factor in acidifying solid tumors, as well as being a pro-invasive factor in well-perfused perivascular regions of a tumor and in metastatic cancer cells.
The potentially important role of HV1 in cancer is not sufficiently recognized in the literature, which is surprising as other H+ extruders—H+-carriers and pumps—are well established (summarized e.g., in [263,264,265]). The pathophysiological role of HV1 is mainly proton extrusion and intracellular pH recovery, which has been described in multiple colorectal cancer [162] and glioblastoma cell lines [163,164]. In these studies, the consequence of channel inhibition by Zn2+ or siRNA restrains intracellular pH-recovery, as well as decreases cell proliferation and transwell chemotaxis. In breast cancer, high mRNA expression levels correlate with an unfavorable prognosis and cancer progression [161]. In a recent study concerning the triple-negative breast cancer cell line MDA-MB-231, both HV1 knockdown (using short hairpin RNA) and knockout (by CRISPR-Cas9) were performed. HV1 knockout, but not knockdown, leads to a higher rate of glycolysis as well as decreased migration and H2O2 production. Xenograft tumor growth rate is decreased in tumors from HV1 KO cells, likely due to accompanying loss of the cell adhesion molecule CD171/LCAM-1 [160]. Additionally, in B-cell chronic lymphocytic leukemia, the high expression of the short variant of HV1 correlates with decreased overall survival [165]. Recurring mutations in the gene encoding HV1 (HVCN1), ranging from nonsense mutations to splice sites, were found in follicular lymphoma patients. These HV1 mutations are associated with a favorable prognosis in lymphoma patients [266]. As the two studies above are somewhat contradictory, it is necessary to conduct investigations in larger patient cohorts to clarify the prognostic role of HV1 variants.

4.4. Other Channels

4.4.1. NaV Channels

Voltage-gated Na+ channels (NaV) have similar modes of interaction with protons as voltage-gated K+ channels (Section 4.1.1). They can conduct protons themselves through their gating pore when a S4 arginine mutation is introduced [267]. In addition, NaV channels are generally inhibited by acidic pHo, although the pH-dependence of NaV1.6–1.9 has not been evaluated yet [268,269]. The most pHo-sensitive NaV channel is NaV1.5 [186,270], while the skeletal muscle isoform NaV1.4 is largely proton-insensitive [271]. How altered pHi affects human NaV channels is not known to date. Similarly, the physiological consequence of an intracellular acidosis on NaV channels in cancer remains to be elucidated. So far, it is only known that changes of pHi do not alter current amplitude, but slow inactivation parameters in NaV channels of frog skeletal muscle and invertebrate giant axons [272].
NaV channels are highly expressed in multiple types of carcinomas, including those of the colon, lung, breast, prostate, ovary and skin [273,274,275]. Primarily, NaV1.5 and NaV1.7 regulate the migratory and invasive properties of cancer cells [275,276]. In breast cancer cells, NaV activity leads to intracellular alkalinization and pericellular acidification. It has been proposed that NaV1.5 forms a functional complex with auxiliary β subunits and the Na+/-H+ exchanger NHE1 and thus affects pHi regulation [276,277]. However, as both NaV1.5 and NHE1 lead to an intracellular Na+ load, it is still unclear how the cells provide the necessary driving force for H+ efflux.
The neonatal splice variant of NaV1.5 is more resistant to extracellular acidosis than the adult NaV1.5: under acidic pHo conditions, voltage-dependent activation of the neonatal form is shifted less towards more depolarized voltages than that of the adult form [278]. This differential pHo-sensitivity may be relevant as the neonatal NaV1.5 is normally not detectable in adult tissues. It becomes functionally expressed in metastatic breast and colon cancer, as well as in neuroblastoma cells, promoting tissue invasion and in vivo metastasis [184,279,280,281]. It remains to be established whether NaV inhibitors with increasing activity at acidic pH, such as ranolazine, would have a beneficial effect in cancer therapy [282].

4.4.2. Piezo1

Since its characterization in 2010 [283], the Piezo channel family has been extensively studied, as Piezo1 and Piezo2 are key mechanosensors in the human body. They are ubiquitously expressed and are able to sense multiple forms of mechanical stimuli and transduce them into appropriate responses depending on the cell type. For example, Piezo1 is vital for sensing vascular shear stress and regulating of red blood cell volume, whereas Piezo2 is crucial for proprioception (for greater detail see [284,285,286]).
The Piezo1 channel enhances the invasive phenotype in multiple tumors, namely breast, colorectal and prostate cancers [193,287,288]. An electrophysiological study described that heterologously expressed Piezo1 channels are inhibited by extracellular protons [188]. These results have been confirmed and extended by our group. We showed channel inhibition also by an intracellular acidification in pancreatic stellate cells that are key stromal cells in pancreatic ductal adenocarcinoma [189]. According to our findings, the Piezo1 channel is likely to be in a nonconducting state in acidic, poorly perfused regions of solid tumors. This could protect the cells from cell death due to harmful Ca2+ overload. On the other hand, Piezo1 would be fully responsive to mechanical stimuli in well-perfused areas of the tumor, when it is relieved from the protonation-dependent channel inhibition.

4.4.3. Acid-Sensing Ion Channels (ASICs)

Acid-sensing ion channels (ASICs) are members of the degenerin/epithelial sodium channel (ENaC) family and are widely expressed in the nervous system [289]. They are sensitive to extracellular acidosis and are involved in acid-induced nociception. In humans, four genes (ASIC1-4) give rise to at least six distinct subunits, [290]. ASICs are sensitive to inhibition by amiloride and are permeable to Na+ to a much greater extent than to K+ and Ca2+.
For a long period, ASIC channels remained enigmatic ”brain” Na+ channels with known ENaC homology. The discovery of H+-gating of these channels propelled subsequent research. Low extracellular pH (pH ~6.5 for ASIC1a and ASIC3; ~6.1 for ASIC1b and ~4.5 for ASIC2a) induces channel opening. The current is transient and followed by rapid channel desensitization [291]. ASIC3 and some heteromers do not fully desensitize, showing small, sustained currents [292]. ASIC3 activation by H+ results from a competition of Ca2+ and multiple protons at the activation site of the channel in a manner that low [Ca2+]o (10 µM) allows channel gating at a pH close to physiological [43].
ASICs sense acidic pH in the tumor environment and are thereby involved in tumor progression [13]. The expression of functional ASIC1 was shown in breast cancer cells, where it induces acidosis-induced ROS production, cancer invasion and growth. The underlying mechanism is possibly based on ASIC1-mediated Ca2+ influx [4]. ASIC1 and ASIC3 are overexpressed in pancreatic cancer and support acidosis-induced epithelial–mesenchymal transition (EMT). In this instance, the authors also suggest the Ca2+ permeability of ASIC1/3 as a pivotal mechanism which leads to subsequent activation of RhoA signaling [194]. In colorectal cancer, ASIC2 promotes metastasis and its expression correlates with poor prognosis [195]. In a nonsmall cell lung cancer line, A549, several ASIC members are expressed and mediate cell proliferation and migration [196]. Overall, based on these few studies, the role of ASIC channels in an acidic tumor environment appears to depend mainly on the Ca2+ permeability of the channel.

4.4.4. P2X

Purinergic receptors are ATP-gated Na+-, K+- and Ca2+-permeable channels. There are seven receptor subunits described (P2X1-7), among which P2X1–4 are modulated by pHo [293]. pH-induced inhibition of P2X7 depends on acidification and duration of the low pHo exposure [198,294]. Antagonists of this channel are already used in therapy of neuropathic pain, and there is evidence of an inhibitory impact on cancer growth and metastasis [295]. In melanoma, colon cancer and PDAC cell lines, P2X7 activity promotes tumor progression [197,200]. P2X7 receptor is gaining more attention, especially due to its expression in tumor-infiltrating immune cells [197,296], detailed in Section 5.1. Whether or not the pH-sensitive P2X1–4 channels are involved in cancer development still remains unclear.

5. pH Dependence of Ion Channels in Tumor Immunity

One aspect which the literature has consistently highlighted in orchestrating a powerful immune reaction is that immune cells require intact ion channels in the cell membrane. Nondisruption of those ion channels is pivotal for the immune system’s ability to efficiently carry out its function of destroying cancer cells. This applies to cells of both the innate and adaptive immunity.

5.1. Innate Immunity

Cells of innate immunity make up an important part of the tumor milieu. Tumor-associated neutrophils and macrophages, dendritic cells (DC) and myeloid-derived suppressor cells (MDSC) are potential targets of immune-based therapies [297,298,299,300]. Although the innate immune response is generally associated with leukocytes of myeloid lineage, descendants of the lymphoid lineage such as natural killer (NK), natural killer T-cells (NKT) and γδ T cells also fall into this category [301]. A high number of myeloid cells in the TME and in the blood circulation predominantly correlates with poor prognosis [302]. In contrast, effector cells of lymphoid lineage (such as NK cells) display beneficial antitumorigenic characteristics [303]. However, different tumors and varied immunomodulatory microenvironments often require tailored approaches.
Low extracellular pH of solid tumors is an important modulator of the immune response. Several Ca2+ permeable, pH-sensitive ion channels are expressed in neutrophils, macrophages and DCs, including STIM/Orai, TRP channels and purinergic P2X receptors. These channels, detailed below and in Table 1, are generally inhibited by protons. Inhibition of the Ca2+ fluxes in the acidic tumor environment may, therefore, result in the retention of compromised (protumorigenic) immune cells in the tumor microenvironment and ultimately in disease aggravation.
Infiltrating neutrophils are unequivocally cellae non gratae in varied tumors [302,304]. Neutrophil migration and recruitment are based on CXCR2 signaling and require concurrent activity of the TRPC6 channel [139]. Inhibition of the TRPC6 channel by extracellular protons may hinder neutrophil chemotaxis. TRPM2, which is inhibited at low extra- and intracellular pH, and TRPM7, may be involved in neutrophil migration [305,306,307,308,309]. However, the underlying mechanisms and consequences of channel inhibition in neutrophils are not well described to date. Other neutrophil functions such as production of ROS and release of neutrophil extracellular traps also rely on undisturbed Ca2+ fluxes. Moreover, the activity of phagocyte NADPH oxidase NOX2 requires constant charge compensation which is provided by HV1 [310]. As mentioned in Section 4.3, HV1 opens when there is a considerable H+ gradient across the plasma membrane. Therefore, an extracellular acidification may inhibit this channel and disrupt NOX2 activity and hence, impair the killing capacity of neutrophils.
In the case of macrophages, plasticity plays an important role in the TME. Macrophage polarization into the so-called M2 type, is one of the mechanisms for rendering the tumor milieu immunosuppressive [311]. The pH-sensitive TRP channels, namely TRPM2, TRPM7 and TRPC1 are involved in modulating macrophage phenotype, and therefore, extracellular pH also indirectly controls this process [312,313,314].
The purinergic P2X7 receptor is expressed in a variety of immune cells in the TME. For instance, P2X7 is involved in phagocyte migration and ROS production [315]. Targeting P2X7 in inflammation-related diseases is already implemented in clinical trials [316]. In dendritic cells, STIM/Orai -mediated Ca2+ fluxes are also pivotal for presenting antigens. Inhibition of STIM/Orai channels in dendritic cells in the acidic TME may contribute to the fact that low pHo disrupts the activation of lymphocytes.
In addition, not-fully differentiated myeloid-derived suppressor cells (MDSCs) contribute to TME immune anergy. Two major populations resemble granulocytes and monocytes/macrophages and have strong immunosuppressive features [317]. In MDSCs, purinergic P2X7 mediates CCL2 release, macrophage recruitment and contributes to MDSC expansion [201]. In the case of hepatitis, TRPV1 activity induced or potentiated by acidic pHo stimulates MDSCs and results in the resolution of the inflammatory process [318]. Not much else is known about pH-sensitive ion channels expressed in MDSCs. It remains to be seen whether the immunosuppressive MDSC function can be overcome by channel modulation through alteration of the microenvironment.

5.2. Acquired Immunity

Ca2+ influx mediated by ion channels is a precondition for triggering lymphocyte metabolism, activation and a range of downstream signaling pathways [319].
Regarding metabolism, naïve T cells are in a “quiescent” state and dependent on oxidative phosphorylation as a source of energy [320]. Antigen-stimulated T cells shift their metabolism to aerobic glycolysis [321]. In this context, it is notable that T cell metabolism is also regulated by Ca2+ signaling through the Ca2+/calcineurin/NFAT pathway, which is involved in controlling the expression of many different elements involved in glycolysis, such as glucose transporters GLUT1 and GLUT3, glycolytic enzymes and transcription factors HIF1α, IRF4 and c-Myc [322]. There is a number of ways in which T cells can be inhibited in the acidic TME due to the fact that both cancer cells and activated T cells use aerobic glycolysis as their primary form of metabolism. For example, having high amounts of lactate production (from the cancer cells) in the TME prevents activated T cells from effluxing lactate thus blocking the glycolytic metabolism in the T cells [323]. Cancer cells and immune cells must, therefore, compete for the available glucose which is limited in the established tumor region. Whilst cancer cells are highly adaptable, T cells are unable to adapt and subsequently enter an anergic state [321,324].
T lymphocyte activation occurs when the T cell antigen receptors (TCRs) recognize and bind antigens, with this antigen–receptor engagement eliciting a response from the phospholipase Cγ (PLCγ)/inositol 1,4,5-triphosphate (IP3) pathway which provokes Ca2+ release from the endoplasmic reticulum (ER). The eventual depletion of those Ca2+ stores stimulates the opening of Ca2+-release activated Ca2+ channels (CRAC) in the plasma membrane [325,326]. The substantial increase in [Ca2+]i activates NFAT transcription complexes which are used to assemble and mediate the transcription of the crucial genes involved in T cell activation, such as IL-2. Rapid Ca2+ influxes, however, cause depolarization of the membrane potential, which, if not regulated, would act to prevent a further build-up of Ca2+. This is prevented by the counterbalancing efflux of potassium ions out of the cell. This K+ efflux, mainly mediated by the (pH-sensitive) voltage-gated KV1.3 and Ca2+-activated KCa3.1 channels, is essential in maintaining the hyperpolarized membrane potential state which is in turn pivotal in sustaining the Ca2+ influx [68,326].
As observed from the examples above, ion channels orchestrating the initiation, intensity and duration of the Ca2+ signal are critical for an effective immune response, also in the case of tumor immunity. With respect to the antigen–receptor engagement for T lymphocytes, the CRAC (ORAI-STIM) channels constitute the dominant pathway involved in mediating the store-operated Ca2+ entry (SOCE) regulated Ca2+ influx [188,327]. Opening of the CRAC channels occurs by the STIM proteins binding to the Orai subunits that form the Ca2+ selective conducting pore. Depending on the activation and the subtype of T lymphocytes, KV1.3 channels or KCa3.1 channels are the main channels involved in regulating the membrane potential in order to sustain SOCE [328].
The general pH-sensitivity of KV and KCa channels is detailed in Section 4.1, where we already pointed out that an acid-induced inhibition of the K+ conductance might be compensated by an upregulation in the respective channels (in lymphocytes), with the result that the net K+ conductance is maintained [49]. Consequently, one reason for K+ remaining in the T lymphocytes is due to this pH blockage of the efflux of K+. Additionally, the increased concentration of K+ in the interstitial fluid leads to an increased intracellular concentration of K+ in T cells. The elevated extracellular K+ concentration, in turn, is due to the fact that in both human and mouse tumors, cell death (necrosis or apoptosis) leads to the release of high levels of K+ ions into the interstitial fluid, with this action impairing the effector function of T cells [92]. The higher concentration of K+ in the T cells inhibits the transcription of the genes involved in stimulating the Akt-mTOR pathway (a central regulator of cell growth and metabolism) which is heavily involved in regulating the effector T cell function. Overexpressing KV1.3 in tumor-specific T cells lowers intracellular K+ concentrations by increasing K+ efflux into the interstitial fluid and thereby improves the functioning of the effector T cells by increasing Akt-mTOR activation and the production of IFN-γ [92].
CRAC channel activity is reduced by both intra- and extracellular acidic pH [329,330]. For the pH sensitivity of Orai1, Glu190 and Glu106 residues are pivotal when Na+ or Ca2+ are the main current contributors, respectively [166,168]. Many studies have shown that blocking/deleting Orai1 attenuates the SOCE process in T cells [331]. Age is associated with a marked decrease in the expression of the Orai and STIM proteins in the CD8+ T cells which reduces Ca2+ current through the Orai channels [332]. This decrease in Ca2+ signaling is linked to the resistance of cytotoxic lymphocyte (CTL) cytotoxicity efficiency. This can be an explanation for the preponderance of tumors with age. It also suggests a reduced efficiency of the antitumorigenic CD8+ response to tumor cells, if Ca2+ influx is consistently blocked by the acidic TME. Many of CTL’s mechanisms of killing tumor cells are dependent on Ca2+ flux [333]. A study proposed a bell-shaped dependence of CTL cytotoxicity on Ca2+, where partial inhibition of Orai1 in CTLs decreases Ca2+ signaling but improves the cytotoxicity of the T cells [181]. Whether such partial inhibition can be achieved therapeutically by fine-tuning the environmental pH of the TME is a subject of further studies.
Other pH-sensitive Ca2+ permeable channels such as members of the TRP family (detailed in Section 4.2) also contribute to Ca2+ influx in lymphocytes [325]. TRPM7 is important in T cell development and functioning [334]. TRPM7 channels can be impaired by the presence of protons. One study shows that TRPM7 channels, functionally cooperating with KCa3.1, are localized at the rear part (uropod) of migrating T lymphocytes and regulate cell migration. Inhibiting KCa3.1 and downregulating TRPM7 hinders T cell motility. It is suggested that this process is executed through the regulation of Ca2+ influx, with KCa3.1 maintaining the membrane potential and with the TRPM7 channel facilitating the influx of Ca2+ [335]. As acidic pHi and pHo inhibit TRPM7 function, the channel is not likely to be functional in lymphocytes residing in acidic areas of tumors, resulting in impaired cell migration. TRPC1 and TRPC3, which are involved in Ca2+ entry in T lymphocytes, have yet to be investigated in this aspect [336,337,338,339]. TRPV1 and TRPV4 are also involved in Ca2+ signaling in T cells [340]. Being activated by both extra- and intracellular protons, they would enable cation influx and possibly Ca2+ signaling in an acidic environment. Overall, the TRP family members contribute to many of the crucial pathways needed in T cell activation and function. Yet, it needs to be further studied how pH may affect their activity and function.
A particular focus in this section was on the role of ion channels in T lymphocytes. However, more information is needed on the different subpopulations of T cells affected by these channels. The expression of K+ channels—such as KV1.3 and KCa3.1-as well as CRAC and other ion channels—varies in different T lymphocyte types and cell subsets also depending on their state of activation. Similarly, evidence of the involvement of cation channels in the effective functioning of B lymphocytes is still scarce. Many of the ion channels discussed above show some evidence of overlapping in B cell function, namely the role of CRAC channels in Ca2+ influx [326]. The proton channel HV1 promotes B cell signaling, and B cells deficient in HV1 have impaired ROS generation, which in turn affects their metabolism, modulating mitochondrial respiration and glycolysis [341].
To summarize, dysregulated pH impairs T-cell activation and T-cell function. The acidic TME may block the CRAC and K+ channels in tumor-infiltrating immune cells. Diminished Ca2+ influx into the T cell in turn blocks the crucial dephosphorylation of NFAT, thereby preventing the production of several cytokines, such as IL-2 and IFN-γ, which are needed for effective T cell functioning [323]. Furthermore, intratumoral acidosis through lactate could prevent the expression of the pivotal proteins needed for the metabolism of activated T cells. In addition, there may be an overlap of metabolic dysfunctions, which are acting together with the acidic pHo to dampen the T cell function. Hypoxia-driven metabolic dysfunctions cause adenosine accumulation in the extracellular tumor microenvironment and together with acidic pHo could negatively affect T-cell function through the immunosuppressive effects of adenosine A2A receptors [342,343]. If this hypothesis is correct, then it could suggest an overlap with KCa3.1, which is adenosine sensitive and required for T cell motility. Along these lines, blockage of KCa3.1 could contribute to causing T cell anergy in the acidic pHo [119,120,342,343].

6. Conclusions

Studying pH and pH-dependent processes in cancer is a timely task. Its relevance becomes particularly evident when considering that some of the cancers with the worst prognosis originate from tissues with an extreme acid-based homeostasis. Prominent examples include gastric and pancreatic cancer whose host organs generate enormous acid-based fluxes after each meal that in turn cause marked changes in the interstitial pH. To understand how these alterations in pH can ultimately push cells into forming a malignant tumor and favor its progression is a prerequisite for developing new pH-oriented diagnostic as well as therapeutic options. To achieve this, a lot of groundwork is still required such as intense pH-imaging at the cellular, tissue and organism levels.
Such information is indispensable for decoding the role of pH-sensitive ion channels in the crosstalk between the cancer cells, cells of the tumor immunity and the components of the TME. Remarkably, the family of ion channels shows a large heterogeneity on how they respond to different pH. For example, most K+ channels are inhibited in an acidic environment typical for hypoxic, poorly vascularized areas inside a tumor, thereby affecting every cell type in this environment. Whether this acidic pH-dependent inhibition of channels results in T cell anergy and ineffective innate tumor immunity still remains an open question. On the other hand, cation flux is probably maintained by an upregulation of many acid-inhibited channels and by a number of channels that are activated by an acidic environment (for example, TRPV1). In this way, Ca2+ signals needed for cell migration to escape the unfavorable acidic environment could still be generated.
In summary, our review was aimed at promoting studies that deepen our understanding of the complex pH landscape of the tumor microenvironment and to understand the role of ion channels in cancer in a more holistic view. This is indeed one of the major conclusions of our review. There is still a large gap of knowledge in understanding the mutual interplay between ion channels in cancer and the pH landscape of solid tumors. The latter applies not only to the channels themselves but how their biophysical properties confer mutual dependencies. There is still a lot to learn about ion channels regulating the function of the different cell types that make up a tumor. Intriguingly, we frequently find the same channel types in different cells of the tumor microenvironment. For example, KCa3.1 channels are expressed not only in tumor cells but also in stromal and endothelial cells as well as in tumor-infiltrating immune cells such as lymphocytes, macrophages and neutrophils. Complex tumor models are needed in order to disclose the therapeutic benefits or side effects of such a “multiple-hit” ion channel-targeting therapy. While this is certainly a challenging endeavor, we should keep in mind that ~15% of all clinically prescribed drugs are already targeting ion transport proteins, many of which can be repurposed for use in cancer therapy. Larger inter-/national consortia should be created to advance this exciting research field further.

Funding

This work was funded by the European Marie Skłodowska-Curie Innovative Training Network (ITN) pH and Ion Transport in Pancreatic Cancer–pHioniC (Grant Agreement number: 813834; H2020-MSCA-ITN-2018). AS thanks the support from the Deutsche Forschungsgemeinschaft (DFG; SCHW 407/17-1 and SCHW 407/22-1; GRK 2515/1, Chembion) and IZKF Münster (Schw2/020/18).

Acknowledgments

The authors wish to thank past and present members of our laboratories whose enthusiastic work made it possible to develop the concepts described in this review. Finally, we would like to mention that this review helped us to keep our spirits up during the COVID-19 pandemic when work in our laboratories was severely restrained or impossible.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Swietach, P.; Vaughan-Jones, R.D.; Harris, A.L.; Hulikova, A. The chemistry, physiology and pathology of pH in cancer. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Swietach, P. What is pH regulation, and why do cancer cells need it? Cancer Metastasis Rev. 2019, 38, 5–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Pedersen, S.F.; Novak, I.; Alves, F.; Schwab, A.; Pardo, L.A. Alternating pH landscapes shape epithelial cancer initiation and progression: Focus on pancreatic cancer. Bioessays 2017, 39. [Google Scholar] [CrossRef] [PubMed]
  4. Gupta, S.C.; Singh, R.; Asters, M.; Liu, J.; Zhang, X.; Pabbidi, M.R.; Watabe, K.; Mo, Y.-Y. Regulation of breast tumorigenesis through acid sensors. Oncogene 2016, 35, 4102–4111. [Google Scholar] [CrossRef] [PubMed]
  5. Harguindey, S.; Alfarouk, K.; Polo Orozco, J.; Hardonnière, K.; Stanciu, D.; Fais, S.; Devesa, J. A New and Integral Approach to the Etiopathogenesis and Treatment of Breast Cancer Based upon Its Hydrogen Ion Dynamics. Int. J. Mol. Sci. 2020, 21. [Google Scholar] [CrossRef] [Green Version]
  6. Böhme, I.; Schönherr, R.; Eberle, J.; Bosserhoff, A.K. Membrane Transporters and Channels in Melanoma. Rev. Physiol. Biochem. Pharm. 2020. [Google Scholar] [CrossRef]
  7. Avagliano, A.; Fiume, G.; Pelagalli, A.; Sanità, G.; Ruocco, M.R.; Montagnani, S.; Arcucci, A. Metabolic Plasticity of Melanoma Cells and Their Crosstalk With Tumor Microenvironment. Front. Oncol. 2020, 10. [Google Scholar] [CrossRef]
  8. Koch, A.; Schwab, A. Cutaneous pH landscape as a facilitator of melanoma initiation and progression. Acta Physiol. 2019, 225, e13105. [Google Scholar] [CrossRef]
  9. Stopa, K.B.; Kusiak, A.A.; Szopa, M.D.; Ferdek, P.E.; Jakubowska, M.A. Pancreatic Cancer and Its Microenvironment—Recent Advances and Current Controversies. Int. J. Mol. Sci. 2020, 21. [Google Scholar] [CrossRef]
  10. Harguindey, S.; Stanciu, D.; Devesa, J.; Alfarouk, K.; Cardone, R.A.; Polo Orozco, J.D.; Devesa, P.; Rauch, C.; Orive, G.; Anitua, E.; et al. Cellular acidification as a new approach to cancer treatment and to the understanding and therapeutics of neurodegenerative diseases. Semin. Cancer Biol. 2017, 43, 157–179. [Google Scholar] [CrossRef]
  11. Holzer, P. Acid-sensitive ion channels and receptors. Handb. Exp. Pharm. 2009, 283–332. [Google Scholar] [CrossRef] [Green Version]
  12. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion channels and the hallmarks of cancer. Trends Mol. Med. 2010, 16, 107–121. [Google Scholar] [CrossRef] [PubMed]
  13. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion Channels in Cancer: Are Cancer Hallmarks Oncochannelopathies? Physiol. Rev. 2018, 98, 559–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Litan, A.; Langhans, S.A. Cancer as a channelopathy: Ion channels and pumps in tumor development and progression. Front. Cell. Neurosci. 2015, 9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Andersen, A.P.; Moreira, J.M.A.; Pedersen, S.F. Interactions of ion transporters and channels with cancer cell metabolism and the tumour microenvironment. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Cone, C.D., Jr. Unified theory on the basic mechanism of normal mitotic control and oncogenesis. J. Theor. Biol. 1971, 30, 151–181. [Google Scholar] [CrossRef]
  17. Payne, S.L.; Levin, M.; Oudin, M.J. Bioelectric Control of Metastasis in Solid Tumors. Bioelectricity 2019, 1, 114–130. [Google Scholar] [CrossRef]
  18. Chernet, B.T.; Adams, D.S.; Lobikin, M.; Levin, M. Use of genetically encoded, light-gated ion translocators to control tumorigenesis. Oncotarget 2016, 7, 19575–19588. [Google Scholar] [CrossRef] [Green Version]
  19. Lyall, V.; Biber, T.U. Potential-induced changes in intracellular pH. Am. J. Physiol. 1994, 266, F685–F696. [Google Scholar] [CrossRef]
  20. Cunderlikova, B.; Moan, J.; Sjaastad, I. pH dependent uptake of porphyrin-type photosensitizers by solid tumor cells in vitro is not induced by modification of transmembrane potential. Cancer Lett. 2005, 222, 39–47. [Google Scholar] [CrossRef]
  21. Honasoge, A.; Shelton, K.A.; Sontheimer, H. Autocrine regulation of glioma cell proliferation via pHe-sensitive K(+) channels. Am. J. Physiol. Cell Physiol. 2014, 306, C493–C505. [Google Scholar] [CrossRef] [PubMed]
  22. Bear, C.E.; Davison, J.S.; Shaffer, E.A. Intracellular pH influences the resting membrane potential of isolated rat hepatocytes. Biochim. Biophys. Acta 1988, 944, 113–120. [Google Scholar] [CrossRef]
  23. Deitmer, J.W.; Szatkowski, M. Membrane potential dependence of intracellular pH regulation by identified glial cells in the leech central nervous system. J. Physiol. 1990, 421, 617–631. [Google Scholar] [CrossRef] [PubMed]
  24. Berzingi, S.; Newman, M.; Yu, H.G. Altering bioelectricity on inhibition of human breast cancer cells. Cancer Cell Int. 2016, 16, 72. [Google Scholar] [CrossRef] [Green Version]
  25. Raimondo, J.V.; Tomes, H.; Irkle, A.; Kay, L.; Kellaway, L.; Markram, H.; Millar, R.P.; Akerman, C.J. Tight Coupling of Astrocyte pH Dynamics to Epileptiform Activity Revealed by Genetically Encoded pH Sensors. J. Neurosci. 2016, 36, 7002–7013. [Google Scholar] [CrossRef]
  26. Javadov, S.; Huang, C.; Kirshenbaum, L.; Karmazyn, M. NHE-1 inhibition improves impaired mitochondrial permeability transition and respiratory function during postinfarction remodelling in the rat. J. Mol. Cell Cardiol. 2005, 38, 135–143. [Google Scholar] [CrossRef]
  27. Bonnet, S.; Archer, S.L.; Allalunis-Turner, J.; Haromy, A.; Beaulieu, C.; Thompson, R.; Lee, C.T.; Lopaschuk, G.D.; Puttagunta, L.; Harry, G.; et al. A mitochondria-K+ channel axis is suppressed in cancer and its normalization promotes apoptosis and inhibits cancer growth. Cancer Cell 2007, 11, 37–51. [Google Scholar] [CrossRef] [Green Version]
  28. Azarias, G.; Perreten, H.; Lengacher, S.; Poburko, D.; Demaurex, N.; Magistretti, P.J.; Chatton, J.Y. Glutamate transport decreases mitochondrial pH and modulates oxidative metabolism in astrocytes. J. Neurosci. 2011, 31, 3550–3559. [Google Scholar] [CrossRef]
  29. Boron, W.F. Regulation of intracellular pH. Adv. Physiol. Educ. 2004, 28, 160–179. [Google Scholar] [CrossRef]
  30. Abdul Kadir, L.; Stacey, M.; Barrett-Jolley, R. Emerging Roles of the Membrane Potential: Action Beyond the Action Potential. Front. Physiol. 2018, 9, 1661. [Google Scholar] [CrossRef]
  31. McLaughlin, K.A.; Levin, M. Bioelectric signaling in regeneration: Mechanisms of ionic controls of growth and form. Dev. Biol. 2018, 433, 177–189. [Google Scholar] [CrossRef] [PubMed]
  32. Urrego, D.; Tomczak, A.P.; Zahed, F.; Stuhmer, W.; Pardo, L.A. Potassium channels in cell cycle and cell proliferation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130094. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Yang, M.; Brackenbury, W.J. Membrane potential and cancer progression. Front. Physiol. 2013, 4, 185. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Flinck, M.; Kramer, S.H.; Pedersen, S.F. Roles of pH in control of cell proliferation. Acta Physiol. 2018, 223, e13068. [Google Scholar] [CrossRef] [PubMed]
  35. Gerson, D.F.; Kiefer, H. High intracellular pH accompanies mitotic activity in murine lymphocytes. J. Cell Physiol. 1982, 112, 1–4. [Google Scholar] [CrossRef]
  36. Gagliardi, L.J.; Shain, D.H. Is intracellular pH a clock for mitosis? Theor. Biol. Med. Model 2013, 10, 8. [Google Scholar] [CrossRef] [Green Version]
  37. Martin, C.; Pedersen, S.F.; Schwab, A.; Stock, C. Intracellular pH gradients in migrating cells. Am. J. Physiol. Cell Physiol. 2011, 300, C490–C495. [Google Scholar] [CrossRef] [Green Version]
  38. Paradise, R.K.; Whitfield, M.J.; Lauffenburger, D.A.; van Vliet, K.J. Directional cell migration in an extracellular pH gradient: A model study with an engineered cell line and primary microvascular endothelial cells. Exp. Cell Res. 2013, 319, 487–497. [Google Scholar] [CrossRef]
  39. Shiraishi, T.; Verdone, J.E.; Huang, J.; Kahlert, U.D.; Hernandez, J.R.; Torga, G.; Zarif, J.C.; Epstein, T.; Gatenby, R.; McCartney, A.; et al. Glycolysis is the primary bioenergetic pathway for cell motility and cytoskeletal remodeling in human prostate and breast cancer cells. Oncotarget 2015, 6, 130–143. [Google Scholar] [CrossRef] [Green Version]
  40. Vyas, S.; Zaganjor, E.; Haigis, M.C. Mitochondria and Cancer. Cell 2016, 166, 555–566. [Google Scholar] [CrossRef]
  41. Porporato, P.E.; Filigheddu, N.; Pedro, J.M.B.; Kroemer, G.; Galluzzi, L. Mitochondrial metabolism and cancer. Cell Res. 2018, 28, 265–280. [Google Scholar] [CrossRef] [PubMed]
  42. Steidl, J.V.; Yool, A.J. Differential sensitivity of voltage-gated potassium channels Kv1.5 and Kv1.2 to acidic pH and molecular identification of pH sensor. Mol. Pharmacol. 1999, 55, 812–820. [Google Scholar] [PubMed]
  43. Immke, D.C.; McCleskey, E.W. Protons open acid-sensing ion channels by catalyzing relief of Ca2+ blockade. Neuron 2003, 37, 75–84. [Google Scholar] [CrossRef] [Green Version]
  44. Sherwood, T.W.; Frey, E.N.; Askwith, C.C. Structure and activity of the acid-sensing ion channels. Am. J. Physiol. Cell Physiol. 2012, 303, C699–C710. [Google Scholar] [CrossRef] [Green Version]
  45. Kazmierczak, M.; Zhang, X.; Chen, B.; Mulkey, D.K.; Shi, Y.; Wagner, P.G.; Pivaroff-Ward, K.; Sassic, J.K.; Bayliss, D.A.; Jegla, T. External pH modulates EAG superfamily K+ channels through EAG-specific acidic residues in the voltage sensor. J. Gen. Physiol. 2013, 141, 721–735. [Google Scholar] [CrossRef] [Green Version]
  46. Chaves, G.; Bungert-Plumke, S.; Franzen, A.; Mahorivska, I.; Musset, B. Zinc modulation of proton currents in a new voltage-gated proton channel suggests a mechanism of inhibition. FEBS J. 2020. [Google Scholar] [CrossRef]
  47. Frankenhaeuser, B.; Hodgkin, A.L. The action of calcium on the electrical properties of squid axons. J. Physiol. 1957, 137, 218–244. [Google Scholar] [CrossRef]
  48. Hille, B. Ion Channels of Excitable Membranes, 3rd ed.; Oxford University Press: Sunderland, MA, USA, 2001; ISBN 0878933212 9780878933211. [Google Scholar]
  49. Deutsch, C.; Lee, S.C. Modulation of K+ currents in human lymphocytes by pH. J. Physiol. 1989, 413, 399–413. [Google Scholar] [CrossRef]
  50. Kehl, S.J.; Eduljee, C.; Kwan, D.C.; Zhang, S.; Fedida, D. Molecular determinants of the inhibition of human Kv1.5 potassium currents by external protons and Zn2+. J. Physiol. 2002, 541, 9–24. [Google Scholar] [CrossRef]
  51. Hille, B. Charges and potentials at the nerve surface: Divalent ions and pH. J. Gen. Physiol. 1968, 51, 221–236. [Google Scholar] [CrossRef] [Green Version]
  52. Becchetti, A.; Pillozzi, S.; Morini, R.; Nesti, E.; Arcangeli, A. New insights into the regulation of ion channels by integrins. In International Review of Cell and Molecular Biology; Elsevier: Amsterdam, The Netherlands, 2010; pp. 135–190. ISBN 1937-6448. [Google Scholar]
  53. Liddington, R.C.; Ginsberg, M.H. Integrin activation takes shape. J. Cell Biol. 2002, 158, 833–839. [Google Scholar] [CrossRef] [PubMed]
  54. Tadokoro, S.; Shattil, S.J.; Eto, K.; Tai, V.; Liddington, R.C.; de Pereda, J.M.; Ginsberg, M.H.; Calderwood, D.A. Talin binding to integrin ß tails: A final common step in integrin activation. Science 2003, 302, 103–106. [Google Scholar] [CrossRef] [PubMed]
  55. Critchley, D.R.; Gingras, A.R. Talin at a glance. J. Cell Sci. 2008, 121, 1345–1347. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Ginsberg, M.H.; Partridge, A.; Shattil, S.J. Integrin regulation. Curr. Opin. Cell Biol. 2005, 17, 509–516. [Google Scholar] [CrossRef] [PubMed]
  57. Krähling, H.; Mally, S.; Eble, J.A.; Noël, J.; Schwab, A.; Stock, C. The glycocalyx maintains a cell surface pH nanoenvironment crucial for integrin-mediated migration of human melanoma cells. Pflügers Arch. Eur. J. Physiol. 2009, 458, 1069–1083. [Google Scholar]
  58. Lehenkari, P.P.; Horton, M.A. Single integrin molecule adhesion forces in intact cells measured by atomic force microscopy. Biochem. Biophys. Res. Commun. 1999, 259, 645–650. [Google Scholar] [CrossRef] [PubMed]
  59. Paradise, R.K.; Lauffenburger, D.A.; van Vliet, K.J. Acidic extracellular pH promotes activation of integrin αvβ3. PLoS ONE 2011, 6, e15746. [Google Scholar] [CrossRef] [Green Version]
  60. Srivastava, J.; Barreiro, G.; Groscurth, S.; Gingras, A.R.; Goult, B.T.; Critchley; Kelly, M.J.S.; Jacobson, M.P.; Barber, D.L. Structural model and functional significance of pH-dependent talin–actin binding for focal adhesion remodeling. Proc. Natl. Acad. Sci. USA 2008, 105, 14436–14441. [Google Scholar] [CrossRef] [Green Version]
  61. Choi, C.H.; Webb, B.A.; Chimenti, M.S.; Jacobson, M.P.; Barber, D.L. pH sensing by FAK-His58 regulates focal adhesion remodeling. J. Cell Biol. 2013, 202, 849–859. [Google Scholar] [CrossRef] [Green Version]
  62. Stock, C.; Schwab, A. Ion channels and transporters in metastasis. Biochim. Biophys. Acta (BBA)-Biomembr. 2015, 1848, 2638–2646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Stock, C.; Gassner, B.; Hauck, C.R.; Arnold, H.; Mally, S.; Eble, J.A.; Dieterich, P.; Schwab, A. Migration of human melanoma cells depends on extracellular pH and Na+/H+ exchange. J. Physiol. 2005, 567, 225–238. [Google Scholar] [CrossRef] [PubMed]
  64. Beaumont, K.G.; Mrksich, M. The mechanostability of isolated focal adhesions is strongly dependent on pH. Chem. Biol. 2012, 19, 711–720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Arcangeli, A.; Becchetti, A. Complex functional interaction between integrin receptors and ion channels. Trends Cell Biol. 2006, 16, 631–639. [Google Scholar] [CrossRef] [PubMed]
  66. Becchetti, A.; Arcangeli, A. Integrins and Ion Channels. Molecular Complexes and Signaling; Springer Science & Business Media: Berlin/Heidelberg, Germany, 2010; ISBN 144196066X. [Google Scholar]
  67. Tanner, M.R.; Pennington, M.W.; Laragione, T.; Gulko, P.S.; Beeton, C. KCa1.1 channels regulate beta1-integrin function and cell adhesion in rheumatoid arthritis fibroblast-like synoviocytes. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2017, 31, 3309–3320. [Google Scholar] [CrossRef] [Green Version]
  68. Cahalan, M.D.; Chandy, K.G. The functional network of ion channels in T lymphocytes. Immunol. Rev. 2009, 231, 59–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Katsumi, A.; Orr, A.W.; Tzima, E.; Schwartz, M.A. Integrins in mechanotransduction. J. Biol. Chem. 2004, 279, 12001–12004. [Google Scholar] [CrossRef] [Green Version]
  70. Arcangeli, A.; Becchetti, A.; Mannini, A.; Mugnai, G.; de Filippi, P.; Tarone, G.; Del Bene, M.R.; Barletta, E.; Wanke, E.; Olivotto, M. Integrin-mediated neurite outgrowth in neuroblastoma cells depends on the activation of potassium channels. J. Cell Biol. 1993, 122, 1131–1143. [Google Scholar] [CrossRef]
  71. Becchetti, A.; Arcangeli, A.; Del Bene, M.R.; Olivotto, M.; Wanke, E. Response to fibronectin–integrin interaction in leukaemia cells: Delayed enhancing of a K+ current. Proc. R. Soc. Lond. Ser. B Biol. Sci. 1992, 248, 235–240. [Google Scholar]
  72. Hofmann, G.; Bernabei, P.A.; Crociani, O.; Cherubini, A.; Guasti, L.; Pillozzi, S.; Lastraioli, E.; Polvani, S.; Bartolozzi, B.; Solazzo, V. HERG K+ channels activation during b1 integrin-mediated adhesion to fibronectin induces an up regulation of avß3 integrin in the preosteoclastic leukemia cell line FLG 29.1. J. Biol. Chem. 2000, 276, 4923–4931. [Google Scholar] [CrossRef] [Green Version]
  73. Rezzonico, R.; Cayatte, C.; Bourget-Ponzio, I.; Romey, G.; Belhacene, N.; Loubat, A.; Rocchi, S.; van Obberghen, E.; Girault, J.-A.; Rossi, B. Focal adhesion kinase pp125FAK interacts with the large conductance calcium-activated hSlo potassium channel in human osteoblasts: Potential role in mechanotransduction. J. Bone Miner. Res. 2003, 18, 1863–1871. [Google Scholar] [CrossRef]
  74. Wei, J.-F.; Wei, L.; Zhou, X.; Lu, Z.-Y.; Francis, K.; Hu, X.-Y.; Liu, Y.; Xiong, W.-C.; Zhang, X.; Banik, N.L. Formation of Kv2.1-FAK complex as a mechanism of FAK activation, cell polarization and enhanced motility. J. Cell. Physiol. 2008, 217, 544–557. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Tian, L.; Chen, L.; McClafferty, H.; Sailer, C.A.; Ruth, P.; Knaus, H.-G.; Shipston, M.J. A noncanonical SH3 domain binding motif links BK channels to the actin cytoskeleton via the SH3 adapter cortactin. Faseb J. 2006, 20, 2588–2590. [Google Scholar] [CrossRef] [PubMed]
  76. Schwab, A.; Nechyporuk-Zloy, V.; Fabian, A.; Stock, C. Cells move when ions and water flow. Pflügers Arch. Eur. J. Physiol. 2007, 453, 421–432. [Google Scholar] [CrossRef]
  77. Nechyporuk-Zloy, V.; Dieterich, P.; Oberleithner, H.; Stock, C.; Schwab, A. Dynamics of single potassium channel proteins in the plasma membrane of migrating cells. Am. J. Physiol. Cell Physiol. 2008, 294, C1096–C1102. [Google Scholar] [CrossRef]
  78. Du, C.; Zheng, Z.; Li, D.; Chen, L.; Li, N.; Yi, X.; Yang, Y.; Guo, F.; Liu, W.; Xie, X. BKCa promotes growth and metastasis of prostate cancer through facilitating the coupling between αvβ3 integrin and FAK. Oncotarget 2016, 7, 40174. [Google Scholar] [CrossRef] [PubMed]
  79. Cherubini, A.; Pillozzi, S.; Hofmann, G.; Crociani, O.; Guasti, L.; Lastraioli, E.; Polvani, S.; Masi, A.; Becchetti, A.; Wanke, E. HERG K+ channels and β1 integrins interact through the assembly of a macromolecular complex. Ann. N. Y. Acad. Sci. 2002, 973, 559–561. [Google Scholar] [CrossRef]
  80. Cherubini, A.; Hofmann, G.; Pillozzi, S.; Guasti, L.; Crociani, O.; Cilia, E.; Di Stefano, P.; Degani, S.; Balzi, M.; Olivotto, M. Human ether-a-go-go-related gene 1 channels are physically linked to β1 integrins and modulate adhesion-dependent signaling. Mol. Biol. Cell 2005, 16, 2972–2983. [Google Scholar] [CrossRef] [Green Version]
  81. Pillozzi, S.; Masselli, M.; de Lorenzo, E.; Accordi, B.; Cilia, E.; Crociani, O.; Amedei, A.; Veltroni, M.; D’Amico, M.; Basso, G. Chemotherapy resistance in acute lymphoblastic leukemia requires hERG1 channels and is overcome by hERG1 blockers. Blood J. Am. Soc. Hematol. 2011, 117, 902–914. [Google Scholar] [CrossRef]
  82. Artym, V.V.; Petty, H.R. Molecular proximity of Kv1.3 voltage-gated potassium channels and β1-integrins on the plasma membrane of melanoma cells: Effects of cell adherence and channel blockers. J. Gen. Physiol. 2002, 120, 29–37. [Google Scholar] [CrossRef] [Green Version]
  83. Levite, M.; Cahalon, L.; Peretz, A.; Hershkoviz, R.; Sobko, A.; Ariel, A.; Desai, R.; Attali, B.; Lider, O. Extracellular K+ and opening of voltage-gated potassium channels activate T cell integrin function: Physical and functional association between Kv1.3 channels and β1 integrins. J. Exp. Med. 2000, 191, 1167–1176. [Google Scholar] [CrossRef]
  84. Lastraioli, E.; Guasti, L.; Crociani, O.; Polvani, S.; Hofmann, G.; Witchel, H.; Bencini, L.; Calistri, M.; Messerini, L.; Scatizzi, M. HERG1 gene and HERG1 protein are overexpressed in colorectal cancers and regulate cell invasion of tumor cells. Cancer Res. 2004, 64, 606–611. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Crociani, O.; Zanieri, F.; Pillozzi, S.; Lastraioli, E.; Stefanini, M.; Fiore, A.; Fortunato, A.; D’Amico, M.; Masselli, M.; de Lorenzo, E. hERG1 channels modulate integrin signaling to trigger angiogenesis and tumor progression in colorectal cancer. Sci. Rep. 2013, 3, 3308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Masi, A.; Becchetti, A.; Restano-Cassulini, R.; Polvani, S.; Hofmann, G.; Buccoliero, A.M.; Paglierani, M.; Pollo, B.; Taddei, G.L.; Gallina, P. hERG1 channels are overexpressed in glioblastoma multiforme and modulate VEGF secretion in glioblastoma cell lines. Br. J. Cancer 2005, 93, 781–792. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Doherty, P.; Ashton, S.V.; Moore, S.E.; Walsh, F.S. Morphoregulatory activities of NCAM and N-cadherin can be accounted for by G protein-dependent activation of L-and N-type neuronal Ca2+ channels. Cell 1991, 67, 21–33. [Google Scholar] [CrossRef]
  88. 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]
  89. Somodi, S.; Varga, Z.; Hajdu, P.; Starkus, J.G.; Levy, D.I.; Gaspar, R.; Panyi, G. pH-dependent modulation of Kv1.3 inactivation: Role of His399. Am. J. Physiol. Cell Physiol. 2004, 287, C1067–C1076. [Google Scholar] [CrossRef] [Green Version]
  90. Somodi, S.; Hajdu, P.; Gaspar, R.; Panyi, G.; Varga, Z. Effects of changes in extracellular pH and potassium concentration on Kv1.3 inactivation. Eur. Biophys. J. 2008, 37, 1145–1156. [Google Scholar] [CrossRef]
  91. Jang, S.H.; Kang, K.-S.; Ryu, P.D.; Lee, S.Y. Kv1.3 voltage-gated K+ channel subunit as a potential diagnostic marker and therapeutic target for breast cancer. BMB Rep. 2009, 42, 535–539. [Google Scholar] [CrossRef] [Green Version]
  92. Eil, R.; Vodnala, S.K.; Clever, D.; Klebanoff, C.A.; Sukumar, M.; Pan, J.H.; Palmer, D.C.; Gros, A.; Yamamoto, T.N.; Patel, S.J.; et al. Ionic immune suppression within the tumour microenvironment limits T cell effector function. Nature 2016, 537, 539–543. [Google Scholar] [CrossRef] [Green Version]
  93. Chimote, A.A.; Hajdu, P.; Sfyris, A.M.; Gleich, B.N.; Wise-Draper, T.; Casper, K.A.; Conforti, L. Kv1.3 Channels Mark Functionally Competent CD8+ Tumor-Infiltrating Lymphocytes in Head and Neck Cancer. Cancer Res. 2017, 77, 53–61. [Google Scholar] [CrossRef] [Green Version]
  94. Trapani, J.G.; Korn, S.J. Effect of external pH on activation of the Kv1.5 potassium channel. Biophys. J. 2003, 84, 195–204. [Google Scholar] [CrossRef] [Green Version]
  95. Xie, J.; Wang, B.S.; Yu, D.H.; Lu, Q.; Ma, J.; Qi, H.; Fang, C.; Chen, H.Z. Dichloroacetate shifts the metabolism from glycolysis to glucose oxidation and exhibits synergistic growth inhibition with cisplatin in HeLa cells. Int. J. Oncol. 2011, 38, 409–417. [Google Scholar] [CrossRef] [PubMed]
  96. Lan, M.; Shi, Y.; Han, Z.; Hao, Z.; Pan, Y.; Liu, N.; Guo, C.; Hong, L.; Wang, J.; Qiao, T.; et al. Expression of delayed rectifier potassium channels and their possible roles in proliferation of human gastric cancer cells. Cancer Biol. Ther. 2005, 4, 1342–1347. [Google Scholar] [CrossRef] [Green Version]
  97. Suzuki, T.; Takimoto, K. Selective expression of HERG and Kv2 channels influences proliferation of uterine cancer cells. Int. J. Oncol. 2004, 25, 153–159. [Google Scholar] [CrossRef] [PubMed]
  98. Shimizu, T.; Fujii, T.; Takahashi, Y.; Takahashi, Y.; Suzuki, T.; Ukai, M.; Tauchi, K.; Horikawa, N.; Tsukada, K.; Sakai, H. Up-regulation of Kv7.1 channels in thromboxane A2-induced colonic cancer cell proliferation. Pflug. Arch. Eur. J. Physiol. 2014, 466, 541–548. [Google Scholar] [CrossRef]
  99. Yamane, T.; Furukawa, T.; Horikawa, S.; Hiraoka, M. External pH regulates the slowly activating potassium current IsK expressed in Xenopus oocytes. FEBS Lett. 1993, 319, 229–232. [Google Scholar] [CrossRef] [Green Version]
  100. Unsöld, B.; Kerst, G.; Brousos, H.; Hübner, M.; Schreiber, R.; Nitschke, R.; Greger, R.; Bleich, M. KCNE1 reverses the response of the human K+ channel KCNQ1 to cytosolic pH changes and alters its pharmacology and sensitivity to temperature. Pflügers Arch. Eur. J. Physiol. 2000, 441, 368–378. [Google Scholar] [CrossRef]
  101. Martínez, R.; Stühmer, W.; Martin, S.; Schell, J.; Reichmann, A.; Rohde, V.; Pardo, L. Analysis of the expression of Kv10.1 potassium channel in patients with brain metastases and glioblastoma multiforme: Impact on survival. BMC Cancer 2015, 15. [Google Scholar] [CrossRef] [Green Version]
  102. Terlau, H.; Ludwig, J.; Steffan, R.; Pongs, O.; Stühmer, W.; Heinemann, S.H. Extracellular Mg2+ regulates activation of rat eag potassium channel. Pflügers Arch. Eur. J. Physiol. 1996, 432, 301–312. [Google Scholar] [CrossRef]
  103. Spitzner, M.; Ousingsawat, J.; Scheidt, K.; Kunzelmann, K.; Schreiber, R. Voltage-gated K+ channels support proliferation of colonic carcinoma cells. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2007, 21, 35–44. [Google Scholar] [CrossRef]
  104. Ding, X.-W.; Luo, H.-s.; Jin, X.; Yan, J.-j.; Ai, Y.-w. Aberrant expression of Eag1 potassium channels in gastric cancer patients and cell lines. Med. Oncol. 2007, 24, 345–350. [Google Scholar] [CrossRef] [PubMed]
  105. Lai, Q.; Wang, T.; Guo, Q.; Zhang, Y.; Wang, Y.; Yuan, L.; Ling, R.; He, Y.; Wang, W. Positive correlation between the expression of hEag1 and HIF-1α in breast cancers: An observational study. BMJ Open 2014, 4, e005049. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Wu, J.; Zhong, D.; Fu, X.; Liu, Q.; Kang, L.; Ding, Z. Silencing of Ether à go-go 1 by shRNA inhibits osteosarcoma growth and cell cycle progression. Int. J. Mol. Sci. 2014, 15, 5570–5581. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Wu, J.; Wu, X.; Zhong, D.; Zhai, W.; Ding, Z.; Zhou, Y. Short Hairpin RNA (shRNA) Ether à go-go 1 (Eag1) Inhibition of Human Osteosarcoma Angiogenesis via VEGF/PI3K/AKT Signaling. Int. J. Mol. Sci. 2012, 13, 12573–12583. [Google Scholar] [CrossRef]
  108. van Slyke, A.C.; Cheng, Y.M.; Mafi, P.; Allard, C.R.; Hull, C.M.; Shi, Y.P.; Claydon, T.W. Proton block of the pore underlies the inhibition of hERG cardiac K+ channels during acidosis. Am. J. Physiol. Cell Physiol. 2012, 302, C1797–C1806. [Google Scholar] [CrossRef] [Green Version]
  109. Wang, Y.; Guo, J.; Perissinotti, L.L.; Lees-Miller, J.; Teng, G.; Durdagi, S.; Duff, H.J.; Noskov, S.Y. Role of the pH in state-dependent blockade of hERG currents. Sci. Rep. 2016, 6, 32536. [Google Scholar] [CrossRef] [Green Version]
  110. Lastraioli, E.; Iorio, J.; Arcangeli, A. Ion channel expression as promising cancer biomarker. Biochim. Biophys. Acta 2015, 1848, 2685–2702. [Google Scholar] [CrossRef] [Green Version]
  111. Klumpp, L.; Sezgin, E.C.; Skardelly, M.; Eckert, F.; Huber, S.M. KCa3.1 Channels and Glioblastoma: In Vitro Studies. Curr. Neuropharmacol. 2018, 16, 627–635. [Google Scholar] [CrossRef]
  112. Doroszewicz, J.; Waldegger, P.; Jeck, N.; Seyberth, H.; Waldegger, S. pH dependence of extracellular calcium sensing receptor activity determined by a novel technique. Kidney Int. 2005, 67, 187–192. [Google Scholar] [CrossRef] [Green Version]
  113. Strupp, M.; Staub, F.; Grafe, P. A Ca2+- and pH-dependent K+ channel of rat C6 glioma cells and its possible role in acidosis-induced cell swelling. Glia 1993, 9, 136–145. [Google Scholar] [CrossRef]
  114. Faouzi, M.; Hague, F.; Geerts, D.; Ay, A.-S.; Potier-Cartereau, M.; Ahidouch, A.; Ouadid-Ahidouch, H. Functional cooperation between KCa3.1 and TRPC1 channels in human breast cancer: Role in cell proliferation and patient prognosis. Oncotarget 2016, 7, 36419–36435. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Ouadid-Ahidouch, H.; Roudbaraki, M.; Delcourt, P.; Ahidouch, A.; Joury, N.; Prevarskaya, N. Functional and molecular identification of intermediate-conductance Ca2+-activated K+ channels in breast cancer cells: Association with cell cycle progression. Am. J. Physiol. Cell Physiol. 2004, 287, C125–C134. [Google Scholar] [CrossRef] [PubMed]
  116. Bonito, B.; Sauter, D.R.P.; Schwab, A.; Djamgoz, M.B.A.; Novak, I. KCa3.1 (IK) modulates pancreatic cancer cell migration, invasion and proliferation: Anomalous effects on TRAM-34. Pflug. Arch. Eur. J. Physiol. 2016, 468, 1865–1875. [Google Scholar] [CrossRef] [PubMed]
  117. Storck, H.; Hild, B.; Schimmelpfennig, S.; Sargin, S.; Nielsen, N.; Zaccagnino, A.; Budde, T.; Novak, I.; Kalthoff, H.; Schwab, A. Ion channels in control of pancreatic stellate cell migration. Oncotarget 2017, 8, 769–784. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Vodnala, S.K.; Eil, R.; Kishton, R.J.; Sukumar, M.; Yamamoto, T.N.; Ha, N.H.; Lee, P.H.; Shin, M.; Patel, S.J.; Yu, Z.; et al. T cell stemness and dysfunction in tumors are triggered by a common mechanism. Science 2019, 363. [Google Scholar] [CrossRef]
  119. Chimote, A.A.; Balajthy, A.; Arnold, M.J.; Newton, H.S.; Hajdu, P.; Qualtieri, J.; Wise-Draper, T.; Conforti, L. A defect in KCa3.1 channel activity limits the ability of CD8+ T cells from cancer patients to infiltrate an adenosine-rich microenvironment. Sci. Signal 2018, 11. [Google Scholar] [CrossRef] [Green Version]
  120. Chimote, A.A.; Gawali, V.S.; Newton, H.S.; Wise-Draper, T.M.; Conforti, L. A Compartmentalized Reduction in Membrane-Proximal Calmodulin Reduces the Immune Surveillance Capabilities of CD8+ T Cells in Head and Neck Cancer. Front. Pharm. 2020, 11, 143. [Google Scholar] [CrossRef] [Green Version]
  121. Guo, Z.; Liu, J.; Zhang, L.; Su, B.; Xing, Y.; He, Q.; Ci, W.; Li, X.; Zhou, L. KCNJ1 inhibits tumor proliferation and metastasis and is a prognostic factor in clear cell renal cell carcinoma. Tumour Biol. 2015, 36, 1251–1259. [Google Scholar] [CrossRef]
  122. Schulte, U.; Hahn, H.; Konrad, M.; Jeck, N.; Derst, C.; Wild, K.; Weidemann, S.; Ruppersberg, J.P.; Fakler, B.; Ludwig, J. pH gating of ROMK (Kir1.1) channels: Control by an Arg-Lys-Arg triad disrupted in antenatal Bartter syndrome. Proc. Natl. Acad. Sci. USA 1999, 96, 15298–15303. [Google Scholar] [CrossRef] [Green Version]
  123. Higashimori, H.; Sontheimer, H. Role of Kir4.1 channels in growth control of glia. Glia 2007, 55, 1668–1679. [Google Scholar] [CrossRef] [Green Version]
  124. Pessia, M.; Imbrici, P.; D’Adamo, M.C.; Salvatore, L.; Tucker, S.J. Differential pH sensitivity of Kir4.1 and Kir4.2 potassium channels and their modulation by heteropolymerisation with Kir5.1. J. Physiol. 2001, 532, 359–367. [Google Scholar] [CrossRef] [PubMed]
  125. Zhang, G.M.; Wan, F.N.; Qin, X.J.; Cao, D.L.; Zhang, H.L.; Zhu, Y.; Dai, B.; Shi, G.H.; Ye, D.W. Prognostic significance of the TREK-1 K2P potassium channels in prostate cancer. Oncotarget 2015, 6, 18460–18468. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Sandoz, G.; Douguet, D.; Chatelain, F.; Lazdunski, M.; Lesage, F. Extracellular acidification exerts opposite actions on TREK1 and TREK2 potassium channels via a single conserved histidine residue. Proc. Natl. Acad. Sci. USA 2009, 106, 14628–14633. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Maingret, F.; Patel, A.J.; Lesage, F.; Lazdunski, M.; Honoré, E. Mechano- or acid stimulation, two interactive modes of activation of the TREK-1 potassium channel. J. Biol. Chem. 1999, 274, 26691–26696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Sauter, D.R.P.; 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 Mol. Basis Dis. 2016, 1862, 1994–2003. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  129. Meuth, S.G.; Herrmann, A.M.; Ip, C.W.; Kanyshkova, T.; Bittner, S.; Weishaupt, A.; Budde, T.; Wiendl, H. The two-pore domain potassium channel TASK3 functionally impacts glioma cell death. J. Neuro-Oncol. 2008, 87, 263–270. [Google Scholar] [CrossRef]
  130. Pei, L.; Wiser, O.; Slavin, A.; Mu, D.; Powers, S.; Lily, Y.J.; Hoey, T. Oncogenic potential of TASK3 (Kcnk9) depends on K+ channel function. Proc. Natl. Acad. Sci. USA 2003, 100, 7803–7807. [Google Scholar] [CrossRef] [Green Version]
  131. Alvarez-Baron, C.P.; Jonsson, P.; Thomas, C.; Dryer, S.E.; Williams, C. The two-pore domain potassium channel KCNK5: Induction by estrogen receptor α and role in proliferation of breast cancer cells. Mol. Endocrinol. 2011, 25, 1326–1336. [Google Scholar] [CrossRef] [Green Version]
  132. Niemeyer, M.I.; Cid, L.P.; Gonzalez, W.; Sepulveda, F.V. Gating, Regulation, and Structure in K2P K+ Channels: In Varietate Concordia? Mol. Pharmacol. 2016, 90, 309–317. [Google Scholar] [CrossRef] [Green Version]
  133. Bustos, D.; Bedoya, M.; Ramírez, D.; Concha, G.; Zúñiga, L.; Decher, N.; Hernández-rodríguez, E.W.; Sepúlveda, F.V.; Martínez, L.; González, W. Elucidating the structural basis of the intracellular ph sensing mechanism of TASK-2 K2P channels. Int. J. Mol. Sci. 2020, 21. [Google Scholar] [CrossRef] [Green Version]
  134. Ratnayake, R.; Covell, D.; Ransom, T.T.; Gustafson, K.R.; Beutler, J.A. Englerin A, a Selective Inhibitor of Renal Cancer Cell Growth, from Phyllanthus engleri. Org. Lett. 2009, 11, 57–60. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Semtner, M.; Schaefer, M.; Pinkenburg, O.; Plant, T.D. Potentiation of TRPC5 by protons. J. Biol. Chem. 2007, 282, 33868–33878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Chigurupati, S.; Venkataraman, R.; Barrera, D.; Naganathan, A.; Madan, M.; Paul, L.; Pattisapu, J.V.; Kyriazis, G.A.; Sugaya, K.; Bushnev, S.; et al. Receptor channel TRPC6 is a key mediator of Notch-driven glioblastoma growth and invasiveness. Cancer Res. 2010, 70, 418–427. [Google Scholar] [CrossRef] [Green Version]
  137. El Boustany, C.; Bidaux, G.; Enfissi, A.; Delcourt, P.; Prevarskaya, N.; Capiod, T. Capacitative calcium entry and transient receptor potential canonical 6 expression control human hepatoma cell proliferation. Hepatology 2008, 47, 2068–2077. [Google Scholar] [CrossRef] [PubMed]
  138. Steele, C.W.; Karim, S.A.; Leach, J.D.G.; Bailey, P.; Upstill-Goddard, R.; Rishi, L.; Foth, M.; Bryson, S.; McDaid, K.; Wilson, Z.; et al. CXCR2 Inhibition Profoundly Suppresses Metastases and Augments Immunotherapy in Pancreatic Ductal Adenocarcinoma. Cancer Cell 2016, 29, 832–845. [Google Scholar] [CrossRef] [Green Version]
  139. Lindemann, O.; Umlauf, D.; Frank, S.; Schimmelpfennig, S.; Bertrand, J.; Pap, T.; Hanley, P.J.; Fabian, A.; Dietrich, A.; Schwab, A. TRPC6 Regulates CXCR2-Mediated Chemotaxis of Murine Neutrophils. J. Immunol. 2013, 190, 5496–5505. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  140. Nielsen, N.; Kondratska, K.; Ruck, T.; Hild, B.; Kovalenko, I.; Schimmelpfennig, S.; Welzig, J.; Sargin, S.; Lindemann, O.; Christian, S.; et al. TRPC6 channels modulate the response of pancreatic stellate cells to hypoxia. Pflug. Arch. Eur. J. Physiol. 2017, 469, 1567–1577. [Google Scholar] [CrossRef]
  141. Gershkovitz, M.; Caspi, Y.; Fainsod-Levi, T.; Katz, B.; Michaeli, J.; Khawaled, S.; Lev, S.; Polyansky, L.; Shaul, M.E.; Sionov, R.V.; et al. TRPM2 Mediates Neutrophil Killing of Disseminated Tumor Cells. Cancer Res. 2018, 78, 2680–2690. [Google Scholar] [CrossRef] [Green Version]
  142. Gershkovitz, M.; Fainsod-Levi, T.; Zelter, T.; Sionov, R.V.; Granot, Z. TRPM2 modulates neutrophil attraction to murine tumor cells by regulating CXCL2 expression. Cancer Immunol. Immunother. 2019, 68, 33–43. [Google Scholar] [CrossRef]
  143. Starkus, J.G.; Fleig, A.; Penner, R. The calcium-permeable non-selective cation channel TRPM2 is modulated by cellular acidification. J. Physiol. 2010, 588, 1227–1240. [Google Scholar] [CrossRef]
  144. Zeng, X.; Sikka, S.C.; Huang, L.; Sun, C.; Xu, C.; Jia, D.; Abdel-Mageed, A.B.; Pottle, J.E.; Taylor, J.T.; Li, M. Novel role for the transient receptor potential channel TRPM2 in prostate cancer cell proliferation. Prostate Cancer Prostatic Dis. 2010, 13, 195–201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Chen, J.-P.; Luan, Y.; You, C.-X.; Chen, X.-H.; Luo, R.-C.; Li, R. TRPM7 regulates the migration of human nasopharyngeal carcinoma cell by mediating Ca2+ influx. Cell Calcium 2010, 47, 425–432. [Google Scholar] [CrossRef] [PubMed]
  146. Chokshi, R.; Matsushita, M.; Kozak, J.A. Detailed examination of Mg2+ and pH sensitivity of human TRPM7 channels. Am. J. Physiol. Cell Physiol. 2012, 302, C1004–C1011. [Google Scholar] [CrossRef] [Green Version]
  147. Gao, H.; Chen, X.; Du, X.; Guan, B.; Liu, Y.; Zhang, H. EGF enhances the migration of cancer cells by up-regulation of TRPM7. Cell Calcium 2011, 50, 559–568. [Google Scholar] [CrossRef] [PubMed]
  148. Rybarczyk, P.; Gautier, M.; Hague, F.; Dhennin-Duthille, I.; Chatelain, D.; Kerr-Conte, J.; Pattou, F.; Regimbeau, J.-M.; Sevestre, H.; Ouadid-Ahidouch, H. Transient receptor potential melastatin-related 7 channel is overexpressed in human pancreatic ductal adenocarcinomas and regulates human pancreatic cancer cell migration. Int. J. Cancer 2012, 131, E851–E861. [Google Scholar] [CrossRef]
  149. Guilbert, A.; Gautier, M.; Dhennin-Duthille, I.; Haren, N.; Sevestre, H.; Ouadid-Ahidouch, H. Evidence that TRPM7 is required for breast cancer cell proliferation. Am. J. Physiol. Cell Physiol. 2009, 297, C493–C502. [Google Scholar] [CrossRef]
  150. Weber, L.V.; Al-Refae, K.; Wölk, G.; Bonatz, G.; Altmüller, J.; Becker, C.; Gisselmann, G.; Hatt, H. Expression and functionality of TRPV1 in breast cancer cells. Breast Cancer Targets Ther. 2016, 8, 243–252. [Google Scholar] [CrossRef] [Green Version]
  151. Caterina, M.J.; Schumacher, M.A.; Tominaga, M.; Rosen, T.A.; Levine, J.D.; Julius, D. The capsaicin receptor: A heat-activated ion channel in the pain pathway. Nature 1997, 389, 816–824. [Google Scholar] [CrossRef]
  152. Hou, N.; He, X.; Yang, Y.; Fu, J.; Zhang, W.; Guo, Z.; Hu, Y.; Liang, L.; Xie, W.; Xiong, H.; et al. TRPV1 Induced Apoptosis of Colorectal Cancer Cells by Activating Calcineurin-NFAT2-p53 Signaling Pathway. Biomed Res. Int. 2019, 2019, 1–8. [Google Scholar] [CrossRef]
  153. Stock, K.; Kumar, J.; Synowitz, M.; Petrosino, S.; Imperatore, R.; Smith, E.S.J.; Wend, P.; Purfürst, B.; Nuber, U.A.; Gurok, U.; et al. Neural precursor cells induce cell death of high-grade astrocytomas through stimulation of TRPV1. Nat. Med. 2012, 18, 1232–1238. [Google Scholar] [CrossRef] [Green Version]
  154. Fonseca, B.M.; Correia-da-Silva, G.; Teixeira, N.A. Cannabinoid-induced cell death in endometrial cancer cells: Involvement of TRPV1 receptors in apoptosis. J. Physiol. Biochem. 2018, 74, 261–272. [Google Scholar] [CrossRef] [PubMed]
  155. Thoppil, R.J.; Cappelli, H.C.; Adapala, R.K.; Kanugula, A.K.; Paruchuri, S.; Thodeti, C.K. TRPV4 channels regulate tumor angiogenesis via modulation of Rho/Rho kinase pathway. Oncotarget 2016, 7, 25849–25861. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Suzuki, M.; Mizuno, A.; Kodaira, K.; Imai, M. Impaired pressure sensation in mice lacking TRPV4. J. Biol. Chem. 2003, 278, 22664–22668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Fiorio Pla, A.; Ong, H.L.; Cheng, K.T.; Brossa, A.; Bussolati, B.; Lockwich, T.; Paria, B.; Munaron, L.; Ambudkar, I.S. TRPV4 mediates tumor-derived endothelial cell migration via arachidonic acid-activated actin remodeling. Oncogene 2012, 31, 200–212. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Fang, Y.; Liu, G.; Xie, C.; Qian, K.; Lei, X.; Liu, Q.; Liu, G.; Cao, Z.; Fu, J.; Du, H.; et al. Pharmacological inhibition of TRPV4 channel suppresses malignant biological behavior of hepatocellular carcinoma via modulation of ERK signaling pathway. Biomed. Pharmacother. 2018, 101, 910–919. [Google Scholar] [CrossRef]
  159. Xie, R.; Xu, J.; Xiao, Y.; Wu, J.; Wan, H.; Tang, B.; Liu, J.; Fan, Y.; Wang, S.; Wu, Y.; et al. Calcium Promotes Human Gastric Cancer via a Novel Coupling of Calcium-Sensing Receptor and TRPV4 Channel. Cancer Res. 2017, 77, 6499–6512. [Google Scholar] [CrossRef] [Green Version]
  160. Bare, D.J.; Cherny, V.V.; DeCoursey, T.E.; Abukhdeir, A.M.; Morgan, D. Expression and function of voltage gated proton channels (Hv1) in MDA-MB-231 cells. PLoS ONE 2020, 15, e0227522. [Google Scholar] [CrossRef]
  161. Wang, Y.; Li, S.J.; Wu, X.; Che, Y.; Li, Q. Clinicopathological and biological significance of human voltage-gated proton channel Hv1 protein overexpression in breast cancer. J. Biol. Chem. 2012, 287, 13877–13888. [Google Scholar] [CrossRef] [Green Version]
  162. Wang, Y.; Wu, X.; Li, Q.; Zhang, S.; Li, S.J. Human voltage-gated proton channel hv1: A new potential biomarker for diagnosis and prognosis of colorectal cancer. PLoS ONE 2013, 8, e70550. [Google Scholar] [CrossRef] [Green Version]
  163. Wang, Y.; Zhang, S.; Li, S.J. Zn2+ induces apoptosis in human highly metastatic SHG-44 glioma cells, through inhibiting activity of the voltage-gated proton channel Hv1. Biochem. Biophys. Res. Commun. 2013, 438, 312–317. [Google Scholar] [CrossRef]
  164. Ribeiro-Silva, L.; Queiroz, F.O.; da Silva, A.M.; Hirata, A.E.; Arcisio-Miranda, M. Voltage-Gated Proton Channel in Human Glioblastoma Multiforme Cells. ACS Chem. Neurosci. 2016, 7, 864–869. [Google Scholar] [CrossRef] [PubMed]
  165. Hondares, E.; Brown, M.A.; Musset, B.; Morgan, D.; Cherny, V.V.; Taubert, C.; Bhamrah, M.K.; Coe, D.; Marelli-Berg, F.; Gribben, J.G.; et al. Enhanced activation of an amino-terminally truncated isoform of the voltage-gated proton channel HVCN1 enriched in malignant B cells. Proc. Natl. Acad. Sci. USA 2014, 111, 18078–18083. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Gavriliouk, D.; Scrimgeour, N.R.; Grigoryev, S.; Ma, L.; Zhou, F.H.; Barritt, G.J.; Rychkov, G.Y. Regulation of Orai1/STIM1 mediated ICRAC by intracellular pH. Sci. Rep. 2017, 7, 9829. [Google Scholar] [CrossRef] [PubMed]
  167. Kim, J.-H.; Lkhagvadorj, S.; Lee, M.-R.; Hwang, K.-H.; Chung, H.C.; Jung, J.H.; Cha, S.-K.; Eom, M. Orai1 and STIM1 are critical for cell migration and proliferation of clear cell renal cell carcinoma. Biochem. Biophys. Res. Commun. 2014, 448, 76–82. [Google Scholar] [CrossRef]
  168. Tsujikawa, H.; Yu, A.S.; Xie, J.; Yue, Z.; Yang, W.; He, Y.; Yue, L. Identification of key amino acid residues responsible for internal and external pH sensitivity of Orai1/STIM1 channels. Sci. Rep. 2015, 5, 16747. [Google Scholar] [CrossRef] [Green Version]
  169. 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] [Green Version]
  170. Zhu, H.; Zhang, H.; Jin, F.; Fang, M.; Huang, M.; Yang, C.S.; Chen, T.; Fu, L.; Pan, Z. Elevated Orai1 expression mediates tumor-promoting intracellular Ca2+ oscillations in human esophageal squamous cell carcinoma. Oncotarget 2014, 5, 3455–3471. [Google Scholar] [CrossRef] [Green Version]
  171. Zhu, M.; Chen, L.; Zhao, P.; Zhou, H.; Zhang, C.; Yu, S.; Lin, Y.; Yang, X. Store-operated Ca2+ entry regulates glioma cell migration and invasion via modulation of Pyk2 phosphorylation. J. Exp. Clin. Cancer Res. 2014, 33, 98. [Google Scholar] [CrossRef] [Green Version]
  172. Motiani, R.K.; Hyzinski-García, M.C.; Zhang, X.; Henkel, M.M.; Abdullaev, I.F.; Kuo, Y.-H.; Matrougui, K.; Mongin, A.A.; Trebak, M. STIM1 and Orai1 mediate CRAC channel activity and are essential for human glioblastoma invasion. Pflug. Arch. Eur. J. Physiol. 2013, 465, 1249–1260. [Google Scholar] [CrossRef] [Green Version]
  173. Zhan, Z.-Y.; Zhong, L.-X.; Feng, M.; Wang, J.-F.; Liu, D.-B.; Xiong, J.-P. Over-expression of Orai1 mediates cell proliferation and associates with poor prognosis in human non-small cell lung carcinoma. Int. J. Clin. Exp. Pathol. 2015, 8, 5080–5088. [Google Scholar]
  174. Umemura, M.; Baljinnyam, E.; Feske, S.; de Lorenzo, M.S.; Xie, L.-H.; Feng, X.; Oda, K.; Makino, A.; Fujita, T.; Yokoyama, U.; et al. Store-operated Ca2+ entry (SOCE) regulates melanoma proliferation and cell migration. PLoS ONE 2014, 9, e89292. [Google Scholar] [CrossRef] [PubMed]
  175. Kondratska, K.; Kondratskyi, A.; Yassine, M.; Lemonnier, L.; Lepage, G.; Morabito, A.; Skryma, R.; Prevarskaya, N. Orai1 and STIM1 mediate SOCE and contribute to apoptotic resistance of pancreatic adenocarcinoma. Biochim. Biophys. Acta 2014, 1843, 2263–2269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Xia, J.; Wang, H.; Huang, H.; Sun, L.; Dong, S.; Huang, N.; Shi, M.; Bin, J.; Liao, Y.; Liao, W. Elevated Orai1 and STIM1 expressions upregulate MACC1 expression to promote tumor cell proliferation, metabolism, migration, and invasion in human gastric cancer. Cancer Lett. 2016, 381, 31–40. [Google Scholar] [CrossRef] [PubMed]
  177. Diez-Bello, R.; Jardin, I.; Salido, G.M.; Rosado, J.A. Orai1 and Orai2 mediate store-operated calcium entry that regulates HL60 cell migration and FAK phosphorylation. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2017, 1864, 1064–1070. [Google Scholar] [CrossRef]
  178. Wu, Z.; Qing, J.; Xia, Y.; Wang, K.; Zhang, F. Suppression of stromal interaction molecule 1 inhibits SMMC7721 hepatocellular carcinoma cell proliferation by inducing cell cycle arrest. Biotechnol. Appl. Biochem. 2015, 62, 107–111. [Google Scholar] [CrossRef]
  179. Gui, L.; Wang, Z.; Han, J.; Ma, H.; Li, Z. High Expression of Orai1 Enhances Cell Proliferation and is Associated with Poor Prognosis in Human Colorectal Cancer. Clin. Lab. 2016, 62, 1689–1698. [Google Scholar] [CrossRef]
  180. Weidinger, C.; Shaw, P.J.; Feske, S. STIM1 and STIM2-mediated Ca2+ influx regulates antitumour immunity by CD8(+) T cells. EMBO Mol. Med. 2013, 5, 1311–1321. [Google Scholar] [CrossRef]
  181. Zhou, X.; Friedmann, K.S.; Lyrmann, H.; Zhou, Y.; Schoppmeyer, R.; Knorck, A.; Mang, S.; Hoxha, C.; Angenendt, A.; Backes, C.S.; et al. A calcium optimum for cytotoxic T lymphocyte and natural killer cell cytotoxicity. J. Physiol. 2018, 596, 2681–2698. [Google Scholar] [CrossRef] [Green Version]
  182. 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] [Green Version]
  183. Holzmann, C.; Kilch, T.; Kappel, S.; Armbrüster, A.; Jung, V.; Stöckle, M.; Bogeski, I.; Schwarz, E.C.; Peinelt, C. ICRAC controls the rapid androgen response in human primary prostate epithelial cells and is altered in prostate cancer. Oncotarget 2013, 4, 2096–2107. [Google Scholar] [CrossRef] [Green Version]
  184. 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]
  185. Yang, M.; James, A.D.; Suman, R.; Kasprowicz, R.; Nelson, M.; O’Toole, P.J.; Brackenbury, W.J. Voltage-dependent activation of Rac1 by Nav1.5 channels promotes cell migration. J. Cell Physiol. 2020, 235, 3950–3972. [Google Scholar] [CrossRef] [Green Version]
  186. Vilin, Y.Y.; Peters, C.H.; Ruben, P.C. Acidosis differentially modulates inactivation in NaV1.2, NaV1.4, and NaV1.5 channels. Front. Pharmacol. 2012, 3, 109. [Google Scholar] [CrossRef] [Green Version]
  187. Yang, X.-N.; Lu, Y.-P.; Liu, J.-J.; Huang, J.-K.; Liu, Y.-P.; Xiao, C.-X.; Jazag, A.; Ren, J.-L.; Guleng, B. Piezo1 is as a novel trefoil factor family 1 binding protein that promotes gastric cancer cell mobility in vitro. Dig. Dis. Sci. 2014, 59, 1428–1435. [Google Scholar] [CrossRef]
  188. Bae, C.; Sachs, F.; Gottlieb, P.A. Protonation of the human PIEZO1 ion channel stabilizes inactivation. J. Biol. Chem. 2015, 290, 5167–5173. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Kuntze, A.; Goetsch, O.; Fels, B.; Najder, K.; Unger, A.; Wilhelmi, M.; Sargin, S.; Schimmelpfennig, S.; Neumann, I.; Schwab, A.; et al. Protonation of Piezo1 Impairs Cell-Matrix Interactions of Pancreatic Stellate Cells. Front. Physiol. 2020, 11, 89. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Suzuki, T.; Muraki, Y.; Hatano, N.; Suzuki, H.; Muraki, K. PIEZO1 Channel Is a Potential Regulator of Synovial Sarcoma Cell-Viability. Int. J. Mol. Sci. 2018, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Chen, X.; Wanggou, S.; Bodalia, A.; Zhu, M.; Dong, W.; Fan, J.J.; Yin, W.C.; Min, H.-K.; Hu, M.; Draghici, D.; et al. A Feedforward Mechanism Mediated by Mechanosensitive Ion Channel PIEZO1 and Tissue Mechanics Promotes Glioma Aggression. Neuron 2018, 100, 799–815. [Google Scholar] [CrossRef] [Green Version]
  192. Li, C.; Rezania, S.; Kammerer, S.; Sokolowski, A.; Devaney, T.; Gorischek, A.; Jahn, S.; Hackl, H.; Groschner, K.; Windpassinger, C.; et al. Piezo1 forms mechanosensitive ion channels in the human MCF-7 breast cancer cell line. Sci. Rep. 2015, 5. [Google Scholar] [CrossRef] [Green Version]
  193. Han, Y.; Liu, C.; Zhang, D.; Men, H.; Huo, L.; Geng, Q.; Wang, S.; Gao, Y.; Zhang, W.; Zhang, Y.; et al. Mechanosensitive ion channel Piezo1 promotes prostate cancer development through the activation of the Akt/mTOR pathway and acceleration of cell cycle. Int. J. Oncol. 2019, 55, 629–644. [Google Scholar] [CrossRef] [Green Version]
  194. 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] [PubMed]
  195. 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. [Google Scholar] [CrossRef] [PubMed]
  196. Wu, Y.; Gao, B.; Xiong, Q.-J.; Wang, Y.-C.; Huang, D.-K.; Wu, W.-N. Acid-sensing ion channels contribute to the effect of extracellular acidosis on proliferation and migration of A549 cells. Tumor Biol. 2017, 39, 101042831770575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Adinolfi, E.; Raffaghello, L.; Giuliani, A.L.; Cavazzini, L.; Capece, M.; Chiozzi, P.; Bianchi, G.; Kroemer, G.; Pistoia, V.; Di Virgilio, F. Expression of P2X7 Receptor Increases In Vivo Tumor Growth. Cancer Res. 2012, 72, 2957–2969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Liu, X.; Ma, W.; Surprenant, A.; Jiang, L.-H. Identification of the amino acid residues in the extracellular domain of rat P2X7 receptor involved in functional inhibition by acidic pH. Br. J. Pharmacol. 2009, 156, 135–142. [Google Scholar] [CrossRef] [Green Version]
  199. Wang, Q.; Wang, L.; Feng, Y.-H.; Li, X.; Zeng, R.; Gorodeski, G.I. P2X7 receptor-mediated apoptosis of human cervical epithelial cells. Am. J. Physiol. Cell Physiol. 2004, 287, C1349–C1358. [Google Scholar] [CrossRef] [Green Version]
  200. Giannuzzo, A.; Pedersen, S.F.; Novak, I. The P2X7 receptor regulates cell survival, migration and invasion of pancreatic ductal adenocarcinoma cells. Mol. Cancer 2015, 14, 203. [Google Scholar] [CrossRef] [Green Version]
  201. Bianchi, G.; Vuerich, M.; Pellegatti, P.; Marimpietri, D.; Emionite, L.; Marigo, I.; Bronte, V.; Di Virgilio, F.; Pistoia, V.; Raffaghello, L. ATP/P2X7 axis modulates myeloid-derived suppressor cell functions in neuroblastoma microenvironment. Cell Death Dis. 2014, 5, e1135. [Google Scholar] [CrossRef] [Green Version]
  202. Starkus, J.G.; Varga, Z.; Schonherr, R.; Heinemann, S.H. Mechanisms of the inhibition of Shaker potassium channels by protons. Pflug. Arch. Eur. J. Physiol. 2003, 447, 44–54. [Google Scholar] [CrossRef]
  203. Ahn, D.S.; Hume, J.R. pH regulation of voltage-dependent K+ channels in canine pulmonary arterial smooth muscle cells. Pflug. Arch. Eur. J. Physiol. 1997, 433, 758–765. [Google Scholar] [CrossRef]
  204. Starace, D.M.; Bezanilla, F. A proton pore in a potassium channel voltage sensor reveals a focused electric field. Nature 2004, 427, 548–553. [Google Scholar] [CrossRef] [PubMed]
  205. Shorthouse, D.; Rahrmann, E.; Kosmidou, C.; Greenwood, B.; Hall, M.; Devonshire, G.; Gilbertson, R.; Fitzgerald, R.C.; Hall, B.A. KCNQ gene family members act as both tumor suppressors and oncogenes in gastrointestinal cancers. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  206. Mohr, C.J.; Steudel, F.A.; Gross, D.; Ruth, P.; Lo, W.Y.; Hoppe, R.; Schroth, W.; Brauch, H.; Huber, S.M.; Lukowski, R. Cancer-Associated Intermediate Conductance Ca2+-Activated K+ Channel KCa3.1. Cancers 2019, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Cook, D.L.; Ikeuchi, M.; Fujimoto, W.Y. Lowering of pHi inhibits Ca2+-activated K+ channels in pancreatic B-cells. Nature 1984, 311, 269–271. [Google Scholar] [CrossRef] [PubMed]
  208. Pedersen, K.A.; Jorgensen, N.K.; Jensen, B.S.; Olesen, S.P. Inhibition of the human intermediate-conductance, Ca2+-activated K+ channel by intracellular acidification. Pflug. Arch. Eur. J. Physiol. 2000, 440, 153–156. [Google Scholar] [CrossRef]
  209. Kume, H.; Takagi, K.; Satake, T.; Tokuno, H.; Tomita, T. Effects of intracellular pH on calcium-activated potassium channels in rabbit tracheal smooth muscle. J. Physiol. 1990, 424, 445–457. [Google Scholar] [CrossRef]
  210. Christensen, O.; Zeuthen, T. Maxi K+ channels in leaky epithelia are regulated by intracellular Ca2+, pH and membrane potential. Pflug. Arch. Eur. J. Physiol. 1987, 408, 249–259. [Google Scholar] [CrossRef]
  211. Paynter, J.J.; Shang, L.; Bollepalli, M.K.; Baukrowitz, T.; Tucker, S.J. Random mutagenesis screening indicates the absence of a separate H+-sensor in the pH-sensitive Kir channels. Channels 2010, 4, 390–397. [Google Scholar] [CrossRef] [Green Version]
  212. Qu, Z.; Zhu, G.; Yang, Z.; Cui, N.; Li, Y.; Chanchevalap, S.; Sulaiman, S.; Haynie, H.; Jiang, C. Identification of a critical motif responsible for gating of Kir2.3 channel by intracellular protons. J. Biol. Chem. 1999, 274, 13783–13789. [Google Scholar] [CrossRef] [Green Version]
  213. Olsen, M.L.; Sontheimer, H. Mislocalization of Kir channels in malignant glia. Glia 2004, 46, 63–73. [Google Scholar] [CrossRef] [Green Version]
  214. Lee, I.; Park, C.; Kang, W.K. Knockdown of inwardly rectifying potassium channel Kir2.2 suppresses tumorigenesis by inducing reactive oxygen species-mediated cellular senescence. Mol. Cancer Ther. 2010, 9, 2951–2959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  215. Tan, G.; Sun, S.Q.; Yuan, D.L. Expression of Kir 4.1 in human astrocytic tumors: Correlation with pathologic grade. Biochem. Biophys. Res. Commun. 2008, 367, 743–747. [Google Scholar] [CrossRef] [PubMed]
  216. DeHart, G.W.; Jin, T.; McCloskey, D.E.; Pegg, A.E.; Sheppard, D. The α9β1 integrin enhances cell migration by polyamine-mediated modulation of an inward-rectifier potassium channel. Proc. Natl. Acad. Sci. USA 2008, 105, 7188–7193. [Google Scholar] [CrossRef] [Green Version]
  217. Williams, S.; Bateman, A.; O’Kelly, I. Altered Expression of Two-Pore Domain Potassium (K2P) Channels in Cancer. PLoS ONE 2013, 8. [Google Scholar] [CrossRef] [PubMed]
  218. Ghatak, S.; Sikdar, S.K. Lactate modulates the intracellular pH sensitivity of human TREK1 channels. Pflug. Arch. Eur. J. Physiol. 2016, 468, 825–836. [Google Scholar] [CrossRef] [PubMed]
  219. Honoré, E. The neuronal background K2P channels: Focus on TREK1. Nat. Rev. Neurosci. 2007, 8, 251–261. [Google Scholar] [CrossRef] [PubMed]
  220. Innamaa, A.; Jackson, L.; Asher, V.; van Schalkwyk, G.; Warren, A.; Keightley, A.; Hay, D.; Bali, A.; Sowter, H.; Khan, R. Expression and effects of modulation of the K2P potassium channels TREK-1 (KCNK2) and TREK-2 (KCNK10) in the normal human ovary and epithelial ovarian cancer. Clin. Transl. Oncol. Off. Publ. Fed. Span. Oncol. Soc. Natl. Cancer Inst. Mex. 2013, 15, 910–918. [Google Scholar] [CrossRef] [PubMed]
  221. Sepúlveda, F.V.; Cid, L.P.; Teulon, J.; Niemeyer, M.I. Molecular aspects of structure, gating, and physiology of pH-sensitive background K2P and Kir K+-transport channels. Physiol. Rev. 2015, 95, 179–217. [Google Scholar] [CrossRef] [Green Version]
  222. Patel, A.J.; Lazdunski, M. The 2P-domain K+ channels: Role in apoptosis and tumorigenesis. Pflug. Arch. Eur. J. Physiol. 2004, 448, 261–273. [Google Scholar] [CrossRef]
  223. Bittner, S.; Budde, T.; Wiendl, H.; Meuth, S.G. From the background to the spotlight: TASK channels in pathological conditions. Brain Pathol. 2010, 20, 999–1009. [Google Scholar] [CrossRef]
  224. Czirják, G.; Enyedi, P. TASK-3 dominates the background potassium conductance in rat adrenal glomerulosa cells. Mol. Endocrinol. 2002, 16, 621–629. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Venkatachalam, K.; Montell, C. TRP Channels. Annu. Rev. Biochem. 2007, 76, 387–417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Glitsch, M. Protons and Ca2+: Ionic Allies in Tumor Progression? Physiology 2011, 26, 252–265. [Google Scholar] [CrossRef]
  227. Fels, B.; Bulk, E.; Petho, Z.; Schwab, A. The Role of TRP Channels in the Metastatic Cascade. Pharmaceuticals 2018, 11. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Ghilardi, J.R. Selective Blockade of the Capsaicin Receptor TRPV1 Attenuates Bone Cancer Pain. J. Neurosci. 2005, 25, 3126–3131. [Google Scholar] [CrossRef]
  229. Tominaga, M.; Caterina, M.J.; Malmberg, A.B.; Rosen, T.A.; Gilbert, H.; Skinner, K.; Raumann, B.E.; Basbaum, A.I.; Julius, D. The Cloned Capsaicin Receptor Integrates Multiple Pain-Producing Stimuli. Neuron 1998, 21, 531–543. [Google Scholar] [CrossRef] [Green Version]
  230. Jordt, S.-E.; Tominaga, M.; Julius, D. Acid potentiation of the capsaicin receptor determined by a key extracellular site. Proc. Natl. Acad. Sci. USA 2000, 97, 8134–8139. [Google Scholar] [CrossRef] [Green Version]
  231. Aneiros, E.; Cao, L.; Papakosta, M.; Stevens, E.B.; Phillips, S.; Grimm, C. The biophysical and molecular basis of TRPV1 proton gating. EMBO J. 2011, 30, 994–1002. [Google Scholar] [CrossRef]
  232. Hellwig, N.; Plant, T.D.; Janson, W.; Schafer, M.; Schultz, G.; Schaefer, M. TRPV1 acts as proton channel to induce acidification in nociceptive neurons. J. Biol. Chem. 2004, 279, 34553–34561. [Google Scholar] [CrossRef] [Green Version]
  233. Capiod, T. Cell proliferation, calcium influx and calcium channels. Biochimie 2011, 93, 2075–2079. [Google Scholar] [CrossRef]
  234. Stewart, T.A.; Yapa, K.T.D.S.; Monteith, G.R. Altered calcium signaling in cancer cells. Biochim. Biophys. Acta 2015, 1848, 2502–2511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Prevarskaya, N.; Ouadid-Ahidouch, H.; Skryma, R.; Shuba, Y. Remodelling of Ca2+ transport in cancer: How it contributes to cancer hallmarks? Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 20130097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Sánchez, M.G.; Sánchez, A.M.; Collado, B.; Malagarie-Cazenave, S.; Olea, N.; Carmena, M.J.; Prieto, J.C.; Díaz-Laviada, I. Expression of the transient receptor potential vanilloid 1 (TRPV1) in LNCaP and PC-3 prostate cancer cells and in human prostate tissue. Eur. J. Pharmacol. 2005, 515, 20–27. [Google Scholar] [CrossRef] [PubMed]
  237. Jankovic, B.; Loblaw, D.A.; Nam, R. Capsaicin may slow PSA doubling time: Case report and literature review. Can. Urol. Assoc. J. 2013, 4, 9. [Google Scholar] [CrossRef] [PubMed]
  238. Montcourrier, P.; Silver, I.; Farnoud, R.; Bird, I.; Rochefort, H. Breast cancer cells have a high capacity to acidify extracellular milieu by a dual mechanism. Clin. Exp. Metastasis 1997, 15, 382–392. [Google Scholar] [CrossRef] [PubMed]
  239. Granlund, K.L.; Tee, S.-S.; Vargas, H.A.; Lyashchenko, S.K.; Reznik, E.; Fine, S.; Laudone, V.; Eastham, J.A.; Touijer, K.A.; Reuter, V.E.; et al. Hyperpolarized MRI of Human Prostate Cancer Reveals Increased Lactate with Tumor Grade Driven by Monocarboxylate Transporter 1. Cell Metab. 2020, 31, 105–114. [Google Scholar] [CrossRef] [PubMed]
  240. Pecze, L.; Jósvay, K.; Blum, W.; Petrovics, G.; Vizler, C.; Oláh, Z.; Schwaller, B. Activation of endogenous TRPV1 fails to induce overstimulation-based cytotoxicity in breast and prostate cancer cells but not in pain-sensing neurons. Biochim. Biophys. Acta (BBA) Mol. Cell Res. 2016, 1863, 2054–2064. [Google Scholar] [CrossRef] [Green Version]
  241. Dömötör, A.; Peidl, Z.; Vincze, Á.; Hunyady, B.; Szolcsányi, J.; Kereskay, L.; Szekeres, G.; Mózsik, G. Immunohistochemical distribution of vanilloid receptor, calcitonin-gene related peptide and substance P in gastrointestinal mucosa of patients with different gastrointestinal disorders. InflammoPharmacology 2005, 13, 161–177. [Google Scholar] [CrossRef]
  242. Lazzeri, M.; Vannucchi, M.G.; Spinelli, M.; Bizzoco, E.; Beneforti, P.; Turini, D.; Faussone-Pellegrini, M.-S. Transient Receptor Potential Vanilloid Type 1 (TRPV1) Expression Changes from Normal Urothelium to Transitional Cell Carcinoma of Human Bladder. Eur. Urol. 2005, 48, 691–698. [Google Scholar] [CrossRef]
  243. Lee, W.H.; Choong, L.Y.; Mon, N.N.; Lu, S.; Lin, Q.; Pang, B.; Yan, B.; Krishna, V.S.R.; Singh, H.; Tan, T.Z.; et al. TRPV4 Regulates Breast Cancer Cell Extravasation, Stiffness and Actin Cortex. Sci. Rep. 2016, 6. [Google Scholar] [CrossRef] [Green Version]
  244. Gaunt, H.J.; Vasudev, N.S.; Beech, D.J. Transient receptor potential canonical 4 and 5 proteins as targets in cancer therapeutics. Eur. Biophys. J. 2016, 45, 611–620. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  245. Zhang, S.-S.; Wen, J.; Yang, F.; Cai, X.-L.; Yang, H.; Luo, K.-J.; Liu, Q.W.; Hu, R.-G.; Xie, X.; Huang, Q.-Y.; et al. High expression of Transient potential receptor C6 correlated with poor prognosis in patients with esophageal squamous cell carcinoma. Med. Oncol. 2013, 30. [Google Scholar] [CrossRef] [PubMed]
  246. Starkus, J.; Beck, A.; Fleig, A.; Penner, R. Regulation of TRPM2 by Extra- and Intracellular Calcium. J. Gen. Physiol. 2007, 130, 427–440. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Yang, W.; Zou, J.; Xia, R.; Vaal, M.L.; Seymour, V.A.; Luo, J.; Beech, D.J.; Jiang, L.-H. State-dependent Inhibition of TRPM2 Channel by Acidic pH. J. Biol. Chem. 2010, 285, 30411–30418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Du, J.; Xie, J.; Yue, L. Modulation of TRPM2 by acidic pH and the underlying mechanisms for pH sensitivity. J. Gen. Physiol. 2009, 134, 471–488. [Google Scholar] [CrossRef] [Green Version]
  249. Schumacker, P.T. Reactive oxygen species in cancer cells: Live by the sword, die by the sword. Cancer Cell 2006, 10, 175–176. [Google Scholar] [CrossRef] [Green Version]
  250. Miller, B.A. TRPM2 in Cancer. Cell Calcium 2019, 80, 8–17. [Google Scholar] [CrossRef]
  251. Li, M.; Du, J.; Jiang, J.; Ratzan, W.; Su, L.-T.; Runnels, L.W.; Yue, L. Molecular Determinants of Mg2+ and Ca2+ Permeability and pH Sensitivity in TRPM6 and TRPM7. J. Biol. Chem. 2007, 282, 25817–25830. [Google Scholar] [CrossRef] [Green Version]
  252. Jiang, J.; Li, M.; Yue, L. Potentiation of TRPM7 Inward Currents by Protons. J. Gen. Physiol. 2005, 126, 137–150. [Google Scholar] [CrossRef] [Green Version]
  253. Schmitz, C.; Perraud, A.-L.; Johnson, C.O.; Inabe, K.; Smith, M.K.; Penner, R.; Kurosaki, T.; Fleig, A.; Scharenberg, A.M. Regulation of Vertebrate Cellular Mg2+ Homeostasis by TRPM7. Cell 2003, 114, 191–200. [Google Scholar] [CrossRef] [Green Version]
  254. Zierler, S.; Yao, G.; Zhang, Z.; Kuo, W.C.; Pörzgen, P.; Penner, R.; Horgen, F.D.; Fleig, A. Waixenicin A Inhibits Cell Proliferation through Magnesium-dependent Block of Transient Receptor Potential Melastatin 7 (TRPM7) Channels. J. Biol. Chem. 2011, 286, 39328–39335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Macianskiene, R.; Almanaityte, M.; Jekabsone, A.; Mubagwa, K. Modulation of Human Cardiac TRPM7 Current by Extracellular Acidic pH Depends upon Extracellular Concentrations of Divalent Cations. PLoS ONE 2017, 12, e0170923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Musset, B.; Decoursey, T. Biophysical properties of the voltage gated proton channel H(V)1. Wiley Interdiscip. Rev. Membr. Transp. Signal 2012, 1, 605–620. [Google Scholar] [CrossRef] [Green Version]
  257. DeCoursey, T.E. Voltage and pH sensing by the voltage-gated proton channel, HV1. J. R. Soc. Interface 2018, 15. [Google Scholar] [CrossRef] [Green Version]
  258. DeCoursey, T.E.; Cherny, V.V. Potential, pH, and arachidonate gate hydrogen ion currents in human neutrophils. Biophys. J. 1993, 65, 1590–1598. [Google Scholar] [CrossRef] [Green Version]
  259. De-la-Rosa, V.; Suarez-Delgado, E.; Rangel-Yescas, G.E.; Islas, L.D. Currents through HV1 channels deplete protons in their vicinity. J. Gen. Physiol. 2016, 147, 127–136. [Google Scholar] [CrossRef]
  260. Berridge, M.J. Calcium microdomains: Organization and function. Cell Calcium 2006, 40, 405–412. [Google Scholar] [CrossRef]
  261. Pathak, M.M.; Tran, T.; Hong, L.; Joos, B.; Morris, C.E.; Tombola, F. The Hv1 proton channel responds to mechanical stimuli. J. Gen. Physiol. 2016, 148, 405–418. [Google Scholar] [CrossRef] [Green Version]
  262. Petho, Z.; Najder, K.; Bulk, E.; Schwab, A. Mechanosensitive ion channels push cancer progression. Cell Calcium 2019, 80, 79–90. [Google Scholar] [CrossRef]
  263. Stock, C.; Schwab, A. Protons make tumor cells move like clockwork. Pflug. Arch. Eur. J. Physiol. 2009, 458, 981–992. [Google Scholar] [CrossRef]
  264. Boedtkjer, E.; Bunch, L.; Pedersen, S.F. Physiology, pharmacology and pathophysiology of the pH regulatory transport proteins NHE1 and NBCn1: Similarities, differences, and implications for cancer therapy. Curr. Pharm. Des. 2012, 18, 1345–1371. [Google Scholar] [CrossRef] [PubMed]
  265. Spugnini, E.P.; Sonveaux, P.; Stock, C.; Perez-Sayans, M.; de Milito, A.; Avnet, S.; Garcia, A.G.; Harguindey, S.; Fais, S. Proton channels and exchangers in cancer. Biochim. Biophys. Acta 2015, 1848, 2715–2726. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  266. Krysiak, K.; Gomez, F.; White, B.S.; Matlock, M.; Miller, C.A.; Trani, L.; Fronick, C.C.; Fulton, R.S.; Kreisel, F.; Cashen, A.F.; et al. Recurrent somatic mutations affecting B-cell receptor signaling pathway genes in follicular lymphoma. Blood 2017, 129, 473–483. [Google Scholar] [CrossRef] [PubMed]
  267. Gosselin-Badaroudine, P.; Delemotte, L.; Moreau, A.; Klein, M.L.; Chahine, M. Gating pore currents and the resting state of Nav1.4 voltage sensor domains. Proc. Natl. Acad. Sci. USA 2012, 109, 19250–19255. [Google Scholar] [CrossRef] [Green Version]
  268. Jones, D.K.; Peters, C.H.; Allard, C.R.; Claydon, T.W.; Ruben, P.C. Proton sensors in the pore domain of the cardiac voltage-gated sodium channel. J. Biol. Chem. 2013, 288, 4782–4791. [Google Scholar] [CrossRef] [Green Version]
  269. Peters, C.H.; Ghovanloo, M.R.; Gershome, C.; Ruben, P.C. pH Modulation of Voltage-Gated Sodium Channels. Handb. Exp. Pharm. 2018, 246, 147–160. [Google Scholar] [CrossRef]
  270. Ghovanloo, M.R.; Peters, C.H.; Ruben, P.C. Effects of acidosis on neuronal voltage-gated sodium channels: Nav1.1 and Nav1.3. Channels 2018, 12, 367–377. [Google Scholar] [CrossRef] [Green Version]
  271. Khan, A.; Kyle, J.W.; Hanck, D.A.; Lipkind, G.M.; Fozzard, H.A. Isoform-dependent interaction of voltage-gated sodium channels with protons. J. Physiol. 2006, 576, 493–501. [Google Scholar] [CrossRef]
  272. Nonner, W.; Spalding, B.C.; Hille, B. Low intracellular pH and chemical agents slow inactivation gating in sodium channels of muscle. Nature 1980, 284, 360–363. [Google Scholar] [CrossRef]
  273. House, C.D.; Vaske, C.J.; Schwartz, A.M.; Obias, V.; Frank, B.; Luu, T.; Sarvazyan, N.; Irby, R.; Strausberg, R.L.; Hales, T.G.; et al. Voltage-gated Na+ channel SCN5A is a key regulator of a gene transcriptional network that controls colon cancer invasion. Cancer Res. 2010, 70, 6957–6967. [Google Scholar] [CrossRef] [Green Version]
  274. Yildirim, S.; Altun, S.; Gumushan, H.; Patel, A.; Djamgoz, M.B.A. Voltage-gated sodium channel activity promotes prostate cancer metastasis in vivo. Cancer Lett. 2012, 323, 58–61. [Google Scholar] [CrossRef]
  275. Fraser, S.P.; Ozerlat-Gunduz, I.; Brackenbury, W.J.; Fitzgerald, E.M.; Campbell, T.M.; Coombes, R.C.; Djamgoz, M.B. Regulation of voltage-gated sodium channel expression in cancer: Hormones, growth factors and auto-regulation. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2014, 369, 20130105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  276. Gillet, L.; Roger, S.; Besson, P.; Lecaille, F.; Gore, J.; Bougnoux, P.; Lalmanach, G.; Le Guennec, J.Y. Voltage-gated Sodium Channel Activity Promotes Cysteine Cathepsin-dependent Invasiveness and Colony Growth of Human Cancer Cells. J. Biol. Chem. 2009, 284, 8680–8691. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  277. Gradek, F.; Lopez-Charcas, O.; Chadet, S.; Poisson, L.; Ouldamer, L.; Goupille, C.; Jourdan, M.L.; Chevalier, S.; Moussata, D.; Besson, P.; et al. Sodium Channel Nav1.5 Controls Epithelial-to-Mesenchymal Transition and Invasiveness in Breast Cancer Cells Through its Regulation by the Salt-Inducible Kinase-1. Sci. Rep. 2019, 9, 18652. [Google Scholar] [CrossRef] [PubMed]
  278. Onkal, R.; Fraser, S.; Djamgoz, M. Cationic Modulation of Voltage-Gated Sodium Channel (Nav1.5): Neonatal Versus Adult Splice Variants—1. Monovalent H+ Ions. Bioelectricity 2019, 1, 139–147. [Google Scholar] [CrossRef] [Green Version]
  279. 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]
  280. Ou, S.W.; Kameyama, A.; Hao, L.Y.; Horiuchi, M.; Minobe, E.; Wang, W.Y.; Makita, N.; Kameyama, M. Tetrodotoxin-resistant Na+ channels in human neuroblastoma cells are encoded by new variants of Nav1.5/SCN5A. Eur. J. Neurosci. 2005, 22, 793–801. [Google Scholar] [CrossRef]
  281. Nelson, M.; Yang, M.; Dowle, A.A.; Thomas, J.R.; Brackenbury, W.J. The sodium channel-blocking antiepileptic drug phenytoin inhibits breast tumour growth and metastasis. Mol. Cancer 2015, 14, 13. [Google Scholar] [CrossRef] [Green Version]
  282. Djamgoz, M.B.A.; Onkal, R. Persistent current blockers of voltage-gated sodium channels: A clinical opportunity for controlling metastatic disease. Recent Pat. Anticancer. Drug Discov. 2013, 8, 66–84. [Google Scholar] [CrossRef]
  283. Coste, B.; Mathur, J.; Schmidt, M.; Earley, T.J.; Ranade, S.; Petrus, M.J.; Dubin, A.E.; Patapoutian, A. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science 2010, 330, 55–60. [Google Scholar] [CrossRef] [Green Version]
  284. Ridone, P.; Vassalli, M.; Martinac, B. Piezo1 mechanosensitive channels: What are they and why are they important. Biophys. Rev. 2019, 11, 795–805. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Beech, D.J.; Xiao, B. Piezo channel mechanisms in health and disease. J. Physiol. 2018, 596, 965–967. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  286. Boris Martinac; Charles, D. Cox. Mechanosensory Transduction: Focus on Ion Channels. In Reference Module in Life Sciences: Comprehensive Biophysics; Elsevier: Amsterdam, The Netherlands, 2016. [Google Scholar] [CrossRef]
  287. Sun, Y.; Li, M.; Liu, G.; Zhang, X.; Zhi, L.; Zhao, J.; Wang, G. The function of Piezo1 in colon cancer metastasis and its potential regulatory mechanism. J. Cancer Res. Clin. Oncol. 2020, 146, 1139–1152. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  288. Luo, M.; Ho, K.K.Y.; Tong, Z.; Deng, L.; Liu, A.P. Compressive Stress Enhances Invasive Phenotype of Cancer Cells via Piezo1 Activation. bioRxiv 2019. [Google Scholar] [CrossRef] [Green Version]
  289. Kellenberger, S.; Schild, L. Epithelial sodium channel/degenerin family of ion channels: A variety of functions for a shared structure. Physiol. Rev. 2002, 82, 735–767. [Google Scholar] [CrossRef] [Green Version]
  290. Pattison, L.A.; Callejo, G.; St John Smith, E. Evolution of acid nociception: Ion channels and receptors for detecting acid. Philos. Trans. R. Soc. Lond. B Biol. Sci. 2019, 374, 20190291. [Google Scholar] [CrossRef]
  291. Waldmann, R.; Champigny, G.; Bassilana, F.; Heurteaux, C.; Lazdunski, M. A proton-gated cation channel involved in acid-sensing. Nature 1997, 386, 173–177. [Google Scholar] [CrossRef]
  292. Vullo, S.; Kellenberger, S. A molecular view of the function and pharmacology of acid-sensing ion channels. Pharmacol. Res. 2020, 154, 104166. [Google Scholar] [CrossRef] [Green Version]
  293. Stoop, R.; Surprenant, A.; North, R.A. Different Sensitivities to pH of ATP-Induced Currents at Four Cloned P2X Receptors. J. Neurophysiol. 1997, 78, 1837–1840. [Google Scholar] [CrossRef] [Green Version]
  294. Sekar, P.; Huang, D.-Y.; Chang, S.-F.; Lin, W.-W. Coordinate effects of P2X7 and extracellular acidification in microglial cells. Oncotarget 2018, 9, 12718–12731. [Google Scholar] [CrossRef]
  295. Giannuzzo, A.; Saccomano, M.; Napp, J.; Ellegaard, M.; Alves, F.; Novak, I. Targeting of the P2X7 receptor in pancreatic cancer and stellate cells. Int. J. Cancer 2016, 139, 2540–2552. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  296. Adinolfi, E.; Capece, M.; Amoroso, F.; Marchi, E.; Franceschini, A. Emerging Roles of P2X Receptors in Cancer. Curr. Med. Chem. 2015, 22, 878–890. [Google Scholar] [CrossRef] [PubMed]
  297. Engblom, C.; Pfirschke, C.; Pittet, M.J. The role of myeloid cells in cancer therapies. Nat. Rev. Cancer 2016, 16, 447–462. [Google Scholar] [CrossRef] [PubMed]
  298. Beatty, G.L.; Chiorean, E.G.; Fishman, M.P.; Saboury, B.; Teitelbaum, U.R.; Sun, W.; Huhn, R.D.; Song, W.; Li, D.; Sharp, L.L.; et al. CD40 Agonists Alter Tumor Stroma and Show Efficacy Against Pancreatic Carcinoma in Mice and Humans. Science 2011, 331, 1612–1616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  299. Vesely, M.D.; Kershaw, M.H.; Schreiber, R.D.; Smyth, M.J. Natural Innate and Adaptive Immunity to Cancer. Annu. Rev. Immunol. 2011, 29, 235–271. [Google Scholar] [CrossRef] [Green Version]
  300. Najder, K.; Rugi, M.; Lebel, M.; Schröder, J.; Oster, L.; Pethő, Z.; Bulk, E.; Schwab, A. Role of the Intracellular Sodium Homeostasis in Chemotaxis of Activated Murine Neutrophils. Front Immunol 2020. [Google Scholar] [CrossRef]
  301. Woo, S.-R.; Corrales, L.; Gajewski, T.F. Innate Immune Recognition of Cancer. Annu. Rev. Immunol. 2015, 33, 445–474. [Google Scholar] [CrossRef]
  302. Gentles, A.J.; Newman, A.M.; Liu, C.L.; Bratman, S.V.; Feng, W.; Kim, D.; Nair, V.S.; Xu, Y.; Khuong, A.; Hoang, C.D.; et al. The prognostic landscape of genes and infiltrating immune cells across human cancers. Nat. Med. 2015, 21, 938–945. [Google Scholar] [CrossRef]
  303. Gajewski, T.F.; Schreiber, H.; Fu, Y.-X. Innate and adaptive immune cells in the tumor microenvironment. Nat. Immunol. 2013, 14, 1014–1022. [Google Scholar] [CrossRef] [Green Version]
  304. Wu, L.; Saxena, S.; Awaji, M.; Singh, R.K. Tumor-Associated Neutrophils in Cancer: Going Pro. Cancers 2019, 11. [Google Scholar] [CrossRef] [Green Version]
  305. Yamamoto, S.; Shimizu, S.; Kiyonaka, S.; Takahashi, N.; Wajima, T.; Hara, Y.; Negoro, T.; Hiroi, T.; Kiuchi, Y.; Okada, T.; et al. TRPM2-mediated Ca2+ influx induces chemokine production in monocytes that aggravates inflammatory neutrophil infiltration. Nat. Med. 2008, 14, 738–747. [Google Scholar] [CrossRef] [PubMed]
  306. Wang, G.; Cao, L.; Liu, X.; Sieracki, N.A.; Di, A.; Wen, X.; Chen, Y.; Taylor, S.; Huang, X.; Tiruppathi, C.; et al. Oxidant Sensing by TRPM2 Inhibits Neutrophil Migration and Mitigates Inflammation. Dev. Cell 2016, 38, 453–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Wang, C.-h.; Rong, M.-y.; Wang, L.; Ren, Z.; Chen, L.-n.; Jia, J.-f.; Li, X.-y.; Wu, Z.-b.; Chen, Z.-n.; Zhu, P. CD147 up-regulates calcium-induced chemotaxis, adhesion ability and invasiveness of human neutrophils via a TRPM-7-mediated mechanism. Rheumatology 2014, 53, 2288–2296. [Google Scholar] [CrossRef] [PubMed]
  308. Nadolni, W.; Zierler, S. The Channel-Kinase TRPM7 as Novel Regulator of Immune System Homeostasis. Cells 2018, 7, 109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  309. Najder, K.; Musset, B.; Lindemann, O.; Bulk, E.; Schwab, A.; Fels, B. The function of TRP channels in neutrophil granulocytes. Pflug. Arch. Eur. J. Physiol. 2018, 470, 1017–1033. [Google Scholar] [CrossRef] [PubMed]
  310. Musset, B.; Cherny, V.V.; Morgan, D.; DeCoursey, T.E. The intimate and mysterious relationship between proton channels and NADPH oxidase. FEBS Lett. 2009, 583, 7–12. [Google Scholar] [CrossRef] [Green Version]
  311. Biswas, S.K.; Mantovani, A. Macrophage plasticity and interaction with lymphocyte subsets: Cancer as a paradigm. Nat. Immunol. 2010, 11, 889–896. [Google Scholar] [CrossRef]
  312. Schilling, T.; Miralles, F.; Eder, C. TRPM7 regulates proliferation and polarisation of macrophages. J. Cell Sci. 2014, 127, 4561–4566. [Google Scholar] [CrossRef] [Green Version]
  313. Beceiro, S.; Radin, J.N.; Chatuvedi, R.; Piazuelo, M.B.; Horvarth, D.J.; Cortado, H.; Gu, Y.; Dixon, B.; Gu, C.; Lange, I.; et al. TRPM2 ion channels regulate macrophage polarization and gastric inflammation during Helicobacter pylori infection. Mucosal Immunol. 2017, 10, 493–507. [Google Scholar] [CrossRef]
  314. Chauhan, A.; Sun, Y.; Sukumaran, P.; Quenum Zangbede, F.O.; Jondle, C.N.; Sharma, A.; Evans, D.L.; Chauhan, P.; Szlabick, R.E.; Aaland, M.O.; et al. M1 Macrophage Polarization Is Dependent on TRPC1-Mediated Calcium Entry. iScience 2018, 8, 85–102. [Google Scholar] [CrossRef] [Green Version]
  315. Di Virgilio, F.; Dal Ben, D.; Sarti, A.C.; Giuliani, A.L.; Falzoni, S. The P2X7 Receptor in Infection and Inflammation. Immunity 2017, 47, 15–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  316. Wiley, J.S.; Sluyter, R.; Gu, B.J.; Stokes, L.; Fuller, S.J. The human P2X7 receptor and its role in innate immunity. Tissue Antigens 2011, 78, 321–332. [Google Scholar] [CrossRef] [PubMed]
  317. Bronte, V.; Brandau, S.; Chen, S.-H.; Colombo, M.P.; Frey, A.B.; Greten, T.F.; Mandruzzato, S.; Murray, P.J.; Ochoa, A.; Ostrand-Rosenberg, S.; et al. Recommendations for myeloid-derived suppressor cell nomenclature and characterization standards. Nat. Commun. 2016, 7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  318. Hegde, V.L.; Nagarkatti, P.S.; Nagarkatti, M. Role of Myeloid-Derived Suppressor Cells in Amelioration of Experimental Autoimmune Hepatitis Following Activation of TRPV1 Receptors by Cannabidiol. PLoS ONE 2011, 6, e18281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  319. Freitas, C.M.T.; Johnson, D.K.; Weber, K.S. T Cell Calcium Signaling Regulation by the Co-Receptor CD5. Int. J. Mol. Sci. 2018, 19. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  320. Fox, C.J.; Hammerman, P.S.; Thompson, C.B. Fuel feeds function: Energy metabolism and the T-cell response. Nat. Rev. Immunol. 2005, 5, 844–852. [Google Scholar] [CrossRef]
  321. Menk, A.V.; Scharping, N.E.; Moreci, R.S.; Zeng, X.; Guy, C.; Salvatore, S.; Bae, H.; Xie, J.; Young, H.A.; Wendell, S.G.; et al. Early TCR Signaling Induces Rapid Aerobic Glycolysis Enabling Distinct Acute T Cell Effector Functions. Cell Rep. 2018, 22, 1509–1521. [Google Scholar] [CrossRef] [Green Version]
  322. Vaeth, M.; Maus, M.; Klein-Hessling, S.; Freinkman, E.; Yang, J.; Eckstein, M.; Cameron, S.; Turvey, S.E.; Serfling, E.; Berberich-Siebelt, F.; et al. Store-Operated Ca2+ Entry Controls Clonal Expansion of T Cells through Metabolic Reprogramming. Immunity 2017, 47, 664–679. [Google Scholar] [CrossRef] [Green Version]
  323. Fischer, K.; Hoffmann, P.; Voelkl, S.; Meidenbauer, N.; Ammer, J.; Edinger, M.; Gottfried, E.; Schwarz, S.; Rothe, G.; Hoves, S.; et al. Inhibitory effect of tumor cell-derived lactic acid on human T cells. Blood 2007, 109, 3812–3819. [Google Scholar] [CrossRef]
  324. Chang, C.H.; Qiu, J.; O’Sullivan, D.; Buck, M.D.; Noguchi, T.; Curtis, J.D.; Chen, Q.; Gindin, M.; Gubin, M.M.; van der Windt, G.J.; et al. Metabolic Competition in the Tumor Microenvironment Is a Driver of Cancer Progression. Cell 2015, 162, 1229–1241. [Google Scholar] [CrossRef] [Green Version]
  325. Feske, S.; Skolnik, E.Y.; Prakriya, M. Ion channels and transporters in lymphocyte function and immunity. Nat. Rev. Immunol. 2012, 12, 532–547. [Google Scholar] [CrossRef] [Green Version]
  326. Feske, S.; Wulff, H.; Skolnik, E.Y. Ion channels in innate and adaptive immunity. Annu. Rev. Immunol. 2015, 33, 291–353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  327. Feske, S.; Prakriya, M.; Rao, A.; Lewis, R.S. A severe defect in CRAC Ca2+ channel activation and altered K+ channel gating in T cells from immunodeficient patients. J. Exp. Med. 2005, 202, 651–662. [Google Scholar] [CrossRef] [PubMed]
  328. Hou, P.; Zhang, R.; Liu, Y.; Feng, J.; Wang, W.; Wu, Y.; Ding, J. Physiological role of Kv1.3 channel in T lymphocyte cell investigated quantitatively by kinetic modeling. PLoS ONE 2014, 9, e89975. [Google Scholar] [CrossRef] [Green Version]
  329. Malayev, A.; Nelson, D.J. Extracellular pH modulates the Ca2+ current activated by depletion of intracellular Ca2+ stores in human macrophages. J. Membr. Biol. 1995, 146. [Google Scholar] [CrossRef] [PubMed]
  330. Demaurex, N.; Nunes, P. The role of STIM and ORAI proteins in phagocytic immune cells. Am. J. Physiol. Cell Physiol. 2016, 310, C496–C508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  331. Feske, S.; Gwack, Y.; Prakriya, M.; Srikanth, S.; Puppel, S.H.; Tanasa, B.; Hogan, P.G.; Lewis, R.S.; Daly, M.; Rao, A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature 2006, 441, 179–185. [Google Scholar] [CrossRef]
  332. Angenendt, A.; Steiner, R.; Knorck, A.; Schwar, G.; Konrad, M.; Krause, E.; Lis, A. Orai, STIM, and PMCA contribute to reduced calcium signal generation in CD8+ T cells of elderly mice. Aging 2020, 12, 3266–3286. [Google Scholar] [CrossRef]
  333. Schwarz, E.C.; Qu, B.; Hoth, M. Calcium, cancer and killing: The role of calcium in killing cancer cells by cytotoxic T lymphocytes and natural killer cells. Biochim. Biophys. Acta 2013, 1833, 1603–1611. [Google Scholar] [CrossRef] [Green Version]
  334. Modiano, J.F.; Kelepouris, E.; Kern, J.A.; Nowell, P.C. Requirement for extracellular calcium or magnesium in mitogen-induced activation of human peripheral blood lymphocytes. J. Cell Physiol. 1988, 135, 451–458. [Google Scholar] [CrossRef]
  335. Kuras, Z.; Yun, Y.H.; Chimote, A.A.; Neumeier, L.; Conforti, L. KCa3.1 and TRPM7 channels at the uropod regulate migration of activated human T cells. PLoS ONE 2012, 7, e43859. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. Philipp, S.; Strauss, B.; Hirnet, D.; Wissenbach, U.; Mery, L.; Flockerzi, V.; Hoth, M. TRPC3 mediates T-cell receptor-dependent calcium entry in human T-lymphocytes. J. Biol. Chem. 2003, 278, 26629–26638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  337. Wu, Q.Y.; Sun, M.R.; Wu, C.L.; Li, Y.; Du, J.J.; Zeng, J.Y.; Bi, H.L.; Sun, Y.H. Activation of calcium-sensing receptor increases TRPC3/6 expression in T lymphocyte in sepsis. Mol. Immunol. 2015, 64, 18–25. [Google Scholar] [CrossRef] [PubMed]
  338. Rao, G.K.; Kaminski, N.E. Induction of intracellular calcium elevation by Delta9-tetrahydrocannabinol in T cells involves TRPC1 channels. J. Leukoc. Biol. 2006, 79, 202–213. [Google Scholar] [CrossRef]
  339. Gamberucci, A.; Giurisato, E.; Pizzo, P.; Tassi, M.; Giunti, R.; McIntosh, D.P.; Benedetti, A. Diacylglycerol activates the influx of extracellular cations in T-lymphocytes independently of intracellular calcium-store depletion and possibly involving endogenous TRP6 gene products. Biochem. J. 2002, 364, 245–254. [Google Scholar] [CrossRef]
  340. Majhi, R.K.; Sahoo, S.S.; Yadav, M.; Pratheek, B.M.; Chattopadhyay, S.; Goswami, C. Functional expression of TRPV channels in T cells and their implications in immune regulation. FEBS J. 2015, 282, 2661–2681. [Google Scholar] [CrossRef]
  341. Capasso, M.; Bhamrah, M.K.; Henley, T.; Boyd, R.S.; Langlais, C.; Cain, K.; Dinsdale, D.; Pulford, K.; Khan, M.; Musset, B.; et al. HVCN1 modulates BCR signal strength via regulation of BCR-dependent generation of reactive oxygen species. Nat. Immunol. 2010, 11, 265–272. [Google Scholar] [CrossRef]
  342. Calcinotto, A.; Filipazzi, P.; Grioni, M.; Iero, M.; de Milito, A.; Ricupito, A.; Cova, A.; Canese, R.; Jachetti, E.; Rossetti, M.; et al. Modulation of microenvironment acidity reverses anergy in human and murine tumor-infiltrating T lymphocytes. Cancer Res. 2012, 72, 2746–2756. [Google Scholar] [CrossRef] [Green Version]
  343. Chimote, A.A.; Hajdu, P.; Kucher, V.; Boiko, N.; Kuras, Z.; Szilagyi, O.; Yun, Y.H.; Conforti, L. Selective inhibition of KCa3.1 channels mediates adenosine regulation of the motility of human T cells. J. Immunol. 2013, 191, 6273–6280. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Response of ion channels to an acidic environment. The left side indicates physiological interstitial pH condition (pHo = 7.4), whereas the right side depicts an acidic environment prevalent in tumors (pHo = 6.5). The size of the arrows represents ion flux rate.
Figure 1. Response of ion channels to an acidic environment. The left side indicates physiological interstitial pH condition (pHo = 7.4), whereas the right side depicts an acidic environment prevalent in tumors (pHo = 6.5). The size of the arrows represents ion flux rate.
Cancers 12 02484 g001
Table 1. pH-dependence of ion channels involved in cancer.
Table 1. pH-dependence of ion channels involved in cancer.
Gene NameChannel NameCancer TypeFunctions in CancerspHo AcidificationpHi Acidification
KCNA3KV1.3colorectal (cell line)↑ migration, adhesiveness [88]slight inhibition [89,90]
breast (cell line)↑ migration, adhesiveness, proliferation [88,91]
melanoma (T lymphocyte)↑ antitumor immunity in tumor-bearing mice [92]
head and neck (T lymphocyte)defects are associated with diminished cytotoxicity [93]
KCNA5KV1.5lung (cell line, tumor)apoptosis, H2O2 production, ↓ tumor growth [27]strong inhibition, pKa = 7.2 [94]inhibition [27,95]
cervical (cell line)apoptosis, H2O2 production [95]
gastric (cell line)↑ proliferation [96]
KCNB1KV2.1uterine (cell line)↑ cell cycle progression [97]slight inhibition [94]
KCNQ1KV.7.1colorectal (cell line)↑ proliferation [98]slight inhibition, pKa = 5.5 [99]inhibition, BUT stimulation in presence of KCNE1 [100]
KCNH1KV10.1 (hEAG1)glioblastoma (cell line, tumor)↑ tumor growth and metastasis [101]inhibition, pKa ≈ 6.5 [102]
colorectal (cell line)↑ proliferation [103]
gastric (cell line)↑ proliferation [104]
breast (cell line, tumor)↑ tumor progression and hypoxia signaling [105]
osteosarcoma (cell line)↑ proliferation, cell cycle progression [106]
osteosarcoma (endothelium)knockdown inhibitis tumor angiogenesis and VEGF-signaling [107]
KCNH2KV11.1 (hERG1)pancreatic (cell line)↑ EGFR-signaling, proliferation, migration [85]slight inhibition, pKa = 5.1 [45,108]slight stimulation [109]
gastric (cell line, tumor)↑ PI3K/Akt pathway, promotes angiogenesis, proliferation, migration and tumor growth [110]
KCNN4KCa3.1glioblastoma (cell line, tumor)↑ migration, swelling, invasion, radiotherapy resistance, tumor growth (reviewed in [111]possibly insensitive between pH 6.5 and pH 8 [112]inhibition [113]
breast (cell line, tumor)↑ cell cycle progression, proliferation and tumor growth [114,115]
pancreatic (cell line)↑ migration, invasion, proliferation [116]inhibition [113]
pancreatic (stromal cell)↑ migration, chemotaxis [117]
melanoma (T lymphocyte)↑ antitumor immunity in tumor-bearing mice [92,118]
head and neck cancer (T lymphocyte)↑ tumor infiltration [119,120]
KCNJ1Kir1.1renal (cell line)↓ cancer cell growth, invasion and promotes cancer cell apoptosis [121]strong inhibition pKa = 6.9 [122]
KCNJ10Kir4.1glioblastoma (cell line)↓ cell cycle progression [123]inhibition pKa = 6 [124]
KCNK2K2P2.1 (TREK1)prostate (cell line)↑ proliferation, contributes to setting the membrane potential [125]strong inhibition pKa = 7.35 [126]slight stimulation [127]
pancreatic (cell line)↑ proliferation, migration, contributes to setting the membrane potential [128]
KCNK3, KCNK9K2P3.1(TASK1), K2P9.1(TASK3)glioblastoma (cell line)↑ apoptosis and hypoxia resistance, contribute to setting the membrane potential [129]inhibition [126]
lung (cell line)↑ proliferation, contribute to setting the membrane potential [130]
KCNK5K2P5.1(TASK2)breast (cell line)↑ proliferation, contribute to setting the membrane potential [131]inhibition [132]inhibition [132,133]
TRPC4, TRPC5TRPC4, TRPC5renal (cell line)activation leads to cancer cell death [134]stimulation [135]
TRPC6TRPC6glioblastoma (U373MG cell line)↑ cancer cell proliferation, invasion, angiogenesis [136]slight inhibition [135]
liver (cell line)↑ proliferation [137]
pancreatic (tumor-associated immune cell)↑ CXCL1/CXCR2 induced neutrophil infiltration [138,139]
pancreatic (stromal cell)↑ migration, hypoxia response [140]
TRPM2TRPM2breast (cell lines and murine model)mediates H2O2 cytotoxicity [141,142]inhibition pKa = 6.5 [143]complete inhibition by pH = 6 [143]
prostate (cell lines)↑ proliferation [144]
TRPM7TRPM7nasopharyngeal (cell lines)↑ migration [145] inhibition pKa = 6.32 [146]
lung (A549 cell line)↑ migration [147]
pancreatic (cell line)↑ migration [148]
breast↑ proliferation and migration (overexpressed in cancer tissue) [149]
TRPV1TRPV1breast (cell line)apoptosis, tumor suppressor (mRNA overexpression) [150]stimulation [151]
gastric and colorectal (cell line)apoptosis, tumor suppressor (decreased expression in cancer tissues) [152]
astrocytoma (murine model)apoptosis, tumor suppressor [153]
endometrial (cell line)apoptosis, tumor suppressor [154]
TRPV4TRPV4lung (murine model)↓ endothelial tumor cell migration, normalization of tumor vascularization [155]stimulation [156]
breast (tumor-derived EC, cancer cell line)↑ promotes tumor-derived endothelial cell migration and metastasis [157]
hepatocellular carcinoma (cell line and murine model)TRPV4 inhibition leads to cell apoptosis [158]
gastric (cell line)↑ cancer proliferation, migration and invasion [159]
HVCN1HV1breast (cell line, tumor)promotes cell migration, tumor growth, pHi regulation [160,161]pHo/pHi relation and membrane-potential dependent, detailed in Section 4.3
colorectal (cell line)promotes cell migration, pHi regulation [162,163]
glioblastoma (cell line)promotes cell migration, survival, pHi regulation [164]
B-cell lymphoma (patient samples, cell line)short variant of HV1 promotes cell migration and proliferation [165]
CRAC channel (ORAI-STIM) inhibition pKa = 8.3 [166]iinhibition [166]
ORAI1/STIM1CRAC channelrenal (primary cells)↑ migration and proliferation [167]inhibition pKa = 8.3 [168]inhibition pKa = 7.5 [166]
breast (primary cells, cell line)↑ migration, invasion [169]
esophageal (cell line)↑ migration and proliferation [170]
glioblastoma (primary cells, cell line)↑ migration, invasion, proliferation and ↓ apoptosis [171,172]
lung (cell line)↑ proliferation [173]
melanoma (cell line)↑ proliferation, migration and invasion; metastasis formation [174]
pancreatic (cell line)apoptosis resistance [175]
gastric (cell line)↑ proliferation, migration, invasion [176]
acut myeloid leukemia (cell line)↑ proliferation and migration [177]
liver (cell line)↑ proliferation [178]
colorectal (primary cells, cell line)↑ proliferation, invasion and metastasis [179]
colorectal (T lymphocyte)↑ tumor cell killing by CD8+ T cells, regulates antitumor immunity [180]
melanoma (T lymphocyte)↑ tumor cell killing by CD8+ T cells, regulates antitumor immunity [180]
leukemia (T lymphocyte and NK cell)↓ tumor cell killing due to verly high Ca2+ signals [181]
ORAI2/STIM1CRAC channelacut myeloid leukemia (cell line)↑ proliferation and migration [177]inhibition pKa = 8.3 [168]
ORAI3/STIM1CRAC channelbreast (cell line)↑ proliferation [182]inhibition pKa = 8.5 [168]
prostate (primary cells, cell line)↑ tumor growth, apoptosis resistance [183]
SCN5ANaV1.5breast (cell line, biopsies)↑ migration, invasion and cancer progression [33,184,185]inhbition pKa = 6.1 [186]
PIEZO1Piezo1gastric (cell line)↑ proliferation, migration, chemotherapy resistance [187]inhibition [188,189]inhibition [189]
synovial (cell line)↑ viability and migration [190]
gliobastoma (cell line)↑ cell proliferation [191]
breast (cell line)↑ invasion, cell migration [192]
prostate (cell line)↑ proliferation, migration [193]
pancreatic (stromal cell)↑ migration and invasiveness [189]
ASIC1ASIC1breast (cell line and murine model)↑ invasion, metastasis and ROS production [4]stimulation (reviewed in [44])
ASIC1, ASIC3ASIC1/3pancreatic (cell lines)↑ epithelial–mesenchymal transition [194,171]
ASIC2ASIC2colorectal (cell line and murine model)↑ cancer proliferation, invasion and metastasis [195]
ASIC1alung (cell line)↑ proliferation and migration [196]
P2X7P2X7melanoma (murine model)↑ tumor growth, reduced apoptosis, increased angiogenesis [197]inhibition pKa = 6.7 [198]
cervical cancer (cell line)↑ cancer cell apoptosis [199]
pancreatic (cell line)↑ tumor progression [200]
neuroblastoma (tumor-associated immune cell)promotion of MDSCs; CCL2 release, macrophage recruitment [201]
↑ indicates increase, whereas ↓ indicates decrease of a given function.

Share and Cite

MDPI and ACS Style

Pethő, Z.; Najder, K.; Carvalho, T.; McMorrow, R.; Todesca, L.M.; Rugi, M.; Bulk, E.; Chan, A.; Löwik, C.W.G.M.; Reshkin, S.J.; et al. pH-Channeling in Cancer: How pH-Dependence of Cation Channels Shapes Cancer Pathophysiology. Cancers 2020, 12, 2484. https://doi.org/10.3390/cancers12092484

AMA Style

Pethő Z, Najder K, Carvalho T, McMorrow R, Todesca LM, Rugi M, Bulk E, Chan A, Löwik CWGM, Reshkin SJ, et al. pH-Channeling in Cancer: How pH-Dependence of Cation Channels Shapes Cancer Pathophysiology. Cancers. 2020; 12(9):2484. https://doi.org/10.3390/cancers12092484

Chicago/Turabian Style

Pethő, Zoltán, Karolina Najder, Tiago Carvalho, Roisin McMorrow, Luca Matteo Todesca, Micol Rugi, Etmar Bulk, Alan Chan, Clemens W. G. M. Löwik, Stephan J. Reshkin, and et al. 2020. "pH-Channeling in Cancer: How pH-Dependence of Cation Channels Shapes Cancer Pathophysiology" Cancers 12, no. 9: 2484. https://doi.org/10.3390/cancers12092484

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