Next Article in Journal
CDK1 Promotes Epithelial–Mesenchymal Transition and Migration of Head and Neck Squamous Carcinoma Cells by Repressing ∆Np63α-Mediated Transcriptional Regulation
Next Article in Special Issue
Achievement of the Selectivity of Cytotoxic Agents against Cancer Cells by Creation of Combined Formulation with Terpenoid Adjuvants as Prospects to Overcome Multidrug Resistance
Previous Article in Journal
Homozygosity for a Novel DOCK7 Variant Due to Segmental Uniparental Isodisomy of Chromosome 1 Associated with Early Infantile Epileptic Encephalopathy (EIEE) and Cortical Visual Impairment
Previous Article in Special Issue
KCNN4 Promotes the Stemness Potentials of Liver Cancer Stem Cells by Enhancing Glucose Metabolism
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ca2+ Signalling and Hypoxia/Acidic Tumour Microenvironment Interplay in Tumour Progression

by
Madelaine Magalì Audero
1,2,
Natalia Prevarskaya
1 and
Alessandra Fiorio Pla
1,2,*
1
U1003—PHYCEL—Laboratoire de Physiologie Cellulaire, Inserm, University of Lille, Villeneuve d’Ascq, 59000 Lille, France
2
Laboratory of Cellular and Molecular Angiogenesis, Department of Life Sciences and Systems Biology, University of Turin, 10123 Turin, Italy
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(13), 7377; https://doi.org/10.3390/ijms23137377
Submission received: 1 June 2022 / Revised: 26 June 2022 / Accepted: 27 June 2022 / Published: 2 July 2022
(This article belongs to the Special Issue Tumor Microenvironment and Its Actors: Are Ion Channels Relevant?)

Abstract

:
Solid tumours are characterised by an altered microenvironment (TME) from the physicochemical point of view, displaying a highly hypoxic and acidic interstitial fluid. Hypoxia results from uncontrolled proliferation, aberrant vascularization and altered cancer cell metabolism. Tumour cellular apparatus adapts to hypoxia by altering its metabolism and behaviour, increasing its migratory and metastatic abilities by the acquisition of a mesenchymal phenotype and selection of aggressive tumour cell clones. Extracellular acidosis is considered a cancer hallmark, acting as a driver of cancer aggressiveness by promoting tumour metastasis and chemoresistance via the selection of more aggressive cell phenotypes, although the underlying mechanism is still not clear. In this context, Ca2+ channels represent good target candidates due to their ability to integrate signals from the TME. Ca2+ channels are pH and hypoxia sensors and alterations in Ca2+ homeostasis in cancer progression and vascularization have been extensively reported. In the present review, we present an up-to-date and critical view on Ca2+ permeable ion channels, with a major focus on TRPs, SOCs and PIEZO channels, which are modulated by tumour hypoxia and acidosis, as well as the consequent role of the altered Ca2+ signals on cancer progression hallmarks. We believe that a deeper comprehension of the Ca2+ signalling and acidic pH/hypoxia interplay will break new ground for the discovery of alternative and attractive therapeutic targets.

1. Introduction

1.1. Cancer Microenvironment: Focus on Tumour Acidic pHe and Hypoxia

Solid tumours are characterised by a dynamic microenvironment constituted by a variety of different non-cellular components, such as the extracellular matrix (ECM) components, circulating free DNA, and cell components, such as aberrant blood vessels, immune cells, tumour-associated fibroblasts (TAFs), endothelial cells, macrophages, pericytes, among others. In turn, the interaction with chemical and physical cues (hypoxia, tumour acidosis, high tumour interstitial stiffness), originates finally a peculiar chemical and physical environment that supports cancer progression [1].
A common feature of almost all advanced solid cancers is the presence of transient or permanent acidic and hypoxic tumour regions, which are a direct outcome of the cancer cells’ metabolic pathways rearrangement, supporting their uncontrolled proliferation. This leads to a significant increase in cancer cells’ anabolic activity and a reduction in the catabolic one, promoting the synthesis of amino acids, nucleotides, and lipids to back their growth. Indeed, according to Otto Warburg’s work [2], cancer cells are characterised by an enhanced glycolytic breakdown of glucose to pyruvate and consequent NADPH and ATP production with respect to healthy cells, even in presence of oxygen and even considering the lower energy yield of pyruvate fermentation compared to oxidative respiration. Nevertheless, it is important to underline that not all cancer cells are characterized by the Warburg effect, but it has been observed that cancer cells may have an opposite phenotype, with an increased mitochondrial oxidative activity [3]. The importance of mitochondrial activity in cancer cells is also explained by the oxidative phosphorylation increase observed when the Warburg effect is inhibited in cancer cells [3,4,5].
The metabolic rearrangement and fermentation of pyruvate resulting from glycolysis lead to high production of lactic acid, acidifying the intracellular environment. Hydrolysis of ATP also determines the release of protons (H+) in the intracellular space, contributing to its acidification [6,7,8]. Protonation has a severe negative impact on several enzymes and lipids, leading to a potential risk of impinging several cellular processes, including cell metabolism. It is therefore not surprising that cells use different systems to maintain the intracellular pH (pHi) within the physiological range of about 7.2. In cancer cells, a slightly more alkaline intracellular pH (pH 7.4) has been observed, and evidence showed that this slight difference in pHi between cancer cells and healthy cells promotes some of the hallmarks of cancer, such as cell death escape and proliferation [9,10,11,12]. To maintain this pHi value, transformed cells have at their disposal an arsenal of overexpressed transporter proteins and pumps for protons and lactic acid extrusion to the extracellular milieu, resulting in its acidification. Examples of this transport system comprise monocarboxylate transporters (MCTs), major players in the transmembrane lactate trafficking, Na+/H+ antiporters (NHE), vacuolar H+ ATPases, and carbonic anhydrases (CAs), mainly CAIX and CAXII, which role in cancer progression is well documented [13,14,15,16,17]. Acidic interstitial fluids are not only the result of the presence of lactic acid and protons but also of the CO2 derived from the cell respiration process in more oxygenated areas. CO2 can passively diffuse through the plasma membrane (PM), or it can be reversibly hydrated to HCO3 by the transmembrane Carbonic Anhydrase IX (CAIX) exofacial site and released in the tumour microenvironment with protons. Na+/HCO3 cotransporters (NBCs) in the proximity of CAIX can mediate HCO3 influx for sustaining intracellular buffering, titrating cytosolic H+ [18,19]. It has to be noticed that the pHi of cancer cells can drop significantly in presence of a strong acidic pHe, giving rise to a heterogeneous pHi landscape (due to the TME), where cancer cells resident in acidic regions will present quite a low pHi, while cancer cells occupying moderate acidic pHe areas will show a moderate alkaline pHi [8]. It is therefore important to consider the tumour microenvironmental complexity to understand how cancer cells adapt to it, in order to possibly find new therapeutical targets. In addition to metabolic rearrangement, tumour acidosis can be further boosted by tumour-associated hypoxia, which leads to higher glycolytic cell metabolism rates. Hypoxia occurs in the context of tumours vascularized by insufficient vessels and/or vessels characterised by a poor capacity to diffuse oxygen and nutrients and to remove the metabolic waste products due to an altered process of angiogenesis, which leads to the formation of aberrant and dysfunctional vessels. Hypoxia is also the result of an increased oxygen demand from highly proliferating areas of the tumour, leading to intratumour hypoxia heterogeneity, with subregions of the tumour characterised by different oxygen concentrations and consumption. The irregular exposure to oxygen fluctuations is associated with adaptive mechanisms set in motion by cancer cells in order to promote their survival in that hostile environment. Indeed, hypoxia adaptation is linked to increased genomic instability and tumourigenesis [20] and to more aggressive cancer phenotypes in terms of tumour growth, drug and cell death resistance, angiogenesis and enhanced metastasis [21].
Hypoxia adaptation processes are initiated by a series of transcription factors belonging to the hypoxia-inducible factor family, in particular hypoxia-inducible factor 1 (HIF-1), which determines a gene expression reprogramming that affects cancer cell metabolism and processes which sustain its progression. HIF-1 is a heterodimer protein constituted by HIF-1α and HIF-1β and this complex is not present in normoxic conditions. Although the β subunit is constitutively expressed in all cells, the HIF-1α subunit is present only at low levels in all cells’ cytoplasm due to the presence of two specific proline residues at positions 402 and 564 in the oxygen-dependent degradation (ODD) domain in the α subunit. These residues are hydroxylated by prolyl hydroxylase protein (PHD) in presence of physiological oxygen levels, and this modification targets the subunit to degradation via the ubiquitin-proteosome pathway [22]. However, in hypoxic conditions, PHD is inhibited and therefore Pro402 and Pro564 are not hydroxylated, allowing HIF-1α and HIF-1β to dimerize and translocate to the cell nucleus and the consequent activation of target genes transcription [22]. In addition to low O2 concentrations, high intracellular lactate levels and specific growth factors or oncogenes can stabilise HIF-1α as well, leading to the activation of HIF-1α-target genes [23,24]. “Hypoxia-adaptive” responsive genes include glucose transporters, such as GLUT1/3, enzymes involved in anaerobic glycolysis, such as lactate dehydrogenase-A (LDHA), aldolase (ALDA), phosphoglycerate kinase-1 (PGK1), enolase (ENOL) and phosphofructokinase-1 (PFK-1), pyruvate dehydrogenase kinase 1 (PDK1), with consequent suppression of mitochondrial oxidative phosphorylation system (OXPHOS).
The hypoxia response is not limited to glycolytic flux, as it also enhances the expression of VEGF and other pro-angiogenic factors and promotes tumour progression by inducing epithelial–mesenchymal transition (EMT) [25], cell survival in moderate hypoxic conditions via autophagy, cell death via the same mechanism but in presence of severe hypoxic conditions [26], and by promoting cancer cells’ invasion and metastasis in acute hypoxic conditions (from minutes up to 72 h exposure in vitro). In particular hypoxia-mediated cell invasion is achieved by sustaining the mesenchymal phenotype, the expression of different metalloproteases (MMPs), lysyl oxidase (LOX), connective tissue growth factor (CTGF) and CAIX, NHE1 and MCTs [21], which contribute to pH regulation and enhance the acidification of the tumour microenvironment.
Hypoxia also plays a key role in chemoresistance, as reduced oxygen availability can affect not only drug delivery but also chemotherapeutics activity [27]. Moreover, the tumour hypoxic core is occupied by cancer cells with a hypoxia-induced stem cell-like phenotype, characterized by cell cycle arrest in the G1 phase and a quiescent state, representing a major problem for those chemotherapy agents which target rapidly proliferating cells [27,28,29]. Hypoxia also upregulates multidrug resistance genes [30,31,32].
Hypoxia sustains all these effects by activating different signalling pathways via HIF-1α, such as the Wnt/β-catenin and TGF-β/SMAD pathways, or in a HIF-1α-independent manner, switching on MAPKs, AMPKs, PI3K/Akt/mTOR, NF-κB and Notch signalling transduction pathways [21,25], which are dependent also on intracellular Ca2+ ions. All these mechanisms of cell adaptation to hypoxia determine the selection of highly aggressive clones, which pave the way for tumour expansion.
In addition to being affected by the hypoxic tumour microenvironment, acidic pHe has been observed to regulate HIF1α and HIF2α levels under normoxic conditions in glioma cells, promoting cancer stem cell maintenance [33], highlighting the feedback regulation and crosstalk between hypoxia and low pHe. Similarly to hypoxia, the acidic tumour microenvironment supports different hallmarks of cancer, such as drug resistance as previously described [34]. In addition, acidic TME plays an important role in immunoreactive processes and inflammation, by promoting the viability and fitness of pro-tumour M2 macrophages with respect to anti-tumour M1 macrophages [35], by inhibiting T and NK cells activation and inducing immune escape [36,37], by inducing a phenotypic shift in macrophages towards a tumour-promoting phenotype [38] and by increasing the tumour-promoting functions of tumour-associated neutrophils [39]. Moreover, acidic pHe fulfils its pro-tumour function through the enhancement of two other important hallmarks: cancer cell invasion and the ability to metastasise [40]. Studies in breast cancer and colon cancer have indeed demonstrated that invasive cell areas co-localise with acidic pHe regions [41], while studies in melanoma cells have shown that acidic pHe exposure increases their invasive abilities in vitro and the formation of pulmonary metastasis in vivo via a low pHe-promoted secretion of proteolytic enzymes and pro-angiogenic factors [42].
An explanatory example of the major role of acidosis in cancer progression is given by the unique pancreatic ductal adenocarcinoma (PDAC) microenvironment [43]. The pancreatic duct is a net acid-base transporting epithelium, in which ductal cells secrete bicarbonate into the ductal lumen across the apical membrane. This transport is coupled to the extrusion of an equal amount of acid across the basolateral membrane, thereby physiological pancreatic interstitium is substantially acidic and epithelial cells are exposed to different extracellular pH (pHe) values. This process is intermittent in the healthy pancreas and associated with food intake. On the other hand, PDAC has been clearly associated with a hypoxic and acidic microenvironment with a dense desmoplastic stroma [43]. A challenging hypothesis is that, in combination with driver mutations, the alternating, but physiological, pHe landscape in the pancreas and the intrinsic ability of pancreatic epithelial cells to adapt to different pH conditions, may act as a “preconditioning phenomenon” favouring the selection of specific cancer aggressive phenotypes which might promote PDAC arising and/or progression. In other words, once the specific mutations drive the ductal pancreatic cells’ transformation, cells would be already adapted and could even benefit from the adverse pH conditions and the combination of these factors may increase cell fitness to survive and become strongly aggressive in the hostile microenvironment [43].
In this context, it would be important to study the “transportome” alterations that are linked to hypoxia and to pHe or pHi alterations as possible targets for therapies. Indeed, many cancer hallmarks, such as cell proliferation, cell migration, invasion, and apoptosis resistance are driven by altered expression/regulation of ion transport proteins or ion channels, including acid–base transporters and O2- and pH-sensitive channels, in particular, Ca2+- and hypoxia- and pH-sensitive ion channels [43,44].

1.2. Calcium Signalling

Among different ions present in the intra- and extracellular environments, Ca2+ ions stand out for their functional importance as second messengers. Ca2+ ions have been observed to crosstalk with several cell signalling pathways by promoting different spatio-temporal Ca2+ patterns to selectively regulate innumerable physiological cell processes, ranging from cell differentiation, proliferation, migration, and programmed cell death to gene transcription, among others [45,46,47]. Its key role in signal transduction translates into the necessity of tight regulation of intracellular Ca2+ homeostasis, maintaining a low cytosolic free Ca2+ concentration (100 nM) with respect to the extracellular milieu (>1 mM) through the orchestrated work of several proteins that constitute the so-called Ca2+ signalling toolkit, including pumps (Ca2+ ATPases PMCA, SERCA) exchangers (Na+/Ca2+ exchanger NCLX in mitochondria and NCX at the plasma membrane) or uniporters (MCU in mitochondria) and PM and ER Ca2+-permeable channels. Calcium signals are modulated in time and space and differences in amplitude, frequency, duration, and location are transduced by cells to activate a specific response.
Thus, considering this multifaced role of Ca2+ ions, it is no surprise that alterations in Ca2+ homeostasis and Ca2+ channels expression and/or activity in cancer progression and vascularisation have been extensively reported by several works [44,48,49,50,51,52,53,54,55,56,57], making ion channels major players in cancer development (“Oncochannelopathies”) [58].
Ca2+-permeable channels represent pivotal molecular devices acting as microenvironmental sensors in the context of tumourigenesis and other diseases, being modulated by microenvironmental physicochemical cues, such as hypoxia and acidic pHe. Thus, it is possible to speculate that acidic and hypoxic TME and Ca2+ signalling may work in synergy for the acquisition of aggressive cancer cell phenotypes. For this reason, a better understanding of the interplay between these players and the remodelling of Ca2+ signals induced by tumour acidic pHe and hypoxia and translated to the cancer cells through the activity of Ca2+-permeable ion channels and pumps may help to provide further comprehension of the mechanisms of cancer progression and novel putative therapeutic approaches. Indeed, hypoxia is often associated with a rise in intracellular Ca2+ levels in several types of cancer, via the upregulation of different Ca2+-permeable channels and eventually potentiating cancer hallmarks [59,60] (see Section 2). In addition to hypoxia, acidic pH has been reported to regulate Ca2+-permeable ion channels in a direct and indirect way, via specific H+ binding sites or via competition with Ca2+ ions for binding sites as discussed below in Section 2. Data regarding the effects of pHi on these channels is limited in the literature and refers principally to normal cells, whereas more data are available on the regulation of acidic pHe, although different publications demonstrated the high variability of the effects of this microenvironmental parameter among different Ca2+ channels [44,57,61] (see also detailed discussion in Section 2).
In the next section, we will present an updated view of the recent literature on the role of Piezo channels, transient receptor potential Ca2+-permeable ions channels (TRPs) and the so-called store-operated Ca2+ channels (SOCs) on normal and cancer cells, which activity and signalling transduction are directly affected by two features of the tumour microenvironment: hypoxia and acidosis. We will moreover illustrate the Ca2+ signalling pathways that may represent potential targets for cancer therapy.

2. Hypoxia and Acidic pHe-Dependent Regulation of Ca2+-Permeable Ion Channels in Normal and Cancer Cells

Hypoxia and acidic pHe regulate the expression and/or activity of several Ca2+-permeable channels, which are linked to tumour aggressiveness. Throughout the years, several publications have revealed the role of these two major players in the tumour microenvironment, proving a marked sensitivity to oxygen and pHe of most TRPs, SOCs and Piezo channels, which affects their functionality in different tissues. A detailed description of the updated literature on both normal and cancer cells will be presented in this section and Table 1 and Table 2.
As summarized in Table 1 and Table 2, the effects of tumour acidosis and tumour hypoxia vary significantly between the different calcium-permeable channels, and the information in the literature regarding certain channels is sometimes contradictory or limited to normal cells.
Table 1. Ca2+-permeable ion channels regulation by acidic pHe.
Table 1. Ca2+-permeable ion channels regulation by acidic pHe.
Ion ChannelCell TypeMethodologyAcidic pH Value and Treatment TimeEffect of Low pH on Channel’s Activity/ExpressionEffect of Low pH on Ca2+ SignalsCellular FunctionRef.
Piezo1Piezo1-transiently transfected HEK293 cellsPatch clamp
Mn2+ quenching assay
pHe 6.3–6.7, acute treatmentStabilization of inactivated state, both acidic pHi and pHe inhibit channel’s activityDecreased Ca2+ influxNot assessed[62]
Murine pancreatic stellate cells (mPSCs)Mn2+ quenching assay
mPSCs spheroids viability and apoptosis assay
pHe 6.6 and pHi 6.77 (obtained by 30 mM propionate) in acute treatment for Mn2+ quenching assay, while 24 h long treatment for spheroid histologyAcidic pHe do not modify Piezo1 activity, while intracellular acidification inhibits channel’s activityAcidic pHe do not modify Ca2+ influx, while intracellular acidification decreases Ca2+ influxAcidic pHe (6.6) impairs PSCs spheroid’s integrity and viability, inducing cell apoptosis[63]
TRPM2Inducible TRPM2-overexpressing HEK293Patch clampExternal solution with pH 5–8 superfused for 200 s. Internal solution with pH 6 superfused for 100 s; External solution with pH 3.5–6.5 in acute treatment or more prolonged periods (≥2 min)Extracellular acidification inactivates the channel in a voltage-dependent manner and [H+]-dependent manner. Intracellular acidification induces channel closureNot assessed, but recovery from acidic pH-induced inactivation requires external Ca2+ ionsNot assessed[64]
Human neutrophilsPatch clampExternal solution with pH 5 in acute treatmentExternal acidification negatively affects open probability and single-channel conductance, inducing channel closureNot assessedNot assessed[64]
TRPM2-overexpressing HEK293Patch clampExternal solution with pH 3.5–6 in acute treatmentExternal acidification (up to pH 4.5) reversely decreases mean current amplitude in a [H+]-dependent manner, decreasing single-channel conductanceNot assessedNot assessed[65]
TRPM2-overexpressing HEK293Patch clampExternal solution with pH 4.0–6.5. Different time exposition based on protocol (from <10 s to ≥2 min)Acidic pHe inactivates open channels in an irreversible manner. Exposition to pHe 4–5 negatively affects channel activation.Not assessedNot assessed[66]
TRPM2-overexpressing HEK293Patch clampExternal solution with pH 5.5, different exposition times (0, 30, 60, 90, and 120 s)Irreversible inhibition after ≤60 s exposureNot assessedNot assessed[67]
TRPM6Pig isolated ventricular myocytesPatch clampExternal solution with pH 5.5 and pH 6.5, ~5–10 min expositionExternal acidification decreases channel’s current amplitude in a pHe-dependent and voltage-independent manner. The inhibitory effect of acidic pHe is prevented by increasing intracellular pH buffering capacityNot assessedNot assessed[68]
TRPM6-overexpressing HEK293 cellsPatch clampExternal solution with pH 3–6, ~10 s-long expositionExternal acidification increases channel’s current amplitude in a pHe-dependent mannerNot assessedNot assessed[69]
TRPM7RBL-2H3 cellsPatch clampAcidification of intracellular side of membrane with ~200 s long 4–40 mM acetate treatmentPre-incubation in 40 mM acetate solution inhibits TRPM7 current in a reversible mannerNot assessedNot assessed[70]
TRPM7-overexpressing Chinese Hamster Ovary (CHO-K1) cellsPatch clampInternal and external solution with pH 5.6 and variable exposition (~200–500 s)Internal and external acidification abolish channels’ currentNot assessedNot assessed[70]
TRPM7-overexpressing HEK293 cellsPatch clampInternal solution with pH 6.1 and ~10 min expositionInternal acidification decreases TRPM7 currents’ densityNot assessedNot assessed[71]
Mouse hippocampal neuronsPatch clampExternal solution with pH 6.5, 2 min expositionExtracellular acidification slows down channel’s activation in a voltage-independent wayNot assessedNot assessed[72]
TRPM7-overexpressing HEK293T cellsPatch clampExternal solution with pH 4 and pH 6, acute treatmentExternal acidification increases channel’s current amplitude in a pHe-dependent mannerNot assessedNot assessed[69]
TRPM7-overexpressing HEK293T cellsPatch clampExternal solution with pH 3–7, ~50 s-long expositionExternal acidification determines a significant increase in TRPM7 inward current in an [H+] in a concentration-dependent mannerNot assessedNot assessed[73]
Pig isolated ventricular myocytesPatch clampExternal solution with pH 5.5 and pH 6.5, ~5–10 min expositionExternal acidification decreases channel’s current amplitude in a pHe-dependent and voltage-independent manner. The inhibitory effect of acidic pHe is prevented increasing intracellular pH buffering capacityNot assessedNot assessed[68]
Rat basophilic leukemia cells (RBL)Patch clampExternal solution with pH 5.5, pH 6 and pH 6.5, ~1-min-long expositionExternal acidification decreases channel’s current amplitude in a pHe-dependent mannerNot assessedNot assessed[68]
HeLa cellsPatch clamp
Cell death assays (fluometric analysis of caspase 3/7 activation, electronic sizing of cell volume, and triple staining with Hoechst/acridine orange and propidium iodide assay.
External solution with pH 4 and pH 6, acute treatment for patch clamp experiments, and 1 h-long treatment with acidic pHe (4 and 6) for cell death assaysExternal acidification increases channel’s current amplitude in a pHe-dependent mannerNot assessedAcidosis promotes HeLa necrotic cell death[74]
Human atrial cardiomyocytesPatch clampExternal solution with pH 4–6, acute treatmentExternal acidification increases channel’s current amplitude in presence of divalent cations in the extracellular milieuNot assessedNot assessed[75]
TRPV1TRPV1-expressing HEK293 cellsPatch clampAcidic solution with pH 5.5 applied intracellularly for ~50 sAcid treatment does not activate the channel in inside-out patches but potentiates 2-APB-evoked currents from the cytoplasmic sideNot assessedNot assessed[76]
hTRPV1-transfected HEK293t cellsCalcium imagingExternal solution with pH 4.3 and pH 6.1, ~4 min-long expositionAcidic pHe activates TRPV1 channelpHe 6.1 determines larger Ca2+ transients with respect to pHe 4.3 in physiological extracellular Ca2+ concentration, while, in presence of low extracellular Ca2+ concentration, cells exposed to pHe 6.1 show reduced Ca2+ entry respect to pHe 4.3 expositionNot assessed[77]
Defolliculated Xenopus laevis oocytes,
TRPV1-expressing HEK293 cells
Patch clampExtracellular solution with pH 6.4, cells pre-treated with acid bath solution for 2 minAcidic pHe potentiates heat-evoked TRPV1 current in oocytes; potentiation of capsaicin and heat-evoked TRPV1 currents in HEK293 cellsNot assessedNot assessed[78]
Primary human adult dermal lymphatic endothelial cell (HDLECs)Cell viability assay
Cell invasion assay
in vitro tube formation assay
Transwell cell migration assay
24 h long exposition to pHe 6.4, and 6 h long exposition for in vitro tube formation assayAcidic pHe activates TRPV1 channelNot assessedAcidic pHe affects HDLECs morphology, increasing their migration and invasive abilities, proliferation and promoting lymphangiogenesis via acidosis-induced TRPV1 activation[79]
TRPV2TRPV2-expressing HEK293 cellsPatch clampAcute administration of extracellular solution with pHe 5.5 and 6Extracellular acidosis potentiates the response of TRPV2 to 2-APB (and analogues) from the cytosolic side, while intracellular acidification and low pHe alone are not able to elicit any detectable currentNot assessedNot assessed[80]
TRPV3TRPV3-expressing HEK293 cellsPatch clamp, calcium imagingAcute administration of extracellular solution with pHe 5.5 and 6Extracellular acidosis potentiates the response of TRPV3 to 2-APB (and analogues) from the cytosolic side. Intracellular acidification activates the channel, eliciting small but detectable currentsExtracellular acidosis increases Ca2+ entry following 2-APB stimulationNot assessed[80]
TRPV3-expressing HEK293 cellsPatch clamp
Cell death assay (PI staining assay)
Intracellular administration of acidic solution with pHe 5.5 and glycolic acid. Extracellular solution with pH 5.5. Intracellular solution with pH 5.5–7.Glycolic acid-induced intracellular proton release in presence of acidic solution activates the channel in a reversible way. Extracellular acidification does not activate TRPV3, while intracellular acidification alone activates the channel in a pH-dependent mannerNot assessedGlycolic acid-induced acidification induces cell toxicity and cell death[81]
Human keratinocytes cells (HaCaT)Patch clamp, cell death assay (PI staining assay)Intracellular administration of acidic solution with pHe 5.5 and glycolic acidGlycolic acid-induced intracellular proton release in presence of acidic solution potentiates the channel’s response to 2-APB in a reversible mannerNot assessedGlycolic acid-induced acidification induces cell toxicity and cell death[81]
TRPV4Chinese hamster ovary cellsPatch clampExternal solution with pHe 4, 5.5 and 6, acute treatmentExtracellular acidosis activates the channel in a pHe-dependent mannerNot assessedNot assessed[82]
mTRPV4-overexpressing primary cultured mouse esophageal epithelial cellsCa2+ imagingExternal solution with pHe 5, acute treatmentNot assessedExtracellular acidic pH decreases Ca2+ entry, lowering cytosolic Ca2+ concentrationNot assessed[83]
TRPV6Jurkat cellsPatch clampExternal solution with pH 6, acute treatmentExtracellular acidosis suppresses TRPV6-mediated currentsExtracellular acidic pH reduces Ca2+ entry, lowering cytosolic Ca2+ concentrationNot assessed[84]
TRPA1HEK-293t cells expressing hTRPA1, mTRPA1, or rTRPA1Patch clamp
Calcium imaging
Acidic solutions with pH 7.0, 6.4, 6.0, and 5.4, 30 s-long treatment in calcium imaging experimentsExtracellular acidosis activates inward currents via hTRPA1 and potentiates acrolein-evoked currents of hTRPA1 in a pHe-dependent and reversible manner, while failing to activate mouse and rodent TRPA1.Extracellular acidosis increases Ca2+ entry in hTRPA1, no effect on mTRPA1 and rTRPA1.Not assessed[85]
DRG neurons derived from TRPV1/TRPA1−/− mice and overexpression hTRPA1Calcium imagingAcidic solutions with pH 5, 60 s-long treatmentNot assessedAcidic pHe induces Ca2+ entryNot assessed[85]
Neuroblastoma ND7/23 cells expressing hTRPA1Patch clampAcidic solution with pH 5, acute treatmentAcidic pHe activates hTRPA1Not assessedNot assessed[85]
TRPC5TRPC5-transiently transfected HEK293 cellsPatch ClampExternal acidic solution with pH 4.2, 5.5, 6.5, 7, ~100 s-long treatmentG protein-activated and spontaneous currents are potentiated by extracellular acidic pH by increasing the channel open probability, with a maximum effect at ~pH 6.5, while more acidic values inhibit the channelNot assessedNot assessed[86]
TRPC4TRPC4-transiently transfected HEK293 cellsPatch ClampExternal acidic solution with pH 4.2, 5.5, 6.5, 7, ~100 s-long treatmentG protein-activated currents are potentiated by extracellular acidic pH, with a maximum effect at ~pH 6.5 and complete inhibition at pHe 5.5Not assessedNot assessed[86]
mTRPC4-stably transfected HEK293 cellsPatch ClampExternal acidic solution with pH 6.8Low pHi (6.75–6.25) accelerates Gi/o-mediated TRPC4 activation, and this requires elevations in intracellular calcium concentration. Intracellular protons inhibit Englerin A-mediated TRPC4 activationNot assessedNot assessed[87]
TRPC6TRPC6-transiently transfected HEK293 cellsPatch ClampExternal acidic solution with pH 4.2, 5.5, 6.5, 7, ~100 s-long treatmentAcidic pHe inhibits channel’s inward and outward currents starting from pHe 6.5 and the inhibition is potentiated by more acidic pHe values.Not assessedNot assessed[86]
ORAI1/STIM1Human macrophagesPatch clampExternal acidic solution with pH 6 and 8, ~200 s-long treatmentExtracellular acidosis inhibits ORAI1 channel in a pHe-dependent and reversible mannerNot assessedNot assessed[88]
H4IIE rat liver cells overexpressing ORAI1 and STIM1Patch clampExternal acidic solutions with pH 5.1 and 5.9ORAI1 and STIM1-mediated ICRAC are inhibited by acidic pHe, with maximal effect at pHe 5.5Not assessedNot assessed[89]
RBL2H3 mast cell line, Jurkat T lymphocytes and heterologous ORAI1-2–3/STIM expressing HEK293 cellsPatch clampExternal and intracellular acidic solutions with pH 6 and 6.6External and internal acidification inhibits IP3-induced ICRAC in RBL2H3 mast cell line, Jurkat T lymphocytes, and in heterologous ORAI/STIM-mediated ICRAC in HEK293 cells in a reversible mannerNot assessedNot assessed[90]
ORAI1/STIM1-transiently transfected HEK293 cellsPatch ClampExternal acidic solution with pH 5.5Acidic pHe inhibits ORAI1-2–3/STIM1 current amplitude in a reversible and pH-dependent manner, with a maximal effect at pHe 4.5Not assessedNot assessed[91]
ORAI1/STIM1-transiently transfected HEK293 cellsPatch ClampIntracellular acidic solution with pH 6.3Intracellular acidosis inhibits ORAI1/STIM1 current, regulating the amplitude of the current and the Ca2+-dependent gating of the CRAC channelsNot assessedNot assessed[92]
SH-SY5Y human neuroblastoma cellsCa2+ signals quantification by Mn2+ quench techniqueExternal acidic solution with pH 6.8 and 7 and 7.2. Different treatment time, ranging from ~3–4 min to ~8 min for carbachol-mediated Ca2+ entry and ~7 min for thapsigargin-mediated Ca2+ entryNot assessedTumour acidic pHe inhibits carbachol- and thapsigargin-mediated Ca2+ entry in a reversible manner, while intracellular acidification or alkalinization leads to no effects in carbachol-mediated Ca2+ entryNot assessed[93]
Table 2. Ca2+-permeable ion channels regulation by hypoxia.
Table 2. Ca2+-permeable ion channels regulation by hypoxia.
Ion ChannelCell TypeMethodologyHypoxia Technique and Treatment TimeEffect of Hypoxia on Channel’s Activity/ExpressionEffect of Hypoxia on Ca2+ SignalsCellular FunctionRef.
Piezo1Mouse and human sickle red blood cells (RBCs)Cell-attached and nystatin-permeabilized patch clamp
Calcium imaging
Deoxygenation obtained by exposure with a superfusate gassed 30 min prior to the experiment with 100% N2Deoxygenation activates a Ca2+- and cation-permeable conductance in a reversible manner, and this current is sensitive to inhibition by GsMTx-4; 1 mMIncreased Ca2+ influxNot assessed[94]
Pulmonary arterial endothelial cells (PASMCs) of patients with pulmonary arterial hypertension (PAH)Calcium imaging
EdU and cell counting proliferation assay
Western Blot
/Piezo1 expression and activity are increased in idiopathic pulmonary arterial hypertension and pulmonary arterial smooth muscle cellsIncreased Ca2+ influx and increased intracellular Ca2+ releaseIncreased PAH-PASMCs’ proliferation[95]
Pulmonary artery smooth muscle cells of mice and rats’ models with experimental chronic hypoxia-induced pulmonary hypertension (PH)
Human pulmonary artery endothelial cells (hPAECs)
Western Blot
Calcium imaging
Hypoxia induced by incubation in 3% O2 for 4 h–12 h or in 10% O2 for a total of 6 weeksPiezo1 is significantly upregulated in the lung tissue of PH rats and in chronic hypoxia-induced PH models. Piezo1 protein is transiently upregulated also in hPAECs after 6 h exposition to hypoxic conditions.
Hypo-osmotic conditions upregulate Piezo1 protein levels in hPAECs
Hypo-osmotic upregulation of Piezo1 promotes Ca2+ influx, promoting Akt and Erk signalling pathways activation, with downstream upregulation of Notch ligandGsMTx4-mediated Piezo1 blockade partially reduces the chronic hypoxia-induced PH in mice with chronic hypoxia-induced pulmonary hypertension[96]
TRPM2TRPM2 WT and knockout (KO) neonatal hypoxic-ischemic (HI) brain injury mouse modelWestern BlotHypoxia damage was induced in ischemic mice models by incubating the pups in a hypoxic chamber for 2 hTRPM2 is acutely overexpressed 24 h after hypoxia-ischemic injury in brain tissue samples from mouse pupsNot assessedBrain damage and inflammation are reduced in TRPM2 KO mice 7 days following hypoxic-ischemic brain injury.
TRPM2 inhibits cell survival pathways after HI injury
[97]
Primary cultures of rat cortical neurons subjected to oxidative stressCalcium imaging
Trypan Blue exclusion assay
Oxidative stress induced by 1 mM or 50 µM H2O2 treatmentNot assessedH2O2 induces TRPM2-mediated intracellular calcium riseSiTRPM2 prevents H2O2-mediated neuronal cell death[98]
TRPM2-overexpressing HEK293 cellsWhole-cell Patch ClampHypoxia induced by cell incubation with gas mixture containing 5% O2 for 30 and 60 minTRPM2 activation is induced by 30- and 60-min exposure to hypoxic conditionsNot assessedHypoxia treatment enhances cell death, probably via TRPM2-mediated Ca2+ influx[99]
ARPE-19 retinal pigment epithelial cellsPatch Clamp
Calcium imaging
Propidium iodide cell death assay
Hypoxia induced by CoCl2 (200 μM) for 24 hHypoxia induces activation of TRPM2 currents and upregulates TRPM2 protein levelsHypoxia induces TRPM2-mediated intracellular calcium riseHypoxia causes mitochondrial oxidative cell cytotoxicity and cell death via TRPM2-mediated Ca2+ signals[100]
Primary IGR39 melanoma cells
TRPM2-overexpressing HEK293 cells
Patch Clamp
Calcium imaging
Treatment with chloramine-T (Chl-T) oxidant agentAmount of 0.5 mM Chl-T activates TRPM2 in IGR39 and in TRPM2-expressing HEK293 cellsChl-T treatments induce a significant increase in cytosolic Ca2+ levelsChl-T-induced TRPM2 activation and increased Ca2+ influx activate BK and KCa3.1 potassium channels[101]
PC3 prostate cancer cellsCalcium imaging
MTT and TUNEL assay
Treatment with 0.5 to 4 mM H2O2 for 6 hH2O2 induces TRPM2 activationH2O2 treatment leads to TRPM2-mediated intracellular Ca2+ increase in a concentration-dependent mannerH2O2 induces TRPM2-Ca2+-CaMKII cascade that promotes ROS production, mitochondrial fragmentation, and inhibition of autophagy, inducing cell death[102]
TRPM2-L and TRPM2-S-expressing SH-SY5Y neuroblastoma cellsCalcium imagingTreatment with 250 μM H2O2 for 20 minNot assessedH2O2 treatment leads to TRPM2-L-mediated intracellular Ca2+ increase and a decrease in TRPM2-STRPM2-L-expressing cells show higher HIF-1/2α levels with respect to TRPM2 short isoform and promote tumour growth in vivo[103]
Human breast cancer cellsCalcium imaging
qPCR
Co-culture with neutrophils or H2O2 treatmentNeutrophil-derived H2O2 induces decrease in TRPM2 expression in H2O2-selected tumour cellsNot assessedTRPM2 activation by neutrophil-derived H2O2 and following Ca2+ entry promotes cancer cells’ death[104]
TRPM6Hepatic ischemia-reperfusion rat modelqPCRIschemia was obtained by 60 min clamping the left hepatic artery and the portal veinTRPM6 expression is increased in liver tissue from ischemia-reperfusion rat modelNot assessedNot assessed[105]
TRPM7TRPM7-overexpressing HEK293T cells
Cortical neurons
Ca2+ imaging
Patch clamp
PI cell death assay
Hypoxia induced by anaerobic chamber containing ˂0.2% O2 atmosphere for 1, 1.5 and 2 h.Hypoxia induces TRPM7 channel activationHypoxia increases Ca2+ entryHypoxia-activated TRPM7 mediated-Ca2+ entry determines cell death in cortical neurons[106]
Hepatic ischemia-reperfusion rat modelqPCRIschemia was obtained by 60 min clamping the left hepatic artery and the portal veinTRPM7 expression is increased in liver tissue from ischemia-reperfusion rat modelNot assessedNot assessed[105]
TRPV1HEK293T cells overexpressing rat TRPV1Patch Clamp
Calcium imaging
Hypoxic solution obtained by bubbling with 100% N2 gas for at least 20 min before the perfusion (PO2, 3%)Acute hypoxia weakly increases TRPV1 activity, but negatively affects capsaicin induced TRPV1 currentsHypoxia leads to a slight increase in cytosolic Ca2+ levelsNot assessed[107]
Rat DRG neurons
hTRPV1/rTRPV1-expressing HEK293 cells
Whole-cell patch-clampOvernight (18–20 h) exposition to hypoxia (4% O2)Overnight exposure to hypoxic/high glucose conditions increases TRPV1 mean peak current densities in both cell lines, without affecting its expressionNot assessedNot assessed[108]
Rat pulmonary artery smooth muscle cells (PASMCs)Calcium imaging
qPCR
Western Blot
Wound Healing assay
BrdU proliferation assay
24–48 h long exposition to hypoxia (1% and 10% O2)Hypoxic conditions do not affect TRPV1 expression, but they increase TRPV1 activityNo assessedHypoxia-mediated TRPV1 activation enhances PASMCs migratory abilities and proliferation[109]
Human pulmonary artery smooth muscle cells (PASMCs)Calcium imaging
qPCR
Western Blot
Cell count proliferation assay
72 h long exposition to hypoxia (3% O2)Chronic hypoxia upregulates both TRPV1 gene and protein levelsChronic hypoxia increases cytosolic Ca2+ levelsThe proliferation of PASMCs is increased under hypoxia[110]
TRPV2HepG2 and Huh-7 human hepatoma cell linesRT-PCR
Western Blot
Flow cytometry
50, 100, 200, and 400 Μm H2O2 treatment for 24 hH2O2 upregulates the expression of TRPV2 at mRNA and protein levelsNot assessedOverexpression of TRPV2 promotes H2O2-induced cell death[111]
TRPV3Rat myocardial cellsMTT and Edu staining assay
Western Blot
Caspase-3 and LDH activity assay
12 h long exposition to hypoxia (1% O2)TRPV3 is overexpressed in myocardial cells induced by ischemia/hypoxiaNot assessedTRPV3 silencing protects cardiomyocytes from hypoxia-induced cell death and decreases the secretion of proinflammatory cytokines[112]
Primary rat pulmonary artery smooth muscle cells (PASMCs)Western Blot
Flow cytometry
MTT assay
24 h long exposition to hypoxia (3% O2)TRPV3 protein expression is enhanced in PASMCs from hypoxic ratsNot assessedTRPV3 mediates hypoxia-induced PASMCs’ proliferation via PI3K/AKT signalling[113]
TRPV3-overexpressing HEK293Patch Clamp12 h long exposition to hypoxia (1% O2)Pre-incubation in hypoxic conditions potentiates TRPV3 currents in response to 2-APB treatmentNot assessedNot assessed[114]
TRPV4Rat cardiomyocytesWestern Blot
qPCR
Calcium imaging
6 h long exposition to hypoxia (95% N2) in a controlled hypoxic chamberTRPV4 gene and protein expression levels are increased after 6 h exposure to hypoxiaHypoxia increases TRPV4-mediated Ca2+ influx responses to 300 nM GSKHypoxia-mediated activation of TRPV4 induces cytosolic Ca2+ overload in cardiomyocytes, leading to ROS production and oxidative injury in vitro and in vivo[115]
Adult rat hippocampal astrocytesPatch Clamp
qPCR
Western Blot
Calcium imaging
Hypoxia/ischemia (H/I) is induced by occlusion of the common carotids in combination with hypoxic conditions (from 1 h up to 7 days, 6% O2)TRPV4 mRNA and protein expression are significantly increased 1 h after H/I. H/I also activates TRPV4 channelH/I enhances the response of 4aPDD, inducing TRPV4-mediated Ca2+ oscillationsNot assessed[116]
TRPA1Several breast and lung cancer cell linesCalcium imaging
Cell viability and apoptosis assay via PI and Annexin IV staining
Treatment with 10 µM H2O2 for 15 min for calcium measurements, 1, 20, and 100 µM for 72–96 h-long exposition for cell viability and cell death assaysH2O2 treatment activates TRPA1 channelH2O2 treatment increases TRPA1-mediated calcium entryTRPA1-mediated calcium entry promotes cell survival by upregulating anti-apoptotic pathways and promoting oxidative stress resistance[117]
OligodendrocytesCalcium imagingIschemia inducing solutionNot assessedIschemia-induced intracellular acidosis promotes Ca2+ entry via TRPA1Ischemia-induced intracellular acidosis and consequent Ca2+ entry via TRPA1 mediate myelin damage[118]
TRPC1U-87 MG glioma cellsqPCR, western blotHypoxia induced by exposition to 1% O2Not assessedNot assessedTRPC1 participates in hypoxia-induced VEGF gene and protein expression[119]
MDA-MB-468 breast cancer cellsqPCR, calcium imagingHypoxia induced by exposition to 1% O2 for 24 hHypoxia upregulates TRPC1 via HIF1αsiTRPC1 reduces non-stimulated Ca2+ entry and increases Store-Operated Ca2+ entry in hypoxic conditionsTRPC1 overexpression promotes Snail EMT marker upregulation and decrease in claudin-4 epithelial marker in hypoxic conditions. TRPC1 regulates HIF-1α protein levels via Akt-dependent pathway and promotes hypoxia-induced STAT3 and EGFR phosphorylation. TRPC1 also regulates hypoxia-induced LC3BII levels via effects on EGFR.[120]
TRPC5MCF-7/WT and adriamycin-treated (MCF-7/ADM) human breast cancer cellsWestern Blot, immunofluorescence,Not assessedNot assessedNot assessedTRPC5 promotes HIF-1alpha translocation to the nucleus and HIF-1alpha-mediated VEGF expression, boosting tumour angiogenesis[121]
SW620 colon cancer cellsWestern blot, transwell invasion, and migration assay, MTT proliferation assayNot assessedNot assessedNot assessedTRPC5 activates HIF-1alpha-Twist signalling to induce EMT, supporting colon cancer cells’ migration, invasion, and proliferation[122]
TRPC6Murine pancreatic stellate cells (mPSCs)Time-lapse single-cell random migration assay
Bead-based cytokine assay
qPCR
Western Blot
Ca2+ signals quantification by Mn2+ quench technique
24 h incubation in hypoxic conditions (1% O2, 5%CO2, and 94% N2) or chemically induced hypoxia by pretreatment with 0.5 mmol/l DMOGHypoxic conditions enhance TRPC6 expression and activate the channelHypoxia stimulates Ca2+ influx mediated by TRPC6 channelsHypoxia-induced TRPC6 activation enhances mPSCs migration via secretion of pro-migratory factors[123]
lx-2 human hepatic stellate cells (HSCs)Calcium imaging
qPCR
Western Blot
Hypoxia induced by 100 μmol/L CoCl2 treatmentHypoxic conditions enhance TRPC6 expression and activate the channelHypoxia stimulates Ca2+ influx mediated by TRPC6 channelsHypoxia-induced TRPC6 activation and consequent calcium entry promote the synthesis of ECM proteins, which facilitate the fibrotic activation of HSCs[124]
Huh7 and HepG2 hepatocellular carcinoma cells (HCCs)Confocal Calcium imaging
Western Blot
Hypoxia induced by cell incubation in a low oxygen atmosphere with 1% O2, 5%CO2, and 94% N2 for 6 hHypoxic conditions activate the channelHypoxia promotes calcium influxHypoxia-induced TRPC6-mediated calcium entry promotes HCCs drug resistance via STAT3 pathway[125]
U373MG and HMEC-1 glioblastoma cell linesqPCR
Western Blot
Calcium imaging
Proliferation assay
Matrigel invasion assay
Endothelial cell tube formation assay
Hypoxia induced by 100 μmol/L CoCl2 treatmentHypoxia enhances TRPC6 expression via Notch pathwayHypoxia stimulates Ca2+ influx mediated by TRPC6 channelsHypoxia-induced TRPC6-mediated calcium entry promotes HCCs proliferation, colony formation, and invasion via NFAT pathway[126]
ORAI1/STIM1Primary Aortic Smooth Muscle Cells and HEK293 cells transfected with ORAI1 and STIM1Patch Clamp
Calcium imaging
Hypoxia was induced with 3 methods: (1) sodium dithionite (Na2S2O4) treatment to 1 mM final concentration, pH adjustment to pH 7.4, and bubbling with 100% N2. (2) cell culture media with 30 min-long bubbling with 100% N2. (3) cell culture media with 30 min-long bubbling with 3% O2Intracellular acidification induced by hypoxia in HEK293 cells leads to inhibition of SOCE by disrupting the electrostatic ORAI1/STIM1 binding and closing ORAI1 channel.Hypoxia-induced intracellular acidification reduces SOCE in Primary Aortic Smooth Muscle Cells and HEK293 cells transfected with ORAI1 and STIM1Not assessed[92]
A549 non-small cell lung cancer cellsWestern Blot
qPCR
BrdU cell proliferation assay
Calcium imaging
Scrape-wound migration assay
Matrigel transwell invasion assay
Hypoxia induced by Nicotine treatment for 48 hNicotine treatment-induced hypoxia determines ORAI1 overexpression at gene and protein levelsNicotine treatment-induced hypoxia increases intracellular basal calcium levels and SOCENicotine treatment-induced hypoxia increases A549 cells’ proliferation and migration[127]
MDA-MB 231 and BT549 breast cancer cell lines and Human Microvascular Endothelial Cell line-1 (HMEC-1)Western Blot
qPCR
Calcium imaging
Migration assay (Wound healing and transwell migration assay)
Matrigel transwell invasion assay
Tube formation assay in vitro
Hypoxia induced by cell incubation in low oxygen atmosphereHypoxia promotes ORAI1 gene and protein upregulation via activation of Notch1 signallingHypoxia increases thapsigargin-induced SOCE, with consequent rise in cytosolic calcium entryHypoxia-induced ORAI1 overexpression and consequent increase in SOCE promote NFAT4 activation and enhance neuroblastoma cells’ migration, invasion, and angiogenesis[128]
HCT-116 and SW480 human colon cancer cells and Human Microvascular Endothelial Cell line-1 (HMEC-1)Western Blot
qPCR
Calcium imaging
Transwell migration assay
Matrigel transwell invasion assay
Tube formation assay in vitro
Cell attachment and detachment assays
Hypoxia induced by 100 μmol/L CoCl2 treatmentHypoxia promotes ORAI1 gene and protein upregulation via activation of Notch1 signallingHypoxia increases thapsigargin-induced SOCEHypoxia-induced ORAI1 overexpression and consequent increase in SOCE promote NFATc3 activation and enhance neuroblastoma cells’ migration, invasion, and angiogenesis[129]

2.1. Piezo Channels

Piezo proteins are mechanically activated non-selective cations channels identified over a decade ago by Patapoutian’s group [130], a discovery that paved the way for various works that further clarified the role of these channels not only in the transduction of mechanical signals, but also in other physiological processes, and culminating in the 2021 Nobel Prize in Physiology or Medicine for David Julius and Ardem Patapoutian “for their discoveries of receptors for temperature and touch”. Piezo channels play a key role in mechanotransduction, directly responding and integrating mechanical stimuli and forces from the cell environment into biological signals, leading to the activation of Ca2+-dependent processes or cell depolarization in different organs [131]. Piezo channels are also involved in other physiological processes, exhaustively reviewed in [131].
The role of Piezo channels in cancer has been deepened in recent years, with several studies evidencing their importance in different types of cancer that originate from tissues subjected to mechanical stress [132]. Piezo1 and/or Piezo2 are overexpressed in several cancers of epithelial origin, where these channels act as oncogenes, enhancing carcinogenesis through different Ca2+-dependent signalling pathways [133,134,135,136,137].
In addition to being sensitive to mechanical stimulation, Piezo1 is also regulated by protons. The work of Bae C. and colleagues in 2015 demonstrated that conditions of acidosis (pHe 6.3) inhibit Piezo1 by stabilizing its inactivated state [62]. A drop in extracellular pH and inactivation of Piezo1 might represent a protective mechanism in specific cell types, especially considering that low extracellular pH can promote intracellular acidification and, therefore, the activation of specific signalling pathways that result in cell death [138] (Table 1). This notion is supported by the study of Kuntze A. et al. of 2020, which demonstrated that extracellular acidosis-induced low intracellular pH (pHi 6.7) of pancreatic stellate cells (PSCs) inhibited the activity of Piezo1, reducing the Ca2+ influx in PSCs (Figure 1 and Table 1). In these conditions, Piezo1 activation with Yoda1 led to a loss of PSCs spheroid integrity and increased fragmentation, resulting virtually from Ca2+ overload and its induction of cell death. Therefore, extracellular acidosis-mediated intracellular pH drop and inactivation of Piezo1 might represent a protective mechanism for PSCs, in which Ca2+ fluxes are decreased and apoptosis is avoided [63].
Information about Piezo channels’ regulation by low oxygen pressures is absent in cancer and the literature data are limited to a few works in red blood cells and pulmonary endothelial and smooth muscle cells, where the pathogenic role of Piezo1 was explored.
Stretch-activated cation channel expression, like Piezo1, has been reported in red blood cells (RBCs), where Piezo1 is expressed and localises at the cell membrane, where it regulates RBCs volume. This role assumes major relevance in pathological conditions, such as Sickle cell disease (SCD), where RBCs showed decreased deformability and higher intracellular Ca2+ levels with respect to normal red blood cells [139]. In sickle RBCs, low oxygen concentration led to the activation of transmembrane mechanosensitive ion channels. This activation originated a non-selective ion current named Psickle, which activated the calcium-activated potassium channel KCa3.1, which initiates the dehydration cascade in sickle RBCs, responsible for potentiating haemoglobin aggregation [140]. Therefore, Ca2+ entry in sickle RBCs is necessary for their sickling, which is mediated by the opening of plasma membrane Ca2+-permeable and mechanosensitive ion channels. In this context, Piezo1 can be suggested as a mediator of the non-selective ion current, as deoxygenation-induced Psickle is abolished by GsMTx4, the tarantula spider toxin which is an inhibitor of mechanosensitive ion channels, including Piezo1 [94].
Recent works have shown that Piezo1 is upregulated in pulmonary arterial endothelial cells of patients with pulmonary arterial hypertension (PAH) and in pulmonary artery smooth muscle cells of mice and rats models with experimental chronic hypoxia-induced pulmonary hypertension (PH) [95,96]. Human pulmonary arterial endothelial cells showed maximal Piezo1 expression at 6 h hypoxic exposure, with a consequent increase in calpain-1 and calpain-2, both involved in PAH development [96]. Moreover, hypo-osmotic conditions upregulated Piezo1 protein levels in the same cells and promoted the activation of Akt and Erk signalling pathways via Piezo1-induced Ca2+ entry, as demonstrated by siPiezo1 treatment, with downstream upregulation of Notch ligands [96]. The potential therapeutic effect of Piezo1 blockade in the mouse model with chronic hypoxia-induced PH was assessed by GsMTx4 treatment, which partially reduced the chronic hypoxia-induced PH [96]. These data support the role of Piezo1 in the remodelling of the pulmonary vasculature in PH.

2.2. Transient Receptor Potential Channels

Transient receptor potential (TRP) ion channels are a family of 28 different proteins in humans, mostly permeable to Ca2+ ions, characterized by a polymodal activation, and whose altered expression and/or functionality have been linked to several cancer types [54]. Several TRP channels are sensitive to changes in intra- and extracellular pH and to hypoxia (see Table 1 and Table 2), altering Ca2+ downstream signalling pathways as described in the following paragraphs.

2.2.1. TRP Melastatin Subfamily

TRPM2 is a non-selective cation channel localized at the plasma membrane and/or in lysosome compartments and permeable to Ca2+, Mg2+, and monovalent cations. TRPM2 is activated by ADP-ribose (ADPR) and by intracellular Ca2+ increase associated with oxidative stress and reactive oxygen species (ROS) production [141,142,143,144].
TRPM2’s activity is regulated by both intra- and extracellular pH. The work of Starkus and colleagues demonstrated that extracellular acidic pH (IC50 pH = 6.5) inhibited both inward and outward TRPM2 currents in TRPM2-overexpressing HEK293 cells in a voltage-dependent manner, by affecting the single-channel conductance, most probably due to the interaction of protons with outer pore and competing for binding sites with extracellular Ca2+ ions, as these ions attenuated the inhibitory effect of pHe on TRPM2 [64]. These results were confirmed the same year by Du J. et al. and on the same cell line, although assuming a non-proton permeation through the channel [65], and by Yang W. et al. in 2010 [66], showing that low pHe-mediated inhibition of TRPM2 might be induced by conformational changes following protons binding. Interestingly, extracellular acidic pH effects are species-dependent, with mTRPM2 channels showing less sensitivity to acidic pHe compared to hTRPM2 [67] (see also Figure 1 and Table 1). Du J et al., as well as Starkus et al., also studied the intracellular acidic pH role showing a reversible inhibitory effect on the channel, inducing its closure without affecting single-channel conductance [64], probably by a mechanism of proton competition with the Ca2+ and ADPr binding site [65].
Several reports have highlighted the importance of the role of hypoxia in TRPM2’s expression and activity, in particular in the context of hypoxia-induced brain damage, but with little evidence in tumour cells (Figure 1 and Table 2).
Concerning TRPM2’s role in cerebral hypoxia, several comprehensive reviews have elucidated the current knowledge on this topic [145,146]. Huang and colleagues explored the TRPM2 pathological role in neonatal hypoxic-ischemic brain injury mouse model, demonstrating that genetic deletion of this channel had a long-term neuroprotective effect, reducing brain damage and inflammatory responses 7 days following the hypoxia-ischemic injury with respect to wild-type littermates. At the molecular level, TRPM2 was acutely overexpressed 24 h following hypoxia-ischemic injury in brain tissue samples from CD1 mouse pups and it inhibited pro-survival pathways in control mice by decreasing pAkt and pGSK-3β levels, exerting its protective role in TRPM2-null mice via Akt/Gsk-3β pathway [97]. In vitro TRPM2 silencing had a neuroprotective effect also in primary cultures of rat cortical neurons subjected to oxidative stress following brief H2O2 exposure, reducing the intracellular calcium levels and preventing H2O2-mediated neuronal cell death [98]. In addition to neuronal cells, other works have explored hypoxia-induced activation of TRPM2 in other cell types, such as HEK-293 cells, where TRPM2 was activated after 30 and 60 min exposure to hypoxic conditions, leading potentially to an increased Ca2+ influx and oxidative stress, resulting in cell death [99], and retinal pigment epithelial cells, where 24 h-long exposure to hypoxia promoted ROS production and cell death via TRPM2-mediated Ca2+ influx [100].
In the context of tumourigenesis, TRPM2 is upregulated in several cancers where it mediates Ca2+-dependent pathways promoting cell survival [147,148,149,150,151,152]. TRPM2 exerts this effect in particular by protecting cancer cells from oxidative stress by increasing their antioxidant defence. Indeed, although high levels of ROS are detected in almost all types of cancer, a precise balancing of their intracellular presence is required to avoid the toxic effect of these reactive species, achieved through the expression of antioxidant proteins. TRM2 channels act as ROS sensors, being activated by oxidative stress and switching on Ca2+-dependent signalling pathways that lead to the enhanced activation of transcription factors involved in antioxidants’ expression, (i.e., HIF-1/2a; CREB; NrF2). These factors promote autophagy, DNA integrity, mitochondrial function, and ATP production [147] (Figure 1). Oxidative-stress-mediated TRPM2 activation and consequent rise in intracellular Ca2+ levels might promote cancer progression also by activating Ca2+-dependent K+ channels, such as the large-conductance voltage-dependent BK channel and the medium-conductance voltage-independent KCa3.1 channel, as reported in melanoma cells [101], which role in cancer hallmarks, such as cell viability and cell migration and invasion have been described [59]. However, other studies correlated TRPM2 expression with a higher sensitivity to chemotherapy. Indeed, an anti-survival role was highlighted in breast and colon cancer, where TRPM2 activation by chemotherapy agents resulted in Ca2+ entry, intracellular Ca2+ overload, and increased mitochondrial depolarisation, leading to cell death [153], and in prostate cancer, where H2O2-induced TRPM2 activation resulted in PC3 cells’ death via Ca2+-dependent inhibition of autophagy [102] (Figure 1). These results could be explained considering the differential role of full-length TRPM2 and its dominant-negative short isoform [103]. Chen and colleagues elucidated their role both in in vitro and in vivo, demonstrating that full-length TRPM2 (TRPM2-L) had a protective role from oxidative stress by increasing antioxidant enzymes and pro-survival transcription factors, while the expression of the dominant-negative short isoform (TRPM2-S) reduced intracellular calcium influx in response to H2O2 treatment and cell survival [103].
Another important consideration is the key role of TRPM2 in neutrophil-mediated cytotoxicity. Neutrophils secrete H2O2, which activates TRPM2 expressed on cancer cells’ surfaces. This activation leads to Ca2+ influx in cancer cells, resulting in intracellular overload and induction of cell death, as was demonstrated in breast cancer cells [104] (Figure 1). Therefore, despite the pro-proliferative role of TRPM2 in several cancer cell lines, inhibition of its activity by tumour extracellular acidic pH might result in cancer cells’ protection from neutrophil cytotoxicity, with overall major efficiency in dissemination. Finally, TRPM2 is also localised in lysosomal membranes, where the highly acidic pH inside the compartment might prevent TRPM2 activation and Ca2+ release from lysosomes to the cytosol, which may trigger apoptosis [154] (Figure 1). These works reinforce the concept that TRPM2 inhibition by acidic pHe and activation by hypoxic conditions represent a protective mechanism for cancer cells.
TRPM6 and TRPM7 are both permeants mainly to Mg2+ ions and contribute to its homeostasis. These channels are also permeant to Ca2+ ions, increasing their intracellular concentration. TRPM7 has a unique structure as a “chanzyme” due to the presence of a kinase domain in its structure. TRPM6 has a tissue-specific expression and it is downregulated in several cancer types, while TRPM7 is ubiquitously expressed and mostly upregulated in different malignancies, where it plays a key role in promoting different cancer hallmarks [155].
TRPM6 and TRPM7 channels’ activity is modulated by both intra- and extracellular pH. For what concerns the effects of intracellular pH on TRPM7, the outward current density was decreased at low intracellular pH in HEK293 cells, with an IC50 of pH 6.32 and in the absence of Mg2+ [71]. This result was previously obtained also in RBL cells, where native TRPM7 currents were inhibited by intracellular acidification obtained by acetate treatment, and in TRPM7-overexpressing HEK293 cells in the same work [70]. Intracellular alkalinisation, induced by NH4+ extracellular application, determined the induction of native TRPM7 current and the enhancement of its activity in RBL cells [70,71] (Table 1).
The effects of acidic pHe (<6.0) on TRPM7 activity are quite controversial, with some evidence showing TRPM7 currents inhibition by low pHe [68,70,71,72], others potentiation of TRPM7 inward current by strong acidic pHe in HEK293 cells [69,73] and HeLa human cervical cells [74] (Table 1). The discrepancy of pHe modulation on TRPM7 described by these works could be explained by taking into consideration the importance of Ca2+ and Mg2+ ions present in the extracellular milieu. According to the work of Mačianskienė R. et al., both TRPM7 outward and inward currents, expressed by cardiomyocytes, were potentiated in the acidic extracellular medium (pH < 6) containing divalent Ca2+ and Mg2+ ions, while the absence of these ions in the acidic extracellular medium resulted in the low pHe-mediated inhibition of TRPM7 currents in a voltage-independent manner [75]. It is hypothesized that protons and divalent cations compete for a binding site within the channel pore, and the absence of these cations might allow protons to permeate the channel and to bind to specific intracellular inhibitory sites, the same bound by intracellular protons, leading to channel inhibition [73,75].
Very few data are available for TRPM6, where contrasting data are reported, as for TRPM7. TRPM6 was indeed inhibited by extracellular acidity similarly to TRPM7 in isolated pig myocytes [68] but potentiated by extracellular acidic pH in HEK293 cells, even though the magnitude of increase in TRPM7 inward current was higher than that of TRPM6 [69] (Table 1).
Concerning TRPM6 and TRPM7 regulation by hypoxia, their expression is increased in a hepatic ischemia-reperfusion rat model [105], while hypoxia-induced TRPM7 overexpression and increased intracellular Mg2+ concentrations were reported in rat hippocampal neurons in vitro [156]. TRPM7 sensitivity to hypoxia was also demonstrated by Tymianski and Mori group, showing that TRPM7 is activated by anoxic conditions or treatment with a hypoxic solution prepared by bubbling N2 gas [106,157] (Table 2).
Although evidence of the effect of acidic pH and hypoxia on TRPM6/7 in cancer cells is still lacking, several works have demonstrated that TRPM7 is upregulated in various cancers and it is involved in the enhancement of a variety of cancer-related processes regulated by Ca2+ signalling, such as proliferation, migration, invasion, cell death escape, and survival and epithelial–mesenchymal transition (EMT), via the activation of the Ras-ERK and the PI3K/AKT/mTOR signalling pathways [155] (Figure 1). Consequently, its activation by cancer-related extracellular acidic pH highlights its pivotal role in cancer progression.

2.2.2. TRP Vanilloid Subfamily

TRPV1 is a non-selective cation channel with relatively high permeability to Ca2+ ions, which is a major player in pain perception activated by different factors, including heat, inflammation, and acidic environment as revealed by David Julius, that shared the 2021 Nobel Prize for Physiology or Medicine [158,159,160]. Indeed, TRPV1 plays a key role in acidosis-induced pain, acting as a proton channel and being directly activated by protons [78,161].
Acidosis modulates TRPV1 activity, promoting its activation and potentiating its response to 2-APB, heat, and capsaicin (TRPV1 selective agonist) [76,78,80] (Table 1). hTRPV1 is indeed activated by mild acidosis (pHe 6.1) increasing intracellular Ca2+ levels, while the channel is blocked in presence of strong acidic conditions [76]. T633 residue in the pore helix and V538 residue in the S3–S4 linker were identified as key residues involved in extracellular pH sensing [77]. Low pHe (<5.9) significantly potentiated heat and capsaicin-evoked response in HEK293 cells by increasing the channel’s open probability at room temperature, therefore lowering the threshold for the channel activation, even in absence of chemical stimuli [78]. These data suggest that the potentiating effect of capsaicin and protons on TRPV1 are independent of each other and they are mediated by different TRPV1 residues (Table 1 and Figure 1). Concerning intracellular acidification, it enhanced TRPV1 currents evoked by 2-APB, without affecting the ones induced by capsaicin [80]. On the other hand, extracellular acidic pH activated TRPV1 and enhanced lymphatic endothelial cells’ proliferative, migratory and invasive abilities via activation of NF-κB transcription factor and consequent upregulation of IL-8, a lymphangiogenic factor, contributing to lymphatic metastasis in tumour acidic microenvironment context [79] (Figure 1). The role of TRPV1 in cancer is dependent on the cancer type. TRPV1 is indeed upregulated in many cancers and it regulates different cancer cell processes, such as proliferation, cell fate, migration, and invasion in a cancer type-specific manner, via the activation of different Ca2+-dependent signalling pathways, such as PI3K/AKT, Ras-ERK and JAK/STAT signalling cascades and NF-κB activation [162], acting both as an anti-proliferative and pro-apoptotic factor in melanoma, colorectal, pancreatic and liver cancer, among others, and exerting a pro-tumour role in highly aggressive types of cancer [162] (Figure 1 and Figure 3). These pro- or anti-tumour effects of TRPV1 can be attributed to the different opening states available to the channel in response of different stimuli, exploiting each opening state’s specific properties for the switching on of specific Ca2+-dependent signalling pathways in different cancer cell types. Consequently, the pHe regulation of TRPV1 might be considered as cell type, ligand, and context-specific, making it more difficult to identify its potential role as a pharmacological target.
Hypoxic TRPV1 modulation is reported in several publications, although cancer-related studies are lacking. Whole-cell patch clamp studies in HEK293T cells overexpressing rat TRPV1 showed that acute hypoxia weakly increased TRPV1 activity, but negatively affected capsaicin-induced TRPV1 currents. Hypoxia did not affect acidic pHe-activated TRPV1 current. These results were confirmed by Ca2+ imaging experiments, where hypoxia induced a slight increase in cytosolic Ca2+ levels [107]. A stronger hypoxia-mediated TRPV1 activation was demonstrated in native rat sensory neurons and HEK293 cells expressing rat or human TRPV1, where chronic hypoxia (>24 h) did not affect the channel expression but potentiated TRPV1’s activity via protein kinase C (PKC)ε- and HIF-1α-dependent signalling pathways [108]. Parpaite and colleagues also showed on rat pulmonary artery smooth muscle cells (PASMCs) that hypoxic conditions did not affect TRPV1 expression, but they increased TRPV1 activity probably via its translocation to the plasma membrane, enhancing PASMCs’ migratory abilities and proliferation [109]. Hypoxia-mediated changes in TRPV1 expression were instead reported in human PASMCs, where chronic hypoxia upregulated both TRPV1 gene and protein levels, increasing intracellular Ca2+ levels and being involved in PASMCs’ proliferation [110] (Table 2).
In addition to TRPV1, also TRPV2, TRPV3, and TRPV4 activity is modulated by acidic pH. TRPV2 is an intracellular-resident non-selective cation channel that translocates to the cell membrane following PI3K activation. Once at the plasma membrane, TRPV2-mediated Ca2+ entry regulates different physiological cellular processes. TRPV2 deregulation has been linked to several types of cancer, where its activity supports its progression, in particular via the activation of the PI3K/AKT and the ERK signalling cascades, by escaping cell death and increasing proliferation, cell migration, and invasion [163]. TRPV2 is known to be insensitive to low pHe alone [76,79], however, it has been demonstrated that acidic pHe (6.0 and 5.5) potentiated TRPV2 currents in transiently transfected HEK293 cells that are evoked by 2-APB, increasing the channel’s sensitivity to this ligand from the cytoplasmic side, as proved by inside-out patch configuration [80] (Table 1 and Figure 1).
Wang and colleagues did not report any change in TRPV2 expression following hypoxia exposure in human PASMCs [110], but H2O2-mediated oxidative stress upregulated TRPV2 gene and protein expression in two different human hepatoma cell lines, with induction or cell death by activating pro-apoptotic proteins and by inhibiting pro-survival ones, and by enhancing the sensitivity of human hepatoma cells to oxidative stress-associated chemicals [111] (Table 2).
TRPV3 is a non-selective calcium permeant cation channel mostly expressed in brain and skin, where it is involved in chemo-somatosensing. TRPV3 oncogenic activity was demonstrated in lung cancer, where TRPV3 expression was associated with short overall survival and Ca2+-mediated increased proliferation via Ca2+/calmodulin-dependent kinase II (CaMKII) [164]. As for TRPV2, acidic pHe alone (pH 5.5) was not able to activate the channel in HEK293 cells, but only to potentiate the TRPV3 response to 2-APB and its analogues via acidification of the intracellular milieu, increasing cytosolic Ca2+ levels [80] (Table 1 and Figure 1). Moreover, cytosolic protons activated the channel, inducing small but detectable currents via different mechanisms with respect to 2-APB response potentiation, and indicating four residues in the S2–S3 linker to be implicated in the acid intracellular activation of TRPV3 [80]. A more recent study from the same group has elucidated the mechanism of TRPV3 acid intracellular activation and extracellular inhibition. The authors identified Asp641 residue, localized in the selectivity filter, as a critical residue involved in TRPV3 extracellular acidic pHe inhibition. Intracellular acidification protonated E682, E689, and D727 residues in the C-terminal, facilitating the channel’s sensitization [165]. These data were also obtained by a previous work on TRPV3-transfected HEK293 and in HaCaT cells, where it was shown that TRPV3 is directly activated by glycolic acid-induced cytosolic acidification, inducing cell death, while extracellular acidification failed to activate the channel, resulting instead in decreased current amplitude [81] (Table 1 and Figure 1 and Figure 3).
Concerning hypoxic regulation, no data on cancer cells is available. As for TRPV2, no differences in TRPV3 expression were reported in hypoxia exposed-PASMCs, but a recent work demonstrated TRPV3 upregulation in rat myocardial cells in response to ischemia/hypoxia treatment. To confirm the role of TRPV3 as a mediator of hypoxia-induced inflammation and apoptosis in rat myocardial cells, the authors reported thatTRPV3 silencing protected cardiomyocytes from hypoxia-induced cell death and decreased the secretion of pro-inflammatory cytokines [112]. TRPV3 mediated hypoxia responses also in pulmonary artery smooth muscle cells, promoting their proliferation via PI3K/AKT signalling pathway [113], and hypoxia also robustly potentiated this channel current in TRPV3-overexpressing HEK293 cells in response to 2-APB treatment, without affecting the protein levels [114] (Table 2).
TRPV4 is a heat-activated and mechanosensitive channel deregulated in different cancer cells, acting mostly as a pro-tumour factor, enhancing cancer cells’ migration and metastasis through the activation of AKT and Rho/ROCK1/cofilin cascade, extracellular remodelling, proliferation and angiogenesis via activation of NFAT and PI3K signalling pathway [166,167,168,169,170,171,172], although a tumour-suppressive role has also been reported, especially in tumours expressing high TRPV4 levels [173,174] (Figure 1 and Figure 3).
For what concerns TRPV4 regulation by pH, low extracellular pH (pHe 6) induced opening of transiently expressed TRPV4 in Chinese hamster ovary cells, with a maximal potentiation at pHe 4, as demonstrated by patch clamp current recording in absence of extracellular Ca2+ ions [82]. An opposite effect was observed in mouse oesophageal epithelial cells, where Ca2+ imaging experiments showed that the Ca2+ influx mediated by TRPV4 was abolished at pHe 5 [83] (Table 1). These opposite results might be explained considering the different techniques used by the two research groups and the experimental conditions in general. TRPV4 is a non-selective ion channel, permeable to protons when the extracellular solution is free from other divalent ions, such as Ca2+ ions. Since protons may compete with Ca2+ ions, a high extracellular proton concentration might lead to a decrease in Ca2+ influx but contributes to the gross TRPV4 current [83]. Collectively, considering the evidence outlined above, TRPV4’s role in tumour biology is cancer type-specific and it might emerge as a potential drug target in the context of cancer treatment.
Regarding hypoxic TRPV4 regulation, no studies on cancer cells are present in the literature, but they are limited to other pathological states. In particular, its mRNA and protein expression levels were increased after 6 h long exposure to hypoxia in cardiomyocytes, inducing Ca2+ overload and enhancing oxidative injury [115]. TRPV4 plays also an important role in mediating hypoxia-induced pulmonary vasoconstriction (HPV) [175] and in the hypoxia/ischemia injury in the brain [116] (Table 2).
pH also regulates the activity of another component of the TRP vanilloid family, TRPV6, a highly Ca2+ selective channel (PCa/PNa~100) that is upregulated in different epithelial cancers, such as prostate, pancreatic, breast and ovarian cancer, in particular during early stages of tumour progression [176]. Several studies has revealed the positive effect of TRPV6 activity on tumour progression through the activation of Ca2+-dependent signalling pathways [176], promoting cancer proliferation and cell survival in prostate cancer cells by activating the Ca2+-dependent NFAT transcription factor [177,178], invasion of breast cancer cells through Ca2+/Calmodulin (CaM)-dependent kinases, such as CaMKII [179,180], cell survival, proliferation, and invasion in pancreatic cancer cells [181] and tumour growth in in vivo ovarian adenocarcinoma xenograft mouse model [182] (Figure 1 and Figure 3).
TRPV6 is a polymodal sensor that is regulated by different chemical and physical stimuli, including acidic pHe and hypoxia. Alkaline pHe positively modulated TRPV6’s activity in Jurkat T-cells, where whole-cell patch-clamp experiments showed that solution at pH 8.2 determined the increase in TRPV6 activity and Ca2+ entry, while opposite effects were experienced for acidic pHe (pH = 6), which reduced inward TRPV6 currents and Ca2+ influx in Jurkat T-cells [84] (Table 1).
Inhibition of TRPV6 by extracellular acidification might be explained considering TRPV6 expression in distinct stages of cancer progression. Tumours can be characterized by a more acidic extracellular microenvironment during late stages [183], phases in which TRPV6 expression is downregulated in some types of cancers, such as colon cancer [184]. However, the lack of further studies focusing on the effect of extracellular acid pH on TRPV6 activity makes a comprehensive understanding of the role of acid pHe on TRPV6 in the context of tumour acidosis difficult.
As for many other TRPV channels, to date, there is extremely little information about TRPV6 modulation by hypoxia, especially in the tumourigenesis context, with a report of placental TRPV6 overexpression in hypoxic rats [185].

2.2.3. TRP Ankyrin Subfamily

TRPA1 is a non-selective cation channel that functions as a polymodal sensor, and deregulation is observed in several malignancies in a tissue-specific manner [186]. TRPA1 has been described as highly sensitive to O2 and oxidants in vagal and sensory neurons. In particular, TRPA1 is activated in hypoxic conditions by the relief of the inhibition of prolyl hydroxylase (PDH) which is O2 sensitive [187].
The role of TRPA1 as a ROS sensor has been demonstrated also in cancer cells. The work of Takahashi et al. has demonstrated a key role of TRPA1 in promoting resistance to ROS-producing chemotherapies and oxidative stress tolerance in breast cancer cells via ROS-mediated TRPA1 activation and Ca2+-CaM/PYK2 signalling pathway [117] (Table 2 and Figure 1). TRPA1 interaction with fibroblast growth factor receptor 2 (FGFR2) induces the activation of the receptor, promoting cancer cells’ proliferation and invasion, prompting lung adenocarcinoma metastasisation to the brain [188]. TRPA1 activation might also promote prostate cancer progression by triggering prostate cancer stromal cells’ secretion of VEGF [189], a known mitogenic factor involved in the proliferation, migration and invasion of prostate cancer cells [190]. In line with those results, TRPA1 is indeed also an important player in promoting angiogenesis both in physiological retinal development as well as in prostate cancer-derived endothelial cells [191].
In addition to being activated by oxidants, TRPA1 is activated by several distinct exogenous and endogenous compounds [192] and by protons in the extracellular environment, on which regulation is specie dependent. HEK293 cells expressing human and rodent TRPA1 showed a specie-specific activation of the channel, where only hTRPA1 generated a membrane current when exposed to different extracellular acidic environments (pHe 6.4-5.4) reaching the maximal response at acidic pHe 5.4; this specific effect for hTRPA1 was confirmed by Ca2+ imaging experiments in HEK293 cells as well as in DRG neurons derived from TRPV1/TRPA1−/− mice and in neuroblastoma ND7/23 cells expressing hTRPA1, where only hTRPA1 induced an increase in Ca2+ entry when exposed to acidic pHe. This specie-specific activation of TRPA1 is due to valine and serine residues within transmembrane domains 5 and 6 [85] (Table 1).
TRPA1 can also be activated by intracellular acidification, as observed in the context of ischemia-induced acidification of the extracellular microenvironment in mice oligodendrocytes, with following acidification of the intracellular space and activation of TRPA1. This led to an increase in Ca2+ influx and damage to myelin [118] (Table 1). Altogether, these results and their role in cancer suggest that TRPA1 activation by acidic pHe and hypoxia may play a significant role in promoting cancer progression, highlighting its potential as a therapeutic target.

2.2.4. TRP Canonical Subfamily

TRPC1 is a non-selective cation channel, which assembles to form homo- and heteromeric channels with other members of the family, such as TRPC3, TRPC4 and TRPC5. TRPC1 upregulation has been reported in several cancers, such as pancreatic cancer, where it potentiates BxPc3 cells’ migration via Ca2+-dependent activation of PKCα [193], and breast cancer, where it exerts a pro-proliferative role in MCF-7 cells by mediating Ca2+ influx induced by KCa3.1 activation [194] and via Ca2+-dependent ERK1/2 activation [195]. TRPC1 also promotes human glioma cancer cells’ proliferation via Ca2+ entry and supports tumour growth in vivo [196], and lung cancer differentiation, by promoting A549 cell proliferation [197]. TRPC1 also participates with ORAI1 channels in the induction of vimentin, a mesenchymal marker for epithelial–mesenchymal transition (EMT) [198] (Figure 2 and Figure 3).
Although TRPC1 is not regulated by protons, several data demonstrated that TRPC1 expression is hypoxia-mediated. The study by Wang B. et al. in 2009 reported that TRPC1 is functionally expressed in U-87 malignant glioma cells under hypoxia, where it promoted the upregulation of VEGF expression, as VEGF mRNA levels were significantly decreased in presence of TRPC1 inhibitor or RNAi in hypoxic conditions [119] (Table 2 and Figure 2). VEGF has a central role in angiogenesis in both physiological and pathological conditions [199] and solid tumours are characterized by a hypoxic microenvironment, in which the lack of oxygen might promote VEGF expression, in order to induce angiogenesis and increase tumour oxygen supply. TRPC1 mediates hypoxia responses also in breast cancer cells, where HIF-1α promoted its upregulation. In MDA-MB-468 breast cancer cells, TRPC1 is involved in the transactivation of the epidermal growth factor receptor (EGFR) during hypoxia, leading to the increase in LC3B autophagy marker [120]. Moreover, hypoxia-induced TRPC1 activation promoted epithelial–mesenchymal transition in the same cells, upregulating the mesenchymal marker snail and downregulating the epithelial marker claudin-4, promoting the hypoxia-induced EMT and, therefore, the aggressive and invasive phenotype of breast cancer cells [120] (Table 2 and Figure 2).
TRPC5 is a Ca2+-activated ion channel that is regulated by components of the tumour microenvironment, such as acidosis and supports hypoxia responses. In long-term adriamycin-treated breast cancer cells, TRPC5-mediated Ca2+ influx promoted HIF-1α translocation in the nucleus and therefore the downstream transcription of HIF-1α-regulated VEGF expression, highlighting its contribution to promoting breast cancer angiogenesis [121] (Table 2 and Figure 2). The same research group also validated TRPC5 role in mediating HIF-1α response in tumour progression in colon cancer, where TRPC5 activated the HIF-1α-Twist signalling pathway to promote EMT, migration and proliferation in SW620 colon cancer cells [122].
TRPC5 also acts as a pHe sensor, as its spontaneous activity and G protein-activated currents are potentiated by extracellular acidic pH by increasing the channel open probability in presence of small changes of extracellular pH, with a maximum activity around pHe 6.5 and current inhibition starting from pH 5.5 [86] (Table 1).
Potentiation by acidic pHe and involvement in HIF-1α regulation demonstrates the interest of cancer cells in keeping TRPC5 channels active and over-expressed, in order to promote cancer-specific hallmarks, in particular chemoresistance. Indeed, the role of TRPC5 in promoting chemoresistance in different types of cancer is well known. In breast carcinoma cells, adriamycin-induced TRPC5 upregulation protects them from chemotherapy treatment, by inducing autophagy via an increase in cytosolic Ca2+ concentrations and activation of the Ca2+-dependent CaMKKβ/AMPKα/mTOR pathway, promoting the cancer cell survival and tumour growth in vivo [200] (Figure 2). The work by Ma X. et al. in 2012 has also demonstrated in vitro and in vivo the critical role of this channel in promoting chemoresistance of adriamycin-resistant MCF-7 breast cancer cell line with the upregulation of another pump linked to drug resistance. TRPC5 overexpression resulted indeed in Ca2+ influx and activation of P-glycoprotein overproduction, a pump in charge of removing cytotoxic drugs from cells via Ca2+/calmodulin/calcineurin-dependent NFATc3 signalling pathway [201]. The role of TRPC5 in therapy resistance is not confined to breast cancer, as its upregulation has been also identified in 5-fluorouracil resistant human colorectal cancer cells, where TRPC5 overexpression determines the overproduction of ATP-binding cassette subfamily B member 1 (ABCB1), a pump involved in drug resistance through the export of cytotoxic drugs, via Ca2+ entry and activation of Ca2+-dependent Wnt/β-catenin signalling pathway and in a glycolysis-dependent manner [201,202,203] (Figure 2).
In addition to chemoresistance, TRPC5 expression was observed to be positively correlated with high proliferative, migratory, and invasive abilities of colon cancer cells, promoting the EMT through the HIF-1α-Twist signalling pathway [122]. Therefore, these results demonstrate that the impact of the acid and hypoxic cancer microenvironment on the TRPC5 channel is aimed at its activation, thus promoting tumour progression (Figure 3).
Similar behaviour is shown by TRPC4, which shares high sequence similarity with TRPC5. TRPC4 is a Ca2+- and G-coupled receptors-activated non-selective Ca2+ permeable cation channel [204]. TRPC4’s role in cancer has been elucidated in the last years, being involved in promoting angiogenesis via cytosolic Na+ and Ca2+ rise [205,206] and proliferation [207], although a negative effect on A-498 renal cell carcinoma cells via Englerin A-mediated channel activation has been documented [208].
TRPC4-mediated currents are two-fold potentiated when exposed to pHe 6.5; however, lower pH leads to current inhibition starting from pHe 6, with complete current inhibition at pHe 5.5 [86]. Gi/o-mediated TRPC4 activation is also accelerated by intracellular protons in an indirect way, regulating the kinetics of Gi/o-dependent TRPC4 activation, and it requires an increase in intracellular Ca2+ concentration. Intracellular protons do not act directly on the channel, as they inhibit TRPC4 activation by its direct agonist, Englerin A, but by acting on PLCδ1 [87] (Table 1; Figure 2 and Figure 3). No data are available to our knowledge about hypoxia regulation of TRPC4 expression and/or activity.
TRPC6 is another TRPC member to be regulated by tumour microenvironmental clues, such as hypoxia. In fact, hypoxic conditions enhanced TRPC6 expression in murine pancreatic stellate cells, which constitute the major cellular components in pancreatic ductal adenocarcinoma’s stroma, and which play a key role in generating PDAC’s characteristic desmoplasia. TRPC6 promoted their activation and it was involved in the secretion of pro-migratory factors in presence of hypoxia [123]. Hypoxia upregulated TRPC6 mRNA expression in hepatocellular carcinoma cells, where TRPC6-mediated Ca2+ influx conferred drug resistance to these cells via the Ca2+-dependent STAT3 signalling pathway in hypoxic conditions [125]. Hypoxia activated Notch1 and downstream TRPC6 expression also in glioma cells, with a consequent rise in cytosolic Ca2+ concentration and Ca2+-dependent activation of the calcineurin-NFAT signalling pathway, promoting proliferation, cell invasion and angiogenesis under hypoxia [126] (Table 2, Figure 1 and Figure 3). Hypoxia also upregulated TRPC6 in hepatic stellate cells via HIF-1α/Notch1 pathway, leading to TRPC6-mediated Ca2+ influx and the downstream activation of Ca2+-dependent nuclear factor of activated T-cells (NFAT) transcription factor and SMAD2/3-dependent TGF-β signalling, which activation resulted in the expression of ECM proteins, such as collagen type I, that facilitate hepatic stellate cells’ fibrotic activation and promotes hepatic fibrosis, strongly linked to arise of hepatocellular carcinoma [124] (Table 2; Figure 1 and Figure 3).
Regarding the direct role of acidic extracellular pH (around pH 6.5), it is sufficient to inhibit TRPC6, and the inhibition increased in a pH-dependent manner, affecting both inward and outwards currents in HEK-transfected cells [86] (Table 1). This inhibitory effect of acidic pHe might be explained considering the high TRPC6 levels expressed by pro-tumourigenic immune cells, such as neutrophils, where TRPC6-mediated calcium entry is required for CXCR2-mediated intermediary chemotaxis [209]. Consequently, inhibition of TRPC6 by acidic pHe may impair neutrophils’ migration and prevent them from leaving the acidic tumour microenvironment, thus contributing to its progression and metastasis by releasing ROS, secreting pro-tumour factors and inducing drug resistance [210]. Altogether, these results suggest that the acidic pHe-mediated potentiation of TRPC channels might be restricted to only some members, and it depends on the cell type expressing the channels.

2.3. Store-Operated Ca2+ Channels

ORAIs are Ca2+ release-activated Ca2+ channels (CRAC) which are major players, with STIM proteins, in the mechanism known as store-operated Ca2+ entry (SOCE), which mediates Ca2+ entry into cells promoting the refilling of ER calcium stores as well as intracellular signalling, controlling both physiological and pathological processes such as inflammation, cell motility, cell proliferation, gene expression, apoptosis escape and cell invasion [211,212,213]. SOCE represents the main route of Ca2+ entry in different types of cancer cells, contributing to several cancer hallmarks. Indeed, different works have highlighted the key role of SOCE in promoting the migration of different cancer cell lines, such as chemoresistant IGROV1 ovarian cancer cells by regulating focal adhesion turnover [214], SW480 colorectal carcinoma cells [215] and oral cancer cells through Akt/mTOR/NF-κB signalling [216]. Moreover, SOCE has been implicated in enhancing invasion of triple-negative breast cancer cells, as well as angiogenesis and migration, through NFAT4 signalling [128] and through NFATc3 in colorectal cancer cells and tissues from patients [129], while in WM793 cells melanoma cells, SOCE-induced Ca2+ oscillations contribute to invadopodia formation via Src activation [217]. SOCE might promote invasion of cancer cells by inducing epithelial–mesenchymal transition, as observed in DU145 and PC3 prostate cancer cells [218] and BGC-803 and MKN-45 gastric cancer cells [219]. Store-operated channels (SOCs) in general play a key role also in the modulation of sensitivity to chemotherapy in a cancer type-specific manner, by promoting chemoresistance in breast cancer cells in the case of ORAI3 [220], ORAI1 and STIM1 in pancreatic ductal adenocarcinoma [212], ORAI1 in hepatocarcinoma [221], among others. SOCE also promotes extracellular vesicle formation, which is a signalling vector involved in the intercellular acquisition of multidrug resistance, in both malignant and non-malignant breast cancer cells via activation of calpain [222].
In addition to all these contributions to cancer progression, SOCE is regulated by hypoxia and pH. Several studies have demonstrated that intracellular and extracellular pH is able to modulate the activity of ORAI channels by affecting its coupling with STIM1 and/or by modifying its gating biophysical properties. The results obtained by numerous studies conducted on ORAI/STIM have clarified the concept that both intra- and extracellular acidic pH have an inhibitory effect on the activity of ORAIs, and on SOCE in general, while intra- and extracellular basic pH potentiate them. The notion that extracellular pH regulates native CRAC currents (ICRAC) was already known in 1995 when the work of Malayev A. and Nelson D.J. showed that acidic extracellular pH (pHe = 6) decreased the amplitude of inward Ca2+ currents while basic pHe (pHe = 8) increased it in macrophages by using the patch clamp technique and that these changes were reversible and voltage-independent [88] (Table 1). More recently, inhibition by acidic pHe was demonstrated in H4IIE rat liver cells overexpressing ORAI1 and STIM1 proteins, in which ICRAC was inhibited completely at pHe 5.5 [89] (Table 1). In the same work, researchers identified E106, located in ORAI1’s pore, as the residue responsible for pH dependence of CRAC currents, as E106D mutation in ORAI1 abolished the inhibition of ICRAC by acidic pHe. These results were also supported by the work of Beck A. and colleagues, which demonstrated that extracellular and intracellular acidification decreased the amplitude of IP3-induced endogenous ICRAC in RBL2H3 mast cell line, in Jurkat T lymphocytes and in heterologous ORAI/STIM-mediated ICRAC in HEK293 cells [90] (Table 1). In contrast to acidification, external alkalinisation increased both endogenous and overexpressed ORAI/STIM amplitude of ICRAC (pKa of 7.8 for RBL2H3 mast cells, 8.0 for Jurkat T lymphocytes and 7.9 for HEK293 cells). Other two key residues (D110 and D112) located in ORAI1’s first extracellular loop have been proposed to contribute to some extent to pHe sensitivity. Indeed, mutations of these residues to alanine prevented the alkalinisation-induced potentiation of ICRAC and increased its amplitude in the presence of acidic pHe [90]. Enhancement and the decrease in SOCs activity by external basic and acidic pH, respectively, were further confirmed in heterologous ORAI/STIM-mediated currents in HEK293 cells by Tsujikawa H. et al. in 2015, who have also demonstrated that E106 mediates pHe sensitivity when Ca2+ is the permeant cation, while E190 when Na+ is the permeant cation [91] (Table 1). However, the effect of alkaline pHi on ICRAC is controversial. Indeed, alkaline pHi-mediated potentiation of ORAI1/STIM1 activity was observed in other papers [91,223] (Table 1). These differences might be explained considering the type of intracellular Ca2+ buffer used [223]. Moreover, cytosolic alkalinisation led to SERCA inhibition, resulting in Ca2+ release from ER stores and activation of SOCs, with Ca2+ influx in NIH 3T3 cells [224]. Residual H155 located in the intracellular loop of ORAI1 is responsible for ORAI1/STIM1 pHi sensitivity, as mutation to phenylalanine decreased low pHi-mediated ICRAC inhibition and alkaline pHi-mediated ICRAC potentiation [91]. Since the effect of pH was the same in presence of all ORAI isoforms [90] and that both extracellular pH sensors (residues E106 and E190) and intracellular one (residue H155) are conserved in all three ORAI isoforms, it might suggest that the residues mentioned before act as common pHi and pHe sensors in ORAI1–2–3/STIM isoforms. In addition to H155, negatively charged amino acid residues in the STIM1 inactivation domain play an important role in pHi sensitivity [223].
For what concerns hypoxia-mediated regulation of SOCs, intracellular acidification of primary aortic smooth muscle cells and HEK293 cells induced by hypoxia led to inhibition of SOCE by disrupting the electrostatic ORAI1/STIM1 binding and closing ORAI1 channel (Table 2 and Figure 2). Nonetheless, STIM1 remained associated with ORAI1 through the second binding site located between ORAI1’s intracellular N-terminal tail and STIM1’s STIM-ORAI activating region (SOAR), preventing the noxious hypoxia-mediated Ca2+ overload [92]. Therefore, intracellular hypoxia-mediated acidification might regulate SOCs activity by uncoupling ORAI11 and STIM1 and, consequently, reducing ICRAC amplitude.
As ORAI/STIM mediate most of the Ca2+ signalling in cancer-induced acidosis, they play a key role in several pathological processes. In the cancer context, it was observed that SOCE is regulated by changes in extracellular pH, as acidification of tumour microenvironment suppressed both the carbachol-(CCH) and thapsigargin (TG)-mediated Ca2+ entry in neuroblastoma cells, while external alkalinisation increased both the CCH- and the TG-induced Ca2+-influx [93], therefore in accordance with the results obtained in non-cancer cells (Table 2 and Figure 2).
However, the work of Liu X. et al. of 2018 on triple-negative breast cancer cells (TNBCs) reported that hypoxia promoted the activation of Notch1 signalling, required for the upregulation of ORAI1 mRNA and protein levels in TNBCs (Table 2 and Figure 2). In addition, the upregulation of ORAI1 in hypoxia determined an increase in basal Ca2+ concentration and of thapsigargin-induced SOCE, which activated the downstream NFAT4 target, known to regulate the expression of cancer-related genes involved in its hallmarks [128]. The results of the same work showed that hypoxia enhanced invasion and TNBC migration via the Notch1/ORAI1/SOCE/NFAT4 pathway; therefore, ORAI1 and SOCE play a key role in promoting an aggressive phenotype. Same results were obtained in colon cancer cells by the same group of researchers, where potentiation of SOCE mediated by hypoxia-induced upregulation of ORAI1 determined the activation of NFATc3, enhancing hypoxia-induced invasion and angiogenesis in colon cancer cells [129]. The hypoxia-induced upregulation of SOCE components was also demonstrated in A549 and NCI-H292 non-small cell lung cancer cells, where nicotine treatment determined the upregulation of HIF-1α, which increased the expression of the SOCE components TRPC1, TRPC6 and ORAI1. This translated into potentiation of SOCE and calcium entry, promoting lung cancer cells’ proliferation [127].
Considering the acidosis and hypoxia-mediated effects presented, the inhibitory action of the tumour intra- and extracellular acidic pH on SOC channels seems to contradict the positive effect of SOCE on tumour progression. In particular, acidosis inhibition of SOCE does not correlate with the different studies demonstrating that ORAI/STIM and SOCE promote cancer development [225] and that also tumour acidosis and hypoxia support it by enhancing tumour cell migration, invasion and, therefore, its aggressive phenotype [40,226]. Possible explanations for the apparent counterproductive blockade of SOCE channels by the tumour’s acidic pH lie in the fact that Ca2+ signalling not only promotes tumour progression and development by potentiation of its hallmarks but also contributes to its suppression by enhancing processes such as cell death, senescence and autophagy [49]. In addition, ORAI members assembly to form different combinations of heteromeric Ca2+ release-activated channels (CRACs) and the ratio of each ORAI member determines specific ICRAC current properties and CRAC effects [227,228]. Therefore, the acidic pH of the tumour microenvironment may differently regulate heteromeric CRACs. Another point to consider is the key role of SOCE in immune cell activation [229]. The requirement of Ca2+ entry for antitumour immunity might explain the inhibitory effect of acidic tumour microenvironment on SOCE, in order to decrease immune cells’ function and protect the tumour.

3. Concluding Remarks

The present review aimed to overview the crosstalk between major chemical components of the tumour microenvironment, hypoxia and acidic pHe, and Ca2+-permeable ion channels, summarizing the major Ca2+-mediated signalling pathways that are involved in hypoxia and acidic pHe responses in cancer cells. We focused our attention mainly on not-voltage gated TRPs, SOCs as well as Piezo channels due to their role as polymodal sensors in the tumour microenvironment. In this perspective, the data presented showed that hypoxia has a positive effect on some TRP and ORAI1 channels (Figure 3), promoting their activation and their expression via different transcription factors. On the contrary, tumour acidosis modulation shows a higher variability, determining loss-of-function of specific Ca2+-permeable ion channels expressed not only by cancer cells but also by immune cells, or potentiation or activation of others.
Although the separate roles of tumour acidosis, hypoxia and Ca2+ signalling in cancer progression are well established, less is known about their crosstalk in this context, with the majority of works cited in this review performed on normal cell lines and with a focus on acidic pHe effect on Ca2+-permeable ion channels’ current via electrophysiology experiments. In addition, investigation of intracellular acidification-mediated regulation of Ca2+ signals and the pHi-dependence of Ca2+-permeable channels in cancer is rare, although it is well known that intracellular pH plays a more decisive role in regulating various biological processes and that this value is highly influenced by the extracellular pH. Based on the established knowledge about Ca2+-dependent signalling pathways involved in tumour progression and on the information about the effects of hypoxia and acidosis on the activity and expression of different Ca2+-permeable channels mostly available from normal or tumourigenic cells studies, the present review critically presented and discussed this crosstalk to elucidate and hypothesize which of TRPs, ORAIs and PIEZO channels Ca2+-dependent signalling pathways could be activated or inhibited by hypoxia and tumour acidosis-mediated regulation of Ca2+-permeable ion channels in cancer cells and involved in tumour progression, for the identification of potential molecular identities to target. In particular, the putative synergistic relationship between hypoxia, tumour acidosis and low pHe-induced intracellular acidification and Ca2+ signals, the mechanisms of their interaction and their interdependence in tumours require further studies and clarifications, in order to fill the gap and promote a better understanding of the crosstalk between three major players in the cancer research field.

Author Contributions

Conceptualisation, M.M.A., A.F.P. and N.P.; methodology, M.M.A., A.F.P. and N.P.; investigation, M.M.A. and A.F.P.; writing—original draft preparation, M.M.A., A.F.P. and N.P.; writing—review and editing, M.M.A., A.F.P. and N.P.; visualisation, M.M.A., A.F.P. and N.P.; supervision, A.F.P. and N.P. All authors have read and agreed to the published version of the manuscript.

Funding

M.M.A. and N.P. are supported by the European Marie Skłodowska—Curie Innovative Training Network (ITN) pH and Ion Transport in Pancreatic Cancer–pHioniC (#GA 813834 pHioniC H 2020 MSCA ITN 2018). A.F.P. and M.M.A. are supported by PRIN 2017 “LIONESS” project (#GA 754345) from the Italian Ministry for Education, University and Research. A.F.P., N.P. and M.M.A. are part of the Associated International Laboratory (CaPANCInv).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wei, R.; Liu, S.; Zhang, S.; Min, L.; Zhu, S. Cellular and Extracellular Components in Tumour Microenvironment and Their Application in Early Diagnosis of Cancers. Anal. Cell. Pathol. 2020, 2020, 6283796. [Google Scholar] [CrossRef] [Green Version]
  2. Warburg, O. On the Origin of Cancer Cells. Science 1956, 123, 309–314. [Google Scholar] [CrossRef]
  3. Cassim, S.; Vučetić, M.; Ždralević, M.; Pouyssegur, J. Warburg and Beyond: The Power of Mitochondrial Metabolism to Collaborate or Replace Fermentative Glycolysis in Cancer. Cancers 2020, 12, 1119. [Google Scholar] [CrossRef]
  4. Marchiq, I.; Le Floch, R.; Roux, D.; Simon, M.-P.; Pouyssegur, J. Genetic Disruption of Lactate/H + Symporters (MCTs) and Their Subunit CD147/BASIGIN Sensitizes Glycolytic Tumour Cells to Phenformin. Cancer Res. 2015, 75, 171–180. [Google Scholar] [CrossRef]
  5. Ždralević, M.; Brand, A.; Di Ianni, L.; Dettmer, K.; Reinders, J.; Singer, K.; Peter, K.; Schnell, A.; Bruss, C.; Decking, S.-M.; et al. Double Genetic Disruption of Lactate Dehydrogenases A and B Is Required to Ablate the “Warburg Effect” Restricting Tumour Growth to Oxidative Metabolism. J. Biol. Chem. 2018, 293, 15947–15961. [Google Scholar] [CrossRef] [Green Version]
  6. Swietach, P.; Hulikova, A.; Vaughan-Jones, R.D.; Harris, A.L. New Insights into the Physiological Role of Carbonic Anhydrase IX in Tumour PH Regulation. Oncogene 2010, 29, 6509–6521. [Google Scholar] [CrossRef] [Green Version]
  7. Petrova, V.; Annicchiarico-Petruzzelli, M.; Melino, G.; Amelio, I. The Hypoxic Tumour Microenvironment. Oncogenesis 2018, 7, 10. [Google Scholar] [CrossRef]
  8. Boedtkjer, E.; Pedersen, S.F. The Acidic Tumour Microenvironment as a Driver of Cancer. Annu. Rev. Physiol. 2020, 82, 103–126. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Webb, B.A.; Chimenti, M.; Jacobson, M.P.; Barber, D.L. Dysregulated PH: A Perfect Storm for Cancer Progression. Nat. Rev. Cancer 2011, 11, 671–677. [Google Scholar] [CrossRef]
  10. Persi, E.; Duran-Frigola, M.; Damaghi, M.; Roush, W.R.; Aloy, P.; Cleveland, J.L.; Gillies, R.J.; Ruppin, E. Systems Analysis of Intracellular PH Vulnerabilities for Cancer Therapy. Nat. Commun. 2018, 9, 2997. [Google Scholar] [CrossRef] [PubMed]
  11. Kazyken, D.; Lentz, S.I.; Fingar, D.C. Alkaline Intracellular PH (PHi) Activates AMPK–MTORC2 Signaling to Promote Cell Survival during Growth Factor Limitation. J. Biol. Chem. 2021, 297, 101100. [Google Scholar] [CrossRef] [PubMed]
  12. 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]
  13. Pérez-Tomás, R.; Pérez-Guillén, I. Lactate in the Tumour Microenvironment: An Essential Molecule in Cancer Progression and Treatment. Cancers 2020, 12, 3244. [Google Scholar] [CrossRef] [PubMed]
  14. Flinck, M.; Kramer, S.H.; Schnipper, J.; Andersen, A.P.; Pedersen, S.F. The Acid-Base Transport Proteins NHE1 and NBCn1 Regulate Cell Cycle Progression in Human Breast Cancer Cells. Cell Cycle 2018, 17, 1056–1067. [Google Scholar] [CrossRef]
  15. Rengifo Rodriguez, J.E.; Garcia-Perdomo, H.A. Role of Monocarboxylate Transporters in the Diagnosis, Progression, Prognosis, and Treatment of Prostate Cancer. Turk. J. Urol. 2020, 46, 413–418. [Google Scholar] [CrossRef]
  16. Ward, C.; Meehan, J.; Gray, M.E.; Murray, A.F.; Argyle, D.J.; Kunkler, I.H.; Langdon, S.P. The Impact of Tumour PH on Cancer Progression: Strategies for Clinical Intervention. Explor. Target. Anti-Tumour Ther. 2020, 1, 71–100. [Google Scholar] [CrossRef] [Green Version]
  17. Mboge, M.; Mahon, B.; McKenna, R.; Frost, S. Carbonic Anhydrases: Role in PH Control and Cancer. Metabolites 2018, 8, 19. [Google Scholar] [CrossRef] [Green Version]
  18. Mookerjee, S.A.; Goncalves, R.L.S.; Gerencser, A.A.; Nicholls, D.G.; Brand, M.D. The Contributions of Respiration and Glycolysis to Extracellular Acid Production. Biochim. Biophys. Acta BBA Bioenerg. 2015, 1847, 171–181. [Google Scholar] [CrossRef] [Green Version]
  19. Swietach, P. What Is PH Regulation, and Why Do Cancer Cells Need It? Cancer Metastasis Rev. 2019, 38, 5–15. [Google Scholar] [CrossRef] [Green Version]
  20. Luoto, K.R.; Kumareswaran, R.; Bristow, R.G. Tumour Hypoxia as a Driving Force in Genetic Instability. Genome Integr. 2013, 4, 5. [Google Scholar] [CrossRef] [Green Version]
  21. Muz, B.; de la Puente, P.; Azab, F.; Azab, A.K. The Role of Hypoxia in Cancer Progression, Angiogenesis, Metastasis, and Resistance to Therapy. Hypoxia 2015, 3, 83–92. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yang, C.; Zhong, Z.-F.; Wang, S.-P.; Vong, C.-T.; Yu, B.; Wang, Y.-T. HIF-1: Structure, Biology and Natural Modulators. Chin. J. Nat. Med. 2021, 19, 521–527. [Google Scholar] [CrossRef]
  23. Doe, M.R.; Ascano, J.M.; Kaur, M.; Cole, M.D. Myc Posttranscriptionally Induces HIF1 Protein and Target Gene Expression in Normal and Cancer Cells. Cancer Res. 2012, 72, 949–957. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Ippolito, L.; Morandi, A.; Giannoni, E.; Chiarugi, P. Lactate: A Metabolic Driver in the Tumour Landscape. Trends Biochem. Sci. 2019, 44, 153–166. [Google Scholar] [CrossRef]
  25. Tam, S.Y.; Wu, V.W.C.; Law, H.K.W. Hypoxia-Induced Epithelial-Mesenchymal Transition in Cancers: HIF-1α and beyond. Front. Oncol. 2020, 10, 486. [Google Scholar] [CrossRef]
  26. Mazure, N.M.; Pouysségur, J. Hypoxia-Induced Autophagy: Cell Death or Cell Survival? Curr. Opin. Cell Biol. 2010, 22, 177–180. [Google Scholar] [CrossRef]
  27. Cosse, J.-P.; Michiels, C. Tumour Hypoxia Affects the Responsiveness of Cancer Cells to Chemotherapy and Promotes Cancer Progression. Anti-Cancer Agents Med. Chem. 2008, 8, 790–797. [Google Scholar] [CrossRef]
  28. Muz, B.; de la Puente, P.; Azab, F.; Luderer, M.; Azab, A.K. Hypoxia Promotes Stem Cell-like Phenotype in Multiple Myeloma Cells. Blood Cancer J. 2014, 4, e262. [Google Scholar] [CrossRef] [Green Version]
  29. Zhang, Q.; Han, Z.; Zhu, Y.; Chen, J.; Li, W. Role of Hypoxia Inducible Factor-1 in Cancer Stem Cells (Review). Mol. Med. Rep. 2020, 23, 17. [Google Scholar] [CrossRef]
  30. Lv, Y.; Zhao, S.; Han, J.; Zheng, L.; Yang, Z.; Zhao, L. Hypoxia-Inducible Factor-1α Induces Multidrug Resistance Protein in Colon Cancer. Oncotargets Ther. 2015, 8, 1941–1948. [Google Scholar] [CrossRef] [Green Version]
  31. Li, D.-W.; Dong, P.; Wang, F.; Chen, X.-W.; Xu, C.-Z.; Zhou, L. Hypoxia Induced Multidrug Resistance of Laryngeal Cancer Cells via Hypoxia-Inducible Factor-1α. Asian Pac. J. Cancer Prev. 2013, 14, 4853–4858. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Li, J.; Wu, X.; Gan, L.; Yang, X.; Miao, Z. Hypoxia Induces Universal but Differential Drug Resistance and Impairs Anticancer Mechanisms of 5-Fluorouracil in Hepatoma Cells. Acta Pharmacol. Sin. 2017, 38, 1642–1654. [Google Scholar] [CrossRef] [PubMed]
  33. Filatova, A.; Seidel, S.; Böğürcü, N.; Gräf, S.; Garvalov, B.K.; Acker, T. Acidosis Acts through HSP90 in a PHD/VHL-Independent Manner to Promote HIF Function and Stem Cell Maintenance in Glioma. Cancer Res. 2016, 76, 5845–5856. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Wojtkowiak, J.W.; Verduzco, D.; Schramm, K.J.; Gillies, R.J. Drug Resistance and Cellular Adaptation to Tumour Acidic PH Microenvironment. Mol. Pharm. 2011, 8, 2032–2038. [Google Scholar] [CrossRef] [PubMed]
  35. Lewis, C.E.; Pollard, J.W. Distinct Role of Macrophages in Different Tumour Microenvironments. Cancer Res. 2006, 66, 605–612. [Google Scholar] [CrossRef] [Green Version]
  36. Pilon-Thomas, S.; Kodumudi, K.N.; El-Kenawi, A.E.; Russell, S.; Weber, A.M.; Luddy, K.; Damaghi, M.; Wojtkowiak, J.W.; Mulé, J.J.; Ibrahim-Hashim, A.; et al. Neutralization of Tumour Acidity Improves Antitumour Responses to Immunotherapy. Cancer Res. 2016, 76, 1381–1390. [Google Scholar] [CrossRef] [Green Version]
  37. Brand, A.; Singer, K.; Koehl, G.E.; Kolitzus, M.; Schoenhammer, G.; Thiel, A.; Matos, C.; Bruss, C.; Klobuch, S.; Peter, K.; et al. LDHA-Associated Lactic Acid Production Blunts Tumour Immunosurveillance by T and NK Cells. Cell Metab. 2016, 24, 657–671. [Google Scholar] [CrossRef] [Green Version]
  38. El-Kenawi, A.; Gatenbee, C.; Robertson-Tessi, M.; Bravo, R.; Dhillon, J.; Balagurunathan, Y.; Berglund, A.; Vishvakarma, N.; Ibrahim-Hashim, A.; Choi, J.; et al. Acidity Promotes Tumour Progression by Altering Macrophage Phenotype in Prostate Cancer. Br. J. Cancer 2019, 121, 556–566. [Google Scholar] [CrossRef] [Green Version]
  39. Díaz, F.E.; Dantas, E.; Cabrera, M.; Benítez, C.A.; Delpino, M.V.; Duette, G.; Rubione, J.; Sanjuan, N.; Trevani, A.S.; Geffner, J. Fever-Range Hyperthermia Improves the Anti-Apoptotic Effect Induced by Low PH on Human Neutrophils Promoting a Proangiogenic Profile. Cell Death Dis. 2016, 7, e2437. [Google Scholar] [CrossRef]
  40. Corbet, C.; Feron, O. Tumour Acidosis: From the Passenger to the Driver’s Seat. Nat. Rev. Cancer 2017, 17, 577–593. [Google Scholar] [CrossRef]
  41. Estrella, V.; Chen, T.; Lloyd, M.; Wojtkowiak, J.; Cornnell, H.H.; Ibrahim-Hashim, A.; Bailey, K.; Balagurunathan, Y.; Rothberg, J.M.; Sloane, B.F.; et al. Acidity Generated by the Tumour Microenvironment Drives Local Invasion. Cancer Res. 2013, 73, 1524–1535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Rofstad, E.K.; Mathiesen, B.; Kindem, K.; Galappathi, K. Acidic Extracellular PH Promotes Experimental Metastasis of Human Melanoma Cells in Athymic Nude Mice. Cancer Res. 2006, 66, 6699–6707. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. 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, 1600253. [Google Scholar] [CrossRef] [PubMed]
  44. 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. [Google Scholar] [CrossRef] [PubMed]
  45. Islam, M.S. Calcium Signaling. In Advances in Experimental Medicine and Biology; Springer International Publishing: Cham, Switzerland, 2020; Volume 1131, ISBN 978-3-030-12456-4. [Google Scholar]
  46. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium Signalling: Dynamics, Homeostasis and Remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [Green Version]
  47. Clapham, D.E. Calcium Signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [Green Version]
  48. Monteith, G.R.; Prevarskaya, N.; Roberts-Thomson, S.J. The Calcium–Cancer Signalling Nexus. Nat. Rev. Cancer 2017, 17, 373–380. [Google Scholar] [CrossRef] [Green Version]
  49. Prevarskaya, N.; Ouadid-Ahidouch, H.; Skryma, R.; Shuba, Y. Remodelling of Ca 2+ Transport in Cancer: How It Contributes to Cancer Hallmarks? Philos. Trans. R. Soc. B 2014, 369, 20130097. [Google Scholar] [CrossRef] [Green Version]
  50. Panda, S.; Chatterjee, O.; Roy, L.; Chatterjee, S. Targeting Ca2+ Signaling: A New Arsenal against Cancer. Drug Discov. Today 2022, 27, 923–934. [Google Scholar] [CrossRef]
  51. Negri, S.; Faris, P.; Berra-Romani, R.; Guerra, G.; Moccia, F. Endothelial Transient Receptor Potential Channels and Vascular Remodeling: Extracellular Ca2+ Entry for Angiogenesis, Arteriogenesis and Vasculogenesis. Front. Physiol. 2020, 10, 1618. [Google Scholar] [CrossRef] [Green Version]
  52. Scarpellino, G.; Munaron, L.; Cantelmo, A.R.; Fiorio Pla, A. Calcium-Permeable Channels in Tumour Vascularization: Peculiar Sensors of Microenvironmental Chemical and Physical Cues. In From Malignant Transformation to Metastasis; Stock, C., Pardo, L.A., Eds.; Reviews of Physiology, Biochemistry and Pharmacology; Springer International Publishing: Cham, Switzerland, 2020; Volume 182, pp. 111–137. ISBN 978-3-030-99799-1. [Google Scholar]
  53. Catacuzzeno, L.; Sforna, L.; Esposito, V.; Limatola, C.; Franciolini, F. Ion Channels in Glioma Malignancy. In Transportome Malfunction in the Cancer Spectrum; Stock, C., Pardo, L.A., Eds.; Reviews of Physiology, Biochemistry and Pharmacology; Springer International Publishing: Cham, Switzerland, 2020; Volume 181, pp. 223–267. ISBN 978-3-030-90919-2. [Google Scholar]
  54. Shapovalov, G.; Ritaine, A.; Skryma, R.; Prevarskaya, N. Role of TRP Ion Channels in Cancer and Tumourigenesis. Semin. Immunopathol. 2016, 38, 357–369. [Google Scholar] [CrossRef] [PubMed]
  55. Seta, K.A.; Yuan, Y.; Spicer, Z.; Lu, G.; Bedard, J.; Ferguson, T.K.; Pathrose, P.; Cole-Strauss, A.; Kaufhold, A.; Millhorn, D.E. The Role of Calcium in Hypoxia-Induced Signal Transduction and Gene Expression. Cell Calcium 2004, 36, 331–340. [Google Scholar] [CrossRef] [PubMed]
  56. Tajada, S. Calcium Permeable Channels in Cancer Hallmarks. Front. Pharmacol. 2020, 11, 19. [Google Scholar] [CrossRef] [PubMed]
  57. Glitsch, M. Protons and Ca2+: Ionic Allies in Tumour Progression? Physiology 2011, 26, 252–265. [Google Scholar] [CrossRef]
  58. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion Channels in Cancer: Are Cancer Hallmarks Oncochannelopathies? Physiol. Rev. 2018, 98, 559–621. [Google Scholar] [CrossRef] [Green Version]
  59. Girault, A.; Ahidouch, A.; Ouadid-Ahidouch, H. Roles for Ca2+ and K+ Channels in Cancer Cells Exposed to the Hypoxic Tumour Microenvironment. Biochim. Biophys. Acta BBA Mol. Cell Res. 2020, 1867, 118644. [Google Scholar] [CrossRef]
  60. Shimoda, L.A.; Polak, J. Hypoxia. 4. Hypoxia and Ion Channel Function. Am. J. Physiol. Cell Physiol. 2011, 300, C951–C967. [Google Scholar] [CrossRef]
  61. Glitsch, M. Mechano- and PH-Sensing Convergence on Ca2+-Mobilising Proteins—A Recipe for Cancer? Cell Calcium 2019, 80, 38–45. [Google Scholar] [CrossRef]
  62. 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] [Green Version]
  63. 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] [Green Version]
  64. Starkus, J.G.; Fleig, A.; Penner, R. The Calcium-Permeable Non-Selective Cation Channel TRPM2 Is Modulated by Cellular Acidification: Proton-Induced Inhibition of TRPM2. J. Physiol. 2010, 588, 1227–1240. [Google Scholar] [CrossRef] [PubMed]
  65. 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] [PubMed] [Green Version]
  66. 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]
  67. Zou, J.; Yang, W.; Beech, D.J.; Jiang, L.-H. A Residue in the TRPM2 Channel Outer Pore Is Crucial in Determining Species-Dependent Sensitivity to Extracellular Acidic PH. Pflug. Arch. Eur. J. Physiol. 2011, 462, 293–302. [Google Scholar] [CrossRef]
  68. Gwanyanya, A.; Amuzescu, B.; Zakharov, S.I.; Macianskiene, R.; Sipido, K.R.; Bolotina, V.M.; Vereecke, J.; Mubagwa, K. Magnesium-Inhibited, TRPM6/7-like Channel in Cardiac Myocytes: Permeation of Divalent Cations and PH-Mediated Regulation: Cardiac TRPM6/7-like Channels. J. Physiol. 2004, 559, 761–776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Li, M.; Jiang, J.; Yue, L. Functional Characterization of Homo- and Heteromeric Channel Kinases TRPM6 and TRPM7. J. Gen. Physiol. 2006, 127, 525–537. [Google Scholar] [CrossRef] [Green Version]
  70. Kozak, J.A.; Matsushita, M.; Nairn, A.C.; Cahalan, M.D. Charge Screening by Internal PH and Polyvalent Cations as a Mechanism for Activation, Inhibition, and Rundown of TRPM7/MIC Channels. J. Gen. Physiol. 2005, 126, 499–514. [Google Scholar] [CrossRef] [Green Version]
  71. 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]
  72. Chu, X.; Zhu, X.; Wei, W.; Li, G.; Simon, R.P.; MacDonald, J.F.; Xiong, Z. Acidosis Decreases Low Ca2+-induced Neuronal Excitation by Inhibiting the Activity of Calcium-sensing Cation Channels in Cultured Mouse Hippocampal Neurons. J. Physiol. 2003, 550, 385–399. [Google Scholar] [CrossRef]
  73. 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]
  74. Numata, T.; Sato-Numata, K.; Okada, Y. TRPM7 Is Involved in Acid-Induced Necrotic Cell Death in a Manner Sensitive to Progesterone in Human Cervical Cancer Cells: Role of TRPM7 in Cervical Cancer Cell Death. Physiol. Rep. 2019, 7, e14157. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Mačianskienė, R.; Almanaitytė, 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]
  76. Heber, S.; Ciotu, C.I.; Hartner, G.; Gold-Binder, M.; Ninidze, N.; Gleiss, A.; Kress, H.-G.; Fischer, M.J.M. TRPV1 Antagonist BCTC Inhibits PH 6.0-Induced Pain in Human Skin. Pain 2020, 161, 1532–1541. [Google Scholar] [CrossRef] [PubMed]
  77. Ryu, S.; Liu, B.; Yao, J.; Fu, Q.; Qin, F. Uncoupling Proton Activation of Vanilloid Receptor TRPV1. J. Neurosci. 2007, 27, 12797–12807. [Google Scholar] [CrossRef]
  78. 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]
  79. Nakanishi, M.; Morita, Y.; Hata, K.; Muragaki, Y. Acidic Microenvironments Induce Lymphangiogenesis and IL-8 Production via TRPV1 Activation in Human Lymphatic Endothelial Cells. Exp. Cell Res. 2016, 345, 180–189. [Google Scholar] [CrossRef]
  80. Gao, L.; Yang, P.; Qin, P.; Lu, Y.; Li, X.; Tian, Q.; Li, Y.; Xie, C.; Tian, J.; Zhang, C.; et al. Selective Potentiation of 2-APB-Induced Activation of TRPV1–3 Channels by Acid. Sci. Rep. 2016, 6, 20791. [Google Scholar] [CrossRef] [Green Version]
  81. Cao, X.; Yang, F.; Zheng, J.; Wang, K. Intracellular Proton-Mediated Activation of TRPV3 Channels Accounts for the Exfoliation Effect of α-Hydroxyl Acids on Keratinocytes. J. Biol. Chem. 2012, 287, 25905–25916. [Google Scholar] [CrossRef] [Green Version]
  82. 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]
  83. Shikano, M.; Ueda, T.; Kamiya, T.; Ishida, Y.; Yamada, T.; Mizushima, T.; Shimura, T.; Mizoshita, T.; Tanida, S.; Kataoka, H.; et al. Acid Inhibits TRPV4-Mediated Ca2+ Influx in Mouse Esophageal Epithelial Cells: Acid-Sensitive TRPV4 in the Esophageal Epithelium. Neurogastroenterol. Motil. 2011, 23, 1020-e497. [Google Scholar] [CrossRef]
  84. Cherezova, A.L.; Negulyaev, J.A.; Zenin, V.V.; Semenova, S.B. Extracellular PH Regulates the Entry of Calcium into Jurkat T-Cells. Cell Tiss. Biol. 2018, 12, 41–47. [Google Scholar] [CrossRef]
  85. De la Roche, J.; Eberhardt, M.J.; Klinger, A.B.; Stanslowsky, N.; Wegner, F.; Koppert, W.; Reeh, P.W.; Lampert, A.; Fischer, M.J.M.; Leffler, A. The Molecular Basis for Species-Specific Activation of Human TRPA1 Protein by Protons Involves Poorly Conserved Residues within Transmembrane Domains 5 and 6. J. Biol. Chem. 2013, 288, 20280–20292. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. 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]
  87. Thakur, D.P.; Wang, Q.; Jeon, J.; Tian, J.; Zhu, M.X. Intracellular Acidification Facilitates Receptor-operated TRPC4 Activation through PLCδ1 in a Ca2+-dependent Manner. J. Physiol. 2020, 598, 2651–2667. [Google Scholar] [CrossRef]
  88. 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, 101–111. [Google Scholar] [CrossRef]
  89. Scrimgeour, N.R.; Wilson, D.P.; Rychkov, G.Y. Glu106 in the Orai1 Pore Contributes to Fast Ca2+-Dependent Inactivation and PH Dependence of Ca2+ Release-Activated Ca2+ (CRAC) Current. Biochem. J. 2012, 441, 743–753. [Google Scholar] [CrossRef]
  90. Beck, A.; Fleig, A.; Penner, R.; Peinelt, C. Regulation of Endogenous and Heterologous Ca2+ Release-Activated Ca2+ Currents by PH. Cell Calcium 2014, 56, 235–243. [Google Scholar] [CrossRef]
  91. 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]
  92. Mancarella, S.; Wang, Y.; Deng, X.; Landesberg, G.; Scalia, R.; Panettieri, R.A.; Mallilankaraman, K.; Tang, X.D.; Madesh, M.; Gill, D.L. Hypoxia-Induced Acidosis Uncouples the STIM-Orai Calcium Signaling Complex. J. Biol. Chem. 2011, 286, 44788–44798. [Google Scholar] [CrossRef] [Green Version]
  93. Laskay, G.; Kálmán, K.; Van Kerkhove, E.; Steels, P.; Ameloot, M. Store-Operated Ca2+-Channels Are Sensitive to Changes in Extracellular PH. Biochem. Biophys. Res. Commun. 2005, 337, 571–579. [Google Scholar] [CrossRef]
  94. Vandorpe, D.H.; Xu, C.; Shmukler, B.E.; Otterbein, L.E.; Trudel, M.; Sachs, F.; Gottlieb, P.A.; Brugnara, C.; Alper, S.L. Hypoxia Activates a Ca2+-Permeable Cation Conductance Sensitive to Carbon Monoxide and to GsMTx-4 in Human and Mouse Sickle Erythrocytes. PLoS ONE 2010, 5, e8732. [Google Scholar] [CrossRef] [PubMed]
  95. Liao, J.; Lu, W.; Chen, Y.; Duan, X.; Zhang, C.; Luo, X.; Lin, Z.; Chen, J.; Liu, S.; Yan, H.; et al. Upregulation of Piezo1 (Piezo Type Mechanosensitive Ion Channel Component 1) Enhances the Intracellular Free Calcium in Pulmonary Arterial Smooth Muscle Cells from Idiopathic Pulmonary Arterial Hypertension Patients. Hypertension 2021, 77, 1974–1989. [Google Scholar] [CrossRef] [PubMed]
  96. Wang, Z.; Chen, J.; Babicheva, A.; Jain, P.P.; Rodriguez, M.; Ayon, R.J.; Ravellette, K.S.; Wu, L.; Balistrieri, F.; Tang, H.; et al. Endothelial Upregulation of Mechanosensitive Channel Piezo1 in Pulmonary Hypertension. Am. J. Physiol. Cell Physiol. 2021, 321, C1010–C1027. [Google Scholar] [CrossRef] [PubMed]
  97. Huang, S.; Turlova, E.; Li, F.; Bao, M.; Szeto, V.; Wong, R.; Abussaud, A.; Wang, H.; Zhu, S.; Gao, X.; et al. Transient Receptor Potential Melastatin 2 Channels (TRPM2) Mediate Neonatal Hypoxic-Ischemic Brain Injury in Mice. Exp. Neurol. 2017, 296, 32–40. [Google Scholar] [CrossRef]
  98. Kaneko, S.; Kawakami, S.; Hara, Y.; Wakamori, M.; Itoh, E.; Minami, T.; Takada, Y.; Kume, T.; Katsuki, H.; Mori, Y.; et al. A Critical Role of TRPM2 in Neuronal Cell Death by Hydrogen Peroxide. J. Pharmacol. Sci. 2006, 101, 66–76. [Google Scholar] [CrossRef] [Green Version]
  99. Duzgun Ergun, D.; Dursun, S.; Pastaci Ozsobaci, N.; Hatırnaz Ng, O.; Naziroglu, M.; Ozcelik, D. The Potential Protective Roles of Zinc, Selenium and Glutathione on Hypoxia-Induced TRPM2 Channel Activation in Transfected HEK293 Cells. J. Recept. Signal Transduct. 2020, 40, 521–530. [Google Scholar] [CrossRef]
  100. Özkaya, D.; Nazıroğlu, M.; Vanyorek, L.; Muhamad, S. Involvement of TRPM2 Channel on Hypoxia-Induced Oxidative Injury, Inflammation, and Cell Death in Retinal Pigment Epithelial Cells: Modulator Action of Selenium Nanoparticles. Biol. Trace Elem. Res. 2021, 199, 1356–1369. [Google Scholar] [CrossRef]
  101. Ferrera, L.; Barbieri, R.; Picco, C.; Zuccolini, P.; Remigante, A.; Bertelli, S.; Fumagalli, M.R.; Zifarelli, G.; La Porta, C.A.M.; Gavazzo, P.; et al. TRPM2 Oxidation Activates Two Distinct Potassium Channels in Melanoma Cells through Intracellular Calcium Increase. Int. J. Mol. Sci. 2021, 22, 8359. [Google Scholar] [CrossRef]
  102. Wang, Q.; Huang, L.; Yue, J. Oxidative Stress Activates the TRPM2-Ca2+-CaMKII-ROS Signaling Loop to Induce Cell Death in Cancer Cells. Biochim. Biophys. Acta BBA Mol. Cell Res. 2017, 1864, 957–967. [Google Scholar] [CrossRef]
  103. Chen, S.; Hoffman, N.E.; Shanmughapriya, S.; Bao, L.; Keefer, K.; Conrad, K.; Merali, S.; Takahashi, Y.; Abraham, T.; Hirschler-Laszkiewicz, I.; et al. A Splice Variant of the Human Ion Channel TRPM2 Modulates Neuroblastoma Tumour Growth through Hypoxia-Inducible Factor (HIF)-1/2α. J. Biol. Chem. 2014, 289, 36284–36302. [Google Scholar] [CrossRef] [Green Version]
  104. 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 Tumour Cells. Cancer Res. 2018, 78, 2680–2690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Bilecik, T.; Karateke, F.; Elkan, H.; Gokce, H. The Effects of TRPM2, TRPM6, TRPM7 and TRPM8 Gene Expression in Hepatic Ischemia Reperfusion Injury. Eur. Rev. Med. Pharmacol. Sci. 2019, 23, 3088–3095. [Google Scholar] [PubMed]
  106. Aarts, M.; Iihara, K.; Wei, W.-L.; Xiong, Z.-G.; Arundine, M.; Cerwinski, W.; MacDonald, J.F.; Tymianski, M. A Key Role for TRPM7 Channels in Anoxic Neuronal Death. Cell 2003, 115, 863–877. [Google Scholar] [CrossRef] [Green Version]
  107. Kim, K.S.; Yoo, H.Y.; Park, K.S.; Kim, J.K.; Zhang, Y.-H.; Kim, S.J. Differential Effects of Acute Hypoxia on the Activation of TRPV1 by Capsaicin and Acidic PH. J. Physiol. Sci. 2012, 62, 93–103. [Google Scholar] [CrossRef] [PubMed]
  108. Ristoiu, V.; Shibasaki, K.; Uchida, K.; Zhou, Y.; Ton, B.-H.T.; Flonta, M.-L.; Tominaga, M. Hypoxia-Induced Sensitization of Transient Receptor Potential Vanilloid 1 Involves Activation of Hypoxia-Inducible Factor-1 Alpha and PKC. Pain 2011, 152, 936–945. [Google Scholar] [CrossRef] [PubMed]
  109. Parpaite, T.; Cardouat, G.; Mauroux, M.; Gillibert-Duplantier, J.; Robillard, P.; Quignard, J.-F.; Marthan, R.; Savineau, J.-P.; Ducret, T. Effect of Hypoxia on TRPV1 and TRPV4 Channels in Rat Pulmonary Arterial Smooth Muscle Cells. Pflug. Arch. Eur. J. Physiol. 2016, 468, 111–130. [Google Scholar] [CrossRef]
  110. Wang, Y.X.; Wang, J.; Wang, C.; Liu, J.; Shi, L.P.; Xu, M.; Wang, C. Functional Expression of Transient Receptor Potential Vanilloid-Related Channels in Chronically Hypoxic Human Pulmonary Arterial Smooth Muscle Cells. J. Membr. Biol. 2008, 223, 151–159. [Google Scholar] [CrossRef]
  111. Ma, W.; Li, C.; Yin, S.; Liu, J.; Gao, C.; Lin, Z.; Huang, R.; Huang, J.; Li, Z. Novel Role of TRPV2 in Promoting the Cytotoxicity of H2O2-Mediated Oxidative Stress in Human Hepatoma Cells. Free. Radic. Biol. Med. 2015, 89, 1003–1013. [Google Scholar] [CrossRef]
  112. Wang, J.; Chen, X.; Huang, W. MicroRNA-369 Attenuates Hypoxia-Induced Cardiomyocyte Apoptosis and Inflammation via Targeting TRPV3. Braz. J. Med. Biol. Res. 2021, 54, e10550. [Google Scholar] [CrossRef]
  113. Zhang, Q.; Cao, Y.; Luo, Q.; Wang, P.; Shi, P.; Song, C.; Mingyao, E.; Ren, J.; Fu, B.; Sun, H. The Transient Receptor Potential Vanilloid-3 Regulates Hypoxia-Mediated Pulmonary Artery Smooth Muscle Cells Proliferation via PI3K/AKT Signaling Pathway. Cell Prolif. 2018, 51, e12436. [Google Scholar] [CrossRef] [Green Version]
  114. Karttunen, S.; Duffield, M.; Scrimgeour, N.R.; Squires, L.; Lim, W.L.; Dallas, M.L.; Scragg, J.L.; Chicher, J.; Dave, K.A.; Whitelaw, M.L.; et al. Oxygen-Dependent Hydroxylation by Factor Inhibiting HIF (FIH) Regulates the TRPV3 Ion Channel. J. Cell Sci. 2015, 128, 225–231. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Wu, Q.-F.; Qian, C.; Zhao, N.; Dong, Q.; Li, J.; Wang, B.-B.; Chen, L.; Yu, L.; Han, B.; Du, Y.-M.; et al. Activation of Transient Receptor Potential Vanilloid 4 Involves in Hypoxia/Reoxygenation Injury in Cardiomyocytes. Cell Death Dis. 2017, 8, e2828. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Butenko, O.; Dzamba, D.; Benesova, J.; Honsa, P.; Benfenati, V.; Rusnakova, V.; Ferroni, S.; Anderova, M. The Increased Activity of TRPV4 Channel in the Astrocytes of the Adult Rat Hippocampus after Cerebral Hypoxia/Ischemia. PLoS ONE 2012, 7, e39959. [Google Scholar] [CrossRef] [Green Version]
  117. Takahashi, N.; Chen, H.-Y.; Harris, I.S.; Stover, D.G.; Selfors, L.M.; Bronson, R.T.; Deraedt, T.; Cichowski, K.; Welm, A.L.; Mori, Y.; et al. Cancer Cells Co-Opt the Neuronal Redox-Sensing Channel TRPA1 to Promote Oxidative-Stress Tolerance. Cancer Cell 2018, 33, 985–1003.e7. [Google Scholar] [CrossRef] [Green Version]
  118. Hamilton, N.B.; Kolodziejczyk, K.; Kougioumtzidou, E.; Attwell, D. Proton-Gated Ca2+-Permeable TRP Channels Damage Myelin in Conditions Mimicking Ischaemia. Nature 2016, 529, 523–527. [Google Scholar] [CrossRef] [Green Version]
  119. Wang, B.; Li, W.; Meng, X.; Zou, F. Hypoxia Up-Regulates Vascular Endothelial Growth Factor in U-87 MG Cells: Involvement of TRPC1. Neurosci. Lett. 2009, 459, 132–136. [Google Scholar] [CrossRef] [PubMed]
  120. Azimi, I.; Milevskiy, M.J.G.; Kaemmerer, E.; Turner, D.; Yapa, K.T.D.S.; Brown, M.A.; Thompson, E.W.; Roberts-Thomson, S.J.; Monteith, G.R. TRPC1 Is a Differential Regulator of Hypoxia-Mediated Events and Akt Signalling in PTEN-Deficient Breast Cancer Cells. J. Cell Sci. 2017, 130, 2292–2305. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Zhu, Y.; Pan, Q.; Meng, H.; Jiang, Y.; Mao, A.; Wang, T.; Hua, D.; Yao, X.; Jin, J.; Ma, X. Enhancement of Vascular Endothelial Growth Factor Release in Long-Term Drug-Treated Breast Cancer via Transient Receptor Potential Channel 5-Ca2+-Hypoxia-Inducible Factor 1α Pathway. Pharmacol. Res. 2015, 93, 36–42. [Google Scholar] [CrossRef] [PubMed]
  122. Chen, Z.; Zhu, Y.; Dong, Y.; Zhang, P.; Han, X.; Jin, J.; Ma, X. Overexpression of TrpC5 Promotes Tumour Metastasis via the HIF-1α–Twist Signaling Pathway in Colon Cancer. Clin. Sci. 2017, 131, 2439–2450. [Google Scholar] [CrossRef]
  123. 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]
  124. Iyer, S.C.; Kannan, A.; Gopal, A.; Devaraj, N.; Halagowder, D. Receptor Channel TRPC6 Orchestrate the Activation of Human Hepatic Stellate Cell under Hypoxia Condition. Exp. Cell Res. 2015, 336, 66–75. [Google Scholar] [CrossRef] [PubMed]
  125. Wen, L.; Liang, C.; Chen, E.; Chen, W.; Liang, F.; Zhi, X.; Wei, T.; Xue, F.; Li, G.; Yang, Q.; et al. Regulation of Multi-Drug Resistance in Hepatocellular Carcinoma Cells Is TRPC6/Calcium Dependent. Sci. Rep. 2016, 6, 23269. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. 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] [PubMed] [Green Version]
  127. Wang, Y.; He, J.; Jiang, H.; Zhang, Q.; Yang, H.; Xu, X.; Zhang, C.; Xu, C.; Wang, J.; Lu, W. Nicotine Enhances Store-operated Calcium Entry by Upregulating HIF-1α and SOCC Components in Non-small Cell Lung Cancer Cells. Oncol. Rep. 2018. [Google Scholar] [CrossRef] [Green Version]
  128. Liu, X.; Wang, T.; Wang, Y.; Chen, Z.; Hua, D.; Yao, X.; Ma, X.; Zhang, P. Orai1 Is Critical for Notch-Driven Aggressiveness under Hypoxic Conditions in Triple-Negative Breast Cancers. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2018, 1864, 975–986. [Google Scholar] [CrossRef]
  129. Liu, X.; Wan, X.; Kan, H.; Wang, Y.; Yu, F.; Feng, L.; Jin, J.; Zhang, P.; Ma, X. Hypoxia-Induced Upregulation of Orai1 Drives Colon Cancer Invasiveness and Angiogenesis. Eur. J. Pharmacol. 2018, 832, 1–10. [Google Scholar] [CrossRef]
  130. 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]
  131. Parpaite, T.; Coste, B. Piezo Channels. Curr. Biol. 2017, 27, R250–R252. [Google Scholar] [CrossRef] [Green Version]
  132. De Felice, D.; Alaimo, A. Mechanosensitive Piezo Channels in Cancer: Focus on Altered Calcium Signaling in Cancer Cells and in Tumour Progression. Cancers 2020, 12, 1780. [Google Scholar] [CrossRef]
  133. Pardo-Pastor, C.; Rubio-Moscardo, F.; Vogel-González, M.; Serra, S.A.; Afthinos, A.; Mrkonjic, S.; Destaing, O.; Abenza, J.F.; Fernández-Fernández, J.M.; Trepat, X.; et al. Piezo2 Channel Regulates RhoA and Actin Cytoskeleton to Promote Cell Mechanobiological Responses. Proc. Natl. Acad. Sci. USA 2018, 115, 1925–1930. [Google Scholar] [CrossRef] [Green Version]
  134. 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.e7. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Yang, H.; Liu, C.; Zhou, R.-M.; Yao, J.; Li, X.-M.; Shen, Y.; Cheng, H.; Yuan, J.; Yan, B.; Jiang, Q. Piezo2 Protein: A Novel Regulator of Tumour Angiogenesis and Hyperpermeability. Oncotarget 2016, 7, 44630–44643. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Zhang, J.; Zhou, Y.; Huang, T.; Wu, F.; Liu, L.; Kwan, J.S.H.; Cheng, A.S.L.; Yu, J.; To, K.F.; Kang, W. PIEZO1 Functions as a Potential Oncogene by Promoting Cell Proliferation and Migration in Gastric Carcinogenesis. Mol. Carcinog. 2018, 57, 1144–1155. [Google Scholar] [CrossRef] [PubMed]
  137. Dombroski, J.A.; Hope, J.M.; Sarna, N.S.; King, M.R. Channeling the Force: Piezo1 Mechanotransduction in Cancer Metastasis. Cells 2021, 10, 2815. [Google Scholar] [CrossRef] [PubMed]
  138. 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]
  139. Hertz, L.; Huisjes, R.; Llaudet-Planas, E.; Petkova-Kirova, P.; Makhro, A.; Danielczok, J.G.; Egee, S.; del Mar Mañú-Pereira, M.; van Wijk, R.; Vives Corrons, J.-L.; et al. Is Increased Intracellular Calcium in Red Blood Cells a Common Component in the Molecular Mechanism Causing Anemia? Front. Physiol. 2017, 8, 673. [Google Scholar] [CrossRef] [Green Version]
  140. Wulff, H.; Castle, N.A. Therapeutic Potential of KCa3.1 Blockers: Recent Advances and Promising Trends. Expert Rev. Clin. Pharmacol. 2010, 3, 385–396. [Google Scholar] [CrossRef] [Green Version]
  141. Hara, Y.; Wakamori, M.; Ishii, M.; Maeno, E.; Nishida, M.; Yoshida, T.; Yamada, H.; Shimizu, S.; Mori, E.; Kudoh, J.; et al. LTRPC2 Ca2+-Permeable Channel Activated by Changes in Redox Status Confers Susceptibility to Cell Death. Mol. Cell 2002, 9, 163–173. [Google Scholar] [CrossRef]
  142. Kolisek, M.; Beck, A.; Fleig, A.; Penner, R. Cyclic ADP-Ribose and Hydrogen Peroxide Synergize with ADP-Ribose in the Activation of TRPM2 Channels. Mol. Cell 2005, 18, 61–69. [Google Scholar] [CrossRef]
  143. McHugh, D.; Flemming, R.; Xu, S.-Z.; Perraud, A.-L.; Beech, D.J. Critical Intracellular Ca2+ Dependence of Transient Receptor Potential Melastatin 2 (TRPM2) Cation Channel Activation. J. Biol. Chem. 2003, 278, 11002–11006. [Google Scholar] [CrossRef] [Green Version]
  144. Perraud, A.-L.; Takanishi, C.L.; Shen, B.; Kang, S.; Smith, M.K.; Schmitz, C.; Knowles, H.M.; Ferraris, D.; Li, W.; Zhang, J.; et al. Accumulation of Free ADP-Ribose from Mitochondria Mediates Oxidative Stress-Induced Gating of TRPM2 Cation Channels. J. Biol. Chem. 2005, 280, 6138–6148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Turlova, E.; Feng, Z.; Sun, H. The Role of TRPM2 Channels in Neurons, Glial Cells and the Blood-Brain Barrier in Cerebral Ischemia and Hypoxia. Acta Pharmacol. Sin. 2018, 39, 713–721. [Google Scholar] [CrossRef] [PubMed]
  146. Mai, C.; Mankoo, H.; Wei, L.; An, X.; Li, C.; Li, D.; Jiang, L. TRPM2 Channel: A Novel Target for Alleviating Ischaemia-reperfusion, Chronic Cerebral Hypo-perfusion and Neonatal Hypoxic-ischaemic Brain Damage. J. Cell. Mol. Med. 2020, 24, 4–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Miller, B.A. TRPM2 in Cancer. Cell Calcium 2019, 80, 8–17. [Google Scholar] [CrossRef]
  148. Zhao, L.-Y.; Xu, W.-L.; Xu, Z.-Q.; Qi, C.; Li, Y.; Cheng, J.; Liu, L.-K.; Wu, Y.-N.; Gao, J.; Ye, J.-H. The Overexpressed Functional Transient Receptor Potential Channel TRPM2 in Oral Squamous Cell Carcinoma. Sci. Rep. 2016, 6, 38471. [Google Scholar] [CrossRef] [Green Version]
  149. 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] [Green Version]
  150. Hopkins, M.M.; Feng, X.; Liu, M.; Parker, L.P.; Koh, D.W. Inhibition of the Transient Receptor Potential Melastatin-2 Channel Causes Increased DNA Damage and Decreased Proliferation in Breast Adenocarcinoma Cells. Int. J. Oncol. 2015, 46, 2267–2276. [Google Scholar] [CrossRef] [Green Version]
  151. Almasi, S.; Sterea, A.M.; Fernando, W.; Clements, D.R.; Marcato, P.; Hoskin, D.W.; Gujar, S.; El Hiani, Y. TRPM2 Ion Channel Promotes Gastric Cancer Migration, Invasion and Tumour Growth through the AKT Signaling Pathway. Sci. Rep. 2019, 9, 4182. [Google Scholar] [CrossRef] [Green Version]
  152. Almasi, S.; Kennedy, B.E.; El-Aghil, M.; Sterea, A.M.; Gujar, S.; Partida-Sánchez, S.; El Hiani, Y. TRPM2 Channel–Mediated Regulation of Autophagy Maintains Mitochondrial Function and Promotes Gastric Cancer Cell Survival via the JNK-Signaling Pathway. J. Biol. Chem. 2018, 293, 3637–3650. [Google Scholar] [CrossRef] [Green Version]
  153. Guler, Y.; Ovey, I.S. Synergic and Comparative Effect of 5-Fluorouracil and Leucoverin on Breast and Colon Cancer Cells through TRPM2 Channels. Bratisl. Lek. Listy 2018, 119, 692–700. [Google Scholar] [CrossRef]
  154. Lange, I.; Yamamoto, S.; Partida-Sanchez, S.; Mori, Y.; Fleig, A.; Penner, R. TRPM2 Functions as a Lysosomal Ca2+-Release Channel in Cells. Sci. Signal. 2009, 2, ra23. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Hantute-Ghesquier, A.; Haustrate, A.; Prevarskaya, N.; Lehen’kyi, V. TRPM Family Channels in Cancer. Pharmaceuticals 2018, 11, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Zhang, J.; Zhao, F.; Zhao, Y.; Wang, J.; Pei, L.; Sun, N.; Shi, J. Hypoxia Induces an Increase in Intracellular Magnesium via Transient Receptor Potential Melastatin 7 (TRPM7) Channels in Rat Hippocampal Neurons In Vitro. J. Biol. Chem. 2011, 286, 20194–20207. [Google Scholar] [CrossRef] [Green Version]
  157. Mori, Y.; Takahashi, N.; Polat, O.K.; Kurokawa, T.; Takeda, N.; Inoue, M. Redox-Sensitive Transient Receptor Potential Channels in Oxygen Sensing and Adaptation. Pflug. Arch. Eur. J. Physiol. 2016, 468, 85–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Caterina, M.J.; Leffler, A.; Malmberg, A.B.; Martin, W.J.; Trafton, J.; Petersen-Zeitz, K.R.; Koltzenburg, M.; Basbaum, A.I.; Julius, D. Impaired Nociception and Pain Sensation in Mice Lacking the Capsaicin Receptor. Science 2000, 288, 306–313. [Google Scholar] [CrossRef] [PubMed]
  159. Reeh, P.W.; Fischer, M.J.M. Nobel Somatosensations and Pain. Pflug. Arch. Eur. J. Physiol. 2022, 474, 405–420. [Google Scholar] [CrossRef] [PubMed]
  160. Storozhuk, M.V.; Moroz, O.F.; Zholos, A.V. Multifunctional TRPV1 Ion Channels in Physiology and Pathology with Focus on the Brain, Vasculature, and Some Visceral Systems. BioMed Res. Int. 2019, 2019, 5806321. [Google Scholar] [CrossRef] [Green Version]
  161. Hellwig, N.; Plant, T.D.; Janson, W.; Schäfer, 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]
  162. Li, L.; Chen, C.; Chiang, C.; Xiao, T.; Chen, Y.; Zhao, Y.; Zheng, D. The Impact of TRPV1 on Cancer Pathogenesis and Therapy: A Systematic Review. Int. J. Biol. Sci. 2021, 17, 2034–2049. [Google Scholar] [CrossRef]
  163. Siveen, K.S.; Nizamuddin, P.B.; Uddin, S.; Al-Thani, M.; Frenneaux, M.P.; Janahi, I.A.; Steinhoff, M.; Azizi, F. TRPV2: A Cancer Biomarker and Potential Therapeutic Target. Dis. Markers 2020, 2020, 8892312. [Google Scholar] [CrossRef]
  164. Li, X.; Zhang, Q.; Fan, K.; Li, B.; Li, H.; Qi, H.; Guo, J.; Cao, Y.; Sun, H. Overexpression of TRPV3 Correlates with Tumour Progression in Non-Small Cell Lung Cancer. Int. J. Mol. Sci. 2016, 17, 437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Wang, H.; Yang, P.; Lu, Y.; Wang, J.; Jeon, J.; Wang, Q.; Tian, J.-B.; Zang, B.; Yu, Y.; Zhu, M.X. Mechanisms of Proton Inhibition and Sensitization of the Cation Channel TRPV3. J. Gen. Physiol. 2021, 153, e202012663. [Google Scholar] [CrossRef] [PubMed]
  166. Lee, W.; Choong, L.; Jin, T.; Mon, N.; Chong, S.; Liew, C.; Putti, T.; Lu, S.; Harteneck, C.; Lim, Y. TRPV4 Plays a Role in Breast Cancer Cell Migration via Ca2+-Dependent Activation of AKT and Downregulation of E-Cadherin Cell Cortex Protein. Oncogenesis 2017, 6, e338. [Google Scholar] [CrossRef] [PubMed]
  167. 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, 27903. [Google Scholar] [CrossRef] [Green Version]
  168. Fiorio Pla, A.; Ong, H.L.; Cheng, K.T.; Brossa, A.; Bussolati, B.; Lockwich, T.; Paria, B.; Munaron, L.; Ambudkar, I.S. TRPV4 Mediates Tumour-Derived Endothelial Cell Migration via Arachidonic Acid-Activated Actin Remodeling. Oncogene 2012, 31, 200–212. [Google Scholar] [CrossRef] [Green Version]
  169. Vellino, S.; Oddou, C.; Rivier, P.; Boyault, C.; Hiriart-Bryant, E.; Kraut, A.; Martin, R.; Coute, Y.; Knölker, H.-J.; Valverde, M.A.; et al. Crosstalk between the Calcium Channel TRPV4 and Reactive Oxygen Species Interlocks Adhesive and Degradative Functions of Invadosomes. J. Cell Biol. 2021, 220, e201910079. [Google Scholar] [CrossRef]
  170. Wang, H.; Zhang, B.; Wang, X.; Mao, J.; Li, W.; Sun, Y.; Yuan, Y.; Ben, Q.; Hua, L.; Qian, A. TRPV4 Overexpression Promotes Metastasis through Epithelial–Mesenchymal Transition in Gastric Cancer and Correlates with Poor Prognosis. Oncotargets Ther. 2020, 13, 8383–8394. [Google Scholar] [CrossRef]
  171. Li, X.; Cheng, Y.; Wang, Z.; Zhou, J.; Jia, Y.; He, X.; Zhao, L.; Dong, Y.; Fan, Y.; Yang, X.; et al. Calcium and TRPV4 Promote Metastasis by Regulating Cytoskeleton through the RhoA/ROCK1 Pathway in Endometrial Cancer. Cell Death Dis. 2020, 11, 1009. [Google Scholar] [CrossRef]
  172. Thoppil, R.J.; Cappelli, H.C.; Adapala, R.K.; Kanugula, A.K.; Paruchuri, S.; Thodeti, C.K. TRPV4 Channels Regulate Tumour Angiogenesis via Modulation of Rho/Rho Kinase Pathway. Oncotarget 2016, 7, 25849–25861. [Google Scholar] [CrossRef] [Green Version]
  173. Azimi, I.; Robitaille, M.; Armitage, K.; So, C.L.; Milevskiy, M.J.G.; Northwood, K.; Lim, H.F.; Thompson, E.W.; Roberts-Thomson, S.J.; Monteith, G.R. Activation of the Ion Channel TRPV4 Induces Epithelial to Mesenchymal Transition in Breast Cancer Cells. Int. J. Mol. Sci. 2020, 21, 9417. [Google Scholar] [CrossRef]
  174. Peters, A.A.; Jamaludin, S.Y.N.; Yapa, K.T.D.S.; Chalmers, S.; Wiegmans, A.P.; Lim, H.F.; Milevskiy, M.J.G.; Azimi, I.; Davis, F.M.; Northwood, K.S.; et al. Oncosis and Apoptosis Induction by Activation of an Overexpressed Ion Channel in Breast Cancer Cells. Oncogene 2017, 36, 6490–6500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Liu, L.; Guo, M.; Lv, X.; Wang, Z.; Yang, J.; Li, Y.; Yu, F.; Wen, X.; Feng, L.; Zhou, T. Role of Transient Receptor Potential Vanilloid 4 in Vascular Function. Front. Mol. Biosci. 2021, 8, 677661. [Google Scholar] [CrossRef] [PubMed]
  176. Stewart, J.M. TRPV6 as A Target for Cancer Therapy. J. Cancer 2020, 11, 374–387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Lehen’kyi, V.; Flourakis, M.; Skryma, R.; Prevarskaya, N. TRPV6 Channel Controls Prostate Cancer Cell Proliferation via Ca2+/NFAT-Dependent Pathways. Oncogene 2007, 26, 7380–7385. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Raphaël, M.; Lehen’kyi, V.; Vandenberghe, M.; Beck, B.; Khalimonchyk, S.; Vanden Abeele, F.; Farsetti, L.; Germain, E.; Bokhobza, A.; Mihalache, A.; et al. TRPV6 Calcium Channel Translocates to the Plasma Membrane via Orai1-Mediated Mechanism and Controls Cancer Cell Survival. Proc. Natl. Acad. Sci. USA 2014, 111, E3870–E3879. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Kärki, T.; Rajakylä, E.K.; Acheva, A.; Tojkander, S. TRPV6 Calcium Channel Directs Homeostasis of the Mammary Epithelial Sheets and Controls Epithelial Mesenchymal Transition. Sci. Rep. 2020, 10, 14683. [Google Scholar] [CrossRef] [PubMed]
  180. Dhennin-Duthille, I.; Gautier, M.; Faouzi, M.; Guilbert, A.; Brevet, M.; Vaudry, D.; Ahidouch, A.; Sevestre, H.; Ouadid-Ahidouch, H. High Expression of Transient Receptor Potential Channels in Human Breast Cancer Epithelial Cells and Tissues: Correlation with Pathological Parameters. Cell. Physiol. Biochem. 2011, 28, 813–822. [Google Scholar] [CrossRef]
  181. Song, H.; Dong, M.; Zhou, J.; Sheng, W.; Li, X.; Gao, W. Expression and Prognostic Significance of TRPV6 in the Development and Progression of Pancreatic Cancer. Oncol. Rep. 2018, 39, 1432–1440. [Google Scholar] [CrossRef] [Green Version]
  182. Xue, H.; Wang, Y.; MacCormack, T.J.; Lutes, T.; Rice, C.; Davey, M.; Dugourd, D.; Ilenchuk, T.T.; Stewart, J.M. Inhibition of Transient Receptor Potential Vanilloid 6 Channel, Elevated in Human Ovarian Cancers, Reduces Tumour Growth in a Xenograft Model. J. Cancer 2018, 9, 3196–3207. [Google Scholar] [CrossRef]
  183. Sutoo, S.; Maeda, T.; Suzuki, A.; Kato, Y. Adaptation to Chronic Acidic Extracellular PH Elicits a Sustained Increase in Lung Cancer Cell Invasion and Metastasis. Clin. Exp. Metastasis 2020, 37, 133–144. [Google Scholar] [CrossRef] [Green Version]
  184. Peleg, S.; Sellin, J.H.; Wang, Y.; Freeman, M.R.; Umar, S. Suppression of Aberrant Transient Receptor Potential Cation Channel, Subfamily V, Member 6 Expression in Hyperproliferative Colonic Crypts by Dietary Calcium. Am. J. Physiol. Gastrointest. Liver Physiol. 2010, 299, G593–G601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Yang, H.; An, B.-S.; Choi, K.-C.; Jeung, E.-B. Change of Genes in Calcium Transport Channels Caused by Hypoxic Stress in the Placenta, Duodenum, and Kidney of Pregnant Rats1. Biol. Reprod. 2013, 88, 1–12. [Google Scholar] [CrossRef] [PubMed]
  186. Park, Y.R.; Chun, J.N.; So, I.; Kim, H.J.; Baek, S.; Jeon, J.-H.; Shin, S.-Y. Data-Driven Analysis of TRP Channels in Cancer: Linking Variation in Gene Expression to Clinical Significance. Cancer Genom. Proteom. 2016, 13, 83–90. [Google Scholar]
  187. Takahashi, N.; Kuwaki, T.; Kiyonaka, S.; Numata, T.; Kozai, D.; Mizuno, Y.; Yamamoto, S.; Naito, S.; Knevels, E.; Carmeliet, P.; et al. TRPA1 Underlies a Sensing Mechanism for O2. Nat. Chem. Biol. 2011, 7, 701–711. [Google Scholar] [CrossRef] [Green Version]
  188. Berrout, J.; Kyriakopoulou, E.; Moparthi, L.; Hogea, A.S.; Berrout, L.; Ivan, C.; Lorger, M.; Boyle, J.; Peers, C.; Muench, S.; et al. TRPA1–FGFR2 Binding Event Is a Regulatory Oncogenic Driver Modulated by MiRNA-142-3p. Nat. Commun. 2017, 8, 947. [Google Scholar] [CrossRef] [Green Version]
  189. Derouiche, S.; Mariot, P.; Warnier, M.; Vancauwenberghe, E.; Bidaux, G.; Gosset, P.; Mauroy, B.; Bonnal, J.-L.; Slomianny, C.; Delcourt, P.; et al. Activation of TRPA1 Channel by Antibacterial Agent Triclosan Induces VEGF Secretion in Human Prostate Cancer Stromal Cells. Cancer Prev. Res. 2017, 10, 177–187. [Google Scholar] [CrossRef] [Green Version]
  190. Roberts, E.; Cossigny, D.A.F.; Quan, G.M.Y. The Role of Vascular Endothelial Growth Factor in Metastatic Prostate Cancer to the Skeleton. Prostate Cancer 2013, 2013, 418340. [Google Scholar] [CrossRef] [Green Version]
  191. Bernardini, M.; Brossa, A.; Chinigo, G.; Grolez, G.P.; Trimaglio, G.; Allart, L.; Hulot, A.; Marot, G.; Genova, T.; Joshi, A.; et al. Transient Receptor Potential Channel Expression Signatures in Tumour-Derived Endothelial Cells: Functional Roles in Prostate Cancer Angiogenesis. Cancers 2019, 11, 956. [Google Scholar] [CrossRef] [Green Version]
  192. Samanta, A.; Hughes, T.E.T.; Moiseenkova-Bell, V.Y. Transient Receptor Potential (TRP) Channels. In Membrane Protein Complexes: Structure and Function; Harris, J.R., Boekema, E.J., Eds.; Springer: Singapore, 2018; Volume 87, pp. 141–165. ISBN 978-981-10-7756-2. [Google Scholar]
  193. Dong, H.; Shim, K.-N.; Li, J.M.J.; Estrema, C.; Ornelas, T.A.; Nguyen, F.; Liu, S.; Ramamoorthy, S.L.; Ho, S.; Carethers, J.M.; et al. Molecular Mechanisms Underlying Ca2+-Mediated Motility of Human Pancreatic Duct Cells. Cell Physiol. 2010, 299, 11. [Google Scholar] [CrossRef] [Green Version]
  194. 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] [Green Version]
  195. El Hiani, Y.; Ahidouch, A.; Lehen’kyi, V.; Hague, F.; Gouilleux, F.; Mentaverri, R.; Kamel, S.; Lassoued, K.; Brûlé, G.; Ouadid-Ahidouch, H. Extracellular Signal-Regulated Kinases 1 and 2 and TRPC1 Channels Are Required for Calcium-Sensing Receptor-Stimulated MCF-7 Breast Cancer Cell Proliferation. Cell. Physiol. Biochem. 2009, 23, 335–346. [Google Scholar] [CrossRef] [PubMed]
  196. Bomben, V.C.; Sontheimer, H. Disruption of Transient Receptor Potential Canonical Channel 1 Causes Incomplete Cytokinesis and Slows the Growth of Human Malignant Gliomas: TRPC1 Channels Are Essential for Glioma Proliferation. Glia 2010, 58, 1145–1156. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Jiang, H.-N.; Zeng, B.; Zhang, Y.; Daskoulidou, N.; Fan, H.; Qu, J.-M.; Xu, S.-Z. Involvement of TRPC Channels in Lung Cancer Cell Differentiation and the Correlation Analysis in Human Non-Small Cell Lung Cancer. PLoS ONE 2013, 8, e67637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Stewart, T.A.; Azimi, I.; Marcial, D.; Peters, A.A.; Chalmers, S.B.; Yapa, K.T.D.S.; Thompson, E.W.; Roberts-Thomson, S.J.; Monteith, G.R. Differential Engagement of ORAI1 and TRPC1 in the Induction of Vimentin Expression by Different Stimuli. Lab. Investig. 2020, 100, 224–233. [Google Scholar] [CrossRef] [PubMed]
  199. Carmeliet, P. VEGF as a Key Mediator of Angiogenesis in Cancer. Oncology 2005, 69, 4–10. [Google Scholar] [CrossRef]
  200. Zhang, P.; Liu, X.; Li, H.; Chen, Z.; Yao, X.; Jin, J.; Ma, X. TRPC5-Induced Autophagy Promotes Drug Resistance in Breast Carcinoma via CaMKKβ/AMPKα/MTOR Pathway. Sci. Rep. 2017, 7, 3158. [Google Scholar] [CrossRef]
  201. Ma, X.; Cai, Y.; He, D.; Zou, C.; Zhang, P.; Lo, C.Y.; Xu, Z.; Chan, F.L.; Yu, S.; Chen, Y.; et al. Transient Receptor Potential Channel TRPC5 Is Essential for P-Glycoprotein Induction in Drug-Resistant Cancer Cells. Proc. Natl. Acad. Sci. USA 2012, 109, 16282–16287. [Google Scholar] [CrossRef] [Green Version]
  202. Wang, T.; Chen, Z.; Zhu, Y.; Pan, Q.; Liu, Y.; Qi, X.; Jin, L.; Jin, J.; Ma, X.; Hua, D. Inhibition of Transient Receptor Potential Channel 5 Reverses 5-Fluorouracil Resistance in Human Colorectal Cancer Cells. J. Biol. Chem. 2015, 290, 448–456. [Google Scholar] [CrossRef] [Green Version]
  203. Wang, T.; Ning, K.; Sun, X.; Zhang, C.; Jin, L.; Hua, D. Glycolysis Is Essential for Chemoresistance Induced by Transient Receptor Potential Channel C5 in Colorectal Cancer. BMC Cancer 2018, 18, 207. [Google Scholar] [CrossRef] [Green Version]
  204. Plant, T.D.; Schaefer, M. TRPC4 and TRPC5: Receptor-Operated Ca2+-Permeable Nonselective Cation Channels. Cell Calcium 2003, 33, 441–450. [Google Scholar] [CrossRef]
  205. Veliceasa, D.; Ivanovic, M.; Hoepfner, F.T.-S.; Thumbikat, P.; Volpert, O.V.; Smith, N.D. Transient Potential Receptor Channel 4 Controls Thrombospondin-1 Secretion and Angiogenesis in Renal Cell Carcinoma: Thrombospondin-1 Loss in Renal Cancer. FEBS J. 2007, 274, 6365–6377. [Google Scholar] [CrossRef] [PubMed]
  206. Song, H.B.; Jun, H.-O.; Kim, J.H.; Fruttiger, M.; Kim, J.H. Suppression of Transient Receptor Potential Canonical Channel 4 Inhibits Vascular Endothelial Growth Factor-Induced Retinal Neovascularization. Cell Calcium 2015, 57, 101–108. [Google Scholar] [CrossRef] [PubMed]
  207. Zeng, B.; Yuan, C.; Yang, X.; Atkin, S.L.; Xu, S.-Z. TRPC Channels and Their Splice Variants Are Essential for Promoting Human Ovarian Cancer Cell Proliferation and Tumourigenesis. Curr. Cancer Drug Targets 2013, 13, 103–116. [Google Scholar] [CrossRef] [PubMed]
  208. Carson, C.; Raman, P.; Tullai, J.; Xu, L.; Henault, M.; Thomas, E.; Yeola, S.; Lao, J.; McPate, M.; Verkuyl, J.M.; et al. Englerin A Agonizes the TRPC4/C5 Cation Channels to Inhibit Tumour Cell Line Proliferation. PLoS ONE 2015, 10, e0127498. [Google Scholar] [CrossRef]
  209. 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] [Green Version]
  210. Wu, L.; Saxena, S.; Awaji, M.; Singh, R.K. Tumour-Associated Neutrophils in Cancer: Going Pro. Cancers 2019, 11, 564. [Google Scholar] [CrossRef] [Green Version]
  211. Prakriya, M. Store-Operated Orai Channels. In Current Topics in Membranes; Elsevier: Amsterdam, The Netherlands, 2013; Volume 71, pp. 1–32. ISBN 978-0-12-407870-3. [Google Scholar]
  212. 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 BBA Mol. Cell Res. 2014, 1843, 2263–2269. [Google Scholar] [CrossRef] [Green Version]
  213. Didiasova, M.; Zakrzewicz, D.; Magdolen, V.; Nagaraj, C.; Bálint, Z.; Rohde, M.; Preissner, K.T.; Wygrecka, M. STIM1/ORAI1-Mediated Ca2+ Influx Regulates Enolase-1 Exteriorization. J. Biol. Chem. 2015, 290, 11983–11999. [Google Scholar] [CrossRef] [Green Version]
  214. Huang, H.-K.; Lin, Y.-H.; Chang, H.-A.; Lai, Y.-S.; Chen, Y.-C.; Huang, S.-C.; Chou, C.-Y.; Chiu, W.-T. Chemoresistant Ovarian Cancer Enhances Its Migration Abilities by Increasing Store-Operated Ca2+ Entry-Mediated Turnover of Focal Adhesions. J Biomed. Sci. 2020, 27, 36. [Google Scholar] [CrossRef]
  215. Zuccolo, E.; Laforenza, U.; Ferulli, F.; Pellavio, G.; Scarpellino, G.; Tanzi, M.; Turin, I.; Faris, P.; Lucariello, A.; Maestri, M.; et al. Stim and Orai Mediate Constitutive Ca2+ Entry and Control Endoplasmic Reticulum Ca2+ Refilling in Primary Cultures of Colorectal Carcinoma Cells. Oncotarget 2018, 9, 31098–31119. [Google Scholar] [CrossRef] [Green Version]
  216. Singh, A.K. Orai-1 and Orai-2 Regulate Oral Cancer Cell Migration and Colonisation by Suppressing Akt/MTOR/NF-κB Signalling. Life Sci. 2020, 261, 118372. [Google Scholar] [CrossRef] [PubMed]
  217. Sun, J.; Lu, F.; He, H.; Shen, J.; Messina, J.; Mathew, R.; Wang, D.; Sarnaik, A.A.; Chang, W.-C.; Kim, M.; et al. STIM1- and Orai1-Mediated Ca2+ Oscillation Orchestrates Invadopodium Formation and Melanoma Invasion. J. Cell Biol. 2014, 207, 535–548. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Xu, Y.; Zhang, S.; Niu, H.; Ye, Y.; Hu, F.; Chen, S.; Li, X.; Luo, X.; Jiang, S.; Liu, Y.; et al. STIM1 Accelerates Cell Senescence in a Remodeled Microenvironment but Enhances the Epithelial-to-Mesenchymal Transition in Prostate Cancer. Sci. Rep. 2015, 5, 11754. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  219. 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 Tumour Cell Proliferation, Metabolism, Migration, and Invasion in Human Gastric Cancer. Cancer Lett. 2016, 381, 31–40. [Google Scholar] [CrossRef]
  220. Hasna, J.; Hague, F.; Rodat-Despoix, L.; Geerts, D.; Leroy, C.; Tulasne, D.; Ouadid-Ahidouch, H.; Kischel, P. Orai3 Calcium Channel and Resistance to Chemotherapy in Breast Cancer Cells: The P53 Connection. Cell Death Differ. 2018, 25, 693–707. [Google Scholar] [CrossRef]
  221. Tang, B.-D.; Xia, X.; Lv, X.-F.; Yu, B.-X.; Yuan, J.-N.; Mai, X.-Y.; Shang, J.-Y.; Zhou, J.-G.; Liang, S.-J.; Pang, R.-P. Inhibition of Orai1-Mediated Ca2+ Entry Enhances Chemosensitivity of HepG2 Hepatocarcinoma Cells to 5-Fluorouracil. J. Cell. Mol. Med. 2017, 21, 904–915. [Google Scholar] [CrossRef]
  222. Taylor, J.; Azimi, I.; Monteith, G.; Bebawy, M. Ca2+ Mediates Extracellular Vesicle Biogenesis through Alternate Pathways in Malignancy. J. Extracell. Vesicles 2020, 9, 1734326. [Google Scholar] [CrossRef] [Green Version]
  223. 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]
  224. Li, S.; Hao, B.; Lu, Y.; Yu, P.; Lee, H.-C.; Yue, J. Intracellular Alkalinization Induces Cytosolic Ca2+ Increases by Inhibiting Sarco/Endoplasmic Reticulum Ca2+-ATPase (SERCA). PLoS ONE 2012, 7, e31905. [Google Scholar] [CrossRef] [Green Version]
  225. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion Channels and the Hallmarks of Cancer. Trends Mol. Med. 2010, 16, 107–121. [Google Scholar] [CrossRef]
  226. Baghban, R.; Roshangar, L.; Jahanban-Esfahlan, R.; Seidi, K.; Ebrahimi-Kalan, A.; Jaymand, M.; Kolahian, S.; Javaheri, T.; Zare, P. Tumour Microenvironment Complexity and Therapeutic Implications at a Glance. Cell Commun. Signal. 2020, 18, 59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  227. Bogeski, I.; Kummerow, C.; Al-Ansary, D.; Schwarz, E.C.; Koehler, R.; Kozai, D.; Takahashi, N.; Peinelt, C.; Griesemer, D.; Bozem, M.; et al. Differential Redox Regulation of ORAI Ion Channels: A Mechanism to Tune Cellular Calcium Signaling. Sci. Signal. 2010, 3, ra24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  228. Dubois, C.; Vanden Abeele, F.; Lehen’kyi, V.; Gkika, D.; Guarmit, B.; Lepage, G.; Slomianny, C.; Borowiec, A.S.; Bidaux, G.; Benahmed, M.; et al. Remodeling of Channel-Forming ORAI Proteins Determines an Oncogenic Switch in Prostate Cancer. Cancer Cell 2014, 26, 19–32. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  229. Xie, J.; Pan, H.; Yao, J.; Zhou, Y.; Han, W. SOCE and Cancer: Recent Progress and New Perspectives. Int. J. Cancer 2016, 138, 2067–2077. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of the effects of acidic pHe and hypoxia on TRP and Piezo Ca2+-permeable channels. The positive or negative effects of hypoxia and acidic pHe on the Ca2+-permeable channels were obtained from both normal and cancer cells-related studies, while the signalling pathways indicated were obtained uniquely from cancer cells-based investigations. The figure depicts Piezo-, TRPV-, TRPA1- and TRPM-mediated Ca2+-dependent signalling pathways activated or inhibited by acidic pHe and hypoxia and linked to tumour progression. TRPV (TRPV1–4, 6), TRPA1, and TRPM (TRPM6, 7) expressed in cell cancer’s plasma membrane are differentially regulated by acidic pHe, being mostly activated by tumour acidosis, and transducing its signals to activate Ca2+-dependent downstream effectors, such as NF-κB, JAK/STAT, PI3K/AKT, NFAT, ERK, and LIMK. TRPA1 is also activated by hypoxia. These effectors promote tumour cell migration, invasion, proliferation, survival, mesenchymal phenotype, and chemoresistance. TRPV6 channels’ activity is inhibited by tumour acidosis, as TRPM2, which inhibition avoids induction of cancer cell death and reduces chemosensitivity. Piezo channels embedded in stellate cells’ plasma membrane are inhibited by acidic pHe, promoting stellate cells’ survival. TRPV1 activation in lymphatic endothelial cells promotes activation of NF-κB and upregulation of IL-8, a lymphangiogenic factor. CaM, calmodulin; CAMKII, Ca2+/calmodulin-dependent protein kinase II; Pyk2, protein tyrosine kinase 2; RAS, Rat sarcoma virus; ERK, extracellular signal-regulated kinase; FAK, Focal Adhesion Kinase; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B; mTOR, mammalian target of rapamycin; NF-κB, nuclear factor-κB; JAK, Janus kinases; STAT, signal transducer and activator of transcription; NFAT, nuclear factor of activated T-cells; RhoA, Ras homolog family member A; ROCK, Rho-associated protein kinase; LIMK, LIM domain kinase; CREB, C-AMP response element-binding protein. The question mark indicates contradictory results in the literature. Created with BioRender.com, accessed on 20 June 2022.
Figure 1. Overview of the effects of acidic pHe and hypoxia on TRP and Piezo Ca2+-permeable channels. The positive or negative effects of hypoxia and acidic pHe on the Ca2+-permeable channels were obtained from both normal and cancer cells-related studies, while the signalling pathways indicated were obtained uniquely from cancer cells-based investigations. The figure depicts Piezo-, TRPV-, TRPA1- and TRPM-mediated Ca2+-dependent signalling pathways activated or inhibited by acidic pHe and hypoxia and linked to tumour progression. TRPV (TRPV1–4, 6), TRPA1, and TRPM (TRPM6, 7) expressed in cell cancer’s plasma membrane are differentially regulated by acidic pHe, being mostly activated by tumour acidosis, and transducing its signals to activate Ca2+-dependent downstream effectors, such as NF-κB, JAK/STAT, PI3K/AKT, NFAT, ERK, and LIMK. TRPA1 is also activated by hypoxia. These effectors promote tumour cell migration, invasion, proliferation, survival, mesenchymal phenotype, and chemoresistance. TRPV6 channels’ activity is inhibited by tumour acidosis, as TRPM2, which inhibition avoids induction of cancer cell death and reduces chemosensitivity. Piezo channels embedded in stellate cells’ plasma membrane are inhibited by acidic pHe, promoting stellate cells’ survival. TRPV1 activation in lymphatic endothelial cells promotes activation of NF-κB and upregulation of IL-8, a lymphangiogenic factor. CaM, calmodulin; CAMKII, Ca2+/calmodulin-dependent protein kinase II; Pyk2, protein tyrosine kinase 2; RAS, Rat sarcoma virus; ERK, extracellular signal-regulated kinase; FAK, Focal Adhesion Kinase; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B; mTOR, mammalian target of rapamycin; NF-κB, nuclear factor-κB; JAK, Janus kinases; STAT, signal transducer and activator of transcription; NFAT, nuclear factor of activated T-cells; RhoA, Ras homolog family member A; ROCK, Rho-associated protein kinase; LIMK, LIM domain kinase; CREB, C-AMP response element-binding protein. The question mark indicates contradictory results in the literature. Created with BioRender.com, accessed on 20 June 2022.
Ijms 23 07377 g001
Figure 2. Overview of the effects of acidic pHe and hypoxia on TRPC and ORAI Ca2+-permeable channels. The positive or negative effects of hypoxia and acidic pHe on the Ca2+-permeable channels were obtained from both normal and cancer cell-related studies, while the signalling pathways indicated were obtained uniquely from cancer cell-based investigations. The figure depicts TRPCs- and SOCs-mediated Ca2+-dependent signalling pathways inhibited or activated by acidic pHe or hypoxia and linked to tumour progression. TRPC (TRPC1, 4, 5) expressed in cancer cells’ plasma membrane is all activated by acidic pHe or hypoxia, transducing their signals to activate Ca2+-dependent downstream effectors, such as SMAD2/3, NFAT, STAT3, HIF1, AMPK and β-catenin. These effectors promote tumour cell migration, angiogenesis, invasion, proliferation, mesenchymal phenotype and chemoresistance and the expression of TRPC1, via EGFR activation, and TRPC6 via Notch1 signalling pathway, in a mechanism of positive feedback regulation for both TRPC1 and TRPC6 channels. Immune cells expressing TRPC6 channels on plasma membrane show TRPC6’s activity that is inhibited by acidic pHe, reducing their migration. ORAI1 channels function in immune cells is also negatively affected by acidic pHe, impairing different processes needed for immune cells’ anti-tumour activity. Hypoxia promotes both ORAI1 expression, via Notch signalling pathway, and activation, leading to increased ROS resistance, migration, invasion, EMT and cell survival. CaM, calmodulin; CAMKII, Ca2+/calmodulin-dependent protein kinase II; Pyk2, protein tyrosine kinase 2; RAS, rat sarcoma virus; ERK, extracellular signal-regulated kinase; FAK, focal adhesion kinase; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B; mTOR, mammalian target of rapamycin; NICD, Notch intracellular domain; CSL, CBF1, suppressor of hairless, Lag-1; NFAT, nuclear factor of activated T-cells; STAT, signal transducers and activators of transcription; EGFR, epidermal growth factor receptor; HIF-1, hypoxia-inducible factor 1. Created with BioRender.com, accessed on 20 June 2022.
Figure 2. Overview of the effects of acidic pHe and hypoxia on TRPC and ORAI Ca2+-permeable channels. The positive or negative effects of hypoxia and acidic pHe on the Ca2+-permeable channels were obtained from both normal and cancer cell-related studies, while the signalling pathways indicated were obtained uniquely from cancer cell-based investigations. The figure depicts TRPCs- and SOCs-mediated Ca2+-dependent signalling pathways inhibited or activated by acidic pHe or hypoxia and linked to tumour progression. TRPC (TRPC1, 4, 5) expressed in cancer cells’ plasma membrane is all activated by acidic pHe or hypoxia, transducing their signals to activate Ca2+-dependent downstream effectors, such as SMAD2/3, NFAT, STAT3, HIF1, AMPK and β-catenin. These effectors promote tumour cell migration, angiogenesis, invasion, proliferation, mesenchymal phenotype and chemoresistance and the expression of TRPC1, via EGFR activation, and TRPC6 via Notch1 signalling pathway, in a mechanism of positive feedback regulation for both TRPC1 and TRPC6 channels. Immune cells expressing TRPC6 channels on plasma membrane show TRPC6’s activity that is inhibited by acidic pHe, reducing their migration. ORAI1 channels function in immune cells is also negatively affected by acidic pHe, impairing different processes needed for immune cells’ anti-tumour activity. Hypoxia promotes both ORAI1 expression, via Notch signalling pathway, and activation, leading to increased ROS resistance, migration, invasion, EMT and cell survival. CaM, calmodulin; CAMKII, Ca2+/calmodulin-dependent protein kinase II; Pyk2, protein tyrosine kinase 2; RAS, rat sarcoma virus; ERK, extracellular signal-regulated kinase; FAK, focal adhesion kinase; PI3K, phosphoinositide 3-kinase; AKT, protein kinase B; mTOR, mammalian target of rapamycin; NICD, Notch intracellular domain; CSL, CBF1, suppressor of hairless, Lag-1; NFAT, nuclear factor of activated T-cells; STAT, signal transducers and activators of transcription; EGFR, epidermal growth factor receptor; HIF-1, hypoxia-inducible factor 1. Created with BioRender.com, accessed on 20 June 2022.
Ijms 23 07377 g002
Figure 3. Schematic illustration of tumour microenvironment landscape. The increased tumour growth, the acidification of the extracellular space and aberrant vascularisation and limited O2 supply origin a tumour core that is hypoxic and acidic, with limited supply of oxygen and nutrients from the blood vessels. Peripheral tumour cells are located in regions with a higher extracellular pH, a result of proximity to blood vessels and the possibility to wash out acidic waste products. TRP, Piezo and SOCs channels expressed in cancer, immune and stromal cells are presented in the corresponding black boxes with indication of their involvement in different cancer hallmarks, such as proliferation, migration, invasion, angiogenesis, and epithelial–mesenchymal transition. The red box contains the up-to-date information regarding the effect (up arrows = positive effect; down arrows = negative effect) of hypoxia and acid pHe on the activity and/or expression of calcium-permeable channels in cancer cells or tumour-associated cells.
Figure 3. Schematic illustration of tumour microenvironment landscape. The increased tumour growth, the acidification of the extracellular space and aberrant vascularisation and limited O2 supply origin a tumour core that is hypoxic and acidic, with limited supply of oxygen and nutrients from the blood vessels. Peripheral tumour cells are located in regions with a higher extracellular pH, a result of proximity to blood vessels and the possibility to wash out acidic waste products. TRP, Piezo and SOCs channels expressed in cancer, immune and stromal cells are presented in the corresponding black boxes with indication of their involvement in different cancer hallmarks, such as proliferation, migration, invasion, angiogenesis, and epithelial–mesenchymal transition. The red box contains the up-to-date information regarding the effect (up arrows = positive effect; down arrows = negative effect) of hypoxia and acid pHe on the activity and/or expression of calcium-permeable channels in cancer cells or tumour-associated cells.
Ijms 23 07377 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Audero, M.M.; Prevarskaya, N.; Fiorio Pla, A. Ca2+ Signalling and Hypoxia/Acidic Tumour Microenvironment Interplay in Tumour Progression. Int. J. Mol. Sci. 2022, 23, 7377. https://doi.org/10.3390/ijms23137377

AMA Style

Audero MM, Prevarskaya N, Fiorio Pla A. Ca2+ Signalling and Hypoxia/Acidic Tumour Microenvironment Interplay in Tumour Progression. International Journal of Molecular Sciences. 2022; 23(13):7377. https://doi.org/10.3390/ijms23137377

Chicago/Turabian Style

Audero, Madelaine Magalì, Natalia Prevarskaya, and Alessandra Fiorio Pla. 2022. "Ca2+ Signalling and Hypoxia/Acidic Tumour Microenvironment Interplay in Tumour Progression" International Journal of Molecular Sciences 23, no. 13: 7377. https://doi.org/10.3390/ijms23137377

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