Skip to main content
Philosophical Transactions of the Royal Society B: Biological Sciences logoLink to Philosophical Transactions of the Royal Society B: Biological Sciences
. 2014 Mar 19;369(1638):20130102. doi: 10.1098/rstb.2013.0102

Ion channels and transporters in tumour cell migration and invasion

Albrecht Schwab 1,, Christian Stock 1
PMCID: PMC3917356  PMID: 24493750

Abstract

Cell migration is a central component of the metastatic cascade requiring a concerted action of ion channels and transporters (migration-associated transportome), cytoskeletal elements and signalling cascades. Ion transport proteins and aquaporins contribute to tumour cell migration and invasion among other things by inducing local volume changes and/or by modulating Ca2+ and H+ signalling. Targeting cell migration therapeutically bears great clinical potential, because it is a prerequisite for metastasis. Ion transport proteins appear to be attractive candidate target proteins for this purpose because they are easily accessible as membrane proteins and often overexpressed or activated in cancer. Importantly, a number of clinically widely used drugs are available whose anticipated efficacy as anti-tumour drugs, however, has now only begun to be evaluated.

Keywords: migration, invasion, ionic signalling, transportome, metastasis

1. Introduction

Tumour progression towards metastatic disease follows a well-defined sequence of events [1]. The metastatic cascade includes several critical steps that rely on the ability of tumour cells to migrate: local invasion of the affected tissue following the disruption of the basement membrane as well as intra- and extravasation of blood or lymph vessels (figure 1). Without their ability to move, tumour cells would not be able to metastasize. However, tumour cells do not necessarily act independently. They also can migrate collectively as a group [2]. Moreover, there is an intense mutual communication between tumour and stromal cells, and the tumour microenvironment that is typically characterized by local acidosis and hypoxia [35]. Thus, in pancreatic ductal adenocarcinoma (PDAC), reciprocal activation of stromal stellate cells and cancer cells strongly promotes tumour progression [6]. Metastases in PDAC may even contain tumour cells and pancreatic stellate cells [7]. Monitoring the collective invasion of co-cultured fibroblasts and squamous cell carcinoma cells revealed that the invading cell group is always led by fibroblasts [8]. Thus, the ability to migrate is equally important for cancer and stromal cells.

Figure 1.

Figure 1.

Involvement of cell migration in the metastatic cascade. During malignant transformation, cells of the primary tumour lose their contacts with the neighbouring, healthy cells and become motile. They degrade basement membranes, invade and cross extracellular matrices in order to eventually intravasate into blood or lymph vessels. Only few cells survive the intravascular milieu and adhere to the vessel wall at a distant site where they extravasate and invade to form a metastasis. Those steps of the metastatic cascade that rely on migration are labelled with a hooklet (tick symbols). (Online version in colour.)

This review highlights the role of ion channels and transporters in the steps of the metastatic cascade that rely on the ability of tumour and stromal cells to migrate. Local fluctuations of the cell volume as well as pH and Ca2+ signalling evolved as common mechanisms linking ion transport proteins to the metastatic behaviour. We refer to three recent reviews [911] for a broader overview of ion transport in cell motility.

2. Ionic mechanisms of cell migration

Most malignant tumours are of epithelial origin. Hence, carcinoma cells have lost their epithelial polarization during epithelial–mesenchymal transition (EMT). Mesenchymal cells can detach and move away from the epithelial layer [12]. EMT involves a coordinated gene expression programme in association with the early steps of transformation which also can include proteins involved in ion transport. Thus, the ectopic expression of carbonic anhydrase (CAIX) is accompanied by a loss of cell–cell contacts. Subsequently, CAIX, pH-regulating transport proteins (e.g. NBCe1, AE2, NHE1, MCTs) and aquaporins are reallocated to the cell front where they elicit functions relevant for cell migration (see below [13]).

Cell migration can be modelled as a repeated cycle of protrusion of the cell front and retraction of the cell rear. This involves a complex interplay of multiple components of the cellular migration machinery that are tightly regulated in space and time. Immediately at the leading edge, actin filaments polymerize while simultaneously depolymerizing towards the lamellipodial rear. This process is controlled by proteins that bind actin monomers, sever and branch existing filaments or nucleate the formation of new ones. The interaction of actin filaments with cytoskeletal motor proteins such as myosin generates force that is transmitted via focal adhesion complexes onto the surrounding extracellular matrix (ECM). In moving cells, new contacts are formed at the cell front, whereas others are released at the rear part. We refer to recent reviews that provide in-depth overviews of the cytoskeleton [14] and cell adhesion [15] in migrating cells. Cytoskeletal and cell adhesion dynamics are regulated by ionic mechanisms, and thereby depend on the activity of the respective transport proteins. These will be described in the following paragraphs with special emphasis on pH- and Ca2+-dependence and the role of local changes of cell volume played therein. However, the involved transport proteins do not only regulate the intra- and extracellular pH and Ca2+ homeostasis. Many of them are pH- and/or Ca2+-sensitive themselves. Moreover, intracellular Ca2+ and H+ concentrations are physico-chemically coupled, and therefore cannot be changed entirely independently from each other [16]. The role of the mutual feedback of functionally cooperating transport proteins, the ‘transportome’, including their signalling pathways in cell migration has only begun to be appreciated. We will not discuss the interaction of K+ channels with integrins as this is covered elsewhere in this issue [17].

(a). pH-dependent regulation of the cytoskeleton in tumour cell migration

Intracellular and extracellular pH homeostasis is particularly important in cancer as outlined elsewhere in this issue [4,5,18]. As the ionization state of all cellular and extracellular proteins including their function in (patho)physiological processes depends on pH, (directional) tumour cell migration is controlled by intra- and extracellular pH [1923].

Cofilin regulates cell motility pH dependently [24]. It generates new sites of actin filament assembly by severing actin filaments and producing free barbed filament ends. This promotes dynamic actin polymerization and membrane protrusion at the cell front or at the tip of invasive structures [25]. The inhibition of cofilin activity by PI-(4,5)-P2 binding is removed upon an intracellular alkalinization [26]. The local intracellular alkalinization in the lamellipodium required for cofilin activation is generated by the activity of the Na+/H+ exchanger NHE1 that accumulates at the front of migrating tumour cells and fibroblasts [2731]. NHE1 is upregulated in numerous tumours [22,32,33] and is required for motility of melanoma, breast cancer and cervix carcinoma cells [34]. Another way of pH-dependent regulation of cofilin involves cortactin. An NHE1-mediated increase in pHi triggers the release of cortactin-bound cofilin. Cofilin then induces barbed end generation, thereby promoting actin polymerization [35]. Gelsolin is another actin-binding protein that is activated by an acidic pH and that controls actin assembly and disassembly [36]. Because NHE1 activity contributes to the generation of an intracellular pH gradient along the moving direction of migrating cells with more alkaline pH values in the lamellipodium [37], we assume that cofilin is more relevant for regulating actin dynamics at the leading edge, whereas gelsolin is more active at the acidic rear end. Finally, actin self-assembly and binding of myosin to actin is promoted by neutral or slightly acidic pHi values [38,39]. These examples demonstrate that actin dynamics underlying the outgrowth of lamellipodia or invadopodia relies on an optimal intracellular pH environment [11,40], which, in turn, is adjusted by the activity of pH regulating H+ (NHE1) and/or HCO3 transporters, possibly in cooperation with carbonic anhydrases or monocarboxylate transporters [13,32,4143].

(b). Ca2+-dependent regulation of the cytoskeleton in tumour cell migration

The intracellular calcium concentration ([Ca2+]i) has a great impact on the migration machinery of ‘normal’ (e.g. keratinocytes), tumour and stromal cells, too, because many of its elements such as myosin II [44], myosin light chain kinase [45], calpain [46], Ca2+/calmodulin-dependent protein kinase II [47,48], focal adhesion kinase [49] or ion channels (e.g. KCa or TMEM16 channels) are Ca2+-sensitive. Ca2+ regulation of cell migration involves a tight spatial and temporal control. In addition to a global front–rear gradient with [Ca2+]i increasing towards the rear end of migrating cells [50,51], there are also local Ca2+ microdomains [45,5254]. Spatial gradients of [Ca2+]i are superimposed by temporal changes, or oscillations of [Ca2+]i that contribute to dynamic responses of the cellular migration apparatus [45,55]. One of the major mechanisms by which [Ca2+]i signalling impacts on cell migration is the modulation of cytoskeletal dynamics.

α-Acitinin is an actin-binding protein controlling lamellipodial dynamics and directional migration [56]. It is Ca2+-sensitive with an increase of [Ca2+]i, causing an inhibition of actin bundling activity [57]. Moreover, [Ca2+]i is involved in the generation of a myosin-II-dependent contractile force during the retraction of the rear part of migrating cells [58]. On a smaller scale, retraction also occurs at the cell front. Here, local Ca2+ pulses of minute amplitudes that originate a few mircometres behind the leading edge induce the retraction of confined areas of the lamellipodium by activating MLCK and myosin-II-mediated contraction [45]. Interestingly, STIM1, a component of the store-operated Ca2+ entry (SOCE) channel, continuously appears close to the leading edge of PDAC cells [59]. Calcium-dependent development of mechanical force also has a large impact on cell adhesion which is detailed below. In neutrophil granulocytes, the ablation of TRPC6 channels caused a severe impairment of the cells to chemotax towards the KC (CXCL1) [60]. Local TRPM7-mediated Ca2+ sparks at the cell front are required for directional migration of fibroblasts towards a source of platelet-derived growth factor (PDGF) [53]. While the two latter studies were not performed with tumour cells, they are nonetheless instructive for tumour pathophysiology. First, they reveal signalling cascades that are also encountered in tumours such as PDAC [61]. Second, the ability of tumour or stromal cells to chemotax secures their mutual communication in response to paracrine growth factor stimulation in the tumour microenvironment. Chemotaxis of tumour cells also requires the activity of ion channels such as TRPC1 and/or KCa3.1 channels [6265].

(c). pH-dependent regulation of tumour cell adhesion

Integrins are integral components of focal adhesions. Depending on their subunit composition, they mediate the interaction with different proteins of the ECM. Multiple integrins, such as α2β1, α5β1 or ανβ3, are pH-dependent in melanoma [19] and in other cells [66,67]. Increased adhesion at acidic extracellular pH values is explained by conformational changes that lead to an enhanced aviditity of the integrin headpieces to ECM proteins [67] or by a pH dependence of the mechanical stability of focal adhesions [68]. Conceptually, it is important that integrins protrude only 20 nm into the extracellular space [69]. Therefore, it is not surprising that the pericellular pHe inside the glycocalyx is more important for cell adhesion than the pH of the extracellular bulk solution surrounding migrating melanoma cells [19,21]. The pericellular pHe even confers asymmetry upon migrating tumour cells, because it is more acidic at the cell front than at the rear end [21,28]. The global pericellular front–rear pH gradient is superimposed by acidic pHe nanodomains generated by NHE1 activity. Because these nanodomains are restricted to focal adhesions, we assume that NHE1 activity locally stabilizes the integrin-mediated interaction between cell surface and ECM [70]. In melanoma and in endothelial cells, this pericellular pHe gradient is accompanied by a complementary intracellular pH gradient. pHi is more alkaline at the cell front than at the rear part [37,71]. The alkaline pHi at the cell front leads to a higher focal adhesion turnover owing to the lower affinity of talin for binding actin [40].

(d). Ca2+-dependent regulation of tumour cell adhesion

The maturation and dynamic turnover of focal adhesions are modulated by the [Ca2+]i [45,49,53,54,72]. This occurs in part because of a global gradient with [Ca2+]i being usually higher at the rear end than at the front [5053] which restricts the disassembly of focal adhesions by Ca2+-sensitive family of calpain phosphatases to the rear part of migrating cells [73,74]. Alternatively, local Ca2+ elevations affect focal adhesion dynamics, possibly by mediating tyrosine phosphorylation of FAK [54]. Adhesion of cancer cells to the ECM is also mediated by invadopodia that share many similarities with podosomes found in ‘normal’ cells such as macrophages [75]. In microglial cells, the formation of invadopodia depends on the presence of extracellular Ca2+ whose transport into invadopodia is likely mediated by Orai1 [76].

(e). Cell volume dynamics during cell migration

As outlined above, cell migration can be viewed as a repetitive cycle of protrusion of the cell front and retraction of the rear part. The rear part often lags behind for quite some time before retracting at a much faster speed. Such shape changes are particularly prominent when cells are moving within a three-dimensional environment where tumour cells or stromal cells may extend processes as long as 100 µm before the cell body eventually catches up. This fast retraction of the rear part coincides with or follows an elevation of the [Ca2+]i [55,77] which is likely to be caused by the activation of mechanosensitive Ca2+ channels. Their molecular identity has not yet been conclusively determined. TRPC1 [78], TRPM7 [53] and TRPV4 [79] channels are possible candidates. In addition to triggering the Ca2+-dependent mechanisms outlined above, the elevation of [Ca2+]i also leads to an activation of Ca2+-sensitive ion channels such as KCa3.1 [80], KCa2.3 [81], ClC3 [48,82] or TMEM16A/ANO1 [8385]. Studies of cell volume regulation revealed that cell shrinkage can be elicited by the simultaneous activation of K+ and Cl channels [86]. Therefore, a hydrodynamic model was postulated according to which ion channels and transporters elicit local changes of cell volume that act in concert with cytoskeletal mechanisms underlying rear end retraction and/or the protrusion of the cell front [80,87,88]. This model was confirmed experimentally by several groups [8993]. The observation that aquaporins are essential components of the cellular migration apparatus [94] lent further strong support to this model. Aquaporins provide the route for osmotically driven water influx or efflux, leading to local cell swelling at the cell front or shrinkage at the rear part, respectively.

3. Ionic mechanisms of tumour cell invasion

Tumour cell invasion is frequently linked to invadopodia [75]. They are sites of proteolytic degradation of the ECM, and thereby facilitate migration through a three-dimensional network of matrix fibres (figure 2). Traditionally, the role of invadopodia in tumour cell invasion is reflected by the presence of several proteases (e.g. MT1MMP, MMP2). The ‘microscopic’ NHE1-mediated acidification at the cell surface of lamellipodia and invadopodia facilitates the action of pH-dependent proteases [28,95,96]. The activity of matrix metalloproteinase 2 (MMP2) requires the protonation of the substrate such as fibrinogen [97], and the expression of MMP9 is upregulated by an acidic extracellular pH [98]. Invasive behaviour of various tumour cells is also triggered by the expression of voltage-gated Na+ channels (NaV) [99]. They appear to cooperate with NHE1 during invasion of breast cancer cells by as yet undefined mechanisms [100] and increase the activity of cysteine cathepsins [101].

Figure 2.

Figure 2.

Hypothetical model of invadopodium formation involving pH, Ca2+ and volume regulation. Phosphorylation of cortactin (P) modulates the interaction between NHE1 and cortactin. Stimulated NHE1 activity increases pHi and triggers the release of cortactin-bound cofilin. Cofilin then promotes actin polymerization. Cdc42, a small, pH-dependent Rho-GTPase, regulates actin polymerization by binding to the neural Wiskott–Aldrich syndrome protein (NWASP) which then activates the Arp2/3 complex. H+ extruded by NHE1 causes an extracellular acidification, thereby facilitating the interaction between integrins and collagen and promotes the activity of matrix metalloproteinases (MMP). Lamellipodia/invadopodia outgrowth requires local volume increase that is mediated by water uptake through aquaporins (AQP) and possibly driven by the osmolytes glucose and Na+ imported by the Na+, glucose co-transporter 1 (SGLT1). By analogy with podosomes, Ca2+ influx through ORAI channels would occur and stimulate both calpain2 (calp2) to cleave cortactin and calmodulin to activate KCa2.3 channels. KCa2.3 could fine-tune Na+ entry through the SGLT1 by keeping the membrane potential stable and controlling the amount of osmolytes entering the invadopodia. ORAI1 contributes to focal adhesion dynamics by co-localizing with PLA2g6. PLA2g6 supports the phosphorylation of the focal adhesion kinase (FAK) by activating ORAI1. (Online version in colour.)

Based on the great similarity between podosomes and invadopodia, we assume that a similar Ca2+ dependence will also be found in cancer cells [76]. This view is supported by the fact that the ability of cancer cells to invade extracellular matrices could be related to the activity of Ca2+ influx channels such as TRPM7 or ORAI1 [102104]. Finally, the activity of matrix metalloproteinases has also been linked to Ca2+ signalling [105,106]. The upregulation of MMP9 requiring Ca2+ influx can therefore be inhibited by blocking voltage-gated Ca2+ channels [107].

(a). Volume dynamics during tumour cell invasion

The ability of tumour cells to locally change their volume and shape, respectively, also facilitates their movement through the tortuous interstitium and through the wall of a blood/lymph vessel (intra-/extravasation). Ion channels and transporters thereby convey a means to the tumour cells to overcome mechanical barriers [108] which is particularly relevant when tumour cells invade the interstitium. Accordingly, migration of glioblastoma cells on a two-dimensional surface is not impaired when the Na+/K+/2Cl cotransporter is inhibited while this is the case when they are invading the brain parenchyma [109]. Possibly, growth and protrusion of invadopodia are also supported by a mechanism similar to the one used by the intracellular pathogen Cryptosporidium parvum that induces SGLT1- and AQP1-driven membrane protrusions in the host cell [110] (figure 2). Glucose transporters are widely expressed in many tumour cells to ensure their metabolic supply [111]. Finally, their role in tumour cell invasion [112] may be due to the fact that the protrusion of lamellipodia is linked to glycolytic energy production [31]. Preclinical and clinical observations lend further support to the importance of cellular volume dynamics during invasion of tumour cells. Expression of aquaporins is associated with poor prognosis and metastatic relapse of a number of tumours (e.g. AQP1 overexpression in patients with lung adenocarcinoma; [113]).

4. Outlook and clinical perspectives

Studies from the past approximately 15 years provided proof of concept that ion channels and transporters are crucial for the metastatic behaviour of tumour cells. It is becoming increasingly clear that ion transport proteins do not act in an isolated manner on their own, but that they act in networks of functionally cooperating units [13,63,114,115]. This is reflected by the concept of the migration-associated transportome that we recently introduced [9]. It implies that ion transport signalling pathways such as pH, Ca2+ or cell volume are closely linked to each other. However, neither the pathophysiological significance of this crosstalk is well understood, let alone the interaction of ionic with kinase-based signalling pathways. For instance, it has not yet been investigated systematically how the altered expression of pH regulatory transporters in solid tumours affect the functional impact of other members of the migration-associated transportome and their crosstalk with growth factor signalling. Similarly, it is unknown whether transport proteins whose upregulation is mediated by hypoxia-induced HIF1α constitute ‘functional units’ involved in cell migration.

Extravasation of tumour cells in tumour-specific target organs is another step of the metastatic process for which the role of ion channels and transporters has not yet been investigated in detail. This process bears similarities to the recruitment of leucocytes from the bloodstream. It is supported by the cooperation of tumour cells with platelets and leucocytes [116]. Thus, it is reasonable to assume that tumour cells use similar ionic mechanisms to leucocytes (e.g. neutrophil granulocytes) in which Ca2+ influx via Orai1 plays an important role in the initial steps of recruitment [117]. Indeed, functional coupling between KCa2.3 and ORAI1 channels in lipid rafts seems to promote bone metastases of breast cancer cells [115]. Along the same lines, one could speculate that the reciprocal activation of integrins and KV11.1 channels in tumour cells contributes to their ability to extravasate, because KV11.1 expression in acute myeloid leukaemia cells correlates with a higher probability of relapse [118].

Many of the transport proteins (e.g. KV10.1) are not only involved in controlling cell migration and invasion, but also in other ‘hallmarks of cancer’ such as proliferation [114,119]. Thus, drugs that target members of the migration-associated transportome are likely to elicit more responses than just the inhibition of tumour cell migration and invasion. Combined effects such as inhibition of migration and proliferation by blocking KCa3.1 channels may be desirable [120,121], whereas the combined inhibition of migration and apoptosis by KV1.3 channel blockade would certainly not be advantageous.

Because most tumour patients die of metastases the inhibition of the mechanisms underlying metastasis offers great therapeutic potential. Targeting tumour cell migration would therefore be a good choice because it is one of the prerequisites for metastasis. Indeed, migration has been targeted in order to treat chronic inflammatory diseases [122] which is another pathological condition strongly relying on the ability of (inflammatory) cells to migrate. Transport proteins that could qualify as a potential anti-migratory target include among others NHE1, KCa3.1 channels or NaVs. Migration and invasion of all tumours cells studied to date relies at least partially on the activity of one of these proteins [9]. Importantly, there are also small molecule inhibitors validated in phase III clinical trials (KCa3.1 blocker senicapoc [123] and NHE1 blocker cariporide [124]) or that are already widely used clinically such as amide-linked local anaesthetics that block NaVs. Presently it is being discussed whether the NaV-mediated metastatic behaviour of tumour cells can be targeted by using amide-linked local anaesthetics during cancer surgery [125]. Similarly, the use of the K+ sparing diuretic amiloride which is also an NHE1 blocker, can be envisaged [126]. Thus, there is an increasing number of functionally relevant ion transport proteins to be targeted. They are easily accessible because they are membrane proteins and often overexpressed or activated in cancer. Moreover, they have a long history of being drug targets in other medical fields such as cardiology, nephrology or anaesthesia. Alternatively, transport-associated proteins such as the tumour marker carbonic anhydrase IX could be targeted by specific antibodies [127].

Funding statement

Work cited from our laboratory was supported by Deutsche Forschungsgemeinschaft, European Commission (ITN ‘IonTraC’), IZKF Münster, IMF Münster.

References

  • 1.Gupta GP, Massague J. 2006. Cancer metastasis: building a framework. Cell 127, 679–695. ( 10.1016/j.cell.2006.11.001) [DOI] [PubMed] [Google Scholar]
  • 2.Hegerfeldt Y, Tusch M, Brocker EB, Friedl P. 2002. Collective cell movement in primary melanoma explants: plasticity of cell-cell interaction, beta1-integrin function, and migration strategies. Cancer Res. 62, 2125–2130. [PubMed] [Google Scholar]
  • 3.Egeblad M, Nakasone ES, Werb Z. 2010. Tumors as organs: complex tissues that interface with the entire organism. Dev. Cell 18, 884–901. ( 10.1016/j.devcel.2010.05.012) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Andersen AP, Moreira JMA, Pedersen SF. 2014. Interactions of ion transporters and channels with cancer cell metabolism and the tumour microenvironment. Phil. Trans. R. Soc. B 369, 20130098 ( 10.1098/rstb.2013.0098) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Swietach P, Vaughan-Jones RD, Harris AL, Hulikova A. 2014. The chemistry, physiology and pathology of pH in cancer. Phil. Trans. R. Soc. B 369, 20130099 ( 10.1098/rstb.2013.0099) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Feig C, Gopinathan A, Neesse A, Chan DS, Cook N, Tuveson DA. 2012. The pancreas cancer microenvironment. Clin. Cancer Res. 18, 4266–4276. ( 10.1158/1078-0432.CCR-11-3114) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7.Xu Z, et al. 2010. Role of pancreatic stellate cells in pancreatic cancer metastasis. Am. J. Pathol. 177, 2585–2596. ( 10.2353/ajpath.2010.090899) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 8.Gaggioli C, Hooper S, Hidalgo-Carcedo C, Grosse R, Marshall JF, Harrington K, Sahai E. 2007. Fibroblast-led collective invasion of carcinoma cells with differing roles for RhoGTPases in leading and following cells. Nat. Cell Biol. 9, 1392–1400. ( 10.1038/ncb1658) [DOI] [PubMed] [Google Scholar]
  • 9.Schwab A, Fabian A, Hanley PJ, Stock C. 2012. Role of ion channels and transporters in cell migration. Physiol. Rev. 92, 1865–1913. ( 10.1152/physrev.00018.2011) [DOI] [PubMed] [Google Scholar]
  • 10.Stock C, Ludwig F, Hanley PJ, Schwab A. 2013. Roles of ion transport in control of cell motility. Compr. Physiol. 3, 59–119. [DOI] [PubMed] [Google Scholar]
  • 11.Stock C, Schwab A. 2009. Protons make tumor cells move like clockwork. Pflügers Arch. 458, 981–992. ( 10.1007/s00424-009-0677-8) [DOI] [PubMed] [Google Scholar]
  • 12.Thiery JP, Sleeman JP. 2006. Complex networks orchestrate epithelial–mesenchymal transitions. Nat. Rev. Mol. Cell Biol. 7, 131–142. ( 10.1038/nrm1835) [DOI] [PubMed] [Google Scholar]
  • 13.Svastova E, et al. 2012. Carbonic anhydrase IX interacts with bicarbonate transporters in lamellipodia and increases cell migration via its catalytic domain. J. Biol. Chem. 287, 3392–3402. ( 10.1074/jbc.M111.286062) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Le Clainche C, Carlier MF. 2008. Regulation of actin assembly associated with protrusion and adhesion in cell migration. Physiol. Rev. 88, 489–513. ( 10.1152/physrev.00021.2007) [DOI] [PubMed] [Google Scholar]
  • 15.Broussard JA, Webb DJ, Kaverina I. 2008. Asymmetric focal adhesion disassembly in motile cells. Curr. Opin. Cell Biol. 20, 85–90. ( 10.1016/j.ceb.2007.10.009) [DOI] [PubMed] [Google Scholar]
  • 16.Swietach P, Youm JB, Saegusa N, Leem CH, Spitzer KW, Vaughan-Jones RD. 2013. Coupled Ca2+/H+ transport by cytoplasmic buffers regulates local Ca2+ and H+ ion signaling. Proc. Natl Acad. Sci. USA 110, E2064–E2073. ( 10.1073/pnas.1222433110) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Arcangeli A, Crociani O, Bencini L. 2014. Interaction of tumour cells with their microenvironment: ion channels and cell adhesion molecules. A focus on pancreatic cancer. Phil. Trans. R. Soc. B 369, 20130101 ( 10.1098/rstb.2013.0101) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Reshkin SJ, Greco MR, Cardone RA. 2014. Role of pHi, and proton transporters in oncogene-driven neoplastic transformation. Phil. Trans. R. Soc. B 369, 20130100 ( 10.1098/rstb.2013.0100) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Krähling H, Mally S, Eble JA, Noel J, Schwab A, Stock C. 2009. The glycocalyx maintains a cell surface pH nanoenvironment crucial for integrin-mediated migration of human melanoma cells. Pflügers Arch. 458, 1069–1083. ( 10.1007/s00424-009-0694-7) [DOI] [PubMed] [Google Scholar]
  • 20.Webb BA, Chimenti M, Jacobson MP, Barber DL. 2011. Dysregulated pH: a perfect storm for cancer progression. Nat. Rev. Cancer 11, 671–677. ( 10.1038/nrc3110) [DOI] [PubMed] [Google Scholar]
  • 21.Stüwe L, Muller M, Fabian A, Waning J, Mally S, Noel J, Schwab A, Stock C. 2007. pH dependence of melanoma cell migration: protons extruded by NHE1 dominate protons of the bulk solution. J. Physiol. 585, 351–360. ( 10.1113/jphysiol.2007.145185) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Chiang Y, Chou CY, Hsu KF, Huang YF, Shen MR. 2008. EGF upregulates Na+/H+ exchanger NHE1 by post-translational regulation that is important for cervical cancer cell invasiveness. J. Cell Physiol. 214, 810–819. ( 10.1002/jcp.21277) [DOI] [PubMed] [Google Scholar]
  • 23.Paradise RK, Whitfield MJ, Lauffenburger DA, Van Vliet KJ. 2013. Directional cell migration in an extracellular pH gradient: a model study with an engineered cell line and primary microvascular endothelial cells. Exp. Cell Res. 319, 487–497. ( 10.1016/j.yexcr.2012.11.006) [DOI] [PubMed] [Google Scholar]
  • 24.Yonezawa N, Nishida E, Sakai H. 1985. pH control of actin polymerization by cofilin. J. Biol. Chem. 260, 14 410–14 412. [PubMed] [Google Scholar]
  • 25.Wang W, Eddy R, Condeelis J. 2007. The cofilin pathway in breast cancer invasion and metastasis. Nat. Rev. Cancer 7, 429–440. ( 10.1038/nrc2148) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Frantz C, Barreiro G, Dominguez L, Chen X, Eddy R, Condeelis J, Kelly MJS, Jacobson MP, Barber DL. 2008. Cofilin is a pH sensor for actin free barbed end formation: role of phosphoinositide binding. J. Cell Biol. 183, 865–879. ( 10.1083/jcb.200804161) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Grinstein S, Woodside M, Waddell TK, Downey GP, Orlowski J, Pouyssegur J, Wong DC, Foskett JK. 1993. Focal localization of the NHE-1 isoform of the Na+/H+ antiport: assessment of effects on intracellular pH. EMBO J. 12, 5209–5218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Stock C, Mueller M, Kraehling H, Mally S, Noel J, Eder C, Schwab A. 2007. pH nanoenvironment at the surface of single melanoma cells. Cell Physiol. Biochem. 20, 679–686. ( 10.1159/000107550) [DOI] [PubMed] [Google Scholar]
  • 29.Clement DL, Mally S, Stock C, Lethan M, Satir P, Schwab A, Pedersen SF, Christensen ST. 2013. PDGFRalpha signaling in the primary cilium regulates NHE1-dependent fibroblast migration via coordinated differential activity of MEK1/2-ERK1/2-p90RSK and AKT signaling pathways. J. Cell Sci. 126, 953–965. ( 10.1242/jcs.116426) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Denker SP, Huang DC, Orlowski J, Furthmayr H, Barber DL. 2000. Direct binding of the Na–H exchanger NHE1 to ERM proteins regulates the cortical cytoskeleton and cell shape independently of H+ translocation. Mol. Cell 6, 1425–1436. ( 10.1016/S1097-2765(00)00139-8) [DOI] [PubMed] [Google Scholar]
  • 31.Lagana A, Vadnais J, Le PU, Nguyen TN, Laprade R, Nabi IR, Noël J. 2000. Regulation of the formation of tumor cell pseudopodia by the Na+/H+ exchanger NHE1. J. Cell Sci. 113, 3649–3662. [DOI] [PubMed] [Google Scholar]
  • 32.Lauritzen G, et al. 2012. The Na+/H+ exchanger NHE1, but not the Na+, HCO3 cotransporter NBCn1, regulates motility of MCF7 breast cancer cells expressing constitutively active ErbB2. Cancer Lett. 317, 172–183. ( 10.1016/j.canlet.2011.11.023) [DOI] [PubMed] [Google Scholar]
  • 33.Reshkin SJ, Bellizzi A, Albarani V, Guerra L, Tommasino M, Paradiso A, Casavola V. 2000. Phosphoinositide 3-kinase is involved in the tumor-specific activation of human breast cancer cell Na+/H+ exchange, motility, and invasion induced by serum deprivation. J. Biol. Chem. 275, 5361–5369. ( 10.1074/jbc.275.8.5361) [DOI] [PubMed] [Google Scholar]
  • 34.Stock C, Ludwig FT, Schwab A. 2012. Is the multifunctional Na+/H+ exchanger isoform 1 a potential therapeutic target in cancer? Curr. Med. Chem. 19, 647–660. ( 10.2174/092986712798992101) [DOI] [PubMed] [Google Scholar]
  • 35.Magalhaes MA, Larson DR, Mader CC, Bravo-Cordero JJ, Gil-Henn H, Oser M, Chen X, Koleske AJ, Condeelis J. 2011. Cortactin phosphorylation regulates cell invasion through a pH-dependent pathway. J. Cell Biol. 195, 903–920. ( 10.1083/jcb.201103045) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Lagarrigue E, Ternent D, Maciver SK, Fattoum A, Benyamin Y, Roustan C. 2003. The activation of gelsolin by low pH: the calcium latch is sensitive to calcium but not pH. Eur. J. Biochem. 270, 4105–4112. ( 10.1046/j.1432-1033.2003.03803.x) [DOI] [PubMed] [Google Scholar]
  • 37.Martin C, Pedersen SF, Schwab A, Stock C. 2011. Intracellular pH gradients in migrating cells. Am. J. Physiol. Cell Physiol. 300, 490–495. ( 10.1152/ajpcell.00280.2010) [DOI] [PubMed] [Google Scholar]
  • 38.Wang F, Sampogna RV, Ware BR. 1989. pH dependence of actin self-assembly. Biophys. J. 55, 293–298. ( 10.1016/S0006-3495(89)82804-8) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Köhler S, Schmoller KM, Crevenna AH, Bausch AR. 2012. Regulating contractility of the actomyosin cytoskeleton by pH. Cell Rep. 2, 433–439. ( 10.1016/j.celrep.2012.08.014) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Srivastava J, Barreiro G, Groscurth S, Gingras AR, Goult BT, Critchley DR, Kelly MJ, Jacobson MP, Barber DL. 2008. Structural model and functional significance of pH-dependent talin–actin binding for focal adhesion remodeling. Proc. Natl Acad. Sci. USA 105, 14 436–14 441. ( 10.1073/pnas.0800915105) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Johnson DE, Casey JR. 2011. Cytosolic H+ microdomain developed around AE1 during AE1-mediated Cl/HCO3 exchange. J. Physiol. 589, 1551–1569. ( 10.1113/jphysiol.2010.201483) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Gallagher SM, Castorino JJ, Wang D, Philp NJ. 2007. Monocarboxylate transporter 4 regulates maturation and trafficking of CD147 to the plasma membrane in the metastatic breast cancer cell line MDA-MB-231. Cancer Res. 67, 4182–4189. ( 10.1158/0008-5472.CAN-06-3184) [DOI] [PubMed] [Google Scholar]
  • 43.Boedtkjer E, Moreira JM, Mele M, Vahl P, Wielenga VT, Christiansen PM, Jensen VED, Pedersen SF, Aalkjaer C. 2013. Contribution of Na+, HCO3-cotransport to cellular pH control in human breast cancer: a role for the breast cancer susceptibility locus NBCn1 (SLC4A7). Int. J. Cancer 132, 1288–1299. ( 10.1002/ijc.27782) [DOI] [PubMed] [Google Scholar]
  • 44.Betapudi V, Rai V, Beach JR, Egelhoff T. 2010. Novel regulation and dynamics of myosin II activation during epidermal wound responses. Exp. Cell Res. 316, 980–991. ( 10.1016/j.yexcr.2010.01.024) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Tsai FC, Meyer T. 2012. Ca2+ pulses control local cycles of lamellipodia retraction and adhesion along the front of migrating cells. Curr. Biol. 22, 837–842. ( 10.1016/j.cub.2012.03.037) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Campbell RL, Davies PL. 2012. Structure–function relationships in calpains. Biochem. J. 447, 335–351. ( 10.1042/BJ20120921) [DOI] [PubMed] [Google Scholar]
  • 47.Daft PG, Yuan K, Warram JM, Klein MJ, Siegal GP, Zayzafoon M. 2013. Alpha-CaMKII plays a critical role in determining the aggressive behavior of human osteosarcoma. Mol. Cancer Res. 11, 349–359. ( 10.1158/1541-7786.MCR-12-0572) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Cuddapah VA, Sontheimer H. 2010. Molecular interaction and functional regulation of ClC-3 by Ca2+/calmodulin-dependent protein kinase II (CaMKII) in human malignant glioma. J. Biol. Chem. 285, 11 188–11 196. ( 10.1074/jbc.O110.000234) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Giannone G, Ronde P, Gaire M, Haiech J, Takeda K. 2002. Calcium oscillations trigger focal adhesion disassembly in human U87 astrocytoma cells. J. Biol. Chem. 277, 26 364–26 371. ( 10.1074/jbc.M203952200) [DOI] [PubMed] [Google Scholar]
  • 50.Hahn K, DeBiasio R, Taylor DL. 1992. Patterns of elevated free calcium and calmodulin activation in living cells. Nature 359, 736–738. ( 10.1038/359736a0) [DOI] [PubMed] [Google Scholar]
  • 51.Schwab A, Finsterwalder F, Kersting U, Danker T, Oberleithner H. 1997. Intracellular Ca2+ distribution in migrating transformed epithelial cells. Pflügers Arch. 434, 70–76. ( 10.1007/s004240050364) [DOI] [PubMed] [Google Scholar]
  • 52.Fabian A, Fortmann T, Dieterich P, Riethmüller C, Schön P, Mally S, Nilius B, Schwab A. 2008. TRPC1 channels regulate directionality of migrating cells. Pflügers Arch. 457, 475–484. ( 10.1007/s00424-008-0515-4) [DOI] [PubMed] [Google Scholar]
  • 53.Wei C, Wang X, Chen M, Ouyang K, Song LS, Cheng H. 2009. Calcium flickers steer cell migration. Nature 457, 901–905. ( 10.1038/nature07577) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54.Giannone G, Ronde P, Gaire M, Beaudouin J, Haiech J, Ellenberg J, Takeda K. 2004. Calcium rises locally trigger focal adhesion disassembly and enhance residency of focal adhesion kinase at focal adhesions. J. Biol. Chem. 279, 28 715–28 723. ( 10.1074/jbc.M404054200) [DOI] [PubMed] [Google Scholar]
  • 55.Espinosa-Tanguma R, O'Neil C, Chrones T, Pickering JG, Sims SM. 2011. Essential role for calcium waves in migration of human vascular smooth muscle cells. Am. J. Physiol. Heart Circ. Physiol. 301, H315–H323. ( 10.1152/ajpheart.00355.2010) [DOI] [PubMed] [Google Scholar]
  • 56.Hamill KJ, Hopkinson SB, Skalli O, Jones JC. 2013. Actinin-4 in keratinocytes regulates motility via an effect on lamellipodia stability and matrix adhesions. FASEB J. 27, 546–556. ( 10.1096/fj.12-217406) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Witke W, Hofmann A, Koppel B, Schleicher M, Noegel AA. 1993. The Ca2+-binding domains in non-muscle type alpha-actinin: biochemical and genetic analysis. J. Cell Biol. 121, 599–606. ( 10.1083/jcb.121.3.599) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Eddy RJ, Pierini LM, Matsumura F, Maxfield FR. 2000. Ca2+-dependent myosin II activation is required for uropod retraction during neutrophil migration. J. Cell Sci. 113, 1287–1298. [DOI] [PubMed] [Google Scholar]
  • 59.Dingsdale H, Okeke E, Awais M, Haynes L, Criddle DN, Sutton R, Tepikin A. 2013. Saltatory formation, sliding and dissolution of ER-PM junctions in migrating cancer cells. Biochem. J. 451, 25–32. ( 10.1042/BJ20121864) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Lindemann O, et al. 2013. TRPC6 regulates CXCR2-mediated chemotaxis of murine neutrophils. J. Immunol. 190, 5496–5505. ( 10.4049/jimmunol.1201502) [DOI] [PubMed] [Google Scholar]
  • 61.Ijichi H, et al. 2011. Inhibiting Cxcr2 disrupts tumor-stromal interactions and improves survival in a mouse model of pancreatic ductal adenocarcinoma. J. Clin. Invest. 121, 4106–4117. ( 10.1172/JCI42754) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 62.Sciaccaluga M, et al. 2010. CXCL12-induced glioblastoma cell migration requires intermediate conductance Ca2+-activated K+ channel activity. Am. J. Physiol. Cell Physiol. 299, C175–C184. ( 10.1152/ajpcell.00344.2009) [DOI] [PubMed] [Google Scholar]
  • 63.Cuddapah VA, Turner KL, Seifert S, Sontheimer H. 2013. Bradykinin-induced chemotaxis of human gliomas requires the activation of KCa3.1 and ClC-3. J. Neurosci. 33, 1427–1440. ( 10.1523/JNEUROSCI.3980-12.2013) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Fabian A, Fortmann T, Bulk E, Bomben VC, Sontheimer H, Schwab A. 2011. Chemotaxis of MDCK-F cells toward fibroblast growth factor-2 depends on transient receptor potential canonical channel 1. Pflügers Arch. 461, 295–306. ( 10.1007/s00424-010-0901-6) [DOI] [PubMed] [Google Scholar]
  • 65.Bomben VC, Turner KL, Barclay TT, Sontheimer H. 2011. Transient receptor potential canonical channels are essential for chemotactic migration of human malignant gliomas. J. Cell Physiol. 226, 1879–1888. ( 10.1002/jcp.22518) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 66.Lehenkari PP, Horton MA. 1999. Single integrin molecule adhesion forces in intact cells measured by atomic force microscopy. Biochem. Biophys. Res. Commun. 259, 645–650. ( 10.1006/bbrc.1999.0827) [DOI] [PubMed] [Google Scholar]
  • 67.Paradise RK, Lauffenburger DA, Van Vliet KJ. 2011. Acidic extracellular pH promotes activation of integrin αvβ3. PLoS ONE 6, e15746 ( 10.1371/journal.pone.0015746) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Beaumont KG, Mrksich M. 2012. The mechanostability of isolated focal adhesions is strongly dependent on pH. Chem. Biol. 19, 711–720. ( 10.1016/j.chembiol.2012.04.016) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Erlandsen SL, Greet Bittermann A, White J, Leith A, Marko M. 2001. High-resolution CryoFESEM of individual cell adhesion molecules (CAMs) in the glycocalyx of human platelets: detection of P-selectin (CD62P), GPI-IX complex (CD42a/CD42b α,b β), and integrin GPIIbIIIa (CD41/CD61) by immunogold labeling and stereo imaging. J. Histochem. Cytochem. 49, 809–819. ( 10.1177/002215540104900702) [DOI] [PubMed] [Google Scholar]
  • 70.Ludwig FT, Schwab A, Stock C. 2013. The Na+/H+ -exchanger (NHE1) generates pH nanodomains at focal adhesions. J. Cell Physiol. 228, 1351–1358. ( 10.1002/jcp.24293) [DOI] [PubMed] [Google Scholar]
  • 71.Rojas JD, et al. 2006. Vacuolar-type H+-ATPases at the plasma membrane regulate pH and cell migration in microvascular endothelial cells. Am. J. Physiol. Heart Circ. Physiol. 291, H1147–H1157. ( 10.1152/ajpheart.00166.2006) [DOI] [PubMed] [Google Scholar]
  • 72.Schäfer C, Rymarczyk G, Ding L, Kirber MT, Bolotina VM. 2012. Role of molecular determinants of store-operated Ca2+ entry (Orai1, phospholipase A2 group 6, and STIM1) in focal adhesion formation and cell migration. J. Biol. Chem. 287, 40 745–40 757. ( 10.1074/jbc.M112.407155) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 73.Svensson L, McDowall A, Giles KM, Stanley P, Feske S, Hogg N. 2010. Calpain 2 controls turnover of LFA-1 adhesions on migrating T lymphocytes. PLoS ONE 5, e15090 ( 10.1371/journal.pone.0015090) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 74.Chan KT, Bennin DA, Huttenlocher A. 2010. Regulation of adhesion dynamics by calpain-mediated proteolysis of focal adhesion kinase (FAK). J. Biol. Chem. 285, 11 418–11 426. ( 10.1074/jbc.M109.090746) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Murphy DA, Courtneidge SA. 2011. The 'ins’ and 'outs’ of podosomes and invadopodia: characteristics, formation and function. Nat. Rev. Mol. Cell Biol. 12, 413–426. ( 10.1038/nrm3141) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 76.Siddiqui TA, Lively S, Vincent C, Schlichter LC. 2012. Regulation of podosome formation, microglial migration and invasion by Ca2+-signaling molecules expressed in podosomes. J. Neuroinflamm. 9, 250 ( 10.1186/1742-2094-9-250) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Lee J, Ishihara A, Oxford G, Johnson B, Jacobson K. 1999. Regulation of cell movement is mediated by stretch-activated calcium channels. Nature 400, 382–386. ( 10.1038/22578) [DOI] [PubMed] [Google Scholar]
  • 78.Fabian A, Bertrand J, Lindemann O, Pap T, Schwab A. 2012. Transient receptor potential canonical channel 1 impacts on mechanosignaling during cell migration. Pflügers Arch. 464, 623–630. ( 10.1007/s00424-012-1169-9) [DOI] [PubMed] [Google Scholar]
  • 79.Matthews BD, Thodeti CK, Tytell JD, Mammoto A, Overby DR, Ingber DE. 2010. Ultra-rapid activation of TRPV4 ion channels by mechanical forces applied to cell surface β1 integrins. Integr. Biol. (Camb.) 2, 435–442. ( 10.1039/c0ib00034e) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Schwab A, Wojnowski L, Gabriel K, Oberleithner H. 1994. Oscillating activity of a Ca2+-sensitive K+ channel. A prerequisite for migration of transformed Madin–Darby canine kidney focus cells. J. Clin. Invest. 93, 1631–1636. ( 10.1172/JCI117144) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 81.Potier M, Joulin V, Roger S, Besson P, Jourdan ML, Leguennec JY, Bougnoux P, Vandier C. 2006. Identification of SK3 channel as a new mediator of breast cancer cell migration. Mol. Cancer Ther. 5, 2946–2953. ( 10.1158/1535-7163.MCT-06-0194) [DOI] [PubMed] [Google Scholar]
  • 82.Cuddapah VA, Turner KL, Sontheimer H. 2013. Calcium entry via TRPC1 channels activates chloride currents in human glioma cells. Cell Calcium 53, 187–194. ( 10.1016/j.ceca.2012.11.013) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 83.Ayoub C, et al. 2010. ANO1 amplification and expression in HNSCC with a high propensity for future distant metastasis and its functions in HNSCC cell lines. Br. J. Cancer 103, 715–726. ( 10.1038/sj.bjc.6605823) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 84.Ruiz C, et al. 2012. Enhanced expression of ANO1 in head and neck squamous cell carcinoma causes cell migration and correlates with poor prognosis. PLoS ONE 7, e43265 ( 10.1371/journal.pone.0043265) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Jacobsen KS, Zeeberg K, Sauter DR, Poulsen KA, Hoffmann EK, Schwab A. 2013. The role of TMEM16A (ANO1) and TMEM16F (ANO6) in cell migration. Pflügers Arch. 465, 1753–1762. ( 10.1007/s00424-013-1315-z) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 86.Hoffmann EK, Lambert IH, Pedersen SF. 2009. Physiology of cell volume regulation in vertebrates. Physiol. Rev. 89, 193–277. ( 10.1152/physrev.00037.2007) [DOI] [PubMed] [Google Scholar]
  • 87.Soroceanu L, Manning TJ, Jr, Sontheimer H. 1999. Modulation of glioma cell migration and invasion using Cl and K+ ion channel blockers. J. Neurosci. 19, 5942–5954. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Klein M, Seeger P, Schuricht B, Alper SL, Schwab A. 2000. Polarization of Na+/H+ and Cl/HCO3 exchangers in migrating renal epithelial cells. J. Gen. Physiol. 115, 599–608. ( 10.1085/jgp.115.5.599) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Schneider SW, Pagel P, Rotsch C, Danker T, Oberleithner H, Radmacher M, Schwab A. 2000. Volume dynamics in migrating epithelial cells measured with atomic force microscopy. Pflügers Arch. 439, 297–303. ( 10.1007/s004240050943) [DOI] [PubMed] [Google Scholar]
  • 90.Happel P, Moller K, Kunz R, Dietzel ID. 2010. A boundary delimitation algorithm to approximate cell soma volumes of bipolar cells from topographical data obtained by scanning probe microscopy. BMC Bioinform. 11, 323 ( 10.1186/1471-2105-11-323) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Happel P, Moller K, Schwering NK, Dietzel ID. 2013. Migrating oligodendrocyte progenitor cells swell prior to soma dislocation. Sci. Rep. 3, 1806. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Watkins S, Sontheimer H. 2011. Hydrodynamic cellular volume changes enable glioma cell invasion. J. Neurosci. 31, 17 250–17 259. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 93.Schneider L, Klausen TK, Stock C, Mally S, Christensen ST, Pedersen SF, Hoffmann EK, Schwab A. 2008. H-ras transformation sensitizes volume-activated anion channels and increases migratory activity of NIH3T3 fibroblasts. Pflügers Arch. 455, 1055–1062. ( 10.1007/s00424-007-0367-3) [DOI] [PubMed] [Google Scholar]
  • 94.Saadoun S, Papadopoulos MC, Watanabe H, Yan D, Manley GT, Verkman AS. 2005. Involvement of aquaporin-4 in astroglial cell migration and glial scar formation. J. Cell Sci. 118, 5691–5698. ( 10.1242/jcs.02680) [DOI] [PubMed] [Google Scholar]
  • 95.Stock C, Cardone RA, Busco G, Krahling H, Schwab A, Reshkin SJ. 2008. Protons extruded by NHE1: digestive or glue? Eur. J. Cell Biol. 87, 591–599. ( 10.1016/j.ejcb.2008.01.007) [DOI] [PubMed] [Google Scholar]
  • 96.Busco G, et al. 2010. NHE1 promotes invadopodial ECM proteolysis through acidification of the peri-invadopodial space. FASEB J. 24, 3903–3915. ( 10.1096/fj.09-149518) [DOI] [PubMed] [Google Scholar]
  • 97.Monaco S, Gioia M, Rodriguez J, Fasciglione GF, Di Pierro D, Lupidi G, Krippahl L, Marini S, Coletta M. 2007. Modulation of the proteolytic activity of matrix metalloproteinase-2 (gelatinase A) on fibrinogen. Biochem. J. 402, 503–513. ( 10.1042/BJ20061064) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Kato Y, et al. 2005. Acidic extracellular pH induces matrix metalloproteinase-9 expression in mouse metastatic melanoma cells through the phospholipase D-mitogen-activated protein kinase signaling. J. Biol. Chem. 280, 10 938–10 944. ( 10.1074/jbc.M411313200) [DOI] [PubMed] [Google Scholar]
  • 99.Fraser SP, et al. 2005. Voltage-gated sodium channel expression and potentiation of human breast cancer metastasis. Clin. Cancer Res. 11, 5381–5389. ( 10.1158/1078-0432.CCR-05-0327) [DOI] [PubMed] [Google Scholar]
  • 100.Brisson L, Gillet L, Calaghan S, Besson P, Le Guennec JY, Roger S, Gore J. 2011. NaV1.5 enhances breast cancer cell invasiveness by increasing NHE1-dependent H+ efflux in caveolae. Oncogene 30, 2070–2076. ( 10.1038/onc.2010.574) [DOI] [PubMed] [Google Scholar]
  • 101.Gillet L, Roger S, Besson P, Lecaille F, Gore J, Bougnoux P, Lalmanach G, Le Guennec J-Y. 2009. Voltage-gated sodium channel activity promotes cysteine cathepsin-dependent invasiveness and colony growth of human cancer cells. J. Biol. Chem. 284, 8680–8691. ( 10.1074/jbc.M806891200) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 102.Yang S, Zhang JJ, Huang XY. 2009. Orai1 and STIM1 are critical for breast tumor cell migration and metastasis. Cancer Cell 15, 124–134. ( 10.1016/j.ccr.2008.12.019) [DOI] [PubMed] [Google Scholar]
  • 103.Rybarczyk P, et al. 2012. Transient receptor potential melastatin-related 7 channel is overexpressed in human pancreatic ductal adenocarcinomas and regulates human pancreatic cancer cell migration. Int. J. Cancer 131, E851–E861. ( 10.1002/ijc.27487) [DOI] [PubMed] [Google Scholar]
  • 104.Motiani RK, Hyzinski-Garcia MC, Zhang X, Henkel MM, Abdullaev IF, Kuo YH, Matrougui K, Mongin AA, Trebak M. 2013. STIM1 and Orai1 mediate CRAC channel activity and are essential for human glioblastoma invasion. Pflügers Arch. 465, 1249–1260. ( 10.1007/s00424-013-1254-8) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 105.Okamoto Y, Ohkubo T, Ikebe T, Yamazaki J. 2012. Blockade of TRPM8 activity reduces the invasion potential of oral squamous carcinoma cell lines. Int. J. Oncol. 40, 1431–1440. [DOI] [PubMed] [Google Scholar]
  • 106.Monet M, et al. 2010. Role of cationic channel TRPV2 in promoting prostate cancer migration and progression to androgen resistance. Cancer Res. 70, 1225–1235. ( 10.1158/0008-5472.CAN-09-2205) [DOI] [PubMed] [Google Scholar]
  • 107.Kato Y, Ozawa S, Tsukuda M, Kubota E, Miyazaki K, St-Pierre Y, Hata R-I. 2007. Acidic extracellular pH increases calcium influx-triggered phospholipase D activity along with acidic sphingomyelinase activation to induce matrix metalloproteinase-9 expression in mouse metastatic melanoma. FEBS J. 274, 3171–3183. ( 10.1111/j.1742-4658.2007.05848.x) [DOI] [PubMed] [Google Scholar]
  • 108.Cuddapah VA, Sontheimer H. 2011. Ion channels and transporters [corrected] in cancer. 2. Ion channels and the control of cancer cell migration. Am. J. Physiol. Cell Physiol. 301, C541–C549. ( 10.1152/ajpcell.00102.2011) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Haas BR, Sontheimer H. 2010. Inhibition of the sodium-potassium-chloride cotransporter isoform-1 reduces glioma invasion. Cancer Res. 70, 5597–5606. ( 10.1158/0008-5472.CAN-09-4666) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Chen XM, O'Hara SP, Huang BQ, Splinter PL, Nelson JB, LaRusso NF. 2005. Localized glucose and water influx facilitates Cryptosporidium parvum cellular invasion by means of modulation of host-cell membrane protrusion. Proc. Natl Acad. Sci. USA 102, 6338–6343. ( 10.1073/pnas.0408563102) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 111.Szablewski L. 2013. Expression of glucose transporters in cancers. Biochim. Biophys. Acta 1835, 164–169. [DOI] [PubMed] [Google Scholar]
  • 112.Fan J, Zhou JQ, Yu GR, Lu DD. 2010. Glucose transporter protein 1-targeted RNA interference inhibits growth and invasion of the osteosarcoma cell line MG63 in vitro. Cancer Biother. Radiopharm. 25, 521–527. ( 10.1089/cbr.2010.0784) [DOI] [PubMed] [Google Scholar]
  • 113.Machida Y, Ueda Y, Shimasaki M, Sato K, Sagawa M, Katsuda S, Sakuma T. 2011. Relationship of aquaporin 1, 3, and 5 expression in lung cancer cells to cellular differentiation, invasive growth, and metastasis potential. Hum. Pathol. 42, 669–678. ( 10.1016/j.humpath.2010.07.022) [DOI] [PubMed] [Google Scholar]
  • 114.Hammadi M, Chopin V, Matifat F, Dhennin-Duthille I, Chasseraud M, Sevestre H, Ouadid-Ahidouch H. 2012. Human ether a-gogo K+ channel 1 (hEag1) regulates MDA-MB-231 breast cancer cell migration through Orai1-dependent calcium entry. J. Cell Physiol. 227, 3837–3846. ( 10.1002/jcp.24095) [DOI] [PubMed] [Google Scholar]
  • 115.Chantome A, et al. 2013. Pivotal role of the lipid raft SK3-Orai1 complex in human cancer cell migration and bone metastases. Cancer Res. 73, 4852–4861. ( 10.1158/0008-5472.CAN-12-4572) [DOI] [PubMed] [Google Scholar]
  • 116.Labelle M, Hynes RO. 2013. The initial hours of metastasis: the importance of cooperative host-tumor cell interactions during hematogenous dissemination. Cancer Discov. 2, 1091–1099. ( 10.1158/2159-8290.CD-12-0329) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 117.Schaff UY, Dixit N, Procyk E, Yamayoshi I, Tse T, Simon SI. 2010. Orai1 regulates intracellular calcium, arrest, and shape polarization during neutrophil recruitment in shear flow. Blood 115, 657–666. ( 10.1182/blood-2009-05-224659) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 118.Pillozzi S, et al. 2007. VEGFR-1 (FLT-1), beta1 integrin, and hERG K+ channel for a macromolecular signaling complex in acute myeloid leukemia: role in cell migration and clinical outcome. Blood 110, 1238–1250. ( 10.1182/blood-2006-02-003772) [DOI] [PubMed] [Google Scholar]
  • 119.Pardo LA, del Camino D, Sanchez A, Alves F, Bruggemann A, Beckh S, Stühmer W. 1999. Oncogenic potential of EAG K+ channels. EMBO J. 18, 5540–5547. ( 10.1093/emboj/18.20.5540) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.Schwab A, Reinhardt J, Schneider SW, Gassner B, Schuricht B. 1999. K+ channel-dependent migration of fibroblasts and human melanoma cells. Cell Physiol. Biochem. 9, 126–132. ( 10.1159/000016309) [DOI] [PubMed] [Google Scholar]
  • 121.Ouadid-Ahidouch H, Roudbaraki M, Delcourt P, Ahidouch A, Joury N, Prevarskaya N. 2004. Functional and molecular identification of intermediate-conductance Ca2+ activated K+ channels in breast cancer cells: association with cell cycle progression. Am. J. Physiol. Cell Physiol. 287, C125–C134. ( 10.1152/ajpcell.00488.2003) [DOI] [PubMed] [Google Scholar]
  • 122.Griffith JW, Luster AD. 2013. Targeting cells in motion: Migrating toward improved therapies. Eur. J. Immunol. 43, 1430–1435. ( 10.1002/eji.201243183) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 123.Ataga KI, et al. 2011. Improvements in haemolysis and indicators of erythrocyte survival do not correlate with acute vaso-occlusive crises in patients with sickle cell disease: a phase III randomized, placebo-controlled, double-blind study of the Gardos channel blocker senicapoc (ICA-17043). Br. J. Haematol. 153, 92–104. ( 10.1111/j.1365-2141.2010.08520.x) [DOI] [PubMed] [Google Scholar]
  • 124.Mentzer RM, Jr, et al. 2008. Sodium-hydrogen exchange inhibition by cariporide to reduce the risk of ischemic cardiac events in patients undergoing coronary artery bypass grafting: results of the EXPEDITION study. Ann. Thorac. Surg. 85, 1261–1270. ( 10.1016/j.athoracsur.2007.10.054) [DOI] [PubMed] [Google Scholar]
  • 125.Heaney A, Buggy DJ. 2012. Can anaesthetic and analgesic techniques affect cancer recurrence or metastasis? Br. J. Anaesth. 109(Suppl. 1), i17–i28. ( 10.1093/bja/aes421) [DOI] [PubMed] [Google Scholar]
  • 126.Matthews H, Ranson M, Kelso MJ. 2011. Anti-tumour/metastasis effects of the potassium-sparing diuretic amiloride: an orally active anti-cancer drug waiting for its call-of-duty? Int. J. Cancer 129, 2051–2061. ( 10.1002/ijc.26156) [DOI] [PubMed] [Google Scholar]
  • 127.Oosterwijk-Wakka JC, Boerman OC, Mulders PFA, Oosterwijk E. 2013. Application of monoclonal antibody G250 recognizing carbonic anhydrase ix in renal cell carcinoma. Int. J. Mol. Sci. 14, 11 402–11 423. ( 10.3390/ijms140611402) [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Philosophical Transactions of the Royal Society B: Biological Sciences are provided here courtesy of The Royal Society

RESOURCES