TRP Channels in Cancer

Chapter 22


TRP Channels in Cancer



Artem Kondratskyi1; Natalia Prevarskaya1,*    1 Inserm U-1003, Equipe labellisée par la Ligue Nationale contre le cancer, Laboratory of Excellence Ion Channels Science and Therapeutics, Université Lille 1, Villeneuve d’Ascq, France
* Corresponding author: Natacha.Prevarskaya@univ-lille1.fr




Introduction


Cancer is a group of diseases characterized by out-of-control growth and dissemination of abnormal cells. Cancer represents one of the leading causes of death worldwide, accounting for 8.2 million deaths in 2012 [1]. Both external (radiation, tobacco, car exhaust fumes, and other carcinogens) and internal factors (inherited mutations, hormones, immune conditions) could be the causes of cancer. Moreover, these factors may act together or in sequence to initiate or promote the development of cancer [2]. There are more than 100 different types of cancer, including breast cancer, skin cancer, lung cancer, colon cancer, prostate cancer, lymphoma, and others. Different cancers are characterized by the different set of mutated genes (either oncogenes or tumor suppressor genes) and thus behave differently. However, Hanahan and Weinberg have proposed several general hallmarks, characteristic to most forms of cancer [3]. These hallmarks include sustaining proliferative signaling, evading growth suppressors, resisting cell death, enabling replicative immortality, inducing angiogenesis, activating invasion and metastasis, avoiding immune destruction, deregulating cellular energetics, tumor-promoting inflammation as well as genome instability and mutation [3].


Interestingly, calcium signaling has been previously reported to regulate most if not all of these phenomena [4,5]. Changes in the cytosolic-free Ca2 + concentration play a central role in many fundamental cellular processes, including muscle contraction, transmitter release, cell proliferation, differentiation, gene transcription, and cell death [6]. Given that Ca2 + controls so many vital processes, disturbance of the Ca2 + homeostasis regulatory mechanisms leads to a vast variety of severe pathologies, including cancer. Indeed, the role of Ca2 + is well established in many cell signaling pathways involved in carcinogenesis [79].


The increase in cytosolic calcium can occur as a result of Ca2 + influx from the extracellular space and/or Ca2 + release from intracellular sources. Both Ca2 + influx and Ca2 + release are tightly controlled by numerous regulatory systems that provide the specific spatial and temporal characteristics of an intracellular calcium signal that are required for sustaining certain cellular functions [6].


Mitochondrial, endoplasmic reticulum (ER), lysosomal, and cytosolic calcium levels are regulated by calcium-permeable ion channels localized either on the membranes of the intracellular organelles or on the plasma membrane [10,11]. The calcium-permeable channels, including families of transient receptor potential (TRP) channels, store-operated channels, voltage-gated calcium channels, two-pore channels, mitochondrial permeability transition pore, mitochondrial calcium uniporter, inositol 1,4,5-trisphosphate receptor (IP3) and ryanodine receptors, and others contribute to changes in [Ca2 +]i by providing Ca2 + entry pathways, by modulating the driving force for the Ca2 + entry, and also by providing intracellular pathways for Ca2 + uptake/release into/from cellular organelles [1013]. Thus, modulation of a calcium-permeable ion channel’s expression/function affects intracellular Ca2 + concentrations and consequently calcium-dependent processes, such as proliferation, apoptosis, and autophagy [1416].


The growing number of studies suggests that malignant transformation is often accompanied by the changes in ion channel expression/function [17]. Among these channels, the members of the TRP channels superfamily have been shown to regulate a variety of calcium-dependent cellular processes altered in cancer [18].


In this chapter, we will summarize the information available on the role of TRP channels in different cancers and discuss the possible use of these channels as biomarkers and therapeutic targets to treat cancer.


The Role of TRP Channels in Cancer


Among 28 members of the mammalian TRP superfamily grouped into six subfamilies (TRPC (canonical), TRPM (melastatin), TRPV (valinoid), TRPA (ankyrin-like), TRPP (polycystin), and TRPML (mucolipin), over 10 members were reported to have a role in cancer [18]. Further, we discuss specific roles of various TRP channels in different cancers, categorized by TRP subfamily.


TRPM Channels


TRPM1 (melastatin), the founding member of TRPM subfamily, was originally identified as a melanoma metastasis suppressor gene based on its decreased expression in metastatic mouse melanoma cell lines [19]. TRPM1 is involved in the regulation of calcium homeostasis in melanocytes and mediate an endogenous current that can be suppressed by TRPM1 knockdown [20,21]. Moreover, it was shown that ultraviolet B radiation could regulate the expression of TRPM1 [20]. Also, TRPM1 expression has been linked to melanocyte proliferation and differentiation in vitro. TRPM1 expression is significantly lower in rapidly proliferating melanocytes compared to the differentiated melanocytes [20].


TRPM1 is alternatively spliced, resulting in the production of long and short isoforms [2123]. Full-length TRPM1 (1533 amino acids) localizes to the plasma membrane and mediates Ca2 + entry when expressed in human embryonic kidney cells, whereas a cytosolic N-terminal truncation variant (500 amino acids) interacts directly with and suppresses the activity of full-length TRPM1. This suppression seems to result from the inhibition of translocation of TRPM1 to the plasma membrane [24]. Thus, alternative splicing of TRPM1 can potentially affect melanoma progression.


It was reported that TRPM1 is expressed at high levels in poorly metastatic variants of B16-F1 melanoma cell line, and its expression was much less in highly metastatic B16-F10 melanoma cell line [19]. Moreover, Duncan and colleagues demonstrated high TRPM1 levels in normal human epidermal melanocytes and benign melanocytic nevi, whereas in primary melanomas TRPM1 expression was decreased with virtually no expression detected in metastatic melanoma [19]. Further, TRPM1 mRNA expression appears to inversely correlate with melanocytic tumor progression, melanoma tumor thickness, and the potential for melanoma metastasis [22,25,26]. These patterns of TRPM1 expression have led to the suggestion that TRPM1 could serve as a valuable diagnostic and prognostic marker for the development and progression of melanomas [27].


TRPM2 is a chanzyme, that is, an ion channel with enzyme activity. Its structure contains an enzymatic region with ADP-ribose (ADPR)-hydrolase activity [28]. TRPM2 is widely expressed in mammalian cell types, including neurons [29], immune cells [30], insulin-secreting cells [31], and neutrophil granulocytes [32]. It is thought that the principal physiological role of TRPM2 is to control cytokine release in human monocytes, including tumor necrosis factor-alpha (TNFα) [33], insulin secretion [31], and others. TRPM2 is activated by a number of signals, such as intracellular ADPR, hydrogen peroxide, and TNFα [3436]. One of the most remarkable features of TRPM2 is its modulation by oxidative stress [34].


Recently it has been shown that TRPM2 could play a role in different cancers. It has been reported that insertion of TRPM2 into human glioblastoma cells facilitates hydrogen peroxide-induced cell death with no effect on proliferation, migration, and invasion [37]. In melanoma cell lines, TRPM2 locus has been found to transcribe TRPM2-AS (antisense) and TRPM2-TE (tumor-enriched dominant negative) transcripts, in addition to full-length TRPM2 [38]. Both TRPM2-AS and TRPM2-TE transcripts are up-regulated in melanoma and several other cancers, and thus, they down-regulate full-length TRPM2. Functional knockout of TRPM2-TE or overexpression of wild-type TRPM2 increases melanoma susceptibility to apoptosis and necrosis. Thus, restoration of full-length TRPM2 activity in cancer could be an attractive therapeutic opportunity. Interestingly, a somewhat different role of TRPM2 was demonstrated in prostate cancer [39]. siRNA-mediated knockdown of TRPM2 inhibits the growth of prostate cancer cells but not of noncancerous cells. This is due to the difference in its subcellular localization. In noncancerous cells TRPM2 is homogenously located near the plasma membrane and in the cytoplasm, whereas in the cancerous cells, a significant amount of the TRPM2 protein is located in the nuclei. Thus, it was suggested that TRPM2 is essential for prostate cancer cell proliferation and may be a potential target for the selective treatment of prostate cancer [39].


TRPM4 and TRPM5 are closely related channels, which are impermeable for calcium, in contrast to all other TRPs [40,41]. TRPM4 is mostly expressed in the pancreas, heart, and placenta, whereas TRPM5 is detected in the lungs, testis, tongue, digestive system, and brain [42]. It was reported that TRPM4 levels are elevated in prostate cancer, diffuse large B-cell non-Hodgkin lymphoma, and cervical cancer [4345]. Further, it was demonstrated that TRPM4 enhances cell proliferation of HeLa cells, a cervical cancer-derived cell line, through up-regulation of the β-catenin signaling pathway [46].


In a recent study on the role of genetic polymorphisms in immune response genes in the development of childhood leukemia, Han and coauthors suggested that some genetic variants of TRPM5 are associated with the risk of childhood leukemia [47]. It is worthwhile to note that the Trpm5 gene is located on chromosome 11, aberrations of which have been linked to a variety of malignances including leukemia, ovarian cancer, and rhabdoid tumors [4850]. Moreover, TRPM5 mRNA was found to be expressed in a large proportion of Wilms’ tumors and rhabdomyosarcomas [49].


TRPM7 represents a Ca2 +-permeable nonselective cation channel with enzyme activity. It contains an atypical serine/threonine protein kinase within the C-terminal domain [51]. It is ubiquitously expressed and involved in the regulation of cellular magnesium homeostasis [52,53]. TRPM7 has been shown to regulate lymphocyte survival [54], anoxic neuronal cell death [55], cell volume in human epithelial cells [56], cell adhesion in neuroblastoma [57], as well as osteoblast proliferation and migration [58].


In cancer, TRPM7 has been reported to be overexpressed in high-grade breast cancer samples. Moreover, its expression positively correlates with tumor size [59], which makes it the putative diagnostic/prognostic marker for the development and progression of breast cancer. Silencing of TRPM7 was shown to inhibit growth and proliferation of human head and neck squamous carcinoma cell lines [60], human retinoblastoma cells [61], as well as breast cancer cells MCF-7 [59]. Recently it was reported that high levels of TRPM7 expression independently predict poor outcome in breast cancer patients and that it is functionally required for metastasis formation in a mouse xenograft model of human breast cancer [62]. Mechanistic studies revealed that TRPM7 regulates myosin II-based cellular tension, thereby modifying focal adhesion number, cell-cell adhesion, and polarized cell movement [62]. Another group of scientists suggested that TRPM7 mediates breast cancer cell migration and invasion through the mitogen-activated protein kinase (MAPK) pathway [63]. The importance of TRPM7 for human nasopharyngeal carcinoma cells migration [64], as well as for proliferation, migration, and invasion of pancreatic adenocarcinoma cells has also been reported [65].


TRPM8, another member of TRPM subfamily, was first cloned from the human prostate as a prostate-specific gene up-regulated in cancer [66]. Later, it was shown that TRPM8 has a major role in the cold sensation in sensory neurons [67,68]. In normal tissues, the expression of the channel could be detected in a subpopulation of primary sensory neurons [67,68], as well as in the male reproductive system [66,6971]. In situ hybridization analysis showed that TRPM8 mRNA expression was at moderate levels in normal prostate tissue and appears to be elevated in prostate cancer [66]. Other than prostate, TRPM8 mRNA was expressed in a number of primary tumors of breast, colon, lung, and skin origin, whereas transcripts encoding TRPM8 were hardly detected or not detected in the corresponding normal human tissues [66]. Moreover, a significant difference in the expression level of TRPM8 mRNA between malignant and nonmalignant tissue specimens has been detected in prostate cancer tumors [70]. It was proposed that in prostate cancer cells TRPM8 could form a calcium-permeable channel at both plasma membrane and ER [7275]. TRPM8 is subjected to alternative splicing, which generates several splice variants of the full-length and truncated forms. The full-length isoform is expressed at the plasma membrane, whereas the short isoform is localized to ER, where it forms a Ca2 +-release channel [72,74,75].


TRPM8 localization and activity was reported to be regulated depending on the differentiation and oncogenic status of prostate cancer cells [72,75]. Thus, only highly differentiated prostate cancer cells expressed functional plasmalemmal TRPM8 channels. This was further confirmed by the finding that the TRPM8-mediated current density was significantly higher in prostate cancer cells obtained from in situ prostate cancer biopsies than in normal prostate or benign prostate hyperplasia cells. However, following dedifferentiation and loss of the luminal secretory phenotype, activity of plasmalemmal TRPM8 was abolished. In contrast, the ER-TRPM8 remained functional regardless of the differentiation status of prostate cells [72]. This differential regulation of TRPM8 activity has been explained by the complex regulation of the TRPM8 isoforms by the androgen receptor [73,75]. Indeed, it has been suggested that TRPM8 is directly regulated by androgen receptor in both normal and cancer prostate cells [73,75]. It appears that Trpm8 gene expression requires a functional androgen receptor and is down-regulated in cells losing androgen sensitivity [73,76]. Moreover, overexpression of androgen receptor in PNT1A cells (human postpubertal prostate normal, immortalized with SV40), that normally lack it, induced TRPM8 expression [73].


TRPM8 overexpression and overactivity in androgen-dependent prostate cancer was reported to correlate with the increased cell growth [77]. On the other hand, antiandrogen therapy as well as transition to androgen independence is associated with the loss of TRPM8 in prostate cancer cells, suggesting that the loss of TRPM8 could be a potential marker of the transition to a more advanced form of the prostate cancer [76]. Furthermore, it has been shown that pharmacological activation of TRPM8 as well as silencing of the channel with siRNA in human prostate carcinoma LNCaP cells negatively influences cell survival, likely by perturbing Ca2 + homeostasis [75]. However, plasma membrane TRPM8 might have a protective effect because stimulation of TRPM8 activity by PSA (prostate-specific antigen) reduced prostate cancer cell motility [78]. Therefore, the gradual loss of the plasmalemmal TRPM8 during tumor progression to late, invasive stages may be an adaptive mechanism of prostate cancer epithelial cells to reduce constant stimulation of the PSA/TRPM8 pathway [78].


Along with prostate cancer, the role of TRPM8 has been also demonstrated in other cancers. Activation of TRPM8 with menthol was reported to decrease the viability of melanoma cells [79]. In breast cancer, TRPM8 is functional at the plasma membrane and expressed in early primitive breast cancers presenting a well-differentiated status [80]. Further, TRPM8 was shown to be overexpressed and required for proliferation of pancreatic adenocarcinoma cells [65]. In human bladder cancer cells, menthol induces cell death via the TRPM8 channel [81]. Thus, TRPM8 seems to have a great potential to be used as a diagnostic marker and/or therapeutic target in the treatment of different cancers.


TRPV Channels


The TRPV1 channel, also known as capsaicin receptor, was the first discovered TRPV channel. TRPV1 was originally identified in sensory neurons as a heat-activated ion channel, which functions as a transducer of painful thermal stimuli in vivo [82]. Increased expression of TRPV1 has been found in different cancers, including cancers of prostate, colon, pancreas, and bladder [8386]. High expression of TRPV1 was reported to be associated with a better prognosis of patients with hepatocellular carcinoma [87]. A progressive loss of TRPV1 expression has been found in the urothelium, as transitional cell carcinoma stage increased and cell differentiation was lower [85]. In glioma, TRPV1 expression was reported to inversely correlate with tumor grade with a complete loss observed in 93% of high grade (IV) glioblastomas [88]. Also, it was reported that TRPV1 suppresses skin carcinogenesis [89]. These results suggest that TRPV1 might function as a tumor suppressor, and down-regulation of TRPV1 mRNA expression may represent an independent negative prognostic factor for cancer patients. Moreover, TRPV1 has been proposed to be involved in cancer pain transduction [84,90].


TRPV2 exhibits 50% sequence identity to TRPV1 and is activated by noxious heat (> 53 °C) [91]. TRPV2 is also expressed in sensory neurons, as well as in other tissues, including smooth muscle cells and GI tract [12]. Overexpression of TRPV2 has been reported in several cancers and cancer cell lines [92]. In prostate cancer, TRPV2 expression levels are higher in metastatic cancers (stage M1) compared with primary solid tumors (stages T2a and T2b). This property makes TRPV2 a potential marker for aggressive advanced prostate cancer [93]. TRPV2 was also demonstrated to promote prostate cancer cell migration by regulating the expression of invasive enzymes MMP2, MMP9, and cathepsin B [93]. In liver cancers, TRPV2 demonstrates high expression levels in human hepatocarcinoma cells (HepG2) [94]. Expression of TRPV2 at both the mRNA and protein levels was also shown to be increased in cirrhotic livers compared with chronic hepatitis, suggesting that it might be a prognostic marker of patients with hepatocellular carcinoma [95]. Moderately and well-differentiated liver tumors were also characterized by the increased TRPV2 expression compared with that of poorly differentiated tumors [95]. In bladder cancer, enhanced TRPV2 mRNA and protein expression was found in high-grade urothelial carcinoma specimens and cell lines. Both the full-length TRPV2 and a short splice-variant were detected in normal human urothelial cells and normal bladder specimens, whereas a progressive loss of short-TRPV2 accompanied by a marked increase of full-length TRPV2 expression was found in high-grade urothelial carcinoma tissues [96]. TRPV2 mRNA was reported to be abundantly expressed in a poorly differentiated urothelial carcinoma T24 cells. Moreover, the TRPV2 agonist cannabidiol induced urothelial carcinoma cell death via apoptosis, suggesting that TRPV2 could serve as a potential therapeutic target for the treatment of bladder cancer [97]. In contrast, TRPV2 mRNA and protein expression progressively decline in high-grade glioma tissues as histological grade increases compared to benign astrocyte tissues [98]. Furthermore, TRPV2 channel negatively controls glioma cell survival and proliferation [98].


TRPV6 channel was first identified as a channel-like transporter mediating intestinal calcium absorption in the duodenum [99]. TRPV6 is expressed in a variety of tissues with predominant expression in the duodenum, jejunum, colon, and kidneys. The expression of TRPV6 is substantially increased in prostate cancer tissue as well as human carcinomas of the colon, breast, thyroid, and ovary (see Ref. [100] for review). In prostate cancer, TRPV6 expression correlates with tumor grade. TRPV6 mRNA levels are elevated in prostate cancer specimens in comparison to benign prostatic hyperplasia (BPH) specimens and positively correlate with the Gleason grade [101,102]. Thus, TRPV6 has been suggested as a prognostic marker and a promising target for new therapeutic strategies to treat advanced prostate cancer [101,102]. TRPV6 was also shown to be directly involved in the control of LNCaP cell proliferation, as siRNA-mediated TRPV6 silencing slowed down the proliferation rate, decreased the accumulation of LNCaP cells in the S phase of the cell cycle, and lowered the expression of the proliferating cell nuclear antigen [103]. An elevated expression of TRPV6 was also detected in colon carcinoma cells, where siRNA-mediated TRPV6 knockdown inhibited proliferation and induced apoptosis [104]. TRPV6 is mainly overexpressed in the invasive breast cancer cells and the selective silencing of TRPV6-inhibited human prostate carcinoma MDA-MB-231 cell migration and invasion, as well as MCF-7 migration [105]. Breast cancer cell proliferation was also reported to be regulated by TRPV6 [106]. Thus, it seems that in many cancers TRPV6 represents an oncogene; however, this is not true for all cancers. Indeed, no TRPV6 up-regulation was observed in pancreatic cancer. Furthermore, TRPV6 has been demonstrated to mediate capsaicin-induced apoptosis in gastric cancer cells [107].


TRPC Channels


TRPC1 is a nonselective ubiquitously expressed cation channel [108]. TRPC1 has been proposed to be involved in store-operated calcium entry in cooperation with Orai1 and STIM1 [109,110]. TRPC1 contributes to multiple important physiological processes and provides an important route for agonist-, growth factor-, and protein kinase C (PKC)-induced Ca2 + entry in a variety of cell types (see Ref. [12] for review). TRPC1 has been also shown to have a role in cancer cell migration and proliferation. Thus, the TRPC1 channel was reported to be essential for glioma cell chemotaxis in response to stimuli, such as Epidermal growth factor (EGF) [111]. Another report has shown that TRPC1 depletion induced cell growth arrest by causing incomplete cytokinesis in gliomas [112]. Further, it was suggested that extracellular signal-regulated kinases 1 and 2 and TRPC1 channels are required for calcium-sensing receptor-stimulated MCF-7 breast cancer cell proliferation [113]. The inhibitory effect of TRPC1 depletion on proliferation was also demonstrated in non-small-cell lung carcinoma cells [114]. TRPC1 has been also associated with proliferative phenotype in breast cancer and was demonstrated to be overexpressed in human breast cancer epithelial primary culture [105]. Recently, the role of TRPC1, C3, C4, and C6 channels in lung and ovarian cancers was demonstrated [115,116]. The expression of TRPC1, C3, C4, and C6 was correlated to the differentiation grade of both non-small-cell lung cancer and ovarian cancer. Moreover, siRNA-mediated knockdown of TRPC1, C3, C4, or C6 channels inhibited cell proliferation in both cancers [115,116].


Recently, a functional expression of TRPC3 has been demonstrated in MCF-7 breast cancer cells. Here, inhibition of TRPC3 by exogenous polyunsaturated fatty acids consistently attenuated breast cancer cell proliferation and migration [117]. Also, high levels of TRPC3 expression in tumor cells have been proposed as an independent predictor of a better prognosis in patients with adenocarcinoma of the lung [118].


The role of another member of TRPC subfamily, TRPC6, has been described in multiple cancers, including prostate [119,120], liver [121], renal [122], and breast cancer [123,124]. Thus, alpha1-adrenergic receptor signaling was reported to require the coupled activation of TRPC6 channels and nuclear factor of activated T-cells (NFAT) to promote proliferation of primary human prostate cancer epithelial cells, thereby suggesting TRPC6 as a novel potential therapeutic target [119]. In liver, TRPC6 expression levels were higher in tumors compared to normal hepatocytes. Moreover, TRPC6 overexpression stimulated proliferation of the human hepatoma cell line Huh-7 [121]. Expression of TRPC6 was demonstrated to be markedly increased in renal cell carcinoma specimens and was associated with renal cell carcinoma histological grade. The positive regulatory role of TRPC6 in proliferation of renal cell carcinoma ACHN cells has been also emphasized [122]. In breast carcinoma specimens, TRPC6 mRNA and protein were found to be elevated in comparison to normal breast tissue. However, no correlation was found between TRPC6 protein levels and tumor grades or metastases [124]. Moreover, activation of TRPC6 significantly reduced the growth and viability of the breast cancer cell lines [123].


Recently, it was reported that TRPC6 promotes cancer cell migration in head and neck squamous cell carcinomas (HNSCC) [125] and glioblastoma [126]. The increase in TRPC6 levels was demonstrated in HNSCC tumor samples and cancer cell lines. siRNA-mediated knockdown of TRPC6 expression in HNSCC-derived cells dramatically inhibited HNSCC-cell invasion but did not significantly alter cell proliferation [125]. In glioblastoma, TRPC6 has been shown to be required for the development of the aggressive phenotype because knockdown of TRPC6 expression inhibits glioma growth, invasion, and angiogenesis [126].


Clinical Implications and Concluding Remarks


It’s now clear that TRP channels play an important role in cancer-related processes, including proliferation, differentiation, apoptosis, migration, and angiogenesis. Thus, it’s not surprising that a number of TRP channels have been proposed as potential anticancer therapeutic targets and/or diagnostic tools.


Recently, synthetic vanilloid arvanil was proposed as therapeutics for high-grade astrocytoma [127]. By activating the ER-localized TRPV1 channel, which is significantly overexpressed in this type of cancer cells, it induces ER-stress thereby promoting tumor cell death [127]. In 2009, the small molecule TRPM8 agonist D-3263 was presented as a potential treatment for BPH alone or in combination with finasteride, a current treatment for BPH. Further, high-affinity TRPM8 agonists WS-12 and WS-12F (WS-12 compound with incorporated fluorine) were proposed as potential drugs for prostate cancer imaging and therapy [128]. Moreover, it has been suggested that incorporation of a radioisotope into WS-12 could allow radiodiagnostics as well as radiotherapy of prostate cancer with all the advantages of selective targeting of TRPM8-expressing cells [128]. At present, there is a huge need in selective pharmacological modulators for most of the TRP channels, and there are no doubts that design of such specific molecules will greatly advance the field toward practical clinical implications of the TRP channel-based anticancer therapies.


Given that calcium overload is detrimental to all cells, prolonged stimulation or overexpression of calcium-permeable channels in cancer cells will lead to substantial increase in cytoplasmic calcium followed by apoptotic/necrotic cell death [129]. In this respect, all the calcium-permeable TRP channels could potentially represent efficient anticancer therapeutic tools, provided that specific approaches will be implemented for selective overexpression (under control of tissue/cell type specific promoter) or targeting of the channel of interest. Indeed, most of the TRP channels have broad expression patterns and thus are not cancer-specific. Therefore, selective targeting is vitally important to prevent toxicity to normal cells induced by the pharmacological impairment of channel function.


Given that the expression level of certain TRP channels differs substantially between normal and cancerous tissues, these channels could be considered as potential diagnostic markers. For example, TRPV6 expression correlates with tumor grades in many tissues, which has led to the suggestion that TRPV6 represents a reliable tool to predict the clinical outcome of different cancers [100]. Nevertheless, further studies are needed to find out whether TRPV6 represents an oncogenic ion channel, or its expression is epigenetically altered during cancer progression. Furthermore, inverse correlation between TRPM1 mRNA expression and melanoma metastatic potential suggests that TRPM1 could serve as a diagnostic/prognostic marker for the development and progression of melanomas [27]. Expression levels of TRPM8 have been also reported to be useful when staging prostate cancer [70,72]. Other TRP channels, including TRPM7, TRPV1, TRPV2, TRPC1, and TRPC3, also show a great potential as diagnostic/prognostic markers in different cancers. However, it should be noted that the use of ion channels for diagnostic purposes involves tissue biopsies, as cellular ion channels normally cannot be detected in blood samples. Thus, the antibody-based approaches should be considered. Indeed, TRP channels localized on plasma membrane could be subjected to antibody-based targeting that can be particularly useful in the case of channel up-regulation in cancer. Fluorescently labeled antibodies could be used for tumor visualization, whereas radionuclide-labeled antibodies provide a powerful therapeutic tool [8].


Importantly, as the contribution of specific TRP channels to cancer-related processes could vary depending on cancer type and stage, the specific therapeutic/diagnostic approach should be implemented for each particular case. Thus, the targeting of TRP channels in cancer treatment should not be regarded as a general approach to treat cancer, but rather as a specific tool for individualized cancer therapy.


Overall, TRP channels have emerged as promising anticancer therapeutic targets and diagnostic/prognostic markers. However, despite the essential progress achieved in this field, there is still a need in highly selective pharmacological modulators, novel drug delivery techniques, as well as siRNA transfection technologies to selectively target TRP channels in cancer cells without significant adverse effects.



References


[1] Ferlay J, Soerjomataram I, Ervik M, Dikshit R, Eser S, Mathers C, et al. GLOBOCAN 2012 v1.0, cancer incidence and mortality worldwide: IARC Cancerbase no. 11 [internet]. Lyon, France: International Agency for Research on Cancer; 2013. Available from: http://globocan.iarc.fr accessed on 23/02/2014.


[2] American Cancer Society. Cancer facts & figures 2013. Atlanta: American Cancer Society; 2013.


[3] Hanahan D, Weinberg RA. Hallmarks of cancer: the next generation. Cell. 2011;144(5):646–674.


[4] Prevarskaya N, Ouadid-Ahidouch H, Skryma R, Shuba Y. Remodelling of Ca2 + transport in cancer: how it contributes to cancer hallmarks? Philos Trans R Soc Lond B Biol Sci. 2014;369(1638):20130097.


[5] Prevarskaya N, Skryma R, Shuba Y. Targeting Ca(2)(+) transport in cancer: close reality or long perspective? Expert Opin Ther Targets. 2013;17(3):225–241.


[6] Berridge MJ, Lipp P, Bootman MD. The versatility and universality of calcium signalling. Nat Rev Mol Cell Biol. 2000;1(1):11–21.


[7] Monteith GR, McAndrew D, Faddy HM, Roberts-Thomson SJ. Calcium and cancer: targeting Ca2 + transport. Nat Rev Cancer. 2007;7(7):519–530.


[8] Prevarskaya N, Skryma R, Shuba Y. Calcium in tumour metastasis: new roles for known actors. Nat Rev Cancer. 2011;11(8):609–618.


[9] Monteith GR, Davis FM, Roberts-Thomson SJ. Calcium channels and pumps in cancer: changes and consequences. J Biol Chem. 2012;287(38):31666–31673.


[10] Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol. 2003;4(7):517–529.


[11] Rizzuto R, De Stefani D, Raffaello A, Mammucari C. Mitochondria as sensors and regulators of calcium signalling. Nat Rev Mol Cell Biol. 2012;13(9):566–578.


[12] Pedersen SF, Owsianik G, Nilius B. TRP channels: an overview. Cell Calcium. 2005;38(3-4):233–252.


[13] Bernardi P, von Stockum S. The permeability transition pore as a Ca(2 +) release channel: new answers to an old question. Cell Calcium. 2012;52(1):22–27.


[14] Flourakis M, Prevarskaya N. Insights into Ca2 + homeostasis of advanced prostate cancer cells. Biochim Biophys Acta. 2009;1793(6):1105–1109.


[15] Decuypere JP, Bultynck G, Parys JB. A dual role for Ca(2 +) in autophagy regulation. Cell Calcium. 2011;50(3):242–250.


[16] Dubois C, Vanden Abeele F, Prevarskaya N. Targeting apoptosis by the remodelling of calcium-transporting proteins in cancerogenesis. FEBS J. 2013;280(21):5430–5440.


[17] Prevarskaya N, Skryma R, Shuba Y. Ion channels and the hallmarks of cancer. Trends Mol Med. 2010;16(3):107–121.


[18] Shapovalov G, Lehen’kyi V, Skryma R, Prevarskaya N. TRP channels in cell survival and cell death in normal and transformed cells. Cell Calcium. 2011;50(3):295–302.


[19] Duncan LM, Deeds J, Hunter J, Shao J, Holmgren LM, Woolf EA, et al. Down-regulation of the novel gene melastatin correlates with potential for melanoma metastasis. Cancer Res. 1998;58(7):1515–1520.


[20] Devi S, Kedlaya R, Maddodi N, Bhat KM, Weber CS, Valdivia H, et al. Calcium homeostasis in human melanocytes: role of transient receptor potential melastatin 1 (TRPM1) and its regulation by ultraviolet light. Am J Physiol Cell Physiol. 2009;297(3):C679–C687.


[21] Oancea E, Vriens J, Brauchi S, Jun J, Splawski I, Clapham DE. TRPM1 forms ion channels associated with melanin content in melanocytes. Sci Signal. 2009;2(70):ra21.


[22] Fang D, Setaluri V. Expression and up-regulation of alternatively spliced transcripts of melastatin, a melanoma metastasis-related gene, in human melanoma cells. Biochem Biophys Res Commun. 2000;279(1):53–61.


[23] Zhiqi S, Soltani MH, Bhat KM, Sangha N, Fang D, Hunter JJ, et al. Human melastatin 1 (TRPM1) is regulated by MITF and produces multiple polypeptide isoforms in melanocytes and melanoma. Melanoma Res. 2004;14(6):509–516.


[24] Xu XZ, Moebius F, Gill DL, Montell C. Regulation of melastatin, a TRP-related protein, through interaction with a cytoplasmic isoform. Proc Natl Acad Sci USA. 2001;98(19):10692–10697.


[25] Deeds J, Cronin F, Duncan LM. Patterns of melastatin mRNA expression in melanocytic tumors. Hum Pathol. 2000;31(11):1346–1356.


[26] Duncan LM, Deeds J, Cronin FE, Donovan M, Sober AJ, Kauffman M, et al. Melastatin expression and prognosis in cutaneous malignant melanoma. J Clin Oncol. 2001;19(2):568–576.


[27] Hammock L, Cohen C, Carlson G, Murray D, Ross JS, Sheehan C, et al. Chromogenic in situ hybridization analysis of melastatin mRNA expression in melanomas from American Joint Committee on Cancer stage I and II patients with recurrent melanoma. J Cutan Pathol. 2006;33(9):599–607.


[28] Perraud AL, Fleig A, Dunn CA, Bagley LA, Launay P, Schmitz C, et al. ADP-ribose gating of the calcium-permeable LTRPC2 channel revealed by Nudix motif homology. Nature. 2001;411(6837):595–599.


[29] Olah ME, Jackson MF, Li H, Perez Y, Sun HS, Kiyonaka S, et al. Ca2 + -dependent induction of TRPM2 currents in hippocampal neurons. J Physiol. 2009;587(Pt 5):965–979.


[30] Sano Y, Inamura K, Miyake A, Mochizuki S, Yokoi H, Matsushime H, et al. Immunocyte Ca2 + influx system mediated by LTRPC2. Science. 2001;293(5533):1327–1330.


[31] Togashi K, Hara Y, Tominaga T, Higashi T, Konishi Y, Mori Y, et al. TRPM2 activation by cyclic ADP-ribose at body temperature is involved in insulin secretion. EMBO J. 2006;25(9):1804–1815.


[32] Heiner I, Eisfeld J, Halaszovich CR, Wehage E, Jungling E, Zitt C, et al. Expression profile of the transient receptor potential (TRP) family in neutrophil granulocytes: evidence for currents through long TRP channel 2 induced by ADP-ribose and NAD. Biochem J. 2003;371(Pt 3):1045–1053.


[33] Wehrhahn J, Kraft R, Harteneck C, Hauschildt S. Transient receptor potential melastatin 2 is required for lipopolysaccharide-induced cytokine production in human monocytes. J Immunol. 2010;184(5):2386–2393.


[34] Hara Y, Wakamori M, Ishii M, Maeno E, Nishida M, Yoshida T, et al. LTRPC2 Ca2 + -permeable channel activated by changes in redox status confers susceptibility to cell death. Mol Cell. 2002;9(1):163–173.


[35] McHugh D, Flemming R, Xu SZ, Perraud AL, Beech DJ. Critical intracellular Ca2 + dependence of transient receptor potential melastatin 2 (TRPM2) cation channel activation. J Biol Chem. 2003;278(13):11002–11006.


[36] Wehage E, Eisfeld J, Heiner I, Jungling E, Zitt C, Luckhoff A. Activation of the cation channel long transient receptor potential channel 2 (LTRPC2) by hydrogen peroxide. A splice variant reveals a mode of activation independent of ADP-ribose. J Biol Chem. 2002;277(26):23150–23156.


[37] Ishii M, Oyama A, Hagiwara T, Miyazaki A, Mori Y, Kiuchi Y, et al. Facilitation of H2°2-induced A172 human glioblastoma cell death by insertion of oxidative stress-sensitive TRPM2 channels. Anticancer Res. 2007;27(6B):3987–3992.


[38] Orfanelli U, Wenke AK, Doglioni C, Russo V, Bosserhoff AK, Lavorgna G. Identification of novel sense and antisense transcription at the TRPM2 locus in cancer. Cell Res. 2008;18(11):1128–1140.


[39] Zeng X, Sikka SC, Huang L, Sun C, Xu C, Jia D, et al. Novel role for the transient receptor potential channel TRPM2 in prostate cancer cell proliferation. Prostate Cancer Prostatic Dis. 2010;13(2):195–201.


[40] Hofmann T, Chubanov V, Gudermann T, Montell C. TRPM5 is a voltage-modulated and Ca(2 +)-activated monovalent selective cation channel. Curr Biol. 2003;13(13):1153–1158.


[41] Launay P, Fleig A, Perraud AL, Scharenberg AM, Penner R, Kinet JP. TRPM4 is a Ca2+-activated nonselective cation channel mediating cell membrane depolarization. Cell. 2002;109(3):397–407.


[42] Ullrich ND, Voets T, Prenen J, Vennekens R, Talavera K, Droogmans G, et al. Comparison of functional properties of the Ca2 + -activated cation channels TRPM4 and TRPM5 from mice. Cell Calcium. 2005;37(3):267–278.


[43] Ashida S, Nakagawa H, Katagiri T, Furihata M, Iiizumi M, Anazawa Y, et al. Molecular features of the transition from prostatic intraepithelial neoplasia (PIN) to prostate cancer: genome-wide gene-expression profiles of prostate cancers and PINs. Cancer Res. 2004;64(17):5963–5972.


[44] Narayan G, Bourdon V, Chaganti S, Arias-Pulido H, Nandula SV, Rao PH, et al. Gene dosage alterations revealed by cDNA microarray analysis in cervical cancer: identification of candidate amplified and overexpressed genes. Genes Chromosomes Cancer. 2007;46(4):373–384.


[45] Suguro M, Tagawa H, Kagami Y, Okamoto M, Ohshima K, Shiku H, et al. Expression profiling analysis of the CD5 + diffuse large B-cell lymphoma subgroup: development of a CD5 signature. Cancer Sci. 2006;97(9):868–874.


[46] Armisen R, Marcelain K, Simon F, Tapia JC, Toro J, Quest AF, et al. TRPM4 enhances cell proliferation through up-regulation of the beta-catenin signaling pathway. J Cell Physiol. 2011;226(1):103–109.


[47] Han S, Koo HH, Lan Q, Lee KM, Park AK, Park SK, et al. Common variation in genes related to immune response and risk of childhood leukemia. Hum Immunol. 2012;73(3):316–319.


[48] Monni O, Knuutila S. 11q deletions in hematological malignancies. Leuk Lymphoma. 2001;40(3-4):259–266.


[49] Prawitt D, Enklaar T, Klemm G, Gartner B, Spangenberg C, Winterpacht A, et al. Identification and characterization of MTR1, a novel gene with homology to melastatin (MLSN1) and the trp gene family located in the BWS-WT2 critical region on chromosome 11p15.5 and showing allele-specific expression. Hum Mol Genet. 2000;9(2):203–216.


[50] Stronach EA, Sellar GC, Blenkiron C, Rabiasz GJ, Taylor KJ, Miller EP, et al. Identification of clinically relevant genes on chromosome 11 in a functional model of ovarian cancer tumor suppression. Cancer Res. 2003;63(24):8648–8655.


[51] Runnels LW, Yue L, Clapham DE. TRP-PLIK, a bifunctional protein with kinase and ion channel activities. Science. 2001;291(5506):1043–1047.


[52] Ryazanova LV, Rondon LJ, Zierler S, Hu Z, Galli J, Yamaguchi TP, et al. TRPM7 is essential for Mg(2 +) homeostasis in mammals. Nat Commun. 2010;1:109.


[53] Schmitz C, Perraud AL, Johnson CO, Inabe K, Smith MK, Penner R, et al. Regulation of vertebrate cellular Mg2 + homeostasis by TRPM7. Cell. 2003;114(2):191–200.


[54] Nadler MJ, Hermosura MC, Inabe K, Perraud AL, Zhu Q, Stokes AJ, et al. LTRPC7 is a Mg.ATP-regulated divalent cation channel required for cell viability. Nature. 2001;411(6837):590–595.


[55] Aarts M, Iihara K, Wei WL, Xiong ZG, Arundine M, Cerwinski W, et al. A key role for TRPM7 channels in anoxic neuronal death. Cell. 2003;115(7):863–877.


[56] Numata T, Shimizu T, Okada Y. TRPM7 is a stretch- and swelling-activated cation channel involved in volume regulation in human epithelial cells. Am J Physiol Cell Physiol. 2007;292(1):C460–C467.


[57] Clark K, Langeslag M, van Leeuwen B, Ran L, Ryazanov AG, Figdor CG, et al. TRPM7, a novel regulator of actomyosin contractility and cell adhesion. EMBO J. 2006;25(2):290–301.


[58] Abed E, Moreau R. Importance of melastatin-like transient receptor potential 7 and magnesium in the stimulation of osteoblast proliferation and migration by platelet-derived growth factor. Am J Physiol Cell Physiol. 2009;297(2):C360–C368.


[59] Guilbert A, Gautier M, Dhennin-Duthille I, Haren N, Sevestre H, Ouadid-Ahidouch H. Evidence that TRPM7 is required for breast cancer cell proliferation. Am J Physiol Cell Physiol. 2009;297(3):C493–C502.


[60] Jiang J, Li MH, Inoue K, Chu XP, Seeds J, Xiong ZG. Transient receptor potential melastatin 7-like current in human head and neck carcinoma cells: role in cell proliferation. Cancer Res. 2007;67(22):10929–10938.


[61] Hanano T, Hara Y, Shi J, Morita H, Umebayashi C, Mori E, et al. Involvement of TRPM7 in cell growth as a spontaneously activated Ca2 + entry pathway in human retinoblastoma cells. J Pharmacol Sci. 2004;95(4):403–419.


[62] Middelbeek J, Kuipers AJ, Henneman L, Visser D, Eidhof I, van Horssen R, et al. TRPM7 is required for breast tumor cell metastasis. Cancer Res. 2012;72(16):4250–4261.


[63] Meng X, Cai C, Wu J, Cai S, Ye C, Chen H, et al. TRPM7 mediates breast cancer cell migration and invasion through the MAPK pathway. Cancer Lett. 2013;333(1):96–102.


[64] Chen JP, Luan Y, You CX, Chen XH, Luo RC, Li R. TRPM7 regulates the migration of human nasopharyngeal carcinoma cell by mediating Ca(2 +) influx. Cell Calcium. 2010;47(5):425–432.


[65] Yee NS, Chan AS, Yee JD, Yee RK. TRPM7 and TRPM8 ion channels in pancreatic adenocarcinoma: potential roles as cancer biomarkers and targets. Scientifica (Cairo). 2012;2012:415158.


[66] Tsavaler L, Shapero MH, Morkowski S, Laus R. Trp-p8, a novel prostate-specific gene, is up-regulated in prostate cancer and other malignancies and shares high homology with transient receptor potential calcium channel proteins. Cancer Res. 2001;61(9):3760–3769.


[67] McKemy DD, Neuhausser WM, Julius D. Identification of a cold receptor reveals a general role for TRP channels in thermosensation. Nature. 2002;416(6876):52–58.


[68] Peier AM, Moqrich A, Hergarden AC, Reeve AJ, Andersson DA, Story GM, et al. A TRP channel that senses cold stimuli and menthol. Cell. 2002;108(5):705–715.


[69] De Blas GA, Darszon A, Ocampo AY, Serrano CJ, Castellano LE, Hernandez-Gonzalez EO, et al. TRPM8, a versatile channel in human sperm. PLoS One. 2009;4(6):e6095.


[70] Fuessel S, Sickert D, Meye A, Klenk U, Schmidt U, Schmitz M, et al. Multiple tumor marker analyses (PSA, hK2, PSCA, trp-p8) in primary prostate cancers using quantitative RT-PCR. Int J Oncol. 2003;23(1):221–228.


[71] Stein RJ, Santos S, Nagatomi J, Hayashi Y, Minnery BS, Xavier M, et al. Cool (TRPM8) and hot (TRPV1) receptors in the bladder and male genital tract. J Urol. 2004;172(3):1175–1178.


[72] Bidaux G, Flourakis M, Thebault S, Zholos A, Beck B, Gkika D, et al. Prostate cell differentiation status determines transient receptor potential melastatin member 8 channel subcellular localization and function. J Clin Invest. 2007;117(6):1647–1657.


[73] Bidaux G, Roudbaraki M, Merle C, Crepin A, Delcourt P, Slomianny C, et al. Evidence for specific TRPM8 expression in human prostate secretory epithelial cells: functional androgen receptor requirement. Endocr Relat Cancer. 2005;12(2):367–382.


[74] Thebault S, Lemonnier L, Bidaux G, Flourakis M, Bavencoffe A, Gordienko D, et al. Novel role of cold/menthol-sensitive transient receptor potential melastatine family member 8 (TRPM8) in the activation of store-operated channels in LNCaP human prostate cancer epithelial cells. J Biol Chem. 2005;280(47):39423–39435.


[75] Zhang L, Barritt GJ. Evidence that TRPM8 is an androgen-dependent Ca2 + channel required for the survival of prostate cancer cells. Cancer Res. 2004;64(22):8365–8373.


[76] Henshall SM, Afar DE, Hiller J, Horvath LG, Quinn DI, Rasiah KK, et al. Survival analysis of genome-wide gene expression profiles of prostate cancers identifies new prognostic targets of disease relapse. Cancer Res. 2003;63(14):4196–4203.


[77] Kiessling A, Fussel S, Schmitz M, Stevanovic S, Meye A, Weigle B, et al. Identification of an HLA-A*0201-restricted T-cell epitope derived from the prostate cancer-associated protein trp-p8. Prostate. 2003;56(4):270–279.


[78] Gkika D, Flourakis M, Lemonnier L, Prevarskaya N. PSA reduces prostate cancer cell motility by stimulating TRPM8 activity and plasma membrane expression. Oncogene. 2010;29(32):4611–4616.


[79] Yamamura H, Ugawa S, Ueda T, Morita A, Shimada S. TRPM8 activation suppresses cellular viability in human melanoma. Am J Physiol Cell Physiol. 2008;295(2):C296–C301.


[80] Chodon D, Guilbert A, Dhennin-Duthille I, Gautier M, Telliez MS, Sevestre H, et al. Estrogen regulation of TRPM8 expression in breast cancer cells. BMC Cancer. 2010;10:212.


[81] Li Q, Wang X, Yang Z, Wang B, Li S. Menthol induces cell death via the TRPM8 channel in the human bladder cancer cell line T24. Oncology. 2009;77(6):335–341.


[82] Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature. 1997;389(6653):816–824.


[83] Domotor A, Peidl Z, Vincze A, Hunyady B, Szolcsanyi J, Kereskay L, et al. Immunohistochemical distribution of vanilloid receptor, calcitonin-gene related peptide and substance P in gastrointestinal mucosa of patients with different gastrointestinal disorders. Inflammopharmacology. 2005;13(1-3):161–177.


[84] Hartel M, di Mola FF, Selvaggi F, Mascetta G, Wente MN, Felix K, et al. Vanilloids in pancreatic cancer: potential for chemotherapy and pain management. Gut. 2006;55(4):519–528.


[85] Lazzeri M, Vannucchi MG, Spinelli M, Bizzoco E, Beneforti P, Turini D, et al. Transient receptor potential vanilloid type 1 (TRPV1) expression changes from normal urothelium to transitional cell carcinoma of human bladder. Eur Urol. 2005;48(4):691–698.


[86] Sanchez MG, Sanchez AM, Collado B, Malagarie-Cazenave S, Olea N, Carmena MJ, et al. Expression of the transient receptor potential vanilloid 1 (TRPV1) in LNCaP and PC-3 prostate cancer cells and in human prostate tissue. Eur J Pharmacol. 2005;515(1-3):20–27.


[87] Miao X, Liu G, Xu X, Xie C, Sun F, Yang Y, et al. High expression of vanilloid receptor-1 is associated with better prognosis of patients with hepatocellular carcinoma. Cancer Genet Cytogenet. 2008;186(1):25–32.


[88] Amantini C, Mosca M, Nabissi M, Lucciarini R, Caprodossi S, Arcella A, et al. Capsaicin-induced apoptosis of glioma cells is mediated by TRPV1 vanilloid receptor and requires p38 MAPK activation. J Neurochem. 2007;102(3):977–990.


[89] Bode AM, Cho YY, Zheng D, Zhu F, Ericson ME, Ma WY, et al. Transient receptor potential type vanilloid 1 suppresses skin carcinogenesis. Cancer Res. 2009;69(3):905–913.


[90] Ghilardi JR, Rohrich H, Lindsay TH, Sevcik MA, Schwei MJ, Kubota K, et al. Selective blockade of the capsaicin receptor TRPV1 attenuates bone cancer pain. J Neurosci. 2005;25(12):3126–3131.


[91] Caterina MJ, Rosen TA, Tominaga M, Brake AJ, Julius D. A capsaicin-receptor homologue with a high threshold for noxious heat. Nature. 1999;398(6726):436–441.


[92] Lehen’kyi V, Prevarskaya N. TRPV2 (transient receptor potential cation channel, subfamily V, member 2). Atlas Genet Cytogenet Oncol Haematol. 2012;16(8):563–567.


[93] Monet M, Lehen’kyi V, Gackiere F, Firlej V, Vandenberghe M, Roudbaraki M, et al. Role of cationic channel TRPV2 in promoting prostate cancer migration and progression to androgen resistance. Cancer Res. 2010;70(3):1225–1235.


[94] Vriens J, Janssens A, Prenen J, Nilius B, Wondergem R. TRPV channels and modulation by hepatocyte growth factor/scatter factor in human hepatoblastoma (HepG2) cells. Cell Calcium. 2004;36(1):19–28.


[95] Liu G, Xie C, Sun F, Xu X, Yang Y, Zhang T, et al. Clinical significance of transient receptor potential vanilloid 2 expression in human hepatocellular carcinoma. Cancer Genet Cytogenet. 2010;197(1):54–59.


[96] Caprodossi S, Lucciarini R, Amantini C, Nabissi M, Canesin G, Ballarini P, et al. Transient receptor potential vanilloid type 2 (TRPV2) expression in normal urothelium and in urothelial carcinoma of human bladder: correlation with the pathologic stage. Eur Urol. 2008;54(3):612–620.


[97] Yamada T, Ueda T, Shibata Y, Ikegami Y, Saito M, Ishida Y, et al. TRPV2 activation induces apoptotic cell death in human T24 bladder cancer cells: a potential therapeutic target for bladder cancer. Urology. 2010;76(2):509e1–509e7.


[98] Nabissi M, Morelli MB, Amantini C, Farfariello V, Ricci-Vitiani L, Caprodossi S, et al. TRPV2 channel negatively controls glioma cell proliferation and resistance to Fas-induced apoptosis in ERK-dependent manner. Carcinogenesis. 2010;31(5):794–803.


[99] Peng JB, Chen XZ, Berger UV, Vassilev PM, Tsukaguchi H, Brown EM, et al. Molecular cloning and characterization of a channel-like transporter mediating intestinal calcium absorption. J Biol Chem. 1999;274(32):22739–22746.


[100] Lehen’kyi V, Raphael M, Prevarskaya N. The role of the TRPV6 channel in cancer. J Physiol. 2012;590(Pt 6):1369–1376.


[101] Fixemer T, Wissenbach U, Flockerzi V, Bonkhoff H. Expression of the Ca2+-selective cation channel TRPV6 in human prostate cancer: a novel prognostic marker for tumor progression. Oncogene. 2003;22(49):7858–7861.


[102] Peng JB, Zhuang L, Berger UV, Adam RM, Williams BJ, Brown EM, et al. CaT1 expression correlates with tumor grade in prostate cancer. Biochem Biophys Res Commun. 2001;282(3):729–734.


[103] Lehen’kyi V, Flourakis M, Skryma R, Prevarskaya N. TRPV6 channel controls prostate cancer cell proliferation via Ca(2 +)/NFAT-dependent pathways. Oncogene. 2007;26(52):7380–7385.


[104] Peleg S, Sellin JH, Wang Y, Freeman MR, 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(3):G593–G601.


[105] Dhennin-Duthille I, Gautier M, Faouzi M, Guilbert A, Brevet M, Vaudry D, et al. High expression of transient receptor potential channels in human breast cancer epithelial cells and tissues: correlation with pathological parameters. Cell Physiol Biochem. 2011;28(5):813–822.


[106] Bolanz KA, Kovacs GG, Landowski CP, Hediger MA. Tamoxifen inhibits TRPV6 activity via estrogen receptor-independent pathways in TRPV6-expressing MCF-7 breast cancer cells. Mol Cancer Res. 2009;7(12):2000–2010.


[107] Chow J, Norng M, Zhang J, Chai J. TRPV6 mediates capsaicin-induced apoptosis in gastric cancer cells—mechanisms behind a possible new “hot” cancer treatment. Biochim Biophys Acta. 2007;1773(4):565–576.


[108] Hofmann T, Schaefer M, Schultz G, Gudermann T. Transient receptor potential channels as molecular substrates of receptor-mediated cation entry. J Mol Med (Berl). 2000;78(1):14–25.


[109] Cheng KT, Liu X, Ong HL, Ambudkar IS. Functional requirement for Orai1 in store-operated TRPC1-STIM1 channels. J Biol Chem. 2008;283(19):12935–12940.


[110] Kim MS, Zeng W, Yuan JP, Shin DM, Worley PF, Muallem S. Native store-operated Ca2 + influx requires the channel function of Orai1 and TRPC1. J Biol Chem. 2009;284(15):9733–9741.


[111] Bomben VC, Turner KL, Barclay TT, Sontheimer H. Transient receptor potential canonical channels are essential for chemotactic migration of human malignant gliomas. J Cell Physiol. 2011;226(7):1879–1888.


[112] Bomben VC, Sontheimer H. Disruption of transient receptor potential canonical channel 1 causes incomplete cytokinesis and slows the growth of human malignant gliomas. Glia. 2010;58(10):1145–1156.


[113] El Hiani Y, Ahidouch A, Lehen’kyi V, Hague F, Gouilleux F, Mentaverri R, et al. 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(4–6):335–346.


[114] Tajeddine N, Gailly P. TRPC1 protein channel is major regulator of epidermal growth factor receptor signaling. J Biol Chem. 2012;287(20):16146–16157.


[115] Jiang HN, Zeng B, Zhang Y, Daskoulidou N, Fan H, Qu JM, et al. Involvement of TRPC channels in lung cancer cell differentiation and the correlation analysis in human non-small cell lung cancer. PLoS One. 2013;8(6):e67637.


[116] Zeng B, Yuan C, Yang X, Atkin SL, Xu SZ. TRPC channels and their splice variants are essential for promoting human ovarian cancer cell proliferation and tumorigenesis. Curr Cancer Drug Targets. 2013;13(1):103–116.


[117] Zhang H, Zhou L, Shi W, Song N, Yu K, Gu Y. A mechanism underlying the effects of polyunsaturated fatty acids on breast cancer. Int J Mol Med. 2012;30(3):487–494.


[118] Saito H, Minamiya Y, Watanabe H, Takahashi N, Ito M, Toda H, et al. Expression of the transient receptor potential channel c3 correlates with a favorable prognosis in patients with adenocarcinoma of the lung. Ann Surg Oncol. 2011;18(12):3377–3383.


[119] Thebault S, Flourakis M, Vanoverberghe K, Vandermoere F, Roudbaraki M, Lehen’kyi V, et al. Differential role of transient receptor potential channels in Ca2 + entry and proliferation of prostate cancer epithelial cells. Cancer Res. 2006;66(4):2038–2047.


[120] Yue D, Wang Y, Xiao JY, Wang P, Ren CS. Expression of TRPC6 in benign and malignant human prostate tissues. Asian J Androl. 2009;11(5):541–547.


[121] El Boustany C, Bidaux G, Enfissi A, Delcourt P, Prevarskaya N, Capiod T. Capacitative calcium entry and transient receptor potential canonical 6 expression control human hepatoma cell proliferation. Hepatology. 2008;47(6):2068–2077.


[122] Song J, Wang Y, Li X, Shen Y, Yin M, Guo Y, et al. Critical role of TRPC6 channels in the development of human renal cell carcinoma. Mol Biol Rep. 2013;40(8):5115–5122.


[123] Aydar E, Yeo S, Djamgoz M, Palmer C. Abnormal expression, localization and interaction of canonical transient receptor potential ion channels in human breast cancer cell lines and tissues: a potential target for breast cancer diagnosis and therapy. Cancer Cell Int. 2009;9:23.


[124] Guilbert A, Dhennin-Duthille I, Hiani YE, Haren N, Khorsi H, Sevestre H, et al. Expression of TRPC6 channels in human epithelial breast cancer cells. BMC Cancer. 2008;8:125.


[125] Bernaldo de Quiros S, Merlo A, Secades P, Zambrano I, de Santa Maria IS, Ugidos N, et al. Identification of TRPC6 as a possible candidate target gene within an amplicon at 11q21-q22.2 for migratory capacity in head and neck squamous cell carcinomas. BMC Cancer. 2013;13:116.


[126] Chigurupati S, Venkataraman R, Barrera D, Naganathan A, Madan M, Paul L, et al. Receptor channel TRPC6 is a key mediator of Notch-driven glioblastoma growth and invasiveness. Cancer Res. 2010;70(1):418–427.


[127] Stock K, Kumar J, Synowitz M, Petrosino S, Imperatore R, Smith ES, et al. Neural precursor cells induce cell death of high-grade astrocytomas through stimulation of TRPV1. Nat Med. 2012;18(8):1232–1238.


[128] Beck B, Bidaux G, Bavencoffe A, Lemonnier L, Thebault S, Shuba Y, et al. Prospects for prostate cancer imaging and therapy using high-affinity TRPM8 activators. Cell Calcium. 2007;41(3):285–294.


[129] Zhang L, Brereton HM, Hahn M, Froscio M, Tilley WD, Brown MP, et al. Expression of Drosophila Ca2 + permeable transient receptor potential-like channel protein in a prostate cancer cell line decreases cell survival. Cancer Gene Ther. 2003;10(8):611–625.

Only gold members can continue reading. Log In or Register to continue

Stay updated, free articles. Join our Telegram channel

Nov 18, 2017 | Posted by in PHARMACY | Comments Off on TRP Channels in Cancer

Full access? Get Clinical Tree

Get Clinical Tree app for offline access