Next Article in Journal
Assessment of the Effect of Sorafenib on Omega-6 and Omega-3 Epoxyeicosanoid Formation in Patients with Hepatocellular Carcinoma
Next Article in Special Issue
On-treatment Serum Mac-2 Binding Protein Glycosylation Isomer (M2BPGi) Level and Risk of Hepatocellular Carcinoma Development in Patients with Chronic Hepatitis B during Nucleot(s)ide Analogue Therapy
Previous Article in Journal
Comparative Analysis of Complete Mitochondrial Genomes of Three Gerres Fishes (Perciformes: Gerreidae) and Primary Exploration of Their Evolution History
Previous Article in Special Issue
Telmisartan Inhibits Cell Proliferation and Tumor Growth of Esophageal Squamous Cell Carcinoma by Inducing S-Phase Arrest In Vitro and In Vivo
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

TRP Channels in Digestive Tract Cancers

Institute of Biochemistry and Molecular Medicine, National Center of Competence in Research NCCR TransCure, University of Bern, 3012 Bern, Switzerland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(5), 1877; https://doi.org/10.3390/ijms21051877
Submission received: 10 February 2020 / Revised: 3 March 2020 / Accepted: 6 March 2020 / Published: 9 March 2020

Abstract

:
Cancers of the digestive tract are among the most prevalent types of cancer. These types of cancers are often diagnosed at a late stage, which results in a poor prognosis. Currently, many biomedical studies focus on the role of ion channels, in particular transient receptor potential (TRP) channels, in cancer pathophysiology. TRP channels show mostly non-selective permeability to monovalent and divalent cations. TRP channels are often dysregulated in digestive tract cancers, which can result in alterations of cancer hallmark functions, such as enhanced proliferation, migration, invasion and the inability to induce apoptosis. Therefore, TRP channels could serve as potential diagnostic biomarkers. Moreover, TRP channels are mostly expressed on the cell surface and ion channel targeting drugs do not need to enter the cell, making them attractive candidate drug targets. In this review, we summarize the current knowledge about TRP channels in connection to digestive tract cancers (oral cancer, esophageal cancer, liver cancer, pancreatic cancer, gastric cancer and colorectal cancer) and give an outlook on the potential of TRP channels as cancer biomarkers or therapeutic targets.

Graphical Abstract

1. Introduction

Digestive malignancies refer to a heterogeneous group of cancers that arise in the gastrointestinal tract and associated organs. Cancers of the colon and rectum, stomach, liver, esophagus, and pancreas are among five of the ten most prevalent cancers and cancer-related causes of death [1]. For 2020, approximately 387,000 new cases of cancer in the digestive system, the oral cavity and the pharynx and 179,000 related deaths have been estimated in the United States [2]. Successful treatment of these malignancies largely depends on effective screening tools, such as diagnostic blood, stool and endoscopic tests [3,4,5]. Therapeutic approaches include surgery, chemo and radiation therapy, hormone therapy; as well as tailored and personalized therapies, which are presently undergoing development [6,7,8,9,10,11,12,13]. To further advance personalized medicine, many research groups have focused their work on the understanding of the molecular nature of cancer hallmark functions, including unlimited proliferation, increased viability, migration and invasion of cancer cells, and decreased ability to induce apoptosis [14]. Ion channels mediate numerous responses in cellular physiology and are often dysregulated in various diseases, especially in most types of cancer including cancers of the gastrointestinal tract [15]. Ion channels are often responsible for switching on and off intracellular signaling pathways, which makes them attractive candidate drug targets. In addition, they are mostly expressed on the cell surface, and, as a consequence, ion channel targeting drugs do not need to enter the cell. In recent decades, members of the transient receptor potential (TRP) channel family have been proposed as potential biomarkers and/or drug targets in cancer therapy.
TRP genes were first reported in 1969, when a mutation in the genome of a visually impaired Drosophila melanogaster fruit fly was described. This mutant, called transient receptor potential, had a transient response to steady light, in opposite to sustained electroretinogram, recorded in the wild-type flies [16]. However, the trp gene was identified and described 20 years later [17]. Since then, numerous homologous TRP channel family members have been identified and classified into six human TRP channel subfamilies, including canonical (TRPC), melastatin (TRPM), vanilloid (TRPV), ankyrin (TRPA), polycystic (TRPP), and mucolipin (TRPML) channels [18,19]. Most TRP channels have an essential role in the influx of monovalent and divalent cations, such as Na+, Mg2+ and Ca2+, as well as trace metal ions [18]. Originally, the TRP channel transduction pathways were described for taste and pungent compound perception, thermo- and mechanosensation, pain, osmoregulation, as well as hormone and pheromone signaling [20,21,22,23,24]. Besides their roles in sensory processes, TRP channels mediate many cellular physiological and pathophysiological functions in cancer and the immune system [25,26,27,28,29,30,31,32,33,34]. One way of how TRP channels contribute to the pathogenesis of different types of cancer is through (dys)regulation of intracellular ion levels. For example, the switch from a quiescent cell to a proliferating cell is characterized by global dynamic Ca2+ elevations and the activation of Ca2+ effectors. Cells progressing through the cell cycle are characterized by Ca2+ oscillations [35,36,37,38]. Additionally, Ca2+ can contribute to the activation or inhibition of apoptosis, as well as the ability of cells to migrate [35,36,37,38]. All TRP proteins share a common topology of six transmembrane segments (S1–S6), with a pore-forming loop between the S5 and S6 segments. The transmembrane segments tend to share the greatest homology within the particular subfamily, and amino acid sequences in the pore region of TRP channels are the most highly conserved [18,19]. The amino (N) and carboxyl (C) terminuses are located intracellularly; they vary in length and sequence, and contain diverse domains and motifs, which play a role in channel assembly, activation and regulation. These domains and motifs can include coiled coils, calmodulin-binding sites, lipid interaction domains, EF hands, or phosphorylation sites, and are highly variable within members of the same subfamily [18,39]. Recent advances in cryogenic electron microscopy (cryo-EM) based structural analysis have provided insights into the architecture of several TRP channels, including TRPA1, TRPPC, TRPM, and TRPV channels [40,41,42,43,44,45,46,47,48,49,50,51,52,53,54]. To date, most changes involving TRP channels in cancer do not involve mutations in the TRP genes but rather increased or decreased levels of expression of functional TRP proteins, depending on the cancer’s stage. Here, we focus on TRP channels, especially members of the TRPC, TRPM, and TRPV subfamilies in digestive malignancies, that are mostly of epithelial origin, including oral, esophageal, pancreatic, gastric, and colorectal cancer.

2. Oral Cancers

Oral squamous cell carcinoma (SCC) accounts for approximately 90% of cases of oral cancer [55]. The oral cavity comes into contact with several sensory stimuli that are known to activate a number of TRP channels. Capsaicin, the chemical compound responsible for the burning sensation of chili peppers, activates TRPV1 channels. Additionally, pungent mustard oil activates TRPA1, and menthol activates TRPM8. It has been suggested that these compounds might exhibit chemoprotective features [56,57,58]. In human oral SCC cells, TRPV1–4, TRPV6, TRPA1, TRPM8, and TRPM2 are expressed [59,60,61,62]. In 2009, the expression levels of TRPV1 were investigated in the human tongue, in tongue SCC, and pre-malignant leukoplakia. Under pre-malignant conditions and in SCC, the TRPV1 protein expression was increased [59]. The expression of TRPV1 protein was also shown in human oral SCC. Capsaicin, a TRPV1 agonist, was shown to induce cytotoxicity in oral SCC cells. However, this effect was independent of TRPV1 conductivity, as these cells did not exhibit an increase in intracellular Ca2+ upon stimulation with capsaicin [63]. Furthermore, it was shown that TRPV1–4 expression levels were elevated on protein level in different areas of the oral cavity, including the tongue, buccal mucosa, gingiva, and the oral floor, compared to normal oral mucosa. In addition, known risk factors for SCC, such as alcohol consumption and smoking, increased the expression levels of TRPV1–4 mRNAs [61]. TRPA1 is expressed in oral SCC, and thymol induces an increase in intracellular Ca2+, which is blocked when a TRPA1 antagonist is present. However, the anti-cancerous effects of thymol are not mediated by TRPA1 [64]. Notably, it has been reported that oral SCC cells may secrete certain lipids that activate TRPA1 and TRPV1 nociceptors and thus mediate SCC-induced neck pain [65]. Cold/menthol-activated TRPM8 is expressed on protein and mRNA level in oral SCC cell lines. Immunohistochemistry (IHC) analysis showed that TRPM8 localizes not only in the plasma membrane, but also in the endoplasmic reticulum (ER) membrane. Menthol activates TRPM8 in the ER plasma membrane, resulting in a Ca2+ release from intracellular Ca2+ stores and store-operated Ca2+ entry. Menthol-activated Ca2+ entry is also partially mediated by TRPM8, and both of these mechanisms can be blocked by a TRPM8-specific inhibitor. Functionally, the inhibition of TRPM8 results in a reduction in oral SCC cell viability, migration and invasion [60]. TRPM2 is another ion channel that contributes to Ca2+ influx and is activated by cADPR, reactive oxygen species (ROS), and 2′-deoxy-ADPR [66,67]. TRPM2 protein expression is significantly increased in tongue SCC compared to non-malignant tongue tissues (control or papilloma). In two human tongue SCC cell lines, TRPM2 expression was upregulated compared to a non-tumorigenic oral epithelial cell line. The downregulation of TRPM2 decreased cancer cell migration and increased ROS induced apoptosis [62]. Table 1 summarizes studies in which the expression of TRP channels in human tongue and oral tumor samples was described.

3. Esophgeal Cancer

Esophageal cancer is the eighth most common cancer worldwide [1] and is characterized by a poor prognosis [68]. This type of cancer is classified into two different cancer entities: esophageal squamous cell carcinoma (ESCC) and esophageal adenocarcinoma (EAC) [1,68,69,70]. The link between the pathophysiology of esophageal cancer and TRP channel expression has been exclusively made for ESCC and includes dysregulation of TRPC6, TRPM7, TRPM8, TRPV1, TRPV2, and TRPV4 expression levels. Table 2 summarizes studies in which the expression of TRP channels in human ESCC samples was described.
TRPC6 mRNA and protein expression levels are increased in human ESCC tissues compared to normal esophageal tissues. Furthermore, the inhibition of TRPC6 in ESCC cells led to decreased Ca2+ signaling and cell cycle arrest via Cdk1. In addition, the inhibition of TRPC6 decreased tumor formation in a mouse xenograft model [71].
TRPM8 mRNA and protein expression have been reported to be upregulated in ESCC tissues and ESCC cell lines. The knockdown and inhibition of TRPM8 decrease proliferation of esophageal cancer cells. Moreover, TRPM8 negatively regulates PD-L1 expression through the calcineurin–NFATc3 pathway, enabling the immune evasion of ESCC cells [72].
The protein expression of the Ca2+/Mg2+ channel TRPM7was reported to be a good independent prognostic factor for patients with ESCC. Additionally, siRNA-based silencing of TRPM7 in TE6 ESCC cells increased their proliferation, migration, and invasion [73].
TRPV2 mRNA and protein have been found to be overexpressed in ESCC cell lines. In ESCC patients, higher expression of TRPV2 protein correlates with a worse 5 year overall survival rate after surgery. The knockdown of TRPV2 in ESCC cells decreased proliferation, cell cycle progression, and the ability to invade and migrate. Moreover, the authors found downregulated WNT/β-catenin or basal cell carcinoma signaling-related genes [74]. TRPV1 and TRPV4 were both found to be expressed on mRNA and protein level in non-tumorous esophageal squamous cells and overexpressed in ESCC cells. Furthermore, TRPV1 and TRPV4 are functional in these cells, as shown with calcium imaging and whole cell patch clamp techniques. Finally, the overactivation of both channels leads to increased proliferation and migration of ESCC cells [75].

4. Liver Cancer

Liver can be affected by two types of cancer: hepatocellular carcinoma (HCC) or metastases from colorectal cancer. Liver cancer tissue has been reported to express several TRP channels. El Boustany and colleagues showed that TRPC1, TRPC6, TRPV1, TRPV2, TRPV4, TRPM4, TRPM6, TRPM7, and TRPM8 are expressed on mRNA level in both Huh-7 and HepG2 human hepatoma cell lines, while TRPV3 and TRPM5 are only expressed in Huh-7 cells [76]. Table 3 summarizes studies in which the expression of TRP channels in human liver cancer samples was described.
TRPV1 is overexpressed in hepatocarcinoma tissues compared to normal liver tissue. In the same study, a correlation between TRPV1 protein expression and histopathologic differentiation was reported. High TRPV1 expression was associated with longer disease-free survival [77]. In addition, TRPV1 was previously shown to modulate migration of HCC cells [81,82]. Furthermore, capsaicin induced an increase in intracellular Ca2+ levels, ROS production, and apoptosis in HepG2 cells [83]. The induction of apoptosis in HepG2 cells by capsaicin-induced TRPV1 activation was later further confirmed [84].
Another member of the TRPV subfamily, TRPV2, is also expressed in hepatocarcinoma tissue. TRPV2 expression on mRNA and protein level are increased in moderately and well-differentiated hepatocarcinoma tissues compared to poorly differentiated tumors. Moreover, a correlation between TRPV2 expression and portal vein invasion was confirmed [78]. Interestingly, the expression of TRPV2 mRNA and protein levels in HepG2 and Huh-7 cells was reported to increase upon exposure to ROS, specifically H2O2, resulting in an activation of cell death. The elevation of TRPV2 expression led to an inhibition of pro-survival signals (Akt, Nrf2) and, in the early stage of apoptosis, promoted the activation of pro-death signals (p38, JNK1) [85]. Another study showed that shRNA-based TRPV2 knockdown in HepG2 cells enhanced spheroid and colony formation, which was restored by the overexpression of TRPV2. Additionally, the expression of TRPV2 protein was linked to the stemness of liver cancer cells, as the knockdown of TRPV2 in HepG2 cells induced the expression of liver cancer stem-like cells (LCSLCs) markers. Moreover, in human hepatoma cell line SMMC-7721, which exhibits lower TRPV2 protein expression than HepG2, reinforced TRPV2 expression led to the reduction of the expression of LCSLCs markers and reduced spheroid and colony formation. In line with these findings, probenecid, a TRPV2 agonist, also reduced LCSLCs markers expression, as well as spheroid and colony formation. LCSLCs are considered to be good targets for liver cancer therapy, as they exhibit stem cell properties and are of a highly tumorigenic nature. Therefore, TRPV2 was proposed to be a potential target in hepatoma therapy [86]. Notably, reduced expression of TRPV2 mRNA and protein in poorly differentiated tumors in comparison to higher differentiated hepatoma [78] supports the idea that reduced TRPV2 expression promotes the stem-cell features of hepatoma cells during the early stages of tumor development [86].
Protein and mRNA levels of TRPV4 are elevated in hepatocellular carcinoma tissue compared with paired non-tumor tissue. Poorly differentiated HCC tumors displayed stronger TRPV4 expression and a positive correlation between TRPV4 expression; the histological grade and number of tumors was described [79]. In addition, with the use of a TRPV4-specific antagonist and agonist, Fang and colleagues showed that TRPV4 is crucial for HCC cell viability and that its inhibition could cause anti-tumor effects through modulation of the expression of apoptosis-related molecules. Furthermore, in a xenograft mouse model system, the blockage of TRPV4 was shown to decrease tumor size and weight compared to the control group [79].
The downregulation of TRPC1 with shRNA-based interference has been shown to suppress cell proliferation but not migration of HCC cells. In addition to the anti-proliferative effect of TRPC1 silencing, Selli and colleagues reported an elevation in store-operated Ca2+ entry [87].
TRPC6 is weakly expressed on mRNA and protein level in normal hepatocytes but strongly expressed in liver carcinoma samples. Moreover, high TRPC6 expression has been suggested to be associated with a higher Tumor Node Metastasis (TNM) classification of tumors [76,80]. TRPC6 overexpression in Huh-7 cells leads to increased proliferation [76]. In addition, TGF-β has been shown to induce a formation of the complex between TRPC6 and the type1 Na+/Ca2+ exchanger (NCX1), leading to an increased migration and invasion of HCC. Moreover, a positive correlation was found between the severity of HCC progression and the expression of TRPC6 and NCX1 [80]. Both the inhibition and RNAi downregulation of TRPC6 or NCX1 was able to attenuate TGF-β induced cell migration or the intrahepatic metastasis of human HCC in an in vivo xenograft mouse model [80]. Moreover, Wen and colleagues recently reported a significantly increased TRPC6 mRNA expression in HCC cells upon hypoxia stimulation, doxorubicin treatment, or ionizing radiation [88]. The inhibition or downregulation of TRPC6 significantly decreased drug resistance to all three stimuli, and TRPC6 inhibition could reverse endothelial–mesenchymal transition (EMT), induced by doxorubicin treatment. In line with these findings, TRPC6 interference in vivo enhanced the sensitivity to doxorubicin and led to slower growth of the cells compared to the control cells [88].
Another study found that bradykinin can activate TRPM7 in HepG2 cells, leading to an influx of Ca2+. TRPM7-mediated Ca2+ influx was necessary for the activity of calpains, which play a role in migration of HepG2 cells. These findings suggest that TRPM7 could be involved in migration of liver cancer cells. Notably, treatment with bradykinin resulted in an increase in MMP-2 secretion and enhanced migration [89].

5. Pancreatic Cancer

Pancreatic cancer is usually detected at an advanced stage and, currently, most patients are diagnosed with distant metastases and, therefore, have a poor 5 year survival rate [90,91]. Detection of pancreatic cancer at an early stage has a favorable impact on long-term survival, as the 5 year survival of localized pancreatic cancer is about 25% but is only 2% for distant disease [90,91].
An analysis performed by Lin and colleagues revealed that somatic mutations in the TRPM2 gene exhibit a negative correlation with pancreatic cancer patient survival rates in comparison to the group without TRPM2 mutations. Additionally, the higher the TRPM2 mRNA expression in the cancerous tissue, the shorter the survival of the patients. TRPM2 expression was also shown to enhance proliferation, migration, and invasion of PANC-1 cells, a human pancreatic cancer cell line [92].
In 2007, it was shown that TRPM8 is expressed on mRNA and protein level and mediates non-selective cation currents in the human pancreatic neuroendocrine tumor (NET) cell line BON, as well as in primary NET cell lines [93]. TRPM8 protein was found to be upregulated in human pancreatic adenocarcinoma cell lines and tissues [94,95]. Other studies further confirmed the upregulation of TRPM8 in pancreatic cancer tissues in human patients compared to adjacent tissues [96,97]. In addition, high TRPM8 protein expression was found to be associated with lower overall survival and poor disease free survival values for pancreatic cancer patients [97], as well as positively correlated with the tumor size and stage of pancreatic cancer [94]. Furthermore, it was shown that TRPM8 is required for proliferation of the pancreatic adenocarcinoma cell lines, PANC-1 and BxPC-3, due to the cell cycle arrest in the G0/G1 phase [95]. Another study showed that TRPM8 protein is expressed in PANC-1 cells in a non-glycosylated form and is functional. Additionally, TRPM8 knockdown with siRNA resulted in a decrease in cell migration [98]. TRPM8 expression in a non-glycosylated form was further confirmed and shown to have an impact on the conductive properties of the channel. Additionally, it was suggested that the glycosylation patterns in PANC-1 cells could have an impact on proliferation [99]. It was also further shown that TRPM8 mediates Ca2+ influx in PANC-1 and BxPC-3 cell lines and plays a role in proliferation as well as migration. Additionally, targeting TRPM8 in PANC-1 and BxPC-3 with siRNA resulted in an enhancement of gemcitabine cytotoxicity, which was accompanied by a change in the expression of apoptosis- and gemcitabine metabolism-related proteins [96]. Another study further showed that targeting TRPM8 with shRNA in BxPC-3 and MIA PaCa-2 resulted in a decrease in invasion [94].
TRPM7 was shown to play a role in the development of the pancreas in a zebra fish model, which was linked to Mg2+ sensitive signaling. These findings could be translated to the pathogenesis of pancreatic adenocarcinoma, in which TRPM7 protein was shown to be overexpressed and necessary for Mg2+-regulated proliferation. Targeting TRPM7 with siRNA in PANC-1 and BxPC-3 resulted in a cell cycle arrest in the G0/G1 phase, accompanied by a change in the expression of p21, cyclin G1, and cyclin B1. The role of TRPM7 in the regulation of pancreatic cancer cells proliferation was also connected to Mg2+, as supplementation of the culture medium with Mg2+ reversed the decrease in proliferation caused by the knockdown of TRPM7 [100]. TRPM7 does not affect apoptosis in PANC-1 and BxPC-3 [100,101]. However, it plays a role in preventing replicative senescence. Targeting TRPM7 enhances cytotoxicity of the conventionally used chemotherapeutic, gemcitabine, which is a pro-apoptotic agent, suggesting that targeting TRPM7 in pancreatic cancer could support standard chemotherapy [101]. Another group showed that TRPM7 mRNA and protein are overexpressed in pancreatic ductal adenocarcinoma compared to healthy pancreatic tissue, and TRPM7 staining intensity is inversely correlated with patients’ survival. Additionally, the study showed that BxPC-3 cells exhibit TRPM7-characteristic non-selective cationic currents and that TRPM7 activity regulates intracellular Mg2+ levels [102]. Contrary to previous findings [100,101], another study based on the siRNA targeting of TRPM7 reported no effect on cell viability and proliferation but showed a decrease in cell migration [102]. The overexpression of TRPM7 protein in pancreatic cancer was further confirmed, and a correlation was found between TRPM7 expression levels and the size and stage of tumors. Additionally, TRPM7 is overexpressed in pre-malignant tissue. Therefore, the expression of TRPM7 could be further investigated as a potential biomarker [103]. Furthermore, high TRPM7 staining in pancreatic ductal adenocarcinoma has been associated with higher TRPM7 protein staining in the lymph nodes, suggesting that TRPM7 might be involved in invasion of pancreatic cancer cells [104]. Indeed, TRPM7 has been shown to be involved in the invasion of BxPC-3 cells [103]. Another study showed that in PANC-1 and MIA PaCa-2 cell lines, the TRPM7-mediated cation current regulates Mg2+ homeostasis and cell invasion, as TRPM7 knockdown was shown to reduce invasion along with a decrease in MMP2, uPA, and Hsp90α secretion [104].
TRPV1 was shown to be expressed on mRNA and protein level and regulate intracellular Ca2+ levels in pancreatic NET BON-1 cells, and its activity triggered secretion of chromogranin A [105]. Furthermore, the TRPV1 agonist, capsaicin, reduced viability and proliferation, and induced apoptosis in NET cells. The cytotoxicity of capsaicin was shown to be due to the disturbance of mitochondrial potential and the inhibition of ATP production [106]. Similar to the observations in NET cells, TRPV1 overexpression in PANC-1 cells inhibited cell proliferation. Either the overexpression or agonist-induced activation of TRPV1 reduced the expression of epidermal growth factor receptor (EGFR), due to its ubiquitination. Additionally, TRPV1 reduced mRNA expression of two oncogenes, KRAS and AKT2 [107]. Another member of the TRPV subfamily, TRPV6, is expressed in BON-1 cell line on mRNA and protein level, and it regulates Ca2+ levels. The downregulation of TRPV6 in BON-1 cells, resulted in an inhibition of proliferation and reduced NFAT activation [108]. Furthermore, TRPV6 protein expression is elevated in pancreatic cancer tumors compared to non-tumor adjacent tissues and its high expression correlated with lower survival of the patients. The knockdown of TRPV6 in pancreatic cancer cell lines resulted in a reduced proliferation, induced cell cycle arrest in the G0/G1 phase, reduced migration and invasion, as well as increased apoptosis [109].
Table 4. summarizes studies in which the expression of TRP channels in human pancreatic cancer samples was described.

6. Gastric Cancer

Gastric cancer was the third most common cause of cancer-related deaths in 2018, just after lung and colorectal cancer [1]. Most patients with early-stage gastric cancer are asymptomatic; therefore, a diagnosis is often made when the cancer is at an advanced stage and shows metastasis [110]. Several TRP channels have been proposed to be involved in the pathogenesis of gastric cancer. The TRPC6 channel was shown to be upregulated on protein and mRNA level in human gastric cancer epithelial cells in comparison to normal gastric epithelial cells. TRPC6-mediated Ca2+ influx in gastric cancer cell lines is responsible for regulation of the cell cycle, as the inhibition of TRPC6 resulted in cell cycle arrest in the G2/M phase and inhibited cell growth. The involvement of TRPC6 conductivity in cell cycle regulation was further confirmed in experiments where the TRPC6 dominant-negative mutant was expressed. Moreover, treatment of nude mice with a TRPC6 blocker resulted in the inhibition of the development of a xenografted human gastric tumor [111]. Another study suggested a role for TRPC 1/3/6 in the regulation of EMT [112]. Additionally, a newly developed TRPC6 inhibitor showed an anti-tumor effect in nude mice with a xenografted human gastric tumor [113].
Capsaicin has been shown to induce apoptosis in a normal human epithelial gastric cell line and the gastric cancer cell line AGS. However, AGS cells were found to be more susceptible to capsaicin-induced apoptosis, which was induced through an increase in mitochondrial permeability and activation of Bax and p53. Surprisingly, the capsaicin-induced apoptosis was dependent on Ca2+ influx mediated by TRPV6, rather than TRPV1, a known capsaicin receptor [114].
TRPM2 was found to be expressed on mRNA level in gastric cancer patients, and its high expression was negatively associated with the overall survival of patients. Functional TRPM2 is expressed in gastric cancer cell lines AGS and MKN-45, and its shRNA based knockdown results in the inhibition of proliferation and enhancement of apoptosis. Additionally, TRPM2 knockdown was shown to alter autophagy in AGS cells, which led to mitochondrial dysfunction. The knockdown of TRPM2 also sensitized AGS and MKN-45 cells to treatment with paclitaxel and doxorubicin, resulting in a further reduction in cell viability. These findings suggest that targeting TRPM2 in combination with standard chemotherapeutic drugs could be beneficial for the treatment of gastric cancer patients [115].
TRPM7 channel was shown to be expressed in gastric cancer cell lines on mRNA and protein level. AGS gastric cancer cells exhibit TRPM7-like currents, and suppression of these currents by the unspecific TRPM7 inhibitors, La3+ or 2-APB, resulted in a decrease in cell viability and higher apoptosis rates. Additionally, Mg2+ was necessary for AGS cells survival and growth [116].
Table 5 summarizes studies in which the expression of TRP channels in human gastric cancer samples was described.

7. Colorectal Cancer

Colorectal cancer (CRC) is one of the most frequent types of cancer and cancer-related cause of death, both in Europe and in the United States [117,118]. Several TRP channels have been shown to be dysregulated in CRC. A study investigating mRNA expression levels of the TRP channels in human CRC tissue versus normal colon mucosa detected an increase in gene expression of TRPM8, TRPV6, and TRPV1 and a lower expression of TRPV4, TRPM4, TRPV3, TRPC6, and TRPV5 in the tumor tissues of CRC compared to normal tissues [119]. Table 6 summarizes studies in which the expression of TRP channels in human gastric cancer samples was described.
TRPC1 ion channel, which displays permeability towards Ca2+ ions, was reported to be upregulated on mRNA and protein level in CRC cells [124], and its higher mRNA expression in CRC patients has been correlated with a poor prognosis [120]. The upregulated expression of TRPC1 in CRC has been shown to contribute to an increased Ca2+ influx via store operated Ca2+ entry and higher Ca2+ signaling, which resulted in an increased proliferation, invasion, and survival of CRC cells [124]. Furthermore, Ca2+ influx through TRPC1 was also linked to an increased migration of CRC cells [125]. Another member of the TRPC subfamily, which conducts Ca2+, TRPC5, was shown to play a role in the drug resistance of human CRC cell lines. HCT-8 and LoVo cells resistant to 5-fluorouracil (5-FU), which is a commonly used chemotherapeutic in CRC therapy, showed higher expression of TRPC5 mRNA and protein in comparison to non-resistant cells. Further investigations revealed that the TRPC5-mediated Ca2+ influx induces the expression of the ATP-binding cassette subfamily B member 1 (ABCB1), a pump overproduced in cancer cells, responsible for the export of cytotoxic drugs. Additionally, TRPC5 promoted nuclear β-catenin localization [126].
Increasingly, evidence suggests an important role for TRPV1 channel in the physiology and pathophysiology of the intestine. Increased expression of TRPV1 protein was found in the colon of patients with inflammatory bowel disease, which was linked to an increased pain sensation [127]. However, several studies suggested that the expression and activity of TRPV1 in the intestine has a protective role in inflammatory states [128,129,130,131]. Consistent with these findings, in a mice model of colitis-associated cancer (CAC), mice that lacked the expression of TRPV1 had a higher incidence and higher number of tumors in the distant colon. The lack of TRPV1 was also accompanied by an increased proliferation of colon cells and higher β-catenin localization to the nuclei. Additionally, tumors from mice lacking TRPV1 showed an increased infiltration of inflammatory cells into the tumors, along with an elevated expression of IL-6 and IL-11 and activation of STAT3 and NF-kB signaling pathways [132]. Further evidence shows that TRPV1 protein expression can have a protective role against tumor development. De Jong and colleagues showed that functional TRPV1 is expressed in intestinal epithelial cells and can be activated by capsaicin. TRPV1 activation inhibits EGFR-induced epithelial cell proliferation via activation of Ca2+/calpain. In a murine model of multiple intestinal neoplasia (ApcMin/+ mice), TRPV1 deficiency promoted intestinal adenoma formation correlated with a reduced lifespan [133], which was consistent with previous findings [132]. In this model, mice that lack TRPV1 showed higher EGFR phosphorylation and proliferation markers in intestinal epithelial cells, and the deletion of TRPV1 increased the expression of the EGFR-regulated oncogenes, c-Fos and c-Myc. Furthermore, the administration of dietary capsaicin increased the survival of ApcMin/+ mice in a process that was dependent on TRPV1 [133]. Another study suggested that TRPV1 protein expression is decreased in CRC tissues. Treatment of HCT116 cell line with capsaicin resulted in an inhibition of proliferation and induced apoptosis through the activation of the tumor suppressor, p53. In this cell line, treatment with capsaicin led to an increase in intracellular Ca2+, possibly through TRPV1 [121]. On the other hand, studies in human CRC cell lines showed that Fibrulin-5, a component of ECM, is downregulated in CRC tissues and cell lines. Fibrulin-5 induces apoptosis through the downregulation of TRPV1 and ROS production [134]. Another study in HT-29 cell line showed that capsaicin induced apoptosis through PPARγ signaling but without the involvement of TRPV1, since capsazepine, a specific antagonist for the vanilloid receptor, did not inhibit capsaicin induced apoptosis [135].
TRPM8 is another ion channel responsible for Ca2+ influx into cells and was found to be highly expressed on protein level in human CRC cell lines, Caco-2, and HCT116. Furthermore, cannabigerol, a non-psychotropic cannabis-derived cannabinoid, reduces colon cancer progression in vivo and selectively inhibits the growth of CRC cells via interaction with TRPM8 [136].
TRPM4 is an ion channel, which is directly activated by an increase in intracellular Ca2+ concentration, however, is not permeable towards Ca2+. In non-excitable cells, under physiological conditions, TRPM4 conducts Na+ into the cell, thereby contributing to plasma membrane depolarization. This, in turn, reduces the driving force for further Ca2+ entry through store operated Ca2+ channels. Therefore, TRPM4 is regarded to be a negative regulator of Ca2+ signaling [137,138,139,140]. Reports on mRNA expression levels of TRPM4 in colorectal cancer either found it to be lower in CRC tissue compared to normal colon tissue [119], or no differences were detected [141]. However, a recent study investigating TRPM4 protein expression levels in CRC tissues showed that its high expression correlates with high numbers of tumor buds and an increased percentage of infiltrative tumor border configuration. Both of these features correlate with an increased frequency of vessel invasion and lymph node metastasis, which ultimately lead to an increased probability of disease reoccurrence and cancer related death. TRPM4 protein was also shown to be highly expressed in tumor buds, which were linked to increased metastasis in CRC. Furthermore, TRPM4-mediated Na+ influx was shown to regulate cell viability and the cell cycle of HCT116 cells. In the same study, TRPM4 was also linked to the regulation of CRC cells migration and invasion [122].
There are two Mg2+ channels in the superfamily of TRP channels, TRPM6 and TRPM7, and their activity is linked to the pathogenesis of CRC. TRPM6 and TRPM7 are unique ion channels that mediate Mg2+ homeostasis, as well as proteins, combining an ion channel with a functional α-kinase domain [29,142,143,144,145]. The role of Mg2+ in the pathogenesis of CRC has been reported. Higher Mg2+ intake seems to be associated with a modest reduction in the risk of CRC, in particular, colon cancer [146]. However, studies showed that mice receiving Mg2+-deficient diet had a significant retardation of their primary tumor growth [147]. Therefore, the relationship between Mg2+ intake and cancer might be more complex, especially since the relationship of Ca2+:Mg2+ intake is important for the regulation of cellular responses [148]. At the cellular level, Mg2+ can exhibit either anti- or pro-tumor effects and has been suggested to contribute to different mechanisms involved in carcinogenesis, such as cancer cell proliferation, metabolic reprogramming, the ability to metastasize, and neo-angiogenesis [149,150]. Mg2+ was shown to contribute to the regulation of cell proliferation and cell cycle, and Mg2+ deficiency has been shown to induce cell cycle arrest in the G0/G1 phase [148]. On these bases, TRPM6 and TRPM7 could be potential players in CRC. Indeed, it was reported that TRPM6 is downregulated in CRC tissues on mRNA level [120,123], and its higher expression is correlated with an increased overall survival [123]. On the other hand, TRPM7 was suggested to be upregulated in CRC on mRNA level [151]. Moreover, a single-nucleotide polymorphism that substitutes TRPM7 threonine 1482 for isoleucine (T1482I) increases the risk of the development of colon cancer, particularly in patients with a high Ca2+/Mg2+ ratio [152]. It must be noted that this mutation does not influence kinase activity, nor the channel’s conductivity, but the channel’s sensitivity to Mg2+ [153]. TRPM7 was also shown to be overexpressed in CRC cell lines. The downregulation of TRPM7 suppressed CRC cell proliferation, migration, and invasion, as well as triggered cell cycle arrest in the G0/G1 phase, reduced the S phase, and promoted apoptosis. Furthermore, the decrease in TRPM7 expression in CRC cells reversed EMT, which was accompanied by a downregulation in N-cadherin and an upregulation of E-cadherin expression [151]. Mg2+ homeostasis regulation via TRPM6 and TRPM7 was also linked to the sensitivity of CRC cells to doxorubicin, a common chemotherapeutic agent. A study in the CRC cell line LoVo showed that cells resistant to doxorubicin have increased total intracellular Mg2+ levels compared to sensitive LoVo cells. However, mRNA expression of TRPM6 in cells resistant to doxorubicin was downregulated compared to cells sensitive to doxorubicin. Further, in LoVo cells resistant to doxorubicin, TRPM7 is downregulated on protein level, but not on mRNA level, compared to doxorubicin-sensitive LoVo. Therefore, the levels of TRPM6 and TRPM7 are inversely related to the amounts of total intracellular Mg2+. This could be partially explained by the fact that TRPM7 was markedly reduced in resistant cells because of the activation of calpains, which are dependent on intracellular Ca2+ levels. Free Ca2+ levels are higher in cells resistant to doxorubicin [154]. These findings further support the idea that pathogenesis of CRC might be dependent on the Ca2+/Mg2+ ratio [152] and suggest a potential interplay between Ca2+ and Mg2+ transport (dys-) regulation in CRC. Nevertheless, TRPM7 is involved in the modulation of drug resistance in LoVo cells, as downregulation of TRPM7, but not TRPM6, expression enhances viability of LoVo cells exposed to doxorubicin [154].

8. Conclusions and Outlook

Recent studies show that digestive malignancies are characterized by different expression patterns of TRP channels (Figure 1). In general, cancers of the gastrointestinal tract are characterized by a poor prognosis, largely due to late diagnosis, when the disease is at an advanced stage. Therefore, there is a need to develop novel biomarkers for the detection of early stage diseases. Whether TRP channels could play such a role remains under investigation, as recent evidence is limited. For example, TRPM7 is upregulated in pre-malignant pancreatic cancer compared to normal pancreatic tissue [103], and so it could potentially serve as a candidate for a pancreatic cancer biomarker. Another candidate for a novel biomarker could be TRPM4, which was shown to be highly expressed in tumor buds in CRC cancer tissues [122]. Currently, there is a need for the introduction of novel additional biomarkers, which could support the TNM staging system [155]. Tumor budding has been suggested to be an additional prognostic factor, as it is strongly predictive of lymph node metastases, recurrence, and cancer-related death at 5 years [156,157]. Therefore, the expression of TRPM4 could potentially add to the detection of tumor buds.
Moreover, TRP channels could be potential candidates for therapeutic targets, especially since they are usually expressed on the cell surface, which makes them accessible to small-molecule inhibitors and biological molecules, such as monoclonal antibodies and fusion proteins [158,159,160]. Given the fact that some TRP channels are ubiquitously expressed, the delivery of TRP channel inhibitors to the tumor sites could be supported by cancer cell-specific drug delivery systems that are currently being developed [161]. TRP channels were shown to regulate cellular responses associated with tumorigenesis (summary in Figure 2). A number of studies show the anti-tumor properties of capsaicin, a TRPV1 channel agonist [106,114,121,133,135]. These findings suggest that the activation of TRPV1 could potentially enhance standard therapies. Furthermore, targeting other TRP channels showed an amplification of the effects of commonly used chemotherapeutics [77,96,101,115,126,154], potentially providing an option for the enhancement of current therapeutic strategies. Currently, several clinical trials evaluating ion channel inhibitors and blockers are ongoing [158,162,163]. Furthermore, phase 1 clinical studies evaluating TRPV6 channel inhibitor, SOR-C13, in patients with advanced solid epithelial tumors are being evaluated (NCT01578564 and NCT03784677). While the second study is in the recruiting stage, the results of the first study show that that SOR-C13 was well tolerated. Moreover, of 22 evaluable patients, 54.5% showed stable disease ranging from 2.8 to 12.5 months, which suggests an anti-tumor activity of SOR-C13 [164]. This is a proof of principle, showing that, in the future, drugs targeting TRP channels in cancer could enter clinical use. Future studies will further investigate the potential of TRP channels as therapeutic targets in cancer.

Author Contributions

Concept and idea C.P. and P.S.; preparation of figures and tables P.S.; A.B., S.K., C.P. and P.S. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

We acknowledge funding by the Swiss National Science Foundation (NCCR TransCure and 31003A_173155/1).

Acknowledgments

Figure 1 was generated with Smart PPT (https://smart.servier.com/).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5-FU5-Fluorouracil
ABCB1ATP-Binding Cassette Subfamily B Member 1
Ca2+Calcium Ions
CACColitis-Associated Cancer
CRCColorectal Cancer
Cryo-EMCryogenic Electron Microscopy
EACEsophageal Adenocarcinoma
EGFREpidermal Growth Factor Receptor
EMTEndothelial–Mesenchymal Transition
EREndoplasmic Reticulum
ESCCEsophageal Squamous Cell Carcinoma
HCCHepatocellular Carcinoma
IHCImmunohistochemistry
La3+Lanthanum Ions
LCSLCsLiver Cancer Stem-Like Cells
Mg2+Magnesium Ion
Na+Sodium Ions
NETNeuroendocrine Tumor
ROSReactive Oxygen Species
SCCSquamus Cell Carcinoma
TNMTumor Node Metastasis
TRP Transient Receptor Potential
TRPATransient Receptor Potential Ankyrin
TRPCTransient Receptor Potential Canonical
TRPMTransient Receptor Potential Melastatin
TRPMLTransient Receptor Potential Mucolipin
TRPPTransient Receptor Potential Polycystic
TRPVTransient Receptor Potential Vanilloid
WBWestern-Blot

References

  1. Bray, F.; Ferlay, J.; Soerjomataram, I.; Siegel, R.L.; Torre, L.A.; Jemal, A. Global cancer statistics 2018: GLOBOCAN estimates of incidence and mortality worldwide for 36 cancers in 185 countries. CA. Cancer J. Clin. 2018, 68, 394–424. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2020. CA. Cancer J. Clin. 2020, 70, 7–30. [Google Scholar] [CrossRef] [PubMed]
  3. Issa, I.A.; NouredDine, M. Colorectal cancer screening: An updated review of the available options. World J. Gastroenterol. 2017, 23, 5086–5096. [Google Scholar] [CrossRef] [PubMed]
  4. Griffin-Sobel, J.P. Gastrointestinal Cancers: Screening and Early Detection. Semin. Oncol. Nurs. 2017, 33, 165–171. [Google Scholar] [CrossRef]
  5. Jitender, S.; Sarika, G.; Varada, H.R.; Omprakash, Y.; Mohsin, K. Screening for oral cancer. J. Exp. Ther. Oncol. 2016, 11, 303–307. [Google Scholar]
  6. Olsen, J.R.; Apisarnthanarax, S.; Murphy, J.D.; Tait, D.; Huguet, F.; Hallemeier, C.L.; Jabbour, S.K. Gastrointestinal Cancers: Fine-Tuning the Management of Rectal, Esophageal, and Pancreas Cancers. Int. J. Radiat. Oncol. Biol. Phys. 2019, 105, 1–10. [Google Scholar] [CrossRef]
  7. Wang, W.; Kandimalla, R.; Huang, H.; Zhu, L.; Li, Y.; Gao, F.; Goel, A.; Wang, X. Molecular subtyping of colorectal cancer: Recent progress, new challenges and emerging opportunities. Semin. Cancer Biol. 2019, 55, 37–52. [Google Scholar] [CrossRef]
  8. Ilson, D.H. Advances in the treatment of gastric cancer. Curr. Opin. Gastroenterol. 2018, 34, 465–468. [Google Scholar] [CrossRef]
  9. Aziz, M.A.; Yousef, Z.; Saleh, A.M.; Mohammad, S.; Al Knawy, B. Towards personalized medicine of colorectal cancer. Crit. Rev. Oncol. Hematol. 2017, 118, 70–78. [Google Scholar] [CrossRef]
  10. Jácome, A.A.; Coutinho, A.K.; Lima, E.M.; Andrade, A.C.; Dos Santos, J.S. Personalized medicine in gastric cancer: Where are we and where are we going? World J. Gastroenterol. 2016, 22, 1160–1171. [Google Scholar] [CrossRef]
  11. Molinari, C.; Marisi, G.; Passardi, A.; Matteucci, L.; De Maio, G.; Ulivi, P. Heterogeneity in colorectal cancer: A challenge for personalized medicine? Int. J. Mol. Sci. 2018, 19, 3733. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Galun, D.; Srdic-Rajic, T.; Bogdanovic, A.; Loncar, Z.; Zuvela, M. Targeted therapy and personalized medicine in hepatocellular carcinoma: Drug resistance, mechanisms, and treatment strategies. J. Hepatocell. Carcinoma 2017, 4, 93–103. [Google Scholar] [CrossRef] [Green Version]
  13. Yu, I.S.; Cheung, W.Y. Metastatic Colorectal Cancer in the Era of Personalized Medicine: A More Tailored Approach to Systemic Therapy. Can. J. Gastroenterol. Hepatol. 2018, 2018, 9450754. [Google Scholar] [CrossRef] [PubMed]
  14. Hanahan, D.; Weinberg, R.A. Hallmarks of cancer: The next generation. Cell 2011, 144, 646–674. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Anderson, K.J.; Cormier, R.T.; Scott, P.M. Role of ion channels in gastrointestinal cancer. World J. Gastroenterol. 2019, 25, 5732–5772. [Google Scholar] [CrossRef]
  16. Cosens, D.J.; Manning, A. Abnormal electroretinogram from a Drosophila mutant. Nature 1969, 224, 285–287. [Google Scholar] [CrossRef]
  17. Montell, C.; Rubin, G.M. Molecular characterization of the drosophila trp locus: A putative integral membrane protein required for phototransduction. Neuron 1989, 2, 1313–1323. [Google Scholar] [CrossRef]
  18. Nilius, B.; Owsianik, G. The transient receptor potential family of ion channels. Genome Biol. 2011, 12, 218. [Google Scholar] [CrossRef] [Green Version]
  19. Ramsey, I.S.; Delling, M.; Clapham, D.E. An Introduction to Trp Channels. Annu. Rev. Physiol. 2006, 68, 619–647. [Google Scholar] [CrossRef] [Green Version]
  20. Vandewauw, I.; De Clercq, K.; Mulier, M.; Held, K.; Pinto, S.; Van Ranst, N.; Segal, A.; Voet, T.; Vennekens, R.; Zimmermann, K.; et al. A TRP channel trio mediates acute noxious heat sensing. Nature 2018, 555, 662–666. [Google Scholar] [CrossRef]
  21. Hung, C.Y.; Tan, C.H. TRP channels in nociception and pathological pain. In Advances in Pain Research: Mechanisms and Modulation of Chronic Pain. Adv. Exp. Med. Biol. 2018, 1099, 13–27. [Google Scholar] [PubMed]
  22. Moore, C.; Gupta, R.; Jordt, S.E.; Chen, Y.; Liedtke, W.B. Regulation of Pain and Itch by TRP Channels. Neurosci. Bull. 2018, 34, 120–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Wagner, T.F.J.; Loch, S.; Lambert, S.; Straub, I.; Mannebach, S.; Mathar, I.; Düfer, M.; Lis, A.; Flockerzi, V.; Philipp, S.E.; et al. Transient receptor potential M3 channels are ionotropic steroid receptors in pancreatic β cells. Nat. Cell Biol. 2008, 10, 1421–1430. [Google Scholar] [CrossRef]
  24. Roper, S.D. TRPs in taste and chemesthesis. Handb. Exp. Pharmacol. 2014, 223, 827–871. [Google Scholar]
  25. Gao, Y.; Liao, P. TRPM4 channel and cancer. Cancer Lett. 2019, 454, 66–69. [Google Scholar] [CrossRef]
  26. Khalil, M.; Alliger, K.; Weidinger, C.; Yerinde, C.; Wirtz, S.; Becker, C.; Engel, M.A. Functional role of transient receptor potential channels in immune cells and epithelia. Front. Immunol. 2018, 9, 174. [Google Scholar] [CrossRef]
  27. Santoni, G.; Maggi, F.; Morelli, M.B.; Santoni, M.; Marinelli, O. Transient Receptor Potential Cation Channels in Cancer Therapy. Med. Sci. 2019, 7, 108. [Google Scholar] [CrossRef] [Green Version]
  28. Wong, K.K.; Banham, A.H.; Yaacob, N.S.; Nur Husna, S.M. The oncogenic roles of TRPM ion channels in cancer. J. Cell. Physiol. 2019, 234, 14556–14573. [Google Scholar] [CrossRef]
  29. Romagnani, A.; Vettore, V.; Rezzonico-Jost, T.; Hampe, S.; Rottoli, E.; Nadolni, W.; Perotti, M.; Meier, M.A.; Hermanns, C.; Geiger, S.; et al. TRPM7 kinase activity is essential for T cell colonization and alloreactivity in the gut. Nat. Commun. 2017, 8, 1–14. [Google Scholar] [CrossRef]
  30. Nadolni, W.; Zierler, S. The Channel-Kinase TRPM7 as Novel Regulator of Immune System Homeostasis. Cells 2018, 7, 109. [Google Scholar] [CrossRef] [Green Version]
  31. Yu, H.; Zhang, Z.; Lis, A.; Penner, R.; Fleig, A. TRPM7 is regulated by halides through its kinase domain. Cell. Mol. Life Sci. 2013, 70, 2757–2771. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Hantute-Ghesquier, A.; Haustrate, A.; Prevarskaya, N.; Lehen’kyi, V. TRPM family channels in cancer. Pharmaceuticals 2018, 11, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Sterea, A.M.; Egom, E.E.; El Hiani, Y. TRP channels in gastric cancer: New hopes and clinical perspectives. Cell Calcium 2019, 82, 102053. [Google Scholar] [CrossRef]
  34. Fels, B.; Bulk, E.; Pethő, Z.; Schwab, A. The role of TRP channels in the metastatic cascade. Pharmaceuticals 2018, 11, 48. [Google Scholar] [CrossRef] [Green Version]
  35. Kim, S.Y.; Yang, D.; Myeong, J.; Ha, K.; Kim, S.H.; Park, E.J.; Kim, I.G.; Cho, N.H.; Lee, K.P.; Jeon, J.H.; et al. Regulation of calcium influx and signaling pathway in cancer cells via TRPV6-Numb1 interaction. Cell Calcium 2013, 53, 102–111. [Google Scholar] [CrossRef]
  36. Prevarskaya, N.; Ouadid-Ahidouch, H.; Skryma, R.; Shuba, Y. Remodelling of Ca2+ transport in cancer: How it contributes to cancer hallmarks? Philos. Trans. R. Soc. B Biol. Sci. 2014, 369, 1–10. [Google Scholar] [CrossRef] [Green Version]
  37. 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, 7380–7385. [Google Scholar] [CrossRef] [Green Version]
  38. Thebault, S.; Flourakis, M.; Vanoverberghe, K.; Vandermoere, F.; Roudbaraki, M.; Lehen’kyi, V.; Slomianny, C.; Beck, B.; Mariot, P.; Bonnal, J.L.; et al. Differential role of transient receptor potential channels in Ca 2+ entry and proliferation of prostate cancer epithelial cells. Cancer Res. 2006, 66, 2038–2047. [Google Scholar] [CrossRef] [Green Version]
  39. Li, M.; Yu, Y.; Yang, J. Structural Biology of TRP Channels. In Transient Receptor Potential Channels. Advances in Experimental Medicine and Biology; Islam, M.S., Ed.; Springer: Dordrecht, The Netherlands, 2011; Volume 704, ISBN 978-94-007-0265-3. [Google Scholar]
  40. Huang, Y.; Fliegert, R.; Guse, A.H.; Lü, W.; Du, J. A structural overview of the ion channels of the TRPM family. Cell Calcium 2020, 85, 102111. [Google Scholar] [CrossRef]
  41. Saotome, K.; Singh, A.K.; Yelshanskaya, M.V.; Sobolevsky, A.I. Crystal structure of the epithelial calcium channel TRPV6. Nature 2016, 534, 506–511. [Google Scholar] [CrossRef] [Green Version]
  42. Guo, J.; She, J.; Zeng, W.; Chen, Q.; Bai, X.-C.; Jiang, Y. Structures of the calcium-activated, non-selective cation channel TRPM4. Nature 2017, 552, 205–209. [Google Scholar] [CrossRef]
  43. Autzen, H.E.; Myasnikov, A.G.; Campbell, M.G.; Asarnow, D.; Julius, D.; Cheng, Y. Structure of the human TRPM4 ion channel in a lipid nanodisc. Science 2018, 359, 228–232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Vinayagam, D.; Mager, T.; Apelbaum, A.; Bothe, A.; Merino, F.; Hofnagel, O.; Gatsogiannis, C.; Raunser, S. Electron cryo-microscopy structure of the canonical TRPC4 ion channel. Elife 2018, 7, e36615. [Google Scholar] [CrossRef] [PubMed]
  45. Fan, C.; Choi, W.; Sun, W.; Du, J.; Lü, W.; Lu, W. Structure of the human lipid-gated cation channel TRPC3. Elife 2018, 7, e36852. [Google Scholar] [CrossRef] [PubMed]
  46. Duan, J.; Li, Z.; Li, J.; Santa-Cruz, A.; Sanchez-Martinez, S.; Zhang, J.; Clapham, D.E. Structure of full-length human TRPM4. Proc. Natl. Acad. Sci. USA 2018, 115, 2377–2382. [Google Scholar] [CrossRef] [Green Version]
  47. Yin, Y.; Wu, M.; Zubcevic, L.; Borschel, W.F.; Lander, G.C.; Lee, S.Y. Structure of the cold- And menthol-sensing ion channel TRPM8. Science 2018, 359, 237–241. [Google Scholar] [CrossRef] [Green Version]
  48. Hughes, T.E.T.; Lodowski, D.T.; Huynh, K.W.; Yazici, A.; Del Rosario, J.; Kapoor, A.; Basak, S.; Samanta, A.; Han, X.; Chakrapani, S.; et al. Structural basis of TRPV5 channel inhibition by econazole revealed by cryo-EM. Nat. Struct. Mol. Biol. 2018, 25, 53–60. [Google Scholar] [CrossRef]
  49. Duan, J.; Li, Z.; Li, J.; Hulse, R.E.; Santa-Cruz, A.; Valinsky, W.C.; Abiria, S.A.; Krapivinsky, G.; Zhang, J.; Clapham, D.E. Structure of the mammalian TRPM7, a magnesium channel required during embryonic development. Proc. Natl. Acad. Sci. USA 2018, 115, E8201–E8210. [Google Scholar] [CrossRef] [Green Version]
  50. Deng, Z.; Paknejad, N.; Maksaev, G.; Sala-Rabanal, M.; Nichols, C.G.; Hite, R.K.; Yuan, P. Cryo-EM and X-ray structures of TRPV4 reveal insight into ion permeation and gating mechanisms. Nat. Struct. Mol. Biol. 2018, 25, 252–260. [Google Scholar] [CrossRef]
  51. Duan, J.; Li, J.; Chen, G.L.; Ge, Y.; Liu, J.; Xie, K.; Peng, X.; Zhou, W.; Zhong, J.; Zhang, Y.; et al. Cryo-EM structure of TRPC5 at 2.8-Å resolution reveals unique and conserved structural elements essential for channel function. Sci. Adv. 2019, 5, eaaw7935. [Google Scholar] [CrossRef] [Green Version]
  52. Duan, J.; Li, J.; Zeng, B.; Chen, G.L.; Peng, X.; Zhang, Y.; Wang, J.; Clapham, D.E.; Li, Z.; Zhang, J. Structure of the mouse TRPC4 ion channel. Nat. Commun. 2018, 9, 1–10. [Google Scholar] [CrossRef] [Green Version]
  53. Singh, A.K.; McGoldrick, L.L.; Sobolevsky, A.I. Structure and gating mechanism of the transient receptor potential channel TRPV3. Nat. Struct. Mol. Biol. 2018, 25, 805–813. [Google Scholar] [CrossRef] [PubMed]
  54. Paulsen, C.E.; Armache, J.P.; Gao, Y.; Cheng, Y.; Julius, D. Structure of the TRPA1 ion channel suggests regulatory mechanisms. Nature 2015, 520, 511–517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Montero, P.H.; Patel, S.G. Cancer of the Oral Cavity. Surg. Oncol. Clin. N. Am. 2015, 24, 491–508. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Surh, Y.J.; Lee, E.; Lee, J.M. Chemoprotective properties of some pungent ingredients present in red pepper and ginger. Mutat. Res. Fundam. Mol. Mech. Mutagen. 1998, 402, 259–267. [Google Scholar] [CrossRef]
  57. Liu, Z.; Shen, C.; Tao, Y.; Wang, S.; Wei, Z.; Cao, Y.; Wu, H.; Fan, F.; Lin, C.; Shan, Y.; et al. Chemopreventive efficacy of menthol on carcinogen-induced cutaneous carcinoma through inhibition of inflammation and oxidative stress in mice. Food Chem. Toxicol. 2015, 82, 12–18. [Google Scholar] [CrossRef]
  58. Dwivedi, C.; Muller, L.A.; Goetz-Parten, D.E.; Kasperson, K.; Mistry, V.V. Chemopreventive effects of dietary mustard oil on colon tumor development. Cancer Lett. 2003, 196, 29–34. [Google Scholar] [CrossRef]
  59. Marincsák, R.; Tóth, B.; Czifra, G.; Márton, I.; Rédl, P.; Tar, I.; Tóth, L.; Kovács, L.; Bíró, T. Increased expression of TRPV1 in squamous cell carcinoma of the human tongue. Oral Dis. 2009, 15, 328–335. [Google Scholar] [CrossRef]
  60. Okamoto, Y.; Ohkubo, T.; Ikebe, T.; Yamazaki, J. Blockade of TRPM8 activity reduces the invasion potential of oral squamous carcinoma cell lines. Int. J. Oncol. 2012, 40, 1431–1440. [Google Scholar]
  61. Sakakibara, A.; Sakakibara, S.; Kusumoto, J.; Takeda, D.; Hasegawa, T.; Akashi, M.; Minamikawa, T.; Hashikawa, K.; Terashi, H.; Komori, T. Upregulated expression of transient receptor potential cation channel subfamily v receptors in mucosae of patients with oral squamous cell carcinoma and patients with a history of alcohol consumption or smoking. PLoS ONE 2017, 12, e0169723. [Google Scholar] [CrossRef]
  62. Zhao, L.Y.; Xu, W.L.; Xu, Z.Q.; Qi, C.; Li, Y.; Cheng, J.; Liu, L.K.; Wu, Y.N.; Gao, J.; Ye, J.H. The overexpressed functional transient receptor potential channel TRPM2 in oral squamous cell carcinoma. Sci. Rep. 2016, 6, 38471. [Google Scholar] [CrossRef] [PubMed]
  63. Gonzales, C.B.; Kirma, N.B.; De La Chapa, J.J.; Chen, R.; Henry, M.A.; Luo, S.; Hargreaves, K.M. Vanilloids induce oral cancer apoptosis independent of TRPV1. Oral Oncol. 2014, 50, 437–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. De La Chapa, J.J.; Singha, P.K.; Lee, D.R.; Gonzales, C.B. Thymol inhibits oral squamous cell carcinoma growth via mitochondria-mediated apoptosis. J. Oral Pathol. Med. 2018, 47, 674–682. [Google Scholar] [CrossRef]
  65. Ruparel, S.; Bendele, M.; Wallace, A.; Green, D. Released lipids regulate transient receptor potential channel (TRP)-dependent oral cancer pain. Mol. Pain 2015, 11, 30. [Google Scholar] [CrossRef] [Green Version]
  66. Fliegert, R.; Bauche, A.; Wolf Pérez, A.M.; Watt, J.M.; Rozewitz, M.D.; Winzer, R.; Janus, M.; Gu, F.; Rosche, A.; Harneit, A.; et al. 2′-Deoxyadenosine 5′-diphosphoribose is an endogenous TRPM2 superagonist. Nat. Chem. Biol. 2017, 13, 1036–1044. [Google Scholar] [CrossRef] [Green Version]
  67. Kolisek, M.; Beck, A.; Fleig, A.; Penner, R. Cyclic ADP-ribose and hydrogen peroxide synergize with ADP-ribose in the activation of TRPM2 channels. Mol. Cell 2005, 18, 61–69. [Google Scholar] [CrossRef]
  68. Zaidi, N.; Kelly, R.J. The management of localized esophageal squamous cell carcinoma: Western approach. Chin. Clin. Oncol. 2017, 6, 46. [Google Scholar] [CrossRef]
  69. Kim, J.J.; Bowlby, R.; Mungall, A.J.; Robertson, A.G.; Odze, R.D.; Cherniack, A.D.; Shih, J.; Pedamallu, C.S.; Cibulskis, C.; Dunford, A.; et al. Integrated genomic characterization of oesophageal carcinoma. Nature 2017, 541, 169–174. [Google Scholar]
  70. Wang, K.; Johnson, A.; Ali, S.M.; Klempner, S.J.; Bekaii-Saab, T.; Vacirca, J.L.; Khaira, D.; Yelensky, R.; Chmielecki, J.; Elvin, J.A.; et al. Comprehensive Genomic Profiling of Advanced Esophageal Squamous Cell Carcinomas and Esophageal Adenocarcinomas Reveals Similarities and Differences. Oncologist 2015, 20, 1132–1139. [Google Scholar] [CrossRef] [Green Version]
  71. Shi, Y.; Ding, X.; He, Z.H.; Zhou, K.C.; Wang, Q.; Wang, Y.Z. Critical role of TRPC6 channels in G2 phase transition and the development of human oesophageal cancer. Gut 2009, 58, 1443–1450. [Google Scholar] [CrossRef]
  72. Lan, X.; Zhao, J.; Song, C.; Yuan, Q.; Liu, X. TRPM8 facilitates proliferation and immune evasion of esophageal cancer cells. Biosci. Rep. 2019, 39. [Google Scholar] [CrossRef] [Green Version]
  73. Nakashima, S.; Shiozaki, A.; Ichikawa, D.; Hikami, S.; Kosuga, T.; Konishi, H.; Komatsu, S.; Fujiwara, H.; Okamoto, K.; Kishimoto, M.; et al. Transient Receptor Potential Melastatin 7 as an Independent Prognostic Factor in Human Esophageal Squamous Cell Carcinoma. Anticancer Res. 2017, 37, 1161–1167. [Google Scholar]
  74. Kudou, M.; Shiozaki, A.; Yamazato, Y.; Katsurahara, K.; Kosuga, T.; Shoda, K.; Arita, T.; Konishi, H.; Komatsu, S.; Kubota, T.; et al. The expression and role of TRPV2 in esophageal squamous cell carcinoma. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Huang, R.; Wang, F.; Yang, Y.; Ma, W.; Lin, Z.; Cheng, N.; Long, Y.; Deng, S.; Li, Z. Recurrent activations of transient receptor potential vanilloid-1 and vanilloid-4 promote cellular proliferation and migration in esophageal squamous cell carcinoma cells. FEBS Open Bio 2019, 9, 206–225. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. El Boustany, C.; Bidaux, G.; Enfissi, A.; Delcourt, P.; Prevarskaya, N.; Capiod, T. Capacitative calcium entry and transient receptor potential canonical 6 expression control human hepatoma cell proliferation. Hepatology 2008, 47, 2068–2077. [Google Scholar] [CrossRef] [PubMed]
  77. Miao, X.; Liu, G.G.; Xu, X.; Xie, C.; Sun, F.; Yang, Y.; Zhang, T.; Hua, S.; Fan, W.; Li, Q.; et al. High expression of vanilloid receptor-1 is associated with better prognosis of patients with hepatocellular carcinoma. Cancer Genet. Cytogenet. 2008, 186, 25–32. [Google Scholar] [CrossRef] [PubMed]
  78. Liu, G.G.; Xie, C.; Sun, F.; Xu, X.; Yang, Y.; Zhang, T.; Deng, Y.; Wang, D.; Huang, Z.; Yang, L.; et al. Clinical significance of transient receptor potential vanilloid 2 expression in human hepatocellular carcinoma. Cancer Genet. Cytogenet. 2010, 197, 54–59. [Google Scholar] [CrossRef]
  79. Fang, Y.; Liu, G.G.; Xie, C.; Qian, K.; Lei, X.; Liu, Q.; Liu, G.G.; Cao, Z.; Fu, J.; Du, H.; et al. Pharmacological inhibition of TRPV4 channel suppresses malignant biological behavior of hepatocellular carcinoma via modulation of ERK signaling pathway. Biomed. Pharmacother. 2018, 101, 910–919. [Google Scholar] [CrossRef]
  80. Xu, J.; Yang, Y.; Xie, R.; Liu, J.; Nie, X.; An, J.; Wen, G.; Liu, X.; Jin, H.; Tuo, B. The ncx1/trpc6 complex mediates tgfb-driven migration and invasion of human hepatocellular carcinoma cells. Cancer Res. 2018, 78, 2564–2576. [Google Scholar] [CrossRef] [Green Version]
  81. Waning, J.; Vriens, J.; Owsianik, G.; Stüwe, L.; Mally, S.; Fabian, A.; Frippiat, C.; Nilius, B.; Schwab, A. A novel function of capsaicin-sensitive TRPV1 channels: Involvement in cell migration. Cell Calcium 2007, 42, 17–25. [Google Scholar] [CrossRef]
  82. 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, 19–28. [Google Scholar] [CrossRef] [PubMed]
  83. Huang, S.P.; Chen, J.C.; Wu, C.C.; Chen, C.T.; Tang, N.Y.; Ho, Y.T.; Lo, C.; Lin, J.P.J.G.; Chung, J.G.; Lin, J.P.J.G. Capsaicin-induced apoptosis in human hepatoma HepG2 cells. Anticancer Res. 2009, 29, 165–174. [Google Scholar] [PubMed]
  84. Chen, W.T.; Lin, G.B.; Lin, S.H.; Lu, C.H.; Hsieh, C.H.; Ma, B.L.; Chao, C.Y. Static magnetic field enhances the anticancer efficacy of capsaicin on HepG2 cells via capsaicin receptor TRPV1. PLoS ONE 2018, 13, 1–15. [Google Scholar] [CrossRef] [Green Version]
  85. Ma, W.; Li, C.; Yin, S.; Liu, J.; Gao, C.; Lin, Z.; Huang, R.; Huang, J.; Li, Z. Novel role of TRPV2 in promoting the cytotoxicity of H2O2-mediated oxidative stress in Human hepatoma cells. Free Radic. Biol. Med. 2015, 89, 1003–1013. [Google Scholar] [CrossRef]
  86. Hu, Z.; Cao, X.; Fang, Y.; Liu, G.; Xie, C.; Qian, K.; Lei, X.; Cao, Z.; Du, H.; Cheng, X.; et al. Transient receptor potential vanilloid-type 2 targeting on stemness in liver cancer. Biomed. Pharmacother. 2018, 105, 697–706. [Google Scholar] [CrossRef]
  87. Selli, C.; Erac, Y.; Kosova, B.; Erdal, E.S.; Tosun, M. Silencing of TRPC1 regulates store-operated calcium entry and proliferation in Huh7 hepatocellular carcinoma cells. Biomed. Pharmacother. 2015, 71, 194–200. [Google Scholar] [CrossRef]
  88. Wen, L.; Liang, C.; Chen, E.; Chen, W.; Liang, F.; Zhi, X.; Wei, T.; Xue, F.; Li, G.; Yang, Q.; et al. Regulation of Multi-drug Resistance in hepatocellular carcinoma cells is TRPC6/Calcium Dependent. Sci. Rep. 2016, 6, 1–14. [Google Scholar] [CrossRef] [Green Version]
  89. Chen, Y.; Yu, Y.; Sun, S.; Wang, Z.; Liu, P.; Liu, S.; Jiang, J. Bradykinin promotes migration and invasion of hepatocellular carcinoma cells through TRPM7 and MMP-2. Exp. Cell Res. 2016, 349, 68–76. [Google Scholar] [CrossRef]
  90. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics, 2015. CA. Cancer J. Clin. 2015, 65, 5–29. [Google Scholar] [CrossRef]
  91. Ansari, D.; Tingstedt, B.; Andersson, B.; Holmquist, F.; Sturesson, C.; Williamsson, C.; Sasor, A.; Borg, D.; Bauden, M.; Andersson, R. Pancreatic cancer: Yesterday, today and tomorrow. Futur. Oncol. 2016, 12, 1929–1946. [Google Scholar] [CrossRef] [Green Version]
  92. Lin, R.; Wang, Y.; Chen, Q.; Liu, Z.; Xiao, S.; Wang, B.; Shi, B. TRPM2 promotes the proliferation and invasion of pancreatic ductal adenocarcinoma. Mol. Med. Rep. 2018, 17, 7537–7544. [Google Scholar] [CrossRef] [PubMed]
  93. Mergler, S.; Strowski, M.Z.; Kaiser, S.; Plath, T.; Giesecke, Y.; Neumann, M.; Hosokawa, H.; Kobayashi, S.; Langrehr, J.; Neuhaus, P.; et al. Transient receptor potential channel TRPM8 agonists stimulate calcium influx and neurotensin secretion in neuroendocrine tumor cells. Neuroendocrinology 2007, 85, 81–92. [Google Scholar] [CrossRef] [PubMed]
  94. Yee, N.; Li, Q.; Kazi, A.; Yang, Z.; Berg, A.; Yee, R. Aberrantly Over-Expressed TRPM8 Channels in Pancreatic Adenocarcinoma: Correlation with Tumor Size/Stage and Requirement for Cancer Cells Invasion. Cells 2014, 3, 500–516. [Google Scholar] [CrossRef] [PubMed]
  95. Yee, N.S.; Zhou, W.; Lee, M. Transient receptor potential channel TRPM8 is over-expressed and required for cellular proliferation in pancreatic adenocarcinoma. Cancer Lett. 2010, 297, 49–55. [Google Scholar] [CrossRef] [Green Version]
  96. Liu, J.F.; Hu, G.H.; Gong, Y.J.; Yu, Q.L.; He, B.; Li, W.H.; He, Z.G.Z.C.; Hao, W.J.; He, Z.G.Z.C.; Liu, Y.P. Silencing of TRPM8 inhibits aggressive tumor phenotypes and enhances gemcitabine sensitivity in pancreatic cancer. Pancreatology 2018, 18, 935–944. [Google Scholar] [CrossRef]
  97. Du, J.D.; Zheng, X.; Chen, Y.L.; Huang, Z.Q.; Cai, S.W.; Jiao, H.B.; Zhu, Z.M.; Hu, B. Elevated transient receptor potential melastatin 8 (Trpm8) expression is correlated with poor prognosis in pancreatic cancer. Med. Sci. Monit. 2018, 24, 3720–3725. [Google Scholar] [CrossRef]
  98. Cucu, D.; Chiritoiu, G.; Petrescu, S.; Babes, A.; Stanica, L.; Duda, D.G.; Horii, A.; Dima, S.O.; Popescu, I. Characterization of functional transient receptor potential melastatin 8 channels in human pancreatic ductal adenocarcinoma cells. Pancreas 2014, 43, 795–800. [Google Scholar] [CrossRef]
  99. Ulăreanu, R.; Chiriţoiu, G.; Cojocaru, F.; Deftu, A.; Ristoiu, V.; Stănică, L.; Mihăilescu, D.F.; Cucu, D. N-glycosylation of the transient receptor potential melastatin 8 channel is altered in pancreatic cancer cells. Tumor Biol. 2017, 39, 1–10. [Google Scholar] [CrossRef] [Green Version]
  100. Yee, N.S.; Zhou, W.; Liang, I.C. Transient receptor potential ion channel Trpm7 regulates exocrine pancreatic epithelial proliferation by Mg2+-sensitive Socs3a signaling in development and cancer. DMM Dis. Model. Mech. 2011, 4, 240–254. [Google Scholar] [CrossRef] [Green Version]
  101. Yee, N.S.; Zhou, W.; Lee, M.; Yee, R.K. Targeted silencing of TRPM7 ion channel induces replicative senescence and produces enhanced cytotoxicity with gemcitabine in pancreatic adenocarcinoma. Cancer Lett. 2012, 318, 99–105. [Google Scholar] [CrossRef] [Green Version]
  102. Rybarczyk, P.; Gautier, M.; Hague, F.; Dhennin-Duthille, I.; Chatelain, D.; Kerr-Conte, J.; Pattou, F.; Regimbeau, J.M.; Sevestre, H.; Ouadid-Ahidouch, H. Transient receptor potential melastatin-related 7 channel is overexpressed in human pancreatic ductal adenocarcinomas and regulates human pancreatic cancer cell migration. Int. J. Cancer 2012, 131, 851–861. [Google Scholar] [CrossRef] [PubMed]
  103. Yee, N.S.; Kazi, A.A.; Li, Q.; Yang, Z.; Berg, A.; Yee, R.K. Aberrant over-expression of TRPM7 ion channels in pancreatic cancer: Required for cancer cell invasion and implicated in tumor growth and metastasis. Biol. Open 2015, 4, 507–514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Rybarczyk, P.; Vanlaeys, A.; Brassart, B.; Dhennin-Duthille, I.; Chatelain, D.; Sevestre, H.; Ouadid-Ahidouch, H.; Gautier, M. The Transient Receptor Potential Melastatin 7 Channel Regulates Pancreatic Cancer Cell Invasion through the Hsp90α/uPA/MMP-2 pathway. Neoplasia (United States) 2017, 19, 288–300. [Google Scholar] [CrossRef] [PubMed]
  105. Mergler, S.; Skrzypski, M.; Sassek, M.; Pietrzak, P.; Pucci, C.; Wiedenmann, B.; Strowski, M.Z. Thermo-sensitive transient receptor potential vanilloid channel-1 regulates intracellular calcium and triggers chromogranin A secretion in pancreatic neuroendocrine BON-1 tumor cells. Cell. Signal. 2012, 24, 233–246. [Google Scholar] [CrossRef]
  106. Skrzypski, M.; Sassek, M.; Abdelmessih, S.; Mergler, S.; Grötzinger, C.; Metzke, D.; Wojciechowicz, T.; Nowak, K.W.; Strowski, M.Z. Capsaicin induces cytotoxicity in pancreatic neuroendocrine tumor cells via mitochondrial action. Cell. Signal. 2014, 26, 41–48. [Google Scholar] [CrossRef]
  107. Huang, J.; Liu, J.; Qiu, L. Transient receptor potential vanilloid 1 promotes EGFR ubiquitination and modulates EGFR/MAPK signalling in pancreatic cancer cells. Cell Biochem. Funct. 2020, 1–8. [Google Scholar] [CrossRef]
  108. Skrzypski, M.; Kołodziejski, P.A.; Mergler, S.; Khajavi, N.; Nowak, K.W.; Strowski, M.Z. TRPV6 modulates proliferation of human pancreatic neuroendocrine BON-1 tumour cells. Biosci. Rep. 2016, 36, 1–9. [Google Scholar] [CrossRef] [Green Version]
  109. Song, H.; Dong, M.; Zhou, J.; Sheng, W.; Li, X.; Gao, W. Expression and prognostic significance of TRPV6 in the development and progression of pancreatic cancer. Oncol. Rep. 2018, 39, 1432–1440. [Google Scholar] [CrossRef] [Green Version]
  110. Van Cutsem, E.; Sagaert, X.; Topal, B.; Haustermans, K.; Prenen, H. Gastric cancer. Lancet 2016, 388, 2654–2664. [Google Scholar] [CrossRef]
  111. Cai, R.; Ding, X.; Zhou, K.; Shi, Y.; Ge, R.; Ren, G.; Jin, Y.; Wang, Y. Blockade of TRPC6 channels induced G2/M phase arrest and suppressed growth in human gastric cancer cells. Int. J. Cancer 2009, 125, 2281–2287. [Google Scholar] [CrossRef]
  112. Ge, P.; Wei, L.; Zhang, M.; Hu, B.; Wang, K.; Li, Y.Y.; Liu, S.; Wang, J.; Li, Y.Y. TRPC1/3/6 inhibition attenuates the TGF-β1-induced epithelial–mesenchymal transition in gastric cancer via the Ras/Raf1/ERK signaling pathway. Cell Biol. Int. 2018, 42, 975–984. [Google Scholar] [CrossRef] [PubMed]
  113. Ding, M.; Wang, H.; Qu, C.; Xu, F.; Zhu, Y.; Lv, G.; Lu, Y.; Zhou, Q.; Zhou, H.; Zeng, X.; et al. Pyrazolo[1,5-a]pyrimidine TRPC6 antagonists for the treatment of gastric cancer. Cancer Lett. 2018, 432, 47–55. [Google Scholar] [CrossRef] [PubMed]
  114. 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 Mol. Cell Res. 2007, 1773, 565–576. [Google Scholar] [CrossRef] [Green Version]
  115. Almasi, S.; Kennedy, B.E.; El-Aghil, M.; Sterea, A.M.; Gujar, S.; Partida-Sánchez, S.; El Hiani, Y. TRPM2 channel–mediated regulation of autophagy maintains mitochondrial function and promotes gastric cancer cell survival via the JNK-signaling pathway. J. Biol. Chem. 2018, 293, 3637–3650. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Kim, B.J.; Park, E.J.; Lee, J.H.; Jeon, J.H.; Kim, S.J.; So, I. Suppression of transient receptor potential melastatin 7 channel induces cell death in gastric cancer. Cancer Sci. 2008, 99, 2502–2509. [Google Scholar] [CrossRef] [PubMed]
  117. Ferlay, J.; Steliarova-Foucher, E.; Lortet-Tieulent, J.; Rosso, S.; Coebergh, J.W.W.; Comber, H.; Forman, D.; Bray, F. Cancer incidence and mortality patterns in Europe: Estimates for 40 countries in 2012. Eur. J. Cancer 2013, 49, 1374–1403. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Siegel, R.; Desantis, C.; Jemal, A. Colorectal Cancer Statistics, 2014. CA. Cancer J. Clin. 2014, 64, 104–117. [Google Scholar] [CrossRef]
  119. Sozucan, Y.; Kalender, M.E.; Sari, I.; Suner, A.; Oztuzcu, S.; Arman, K.; Yumrutas, O.; Bozgeyik, I.; Cengiz, B.; Igci, Y.Z.; et al. TRP genes family expression in colorectal cancer. Exp. Oncol. 2015, 37, 208–212. [Google Scholar] [CrossRef]
  120. Ibrahim, S.; Dakik, H.; Vandier, C.; Chautard, R.; Paintaud, G.; Mazurier, F.; Lecomte, T.; Guéguinou, M.; Raoul, W. Expression profiling of calcium channels and calcium-activated potassium channels in colorectal cancer. Cancers (Basel) 2019, 11, 561. [Google Scholar] [CrossRef] [Green Version]
  121. Hou, N.; He, X.; Yang, Y.; Fu, J.; Zhang, W.; Guo, Z.; Hu, Y.; Liang, L.; Xie, W.; Xiong, H.; et al. TRPV1 Induced apoptosis of colorectal cancer cells by activating calcineurin-NFAT2-p53 signaling pathway. Biomed Res. Int. 2019, 2019, 1–9. [Google Scholar] [CrossRef]
  122. Kappel, S.; Stokłosa, P.; Hauert, B.; Ross-Kaschitza, D.; Borgström, A.; Baur, R.; Galván, J.A.; Zlobec, I.; Peinelt, C. TRPM4 is highly expressed in human colorectal tumor buds and contributes to proliferation, cell cycle, and invasion of colorectal cancer cells. Mol. Oncol. 2019, 13, 2393–2405. [Google Scholar] [CrossRef] [PubMed]
  123. Xie, B.; Zhao, R.; Bai, B.; Wu, Y.; Xu, Y.; Lu, S.; Fang, Y.; Wang, Z.; Maswikiti, E.P.; Zhou, X.; et al. Identification of key tumorigenesis-related genes and their microRNAs in colon cancer. Oncol. Rep. 2018, 40, 3551–3560. [Google Scholar] [CrossRef] [PubMed]
  124. Sobradillo, D.; Hernández-Morales, M.; Ubierna, D.; Moyer, M.P.; Núñez, L.; Villalobos, C. A reciprocal shift in transient receptor potential channel 1 (TRPC1) and stromal interaction molecule 2 (STIM2) contributes to Ca2+remodeling and cancer hallmarks in colorectal carcinoma cells. J. Biol. Chem. 2014, 289, 28765–28782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Guéguinou, M.; Harnois, T.; Crottes, D.; Uguen, A.; Deliot, N.; Gambade, A.; Chantôme, A.; Haelters, J.P.; Jaffrès, P.A.; Jourdan, M.L.; et al. SK3/TRPC1/Orai1 complex regulates SOCE-dependent colon cancer cell migration: A novel opportunity to modulate anti- EGFR mAb action by the alkyl-lipid Ohmline. Oncotarget 2016, 7, 36168–36184. [Google Scholar] [CrossRef] [Green Version]
  126. Wang, T.; Chen, Z.; Zhu, Y.; Pan, Q.; Liu, Y.; Qi, X.; Jin, L.; Jin, J.; Ma, X.; Hua, D. Inhibition of transient receptor potential channel 5 reverses 5-fluorouracil resistance in human colorectal cancer cells. J. Biol. Chem. 2015, 290, 448–456. [Google Scholar] [CrossRef] [Green Version]
  127. Akbar, A.; Yiangou, Y.; Facer, P.; Brydon, W.G.; Walters, J.R.F.; Anand, P.; Ghosh, S. Expression of the TRPV1 receptor differs in quiescent inflammatory bowel disease with or without abdominal pain. Gut 2010, 59, 767–774. [Google Scholar] [CrossRef]
  128. Devesa, I.; Planells-Cases, R.; Fernández-Ballester, G.; González-Ros, J.M.; Ferrer-Montiel, A.; Fernández-Carvajal, A. Role of the transient receptor potential vanilloid 1 in inflammation and sepsis. J. Inflamm. Res. 2011, 4, 67–81. [Google Scholar]
  129. Lee, J.; Yamamoto, T.; Kuramoto, H.; Kadowaki, M. TRPV1 expressing extrinsic primary sensory neurons play a protective role in mouse oxazolone-induced colitis. Auton. Neurosci. Basic Clin. 2012, 166, 72–76. [Google Scholar] [CrossRef]
  130. Massa, F.; Sibaev, A.; Marsicano, G.; Blaudzun, H.; Storr, M.; Lutz, B. Vanilloid receptor (TRPV1)-deficient mice show increased susceptibility to dinitrobenzene sulfonic acid induced colitis. J. Mol. Med. 2006, 84, 142–146. [Google Scholar] [CrossRef]
  131. Martelli, L.; Ragazzi, E.; Di Mario, F.; Martelli, M.; Castagliuolo, I.; Dal Maschio, M.; Palù, G.; Maschietto, M.; Scorzeto, M.; Vassanelli, S.; et al. A potential role for the vanilloid receptor TRPV1 in the therapeutic effect of curcumin in dinitrobenzene sulphonic acid-induced colitis in mice. Neurogastroenterol. Motil. 2007, 19, 668–674. [Google Scholar] [CrossRef]
  132. Vinuesa, A.G.; Sancho, R.; García-Limones, C.; Behrens, A.; Ten Dijke, P.; Calzado, M.A.; Muñoz, E. Vanilloid receptor-1 regulates neurogenic inflammation in colon and protects mice from colon cancer. Cancer Res. 2012, 72, 1705–1716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. De Jong, P.R.; Takahashi, N.; Harris, A.R.; Lee, J.J.; Bertin, S.; Jeffries, J.; Jung, M.; Duong, J.; Triano, A.I.; Lee, J.J.; et al. Ion channel TRPV1-dependent activation of PTP1B suppresses EGFR-associated intestinal tumorigenesis. J. Clin. Invest. 2014, 124, 3793–3806. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Chen, Y.; Li, J.; Jin, L.; Lei, K.; Liu, H.; Yang, Y. Fibulin-5 contributes to colorectal cancer cell apoptosis via the ROS/MAPK and Akt signal pathways by downregulating transient receptor potential cation channel subfamily V member 1. J. Cell. Biochem. 2019, 120, 17838–17846. [Google Scholar] [CrossRef]
  135. Kim, C.S.; Park, W.H.; Park, J.Y.; Kang, J.H.; Kim, M.O.; Kawada, T.; Yoo, H.; Han, I.S.; Yu, R. Capsaicin, a spicy component of hot pepper, induces apoptosis by activation of the peroxisome proliferator-activated receptor γ in HT-29 human colon cancer cells. J. Med. Food 2004, 7, 267–273. [Google Scholar] [CrossRef] [PubMed]
  136. Borrelli, F.; Pagano, E.; Romano, B.; Panzera, S.; Maiello, F.; Coppola, D.; De Petrocellis, L.; Buono, L.; Orlando, P.; Izzo, A.A. Colon carcinogenesis is inhibited by the TRPM8 antagonist cannabigerol, a Cannabis-derived non-psychotropic cannabinoid. Carcinogenesis 2014, 35, 2787–2797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Launay, P.; Cheng, H.; Srivatsan, S.; Penner, R.; Fleig, A.; Kinet, J.P. TRPM4 regulates calcium oscillations after T cell activation. Science 2004, 306, 1374–1377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Launay, P.; Fleig, A.; Perraud, A.L.; Scharenberg, A.M.; Penner, R.; Kinet, J.P. TRPM4 is a Ca2+-activated nonselective cation channel mediating cell membrane depolarization. Cell 2002, 109, 397–407. [Google Scholar] [CrossRef] [Green Version]
  139. Fleig, A.; Penner, R. The TRPM ion channel subfamily: Molecular, biophysical and functional features. Trends Pharmacol. Sci. 2004, 25, 633–639. [Google Scholar] [CrossRef]
  140. Holzmann, C.; Kappel, S.; Kilch, T.; Jochum, M.M.; Urban, S.K.; Jung, V.; Stöckle, M.; Rother, K.; Peinelt, C. Transient receptor potential melastatin 4 channel contributes to migration of androgen-insensitive prostate cancer cells. Oncotarget 2015, 6, 41783–41793. [Google Scholar] [CrossRef]
  141. Pérez-Riesgo, E.; Gutiérrez, L.G.; Ubierna, D.; Acedo, A.; Moyer, M.P.; Núñez, L.; Villalobos, C. Transcriptomic analysis of calcium remodeling in colorectal cancer. Int. J. Mol. Sci. 2017, 18, 922. [Google Scholar] [CrossRef]
  142. Schlingmann, K.P.; Weber, S.; Peters, M.; Nejsum, L.N.; Vitzthum, H.; Klingel, K.; Kratz, M.; Haddad, E.; Ristoff, E.; Dinour, D.; et al. Hypomagnesemia with secondary hypocalcemia is caused by mutations in TPRM6, a new member of the TPRM gene family. Nat. Genet. 2002, 31, 166–170. [Google Scholar] [CrossRef] [PubMed]
  143. Schlingmann, K.P.; Waldegger, S.; Konrad, M.; Chubanov, V.; Gudermann, T. TRPM6 and TRPM7-Gatekeepers of human magnesium metabolism. Biochim. Biophys. Acta Mol. Basis Dis. 2007, 1772, 813–821. [Google Scholar] [CrossRef] [Green Version]
  144. Zou, Z.; Rios, F.J.; Montezano, A.C.; Touyz, R.M. TRPM7, Magnesium, and Signaling. Int. J. Mol. Sci. 2019, 20, 1877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Camprubí-Robles, M.; Planells-Cases, R.; Ferrer-Montiel, A. Differential contribution of SNARE-dependent exocytosis to inflammatory potentiation of TRPV1 in nociceptors. FASEB J. 2009, 23, 3722–3733. [Google Scholar] [CrossRef] [PubMed]
  146. Chen, G.C.; Pang, Z.; Liu, Q.F. Magnesium intake and risk of colorectal cancer: A meta-analysis of prospective studies. Eur. J. Clin. Nutr. 2012, 66, 1182–1186. [Google Scholar] [CrossRef]
  147. Nasulewicz, A.; Wietrzyk, J.; Wolf, F.I.; Dzimira, S.; Madej, J.; Maier, J.A.M.; Rayssiguier, Y.; Mazur, A.; Opolski, A. Magnesium deficiency inhibits primary tumor growth but favors metastasis in mice. Biochim. Biophys. Acta Mol. Basis Dis. 2004, 1739, 26–32. [Google Scholar] [CrossRef] [Green Version]
  148. Wolf, F.I.; Trapani, V.; Cittadini, A. Magnesium and the control of cell proliferation: Looking for a needle in a haystack. Magnes. Res. 2008, 21, 83–91. [Google Scholar]
  149. Trapani, V.; Arduini, D.; Cittadini, A.; Wolf, F.I. From magnesium to magnesium transporters in cancer: TRPM7, a novel signature in tumour development. Magnes. Res. 2014, 26, 149–155. [Google Scholar] [CrossRef] [Green Version]
  150. Wolf, F.I.; Maier, J.A.M.; Nasulewicz, A.; Feillet-Coudray, C.; Simonacci, M.; Mazur, A.; Cittadini, A. Magnesium and neoplasia: From carcinogenesis to tumor growth and progression or treatment. Arch. Biochem. Biophys. 2007, 458, 24–32. [Google Scholar] [CrossRef]
  151. Su, F.; Wang, B.-F.; Zhang, T.; Hou, X.-M.; Feng, M.-H. TRPM7 deficiency suppresses cell proliferation, migration, and invasion in human colorectal cancer via regulation of epithelial-mesenchymal transition. Cancer Biomark. 2019, 26, 451–460. [Google Scholar] [CrossRef] [PubMed]
  152. Dai, Q.; Shrubsole, M.J.; Ness, R.M.; Schlundt, D.; Cai, Q.; Smalley, W.E.; Li, M.; Shyr, Y.; Zheng, W.; John, R.; et al. The relation of magnesium and calcium intakes and a genetic polymorphism in the magnesium transporter to colorectal neoplasia risk. Am. J. Clin. Nutr. 2007, 86, 743–751. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Hermosura, M.C.; Nayakanti, H.; Dorovkov, M.V.; Calderon, F.R.; Ryazanov, A.G.; Haymer, D.S.; Garruto, R.M. A TRPM7 variant shows altered sensitivity to magnesium that may contribute to the pathogenesis of two Guamanian neurodegenerative disorders. Proc. Natl. Acad. Sci. USA 2005, 102, 11510–11515. [Google Scholar] [CrossRef] [Green Version]
  154. Castiglioni, S.; Cazzaniga, A.; Trapani, V.; Cappadone, C.; Farruggia, G.; Merolle, L.; Wolf, F.I.; Iotti, S.; Maier, J.A.M. Magnesium homeostasis in colon carcinoma LoVo cells sensitive or resistant to doxorubicin. Sci. Rep. 2015, 5, 1–9. [Google Scholar] [CrossRef] [PubMed]
  155. Reimers, M.S.; Zeestraten, E.C.M.; Kuppen, P.J.K.; Liefers, G.J.; van de Velde, C.J.H. Biomarkers in precision therapy in colorectal cancer. Gastroenterol. Rep. 2013, 1, 166–183. [Google Scholar] [CrossRef] [PubMed]
  156. Rogers, A.C.; Winter, D.C.; Heeney, A.; Gibbons, D.; Lugli, A.; Puppa, G.; Sheahan, K. Systematic review and meta-analysis of the impact of tumour budding in colorectal cancer. Br. J. Cancer 2016, 115, 831–840. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Zlobec, I.; Lugli, A. Tumour budding in colorectal cancer: Molecular rationale for clinical translation. Nat. Rev. Cancer 2018, 18, 203–204. [Google Scholar] [CrossRef]
  158. Hutchings, C.J.; Colussi, P.; Clark, T.G. Ion channels as therapeutic antibody targets. MAbs 2019, 11, 265–296. [Google Scholar] [CrossRef]
  159. Haustrate, A.; Hantute-Ghesquier, A.; Prevarskaya, N.; Lehen’kyi, V. Monoclonal antibodies targeting ion channels and their therapeutic potential. Front. Pharmacol. 2019, 10, 1–12. [Google Scholar] [CrossRef]
  160. Edwards, W.; Fung-Leung, W.P.; Huang, C.; Chi, E.; Wu, N.; Liu, Y.; Maher, M.P.; Bonesteel, R.; Connor, J.; Fellows, R.; et al. Targeting the ion channel Kv1.3 with scorpion venom peptides engineered for potency, selectivity, and half-life. J. Biol. Chem. 2014, 289, 22704–22714. [Google Scholar] [CrossRef] [Green Version]
  161. Wang, A.Z.; Langer, R.; Farokhzad, O.C. Nanoparticle Delivery of Cancer Drugs. Annu. Rev. Med. 2012, 63, 185–198. [Google Scholar] [CrossRef]
  162. Ataga, K.I.; Stocker, J. Senicapoc (ICA-17043): A potential therapy for the prevention and treatment of hemolysis-associated complications in sickle cell anemia. Expert Opin. Investig. Drugs 2009, 18, 231–239. [Google Scholar] [CrossRef] [PubMed]
  163. Arulkumaran, N.; Unwin, R.J.; Tam, F.W.K. A potential therapeutic role for P2X7 receptor (P2X7R) antagonists in the treatment of inflammatory diseases. Expert Opin. Investig. Drugs 2011, 20, 897–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Fu, S.; Hirte, H.; Welch, S.; Ilenchuk, T.T.; Lutes, T.; Rice, C.; Fields, N.; Nemet, A.; Dugourd, D.; Piha-Paul, S.; et al. First-in-human phase I study of SOR-C13, a TRPV6 calcium channel inhibitor, in patients with advanced solid tumors. Invest. New Drugs 2017, 35, 324–333. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Expression of TRP channels in gastrointestinal tract. Overview of TRP channels expressed in cancers of the gastrointestinal tract or in the cell lines originating from these types of cancers. ↑Arrows indicates that the channel is upregulated, while ↓ arrows indicates that the channel is downregulated. For some channels, there was more than one study, showing that they are either upregulated or downregulated, which is indicated by ↑↓.
Figure 1. Expression of TRP channels in gastrointestinal tract. Overview of TRP channels expressed in cancers of the gastrointestinal tract or in the cell lines originating from these types of cancers. ↑Arrows indicates that the channel is upregulated, while ↓ arrows indicates that the channel is downregulated. For some channels, there was more than one study, showing that they are either upregulated or downregulated, which is indicated by ↑↓.
Ijms 21 01877 g001
Figure 2. Overview of the role of TRP channels in cell functions of cancers of the gastrointestinal tract. TRP channels, which are permeable to monovalent and divalent cations, such as Na+, Ca2+ and Mg2+, are often dysregulated in cancer cells. These can lead to enhancement/suppression of proliferation, migration and invasion, cell cycle progression, and apoptosis. Channels with activity promoting a particular function are marked in green. Channels with activity suppressing a particular function are marked in red.
Figure 2. Overview of the role of TRP channels in cell functions of cancers of the gastrointestinal tract. TRP channels, which are permeable to monovalent and divalent cations, such as Na+, Ca2+ and Mg2+, are often dysregulated in cancer cells. These can lead to enhancement/suppression of proliferation, migration and invasion, cell cycle progression, and apoptosis. Channels with activity promoting a particular function are marked in green. Channels with activity suppressing a particular function are marked in red.
Ijms 21 01877 g002
Table 1. Expression of transient receptor potential (TRP) channels in human oral cancers.
Table 1. Expression of transient receptor potential (TRP) channels in human oral cancers.
Type of CancerChannelmRNA/Protein (+ Assessment Method)Sample SizeAim/Outcome + Reference
OralTRPV1–4mRNA (qPCR)37 oral SCC + tissues samples/compared to normal adjacent tissueTRPV1–4 mRNA and protein expression upregulated in oral SCC tissue samples in comparison to normal tissue [61]
protein (IHC)
TRPV1Protein
(IHC + WB)
18 tongue SCC +
8 leukoplakia +
7 normal tongue tissues samples
TRPV1 protein and mRNA expression upregulated in tongue SCC tissue samples in comparison to normal tissue [59] 1
mRNA (qPCR)
Protein (IHC)3 oral SCC + 3 normal oral mucosa tissue samplesTRPV1 protein is expressed in oral SCC [63]
TRPM2Protein (IHC)9 normal tongue +
12 papilloma tongue +
23 tongue carcinoma tissue samples
TRPM2 protein is overexpressed in tongue carcinoma in comparison to normal and papilloma samples [62]
1 Only IHC assessment of protein expression showed upregulation of TRPV1 in leukoplakia in comparison to normal tissue; data not available for qPCR and WB. IHC, immunohistochemistry; SCC, squamous cell carcinoma; qPCR, quantitative polymerase chain reaction; WB, Western-Blot.
Table 2. Expression of TRP channels in human esophageal squamous cell carcinoma (ESCC).
Table 2. Expression of TRP channels in human esophageal squamous cell carcinoma (ESCC).
Type of CancerChannelmRNA/Protein (+ Assessment Method)Sample SizeAim/Outcome + Reference
EsophagealTRPC6mRNA (in situ hybridization)55 paraffin-embedded ESSC + 21 fresh ESCC tissue samples/compared to normal adjacent tissueTRPC6 mRNA and protein are overexpressed compared to normal adjacent tissue [71]
Protein (IHC)
TRPM8mRNA (qPCR)10 ESCC tissue samples/compared to normal adjacent tissueTRPM8 mRNA and protein are overexpressed compared to normal adjacent tissue [72]
Protein (WB)
TRPM7Protein (IHC)52 ESCC tissue samples/compared to non-cancerous esophageal epitheliaTRPM7 protein is overexpressed compared to non-cancerous esophageal epithelia (no TRPM7 expression detected) [73] 1
TRPV2Protein (IHC)62 ESCC tissue samples Analysis of TRPV2 expression (low/high); worse overall survival and 5 year survival of patients with high TRPV2 protein expression [74]
1 The 5 year survival rate of patients with high TRPM7 expression (82.6%) was significantly higher than that of the patients with low expression (54.6%). ESCC, esophageal squamous cell carcinoma; IHC, immunohistochemistry; qPCR, quantitative polymerase chain reaction; WB, Western-Blot.
Table 3. Expression of TRP channels in human liver cancer samples.
Table 3. Expression of TRP channels in human liver cancer samples.
Type of CancerChannelmRNA/Protein (+ Assessment Method)Sample SizeAim/Outcome + Reference
LiverTRPV1mRNA
(RT-PCR)
6 pairs of HCC tissue samples/compared to normal adjacent tissue6 non-tumor tissues showed TRPV1 mRNA overexpression; HCC tissue samples showed downregulation in 4/6 tested [77]
mRNA (in situ hybridization)15 HCC samples/compared to normal adjacent tissueTRPV1 expressed in some HCC and normal tissue samples; data non-conclusive [77]
Protein (IHC)62 HCC tissue samples + 62 non-tumor control tissuesHigh TRPV1 expression was observed in 30/62 HCC samples; high TRPV1 expression was associated with longer disease-free survival [77]
TRPV2Protein (IHC) 55 HCC cancer tissue samplesUpregulation of TRPV2 on mRNA and protein levels inversely correlated with histopathologic differentiation [78]
mRNA (RT-PCR)13 paired HCC tumor mRNA extracts
TRPV4Protein (IHC)45 HCC tissue samples/compared to normal adjacent tissueTRPV4 protein and mRNA levels higher in HCC tissues than in normal tissues; positive correlation between TRPV4 expression; the histological grade and number of tumors [79]
mRNA (qPCR)
TRPC6Protein (IHC)150 HCC tissue samples/compared to normal tissuesTRPC protein upregulated in HCC tissues in comparison to normal tissues [80]
IHC, immunohistochemistry; qPCR, quantitative polymerase chain reaction; RT-PCR, reverse transcription polymerase chain reaction; WB, Western-Blot.
Table 4. Expression of TRP channels in human pancreatic cancer samples.
Table 4. Expression of TRP channels in human pancreatic cancer samples.
Type of CancerChannelmRNA/Protein (+ Assessment Method)Sample SizeAim/Outcome + Reference
PancreaticTRPM2mRNA
(analysis of previously published cancer genome studies)
91 pancreatic cancer patients High TRPM2 expression correlated with lower overall survival [92]
TRPM8Protein (IHC)280 pancreatic adenocarcinoma tissue microarrays Moderate or high level of TRPM8 protein expression in 92% of pancreatic adenocarcinoma; the expression levels of TRPM8 positively correlate with the size of the primary tumor and tumor stages [94]
Protein (IHC)5 pancreatic adenocarcinoma tissue samples/compared to normal adjacent tissueTRPM8 protein expression upregulated compared to normal tissue [95]
Protein (IHC)
mRNA (qPCR)
44 pancreatic adenocarcinoma tissue samples/compared to normal adjacent tissueTRPM8 protein and mRNA upregulated compared to normal tissue [96]
mRNA (qPCR)110 pancreatic adenocarcinoma tissue samples/compared to normal adjacent tissueTRPM8 mRNA upregulated compared to normal tissue; high TRPM8 protein expression was found to be associated with lower overall survival and poor disease free survival values for pancreatic cancer patients [97]
TRPM7Protein (IHC)5 pancreatic adenocarcinoma tissue samples/compared to normal pancreatic tissue samplesTRPM7 protein upregulated compared to normal tissue [100]
Protein (IHC)282 pancreatic adenocarcinoma tissue microarrays/compared to normal pancreatic tissue microarraysTRPM7 protein upregulated compared to normal tissue; TRPM7 expression correlates with the tumor stage [103]
Protein (IHC)
mRNA (RT-PCR)
8 tumor pancreatic ductal adenocarcinoma/compared to 6 normal pancreatic tissues TRPM7 protein and mRNA upregulated compared to normal pancreatic tissue [102]
TRPV6Protein (IHC)76 tumor pancreatic tissue samples compared to adjacent normal pancreatic tissues TRPV6 protein upregulated compared to normal pancreatic tissue [109]
IHC, immunohistochemistry; qPCR, quantitative polymerase chain reaction; RT-PCR, reverse transcription polymerase chain reaction.
Table 5. Expression of TRP channels in human gastric cancer samples.
Table 5. Expression of TRP channels in human gastric cancer samples.
Type of CancerChannelmRNA/Protein (+ Assessment method)Sample SizeAim/Outcome + Reference
GastricTRPC6Protein (IHC)25 primary gastric cancer samples/compared to 4 gastritis samplesTRPC6 mRNA and protein expression upregulated compared to gastritis samples [111]
mRNA (in situ hybridization)10 primary gastric cancer samples
TRPM2mRNA (analysis of online gastric cancer databases)896 gastric cancer patients; analysis of low TRPM2 vs high TRPM2 expressionHigh TRPM2 mRNA expression high expression negatively associated with the overall survival of patients [115]
IHC, immunohistochemistry.
Table 6. Expression of TRP channels in human colorectal cancer (CRC) samples.
Table 6. Expression of TRP channels in human colorectal cancer (CRC) samples.
Type of CancerChannelmRNA/Protein (+ Assessment Method)Sample SizeAim/Outcome + Reference
Colorectal (CRC)TRPC1mRNA (analysis of CRC datasets, available from public databases)656 CRC samples including 47 normal samplesHigh TRPC1 expression correlated with poor prognosis for the patients [120]
585 CRC samples including 19 normal samples
TRPV1Protein (IHC)10 CRC tissue samples, 10 CRC-adjacent tissue samples, and 6 normal subjectsTRPV1 protein expression decreased in CRC tissues compared to normal tissues [121]
TRPM4Protein (IHC)CRC tumor tissue microarrays from 379 patientsHigh TRPM4 protein expression was associated with unfavorable tumor features characteristic for epithelial-mesenchymal transition and infiltrative growth patterns [122]
TRPM6mRNA (analysis of CRC datasets, available from public databases)656 CRC samples including 47 normal samplesTRPM6 mRNA expression decreased compared to normal tissue [120]
585 CRC samples including 19 normal samples
mRNA (analysis of CRC dataset, available from public databases)585 CRC samples including 19 normal samplesTRPM6 mRNA expression decreased compared to normal tissue; high TRPM6 mRNA expression positively correlated with overall survival [123].
CRC, colorectal cancer; IHC, immunohistochemistry.

Share and Cite

MDPI and ACS Style

Stokłosa, P.; Borgström, A.; Kappel, S.; Peinelt, C. TRP Channels in Digestive Tract Cancers. Int. J. Mol. Sci. 2020, 21, 1877. https://doi.org/10.3390/ijms21051877

AMA Style

Stokłosa P, Borgström A, Kappel S, Peinelt C. TRP Channels in Digestive Tract Cancers. International Journal of Molecular Sciences. 2020; 21(5):1877. https://doi.org/10.3390/ijms21051877

Chicago/Turabian Style

Stokłosa, Paulina, Anna Borgström, Sven Kappel, and Christine Peinelt. 2020. "TRP Channels in Digestive Tract Cancers" International Journal of Molecular Sciences 21, no. 5: 1877. https://doi.org/10.3390/ijms21051877

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

Article Metrics

Back to TopTop