Next Article in Journal
Neuroprotective Role of Dietary Supplementation with Omega-3 Fatty Acids in the Presence of Basal Forebrain Cholinergic Neurons Degeneration in Aged Mice
Next Article in Special Issue
Long-Tailed Unconventional Class I Myosins in Health and Disease
Previous Article in Journal
Correction: Cherkaoui, M., Lollier, V., Geairon, A., Bouder, A., Larré, C., Rogniaux, H., Jamet, E., Guillon, F. and Francin-Allami, M. Cell Wall Proteome of Wheat Grain Endosperm and Outer Layers at Two Key Stages of Early Development. Int. J. Mol. Sci. 2020, 21, 239
Previous Article in Special Issue
Junctional Adhesion Molecule-C Mediates the Recruitment of Embryonic-Endothelial Progenitor Cells to the Perivascular Niche during Tumor Angiogenesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Transient Receptor Potential Canonical (TRPC) Channels as Modulators of Migration and Invasion

by
Muhammad Yasir Asghar
1,2 and
Kid Törnquist
1,2,*
1
Minerva Foundation Institute for Medical Research, Biomedicum Helsinki 2U, Tukholmankatu 8, 00290 Helsinki, Finland
2
Faculty of Science and Engineering, Cell Biology, Åbo Akademi University, Tykistökatu 6A, 20520 Turku, Finland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2020, 21(5), 1739; https://doi.org/10.3390/ijms21051739
Submission received: 11 February 2020 / Revised: 25 February 2020 / Accepted: 26 February 2020 / Published: 3 March 2020
(This article belongs to the Special Issue Cell Adhesion and Migration in Health and Diseases)

Abstract

:
Calcium (Ca2+) is perhaps the most versatile signaling molecule in cells. Ca2+ regulates a large number of key events in cells, ranging from gene transcription, motility, and contraction, to energy production and channel gating. To accomplish all these different functions, a multitude of channels, pumps, and transporters are necessary. A group of channels participating in these processes is the transient receptor potential (TRP) family of cation channels. These channels are divided into 29 subfamilies, and are differentially expressed in man, rodents, worms, and flies. One of these subfamilies is the transient receptor potential canonical (TRPC) family of channels. This ion channel family comprises of seven isoforms, labeled TRPC1–7. In man, six functional forms are expressed (TRPC1, TRPC3–7), whereas TRPC2 is a pseudogene; thus, not functionally expressed. In this review, we will describe the importance of the TRPC channels and their interacting molecular partners in the etiology of cancer, particularly in regard to regulating migration and invasion.

1. Introduction

Increasing evidence during the past decade indicates that different ion channels are expressed in several cancers in humans, and regulate a multitude of cellular processes, including migration, invasion and proliferation [1,2,3]. Hence, ion channels have emerged as potential targets for cancer therapy. Of the channels considered important for cancer progression, special interest has been given to Ca2+ channels, as Ca2+ signaling is of crucial importance in the regulation of many cellular processes. Different Ca2+ channels gate Ca2+ ions into the cell, where after Ca2+ bind to Ca2+ binding proteins and activate downstream signaling pathways, resulting in specific cellular responses. These include motility and contraction, energy production, and gene transcription. To make this versatility possible, several different, pumps, transporters, and channels are necessary [4,5,6].
In many cancer forms, Ca2+ signaling plays an essential role in cell migration, invasion, and proliferation through transient receptor potential canonical (TRPC) channels, a subfamily of the transient receptor potential (TRP) superfamily of Ca2+ ion channels. The TRPC subfamily comprises six members in humans, TRPC1 and TRPC3–7. Many of these channels are ubiquitously expressed in human tissues and modulate a multitude of cellular responses [7,8,9,10,11,12]. Some of these channels function both as receptor operated Ca2+ entry (ROCE) and as store operated Ca2+ entry (SOCE) channels [1,13,14]. In the present review, we want to highlight the importance of different TRPC channels in the migration and invasion of cancer cells. For a recent, comprehensive overview of the importance of different ion channels, and not exclusively TRPC channels, see [2,3].

2. Calcium Signaling

Ca2+ is a ubiquitous intracellular second messenger and activates numerous cellular processes, including fluid secretion, muscle contraction, exocytosis, gene transcription, fertilization, cell differentiation, proliferation and migration [4,15,16].
Ca2+ entry into the cells can be evoked by depolarization of the plasma membrane and the activation of voltage-operated calcium channels (VOCs). These channels are mainly found in excitable cells, such as neurons and muscle cells. An increase in calcium entry can also be obtained in response to an agonist stimulus that, through G protein coupled receptors (GPCR), activates membrane-associated phospholipase C (PLC), and generates two important second messengers, inositol 1,4,5-triphosphate (IP3) and diacylglycerol (DAG). DAG has been shown to activate receptor-operated calcium entry (ROCE) by activation of TRPC channels, which enables Ca2+ influx into the cells. The IP3 diffuses into the cytoplasm and activates IP3 receptors on the endoplasmic reticulum (ER), resulting in the release of Ca2+ from the ER stores [4,17,18,19,20]. The extracellular Ca2+ concentration, which usually is 2–3 mM, is huge as compared to the intracellular free Ca2+ concentration at resting state. The resting level of free Ca2+ in the cytoplasm is maintained at low concentration, about 100 nM. The steep concentration gradient facilitates rapid Ca2+ entry into the cells. Cells maintain the low level of cytosolic Ca2+ by strict regulation of Ca2+ channels, pumps, and Ca2+ binding proteins (Figure 1). The Ca2+ signals in cells are normally in the form of rapid transients in the free cytosolic Ca2+ concentrations. The abnormality or impairment of the cytosolic Ca2+ transients may induce severe diseases [6,20,21].
The increase in the cytoplasmic Ca2+ levels enables Ca2+ binding proteins such as EF-hand proteins, annexins, and C2-domain proteins to become active and to regulate several cellular processes, such as muscle contraction, exocytosis, metabolism, gene transcription, fertilization, proliferation, and hypertrophy. To avoid too much of Ca2+ in the cytoplasm, the levels of Ca2+ in the cytoplasm are strictly regulated. However, the Ca2+ signals may have a duration ranging from microseconds to hours, depending upon the cell type and the specific function of the signal. Persistently elevated Ca2+ levels in the cytoplasm can lead to apoptosis and cell death. Moreover, irregular, e.g., high, or low amplitude Ca2+ signals have been associated with diseases [6,21].

3. Transient Receptor Potential (TRP) Channels

In the fly, Drosophila melanogaster, a mutant that caused temporary blindness in bright light stimulus was detected, and later it was found that this mutation encodes a cation transient receptor potential (TRP) channel [22,23,24]. The TRP channel-superfamily is the largest family of cation channels and is comprised of 17 different channels in worms, 13 in flies, 28 in mice, and 27 in humans, as shown in Table 1.
The expression and function of TRP channels is diverse throughout the animal kingdom, including worms, flies, mice, and humans. There are seven subfamilies of TRP channels based on amino acid sequence homology, and include TRPC (canonical), TRPV (Vanilloid), TRPML (Mucolipin), TRPP (Polycystin), TRPM (Melastatin), TRPA (Ankyrin), and TRPN (no mechanoreceptor potential C), as shown in Figure 2.
TRP channels are selective to cations and some subtypes are highly selective for Ca2+ and Mg2+. The channels are ubiquitously expressed in human tissues. Many TRP channels have been indicated to participate in a multitude of physiological processes including perception of heat, touch, pain, odor and smell, cellular regulation of osmolarity, fluid secretion, inflammation, cell adhesion, proliferation, cell differentiation, migration, and apoptosis [15,25,26,27,28,29,30,31,32,33,34]. In addition to the physiological importance, TRP channels have been found to regulate many diseases, including cancer [35,36,37,38,39].

3.1. TRPC Channels

The transient receptor potential canonical (TRPC) subfamily consists of seven members (TRPC1-7). In humans, all isoforms are expressed except TRPC2, which is a pseudogene. TRPCs are non-selective cation channels, with a preference for Ca2+ over Na+ and K+ ions. The structure of the TRPC channels consists of six transmembrane segments (S1–S6) connected through loops, and the amino (N) and the carboxyl (C) terminals are located in the cytoplasm. The length and amino acid sequence of the C or N terminals varies. The loop region between segments 5 and 6 makes the pore, which conducts ions as shown in Figure 3. The N-terminus has a coiled-coil domain and four ankyrin domains, which facilitates protein interactions and has been shown to be involved in the regulation of TRPC channel function by tetramerization of TRPC subunits [40]. The C-terminus has a TRP domain, which is the site for other TRP channel isoforms to bind and form channel complexes, a coiled-coil domain, and a calmodulin and IP3R binding site, which regulates the activation and inhibition of the channel [41]. TRPCs regulate several Ca2+ dependent cellular processes. TRPC1 can co-assemble with all other TRPC isoforms (TRPC3-7). The TRPC1/TRPC3 complex enhances SOCE, and regulates the differentiation of e.g., H19-7 hippocampal neuronal cells [42]. TRPC channels participate in both ROCE as well as SOCE together with Orai1 and STIM1 proteins.

3.1.1. TRPC1

TRPC1 channels are ubiquitously expressed in human tissues and participate in carrying out many cellular processes, including cell migration and proliferation [11,13,14,43]. TRPC1 is the first mammalian TRPC channel identified and cloned [44]. TRPC1 has been established as a potent molecular component of SOCE in several distinct cell types [19,45,46,47]. In addition, TRPC1 has been shown to function as a ROCE channel in several cell lines [48,49]. TRPC1 can make complexes with all other TRPC channels. In HSY cells, TRPC1 forms a complex with TRPC3. In human embryonic kidney (HEK-293) cells, TRPC1 forms a heterotrimeric complex with both TRPC3 and TRPC7. In mesangial cells, TRPC1 forms a complex with TRPC4 and in neuronal cells TRPC1 has been shown in complex with TRPC5 [11,50,51,52,53]. TRPC1 channels also interact with the Ca2+ signaling proteins Gq/11, PLC, calmodulin, IP3 receptors (IP3R), PMCA, SERCA, and STIM1 to regulate cellular processes. STIM1 and the TRPC1-TRPC4 complex are essential for store refilling and differentiation in myotubes [54].

3.1.2. TRPC2

TRPC2 is a pseudogene and does not form a functional channel in humans [44,55]. Relatively little is known about this channel’s physiology and pathology. However, in other mammals, TRPC2 forms a functional channel in distinct tissues, such as the vomeronasal organ (VNO), testis, spleen and liver [55,56]. The importance of TRPC2 in the vomeronasal organ has obtained much interest in regard to pheromone detection, see reviews [57,58]. TRPC2 has been shown to interact with Homer 1, calmodulin, Gq/11, IP3R, and receptor transporting protein 1 (RTP1) [59,60,61]. We detected TRPC2 in the rat thyroid FRTL-5 cell line. Knockdown of TRPC2, or expression of a dominant negative form of TRPC2 in these cells resulted in a significant decrease in proliferation, migration, adhesion and invasion [62]. In addition, TRPC2 has the ability to mediate both ROCE and SOCE in FRTL-5 cells [63].

3.1.3. TRPC3

TRPC3 channels are expressed in the human brain, kidney, skeletal muscle, mammary, ovary and cardiovascular tissues [14]. TRPC3 interacts and forms heterodimers with TRPC1, 4, 5, 6, and 7. Despite a considerable homology of the amino acid sequence between TRPC3, TRPC6, and TRPC7, these channels regulate many different functions in humans [64,65]. TRPC3 also interacts with the Orai1 calcium channel [66,67]. TRPC3, either in complex with TRPC1, or in complex with both TRPC1 and TRPC7, forms a SOCE channel. TRPC3 interacts with several signaling proteins in response to receptor-evoked Ca2+ mobilization, including PLCβ, Gq/11, and IP3R. TRPC3 has been reported to be directly activated by DAG [68].

3.1.4. TRPC4

TRPC4 shares a 65% homology in the amino acid sequence with TRPC5 [19,45]. Mostly, TRPC4 channels form SOCE channels, but in some cell types, they form store-independent channels as well. It can make heteromeric channels with TRPC6. STIM1 binds with TRPC4 and regulates its function as a SOCE channel [69]. Furthermore, TRPC4 has been studied extensively in regard to endothelial cell function and has been shown to regulate proliferation of these cells [14,70].

3.1.5. TRPC5

TRPC5 is highly expressed in neuronal cells and regulates their function. In HEK-293 cells, TRPC5 is a non-selective channel and functions as a ROCE channel but not as a SOCE channel [71]. Recently, it has been shown that TRPC5 is activated in response to mechanical stress, and this increase in the activity of the channel is dependent on actin filaments [72]. TRPC5-TRPC1 complexes are found in neurons, vascular endothelial cells, and smooth muscle cells. In vascular smooth muscle (VSM) cells, TRPC5 has been suggested to act as SOCE channel by forming complexes with TRPC1, TRPC6, and TRPC7 [73].

3.1.6. TRPC6

TRPC6 is expressed in pulmonary and vascular smooth muscle cells. These channels are directly activated by DAG and do regulate ROCE [74]. However, activation of TRPC6 due to Ca2+ release from the ER is mediated by Orai1 and TRPC4 [75]. TRPC6 has been shown to make a complex with TRPC3 in neuronal cells and prostate cancer epithelial cells. In astrocytes, sphingosine 1-phosphate (S1P)-evoked secretion of the C-X-C motif chemokine ligand 1 (CXCL1) is mediated through activation of TRPC6 and the mitogen-activated protein kinase (MAPK) signaling pathway [76]. Furthermore, TRPC6 is important in transforming growth factor β1 (TGFβ1) signaling in vascular smooth muscles. TGFβ1 induces stress fiber formation in these cells via upregulation of TRPC6 [77]. The interplay of leukocytes with platelet/endothelial cell adhesion molecule-1 (PECAM) result in activation of TRPC6 which modulates subsequent leukocyte transendothelial migration (TEM) [78].

3.1.7. TRPC7

In humans, TRPC7 is widely expressed in many tissues including the brain, skin, cartilage, pituitary gland, intestine, kidney, and prostate [14]. TRPC7, a non-selective cation channel, is the seventh identified member of mammalian TRPC family. It was isolated through molecular cloning of the mouse fetal brain and caudate nucleus cDNA libraries [79,80]. The role of TRPC7 in the regulation of normal cell physiology and pathology is still indefinable. TRPC7 is activated by Gq-coupled protein receptors and the PLC pathway and is directly activated by DAG. In some cell types, TRPC7s are constitutively active proteins and may function as SOCE channels by forming a TRPC1-TRPC3-TRPC7 complex. However, in HEK-293 cells, a TRPC3 and TRPC7 complex is activated by DAG and functions as ROCE channels [81]. Furthermore, cGMP-dependent protein kinase 1α, calmodulin, IP3R, and phosphatidylinositol 4,5,-bisphosphate (PIP2) have all been reported to regulate the function of TRPC7 [82]. In addition, activation of TRPC7 potently induces myocardial apoptosis [83].

4. TRPC Channels as Regulators of Migration and Invasion

4.1. TRPC Channels, Migration and Invasion

Several TRPC channels are involved in regulating migration and invasion, and all are involved in enhancing proliferation of cancer cells [84,85,86]. TRPC1 channels are essential for the polarity and direction of migrating cells both in vitro and in vivo [49,51,52]. The directionality of migrating cells, as shown using renal transformed epithelial cells, was dependent on TRPC1 [87]. In aggressive glioma cells, TRPC1 regulates epidermal growth factor (EGF)-evoked migration [88]. In addition, TRPC1 has been shown to localize to lipid rafts at the leading edge of migrating cells. TRPC1 mediated Ca2+ entry has been shown to activate the MAPK and phosphoinositide 3-kinase/Akt (PI3K/Akt) signaling pathways, calpains, hypoxia-induced factor 1α (HIF1α) and several matrix-metalloproteinases (MMPs). Thus, TRPC1 has emerged as an important player involved both in normal and cancer cell function [49]. Furthermore, TRPC1 may be in complex with the small conductance Ca2+ activated potassium channel (SK3) and Orai1 to enhance SOCE-dependent colon cancer migration [89]. In skeletal myoblast migration, TRPC1 is of importance to evoke Ca2+ -mediated activation of calpains [11]. Silencing TRPC1 by siRNA inhibited invasion of CNE2 nasopharyngeal tumor cells [90].
We have shown that TRPC1 is a potent regulator of both migration and invasion in follicular thyroid cancer ML-1 cells [91]. The mechanisms involved appear to depend on the at least MAPK/ERK1/2, MMP2 and-9, and HIF1α. However, so far, no studies regarding TRPC1 in thyroid cancer have been performed. The importance of TRPC channels in normal thyroid cells has, to the best of our knowledge, not been investigated.
The importance of TRPC2 as a regulator of invasion and migration has not been extensively studied. In thyroid cells, Ca2+ has many important functions [63,92], and we have shown that, in normal rat thyroid FRTL-5 cells, TRPC2 potently regulates migration, invasion, and adhesion [62]. Rac, calpain and MMP2 seem to be of importance in TRPC2-regulated migration. Considering that thyroid cells migrate during embryogenesis [93], our data suggest that TRPC2 might have a role in this event in rodents. Although TRPC2 knock-out mice appear normal (except for their behavioral changes, see e.g., [94]), it would also be of interest to investigate the thyroid status in these mice.
The importance of TRPC3 as a regulator of migration and invasion has been studied in several different tumor cells. In human ovarian cancer, the expression of TRPC3 was increased, and injection of TRPC3-knock down SKOV3 cells decreased tumor formation in nude mice [95]. In pancreatic stellar cells, activation of KCa3.1 results in activation of TRPC3-evoked Ca2+ entry, calpain activity, and concomitant cell migration [96]. TRPC3 also enhanced melanoma cell migration and tumor formation both in vitro and in vivo [97]. Furthermore, TRPC3, together with mGluR5, has been shown to enhance motility in embryonic neuronal cells as a result of endocannabinoid stimulation [98]. It has also been shown that the invasion of bladder cancer cells is dependent on both TRPC3 and TRPC6. Interestingly, in these cells the expressional level of these channels is under the control of the histone variant macroH2A [99].
Not much is known regarding the importance of TRPC4 or TRPC5 in enhancing migration and invasion of cancer cells, although at least TRPC4 seems to have a role in proliferation and tumor formation [100]. Furthermore, TRPC4 seems to enhance invasion of certain forms of medulloblastoma cells. In these cells, the expression of TRPC4 appears to be regulated by the proton sensing G-protein coupled receptor 1 (OGR1) [101]. Several TRPC isoforms, including TRPC4, have also been suggested to enhance proliferation of non-small cell lung cancer [102]. The importance of TRPC5 has been shown in regulating hippocampal neurite length and growth cone morphology, by having an inhibiting role in neurite extension [103]. Furthermore, in fibroblasts and kidney podocytes, TRPC5 is in complex with Rac1, enhancing cell motility, whereas TRPC6-evoked Ca2+ entry enhances RhoA activity; thus, inhibiting cell motility [104]. Interestingly, the plant derived compound Englerin A induces cytotoxicity in some cancer cell lines expressing TRPC4 and TRPC5 by enhancing channel activity; thus, increasing cytosolic Ca2+ and so Na+ concentrations [105,106].
In addition to TRPC1 and TRPC3, the involvement of TRPC6 in cancer cell invasion and migration has been extensively investigated [3]. In human head and neck squamous cell carcinomas (HNSCC), TRPC6 was overexpressed, and knock-down of TRPC6 in HNSCC potently attenuated invasion [107]. TRPC6 was also shown to be of significant importance in regulating Notch-driven glioblastoma growth and invasiveness [108]. In human prostate cancer epithelial (hPCE) cells, active TRPC6 and nuclear factor of activate T cells (NFAT) promote proliferation via alpha 1-adreniergic receptor signaling [109]. In non-small cell lung cancer A549, TRPC6 was important for invasion [110]. Interestingly, knock-down of TRPC6 not only attenuated invasion, but also decreased the expression of the adhesion protein fibronectin and the tight junction protein ZO-1 [110]. Furthermore, TRPC6 was of importance in regulating breast cancer MDA-MB-231 cell migration [111]. TRPC6 expression has been found upregulated in breast cancer cell lines, MCF-7 and MDA-MB-231, compared with normal breast epithelial MCF10A cells. In vitro knockdown of TRPC6 decreased proliferation and migration in these cancer cells. In addition, TRPC6 has been shown to interact and translocate Orai1 and Orai3 to the plasma membrane upon ER Ca2+ store depletion [112].
An overview of TRPC channels expressed in tissues and their respective reported function regarding invasion and migration has been summarized in Table 2.

4.2. TRPC Channels and Angiogenesis

The regulation of angiogenesis is a complex event where calcium signaling plays a significant role. Endothelial cells express several members of the TRP superfamily, many of which have been shown to be involved in angiogenesis. Knockdown or inhibition of TRPC3 concealed the endothelial tube formation by attenuation of VEGF-evoked Ca2+ entry and phosphorylation of MAPK [113], and inhibition of TRPC4 suppressed the VEGF evoked neovascularization in human retina microvascular endothelial cells (HRMECs) [114]. TRPC3, 4 and 5 have been shown to participate in in vitro tube formation of the human umbilical vein endothelial cell-derived cell line EA.hy926 [115], and in mice with ischemic injury, the knockdown of TRPC5 in endothelial cells (ECs) under-hypoxia inhibited sprouting and tube formation. However, activation of TRPC5 in these cells restored angiogenesis [116]. TRPC6 has been shown to participate in VEGF evoked sprouting and angiogenesis in human microvascular endothelial cells (HMVEC) [117]. Furthermore, phosphatase and tensin homolog (PTEN) regulates TRPC6 activation and subsequent promotion of angiogenesis in primary human pulmonary artery endothelial cells (HPAEs) [118]. As several comprehensive reviews regarding TRP channels and angiogenesis have recently been published, we recommend the readers to turn to these for further information [119,120,121].

5. Conclusions

It seems clear that TRPC channels (and apparently, also, other members of the TRP superfamily), can have a profound role in the migration and invasion of cancer cells, and thus in the process of metastasis. Furthermore, as pointed out by a recent review from the Prevarskaya lab, in many of the common cancer hallmarks, i.e., uncontrolled proliferation, resistance to programmed cell death, invasion and metastasis, and angiogenesis, ion channels (including TRP channels) have a clear role [3]. Thus, it seems obvious that different ion channels are considered putative targets for pharmacological interventions. However, as the TRPC channels often are ubiquitously expressed, it will be a challenge to find suitable pharmacological tools to block these channels and curtail cancer without severe side effects. One possibility could be functionalized nanoparticles that specifically target cancer cells, see e.g., [122,123]. This area of research is proceeding rapidly, and novel approaches are constantly developed. It will be interesting to see if these will be used to e.g., carry siRNA to target ion channels in cancer cells, and whether this approach can be used in patients.

Author Contributions

M.Y.A. and K.T. wrote the text; figures were made by M.Y.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

There exists an extensive literature regarding the importance of calcium and TRPC channels in health and disease. We apologize to all our colleagues whose work we have not been able to cite. The studies emanating from the Törnquist lab were generously supported by the Academy of Finland, the Sigrid Juselius Foundation, the Minerva Foundation, the Centre of Excellence in cell Stress and Molecular Ageing (Åbo Akademi University), the Magnus Ehrnrooth Foundation, the Liv och Hälsa Foundation, K Albin Johansson Foundation and the Svenska Kulturfonden, which is gratefully acknowledged.

Conflicts of Interest

The Authors declare no conflict of interest.

References

  1. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion channels and the hallmarks of cancer. Trends Mol. Med. 2010, 16, 107–121. [Google Scholar] [CrossRef] [PubMed]
  2. Monteith, G.R.; Prevarskaya, N.; Roberts-Thomson, S.J. The calcium-cancer signalling nexus. Nat. Rev. Cancer 2017, 17, 373–380. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Prevarskaya, N.; Skryma, R.; Shuba, Y. Ion Channels in Cancer: Are Cancer Hallmarks Oncochannelopathies? Physiol. Rev. 2018, 98, 559–621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Berridge, M.J.; Lipp, P.; Bootman, M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000, 1, 11–21. [Google Scholar] [CrossRef] [PubMed]
  5. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium signalling: Dynamics, homeostasis and remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [Green Version]
  6. Brini, M.; Carafoli, E. Calcium pumps in health and disease. Physiol. Rev. 2009, 89, 1341–1378. [Google Scholar] [CrossRef] [Green Version]
  7. Goodman, M.B.; Schwarz, E.M. Transducing Touch in Caenorhabditis elegans. Annu. Rev. Physiol. 2003, 65, 429–452. [Google Scholar] [CrossRef] [Green Version]
  8. Liu, X.; Cheng, K.T.; Bandyopadhyay, B.C.; Pani, B.; Dietrich, A.; Paria, B.C.; Swaim, W.D.; Beech, D.; Yildrim, E.; Singh, B.B.; et al. Attenuation of store-operated Ca2+ current impairs salivary gland fluid secretion in TRPC1(-/-) mice. Proc. Natl. Acad. Sci. USA 2007, 104, 17542–17547. [Google Scholar] [CrossRef] [Green Version]
  9. Montell, C. The TRP superfamily of cation channels: Science’s STKE: Signal transduction knowledge environment. Science 2005, 2005, re3. [Google Scholar] [CrossRef] [Green Version]
  10. Pedersen, S.F.; Owsianik, G.; Nilius, B. TRP channels: An overview. Cell Calcium 2005, 38, 233–252. [Google Scholar] [CrossRef]
  11. Louis, M.; Zanou, N.; Van Schoor, M.; Gailly, P. TRPC1 regulates skeletal myoblast migration and differentiation. J. Cell Sci. 2008, 121, 3951–3959. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Nilius, B.; Owsianik, G. The transient receptor potential family of ion channels. Genome Biol. 2011, 12, 218. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Riccio, A.; Medhurst, A.D.; Mattei, C.; Kelsell, R.E.; Calver, A.R.; Randall, A.D.; Benham, C.D.; Pangalos, M.N. mRNA distribution analysis of human TRPC family in CNS and peripheral tissues. Mol. Brain Res. 2002, 109, 95–104. [Google Scholar] [CrossRef]
  14. Abramowitz, J.; Birnbaumer, L. Physiology and pathophysiology of canonical transient receptor potential channels. FASEB J. 2009, 23, 297–328. [Google Scholar] [CrossRef] [Green Version]
  15. Clapham, D.E. Calcium Signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [Green Version]
  16. Pinto, M.C.X.; Kihara, A.H.; Goulart, V.A.M.; Tonelli, F.M.P.; Gomes, K.N.; Ulrich, H.; Resende, R.R. Calcium signaling and cell proliferation. Cell. Signal. 2015, 27, 2139–2149. [Google Scholar] [CrossRef]
  17. Venkatachalam, K.; Zheng, F.; Gill, D.L. Regulation of canonical transient receptor potential (TRPC) channel function by diacylglycerol and protein kinase C. J. Biol. Chem. 2003, 278, 29031–29040. [Google Scholar] [CrossRef] [Green Version]
  18. Montell, C.; Birnbaumer, L.; Flockerzi, V. The TRP channels, a remarkably functional family. Cell 2002, 108, 595–598. [Google Scholar] [CrossRef] [Green Version]
  19. Venkatachalam, K.; Montell, C. TRP channels. Annu. Rev. Biochem. 2007, 76, 387–417. [Google Scholar] [CrossRef] [Green Version]
  20. Berridge, M.J. The Inositol Trisphosphate/Calcium Signaling Pathway in Health and Disease. Physiol. Rev. 2016, 96, 1261–1296. [Google Scholar] [CrossRef] [Green Version]
  21. Stutzmann, G.E.; Mattson, M.P. Endoplasmic reticulum Ca(2+) handling in excitable cells in health and disease. Pharmacol. Rev. 2011, 63, 700–727. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Cosens, D.J.; Manning, A. Abnormal electroretinogram from a Drosophila mutant. Nature 1969, 224, 285–287. [Google Scholar] [CrossRef] [PubMed]
  23. Minke, B.; Wu, C.-F.; Pak, W.L. Induction of photoreceptor voltage noise in the dark in Drosophila mutant. Nature 1975, 258, 84–87. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. Clapham, D.E. TRP channels as cellular sensors. Nature 2003, 426, 517–524. [Google Scholar] [CrossRef]
  26. Julius, D. TRP channels and pain. Annu. Rev. Cell Dev. Biol. 2013, 29, 355–384. [Google Scholar] [CrossRef] [Green Version]
  27. Montell, C. Physiology, Phylogeny, and Functions of the TRP Superfamily of Cation Channels. Sci. Signal. 2001, 2001, re1. [Google Scholar] [CrossRef]
  28. Minke, B.; Cook, B. TRP channel proteins and signal transduction. Physiol. Rev. 2002, 82, 429–472. [Google Scholar] [CrossRef] [Green Version]
  29. Minke, B. The history of the Drosophila TRP channel: the birth of a new channel superfamily. J. Neurogenet. 2010, 24, 216–233. [Google Scholar] [CrossRef] [Green Version]
  30. Nilius, B.; Mahieu, F. A Road Map for TR(I)Ps. Mol. Cell 2006, 22, 297–307. [Google Scholar] [CrossRef]
  31. Benemei, S.; Patacchini, R.; Trevisani, M.; Geppetti, P. TRP channels. Curr. Opin. Pharmacol. 2015, 22, 18–23. [Google Scholar] [CrossRef] [PubMed]
  32. Nilius, B.; Flockerzi, V. Mammalian Transient Receptor Potential (TRP) Cation Channels. Handb. Exp. Pharmacol. 2014, 223, 795–826. [Google Scholar] [CrossRef]
  33. Kaneko, Y.; Szallasi, A. Transient receptor potential (TRP) channels: A clinical perspective. Br. J. Pharmacol. 2014, 171, 2474–2507. [Google Scholar] [CrossRef]
  34. Gees, M.; Owsianik, G.; Voets, T. TRP Channels. Compr. Physiol. 2012, 2, 563–608. [Google Scholar] [CrossRef]
  35. Nilius, B. TRP channels in disease. Biochim. Biophys. Acta Mol. Basis Dis. 2007, 1772, 805–812. [Google Scholar] [CrossRef] [Green Version]
  36. Prevarskaya, N.; Zhang, L.; Barritt, G. TRP channels in cancer. Biochim. Biophys. Acta Mol. Basis Dis. 2007, 1772, 937–946. [Google Scholar] [CrossRef] [Green Version]
  37. Van Haute, C.; De Ridder, D.; Nilius, B. TRP channels in human prostate. Sci. World J. 2010, 10, 1597–1611. [Google Scholar] [CrossRef] [Green Version]
  38. Lehen’Kyi, V.; Prevarskaya, N. Oncogenic TRP channels. Adv. Exp. Med. Biol. 2010, 704, 929–945. [Google Scholar] [CrossRef]
  39. Ouadid-Ahidouch, H.; Dhennin-Duthille, I.; Gautier, M.; Sevestre, H.; Ahidouch, A. TRP channels: Diagnostic markers and therapeutic targets for breast cancer? Trends Mol. Med. 2013, 19, 117–124. [Google Scholar] [CrossRef]
  40. Lepage, P.K.; Lussier, M.P.; Barajas-Martinez, H.; Bousquet, S.M.; Blanchard, A.P.; Francoeur, N.; Dumaine, R.; Boulay, G. Identification of two domains involved in the assembly of transient receptor potential canonical channels. J. Biol. Chem. 2006, 281, 30356–30364. [Google Scholar] [CrossRef] [Green Version]
  41. Birnbaumer, L. The TRPC class of ion channels: a critical review of their roles in slow, sustained increases in intracellular Ca(2+) concentrations. Annu. Rev. Pharmacol. Toxicol. 2009, 49, 395–426. [Google Scholar] [CrossRef] [PubMed]
  42. Wu, X.; Zagranichnaya, T.K.; Gurda, G.T.; Eves, E.M.; Villereal, M.L. A TRPC1/TRPC3-mediated increase in store-operated calcium entry is required for differentiation of H19-7 hippocampal neuronal cells. J. Biol. Chem. 2004, 279, 43392–43402. [Google Scholar] [CrossRef] [Green Version]
  43. Kuang, C.; Yu, Y.; Wang, K.; Qian, D.; Den, M.; Huang, L. Knockdown of Transient Receptor Potential Canonical-1 Reduces the Proliferation and Migration of Endothelial Progenitor Cells. Stem. Cells Dev. 2012, 21, 487–496. [Google Scholar] [CrossRef] [PubMed]
  44. Wes, P.D.; Chevesich, J.; Jeromin, A.; Rosenberg, C.; Stetten, G.; Montell, C. TRPC1, a human homolog of a Drosophila store-operated channel. Proc. Natl. Acad. Sci. USA 1995, 92, 9652–9656. [Google Scholar] [CrossRef] [Green Version]
  45. Parekh, A.B.; Putney, J.W. Store-operated calcium channels. Physiol. Rev. 2005, 85, 757–810. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Beech, D.J. TRPC1: Store-operated channel and more. Pflug. Arch. Eur. J. Physiol. 2005, 451, 53–60. [Google Scholar] [CrossRef] [PubMed]
  47. Ambudkar, I.S. TRPC1: a core component of store-operated calcium channels. Biochem. Soc. Trans. 2007, 35, 96–100. [Google Scholar] [CrossRef] [Green Version]
  48. Storch, U.; Forst, A.L.; Philipp, M.; Gudermann, T.; Mederos, Y.; Schnitzler, M. Transient receptor potential channel 1 (TRPC1) reduces calcium permeability in heteromeric channel complexes. J. Biol. Chem. 2012, 287, 3530–3540. [Google Scholar] [CrossRef] [Green Version]
  49. Nesin, V.; Tsiokas, L. TRPC1. Handb. Exp. Pharmacol. 2014, 222, 15–51. [Google Scholar] [CrossRef]
  50. Liu, X.; Bandyopadhyay, B.C.; Singh, B.B.; Groschner, K.; Ambudkar, I.S. Molecular analysis of a store-operated and OAG sensitive non-selective cation channel: Heteromeric assembly of TRPC1-TRPC3. J. Biol. Chem. 2005, 280, 21600–21606. [Google Scholar] [CrossRef] [Green Version]
  51. Rao, J.N.; Platoshyn, O.; Golovina, V.; Liu, L.; Zou, T.; Marasa, B.S.; Turner, D.J.; Yuan, J.X.J.; Wang, J.Y. TRPC1 functions as a store-operated Ca2+ channel in intestinal epithelial cells and regulates early mucosal restitution after wounding. Am. J. Physiol. Gastrointest Liver Physiol. 2006, 290, G782–G792. [Google Scholar] [CrossRef] [Green Version]
  52. Bomben, V.C.; Sontheimer, H.W. Inhibition of transient receptor potential canonical channels impairs cytokinesis in human malignant gliomas. Cell Prolif. 2008, 41, 98–121. [Google Scholar] [CrossRef]
  53. Dhennin-Duthille, I.; Gautier, M.; Faouzi, M.; Guilbert, A.; Brevet, M.; Vaudry, D.; Ahidouch, A.; Sevestre, H.; Ouadid-Ahidouch, H. High expression of transient receptor potential channels in human breast cancer epithelial cells and tissues: Correlation with pathological parameters. Cell. Physiol. Biochem. 2011, 28, 813–822. [Google Scholar] [CrossRef]
  54. Antigny, F.; Sabourin, J.; Saüc, S.; Bernheim, L.; Koenig, S.; Frieden, M. TRPC1 and TRPC4 channels functionally interact with STIM1L to promote myogenesis and maintain fast repetitive Ca2??+ release in human myotubes. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 806–813. [Google Scholar] [CrossRef]
  55. Wissenbach, U.; Schroth, G.; Philipp, S.; Flockerzi, V. Structure and mRNA expression of a bovine trp homologue related to mammalian trp2 transcripts. FEBS Lett. 1998, 429, 61–66. [Google Scholar] [CrossRef] [Green Version]
  56. Vannier, B.; Peyton, M.; Boulay, G.; Brown, D.; Qin, N.; Jiang, M.; Zhu, X.; Birnbaumer, L. Mouse trp2, the homologue of the human trpc2 pseudogene, encodes mTrp2, a store depletion-activated capacitative Ca2+ entry channel. Proc. Natl. Acad. Sci. USA 1999, 96, 2060–2064. [Google Scholar] [CrossRef] [Green Version]
  57. Zufall, F. TRPs in olfaction. Handb. Exp. Pharmacol. 2014, 223, 917–933. [Google Scholar] [CrossRef]
  58. Miller, B.A. TRPC2. Handb. Exp. Pharmacol. 2014, 222, 53–65. [Google Scholar] [CrossRef] [PubMed]
  59. Yuan, J.P.; Kiselyov, K.; Shin, D.M.; Chen, J.; Shcheynikov, N.; Kang, S.H.; Dehoff, M.H.; Schwarz, M.K.; Seeburg, P.H.; Muallem, S.; et al. Homer binds TRPC family channels and is required for gating of TRPC1 by IP3 receptors. Cell 2003, 114, 777–789. [Google Scholar] [CrossRef] [Green Version]
  60. Lockwich, T.P.; Liu, X.; Singh, B.B.; Jadlowiec, J.; Weiland, S.; Ambudkar, I.S. Assembly of Trp1 in a signaling complex associated with caveolin-scaffolding lipid raft domains. J. Biol. Chem. 2000, 275, 11934–11942. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  61. Mast, T.G.; Brann, J.H.; Fadool, D.A. The TRPC2 channel forms protein-protein interactions with Homer and RTP in the rat vomeronasal organ. BMC Neurosci. 2010, 11, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Sukumaran, P.; Löf, C.; Pulli, I.; Kemppainen, K.; Viitanen, T.; Törnquist, K. Significance of the transient receptor potential canonical 2 (TRPC2) channel in the regulation of rat thyroid FRTL-5 cell proliferation, migration, adhesion and invasion. Mol. Cell Endocrinol. 2013, 374, 10–21. [Google Scholar] [CrossRef] [PubMed]
  63. Törnquist, K.; Sukumaran, P.; Kemppainen, K.; Löf, C.; Viitanen, T. Canonical transient receptor potential channel 2 (TRPC2): old name-new games. Importance in regulating of rat thyroid cell physiology. Pflug. Arch. Eur. J. Physiol. 2014, 466, 2025–2034. [Google Scholar] [CrossRef] [PubMed]
  64. Putney, J.W. Physiological mechanisms of TRPC activation. Pflug. Arch. Eur. J. Physiol. 2005, 451, 29–34. [Google Scholar] [CrossRef] [PubMed]
  65. Owsianik, G.; Talavera, K.; Voets, T.; Nilius, B. Permeation and selectivity of trp channels. Annu. Rev. Physiol. 2006, 68, 685–717. [Google Scholar] [CrossRef] [PubMed]
  66. Liao, Y.; Plummer, N.W.; George, M.D.; Abramowitz, J.; Zhu, M.X.; Birnbaumer, L. A role for Orai in TRPC-mediated Ca2+ entry suggests that a TRPC:Orai complex may mediate store and receptor operated Ca2+ entry. Proc. Natl. Acad. Sci. USA 2009, 106, 3202–3206. [Google Scholar] [CrossRef] [Green Version]
  67. Woodard, G.E.; López, J.J.; Jardín, I.; Salido, G.M.; Rosado, J.A. TRPC3 regulates agonist-stimulated Ca2+ mobilization by mediating the interaction between type I inositol 1,4,5-trisphosphate receptor, RACK1, and Orai1. J. Biol. Chem. 2010, 285, 8045–8053. [Google Scholar] [CrossRef] [Green Version]
  68. Trebak, M.; Bird, G.S.J.; McKay, R.R.; Birnbaumer, L.; Putney, J.W. Signaling mechanism for receptor-activated canonical transient receptor potential 3 (TRPC3) channels. J. Biol. Chem. 2003, 278, 16244–16252. [Google Scholar] [CrossRef] [Green Version]
  69. Yuan, J.P.; Zeng, W.; Huang, G.N.; Worley, P.F.; Muallem, S. STIM1 heteromultimerizes TRPC channels to determine their function as store-operated channels. Nat. Cell Biol. 2007, 9, 636–645. [Google Scholar] [CrossRef]
  70. Qin, W.; Xie, W.; Xia, N.; He, Q.; Sun, T. Silencing of transient receptor potential channel 4 alleviates oxLDL-induced angiogenesis in human coronary artery endothelial cells by inhibition of VEGF and NF-κB. Med. Sci. Monit. 2016, 22, 930–936. [Google Scholar] [CrossRef] [Green Version]
  71. Schaefer, M.; Plant, T.D.; Obukhov, A.G.; Hofmann, T.; Gudermann, T.; Schultz, G. Receptor-mediated regulation of the nonselective cation channels TRPC4 and TRPC5. J. Biol. Chem. 2000, 275, 17517–17526. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Shen, B.; Wong, C.O.; Lau, O.C.; Woo, T.; Bai, S.; Huang, Y.; Yao, X. Plasma membrane mechanical stress activates TRPC5 channels. PLoS ONE 2015, 10, e0122227. [Google Scholar] [CrossRef] [Green Version]
  73. Earley, S.; Brayden, J.E. Transient receptor potential channels in the vasculature. Physiol. Rev. 2015, 95, 645–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Dietrich, A.; Mederos YSchnitzler, M.; Gollasch, M.; Gross, V.; Storch, U.; Dubrovska, G.; Obst, M.; Yildirim, E.; Salanova, B.; Kalwa, H.; et al. Increased vascular smooth muscle contractility in TRPC6-/- mice. Mol. Cell Biol. 2005, 25, 6980–6989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Liao, Y.; Erxleben, C.; Yildirim, E.; Abramowitz, J.; Armstrong, D.L.; Birnbaumer, L. Orai proteins interact with TRPC channels and confer responsiveness to store depletion. Proc. Natl. Acad. Sci. USA 2007, 104, 4682–4687. [Google Scholar] [CrossRef] [Green Version]
  76. Shirakawa, H.; Katsumoto, R.; Iida, S.; Miyake, T.; Higuchi, T.; Nagashima, T.; Nagayasu, K.; Nakagawa, T.; Kaneko, S. Sphingosine-1-phosphate induces Ca2+ signaling and CXCL1 release via TRPC6 channel in astrocytes. Glia 2017, 65, 1005–1016. [Google Scholar] [CrossRef] [Green Version]
  77. Park, S.; Lee, S.; Park, E.J.; Kang, M.J.; So, I.; Jeon, J.H.; Chun, J.N. TGFβ1 induces stress fiber formation through upregulation of TRPC6 in vascular smooth muscle cells. Biochem. Biophys. Res. Commun. 2017, 483, 129–134. [Google Scholar] [CrossRef]
  78. Weber, E.W.; Han, F.; Tauseef, M.; Birnbaumer, L.; Mehta, D.; Muller, W.A. TRPC6 is the endothelial calcium channel that regulates leukocyte transendothelial migration during the inflammatory response. J. Exp. Med. 2015, 212, 1883–1899. [Google Scholar] [CrossRef]
  79. Nagamine, K.; Kudoh, J.; Minoshima, S.; Kawasaki, K.; Asakawa, S.; Ito, F.; Shimizu, N. Molecular cloning of a novel putative Ca2+ channel protein (TRPC7) highly expressed in brain. Genomics 1998, 54, 124–131. [Google Scholar] [CrossRef]
  80. Okada, T.; Inoue, R.; Yamazaki, K.; Maeda, A.; Kurosaki, T.; Yamakuni, T.; Tanaka, I.; Shimizu, S.; Ikenaka, K.; Imoto, K.; et al. Molecular and functional characterization of a novel mouse transient receptor potential protein homologue TRP7. J. Biol. Chem. 1999, 274, 27359–27370. [Google Scholar] [CrossRef] [Green Version]
  81. Zagranichnaya, T.K.; Wu, X.; Villereal, M.L. Endogenous TRPC1, TRPC3, and TRPC7 proteins combine to form native store-operated channels in HEK-293 cells. J. Biol. Chem. 2005, 280, 29559–29569. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Yuasa, K.; Matsuda, T.; Tsuji, A. Functional regulation of transient receptor potential canonical 7 by cGMP-dependent protein kinase Iα. Cell. Signal. 2011, 23, 1179–1187. [Google Scholar] [CrossRef] [PubMed]
  83. Satoh, S.; Tanaka, H.; Ueda, Y.; Oyama, J.I.; Sugano, M.; Sumimoto, H.; Mori, Y.; Makino, N. Transient receptor potential (TRP) protein 7 acts as a G protein-activated Ca2+ channel mediating angiotensin II-induced myocardial apoptosis. Mol. Cell. Biochem. 2007, 294, 205–215. [Google Scholar] [CrossRef]
  84. Schwab, A.; Fabian, A.; Hanley, P.J.; Stock, C. Role of Ion Channels and Transporters in Cell Migration. Physiol. Rev. 2012, 92, 1865–1913. [Google Scholar] [CrossRef]
  85. Déliot, N.; Constantin, B. Plasma membrane calcium channels in cancer: Alterations and consequences for cell proliferation and migration. Biochim. Biophys. Acta Biomembr. 2015, 1848, 2512–2522. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Vrenken, K.S.; Jalink, K.; van Leeuwen, F.N.; Middelbeek, J. Beyond ion-conduction: Channel-dependent and -independent roles of TRP channels during development and tissue homeostasis. Biochim. Biophys. Acta Mol. Cell Res. 2016, 1863, 1436–1446. [Google Scholar] [CrossRef] [PubMed]
  87. Fabian, A.; Fortmann, T.; Dieterich, P.; Riethmüller, C.; Schön, P.; Mally, S.; Nilius, B.; Schwab, A. TRPC1 channels regulate directionality of migrating cells. Pflug. Arch. Eur. J. Physiol. 2008, 457, 475–484. [Google Scholar] [CrossRef]
  88. Bomben, V.C.; Turner, K.L.; Barclay, T.T.C.; Sontheimer, H. Transient receptor potential canonical channels are essential for chemotactic migration of human malignant gliomas. J. Cell Physiol. 2011, 226, 1879–1888. [Google Scholar] [CrossRef] [Green Version]
  89. 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]
  90. He, B.; Liu, F.; Ruan, J.; Li, A.; Chen, J.; Li, R.; Shen, J.; Zheng, D.; Luo, R. Silencing TRPC1 expression inhibits invasion of CNE2 nasopharyngeal tumor cells. Oncol. Rep. 2012, 27. [Google Scholar] [CrossRef]
  91. Asghar, M.Y.; Magnusson, M.; Kemppainen, K.; Sukumaran, P.; Löf, C.; Pulli, I.; Kalhori, V.; Törnquist, K. Transient receptor potential canonical 1 (TRPC1) channels as regulators of sphingolipid and VEGF receptor expression: Implications for thyroid cancer cell migration and proliferation. J. Biol. Chem. 2015, 290, 16116–16131. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Törnquist, K. Sphingosine 1-phosphate, sphingosine kinase and autocrine calcium signalling in thyroid cells. Acta Physiol. 2012, 204, 151–157. [Google Scholar] [CrossRef] [PubMed]
  93. Wildi-Runge, S.; Stoppa-Vaucher, S.; Lambert, R.; Turpin, S.; Van Vliet, G.; Deladoëy, J. A high prevalence of dual thyroid ectopy in congenital hypothyroidism: Evidence for insufficient signaling gradients during embryonic thyroid migration or for the polyclonal nature of the thyroid gland? J. Clin. Endocrinol. Metab. 2012, 97, E978–E981. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Kimchi, T.; Xu, J.; Dulac, C. A functional circuit underlying male sexual behaviour in the female mouse brain. Nature 2007, 448, 1009–1014. [Google Scholar] [CrossRef]
  95. Yang, S.L.; Cao, Q.; Zhou, K.C.; Feng, Y.J.; Wang, Y.Z. Transient receptor potential channel C3 contributes to the progression of human ovarian cancer. Oncogene 2009, 28, 1320–1328. [Google Scholar] [CrossRef] [Green Version]
  96. Storck, H.; Hild, B.; Schimmelpfennig, S.; Sargin, S.; Nielsen, N.; Zaccagnino, A.; Budde, T.; Novak, I.; Kalthoff, H.; Schwab, A. Ion channels in control of pancreatic stellate cell migration. Oncotarget 2016, 8, 769–784. [Google Scholar] [CrossRef] [Green Version]
  97. Oda, K.; Umemura, M.; Nakakaji, R.; Tanaka, R.; Sato, I.; Nagasako, A.; Oyamada, C.; Baljinnyam, E.; Katsumata, M.; Xie, L.H.; et al. Transient receptor potential cation 3 channel regulates melanoma proliferation and migration. J. Physiol. Sci. 2016, 67, 497–505. [Google Scholar] [CrossRef]
  98. Turunen, P.M.; Louhivuori, L.M.; Louhivuori, V.; Kukkonen, J.P.; Åkerman, K.E. Endocannabinoid Signaling in Embryonic Neuronal Motility and Cell–Cell Contact—Role of mGluR5 and TRPC3 Channels. Neuroscience 2018, 375, 135–148. [Google Scholar] [CrossRef] [Green Version]
  99. Kim, J.M.; Heo, K.; Choi, J.; Kim, K.; An, W. The histone variant MacroH2A regulates Ca2+ influx through TRPC3 and TRPC6 channels. Oncogenesis 2013, 2, e77. [Google Scholar] [CrossRef]
  100. Zeng, B.; Yuan, C.; Yang, X.; LAtkin, S.; Xu, S.-Z. TRPC Channels and Their Splice Variants are Essential for Promoting Human Ovarian Cancer Cell Proliferation and Tumorigenesis. Curr. Cancer Drug Targets 2012, 13, 103–116. [Google Scholar] [CrossRef]
  101. Wei, W.C.; Huang, W.C.; Lin, Y.P.; Becker, E.B.E.; Ansorge, O.; Flockerzi, V.; Conti, D.; Cenacchi, G.; Glitsch, M.D. Functional expression of calcium-permeable canonical transient receptor potential 4-containing channels promotes migration of medulloblastoma cells. J. Physiol. 2017, 595, 5525–5544. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Jiang, H.N.; Zeng, B.; Zhang, Y.; Daskoulidou, N.; Fan, H.; Qu, J.M.; Xu, S.Z. Involvement of TRPC Channels in Lung Cancer Cell Differentiation and the Correlation Analysis in Human Non-Small Cell Lung Cancer. PLoS ONE 2013, 8, e67637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Greka, A.; Navarro, B.; Oancea, E.; Duggan, A.; Clapham, D.E. TRPC5 is a regulator of hippocampal neurite length and growth cone morphology. Nat. Neurosci. 2003, 6, 837–845. [Google Scholar] [CrossRef] [PubMed]
  104. Tian, D.; Jacobo, S.M.P.; Billing, D.; Rozkalne, A.; Gage, S.D.; Anagnostou, T.; Pavenstaedt, H.; Hsu, H.H.; Schlondorff, J.; Ramos, A.; et al. Antagonistic regulation of actin dynamics and cell motility by TRPC5 and TRPC6 channels. Sci. Signal. 2010, 3, ra77. [Google Scholar] [CrossRef] [Green Version]
  105. Carson, C.; Raman, P.; Tullai, J.; Xu, L.; Henault, M.; Thomas, E.; Yeola, S.; Lao, J.; McPate, M.; Martin Verkuyl, J.; et al. Englerin a agonizes the TRPC4/C5 cation channels to inhibit tumor cell line proliferation. PLoS ONE 2015, 10, e0127498. [Google Scholar] [CrossRef]
  106. Muraki, K.; Ohnishi, K.; Takezawa, A.; Suzuki, H.; Hatano, N.; Muraki, Y.; Hamzah, N.; Foster, R.; Waldmann, H.; Nussbaumer, P.; et al. Na+ entry through heteromeric TRPC4/C1 channels mediates (-)Englerin A-induced cytotoxicity in synovial sarcoma cells. Sci. Rep. 2017, 7, 16988. [Google Scholar] [CrossRef] [Green Version]
  107. Bernaldo de Quirós, S.; Merlo, A.; Secades, P.; Zambrano, I.; de Santa María, I.S.; Ugidos, N.; Jantus-Lewintre, E.; Sirera, R.; Suarez, C.; Chiara, M.D. Identification of TRPC6 as a possible candidate target gene within an amplicon at 11q21-q22.2 for migratory capacity in head and neck squamous cell carcinomas. BMC Cancer 2013, 13, 116. [Google Scholar] [CrossRef] [Green Version]
  108. Chigurupati, S.; Venkataraman, R.; Barrera, D. Receptor channel TRPC6 is a key mediator of Notch-driven glioblastoma growth and invasiveness. Cancer Res. 2010, 70, 418–427. [Google Scholar] [CrossRef] [Green Version]
  109. 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 Ca2+ entry and proliferation of prostate cancer epithelial cells. Cancer Res. 2006, 66, 2038–2047. [Google Scholar] [CrossRef] [Green Version]
  110. Yang, L.L.; Liu, B.C.; Lu, X.Y.; Yan, Y.; Zhai, Y.J.; Bao, Q.; Doetsch, P.W.; Deng, X.; Thai, T.L.; Alli, A.A.; et al. Inhibition of TRPC6 reduces non-small cell lung cancer cell proliferation and invasion. Oncotarget 2016, 8, 5123–5134. [Google Scholar] [CrossRef]
  111. Diez-Bello, R.; Jardin, I.; Lopez, J.J.; El Haouari, M.; Ortega-Vidal, J.; Altarejos, J.; Salido, G.M.; Salido, S.; Rosado, J.A. (−)-Oleocanthal inhibits proliferation and migration by modulating Ca2+ entry through TRPC6 in breast cancer cells. Biochim. Biophys. Acta Mol. Cell Res. 2019, 1866, 474–485. [Google Scholar] [CrossRef] [PubMed]
  112. Jardin, I.; Diez-Bello, R.; Lopez, J.J.; Redondo, P.C.; Salido, G.M.; Smani, T.; Rosado, J.A. Trpc6 channels are required for proliferation, migration and invasion of breast cancer cell lines by modulation of orai1 and orai3 surface exposure. Cancers 2018, 10, 331. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  113. Andrikopoulos, P.; Eccles, S.A.; Yaqoob, M.M. Coupling between the TRPC3 ion channel and the NCX1 transporter contributed to VEGF-induced ERK1/2 activation and angiogenesis in human primary endothelial cells. Cell. Signal. 2017, 37, 12–30. [Google Scholar] [CrossRef] [PubMed]
  114. Song, H.B.; Jun, H.O.; Kim, J.H.; Fruttiger, M.; Kim, J.H. Suppression of transient receptor potential canonical channel 4 inhibits vascular endothelial growth factor-induced retinal neovascularization. Cell Calcium 2015, 57, 101–108. [Google Scholar] [CrossRef] [PubMed]
  115. Antigny, F.; Girardin, N.; Frieden, M. Transient receptor potential canonical channels are required for in vitro endothelial tube formation. J. Biol. Chem. 2011, 287, 5917–5927. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Zhu, Y.; Gao, M.; Zhou, T.; Xie, M.; Mao, A.; Feng, L.; Yao, X.; Wong, W.T.; Ma, X. The TRPC5 channel regulates angiogenesis and promotes recovery from ischemic injury in mice. J. Biol. Chem. 2018, 294, 28–37. [Google Scholar] [CrossRef] [Green Version]
  117. Hamdollah Zadeh, M.A.; Glass, C.; Magnussen, A.; Hancox, J.; Bates, D. VEGF-Mediated elevated intracellular calcium and angiogenesis in human microvascular endothelial cells In vitro are inhibited by dominant negative TRPC6. Microcirculation 2008, 15, 605–614. [Google Scholar] [CrossRef] [Green Version]
  118. Kini, V.; Chavez, A.; Mehta, D. A new role for PTEN in regulating transient receptor potential canonical channel 6-mediated Ca2+ entry, endothelial permeability, and angiogenesis. J. Biol. Chem. 2010, 285, 33082–33091. [Google Scholar] [CrossRef] [Green Version]
  119. Smani, T.; Gómez, L.J.; Regodon, S.; Woodard, G.E.; Siegfried, G.; Khatib, A.M.; Rosado, J.A. Trp channels in angiogenesis and other endothelial functions. Front. Physiol. 2018, 9. [Google Scholar] [CrossRef]
  120. Moccia, F.; Negri, S.; Shekha, M.; Faris, P.; Guerra, G. Endothelial Ca2+ signaling, angiogenesis and vasculogenesis: Just what it takes to make a blood vessel. Int. J. Mol. Sci. 2019, 20, 3962. [Google Scholar] [CrossRef] [Green Version]
  121. Thakore, P.; Earley, S. Transient Receptor Potential Channels and Endothelial Cell Calcium Signaling. Compr. Physiol. 2019, 9, 1249–1277. [Google Scholar] [CrossRef] [PubMed]
  122. Petros, R.A.; Desimone, J.M. Strategies in the design of nanoparticles for therapeutic applications. Nat. Rev. Drug Discov. 2010, 9, 615–627. [Google Scholar] [CrossRef] [PubMed]
  123. Rahikkala, A.; Rosenholm, J.M.; Santos, H.A. Biofunctionalized Mesoporous Silica Nanomaterials for Targeted Drug Delivery. Biomed. Appl. Funct. Nanomater. 2018, 489–520. [Google Scholar] [CrossRef]
Figure 1. Mechanisms of Ca2+ signaling. Upon activation of a G-protein coupled receptor (GPCR) by an agonist, phospholipase C (PLC) is activated, which generates two second messengers; diacylglycerol (DAG) and inositol 1,4,5-triphosphate (IP3). DAG is capable of activating TRPC channels in the plasma membrane and Ca2+ influx is triggered. IP3 diffuses through the cytoplasm and binds to IP3 receptors on the endoplasmic reticulum (ER) membranes. This binding enables ER depletion, resulting in a rapid Ca2+ transient. The depletion of ER is sensed by stromal interaction protein 1 (STIM1) proteins which act as sensors. STIM1 makes a complex with Orai1 channels in the plasma membrane and induces store operated Ca2+ entry through Orai1. Voltage-operated calcium channels open in response to a depolarization of the plasma membrane in excitable neural and muscle cells. Ca2+ in the cytoplasm activates ryanodine receptors and Ca2+ is released from the endoplasmic reticulum (ER). To avoid a Ca2+ flood in the cytoplasm, the pumps sarcoplasmic-endoplasmic reticulum Ca2+ATPase (SERCA) and plasma membrane Ca2+ATPase (PMCA), and the sodium calcium exchangers (NCX), are activated which export Ca2+ out of the cell or into the ER. The secretory pathway Ca2+ ATPase (SPCA) transports Ca2+ ions into the Golgi apparatus.
Figure 1. Mechanisms of Ca2+ signaling. Upon activation of a G-protein coupled receptor (GPCR) by an agonist, phospholipase C (PLC) is activated, which generates two second messengers; diacylglycerol (DAG) and inositol 1,4,5-triphosphate (IP3). DAG is capable of activating TRPC channels in the plasma membrane and Ca2+ influx is triggered. IP3 diffuses through the cytoplasm and binds to IP3 receptors on the endoplasmic reticulum (ER) membranes. This binding enables ER depletion, resulting in a rapid Ca2+ transient. The depletion of ER is sensed by stromal interaction protein 1 (STIM1) proteins which act as sensors. STIM1 makes a complex with Orai1 channels in the plasma membrane and induces store operated Ca2+ entry through Orai1. Voltage-operated calcium channels open in response to a depolarization of the plasma membrane in excitable neural and muscle cells. Ca2+ in the cytoplasm activates ryanodine receptors and Ca2+ is released from the endoplasmic reticulum (ER). To avoid a Ca2+ flood in the cytoplasm, the pumps sarcoplasmic-endoplasmic reticulum Ca2+ATPase (SERCA) and plasma membrane Ca2+ATPase (PMCA), and the sodium calcium exchangers (NCX), are activated which export Ca2+ out of the cell or into the ER. The secretory pathway Ca2+ ATPase (SPCA) transports Ca2+ ions into the Golgi apparatus.
Ijms 21 01739 g001
Figure 2. The human phylogenetic tree of the TRP channel superfamily. TRPC (canonical), TRPV (Vanilloid), TRPML (Mucolipin), TRPP (Polycystin), TRPM (Melastatin), TRPA (Ankyrin), and TRPN (NOMPC). TRPC2 is a pseudogene in human. TRPN is expressed in fish.
Figure 2. The human phylogenetic tree of the TRP channel superfamily. TRPC (canonical), TRPV (Vanilloid), TRPML (Mucolipin), TRPP (Polycystin), TRPM (Melastatin), TRPA (Ankyrin), and TRPN (NOMPC). TRPC2 is a pseudogene in human. TRPN is expressed in fish.
Ijms 21 01739 g002
Figure 3. TRPC channel transmembrane structure and the domain organization of the subunits.
Figure 3. TRPC channel transmembrane structure and the domain organization of the subunits.
Ijms 21 01739 g003
Table 1. Expression of transient receptor potential (TRP) channels in worms, flies, mice, and humans.
Table 1. Expression of transient receptor potential (TRP) channels in worms, flies, mice, and humans.
Ijms 21 01739 i001 Ijms 21 01739 i002 Ijms 21 01739 i003 Ijms 21 01739 i004
TRPC3376
TRPV5266
TRPM4188
TRPA2411
TRPML1133
TRPP1133
TRPN1100
Total17132827
Table 2. Effects of TRPC channels on invasion, migration, and proliferation.
Table 2. Effects of TRPC channels on invasion, migration, and proliferation.
ChannelTissueCell LineCell functionReferences
TRPC1KidneyMDCK-FMigration[87]
Brain cancerD54MGMigration[88]
Colon cancerHCT-116Migration[89]
Skeletal MuscleC2C12Migration, differentiation[11]
Nasopharyngeal cancerCNE2Migration, Invasion[90]
Thyroid cancerML-1Migration and Invasion[91]
TRPC2Rat ThyroidFRTL5Proliferation, adhesion, migration and invasion[62]
TRPC3Ovarian cancerSKOV3Migration[95]
Pancreatic stellate cellsRLT-PSCMigration[96]
MelanomaC8161Migration[97]
Neuronal progenitor cellNPCsMigration[98]
Bladder CancerLD611Migration and Invasion[99]
TRPC4Coronary arteryHCAECsProliferation[70]
Ovarian cancerSKOV3Proliferation[100]
MedulloblastomaDAOYMigration, Invasion[101]
Lung cancerA549Proliferation[102]
TRPC5Rat hippocampal neuronE18Inhibition of neurite extension[103]
Mouse KidneyPodocytesPromote migration[104]
TRPC6Mouse KidneyPodocytesInhibit migration[109]
Head and neck cancerHNSCCPromote Invasion[107]
Brain cancerU373MGMigration, cell growth[108]
Prostate cancerhPCEProliferation[109]
Lung cancerA549Proliferation[110]
Breast CancerMDA-MB-231Migration[111]

Share and Cite

MDPI and ACS Style

Asghar, M.Y.; Törnquist, K. Transient Receptor Potential Canonical (TRPC) Channels as Modulators of Migration and Invasion. Int. J. Mol. Sci. 2020, 21, 1739. https://doi.org/10.3390/ijms21051739

AMA Style

Asghar MY, Törnquist K. Transient Receptor Potential Canonical (TRPC) Channels as Modulators of Migration and Invasion. International Journal of Molecular Sciences. 2020; 21(5):1739. https://doi.org/10.3390/ijms21051739

Chicago/Turabian Style

Asghar, Muhammad Yasir, and Kid Törnquist. 2020. "Transient Receptor Potential Canonical (TRPC) Channels as Modulators of Migration and Invasion" International Journal of Molecular Sciences 21, no. 5: 1739. https://doi.org/10.3390/ijms21051739

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