NCBI Bookshelf. A service of the National Library of Medicine, National Institutes of Health.
Zhu MX, editor. TRP Channels. Boca Raton (FL): CRC Press/Taylor & Francis; 2011.
17.1. INTRODUCTION
Many proteins in cancer cells exhibit increased or decreased expression compared to their levels in normal cells. Some of these proteins, i.e., those encoded by oncogenes and tumor suppressor genes, play key roles in tumorigenesis and in the development of metastases, while others, most likely including those involved in intracellular Ca2+ homeostasis (reviewed in Refs. 1 and 2), are associated with cancer progression but are not causative in further development of the tumor and/or malignant cells (reviewed in Ref. 3). Most cancers are heterogeneous with respect to rates of growth and degrees of aggression. This most likely reflects the fact that, for a given cancer, there can be different combinations of oncogenes and tumor suppressor genes that are mutated, different sequences in which these mutations occur, and variations in the time over which the mutations accumulate.3
Some of the most important signaling pathways altered in tumorigenesis enhance cell proliferation and inhibit apoptosis. Ca2+ homeostasis controls these cellular processes, including proliferation, apoptosis, gene transcription, and angiogenesis.4 Ca2+ signaling is thus required for cell proliferation in all eukaryote cells, while some transformed cells and tumor cell lines show reduced dependence on Ca2+ to maintain proliferation.5, 6 Furthermore, the regulation of cell cycles, apoptosis, or proliferation depends on the amplitude and temporal-spatial aspects of the Ca2+ signal,7, 8 thus highlighting the importance of Ca2+ signaling components such as Ca2+ channels. Indeed, dysfunctions in Ca2+ channels are involved in tumorigenesis because increased expression of plasma membrane Ca2+ channels amplifies Ca2+ influx with consequent promotion of Ca2+-dependent proliferative pathways.4, 7, 9
Transient receptor potential (TRP) channels contribute to changes in intracellular Ca2+ concentrations, either by acting as Ca2+ entry pathways in the plasma membrane or via changes in membrane polarization, modulating the driving force for Ca2+ entry mediated by alternative pathways,10 as well as the activity of voltage-gated Ca2+ channels. In addition, TRP channels are expressed on the membranes of internal Ca2+ stores11– 13 where they may act as triggers for enhanced proliferation, aberrant differentiation, and impaired ability to die, leading to the uncontrolled expansion and invasion characteristic of cancer.
All these normal as well as abnormal misregulations are within the scope of this chapter, which is aimed at the study of TRP channels at all known levels, such as mRNA transcription, splicing variants, mRNA translation into protein, further protein processing in Golgi, regulation of channel trafficking to the plasma membrane, and the final modulation of TRP channel activity at the plasma membrane.
17.1.1. TRP Channels and Cancer
The extent to which TRP channels are associated with cancer has been increasingly clarified in recent years. The approximately 30 TRPs identified to date are classified in six different families: TRPC (canonical), TRPV (vanilloid), TRPM (melastatin), TRPML (mucolipin), TRPP (polycystin), and TRPA (ankyrin transmembrane protein).14 The expression levels of members of the TRPC, TRPM, and TRPV families are correlated with the emergence and/or progression of certain epithelial cancers.15– 18 It has not yet been established whether these expression changes are drivers, required to sustain the transformed phenotype. Usually, the progression of cells from a normal, differentiated state to a tumorigenic, metastatic state involves the accumulation of mutations in multiple key signaling proteins, encoded by oncogenes and tumor suppressor genes, together with the evolution and clonal selection of more aggressive cell phenotypes.
Several recent works have shown that changes in the expression of TRP channels contribute to malignancy. The first evidence that TRP channel expression is correlated with different types of cancers came from the analysis of the expression of TRPM1. The expression of the TRPM1 gene is inversely correlated with the aggressiveness of melanoma malignant cells, which suggests that TRPM1 may behave as a tumor suppressor gene.18, 19 A second member of the TRPM subfamily, TRPM5, was shown to be responsible for the Beckwith-Wiedemann syndrome, a disease characterized by a childhood predisposition to tumors.20 TRPC channels are often coupled to a membrane receptor with which they work in synergy. Thus, it was demonstrated in prostate cancer cells that a Ca2+ signal can promote either cell proliferation or apoptosis,12, 15 depending on the type of TRPC channel involved: Ca2+ entry via TRPC6 channels stimulates cell proliferation, whereas TRPC1 and TRPC4 are mostly involved in apoptosis induction.
In addition, TRP channels could also play a key role in cancer progression. Indeed, this seems to be the case for TRPM8 and TRPV6 channels in prostate cancer. TRPM8 has originally been cloned from cancerous prostate tissue16 and was thereafter identified as an ion channel responding to cold stimuli.21 TRPM8 is expressed in normal prostate; however, its expression is increased in ADCaP. TRPV6 is strongly expressed in advanced prostate cancer, with little or no expression in healthy and benign prostate tissues.17, 22
Thus, to date, most changes involving TRP proteins do not involve mutations in the TRP gene but rather increased or decreased expression levels of the wild-type TRP protein, depending on the stage of the cancer. Table 17.1 summarizes these changes in cancer and metastatic cells.
17.1.2. TRP Channels as Potential Pharmaceutical Targets
Two aspects of the properties of TRP proteins and the association of increased or decreased expression of a given TRP protein with cancer and the progression of cancer have been used to try to develop strategies to kill cancer cells. One approach uses Ca2+ and Na+ entry through TRP channels expressed in cancer cells, which leads to a sustained high [Ca2+]cyt and cytoplasmic Na+ concentration ([Na+]cyt), a condition that kills cells by apoptosis and necrosis. This strategy requires the selective expression and activation of a given TRP channel in the targeted cancer cells. New strategies for killing cancer cells by activation of the apoptotic pathway are valuable because for many cancer cells, including androgen-insensitive prostate cancer cells, the normal pathways of apoptosis are inhibited, and the cells are resistant to apoptosis.23– 25 The other aspect makes use of the high expression of some TRP channels in cancer cells to provide a target for delivering a toxic payload (e.g., a radioactive nuclide or toxic chemical) to the cancer cells. Recognition of the TRP protein could be achieved through a tight-binding agonist or an anti-TRP antibody (reviewed in Ref. 26).
Studies with some other cancers also suggest that TRPV1 may be a useful target for killing cancer cells through a sustained increase in [Ca2+]cyt and [Na+]cyt. Resiniferatoxin, an analogue of capsaicin and an agonist of TRPV1, was shown to cause inhibition of mitochondrial function and induce apoptosis in pancreatic cancer cells, presumably via endogenous TRPV1 channels in the plasma membrane.27 It was suggested that vanilloids might be used to treat pancreatic cancer.27
Lignesti and colleagues tested the ability of various plant cannabinoids, which bind to the cannabinoid receptors (CB) and TRPV1 channels, to inhibit tumor cell growth.32 Using a panel of tumor cell lines as well as a xenograft mouse model of breast cancer (MDA-MB-231 cells), these investigators found that cannabinoids, of which cannabidiol was the most potent, inhibited cell and tumor growth. They suggested that cannabinoids may act through CB2 receptors and the TRPV1 channels. Endogenous cannabinoids play an important role in the neuronal control of the digestive tract.33 It has been suggested that the pharmacological administration of cannabinoids, which in part act through TRPV1 channels, could be used to treat colon cancer.33
As discussed above, TRPM8 is expressed in prostate cancer cells, and its expression decreases as the cancer progresses to a more metastatic state. Hence, TRPM8 is considered potentially useful both for the diagnosis of prostate cancer and as a target for cancer therapy. The treatment of prostate cancer would be greatly enhanced by better prediction of the course of the disease, including the likelihood of the development of androgen insensitivity and metastases, and by new strategies to kill androgen-insensitive prostate cancer cells, which, as mentioned above, are resistant to apoptosis.34, 35
Recently, a menthol analogue and TRPM8 agonist, WS-12, has been synthesized and characterized. WS-12 has an affinity for the TRPM8 menthol binding site, which is about 2000 times higher than that for menthol itself.36 Incorporation of a fluorine atom into WS-12 resulted in an analogue (WS-12F) that activated TRPM8 by 75% of the activation induced by WS-12 and retained a high affinity for TRPM8. It has been suggested that WS-12 and WS-12F offer potential possibilities in the detection of micro metastases and in killing prostate cancer cells. Thus, the incorporation of 18F into WS-12 may permit radio-imaging of micrometastases, and/or the delivery of a radionuclide, which could kill target cells, to specific locations of cancer cells in both the prostate and in metastases.36
17.2. ROLE OF TRP CHANNEL ISOFORM EXPRESSION AND FUNCTION IN THE STUDY OF THE CANCER INITIATION AND PROGRESSION
TRP proteins form tetrameric channel complexes, and at least the closely related members of one subfamily are capable of building heteromeric channels.37 The diversity of native TRP-related channels might be considerably increased by combining different TRP channel subunits to build a common ion-conducting pore. There is growing evidence that transcriptional regulation and alternative mRNA processing also contribute to the diversity of TRP channels. Some TRPs are expressed in two or more short splice variants, which may also exhibit different expression profiles in cancer as compared to the full-length forms. Alternative splicing enables the same gene to generate multiple mature mRNA types for translation, resulting in multiple-channel proteins. This leads to functional diversity, which, in turn, may have consequences for cellular function. Alternative splicing generates protein isoforms with different biological properties, such as a change in functionality, protein/protein interaction, or subcellular localization.38 Many of the splice variants are not functional and may not even be efficiently translated, so they may be considered negligible populations of incomplete or aberrantly spliced transcripts. Nevertheless, alternative splicing, as a regulatory process, contributes to biological complexity, not only by proteome expansion but also through its ability to control the expression of functional proteins. This may be achieved by producing nonfunctional isoforms of the gene by altering the domains necessary for TRP channel opening, membrane localization, or association.
This is the case for TRPV2, which expresses two transcripts in normal human urothelial cells and bladder tissue specimens: full-length TRPV2 and a short-splice variant, s-TRPV2. Analysis of TRPV2 gene and protein expressions in distinct superficial and invasive grades and stages indicates that the mRNA of TRPV2 increases gradually at increasing grades and stages, while that of s-TRPV2 gradually decreases.20 The authors suggested that the differences observed in the short/full TRPV2 ratio during tumorigenesis implied that s-TRPV2 was lost as an early event in bladder carcinogenesis, whereas the enhanced expression of full-length TRPV2 in high-stage muscle-invasive urothelial cancer is a secondary event. In a similar study, a different s-TRPV2, lacking the pore-forming region and the fifth and sixth transmembrane domains, was characterized in human macrophages.39 As for TRPV1, these naturally occurring alternative splice variants may act as dominant-negative mutants40 by forming a heterodimer with TRPV2 and inhibiting its trafficking and translocation to the plasma membrane.
In the case of TRPV1, the short isoform (TRPV1β) is produced by alternative splicing of the TRPV1 gene, with 10 amino acids missing near the end of the cytoplasmic N-terminus.40 TRPV1β does not form a functional channel when it is heterologously expressed alone, but exerts a dominant-negative effect on TRPV1 when they are co-expressed. Stability is affected when TRPV1β is assembled with the full-length channel, making less TRPV1 protein available at the plasma membrane. The residual amount of TRPV1β on the plasma membrane is not activated by factors known to stimulate TRPV1, but there are two other possibilities.40 Either the residual proteins are not properly assembled into tetrameric channels or channels that contain TRPV1β subunits cannot be opened. It should be noted, however, that TRPV1 Western blot analysis of the urothelium revealed two bands of equal intensity at 100 and 95 kDa, which decreased as the cancer progressed.19 Further investigation is required to determine whether these are the two splice TRPV1 isoforms and to analyze their expression regulation as cancer progresses. A similar mechanism is present in normal and benign melanocytes, which express the full-length TRPM1 mRNA, along with some shorter products.18 Heterologous co-expression of the full-length and short TRPM1 isoforms results in retention of the full-length channel in the endoplasmic reticulum (ER).13 However, it is currently unclear whether TRPM1 expression in metastasizing lines inhibits their growth. Metastatic melanomas lack the full-length transcript, but express several short fragments of TRPM1,18 probably owing to proteolysis of the full-length protein.41
TRPM8 also encodes some splice variants, comprising an altered N-terminus cloned from lung epithelia42 and cancerous prostate.43 The lung epithelia splice variant localizes preferentially to the ER, and its activation controls cell responses to cold air-induced inflammation.42 It has not yet been clarified whether this newly identified variant is implicated in cancer, whereas it may constitute a regulatory mechanism for the full-length TRPM8 in tissues where they both localize, such as liver, colon, and testis.42 Little information is available concerning the cancerous prostate TRPM8 isoform. It has a truncated N-terminus43 and may serve as a dominant negative regulator of full-length TRPM8, as suggested for TRPM1 truncated variants.13 Furthermore, a recent study by Bidaux and coworkers identified two TRPM8 isoforms with different androgen sensitivity and distinct localization on the plasma and ER membranes.11 Figure 17.1 shows schematically the TRPM8 gene, TRPM8 mRNA, and the position of Q-PCR primers used to distinguish between the two isoforms. The relative ratio between the two PCR products is used to calculate the predominance of one isoform over the other. This strategy may be also used to selectively knock down either of the isoforms. The differential regulation of TRPM8 activity may be due to complex regulation of the two isoforms by androgen receptors: An alternative TRPM8 gene promoter may make the ER TRPM8 isoform less sensitive to androgens. However, this ER localization may also result from a variation in the primary sequence leading to the appearance of an ER retention signal or the involvement of other associated proteins affecting its trafficking. It should be noted that there is a controversy in the literature concerning the localization of ER TRPM8. Two studies proposed a TRPM8-independent ER Ca2+ release mechanism in LNCaP 44 and PC345 cells when using high doses of menthol (3 mM)44 versus the ER TRPM8 activation with 100–250 μM menthol.12, 46
Furthermore, immunocytochemistry experiments in LNCaP revealed contradictory results concerning the presence of TRPM8 on the ER.12, 47 Two scenarios may explain this incongruity: firstly, as TRPM8 is under androgenic control, culture conditions of LNCaP cells, especially with respect of the serum used, may be critical for channel expression and localization, and, secondly, the putative ER TRPM8 isoform is not necessarily detected by the different antibodies used in these studies. In any case, the presence of ER TRPM8 was demonstrated in freshly isolated primary epithelial prostate cancer cells.11 Consequently, to clarify whether TRPM8 localizes into the ER, it is necessary to clone this putative ER-specific TRPM8 isoform and identify its distinguishing features, as compared to the two previously cloned variants.
Therefore, the abundant short or long mRNA forms in some cancers arise from a regulatory mechanism that produces either spliced or partially degraded nonproductive RNAs. These spliced transcripts form multimers and regulate targeting to the plasma or ER membrane, and consequently protein activity. Changes in TRP localization may have a causal or promoting role in cancer. The increases in constitutively active channels, such as TRPV6, in the plasma membrane of prostate cancer cells17, 30 may augment Ca2+ in the cytosol, thus promoting Ca2+-dependent proliferative pathways. The same may hold true for TRPM1 in melanocytes13, 18, 23 because Ca2+ imaging experiments on transfected HEK293 cells revealed an increase in intracellular Ca2+ concentrations in comparison to the nontransfected cells.13 However, in the absence of electrophysiological data, it would be premature to conclude that TRPM1 is a constitutively active channel. Alternatively, altered expression of the channels localizing on the internal stores such as the membranes of the ER may be an adaptive response or may offer a survival advantage, such as resistance to apoptosis. In that respect, the decrease in urothelial TRPV119 and prostatic TRPM812, 46 in intracellular stores in aggressive tumors probably reduces the Ca2+ release content and confers resistance to apoptosis.
17.3. TRP PROTEIN TRANSLATION FROM ALTERNATIVE START CODONS: AN UNKNOWN MECHANISM OF CANCER CELL-ENHANCED RESISTANCE AND SURVIVAL
There is not so much known so far about TRP protein translation from alternative start codons (ATGs) situated downstream of the first methionine codon of the predicted full-length sequence. These ATG codons are preceded by the Kozak consensus sequence (gcc)gccRccAUGG, where R is a purine (adenine or guanine), three bases upstream of the start codon (AUG), which is followed by another “G.”48 The Kozak consensus sequence occurs on eukaryotic mRNAs and plays a major role in the initiation of the translation process.49 In vivo, this site is often not matched exactly on different mRNAs, and the amount of protein synthesized from a given mRNA is dependent on the strength of the Kozak sequence. There are examples in vivo of each type of Kozak consensus, and they probably evolved as yet another mechanism of gene regulation. Lmx1b is an example of a gene with a weak Kozak consensus sequence.50 For initiation of translation from such a site, other features are required in the mRNA sequence in order for the ribosome to recognize the initiation codon. It seems likely that the above mechanisms may be engaged by the cancer cell to enhance its survival and resistance to apoptosis. It may also be an alternative mechanism to increase the cancer cell plasticity in the accelerated process of evolution and adaptation to the environment.
The cloning and expression pattern of bCCE 1Δ514, a 5′ truncated splice variant of the bovine TRPC4 (formally bCCE1, as supported by the evidence that it functioned as a capacitative calcium entry channel), provided the first example of native TRP protein expression using a downstream ATG codon,51 although it was shown to represent an alternatively spliced product of the TRPC4 gene that gives rise to an ~1.9-kb transcript rather than protein translation from an alternative ATG. It is interesting that in contrast to the six transmembrane segments predicted to be present in the full-length bTRPC4, bCCE 1Δ514 contains only three hydrophobic segments corresponding to transmembrane segments 5 and 6 and the putative pore-forming region in between. This membrane topology is reminiscent to that of the inward rectifier-type K+ channel (Kir) gene family, which also encodes proteins of less than 500 amino acids with two putative transmembrane spanning domains M1 and M2 and a hydrophobic segment in between contributing to ion conducting pore formation.52 Accordingly, the ion conducting properties of the bTRPC4 protein might still be preserved in bCCE 1Δ514, although with different regulation mechanisms arising from the lack of the greater part of the N-terminus of bTRPC4.
By utilizing recently developed full-length cDNA technologies, large-scale cDNA sequencing is carried out by several cDNA projects. Now full-length cDNA resources cover the major part of the protein-coding human genes. Comprehensive analyses of the collected full-length cDNA data reveal not only the complete sequences of thousands of novel gene transcripts but also novel alternatively spliced isoforms of hitherto identified genes. However, it is not as easy as expected to deduce their encoded amino acid sequences based solely on the full-length cDNA sequences. It is neither always the case that the longest open reading frame corresponds to the real protein coding region nor that the first ATG should be the translation initiation codon. Also, proteome-wide mass spectrometry analysis has shown that there is an unexpectedly large population of small proteins, encoded by so-called upstream open reading frames, within the cell (for review see Ref. 53). Figure 17.2 shows the methodological approach to study the expression of the TRPV6 channel from the cloned cDNA. The TRPV6 molecule contains several ATG codons preceded by the functional consensus Kozak sequences theoretically capable of producing different size proteins. As can be seen from the blots, at least five alternatively translated protein molecules may be produced. Site-directed mutagenesis should be used thereafter to prove that putative protein molecules result from proper alternative ATG codons. This method may also be used to verify the specificity of the antibody. Some other examples follow hereafter.
Sequence analysis of various human papillomavirus types associated with particular clinical outcomes has revealed that L1 protein sequences of the major cervical cancer-associated viruses generally possess the ability to encode a longer translation product whilst the non-cancer-causing viruses do not.54 Equally intriguing, the upstream initiation codon is always separated by 78 nucleotides from the initiation codon that produces L1 protein, which efficiently assembles into viral particles. The authors conclude that the longer L1 protein could play a role in the development of cervical carcinoma and that human papillomaviruses with the potential to cause cervical cancer may be identified by the presence of an in-frame ATG situated 78 nucleotides upstream.
A novel form of the E2F-3 protein, termed E2F-3B, has been identified.55 In contrast to the full-length E2F-3, which is expressed only at the G1/S boundary, E2F-3B is detected throughout the cell cycle with peak levels in G0 where it is associated with Rb. Transfection and in vitro translation experiments demonstrated that a protein identical to E2F-3B in size and isoelectric point is produced from the E2F-3 mRNA via the use of an alternative translational start site. Owing to this alternative initiation codon, the E2F-3B is missing 101 N-terminal amino acids relative to the full-length E2F-3. This region includes a moderately conserved sequence of unknown function that is present only in the growth-promoting E2F family members, including E2F-1, E2F-2, and full-length E2F-3. These observations make E2F-3B the first example of an E2F gene giving rise to two different protein species and also suggest that E2F-3 and E2F-3B may have opposing roles in the cell cycle control.55
Therefore, the alternative initiation of translation may represent an important evolutionary mechanism for a cancer cell to survive, to escape apoptosis, and to invade the body. TRP channels represent a potential “cancer target” because they are involved in all of them and therefore are subject to additional variability and regulation.
17.4. TRP PROTEIN POSTTRANSLATIONAL MODIFICATIONS: DIRECT OR INDIRECT MODULATION OF THE CHANNEL ACTIVITY
Protein posttranslational modifications are the chemical modifications of a protein after its translation. These modifications are largely used in the cell to regulate not only the protein activity but also its trafficking and stability at the plasma membrane in the case of the channel. Several types of posttranslational modifications exist, including the addition of functional groups such as acetylation, glycosylation, and phosphorylation; the addition of other proteins and peptides such as SUMOylation and ubiquitination; modifications involving changing the chemical nature of amino acids such as deamidation and deimination, as well as modifications involving structural changes such as disulfide bridges, proteolytic cleavage, etc. Modification and/or misregulation of posttranslational modifications may play a significant role in carcinogenesis and therefore should be considered in the case of TRP channels and cancer.
Nothing is known so far as to whether particular posttranslational modifications are explicitly used by the cancer cell for its needs. However, some data exist as to the posttranslational changes the channels are subjected to in cancer. In the case of Morris hepatoma H5123 cells, the cell-to-cell channel protein, connexin-43 (Cx43), is little expressed, and these cells lack gap junctional communication.56 The authors found that the inhibition of glycosylation by tunicamycin induced open channels in these cells. Although tunicamycin caused the formation of open channels, channels were not found aggregated into gap junctional plaques, as they are when they have been induced by elevation of intracellular cAMP. The results suggest that although Cx43 itself is not glycosylated, other glycosylated proteins influence Cx43 posttranslational modification and the formation of Cx43 cell-to-cell channels.56 Below, we will consider the two most frequent posttranslational modifications, namely, glycosylation and phosphorylation.
17.4.1. Glycosylatlon to Stabilize Channel Expression at the Plasma Membrane
N-linked glycosylation is considered to be important for channel trafficking to the plasma membrane. For instance, in human two-pore domain K+ channel TRESK subunits, one or two N-glycosylation consensus sites were identified.57 Using site-directed mutagenesis and Western immunoblotting, a single residue of both orthologues was found to be glycosylated upon heterologous expression. Two-electrode voltage-clamp recordings from Xenopus oocytes revealed that current amplitudes of N-glycosylation mutants were reduced by 80% as compared to the wild-type TRESK, so that their lower current amplitudes substantially result from inadequate, very low surface expression of the channel.57
Glycosylation is the covalent addition of sugar, or saccharide, moieties to a macromolecule via enzymatic action; glycation, in contrast, is its nonenzymatic counterpart. Mammalian cells can call upon a repertoire of nine distinct monosaccharides to enzymatically modify proteins and lipids.58 N-linked oligosaccharides serve many functions: they promote folding of glycoproteins; help target proteins reach the correct cellular compartments; contribute to protein–protein interactions and other ligand recognition processes; stabilize proteins against denaturation and proteases; increase protein solubility; facilitate proper orientation in membranes; confer structural rigidity; influence protein turnover; and modify the charge and isoelectric point of proteins.58 Among membrane-associated glycoproteins in epithelial cells, a role for glycosylation in intracellular trafficking has received the most attention. The plasma membrane is comprised of two functional compartments: the apical membrane compartment, facing the lumen, and the basolateral membrane compartment, generally apposed to a capillary. Targeting of cell surface proteins (e.g., ion channels) to the basolateral membrane is usually mediated via specific sorting motifs encoded within the amino acid sequence itself; generally, these motifs are in cytosolic regions of the protein.
Posttranslational modification of parts of the cardiac L-type Ca2+ channel by N-glycosylation is an important determinant for the binding of the dihydropyridine type of antagonists to Ca2+ channel α1 subunit, which itself is not glycosylated.59 The results suggest a participation of N-glycosylation in subunit assembling of the functional channel and/or its turnover. However, a possible effect of tunicamycin on the expression of the Ca2+ channel as an alternative mechanism cannot be excluded.
Cystic fibrosis transmembrane conductance regulator (CFTR) is a cAMP-regulated C1- channel. Malfunction of CFTR causes cystic fibrosis (CF).60 CFTR belongs to the ATP-binding cassette transporter superfamily, which includes P-glycoprotein (Pgp), the molecule that is responsible for multidrug resistance in cancer cells. It has been suggested that the membrane targeting and insertion of CFTR and Pgp may take the same pathway, i.e., the signal recognition particle (SRP) dependent pathway, but the membrane folding mechanism of these two proteins in microsomal membranes is probably different.60
The best studied example of TRP channel N-linked glycosylation, and the most striking example of its potential physiological impact, began with a gene called klotho. Klotho is highly expressed in the distal convoluted tubule of the kidney and in the hormone-secreting cells of the parathyroid gland;61 both are sites of active regulation of systemic Ca2+ balance. Like Klotho, the Ca2+ transporting channel TRPV5 is expressed in the distal convoluted tubule of the kidney where it is involved in Ca2+ resorption from the glomerular filtrate.62 Chang et al. found that Klotho cleaves an N-linked oligosaccharide from TRPV5, thereby trapping the channel in the apical plasma membrane.63 This deglycosylation of TRPV5 is accomplished by an N-terminal ectodomain of the membrane-associated Klotho enzyme that is itself cleaved and liberated into the urinary space, and into the plasma and cerebrospinal fluid.63 This model of regulation by compartmentalization is well suited to a channel such as TRPV5, which exhibits constitutive activity in heterologous expression systems. Whereas the channel may remain perpetually active, it is only physiologically active (that is, reabsorbing urinary Ca2+) when it is confined to the tubular apical membrane. Thus, the above regulation is crucial for the proper function of TRPV5. The very similar channel TRPV6, which possesses 75% of amino acid identity to TRPV5, has already been shown to be implicated in some cancers.21, 25
The alignment of TRPV5 with all members of the TRPC family was performed to uncover other analogous glycosylation motifs within the first extracellular loop. Of note, for this alignment, only the membrane-spanning regions were subjected to CLUSTAL analysis (http://www.ebi.ac.uk/clustalw/#).64 Most members of the TRPC family, in contrast, have eight hydrophobic regions. Vannier et al., investigating human TRPC3, surmised that the first hydrophobic region was not a transmembrane segment,65 and this is supported by the work of Dietrich et al.66 Dohke et al., analyzing the closely related TRPC1 channel, concluded that the third hydrophobic region may not be a true transmembrane segment.67 Accurate assignment of the first transmembrane segment is essential for investigating potential glycosylation motifs in the first extracellular loop. For our protein sequence alignment, we excluded the first hydrophobic domain from the TRPC family members and included only domains two through eight; this resulted in alignment of all of the known “e1” glycosylation motifs (i.e., TRPV5, TRPC3, and TRPC6) and uncovered a similar site in TRPC7. The previously unrecognized TRPC7 potential glycosylation site is evolutionarily conserved, and although there are no direct experimental data, migration of the heterologously expressed channel is consistent with glycosylation (e.g., see Ref. 68). In the case of TRPC6, the extensive glycosylation has also been shown.26, 69 The authors used peptide N-glycosidase F (PNGase) to treat the protein sample to deglycosylate the TRPC6 channel and to show that its real size corresponds to the predicted one of 106 kDa.
Glycosylation has been reported to influence trafficking and/or function of a variety of voltage-gated ion channels. This protein superfamily includes the focus of this review, the TRP channels, as well as the two-pore segment (TPC), voltage-gated sodium (NaV), voltage-gated calcium (CaV), hyperpolarization-activated cyclic nucleotide gated (HCN), cyclic nucleotide-gated (CNG), voltage-gated potassium (KV), two-pore potassium (K2p), calcium-activated potassium (KCa), and inwardly rectifying potassium (Kir) channel families.70 Of these, the architecture of the KV, HCN, CNG, and KCa channels most closely resembles that of the TRP channels, with six membrane-spanning domains and a pore-forming loop between helices five and six. Interestingly, the most abundant evidence for a functional role of glycosylation comes from this TRP-like subgroup. Two members—HCN2 and the Human Ether-a-go-go Related Gene (HERG) potassium channel—share N-linked glycosylation sites, adjacent to the pore-forming loop, that influence membrane trafficking and are potentially analogous to those of TRPV1 and TRPV4. In HERG channels (also known as KV11.1), mutation of this glycosylation site disrupts targeting.71, 72 A human mutation in this channel gene confers a heritable variant of the “long QT syndrome,” which is associated with potentially lethal cardiac dysrhythmias.73 In HCN2, a channel mutated for the putative glycosylation site similarly fails to traffic to the plasma membrane.74 TRPV4 appears to be unique among this group in that membrane trafficking is down-regulated rather than facilitated by glycosylation.75
At least a subset of TRP channel proteins undergo regulatory N-linked glycosylation. The function and/or subcellular localization of TRP channels are influenced by glycosylation, and other members of the TRP family share the motif. It is likely that N-linked glycosylation, and the dynamic regulation of this process, will play major roles in the function and targeting of a wide range of TRP and closely related ion channels.
17.4.2. Phosphorylation of TRP Channels to Directly Modulate Channel Activity
Protein phosphorylation and dephosphorylation are common, reversible, posttranslational modifications that can regulate the structure and function of ion channels. A particular phosphorylation/dephosphorylation state can modify channel activity and thus alter the electrophysiological properties of excitable and nonexcitable cells. A few well-known examples include protein kinase G (PKG) regulation of large conductance Ca2+-dependent K+ channels (BKCa) and the regulation of NMDA receptors by tyrosine phosphorylation. The BKCa channel is composed of four α-subunits that form the pore and a regulatory β-subunit; the channel is regulated by voltage in a Ca2+-dependent manner. PKG phosphorylates the α-subunit at Ser-1072 near the C-terminus, shifting the Ca2+ sensitivity of the channel and producing hyperpolarization.76
In the past few years, rapid development in the field of TRP channel research has demonstrated important roles for protein phosphorylation in the regulation of TRP channels, particularly for members of the TRPC, TRPV, TRPM, and TRPP subfamilies.
The TRPC subfamily contains seven members, which can be further divided into four subgroups: TRPC1, TRPC2, TRPC4,5, and TRPC3,6,7.77 TRPC1 needs to form heteromultimeric complexes with TRPC4 or TRPC5 for its proper trafficking to the plasma membrane in order to form functional channels.78 Zhang and Saffen provided the first evidence that TRPC6 activity is negatively regulated by PKC. They found that TRPC6, when overexpressed in CHO cells, was inhibited by PKC-activating phorbol 12-myristate 13-acetate (PMA).26 Recently, Trebak et al. identified Ser-712 in the TRPC3 amino acid sequences to be a specific PKC phosphorylation site.79 A point mutation at this site abolished the PKC phosphorylation on TRPC3 proteins and also markedly reduced the inhibitory effect of PKC activation on TRPC3-mediated Ca2+ influx. Another study by Zhu et al. found that PKC phosphorylates Thr-972 of mouse TRPC5, causing channel desensitization.80 TRPC6 and 7 are also desensitized by PKC.81 Channel desensitization is expected to cause an overall reduction in Ca2+ influx. However, several lines of evidence suggest that different mechanisms may govern the channel desensitization observed by Zhu et al.80
PKG is another kinase capable of inhibiting TRPC3 activity.82 Disruption of two consensus PKG phosphorylation sites, Thr-11 and Ser-263, markedly reduces the inhibitory effect of cGMP. These data indicate that PKG phosphorylates TRPC3 at Thr-11 and Ser-263 and, as a consequence, inactivates TRPC3.82 The inhibitory action of PKC and PKG on TRPC may represent important negative feedback mechanisms in the control of cytosolic Ca2+ levels, thereby influencing Ca2+-dependent processes in a variety of different cell types. In these negative feedback pathways, the activation of TRPC results in Ca2+ entry; a rise in cytosolic Ca2+, together with elevated diacylglycerol (DAG) levels, stimulates PKC activity, which feeds back to inactivate the TRPC channels.83
It has been known for a long time that tyrosine kinases are involved in the activation of capacitative Ca2+ influx, of which TRPC channels are among the major molecular candidates.84 Vazquez et al. found that inhibition of Src kinases by genistein and erbstatin abolished the receptor- and OAG (a DAG analog)-induced activation of TRPC3.85 In addition, OAG failed to activate TRPC3 in cells that were either Src-deficient or expressed a dominant-negative mutant of Src, and furthermore, OAG activation of TRPC3 was restored after the cells were transfected with a Src-expressing construct. These results indicate an obligatory requirement for Src kinase in DAG-induced activation of TRPC3. Note that Src may not directly act on TRPC3. Instead, a concerted role for both DAG and Src seems to be necessary for TRPC3 activation, perhaps through a mechanism involving Src-dependent phosphorylation and/or recruitment of a yet unknown accessory/regulatory protein within the vicinity of TRPC3.85
CaM-kinase II can activate TRPC6. In patch-clamp studies, Shi et al. found that TRPC6, expressed in HEK293 cells, was activated by extracellular Ca2+, which could be prevented by either the organic CaM-kinase II inhibitor KN-62 or a CaM-kinase II-specific inhibitory peptide.81 These results suggest that CaM-kinase II-mediated phosphorylation is an obligatory step for TRPC6 channel activation.
The TRPV subfamily contains six members (TRPV1–6). TRPV1–4 channels are temperature-sensitive. TRPV5 and TRPV6 are only distinctly related to TRPV1–4 with a 30–40% sequence homology. TRPV5 and TRPV6 have high selectivity for Ca2+ over Na+ (PCa/PNa = 100/1), are mainly expressed in Ca2+-transporting epithelia, and are assumed to play an important role in Ca2+ (re)absorption by the kidney and intestine.86
Multiple kinases are known to regulate TRPV1. PKC phosphorylates Ser-502 and Ser-800 in rat TRPV1 and, as a result, either potentiates or sensitizes the responses of this channel to capsaisin, heat, and anandamide.87, 88 TRPV2 and TRPV4 are two temperature-sensitive channels with activation thresholds of ≥53 and ≥25°–27°C, respectively.89 TRPV2 is a substrate of PKA. In mast cells, PKA interacts with TRPV2 through a PKA-binding protein named ACBD3.90 PKA phosphorylation enhances TRPV2-mediated Ca2+ influx in response to heat.90 On the other hand, the activity of TRPV4 is stimulated by PKC,91 although it is unclear whether this stimulation is due to direct PKC phosphorylation on the TRPV4 proteins.
TRPV5 is activated by serum and glucocorticoid-inducible kinase, SGK1. This stimulatory effect is due to enhanced TRPV5 abundance in the plasma membrane, requiring the presence of the scaffold protein, NHERF2.92 On the other hand, the activity of TRPV6 can be regulated by calmodulin and PKC. Binding of Ca2+-dependent calmodulin to TRPV6 inactivates the channel, which is countered by PKC-mediated phosphorylation of TRPV6.93 Thus, by altering the inactivation behavior of TRPV6, PKC-mediated phosphorylation acts as a switch to regulate the amount of Ca2+ influx through TRPV6.94 TRPV6 can also be activated by the Src tyrosine kinase, which is counterbalanced by the protein tyrosine phosphatase 1B.94 Taken together, TRPV6 activity is closely controlled by both the calmodulin/PKC system and the tyrosine kinase/phosphatase system.93, 94
The TRPM subfamily consists of eight members (TRPM1–8). TRPM4 is a voltage-dependent, Ca2+-impermeable cation channel. Opening of this channel depolarizes the membrane. While the channel is activated by intracellular Ca2+, the currents decay rapidly due to decreased sensitivity of the channel to Ca2+. PMA, an activator of PKC, increases the activity of TRPM495 by enhancing the sensitivity of TRPM4 to Ca2+.96
The physiological significance of the regulation of TRP channels by phosphorylation is fascinating. TRP channels play diverse functional roles, including thermal sensation, nociception, mechanosensing, growth cone guidance, inflammatory responses, membrane potential control, and Mg2+ homeostasis. Various kinases and phosphatases may regulate the activities of different TRP channel isoforms, providing enormous control on diverse cellular processes.
17.5. CONCLUSIONS
The progression of cells from a normal, differentiated state to a tumorigenic, metastatic state involves the accumulation of mutations in multiple key signaling proteins, encoded by oncogenes and tumor suppressor genes, together with the evolution and clonal selection of more aggressive cell phenotypes. These events are associated with changes in the expression of numerous other proteins. To date, most changes involving TRP proteins do not involve mutations in the TRP genes but rather increased or decreased expression levels of the wild-type TRP proteins, depending on the stage of the cancer. On the other hand, several common tuning pathways lead to this divergence in expression. In this respect, TRP channels may be regulated at different levels: (1) transcriptional, (2) translational, (3) trafficking to the plasma membrane, or (4) direct channel modulation on the plasma membrane. Modulation of TRP expression/activity on one of these levels affects intracellular Ca2+ signaling and, consequently, the processes involved in carcinogenesis, such as proliferation, apoptosis, and migration.
REFERENCES
- 1.
- Prevarskaya N, Skryma R, Shuba Y. Ca2+ homeostasis in apoptotic resistance of prostate cancer cells. Biochemical and Biophysical Research Communications. 2004;322:1326–1335. [PubMed: 15336979]
- 2.
- Rosado J.A, Redondo P.C, Pariente J.A, Salido G.M. Calcium signalling and tumorigenesis. Cancer Therapy. 2004;2:263–270.
- 3.
- Weinberg R.A. In The Biology of Cancer. New York: Garland Science; 2006. Multi-step tumorigenesis in the biology of cancer; pp. 399–461.
- 4.
- Roderick H.L, Cook S.J. Ca2+ signalling checkpoints in cancer: remodelling Ca2+ for cancer cell proliferation and survival. Nature Reviews. Cancer. 2008;8:361–375. [PubMed: 18432251]
- 5.
- Cook S.J, Lockyer P.J. Recent advances in Ca2+-dependent Ras regulation and cell proliferation. Cell Calcium. 2006;39:101–112. [PubMed: 16343616]
- 6.
- Whitfield J.F. Calcium signals and cancer. Critical Reviews in Oncogenesis. 1992;3:55–90. [PubMed: 1550862]
- 7.
- Berridge M.J, Bootman M.D, Roderick H.L. Calcium signalling: dynamics, homeostasis and remodelling. Nature Reviews. Molecular Cell Biology. 2003;4:517–529. [PubMed: 12838335]
- 8.
- Rizzuto R, Pinton P, Ferrari D. et al. Calcium and apoptosis: facts and hypotheses. Oncogene. 2003;22:8619–8627. [PubMed: 14634623]
- 9.
- Monteith G.R, McAndrew D, Faddy H.M, Roberts-Thomson S.J. Calcium and cancer: targeting Ca2+ transport. Nature Reviews. Cancer. 2007;7:519–530. [PubMed: 17585332]
- 10.
- Nilius B, Owsianik G, Voets T, Peters J.A. Transient receptor potential cation channels in disease. Physiological Reviews. 2007;87:165–217. [PubMed: 17237345]
- 11.
- Bidaux G, Flourakis M, Thebault S. et al. Prostate cell differentiation status determines transient receptor potential melastatin member 8 channel subcellular localization and function. Journal of Clinical Investigation. 2007;117:1647–1657. [PMC free article: PMC1866249] [PubMed: 17510704]
- 12.
- Thebault S, Lemonnier L, Bidaux G. et al. Novel role of cold/menthol-sensitive transient receptor potential melastatine family member 8 (TRPM8) in the activation of store-operated channels in LNCaP human prostate cancer epithelial cells. Journal of Biological Chemistry. 2005;280:39423–39435. [PubMed: 16174775]
- 13.
- Xu X.Z, Moebius F, Gill D.L, Montell C. Regulation of melastatin, a TRP-related protein, through interaction with a cytoplasmic isoform. Proceedings of the National Academy of Science of the United States of America. 2001;98:10692–10697. [PMC free article: PMC58528] [PubMed: 11535825]
- 14.
- Montell C, Birnbaumer L, Flockerzi V. et al. A unified nomenclature for the superfamily of TRP cation channels. Molecular Cell. 2002;9:229–231. [PubMed: 11864597]
- 15.
- Thebault S, Flourakis M, Vanoverberghe K. et al. Differential role of transient receptor potential channels in Ca2+ entry and proliferation of prostate cancer epithelial cells. Cancer Research. 2006;66:2038–2047. [PubMed: 16489003]
- 16.
- Tsavaler L, Shapero M.H, Morkowski S. TRP-p8, a novel prostate-specific gene, is up-regulated in prostate cancer and other malignancies and shares high homology with transient receptor potential calcium channel proteins. Cancer Research. 2001;61:3760–3769. [PubMed: 11325849]
- 17.
- Wissenbach U, Niemeyer B.A, Fixemer T. et al. Expression of CaT-like, a novel calcium-selective channel, correlates with the malignancy of prostate cancer. Journal of Biological Chemistry. 2001;276:19461–19468. [PubMed: 11278579]
- 18.
- Duncan L.M, Deeds J, Hunter J. et al. Down-regulation of the novel gene melastatin correlates with potential for melanoma metastasis. Cancer Research. 1998;58:1515–1520. [PubMed: 9537257]
- 19.
- Lazzeri M, Vannucchi M.G, Spinelli M.E. et al. Transient receptor potential vanilloid type 1 (TRPV1) expression changes from normal urothelium to transitional cell carcinoma of human bladder. European Urology. 2005;48:691–698. [PubMed: 15992990]
- 20.
- Caprodossi S, Lucciarini R, Amantini C. et al. Transient receptor potential vanilloid type 2 (TRPV2) expression in normal urothelium and in urothelial carcinoma of human bladder: correlation with the pathologic stage. European Urology. 2007;72:440–448. [PubMed: 17977643]
- 21.
- Fixemer T, Wissenbach U, Flockerzi V, Bonkhoff H. Expression of the Ca2+-selective cation channel TRPV6 in human prostate cancer: a novel prognostic marker for tumor progression. Oncogene. 2003;22:7858–7861. [PubMed: 14586412]
- 22.
- Prawitt D, Enklaar T, Klemm G. et al. Identification and characterization of MTR1, a novel gene with homology to melastatin (MLSN1) and the TRP gene family located in the BWS-WT2 critical region on chromosome 11p15.5 and showing allele-specific expression. Human Molecular Genetics. 2000;9:203–216. [PubMed: 10607831]
- 23.
- Fang D, Setaluri V. Expression and up-regulation of alternatively spliced transcripts of melastatin, a melanoma metastasis-related gene, in human melanoma cells. Biochemical and Biophysical Research Communications. 2000;279:53–61. [PubMed: 11112417]
- 24.
- Fuessel S, Sickert D, Meye A. Multiple tumor marker analyses (PSA, hK2, PSCA, TRP-p8) in primary prostate cancers using quantitative RT-PCR. International Journal of Oncology. 2003;23:221–228. [PubMed: 12792797]
- 25.
- Lehen’kyi V, Flourakis M, Skrym R, Prevarskaya N. TRPV6 channel controls prostate cancer cell proliferation via Ca2+/NFAT-dependent pathways. Oncogene. 2007;26:7380–7385. [PubMed: 17533368]
- 26.
- Zhang L, Saffen D. Muscarinic acetylcholine receptor regulation of TRP6 Ca2+ channel isoforms. Molecular structures and functional characterization. Journal of Biological Chemistry. 2001;276:13331–13339. [PubMed: 11278449]
- 27.
- Monet M, Gkika D, Lehen’kyi V. Lysophospholipids stimulate prostate cancer cell migration via TRPV2 channel activation. Biochimica et Biophysica Acta. 2009;1793:528–539. [PubMed: 19321128]
- 28.
- Bolanz K.A, Hediger M.A, Landowski C.P. The role of TRPV6 in breast carcinogenesis. Molecular Cancer Therapeutics. 2008;7:271–279. [PubMed: 18245667]
- 29.
- Zhuang L, Peng J.B, Tou L. et al. Calcium-selective ion channel, CaT1, is apically localized in gastrointestinal tract epithelia and is aberrantly expressed in human malignancies. Lab Investigations. 2002;82:1755–1764. [PubMed: 12480925]
- 30.
- Peng J.B, Zhuang L, Berger U.V. et al. CaT1 expression correlates with tumor grade in prostate cancer. Biochemical and Biophysical Research Communications. 2001;282:729–734. [PubMed: 11401523]
- 31.
- Peng J.B, Chen X.Z, Berger U.V. et al. Molecular cloning and characterization of a channel-like transporter mediating intestinal calcium absorption. Journal of Biological Chemistry. 1999;274:22739–22746. [PubMed: 10428857]
- 32.
- Ligresti A, Moriello A.S, Starowicz K. et al. Antitumour activity of plant cannabinoids with emphasis on the effect of cannabidiol on human breast carcinoma. Journal of Pharmacology and Experimental Therapeutics. 2006;318:1375–1387. [PubMed: 16728591]
- 33.
- Izzo A.A, Coutts A.A. Cannabinoids and the digestive tract. Handbook of Experimental Pharmacology. 2005;168:573–598. [PubMed: 16596788]
- 34.
- Wertz I.E, Dixit V.M. Characterization of calcium release-activated apoptosis of LNCaP prostate cancer cells. Journal of Biological Chemistry. 2000;275:11470–11478. [PubMed: 10753965]
- 35.
- Costa-Pereira A.P, Cotter T.G. Molecular and cellular biology of prostate cancer-the role of apoptosis as a target for therapy. Prostate Cancer Prostatic Diseases. 1999;2:126–139. [PubMed: 12496822]
- 36.
- Beck B, Bidaux G, Bavencoffe A. et al. Prospects for prostate cancer imaging and therapy using high-affinity TRPM8 activators. Cell Calcium. 2007;41:285–294. [PubMed: 16949669]
- 37.
- Clapham D.E, Runnels L.W, Strubing C. The TRP ion channel family. Nature Reviews. Neuroscience. 2001;2:387–396. [PubMed: 11389472]
- 38.
- Stamm S, Ben-Ari I, Rafalska Y. et al. Function of alternative splicing. Gene. 2005;344:1–20. [PubMed: 15656968]
- 39.
- Nagasawa M, Nakagawa Y, Tanaka S, Kojima I. Chemotactic peptide fMetLeuPhe induces translocation of the TRPV2 channel in macrophages. Journal of Cell Physiology. 2007;210:692–702. [PubMed: 17154364]
- 40.
- Wang C, Hu H.Z, Colton C.K, Wood J.D, Zhu M.X. An alternative splicing product of the murine TRPV1 gene dominant negatively modulates the activity of TRPV1 channels. Journal of Biological Chemistry. 2004;279:37423–37430. [PubMed: 15234965]
- 41.
- Zhiqi S, Soltani M.H, Bhat K.M. et al. Human melastatin 1 (TRPM1) is regulated by MITF and produces multiple polypeptide isoforms in melanocytes and melanoma. Melanoma Research. 2004;14:509–516. [PubMed: 15577322]
- 42.
- Sabnis A.S, Shadid M, Yost G.S, Reilly C.A. Human lung epithelial cells express a functional cold-sensing TRPM8 variant. American Journal of Respiratory Cell and Molecular Biology. 2008;39:466–474. [PMC free article: PMC2551706] [PubMed: 18458237]
- 43.
- Lis A, Wissenbach U, Philipp S.E. Transcriptional regulation and processing increase the functional variability of TRPM channels. Naunyn-Schmiedeberg's Archives of Pharmacology. 2005;371:315–324. [PubMed: 15856355]
- 44.
- Mahieu F, Owsianik G, Verbert L. et al. TRPM8-independent menthol-induced Ca2+ release from endoplasmic reticulum and Golgi. Journal of Biological Chemistry. 2007;282:3325–3336. [PubMed: 17142461]
- 45.
- Kim S.H, Nam J.H, Park E.J. et al. Menthol regulates TRPM8-independent processes in PC-3 prostate cancer cells. Biochimica et Biophysica Acta. 2008;1792:33–38. [PubMed: 18955132]
- 46.
- Zhang L, Barritt G.J. Evidence that TRPM8 is an androgen-dependent Ca2+ channel required for the survival of prostate cancer cells. Cancer Research. 2004;64:8365–8373. [PubMed: 15548706]
- 47.
- Harteneck C, Plant T.D, Schultz G. From worm to man: three subfamilies of TRP channels. Trends in Neurosciences. 2000;23:159–166. [PubMed: 10717675]
- 48.
- Kozak M. An analysis of 5'-noncoding sequences from 699 vertebrate messenger RNAs. Nucleic Acids Research. 1987;15:8125–8148. [PMC free article: PMC306349] [PubMed: 3313277]
- 49.
- De Angioletti M, Lacerra G, Sabato V, Carestia C. Beta+45 G --> C: a novel silent beta-thalassaemia mutation, the first in the Kozak sequence. British Journal of Haematology. 2004;124:224–231. [PubMed: 14687034]
- 50.
- Dunston J.A, Hamlington J.D, Zaveri J. et al. The human LMX1B gene: transcription unit, promoter, and pathogenic mutations. Genomics. 2004;84:565–576. [PubMed: 15498463]
- 51.
- Freichel M, Wissenbach U, Philipp S, Flockerzi V. Alternative splicing and tissue specific expression of the 5' truncated bCCE 1 variant bCCE 1delta514. FEBS Letters. 1998;422:354–358. [PubMed: 9498815]
- 52.
- Doupnik C.A, Davidson N, Lester H.A. The inward rectifier potassium channel family. Current Opinion in Neurobiology. 1995;5:268–277. [PubMed: 7580148]
- 53.
- Suzuki Y, Sugano S. Transcriptome analyses of human genes and applications for proteome analyses. Current Protein and Peptide Science. 2006;7:147–163. [PubMed: 16611140]
- 54.
- Webb E, Cox J, Edwards S. Cervical cancer-causing human papillomaviruses have an alternative initiation site for the L1 protein. Virus Genes. 2005;30:31–35. [PubMed: 15744560]
- 55.
- He Y, Armanious M.K, Thomas M.J, Cress W.D. Identification of E2F-3B, an alternative form of E2F-3 lacking a conserved N-terminal region. Oncogene. 2000;19:3422–3433. [PubMed: 10918599]
- 56.
- Wang Y, Mehta P.P, Rose B. Inhibition of glycosylation induces formation of open connexin-43 cell-to-cell channels and phosphorylation and triton X-100 insolubility of connexin-43. Journal of Biological Chemistry. 1995;270:26581–26585. [PubMed: 7592880]
- 57.
- Egenberger B, Polleichtner G, Wischmeyer E, Döring F. N-linked glycosylation determines cell surface expression of two-pore-domain K+ channel TRESK. Biochemical and Biophysical Research Communications. 2009;391:1262–1267. [PubMed: 20006580]
- 58.
- Ohtsubo K, Marth J.D. Glycosylation in cellular mechanisms of health and disease. Cell. 2006;126:855–867. [PubMed: 16959566]
- 59.
- Henning U, Wolf W.P, Holtzhauer M. Influence of glycosylation inhibitors on dihydropyridine binding to cardiac cells. Molecular Cell Biochemistry. 1996;(160-161):41–46. [PubMed: 8901454]
- 60.
- Chen M, Zhang J.T. Membrane insertion, processing, and topology of cystic fibrosis transmembrane conductance regulator (CFTR) in microsomal membranes. Molecular Membrane Biology. 1996;13:33–40. [PubMed: 9147660]
- 61.
- Kuro-o M, Matsumura Y, Aizawa H. et al. Mutation of the mouse klotho gene leads to a syndrome resembling ageing. Nature. 1997;390:45–51. [PubMed: 9363890]
- 62.
- Hoenderop J.G, van der Kemp A.W., Harto A. et al. Molecular identification of the apical Ca2+ channel in 1, 25-dihydroxyvitamin D3-responsive epithelia. Journal of Biological Chemistry. 1999;274:8375–8378. [PubMed: 10085067]
- 63.
- Chang Q, Hoefs S, van der Kemp A.W, Topala C.N, Bindels R.J, Hoenderop J.G. The beta-glucuronidase klotho hydrolyzes and activates the TRPV5 channel. Science. 2005;310:490–493. [PubMed: 16239475]
- 64.
- Thompson J.D, Higgins D.G, Gibson T.J, Clustal W. Improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice. Nucleic Acids Research. 1994;22:4673–4680. [PMC free article: PMC308517] [PubMed: 7984417]
- 65.
- Vannier B, Zhu M.X, Brown D, Birnbaumer L. The membrane topology of human transient receptor potential 3 as inferred from glycosylation-scanning mutagenesis and epitope immunocytochemistry. Journal of Biological Chemistry. 1998;273:8675–8679. [PubMed: 9535843]
- 66.
- Dietrich A, Mederos y Schnitzler M, Emmel J, Kalwa H, Hofmann T, Gudermann T. N-linked protein glycosylation is a major determinant for basal TRPC3 and TRPC6 channel activity. Journal of Biological Chemistry. 2003;278:47842–47852. [PubMed: 12970363]
- 67.
- Dohke Y, Oh Y.S, Ambudkar I.S, Turner R.J. Biogenesis and topology of the transient receptor potential Ca2+ channel TRPC1. Journal of Biological Chemistry. 2004;279:12242–12248. [PubMed: 14707123]
- 68.
- Maruyama Y, Nakanishi Y, Walsh E.J, Wilson D.P, Welsh D.G, Cole W.C. Heteromultimeric TRPC6-TRPC7 channels contribute to arginine vasopressin-induced cation current of A7r5 vascular smooth muscle cells. Circulation Research. 2006;98:1520–1527. [PubMed: 16690880]
- 69.
- 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. [PubMed: 18506892]
- 70.
- Yu F.H, Catterall W.A. The VGL-chanome: a protein superfamily specialized for electrical signaling and ionic homeostasis. Science STKE. 2004;2004:re15. [PubMed: 15467096]
- 71.
- Petrecca K, Atanasiu R, Akhavan A, Shrier A. N-linked glycosylation sites determine HERG channel surface membrane expression. Journal of Physiology. 1999;515(Part 1):41–48. [PMC free article: PMC2269130] [PubMed: 9925876]
- 72.
- Gong Q, Anderson C.L, January C.T, Zhou Z. Role of glycosylation in cell surface expression and stability of HERG potassium channels. American Journal of Physiology. Heart and Circulatory Physiology. 2002;283:H77–H84. [PubMed: 12063277]
- 73.
- Satler C.A, Vesely M.R, Duggal P, Ginsburg G.S, Beggs A.H. Multiple different missense mutations in the pore region of HERG in patients with long QT syndrome. Human Genetics. 1998;102:265–272. [PubMed: 9544837]
- 74.
- Much B, Wahl-Schott C, Zong X. et al. Role of subunit heteromerization and N-linked glycosylation in the formation of functional hyperpolarization-activated cyclic nucleotide-gated channels. Journal of Biological Chemistry. 2003;278:43781–43786. [PubMed: 12928435]
- 75.
- Xu H, Fu Y, Tian W, Cohen D.M. Glycosylation of the osmoresponsive transient receptor potential channel TRPV4 on Asn-651 influences membrane trafficking. American Journal of Physiology. Renal Physiology. 2006;290:F1103–F1109. [PubMed: 16368742]
- 76.
- Fukao M, Mason H.S, Britton F.C, Kenyon J.L, Horowitz B, Keef K.D. Cyclic GMP-dependent protein kinase activates cloned BKCa channels expressed in mammalian cells by direct phosphorylation at serine 1072. Journal of Biological Chemistry. 1999;274:10927–10935. [PubMed: 10196172]
- 77.
- Minke B, Cook B. TRP channel proteins and signal transduction. Physiological Reviews. 2002;82:429–472. [PubMed: 11917094]
- 78.
- Strubing C, Krapivinsky G, Krapivinsky L, Clapham D.E. TRPC1 and TRPC5 form a novel cation channel in mammalian brain. Neuron. 2001;29:645–655. [PubMed: 11301024]
- 79.
- Trebak M, Hempel N, Wedel B.J, Smyth J.T, Bird G.S.J, Putney J.W Jr. Negative regulation of TRPC3 channels by protein kinase C-mediated phosphorylation of serine 712. Molecular Pharmacology. 2005;67:558–563. [PubMed: 15533987]
- 80.
- Zhu M.H, Chae M, Kim H.J. et al. Desensitization of canonical transient receptor potential channel 5 (TRPC5) by protein kinase C. American Journal of Physiology. 2005;289:C591–C600. [PubMed: 15843439]
- 81.
- Shi J, Mori E, Mori Y. et al. Multiple regulation by calcium of murine homologues of transient receptor potential proteins TRPC6 and TRPC7 expressed in HEK293 cells. Journal of Physiology. 2004;561:415–432. [PMC free article: PMC1665365] [PubMed: 15579537]
- 82.
- Kwan H.Y, Huang Y, Yao X. Regulation of canonical transient receptor potential isoform 3 (TRPC3) channel by protein kinase G. Regulation of canonical transient receptor potential isoform 3 (TRPC3) channel by protein kinase G. Proceedings of the National Academy of Sciences of the United States of America. 2004;101:2625–2630. [PMC free article: PMC357000] [PubMed: 14983059]
- 83.
- Venkatachalam K, Zheng F, Gill D.L. Regulation of canonical transient receptor potential channel function by diacylglycerol and protein kinase C. Journal of Biological Chemistry. 2003;278:29031–29040. [PubMed: 12721302]
- 84.
- Parekh A.B, Putney J.W Jr. Store depletion and calcium influx. Physiological Reviews. 2005;85:757–810. [PubMed: 15788710]
- 85.
- Vazquez G, Wedel B.J, Kawasaki B.T, Bird G.S, Putney J.W Jr. Obligatory role of Src kinase in the signaling mechanism for TRPC3 cation channels. Journal of Biological Chemistry. 2004;279:40521–40528. [PubMed: 15271991]
- 86.
- Montell C. The TRP superfamily of cation channels. The TRP superfamily of cation channels. Science STKE. 2005;272:re3. [PubMed: 15728426]
- 87.
- Vellani V, Mappleback S, Moriondo A, Davis J.B, McNaughton P.A. Protein kinase C activation potentiates gating of the vanilloid receptor VR1 by capsaicin, protons, heat and anandamide. Journal of Physiology. 2001;534:813–825. [PMC free article: PMC2278732] [PubMed: 11483711]
- 88.
- Premkumar L.S, Ahern G.P. Induction of vanilloid receptor channel activity by protein kinase C. Nature. 2000;408:985–990. [PubMed: 11140687]
- 89.
- Tominaga M, Caterina M.J. Thermosensation and pain. Journal of Neurobiology. 2004;61:3–12. [PubMed: 15362149]
- 90.
- Stokes A.J, Shimoda L.M, Koblan-Huberson M, Adra C.N, Turner H. A TRPV2-PKA signalling module for transduction of physical stimuli in mast cells. Journal of Experimental Medicine. 2004;200:137–147. [PMC free article: PMC2212017] [PubMed: 15249591]
- 91.
- Xu F, Satoh E, Iijima T. Protein kinase C-mediated Ca2+ entry in HEK 293 cells transiently expressing human TRPV4. British Journal of Pharmacology. 2003;140:413–421. [PMC free article: PMC1574039] [PubMed: 12970074]
- 92.
- Embark H.M, Setiawan I, Poppendieck S. et al. Regulation of the epithelial Ca2+ channel TRPV5 by the NHE regulating factor NHERF2 and the serum and glucocorticoid inducible kinase isoforms SGK1 and SGK3 expressed in Xenopus oocytes. Cellular Physiology and Biochemistry. 2004;14:203–212. [PubMed: 15319523]
- 93.
- Niemeyer B.A, Bergs C, Wissenbach U, Flockerzi V, Trost C. Competitive regulation of CaT-like-mediated Ca2+ entry by protein kinase C and calmodulin. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:3600–3605. [PMC free article: PMC30699] [PubMed: 11248124]
- 94.
- Sternfeld L, Krause E, Schmid A. Tyrosine phosphatase PTP1B interacts with TRPV6 in vivo and plays a role in TRPV6-mediated calcium influx in HEK293 cells. Cellular Signalling. 2005;17:951–960. [PubMed: 15894168]
- 95.
- Guinamard R, Chatelier A, Lenfant J, Bios P. Activation of the Ca2+-activated nonselective cation channel by diacylglycerol analogues in rat cardiomyocytes. Journal of Cardiovascular Electrophysiology. 2004;15:342–348. [PubMed: 15030426]
- 96.
- Nilius B, Prenen J, Tang J. et al. Regulation of the Ca2+ sensitivity of the nonselective cation channel TRPM4. Journal of Biological Chemistry. 2005;280:6423–6433. [PubMed: 15590641]
- INTRODUCTION
- ROLE OF TRP CHANNEL ISOFORM EXPRESSION AND FUNCTION IN THE STUDY OF THE CANCER INITIATION AND PROGRESSION
- TRP PROTEIN TRANSLATION FROM ALTERNATIVE START CODONS: AN UNKNOWN MECHANISM OF CANCER CELL-ENHANCED RESISTANCE AND SURVIVAL
- TRP PROTEIN POSTTRANSLATIONAL MODIFICATIONS: DIRECT OR INDIRECT MODULATION OF THE CHANNEL ACTIVITY
- CONCLUSIONS
- REFERENCES
- Study of TRP Channels in Cancer Cells - TRP ChannelsStudy of TRP Channels in Cancer Cells - TRP Channels
- PREDICTED: Homo sapiens phosphatidylinositol glycan anchor biosynthesis class G ...PREDICTED: Homo sapiens phosphatidylinositol glycan anchor biosynthesis class G (EMM blood group) (PIGG), transcript variant X2, misc_RNAgi|2217351091|ref|XR_924965.4|Nucleotide
- Proteomic Analysis of TRPC Channels - TRP ChannelsProteomic Analysis of TRPC Channels - TRP Channels
- Protomelas spilopterus bio-material Malawi 1715 distal-less homeobox 2 (dlx2) ge...Protomelas spilopterus bio-material Malawi 1715 distal-less homeobox 2 (dlx2) gene, partial cdsgi|298353689|gb|GU936530.1|Nucleotide
- protein SDA1 homolog isoform X1 [Homo sapiens]protein SDA1 homolog isoform X1 [Homo sapiens]gi|2462597922|ref|XP_054206343.1|Protein
Your browsing activity is empty.
Activity recording is turned off.
See more...