Abstract
Reactive oxygen species (ROS) and reactive nitrogen species (RNS) are important mediators of signaling pathways in the endothelium. Specific members of the transient receptor potential (TRP) superfamily of cation channels act as important Ca2+ influx pathways in endothelial cells, and are involved in endothelium-dependent vasodilation, regulation of barrier permeability, and angiogenesis. ROS and RNS can modulate the activity of certain TRP channels mainly by modifying specific cysteine residues or by stimulating the production of second messengers. In this review we highlight the recent literature describing in redox regulation of TRP channel activity in endothelial cells as well as the physiological importance of these pathways and implication for cardiovascular diseases.
Keywords: redox signaling, TRP channels, endothelial cells, Ca2+ influx
Introduction
Endothelial cells that collectively form the vascular endothelium are actively involved in the regulation of vascular tone, the exchange of nutrients between tissue and blood, and the formation of semi-permeable barriers to limit blood-borne pathogens and toxic substances from accessing sensitive tissues, such as the brain parenchyma. Ca2+ ions are ubiquitous second messengers that act as a unifying mediator in maintenance of endothelial function. Increases in intracellular [Ca2+] induce endothelium-dependent vasodilation1, regulate endothelial cell permeability2 and promotes the formation and maintenance of the blood-brain barrier3. Ca2+ released from intracellular stores following the opening of inositol-triphosphate receptors or Ca2+ influx through Ca2+-permeable ion channels located in the plasma membrane increase global or localized intracellular endothelial [Ca2+] to stimulate Ca2+-dependent signaling pathways. Endothelial cells are non-excitable cells (i.e., lacking voltage-gated ion channels) and the activation of Ca2+-permeable channels occurs by chemical and physical stimuli, such as membrane stretch caused by intraluminal pressure and shear stress4-6, or by soluble substances produced by autocrine or paracrine mechanisms. This review is focused on the control of endothelial cell Ca2+ influx by reduction-oxidation reactions (redox) regulation, defined here as reactions in which the oxidation states of atoms are changed. In biological systems, redox regulation is often mediated by reactive oxygen species (ROS) and reactive nitrogen species (RNS).
Redox Signaling
ROS and RNS are small, inorganic and highly reactive compounds with short half-lives in biological systems. Due to their strongly electrophilic nature, ROS and RNS rapidly react with and modify organic molecules, such as nucleic acids, lipids and proteins. Under physiological conditions, ROS and RNS take part in numerous signaling pathways. Tissue oxidative stress occurs when generation of these agents outstrips removal and detoxification, leading to cellular damage and, in the vasculature, impaired endothelial function7-9. ROS and RNS are constantly generated by mitochondrial oxidative metabolism10, activity of nicotinamide adenine dinucleotide phosphate (NADPH) oxidases (NOX)11 and coupled or uncoupled activity of other oxidative enzymes, most predominately endothelial nitric oxide synthase (eNOS)12 (Figure 1). NOX enzymes hydrolyze NADPH into NADP+ and transfer an electron to molecular oxygen, generating superoxide (O2−)13, which can in turn be dismutated to H2O2 and, in the presence of iron, converted to the highly reactive hydroxyl radical (OH−) by the Fenton reaction. OH− reacts strongly with proteins, nucleic acids and lipids, the latter generating lipid peroxides compounds14 which can act as signaling molecules. O2− is also produced by the ubiquinone in the interface between complexes I, II and III within the electron transport chain pathway of the mitochondria15, which can then leak to the cytoplasm via voltage-dependent anion channels located in the outer mitochondrial membrane16. Under physiological conditions, eNOS generates nitric oxide (NO) that can react with amino acids containing thiol moieties, such as cysteine, leading to S-nitrosylation17. However, under pathological conditions or when the intracellular pool of the co-factor tetrahydrobiopterin is limited, eNOS uses L-arginine to produce O2−18, a state known as uncoupled NOS.
Figure 1. Major pathways generating ROS and RNS in endothelial cells.
A) The enzyme NADPH oxidase (left) catalyzes the transfer of an electron from NADPH to molecular oxygen, generating superoxide (O2−) in the extracellular environment. The enzyme endothelial nitric oxide synthase (eNOS, middle) generates nitric oxide in the presence of the co-factor tetrahydrobiopterin (BH4, coupled eNOS) or O2− in the absence of BH4 (uncoupled eNOS). O2− is constantly produced in the mitochondria (right) by members of the oxidative phosphorylation complexes in the inner mitochondrial membrane. O2− located in the intermembrane space can then leak to the cytoplasm through voltage-dependent anion channels immersed in the outer mitochondrial membrane. B) Ubiquinone in the mitochondrial respiratory chain can be reduced by electron leak from Complex I or II, generating ubiquinol (UQ), and transports electrons to Complex III. Electron leak from ubiquinol and Complex III can interact with O2 present in the intermembrane space and generate O2− that flows to the cytoplasm. Most of the O2− generated in the mitochondrial matrix is inactivated by antioxidant defenses.
ROS and RNS act on molecular entities that serve as oxidative sensors to stimulate downstream signaling pathways. Proteins containing exposed cysteine residues are the best-characterized ROS and RNS sensors. Cysteine residues are particularly susceptible to redox reactions due to the presence of sulfhydryl radicals that can be converted to thiolate (R-S−), a highly nucleophilic radical, at near physiological pH19. Due to its physicochemical properties, thiolate can easily react with oxidants in the protein microenvironment, either reversibly or irreversibly, including S-nitrosylation and prenylation reactions with farnesyl or geranylgeranyl20. Redox molecules, such as H2O2, O2−, lipid peroxides, peroxynitrite and nitric oxide, covalently modify cysteine residues to alter function. Reactive cysteine residues are primarily found in specialized protein domains, including catalytic pockets, regulatory segments, and metal-binding regions21. Numerous proteins present in mammalian tissues possess reactive cysteines in their structure, including transcription factors, such as nuclear factor-κB (NF-κB), myofilaments, protein kinases10 and ion channels22. Modification of cysteine residues by ROS has been shown to alter the activity of voltage-gated Ca2+ channels23, voltage-gated potassium channels24 and, the primary subject of this review, transient receptor potential (TRP) channels25.
Redox-Sensitive TRP Channels in the Endothelium
TRP channels are a superfamily of cation-permeable ion channels first discovered in Drosophila melanogaster26, 27 that are now recognized as broadly expressed polymodal cellular sensors involved in thermoregulation, mechanotransduction, control of vascular tone1, nociception, and Ca2+ and Mg2+ homeostasis, among other functions28. The mammalian TRP superfamily consists of 28 distinct gene products (27 are present in humans29), divided into 6 subfamilies: canonical (TRPC), vanilloid (TRPV), ankyrin (TRPA), mucolipin (TRPML), melastatin (TRPM) and polycystin (TRPP)1, 29. Various reports indicate detection of mRNA or protein of at least 20 TRP channels in cultured and native endothelial cells, including the redox-sensitive channels TRPM230-32, TRPC533, 34, TRPA135, 36, TRPV137, TRPV338, 39, and TRPV440, 41. TRPC5, TRPA1, TRPV1, TRPV3, and TRPV4 channels are modulated by covalent modification of cysteine residues by ROS and/or RNS42, 43. TRPML1, a Ca2+-permeable non-selective cation channel that localizes to late endosomes and lysosomes 44, 45 is also activated by ROS in vitro to regulate autophagy46. TRPML1 appears to be ubiquitously expressed but is not known to be specifically involved in endothelial function. Regulation of TRPM2 is distinctive and involves direct binding of oxidant-derived second messengers47, 48 (Figure 2). TRPA1, TRPV3, and TRPV4 are of particular interest in terms of endothelial cell function because these channels exhibit high Ca2+ permeability1, 36, 49, 50.
Figure 2. Redox modifications of TRP channel activity.
ROS generated by NADPH oxidase activity generate the lipid peroxide 4-hydroxynonenal (4-HNE), which activated the TRPA1 channel. Further, many reactive cysteine radicals (red circles) are present in the ankyrin repeat domains of the N-terminal tail of the TRPA1 channel, and they have the potential to alter TRPA1 function. Nitric oxide (NO) mediated S-nitrosylation of reactive cysteines located near the pore-forming has also been shown to activate TRPV3, TRPV4 and TRPC5. However, S-nitrosylation of a cysteine in the C-terminal tail of TRPV4 inhibits the channel. TRPM2 is activated by ADP-ribose, which binds to the NUDIX motif in the C-terminal tail of the channel. Oxidative stress, such as the generated by accumulation of β-amyloid, increase ADP-ribose in cells, leading to prolonged TRPM2 activation.
The remainder of this review will discuss the current literature regarding redox regulation of specific TRP channels and how these pathways influence endothelial function under normal conditions and during cardiovascular disease. Here we attempt to specify if findings were derived from experiments performed in heterologous overexpression system, cultured endothelial cells, or in native tissues. Consideration of these factors is critical because cell culture conditions and overexpression of proteins often leads to phenotypic changes and experimental artifacts. We encourage the reader to carefully consider such issues when interpreting studies that are not corroborated by the use of freshly isolated endothelial cells or in vivo methods.
Redox Regulation of TRPM2
TRPM2 was initially described as a non-selective cation channel regulated by the second messenger adenosine diphosphate ribose (ADPR)51. A later study demonstrating that TRPM2 currents are stimulated by H2O2 is the first evidence of redox-dependent TRP channel activation52. Subsequent reports linked these two pathways and demonstrated that H2O2 indirectly activates TRPM2 by inducing the production of NAD+ that is promptly converted to poly(ADPR) in the mitochondria53. ADPR binds to a nudix box phosphohydrolase enzymatic domain (NUDT9-H) in the TRPM2 C-terminal tail, consequently activating the channel54. Cyclic ADPR indirectly activates TRPM2 channels in a similar manner55. Interestingly, the NUDT9-H domain of TRPM2 channels has slow hydrolytic kinetics, thus it is likely involved in ADPR binding and channel gating, rather than ADPR hydrolysis51, 56, 57, as opposed to the homologous mitochondrial pyrophosphatase NUDIT957. ADPR-induced gating of TRPM2 requires co-binding of Ca2+ 58, suggesting that the channel acts as a co-incidence detector of upstream signaling events involving increases in intracellular [Ca2+] and generation of ADPR. TRPM2 currents are difficult to isolate and study in native cells and the in vivo setting because the current-voltage relationship is linear and no selective pharmacological inhibitors are currently available.
Several recent studies show that activation of TRPM2 channels by ROS in the endothelium could be involved in vascular disease. For example, H2O2 induced TRPM2-like currents and transient Ca2+ entry in “Ca2+ add-back experiments” (similar to store operated Ca2+ entry protocols) in cultured human pulmonary artery endothelial cells. This study also linked TRPM2 activity to a reduction in transendothelial electrical resistance in cultured monolayers treated with H2O2, suggesting a role for the channel in the failure of tight junctions forming endothelial barriers30. This mechanism could potentially be involved in edema formation59, 60 and breakdown of the blood-brain barrier and neurotoxicity during oxidative stress61, although this has not been demonstrated experimentally. Another study links sustained activation of TRPM2 to endothelial cell apoptosis31. Apoptotic mechanisms may underlie cerebral capillary rarefaction often observed in chronic vascular diseases associated with oxidative stress, such as hypertension9, but this process has not yet been associated with TRPM2 activity. Recently, TRPM2 was shown to be involved in β-amyloid-induced neurovascular dysfunction, an important pathology contributing to Alzheimer’s dementia32. This study suggests that that β-amyloid binds to CD36, which activates NOX and ultimately leads to TRPM2 activation by increasing the intracellular pool of ADPR32. Increased TRPM2 activity was linked to endothelial dysfunction and impairment in neurovascular coupling caused by β-amyloid, and this response was not observed in TRPM2−/− mice32.
The role of TRPM2 on the pathophysiology of ischemic strokes has been investigated in vivo. TRPM2 mRNA is increased in the brain of rats after 1 and 4 weeks following transient middle cerebral artery occlusion62, a well-accepted model of cerebral ischemia/reperfusion injury63. Increased expression appears to occur primarily in microglial cells. Clotrimazole, a purported TRPM2 blocker, successfully reduced cerebral infarct in male mice after transient focal ischemia, possibly a consequence of improved neuronal survival after oxygen-glucose deprivation64. Clotrimazole did not reduce ischemic damage in the brain of female mice64. A sexual dimorphic response was also observed after global cerebral ischemia caused by cardiac arrest in mice. In this study, inhibition of TRPM2 with clotrimazole was protective in males but not in females65. Interpretation of in vivo experiments using this compound is difficult. The sexual dimorphic nature of the effects of clotrimazole on stroke is claimed to be caused by androgen-dependent production of ADPR after an ischemic stroke66, although simply administering androgens to females did not cause a TRPM2-dependent increase in cerebral infarct67. It should be noted here that clotrimazole is used medically as an anti-fungal compound and is not a selective TRPM2 blocker. Clotrimazole impedes TRPM2 activity but has many additional effects including inhibition of cytochrome P450 enzymes68, block of intermediate conductance Ca2+ activated K+ channels69, L-type voltage gated Ca2+ channels70, ATP-gated K+ channels71, and TRPM872 channels, and activation of TRPV1 and TRPA1 channels72. A more convincing study using TRPM2−/− mice showed reduced infarct after ischemia/reperfusion injury, which was a consequence of reduced neutrophil infiltration73. Interestingly, transplantation of wild type bone marrow into TRPM2−/− mice reversed the protective effect of TRPM2 deletion, possibly due to increased neutrophil infiltration73. Recently, a novel peptide inhibitor of TRPM2, tat-M2NX, was shown to reduce infarct size in male mice, even when administered 3 hours after reperfusion74. These findings identify TRPM2 as a possible therapeutic target to reduce ischemic damage, although direct involvement of the endothelium was not considered in these studies.
The overall importance of TRPM2 as a redox-regulated Ca2+ entry channel in native endothelial cells remains undetermined. TRPM2 is a non-selective cation channel that is reported more permeable to Na+ versus Ca2+ ions (PCa:PNa ~ 0.6)47, 51, although a single report suggests that the relative Ca2+ to Na+ is permeability is ~6:1 in HEK293 cells at physiological temperatures (>35ºC)75. Regardless of the PCa:PNa, the Ca2+ fraction of the mixed cation current conducted by TRPM2 is expected to be very small because the channel appears to be equally permeable to K+ and Na+76 leading to a localized depolarization-induced reduction in the driving force for Ca2+ entry in these non-excitable cells77. In contrast, TRPV4 channels, which support nonselective cation currents with a high Ca2+ fraction78, are more permeable for K+ versus Na+77. On this basis, we suggest that the biophysical properties of TRPM2 channels are inconsistent with a potential role as an oxidant-dependent Ca2+ influx pathway in the endothelium. However, it is conceivable that Na+ influx through TRPM2 could drive the Na+/Ca2+ exchanger into reverse mode to initiate Ca2+ influx79, but this has yet to be demonstrated experimentally. It is also possible that oxidative stress-dependent activation of TRPM2 underlies pathological loss of endothelial cell homeostasis by destabilizing of the membrane potential, thereby contributing to loss of barrier function and other vascular pathologies.
Regulation of TRPC5 Activity by S-nitrosylation
TRPC5 is a receptor-operated channel responsive to stimulation of Gq/11-protein coupled receptors upstream of PLCβ and receptor tyrosine kinases linked to PLCγ80, 81. Activity of the channel is potentiated by increases in intracellular Ca2+ 82. TRPC5 is a non-selective cation channel permeable to Ca2+, Na+ and K+, with reports of relative permeability Ca2+ to Na+ permeability ranging from PCa:PNa ~ 283, 84 to PCa:PNa ~ 1485. The extracellular linker domain between the transmembrane domain S5 and S6 contains two cysteine residues (Cys553 and Cys558) that are potential targets for NO-dependent S-nitrosylation43. It has been suggested that when these residues are S-nitrosylated a disulfide bond forms in the linker region stabilizing the channel in the open configuration86. This NO-dependent mechanism could potentially have important physiological consequences in the endothelium. In cultured bovine aortic endothelial cells, TRPC5 activation was linked to inhibition of endothelial cell migration87 as a consequence of a Ca2+ influx-dependent rearrangement of the cytoskeleton88. These findings could have implications in the context of angiogenesis and recovery of endothelial cell lesions in arteries of patients with atherosclerosis. However, activation of TRPC5 by S-nitrosylation remains controversial and no in vivo studies have demonstrated involvement of TRPC5 in endothelial function under normal conditions or in cardiovascular disease models.
Physiological Redox Signaling Regulates TRPA1 Activity
The sole member of the ankyrin subfamily of TRP channels, TRPA1, is a large-conductance (~100 pS) cation channel that is more permeable to Ca2+ compared with Na+ ions (PCa:PNa ~ 7.9)1, 89. TRPA1 conducts the highest fractional Ca2+ current of all TRP channels with the exception of the epithelial cell Ca2+-selective channels TRPV5 and TRPV690. TRPA1 is activated by a host of electrophilic compounds including the dietary molecules allyl isothiocyanate (from mustard), cinnamaldehyde (from cinnamon), and allicin (from garlic)1. Trevisani et al. reported the first evidence of a redox-dependent regulation when they showed that TRPA1 currents can be elicited by an endogenous-produced lipid peroxidation product, 4-hydroxy-2-nonenal (4-HNE), in an HEK293 expression system and dorsal root ganglia neurons91. Interestingly, a recent study shows that stimulation of TRPA1 activity with 4-HNE translated to an increase in release of the algesic neuropeptides substance P and calcitonin-gene related peptide (CGRP) from nerve terminals in the spinal cord and pain behavior in rodents91. Adding complexity to redox regulation of TRPA1 channels, a recent report showed that activation of TRPA1 with cinnamaldehyde causes neurogenic vasodilation in the mouse ear by activation of perivascular CGRP and nitrergic nerve terminals92. Interestingly, TRPA1 activation was upstream of ROS and RNS generation, and the resulting vasodilation was likely mediated by an increase in peroxynitrite92, which has been shown to induce smooth muscle relaxation by stimulation of cGMP production93, 94 and activation of KATP channels94, 95.
A recent report by Sullivan et al. showed that ROS generated by NOX2 activity activates TRPA1 in endothelial cells of cerebral arteries in an autocrine manner, leading to vasodilation35. This study showed that TRPA1 and NOX2 co-localize within myendothelial projections in cerebral arteries, likely forming signaling complexes, and dissected a pathway in which NOX activity generates O2− that, through a series of reactions, leads to production of 4-HNE. Administration of NADPH, the substrate for NOX, to primary endothelial cells increased TRPA1 elemental activity, leading to a localized influx of Ca2+, recorded as TRPA1 sparklets, that activates nearby intermediate conductance Ca2+-activated K+ (IK) channels causing vasodilation of isolated cerebral arteries35. NADPH and exogenous 4-HNE failed to cause dilation of cerebral arteries pre-incubated with the TRPA1 inhibitor HC-030031 and cerebral arteries from endothelial cell-specific TRPA1−/− mice, demonstrating that NOX-derived ROS activates TRPA1 in endothelial cells35. Interestingly, TRPA1 seems to have selective expression in different vascular beds, as TRPA1 is found in cerebral artery endothelial cells, but not in mesenteric, coronary or renal arteries35. This unique distribution pattern may have important implications for therapeutic interventions in ischemic stroke and chronic cerebral hypoperfusion.
Two studies investigated the hypothesis that TRPA1 channels, which are present in sensory and vagal afferent neurons in the lungs and the heart, are cellular sensors of local oxygen concentration96, 97. Measuring Ca2+ influx in an HEK expression system, Takahashi et al. showed that during “normoxic” conditions, defined by the authors as a partial pressure of O2 (pO2) of ~150 mmHg, TRPA1 activity was undetectable. In contrast, when pO2 was ~100 mmHg, TRPA1 activity was present and at a pO2 of 80 mmHg (defined by the authors as hypoxia), TRPA1-mediated Ca2+ influx was nearly maximal96. These data suggest that a shift in pO2 from 150 mmHg to 80 mmHg strongly stimulates TRPA1-mediated Ca2+ influx, which the authors interpreted as hypoxia-induced activation. However, at sea level, pO2 within the lungs is ~100 mmHg, dropping to ~80 mmHg at 1500 m (the elevation of our laboratory and many other cities in Western North America). Hemoglobin saturation is ~100% when the pO2 of inspired air is 80 mmHg and as such, this condition does not stimulate hypoxic adaptation in healthy individuals. Furthermore, under normal physiological conditions in healthy humans, pO2 of arterial blood is ~80-100 mmHg within the aorta, ~30-70 mmHg in capillaries and ~20-40 mmHg in the venous system98. Thus, under normoxic conditions at sea level, physiological pO2 levels vary between ~100 mmHg within the lungs and ~20-40 mm Hg in venous blood. Tissue pO2 levels are typically reported to in the range of venous blood, or lower98. If the model proposed by Takahashi et al. is correct, these facts suggest that TRPA1 channels are tonically active in all tissues all of the time. This possibility is incompatible with our observations and the majority of the literature on this topic. Thus, rather than being reflective of hypoxia-induced activation, the phenomena reported by Takahashi et al. may represent oxidant-induced activation or sensitization of TRPA1 during recovery from hyperoxic conditions.
Although it seems unlikely that hypoxia per se regulates TRPA1 activity, one study suggests that TRPA1 gene expression may be regulated by hypoxia. Using a bioinformatic approach, Hatano et al. identified ten consensus hypoxia-responsive elements in the promoter region for the human TRPA1 gene. The authors used chromatin immunoprecipitation to show that the transcription factor hypoxia inducible factor-1α (HIF-1α) binds to these sites in synoviocytes. Further, a luciferase reporter assay identified that one of the reverse hypoxia-responsive element within the promoter is sufficient to activate transcription99. However, to date the effects of hypoxia on TRPA1 expression levels in the endothelium have not been reported.
S-nytrosylation of TRPV3 channels
TRPV3 has the highest unitary conductance in the vanilloid subfamily (~200 pS) and is more permeable to Ca2+ ion compared with Na+ ions (PCa:PNa ~10)49. TRPV3 is activated by innocuous heat (>30º C)100, suggesting a role in thermosensation and tonic activity at physiological temperatures. Plant-derived monoterpenoids, such as carvacrol (from Origanum vulgare) and camphor (from Cinnamomum camphora) are chemical activators of TRPV3101, 102. It has been postulated that intermediate metabolites of the mevalonate biosynthetic pathway involved in cholesterol biosynthesis are endogenous TRPV3 agonists103. Farnesyl pyrophosphate was shown to activate TRPV3 in heterologous expression systems104, whereas its precursor, isopentenyl pyrophosphate, inhibits the channel39, 105. Interestingly, one report indicates that TRPV3 subunits can form active heteromultimeric channels with TRPV1 subunits in HEK293 cells expressing a cDNA construct encoding a fused TRPV1/TRPV3 protein106. The resulting channels exhibit a unitary conductance and temperature sensitivity that is intermediate compared to the respective homomeric channels and increased sensitivity to capsaicin compare to TRPV1 homomeric channels106. It is not known if TRPV3 forms heteromultimeric channels with TRPV1 or other TRPV channels in vivo.
TRPV3 is present in the cerebral endothelium and activation by carvacrol results in endothelium-dependent vasodilation of cerebral pial arteries and parenchymal arterioles38, 39, 107 by a Ca2+-dependent signaling pathway requiring the activity of IK and small conductance Ca2+-activated K+ (SK) channel38. Interestingly, although TRPV3 might be expected to be consitutively active at physiological temperatures in mammals based theromosenitivity100, does not seem to be involved in maintenance of basal myogenic tone in cerebral parenchymal arterioles, as inhibition with isopentenyl pyrophosphate had no effect in pressure myograph experiments39. However, it is possible that during disease states, TRPV3 activity may be altered as result of possible excessive production of farnesyl pyrophosphate and isopentenyl pyrophosphate, that may occur when cholesterol biosynthesis is increased, such as during obesity108. TRPV3 activity has been shown to be modulated by RNS. NO causes S-nytrosylation of cysteine residues near the pore-forming region of TRPV3 expressed in HEK293 cells, leading to channel opening and Ca2+ influx43. H2O2 failed to induce TRPV3 activation in HEK293 cells, suggesting a NO specific mechanism43. NO-dependent activation of TRPV3 channel could have important implications for vascular function and cardiovascular disease, but this mode of regulation has yet to be demonstrated in native endothelial cells.
Oxidant-dependent regulation of TRPV3 was demonstrated by a study showing that severe prolonged hypoxia (24 hrs., 0.1% O2) potentiated 2-aminoethoxydiphenyl borate induced currents in HEK293 cells expressing recombinant TRPV3 channels109. This study also showed that the enzyme Factor Inhibiting HIF (FIH), a 2-oxoglutarate-dependent dioxygenase, constitutively hydroxylates asparagine residues within the ankyrin repeat domain of TRPV3, thereby reducing channel activity. Low cellular O2 reduces the activity of FIH, releasing the inhibition and increasing TRPV3 current density upon activation109. TRPV3 channels are also activated by intracellular protons H+ 110, which could contribute to activation of TRPV3 under hypoxic conditions channels in vivo. The effects of hypoxia on TRPV3 activity in the endothelium have not been reported.
TRPV4: Regulated by and Regulator of Redox Signaling
TRPV4 channels are present in the smooth muscle and endothelial layer of arterial wall in many vascular beds111, 112. TRPV4 has a single channel conductance of approximately 60 pS at negative membrane potentials50, and a higher permeability for Ca2+ versus Na+ ions (PCa:PNa ~6)50, 113. TRPV4 channels were initially described as osmotically regulated channels acting as mechanosensors114. Subsequent studies suggest that TRPV4 is not inherently mechanosensitive, but can be indirectly activated by force-dependent signaling pathways5. For example, in cultured endothelial cells a complex formed between TRPV4 and integrins responds to mechanical strain115. TRPV4 is activated by epoxyeicosatrienoic acids and anandamide111, 116, 117, which may be endogenous agonists of the channel, and activity is initially potentiated118 and subsequently inhibited by increases in intracellular Ca2+ 119.
TRPV4 activity has been linked to endothelium-dependent arterial dilation in different vascular beds. Activation of TRPV4 was shown to induce dilation of mesenteric arteries112, 120, cerebral arteries41, 111 and coronary arterioles121. In mesenteric arteries, TRPV4 activation with the highly selective pharmacological activator GSK1016790A leads to vasodilation associated with elemental Ca2+ entry through TRPV4 channels, recorded as TRPV4 sparklets. Interestingly, stimulation of the muscarinic receptor with acetylcholine also increased TRPV4 sparklet activity in endothelial cells of mesenteric arteries, suggesting that downstream phospholipase C and protein kinase C signaling pathways influence TRPV4 activity120.
A few studies suggest that TRPV4 activity may be regulated by ROS and RNS. NO promotes S-nytrosylation of cysteines close to the pore-forming domain of TRPV4 monomers, which may increase activity43. In contrast, a study by Lee et al. showed that S-nytrosylation of the Cys853 residue, located on the intracellular C-terminal tail, prevented channel activation by 4α-phorbol-12,13-didecanoate in HEK293 cells122. Desensitization appears to be a consequence of reduced interaction between TRPV4 and calmodulin122, a previously described Ca2+-dependent mechanism of TRPV4 inactivation118. In cultured pulmonary endothelial cells, H2O2 induces Ca2+ influx through a mechanism that is dependent of TRPV4 activity123. Although not directly evaluated in the study, it is possible that in the short-term (1 hour of exposure) H2O2 may act directly on the TRPV4 channel, whereas on the long-term (24 hours of exposure to H2O2) TRPV4 activity is regulated by the Src kinase Fyn 123. Interestingly, Ca2+ entry through TRPV4 channels may elevate ROS production in the coronary endothelium, providing a positive feedback mechanism to stimulate flow-induced dilation121. This study shows that in human coronary endothelial cells, TRPV4-mediated Ca2+ influx stimulated with 4α-phorbol-12,13-didecanoate induced O2− and H2O2 production in the mitochondria of these cells. This response and flow-induced dilation of coronary arteries were blocked by TRPV4 inhibition, linking the ROS and TRPV4 pathways121.
Concluding remarks
Controlled production of reactive oxidant radicals as cellular signaling molecules is now an accepted paradigm, whereas excessive generation of these molecules is associated with numerous pathological conditions. This topic is of considerable interest in the context of chronic cardiovascular diseases, where oxidative stress is a strong predictive marker124, but antioxidant therapy has no demonstrable protective effects125. The literature reviewed here shows that ROS and RNS can stimulate the activity of several different TRP channels present in the endothelium to locally or globally increase in intracellular [Ca2+] and stimulate Ca2+-dependent signaling pathways. In healthy tissues, the outcomes of these redox pathways can be adaptive, such as endothelium-dependent dilation and increased tissue perfusion, whereas during oxidative stress, elevated and/or prolonged activation of Ca2+ signaling pathways can have detrimental effects including loss of barrier function and apoptosis. Failure of antioxidant therapy to improve cardiovascular disease outcomes could be a consequence of indiscriminately inhibiting both positive and negative aspects of redox-dependent signaling networks. The studies reviewed here provide a foundation for future integrative in vivo experiments exploring new therapeutic avenues for vascular diseases related to oxidative stress that target specific redox-sensitive TRP channels in the endothelium rather than blanket suppression of oxidant radical production.
Acknowledgments
Sources of funding
The present manuscript was supported by the American Heart Association (15POST24720002 to PWP) and the National Heart, Lung and Blood Institute (R01HL091905 to SE).
List of abbreviations
- 4-HNE
4-hydroxynonenal
- ADPR
adenosine diphosphate ribose
- CGRP
calcitonin-gene related peptide
- eNOS
endothelial nitric oxide synthase
- FIH
Factor Inhibiting HIF
- H2O2
hydrogen peroxide
- HIF-1α
hypoxia inducible factor-1α
- IK
intermediate conductance Ca2+-activated K+ channel
- KATP
ATP-sensitive K+ channel
- NADPH
nicotinamide adenine dinucleotide phosphate
- NO
nitric oxide
- NOX
nicotinamide adenine dinucleotide phosphate oxidases
- O2−
superoxide anion
- ROS
reactive oxygen species
- RNS
reactive nitrogen species
- SK
small conductance Ca2+-activated K+ channel
- TRP
transient receptor potential channel
- TRPA
transient receptor potential ankyrin
- TRPC
transient receptor potential canonical
- TRPM
transient receptor potential melastatin
- TRPML
transient receptor potential mucolipin
- TRPV
transient receptor potential vanilloid
Footnotes
Disclosures
The authors have no conflict of interest to disclose, financial or otherwise.
Author contributions
PWP and SE conceived the review; PWP wrote the initial draft of the manuscript; PWP constructed the figures; PWP and SE revised and edited the manuscript; PWP and SE accepted the final version of the manuscript.
REFERENCES
- 1.Earley S, Brayden JE. Transient receptor potential channels in the vasculature. Physiol Rev. 2015;95:645–90. doi: 10.1152/physrev.00026.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Tiruppathi C, Ahmmed GU, Vogel SM, Malik AB. Ca2+ signaling, TRP channels, and endothelial permeability. Microcirculation. 2006;13:693–708. doi: 10.1080/10739680600930347. [DOI] [PubMed] [Google Scholar]
- 3.Brown RC, Davis TP. Calcium modulation of adherens and tight junction function: a potential mechanism for blood-brain barrier disruption after stroke. Stroke. 2002;33:1706–11. doi: 10.1161/01.str.0000016405.06729.83. [DOI] [PubMed] [Google Scholar]
- 4.Kohler R, Heyken WT, Heinau P, Schubert R, Si H, Kacik M, Busch C, Grgic I, Maier T, Hoyer J. Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation. Arterioscler Thromb Vasc Biol. 2006;26:1495–502. doi: 10.1161/01.ATV.0000225698.36212.6a. [DOI] [PubMed] [Google Scholar]
- 5.Hill-Eubanks DC, Gonzales AL, Sonkusare SK, Nelson MT. Vascular TRP channels: performing under pressure and going with the flow. Physiology. 2014;29:343–60. doi: 10.1152/physiol.00009.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Shen B, Wong CO, Lau OC, Woo T, Bai S, Huang Y, Yao X. Plasma membrane mechanical stress activates TRPC5 channels. PLoS One. 2015;10:e0122227. doi: 10.1371/journal.pone.0122227. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Nishida M, Kumagai Y, Ihara H, Fujii S, Motohashi H, Akaike T. Redox signaling regulated by electrophiles and reactive sulfur species. J Clin Biochem Nutr. 2016;58:91–8. doi: 10.3164/jcbn.15-111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Cai H, Harrison DG. Endothelial dysfunction in cardiovascular diseases: the role of oxidant stress. Circ Res. 2000;87:840–4. doi: 10.1161/01.res.87.10.840. [DOI] [PubMed] [Google Scholar]
- 9.Pires PW, Dams Ramos CM, Matin N, Dorrance AM. The effects of hypertension on the cerebral circulation. Am J Physiol Heart Circ Physiol. 2013;304:H1598–614. doi: 10.1152/ajpheart.00490.2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Droge W. Free radicals in the physiological control of cell function. Physiol Rev. 2002;82:47–95. doi: 10.1152/physrev.00018.2001. [DOI] [PubMed] [Google Scholar]
- 11.Lambeth JD. NOX enzymes and the biology of reactive oxygen. Nat Rev Immunol. 2004;4:181–9. doi: 10.1038/nri1312. [DOI] [PubMed] [Google Scholar]
- 12.Panieri E, Santoro MM. ROS signaling and redox biology in endothelial cells. Cell Mol Life Sci. 2015;72:3281–303. doi: 10.1007/s00018-015-1928-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Bedard K, Krause KH. The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev. 2007;87:245–313. doi: 10.1152/physrev.00044.2005. [DOI] [PubMed] [Google Scholar]
- 14.Winterbourn CC. Toxicity of iron and hydrogen peroxide: the Fenton reaction. Toxicol Lett. 1995;82-83:969–74. doi: 10.1016/0378-4274(95)03532-x. [DOI] [PubMed] [Google Scholar]
- 15.Nishikawa T, Edelstein D, Du XL, Yamagishi S, Matsumura T, Kaneda Y, Yorek MA, Beebe D, Oates PJ, Hammes HP, Giardino I, Brownlee M. Normalizing mitochondrial superoxide production blocks three pathways of hyperglycaemic damage. Nature. 2000;404:787–90. doi: 10.1038/35008121. [DOI] [PubMed] [Google Scholar]
- 16.Han D, Antunes F, Canali R, Rettori D, Cadenas E. Voltage-dependent anion channels control the release of the superoxide anion from mitochondria to cytosol. J Biol Chem. 2003;278:5557–63. doi: 10.1074/jbc.M210269200. [DOI] [PubMed] [Google Scholar]
- 17.Jaffrey SR, Erdjument-Bromage H, Ferris CD, Tempst P, Snyder SH. Protein S-nitrosylation: a physiological signal for neuronal nitric oxide. Nat Cell Biol. 2001;3:193–7. doi: 10.1038/35055104. [DOI] [PubMed] [Google Scholar]
- 18.Vasquez-Vivar J, Kalyanaraman B, Martasek P, Hogg N, Masters BS, Karoui H, Tordo P, Pritchard KA., Jr. Superoxide generation by endothelial nitric oxide synthase: the influence of cofactors. Proc Natl Acad Sci U S A. 1998;95:9220–5. doi: 10.1073/pnas.95.16.9220. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Marino SM, Gladyshev VN. Analysis and functional prediction of reactive cysteine residues. J Biol Chem. 2012;287:4419–25. doi: 10.1074/jbc.R111.275578. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Zhang FL, Casey PJ. Protein prenylation: molecular mechanisms and functional consequences. Annu Rev Biochem. 1996;65:241–69. doi: 10.1146/annurev.bi.65.070196.001325. [DOI] [PubMed] [Google Scholar]
- 21.Marino SM, Gladyshev VN. Cysteine function governs its conservation and degeneration and restricts its utilization on protein surfaces. Journal of molecular biology. 2010;404:902–16. doi: 10.1016/j.jmb.2010.09.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Bogeski I, Kappl R, Kummerow C, Gulaboski R, Hoth M, Niemeyer BA. Redox regulation of calcium ion channels: chemical and physiological aspects. Cell Calcium. 2011;50:407–23. doi: 10.1016/j.ceca.2011.07.006. [DOI] [PubMed] [Google Scholar]
- 23.Hidalgo C, Donoso P. Crosstalk between calcium and redox signaling: from molecular mechanisms to health implications. Antioxid Redox Signal. 2008;10:1275–312. doi: 10.1089/ars.2007.1886. [DOI] [PubMed] [Google Scholar]
- 24.Sahoo N, Hoshi T, Heinemann SH. Oxidative modulation of voltage-gated potassium channels. Antioxid Redox Signal. 2014;21:933–52. doi: 10.1089/ars.2013.5614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Kozai D, Ogawa N, Mori Y. Redox regulation of transient receptor potential channels. Antioxid Redox Signal. 2014;21:971–86. doi: 10.1089/ars.2013.5616. [DOI] [PubMed] [Google Scholar]
- 26.Montell C, Rubin GM. Molecular characterization of the Drosophila trp locus: a putative integral membrane protein required for phototransduction. Neuron. 1989;2:1313–23. doi: 10.1016/0896-6273(89)90069-x. [DOI] [PubMed] [Google Scholar]
- 27.Minke B, Wu C, Pak WL. Induction of photoreceptor voltage noise in the dark in Drosophila mutant. Nature. 1975;258:84–7. doi: 10.1038/258084a0. [DOI] [PubMed] [Google Scholar]
- 28.Gees M, Owsianik G, Nilius B, Voets T. TRP channels. Compr Physiol. 2012;2:563–608. doi: 10.1002/cphy.c110026. [DOI] [PubMed] [Google Scholar]
- 29.Flockerzi V, Nilius B. TRPs: truly remarkable proteins. Handb Exp Pharmacol. 2014;222:1–12. doi: 10.1007/978-3-642-54215-2_1. [DOI] [PubMed] [Google Scholar]
- 30.Hecquet CM, Ahmmed GU, Vogel SM, Malik AB. Role of TRPM2 channel in mediating H2O2-induced Ca2+ entry and endothelial hyperpermeability. Circ Res. 2008;102:347–55. doi: 10.1161/CIRCRESAHA.107.160176. [DOI] [PubMed] [Google Scholar]
- 31.Hecquet CM, Zhang M, Mittal M, Vogel SM, Di A, Gao X, Bonini MG, Malik AB. Cooperative interaction of trp melastatin channel transient receptor potential (TRPM2) with its splice variant TRPM2 short variant is essential for endothelial cell apoptosis. Circ Res. 2014;114:469–79. doi: 10.1161/CIRCRESAHA.114.302414. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Park L, Wang G, Moore J, Girouard H, Zhou P, Anrather J, Iadecola C. The key role of transient receptor potential melastatin-2 channels in amyloid-beta-induced neurovascular dysfunction. Nat Commun. 2014;5:5318. doi: 10.1038/ncomms6318. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Beech DJ. Canonical transient receptor potential 5. Handb Exp Pharmacol. 2007:109–23. doi: 10.1007/978-3-540-34891-7_6. [DOI] [PubMed] [Google Scholar]
- 34.Garcia RL, Schilling WP. Differential expression of mammalian TRP homologues across tissues and cell lines. Biochem Biophys Res Commun. 1997;239:279–83. doi: 10.1006/bbrc.1997.7458. [DOI] [PubMed] [Google Scholar]
- 35.Sullivan MN, Gonzales AL, Pires PW, Bruhl A, Leo MD, Li W, Oulidi A, Boop FA, Feng Y, Jaggar JH, Welsh DG, S E. Localized TRPA1 Channel Ca2+ Signals Stimulated by Reactive Oxygen Species Promote Cerebral Artery Dilation. Science Signaling. 2015:8. doi: 10.1126/scisignal.2005659. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Earley S, Gonzales AL, Crnich R. Endothelium-dependent cerebral artery dilation mediated by TRPA1 and Ca2+-Activated K+ channels. Circ Res. 2009;104:987–94. doi: 10.1161/CIRCRESAHA.108.189530. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Yang D, Luo Z, Ma S, Wong WT, Ma L, Zhong J, He H, Zhao Z, Cao T, Yan Z, Liu D, Arendshorst WJ, Huang Y, Tepel M, Zhu Z. Activation of TRPV1 by dietary capsaicin improves endothelium-dependent vasorelaxation and prevents hypertension. Cell metabolism. 2010;12:130–41. doi: 10.1016/j.cmet.2010.05.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 38.Earley S, Gonzales AL, Garcia ZI. A dietary agonist of transient receptor potential cation channel V3 elicits endothelium-dependent vasodilation. Molecular pharmacology. 2010;77:612–20. doi: 10.1124/mol.109.060715. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Pires PW, Sullivan MN, Pritchard HA, Robinson JJ, Earley S. Unitary TRPV3 channel Ca2+ influx events elicit endothelium-dependent dilation of cerebral parenchymal arterioles. Am J Physiol Heart Circ Physiol. 2015;309:H2031–41. doi: 10.1152/ajpheart.00140.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Hartmannsgruber V, Heyken WT, Kacik M, Kaistha A, Grgic I, Harteneck C, Liedtke W, Hoyer J, Kohler R. Arterial response to shear stress critically depends on endothelial TRPV4 expression. PLoS One. 2007;2:e827. doi: 10.1371/journal.pone.0000827. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Marrelli SP, O'Neil RG, Brown RC, Bryan RM., Jr. PLA2 and TRPV4 channels regulate endothelial calcium in cerebral arteries. American journal of physiology Heart and circulatory physiology. 2007;292:H1390–7. doi: 10.1152/ajpheart.01006.2006. [DOI] [PubMed] [Google Scholar]
- 42.Sawada Y, Hosokawa H, Matsumura K, Kobayashi S. Activation of transient receptor potential ankyrin 1 by hydrogen peroxide. Eur J Neurosci. 2008;27:1131–42. doi: 10.1111/j.1460-9568.2008.06093.x. [DOI] [PubMed] [Google Scholar]
- 43.Yoshida T, Inoue R, Morii T, Takahashi N, Yamamoto S, Hara Y, Tominaga M, Shimizu S, Sato Y, Mori Y. Nitric oxide activates TRP channels by cysteine S-nitrosylation. Nature chemical biology. 2006;2:596–607. doi: 10.1038/nchembio821. [DOI] [PubMed] [Google Scholar]
- 44.Wang W, Zhang X, Gao Q, Xu H. TRPML1: an ion channel in the lysosome. Handb Exp Pharmacol. 2014;222:631–45. doi: 10.1007/978-3-642-54215-2_24. [DOI] [PubMed] [Google Scholar]
- 45.Dong XP, Shen D, Wang X, Dawson T, Li X, Zhang Q, Cheng X, Zhang Y, Weisman LS, Delling M, Xu H. PI(3,5)P(2) controls membrane trafficking by direct activation of mucolipin Ca(2+) release channels in the endolysosome. Nat Commun. 2010;1:38. doi: 10.1038/ncomms1037. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 46.Zhang X, Cheng X, Yu L, Yang J, Calvo R, Patnaik S, Hu X, Gao Q, Yang M, Lawas M, Delling M, Marugan J, Ferrer M, Xu H. MCOLN1 is a ROS sensor in lysosomes that regulates autophagy. Nat Commun. 2016;7:12109. doi: 10.1038/ncomms12109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Kraft R, Grimm C, Grosse K, Hoffmann A, Sauerbruch S, Kettenmann H, Schultz G, Harteneck C. Hydrogen peroxide and ADP-ribose induce TRPM2-mediated calcium influx and cation currents in microglia. Am J Physiol Cell Physiol. 2004;286:C129–37. doi: 10.1152/ajpcell.00331.2003. [DOI] [PubMed] [Google Scholar]
- 48.Naziroglu M, Luckhoff A. A calcium influx pathway regulated separately by oxidative stress and ADP-Ribose in TRPM2 channels: single channel events. Neurochem Res. 2008;33:1256–62. doi: 10.1007/s11064-007-9577-5. [DOI] [PubMed] [Google Scholar]
- 49.Gees M, Colsoul B, Nilius B. The role of transient receptor potential cation channels in Ca2+ signaling. Cold Spring Harb Perspect Biol. 2010;2:a003962. doi: 10.1101/cshperspect.a003962. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Strotmann R, Harteneck C, Nunnenmacher K, Schultz G, Plant TD. OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity. Nat Cell Biol. 2000;2:695–702. doi: 10.1038/35036318. [DOI] [PubMed] [Google Scholar]
- 51.Perraud AL, Fleig A, Dunn CA, Bagley LA, Launay P, Schmitz C, Stokes AJ, Zhu Q, Bessman MJ, Penner R, Kinet JP, Scharenberg AM. ADP-ribose gating of the calcium-permeable LTRPC2 channel revealed by Nudix motif homology. Nature. 2001;411:595–9. doi: 10.1038/35079100. [DOI] [PubMed] [Google Scholar]
- 52.Hara Y, Wakamori M, Ishii M, Maeno E, Nishida M, Yoshida T, Yamada H, Shimizu S, Mori E, Kudoh J, Shimizu N, Kurose H, Okada Y, Imoto K, Mori Y. LTRPC2 Ca2+-permeable channel activated by changes in redox status confers susceptibility to cell death. Mol Cell. 2002;9:163–73. doi: 10.1016/s1097-2765(01)00438-5. [DOI] [PubMed] [Google Scholar]
- 53.Schreiber V, Dantzer F, Ame JC, de Murcia G. Poly(ADP-ribose): novel functions for an old molecule. Nat Rev Mol Cell Biol. 2006;7:517–28. doi: 10.1038/nrm1963. [DOI] [PubMed] [Google Scholar]
- 54.Perraud AL, Takanishi CL, Shen B, Kang S, Smith MK, Schmitz C, Knowles HM, Ferraris D, Li W, Zhang J, Stoddard BL, Scharenberg AM. Accumulation of free ADP-ribose from mitochondria mediates oxidative stress-induced gating of TRPM2 cation channels. J Biol Chem. 2005;280:6138–48. doi: 10.1074/jbc.M411446200. [DOI] [PubMed] [Google Scholar]
- 55.Toth B, Csanady L. Identification of direct and indirect effectors of the transient receptor potential melastatin 2 (TRPM2) cation channel. The Journal of biological chemistry. 2010;285:30091–102. doi: 10.1074/jbc.M109.066464. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 56.Perraud AL, Shen B, Dunn CA, Rippe K, Smith MK, Bessman MJ, Stoddard BL, Scharenberg AM. NUDT9, a member of the Nudix hydrolase family, is an evolutionarily conserved mitochondrial ADP-ribose pyrophosphatase. J Biol Chem. 2003;278:1794–801. doi: 10.1074/jbc.M205601200. [DOI] [PubMed] [Google Scholar]
- 57.Iordanov I, Mihalyi C, Toth B, Csanady L. The proposed channel-enzyme transient receptor potential melastatin 2 does not possess ADP ribose hydrolase activity. Elife. 2016:5. doi: 10.7554/eLife.17600. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Csanady L, Torocsik B. Four Ca2+ ions activate TRPM2 channels by binding in deep crevices near the pore but intracellularly of the gate. The Journal of general physiology. 2009;133:189–203. doi: 10.1085/jgp.200810109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Di A, Gao XP, Qian F, Kawamura T, Han J, Hecquet C, Ye RD, Vogel SM, Malik AB. The redox-sensitive cation channel TRPM2 modulates phagocyte ROS production and inflammation. Nat Immunol. 2012;13:29–34. doi: 10.1038/ni.2171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Rosenberg GA, Yang Y. Vasogenic edema due to tight junction disruption by matrix metalloproteinases in cerebral ischemia. Neurosurg Focus. 2007;22:E4. doi: 10.3171/foc.2007.22.5.5. [DOI] [PubMed] [Google Scholar]
- 61.Ballabh P, Braun A, Nedergaard M. The blood-brain barrier: an overview: structure, regulation, and clinical implications. Neurobiol Dis. 2004;16:1–13. doi: 10.1016/j.nbd.2003.12.016. [DOI] [PubMed] [Google Scholar]
- 62.Fonfria E, Mattei C, Hill K, Brown JT, Randall A, Benham CD, Skaper SD, Campbell CA, Crook B, Murdock PR, Wilson JM, Maurio FP, Owen DE, Tilling PL, McNulty S. TRPM2 is elevated in the tMCAO stroke model, transcriptionally regulated, and functionally expressed in C13 microglia. J Recept Signal Transduct Res. 2006;26:179–98. doi: 10.1080/10799890600637522. [DOI] [PubMed] [Google Scholar]
- 63.Longa EZ, Weinstein PR, Carlson S, Cummins R. Reversible middle cerebral artery occlusion without craniectomy in rats. Stroke. 1989;20:84–91. doi: 10.1161/01.str.20.1.84. [DOI] [PubMed] [Google Scholar]
- 64.Jia J, Verma S, Nakayama S, Quillinan N, Grafe MR, Hurn PD, Herson PS. Sex differences in neuroprotection provided by inhibition of TRPM2 channels following experimental stroke. J Cereb Blood Flow Metab. 2011;31:2160–8. doi: 10.1038/jcbfm.2011.77. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Nakayama S, Vest R, Traystman RJ, Herson PS. Sexually dimorphic response of TRPM2 inhibition following cardiac arrest-induced global cerebral ischemia in mice. J Mol Neurosci. 2013;51:92–8. doi: 10.1007/s12031-013-0005-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Yuan M, Siegel C, Zeng Z, Li J, Liu F, McCullough LD. Sex differences in the response to activation of the poly (ADP-ribose) polymerase pathway after experimental stroke. Exp Neurol. 2009;217:210–8. doi: 10.1016/j.expneurol.2009.02.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Quillinan N, Grewal H, Klawitter J, Herson PS. Sex Steroids Do Not Modulate TRPM2-Mediated Injury in Females following Middle Cerebral Artery Occlusion(1,2,3) eNeuro. 2014:1. doi: 10.1523/ENEURO.0022-14.2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Rendic S, Guengerich FP. Update information on drug metabolism systems--2009, part II: summary of information on the effects of diseases and environmental factors on human cytochrome P450 (CYP) enzymes and transporters. Curr Drug Metab. 2010;11:4–84. doi: 10.2174/138920010791110917. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Wulff H, Gutman GA, Cahalan MD, Chandy KG. Delineation of the clotrimazole/TRAM-34 binding site on the intermediate conductance calcium-activated potassium channel, IKCa1. J Biol Chem. 2001;276:32040–5. doi: 10.1074/jbc.M105231200. [DOI] [PubMed] [Google Scholar]
- 70.Xiao YF, Huang L, Morgan JP. Cytochrome P450: a novel system modulating Ca2+ channels and contraction in mammalian heart cells. J Physiol. 1998;508:777–92. doi: 10.1111/j.1469-7793.1998.777bp.x. Pt 3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Edwards G, Zygmunt PM, Hogestatt ED, Weston AH. Effects of cytochrome P450 inhibitors on potassium currents and mechanical activity in rat portal vein. Br J Pharmacol. 1996;119:691–701. doi: 10.1111/j.1476-5381.1996.tb15728.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Meseguer V, Karashima Y, Talavera K, D'Hoedt D, Donovan-Rodriguez T, Viana F, Nilius B, Voets T. Transient receptor potential channels in sensory neurons are targets of the antimycotic agent clotrimazole. J Neurosci. 2008;28:576–86. doi: 10.1523/JNEUROSCI.4772-07.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Gelderblom M, Melzer N, Schattling B, Gob E, Hicking G, Arunachalam P, Bittner S, Ufer F, Herrmann AM, Bernreuther C, Glatzel M, Gerloff C, Kleinschnitz C, Meuth SG, Friese MA, Magnus T. Transient receptor potential melastatin subfamily member 2 cation channel regulates detrimental immune cell invasion in ischemic stroke. Stroke. 2014;45:3395–402. doi: 10.1161/STROKEAHA.114.005836. [DOI] [PubMed] [Google Scholar]
- 74.Shimizu T, Dietz RM, Cruz-Torres I, Strnad F, Garske AK, Moreno M, Venna VR, Quillinan N, Herson PS. Extended therapeutic window of a novel peptide inhibitor of TRPM2 channels following focal cerebral ischemia. Exp Neurol. 2016;283:151–6. doi: 10.1016/j.expneurol.2016.06.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Togashi K, Hara Y, Tominaga T, Higashi T, Konishi Y, Mori Y, Tominaga M. TRPM2 activation by cyclic ADP-ribose at body temperature is involved in insulin secretion. The EMBO journal. 2006;25:1804–15. doi: 10.1038/sj.emboj.7601083. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Sano Y, Inamura K, Miyake A, Mochizuki S, Yokoi H, Matsushime H, Furuichi K. Immunocyte Ca2+ influx system mediated by LTRPC2. Science. 2001;293:1327–30. doi: 10.1126/science.1062473. [DOI] [PubMed] [Google Scholar]
- 77.Perraud AL, Schmitz C, Scharenberg AM. TRPM2 Ca2+ permeable cation channels: from gene to biological function. Cell Calcium. 2003;33:519–31. doi: 10.1016/s0143-4160(03)00057-5. [DOI] [PubMed] [Google Scholar]
- 78.Mercado J, Baylie R, Navedo MF, Yuan C, Scott JD, Nelson MT, Brayden JE, Santana LF. Local control of TRPV4 channels by AKAP150-targeted PKC in arterial smooth muscle. The Journal of general physiology. 2014;143:559–75. doi: 10.1085/jgp.201311050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Andrikopoulos P, Baba A, Matsuda T, Djamgoz MB, Yaqoob MM, Eccles SA. Ca2+ influx through reverse mode Na+/Ca2+ exchange is critical for vascular endothelial growth factor-mediated extracellular signal-regulated kinase (ERK) 1/2 activation and angiogenic functions of human endothelial cells. J Biol Chem. 2011;286:37919–31. doi: 10.1074/jbc.M111.251777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Philipp S, Hambrecht J, Braslavski L, Schroth G, Freichel M, Murakami M, Cavalie A, Flockerzi V. A novel capacitative calcium entry channel expressed in excitable cells. The EMBO journal. 1998;17:4274–82. doi: 10.1093/emboj/17.15.4274. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Plant TD, Schaefer M. Receptor-operated cation channels formed by TRPC4 and TRPC5. Naunyn Schmiedebergs Arch Pharmacol. 2005;371:266–76. doi: 10.1007/s00210-005-1055-5. [DOI] [PubMed] [Google Scholar]
- 82.Blair NT, Kaczmarek JS, Clapham DE. Intracellular calcium strongly potentiates agonist-activated TRPC5 channels. The Journal of general physiology. 2009;133:525–46. doi: 10.1085/jgp.200810153. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Schaefer M, Plant TD, Obukhov AG, Hofmann T, Gudermann T, Schultz G. Receptor-mediated regulation of the nonselective cation channels TRPC4 and TRPC5. J Biol Chem. 2000;275:17517–26. doi: 10.1074/jbc.275.23.17517. [DOI] [PubMed] [Google Scholar]
- 84.Jung S, Muhle A, Schaefer M, Strotmann R, Schultz G, Plant TD. Lanthanides potentiate TRPC5 currents by an action at extracellular sites close to the pore mouth. J Biol Chem. 2003;278:3562–71. doi: 10.1074/jbc.M211484200. [DOI] [PubMed] [Google Scholar]
- 85.Okada T, Shimizu S, Wakamori M, Maeda A, Kurosaki T, Takada N, Imoto K, Mori Y. Molecular cloning and functional characterization of a novel receptor-activated TRP Ca2+ channel from mouse brain. J Biol Chem. 1998;273:10279–87. doi: 10.1074/jbc.273.17.10279. [DOI] [PubMed] [Google Scholar]
- 86.Takahashi N, Kozai D, Mori Y. TRP channels: sensors and transducers of gasotransmitter signals. Front Physiol. 2012;3:324. doi: 10.3389/fphys.2012.00324. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Chaudhuri P, Colles SM, Bhat M, Van Wagoner DR, Birnbaumer L, Graham LM. Elucidation of a TRPC6-TRPC5 channel cascade that restricts endothelial cell movement. Mol Biol Cell. 2008;19:3203–11. doi: 10.1091/mbc.E07-08-0765. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Chaudhuri P, Colles SM, Damron DS, Graham LM. Lysophosphatidylcholine inhibits endothelial cell migration by increasing intracellular calcium and activating calpain. Arterioscler Thromb Vasc Biol. 2003;23:218–23. doi: 10.1161/01.atv.0000052673.77316.01. [DOI] [PubMed] [Google Scholar]
- 89.Earley S. TRPA1 channels in the vasculature. Br J Pharmacol. 2012;167:13–22. doi: 10.1111/j.1476-5381.2012.02018.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 90.den Dekker E, Hoenderop JG, Nilius B, Bindels RJ. The epithelial calcium channels, TRPV5 & TRPV6: from identification towards regulation. Cell Calcium. 2003;33:497–507. doi: 10.1016/s0143-4160(03)00065-4. [DOI] [PubMed] [Google Scholar]
- 91.Trevisani M, Siemens J, Materazzi S, Bautista DM, Nassini R, Campi B, Imamachi N, Andre E, Patacchini R, Cottrell GS, Gatti R, Basbaum AI, Bunnett NW, Julius D, Geppetti P. 4-Hydroxynonenal, an endogenous aldehyde, causes pain and neurogenic inflammation through activation of the irritant receptor TRPA1. Proc Natl Acad Sci U S A. 2007;104:13519–24. doi: 10.1073/pnas.0705923104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Aubdool AA, Kodji X, Abdul-Kader N, Heads R, Fernandes ES, Bevan S, Brain SD. TRPA1 activation leads to neurogenic vasodilatation: involvement of reactive oxygen nitrogen species in addition to CGRP and NO. Br J Pharmacol. 2016;173:2419–33. doi: 10.1111/bph.13519. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Casey DB, Pankey EA, Badejo AM, Bueno FR, Bhartiya M, Murthy SN, Uppu RM, Nossaman BD, Kadowitz PJ. Peroxynitrite has potent pulmonary vasodilator activity in the rat. Can J Physiol Pharmacol. 2012;90:485–500. doi: 10.1139/y2012-012. [DOI] [PubMed] [Google Scholar]
- 94.Zhao KS, Liu J, Yang GY, Jin C, Huang Q, Huang X. Peroxynitrite leads to arteriolar smooth muscle cell membrane hyperpolarization and low vasoreactivity in severe shock. Clin Hemorheol Microcirc. 2000;23:259–67. [PubMed] [Google Scholar]
- 95.Villa LM, Salas E, Darley-Usmar VM, Radomski MW, Moncada S. Peroxynitrite induces both vasodilatation and impaired vascular relaxation in the isolated perfused rat heart. Proc Natl Acad Sci U S A. 1994;91:12383–7. doi: 10.1073/pnas.91.26.12383. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Takahashi N, Kuwaki T, Kiyonaka S, Numata T, Kozai D, Mizuno Y, Yamamoto S, Naito S, Knevels E, Carmeliet P, Oga T, Kaneko S, Suga S, Nokami T, Yoshida J, Mori Y. TRPA1 underlies a sensing mechanism for O2. Nature chemical biology. 2011;7:701–11. doi: 10.1038/nchembio.640. [DOI] [PubMed] [Google Scholar]
- 97.So K, Tei Y, Zhao M, Miyake T, Hiyama H, Shirakawa H, Imai S, Mori Y, Nakagawa T, Matsubara K, Kaneko S. Hypoxia-induced sensitisation of TRPA1 in painful dysesthesia evoked by transient hindlimb ischemia/reperfusion in mice. Sci Rep. 2016;6:23261. doi: 10.1038/srep23261. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Tsai AG, Johnson PC, Intaglietta M. Oxygen gradients in the microcirculation. Physiological reviews. 2003;83:933–63. doi: 10.1152/physrev.00034.2002. [DOI] [PubMed] [Google Scholar]
- 99.Hatano N, Itoh Y, Suzuki H, Muraki Y, Hayashi H, Onozaki K, Wood IC, Beech DJ, Muraki K. Hypoxia-inducible factor-1alpha (HIF1alpha) switches on transient receptor potential ankyrin repeat 1 (TRPA1) gene expression via a hypoxia response element-like motif to modulate cytokine release. J Biol Chem. 2012;287:31962–72. doi: 10.1074/jbc.M112.361139. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Grandl J, Hu H, Bandell M, Bursulaya B, Schmidt M, Petrus M, Patapoutian A. Pore region of TRPV3 ion channel is specifically required for heat activation. Nature neuroscience. 2008;11:1007–13. doi: 10.1038/nn.2169. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Moqrich A, Hwang SW, Earley TJ, Petrus MJ, Murray AN, Spencer KS, Andahazy M, Story GM, Patapoutian A. Impaired thermosensation in mice lacking TRPV3, a heat and camphor sensor in the skin. Science. 2005;307:1468–72. doi: 10.1126/science.1108609. [DOI] [PubMed] [Google Scholar]
- 102.Xu H, Delling M, Jun JC, Clapham DE. Oregano, thyme and clove-derived flavors and skin sensitizers activate specific TRP channels. Nature neuroscience. 2006;9:628–35. doi: 10.1038/nn1692. [DOI] [PubMed] [Google Scholar]
- 103.Nilius B, Biro T, Owsianik G. TRPV3: time to decipher a poorly understood family member! J Physiol. 2014;592:295–304. doi: 10.1113/jphysiol.2013.255968. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Bang S, Yoo S, Yang TJ, Cho H, Hwang SW. Farnesyl pyrophosphate is a novel pain-producing molecule via specific activation of TRPV3. J Biol Chem. 2010;285:19362–71. doi: 10.1074/jbc.M109.087742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Bang S, Yoo S, Yang TJ, Cho H, Hwang SW. Isopentenyl pyrophosphate is a novel antinociceptive substance that inhibits TRPV3 and TRPA1 ion channels. Pain. 2011;152:1156–64. doi: 10.1016/j.pain.2011.01.044. [DOI] [PubMed] [Google Scholar]
- 106.Cheng W, Yang F, Liu S, Colton CK, Wang C, Cui Y, Cao X, Zhu MX, Sun C, Wang K, Zheng J. Heteromeric heat-sensitive transient receptor potential channels exhibit distinct temperature and chemical response. J Biol Chem. 2012;287:7279–88. doi: 10.1074/jbc.M111.305045. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Earley S. Vanilloid and melastatin transient receptor potential channels in vascular smooth muscle. Microcirculation. 2010;17:237–49. doi: 10.1111/j.1549-8719.2010.00026.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Sikaris KA. The clinical biochemistry of obesity. Clin Biochem Rev. 2004;25:165–81. [PMC free article] [PubMed] [Google Scholar]
- 109.Karttunen S, Duffield M, Scrimgeour NR, Squires L, Lim WL, Dallas ML, Scragg JL, Chicher J, Dave KA, Whitelaw ML, Peers C, Gorman JJ, Gleadle JM, Rychkov GY, Peet DJ. Oxygen-dependent hydroxylation by FIH regulates the TRPV3 ion channel. J Cell Sci. 2015;128:225–31. doi: 10.1242/jcs.158451. [DOI] [PubMed] [Google Scholar]
- 110.Cao X, Yang F, Zheng J, Wang K. Intracellular proton-mediated activation of TRPV3 channels accounts for the exfoliation effect of alpha-hydroxyl acids on keratinocytes. J Biol Chem. 2012;287:25905–16. doi: 10.1074/jbc.M112.364869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Earley S, Heppner TJ, Nelson MT, Brayden JE. TRPV4 forms a novel Ca2+ signaling complex with ryanodine receptors and BKCa channels. Circ Res. 2005;97:1270–9. doi: 10.1161/01.RES.0000194321.60300.d6. [DOI] [PubMed] [Google Scholar]
- 112.Earley S, Pauyo T, Drapp R, Tavares MJ, Liedtke W, Brayden JE. TRPV4-dependent dilation of peripheral resistance arteries influences arterial pressure. Am J Physiol Heart Circ Physiol. 2009;297:H1096–102. doi: 10.1152/ajpheart.00241.2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Watanabe H, Davis JB, Smart D, Jerman JC, Smith GD, Hayes P, Vriens J, Cairns W, Wissenbach U, Prenen J, Flockerzi V, Droogmans G, Benham CD, Nilius B. Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J Biol Chem. 2002;277:13569–77. doi: 10.1074/jbc.M200062200. [DOI] [PubMed] [Google Scholar]
- 114.Liedtke W, Choe Y, Marti-Renom MA, Bell AM, Denis CS, Sali A, Hudspeth AJ, Friedman JM, Heller S. Vanilloid receptor-related osmotically activated channel (VR-OAC), a candidate vertebrate osmoreceptor. Cell. 2000;103:525–35. doi: 10.1016/s0092-8674(00)00143-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Matthews BD, Thodeti CK, Tytell JD, Mammoto A, Overby DR, Ingber DE. Ultra-rapid activation of TRPV4 ion channels by mechanical forces applied to cell surface beta1 integrins. Integr Biol (Camb) 2010;2:435–42. doi: 10.1039/c0ib00034e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature. 2003;424:434–8. doi: 10.1038/nature01807. [DOI] [PubMed] [Google Scholar]
- 117.Fernandes J, Lorenzo IM, Andrade YN, Garcia-Elias A, Serra SA, Fernandez-Fernandez JM, Valverde MA. IP3 sensitizes TRPV4 channel to the mechano- and osmotransducing messenger 5'-6'-epoxyeicosatrienoic acid. J Cell Biol. 2008;181:143–55. doi: 10.1083/jcb.200712058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Strotmann R, Schultz G, Plant TD. Ca2+-dependent potentiation of the nonselective cation channel TRPV4 is mediated by a C-terminal calmodulin binding site. J Biol Chem. 2003;278:26541–9. doi: 10.1074/jbc.M302590200. [DOI] [PubMed] [Google Scholar]
- 119.Watanabe H, Vriens J, Suh SH, Benham CD, Droogmans G, Nilius B. Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells. J Biol Chem. 2002;277:47044–51. doi: 10.1074/jbc.M208277200. [DOI] [PubMed] [Google Scholar]
- 120.Sonkusare SK, Bonev AD, Ledoux J, Liedtke W, Kotlikoff MI, Heppner TJ, Hill-Eubanks DC, Nelson MT. Elementary Ca2+ signals through endothelial TRPV4 channels regulate vascular function. Science. 2012;336:597–601. doi: 10.1126/science.1216283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Bubolz AH, Mendoza SA, Zheng X, Zinkevich NS, Li R, Gutterman DD, Zhang DX. Activation of endothelial TRPV4 channels mediates flow-induced dilation in human coronary arterioles: role of Ca2+ entry and mitochondrial ROS signaling. Am J Physiol Heart Circ Physiol. 2012;302:H634–42. doi: 10.1152/ajpheart.00717.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Lee EJ, Shin SH, Hyun S, Chun J, Kang SS. Mutation of a putative S-nitrosylation site of TRPV4 protein facilitates the channel activates. Animal Cells Syst (Seoul) 2011;15:95–106. doi: 10.1080/19768354.2011.555183. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Suresh K, Servinsky L, Reyes J, Baksh S, Undem C, Caterina M, Pearse DB, Shimoda LA. Hydrogen peroxide-induced calcium influx in lung microvascular endothelial cells involves TRPV4. Am J Physiol Lung Cell Mol Physiol. 2015;309:L1467–77. doi: 10.1152/ajplung.00275.2015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Madamanchi NR, Vendrov A, Runge MS. Oxidative stress and vascular disease. Arterioscler Thromb Vasc Biol. 2005;25:29–38. doi: 10.1161/01.ATV.0000150649.39934.13. [DOI] [PubMed] [Google Scholar]
- 125.Vivekananthan DP, Penn MS, Sapp SK, Hsu A, Topol EJ. Use of antioxidant vitamins for the prevention of cardiovascular disease: meta-analysis of randomised trials. Lancet. 2003;361:2017–23. doi: 10.1016/S0140-6736(03)13637-9. [DOI] [PubMed] [Google Scholar]


