Europe PMC

This website requires cookies, and the limited processing of your personal data in order to function. By using the site you are agreeing to this as outlined in our privacy notice and cookie policy.

Abstract 


The mammalian genome encodes 28 distinct members of the transient receptor potential (TRP) superfamily of cation channels, which exhibit varying degrees of selectivity for different ionic species. Multiple TRP channels are present in all cells and are involved in diverse aspects of cellular function, including sensory perception and signal transduction. Notably, TRP channels are involved in regulating vascular function and pathophysiology, the focus of this review. TRP channels in vascular smooth muscle cells participate in regulating contractility and proliferation, whereas endothelial TRP channel activity is an important contributor to endothelium-dependent vasodilation, vascular wall permeability, and angiogenesis. TRP channels are also present in perivascular sensory neurons and astrocytic endfeet proximal to cerebral arterioles, where they participate in the regulation of vascular tone. Almost all of these functions are mediated by changes in global intracellular Ca(2+) levels or subcellular Ca(2+) signaling events. In addition to directly mediating Ca(2+) entry, TRP channels influence intracellular Ca(2+) dynamics through membrane depolarization associated with the influx of cations or through receptor- or store-operated mechanisms. Dysregulation of TRP channels is associated with vascular-related pathologies, including hypertension, neointimal injury, ischemia-reperfusion injury, pulmonary edema, and neurogenic inflammation. In this review, we briefly consider general aspects of TRP channel biology and provide an in-depth discussion of the functions of TRP channels in vascular smooth muscle cells, endothelial cells, and perivascular cells under normal and pathophysiological conditions.

Free full text 


Logo of physrevLink to Publisher's site
Physiol Rev. 2015 Apr; 95(2): 645–690.
PMCID: PMC4551213
PMID: 25834234

Transient Receptor Potential Channels in the Vasculature

Abstract

The mammalian genome encodes 28 distinct members of the transient receptor potential (TRP) superfamily of cation channels, which exhibit varying degrees of selectivity for different ionic species. Multiple TRP channels are present in all cells and are involved in diverse aspects of cellular function, including sensory perception and signal transduction. Notably, TRP channels are involved in regulating vascular function and pathophysiology, the focus of this review. TRP channels in vascular smooth muscle cells participate in regulating contractility and proliferation, whereas endothelial TRP channel activity is an important contributor to endothelium-dependent vasodilation, vascular wall permeability, and angiogenesis. TRP channels are also present in perivascular sensory neurons and astrocytic endfeet proximal to cerebral arterioles, where they participate in the regulation of vascular tone. Almost all of these functions are mediated by changes in global intracellular Ca2+ levels or subcellular Ca2+ signaling events. In addition to directly mediating Ca2+ entry, TRP channels influence intracellular Ca2+ dynamics through membrane depolarization associated with the influx of cations or through receptor- or store-operated mechanisms. Dysregulation of TRP channels is associated with vascular-related pathologies, including hypertension, neointimal injury, ischemia-reperfusion injury, pulmonary edema, and neurogenic inflammation. In this review, we briefly consider general aspects of TRP channel biology and provide an in-depth discussion of the functions of TRP channels in vascular smooth muscle cells, endothelial cells, and perivascular cells under normal and pathophysiological conditions.

I. INTRODUCTION

Since the initial discovery and characterization of the transient receptor potential (trp) channel in Drosophila phototransduction mutants, a growing body of work has demonstrated the diversity, tissue distribution, and remarkable range of functions performed by these cation channels in mammals. Twenty-eight mammalian TRP homologs, distributed throughout the body, have been described. TRP channels are critically involved in sensory and signal transduction processes in excitable and nonexcitable cells present in virtually every organ system. A common theme regarding the specific physiological roles of these channels is their significant contribution to the regulation of intracellular Ca2+ concentration ([Ca2+]i) and distribution, which directly influence a multitude of cellular functions. The biophysics, molecular and cellular biology, and pathophysiological involvement of these “truly remarkable proteins” have recently been reviewed in considerable depth (95, 249).

The importance of TRP channels in the cardiovascular system has been clearly demonstrated over the past 15 years. TRP channels in the heart are involved in normal pacemaker function and contractility and contribute to mechanisms underlying pathologies such as cardiac hypertrophy, fibrotic disease, and arrhythmias. In the circulatory system, TRP channels are present and functional in smooth muscle cells and endothelial cells constituting the vascular wall, and are also expressed in perivascular neurons and astrocytes, which are closely associated with cerebral parenchymal arterioles. TRP channels contribute to mechanosensation and G protein-coupled receptor (GPCR)-initiated signaling pathways that modulate vasoconstrictor and vasodilator activity and cellular proliferation. In vascular endothelial cells, Ca2+ entry through TRP channels is an important contributor to endothelium-dependent vasodilation, vascular wall permeability, and angiogenesis. TRP channels present in perivascular sensory nerves and astrocytes participate in vasodilator responses in some tissues. Altered TRP channel function has been linked to a number of common vascular disorders, including hypertension, vascular occlusive disease, neurogenic inflammation, neointimal injury, and pulmonary edema.

This review is organized into three main sections. The first provides a brief overview of TRP channel structure and function. The second considers the normal physiological roles of TRP channels in vascular smooth muscle cells, endothelial cells, and perivascular cells (i.e., neurons and astrocytes). And the last reviews evidence supporting the involvement of TRP channels in vascular disease.

II. A BRIEF REVIEW OF TRP CHANNELS

A. Discovery

The discovery of the TRP superfamily can be traced to Cosens and Manning's seminal study of Drosophila phototransduction mutants that were able to navigate normally under low light conditions but behaved as if blind under bright lights (63). These flies also displayed characteristic abnormalities in electrical responses in the retina during prolonged light exposure. Minke et al. (224) further characterized the electrophysiological properties of photoreceptors from these mutants and reported that, in contrast to the steady-state depolarization induced by bright light recorded from wild-type flies, light-induced depolarization of the photoreceptor membrane potential in mutant flies was transient in nature. Consequently, the strain was designated transient receptor potential (trp) (224). Isolation of DNA encoding a portion of the affected gene by Montell et al. (230) and subsequent cloning of the full-length trp cDNA by Montell and Rubin (231) and Wong et al. (394) provided additional insight into the function of the trp gene product. Hydropathy plots of the predicted amino acid sequence suggested that trp encoded a membrane protein with intracellular NH2 and COOH termini and six to eight putative transmembrane domains. Montell et al. (231) noted that the membrane topology of the protein encoded by trp was reminiscent of previously characterized voltage-dependent ion channel proteins, and predicted that trp also encoded an ion channel. Vaca et al. (366) validated this hypothesis, reporting an electrophysiological analysis of trp and the related gene trp-like (trpl) in a heterologous expression system. When expressed in Sf9 insect cells, trp formed a Ca2+-permeable, nonselective cation channel that was activated by the sarco(endo)plasmic ATPase (SERCA) inhibitor thapsigargin. Evidence that trp homologs are present in vertebrates was reported almost simultaneously by Petersen et al. (268), who found that sequences related to trp are present in Xenopus oocytes and mouse brain. More comprehensive studies were performed by Wes et al. (388) and Zhu et al. (425), who cloned and characterized the first mammalian trp homologs, and Zhu et al. (426), who reported the first evidence of a receptor-operated mammalian TRP channel. The first mammalian TRP to be cloned is now known as TRPC1. Wes et al. (388) were prescient in predicting a larger role for multiple trp homologs in humans, concluding that “it will be interesting to determine the full size of the human TRP family and whether the human TRPs have different functional characteristics . . .” Indeed, the discovery of the first human TRPs spurred a decade-long race to clone all of the TRP genes and characterize the functional properties of the encoded proteins in heterologous expression systems. All members of the TRP superfamily now appear to be accounted for, and current efforts are focused on understanding how TRP channels affect the physiology and pathophysiology of different organs and tissues. Detailed insight into the early days of TRP channel research is provided by reviews and commentaries by the pioneers in this field (130, 223, 229).

B. TRP Superfamily Organization and Channel Structure

Mammalian genomes encode 28 distinct TRP protein subunits, which together comprise the TRP superfamily (Figure 1, Table 1). A standard nomenclature was established (61), grouping these genes and their products into six subfamilies based on amino acid sequence homology. These subfamilies are designated canonical (TRPC), vanilloid (TRPV), melastatin (TRPM), ankyrin (TRPA), mucolipin (TRPML), and polycystin (TRPP) (Figure 1, Table 1). One member of the TRPC subfamily, TRPC2, is present in rodents and is important for pheromone sensing (206), but is a nonexpressed pseudogene in humans (409).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240001.jpg

Phylogenetic tree of the mouse TRP superfamily.

Table 1.

The human transient receptor potential cation channel superfamily

TRPGene NameDescriptionUniGene IDChromosome
TRPC1TRPC1Transient receptor potential cation channel, subfamily C, member 1Hs.2506873
TRPC2TRPC2Transient receptor potential cation channel, subfamily C, member 2Hs.13191017
TRPC3TRPC3Transient receptor potential cation channel, subfamily C, member 3Hs.1509814
TRPC4TRPC4Transient receptor potential cation channel, subfamily C, member 4Hs.26296013
TRPC5TRPC5Transient receptor potential cation channel, subfamily C, member 5Hs.657709X
TRPC6TRPC6Transient receptor potential cation channel, subfamily C, member 6Hs.1590031
TRPC7TRPC7Transient receptor potential cation channel, subfamily C, member 7Hs.5912635
TRPV1TRPV1Transient receptor potential cation channel, subfamily V, member 1Hs.57921717
TRPV2TRPV2Transient receptor potential cation channel, subfamily V, member 2Hs.27974617
TRPV3TRPV3Transient receptor potential cation channel, subfamily V, member 3Hs.44625517
TRPV4TRPV4Transient receptor potential cation channel, subfamily V, member 4Hs.50671312
TRPV5TRPV5Transient receptor potential cation channel, subfamily V, member 5Hs.2833697
TRPV6TRPV6Transient receptor potential cation channel, subfamily V, member 6Hs.3027407
TRPA1TRPA1Transient receptor potential cation channel, subfamily A, member 1Hs.6671568
TRPM1TRPM1Transient receptor potential cation channel, subfamily M, member 1Hs.15594215
TRPM2TRPM2Transient receptor potential cation channel, subfamily M, member 2Hs.36975921
TRPM3TRPM3Transient receptor potential cation channel, subfamily M, member 3Hs.472889
TRPM4TRPM4Transient receptor potential cation channel, subfamily M, member 4Hs.46710119
TRPM5TRPM5Transient receptor potential cation channel, subfamily M, member 5Hs.27228711
TRPM6TRPM6Transient receptor potential cation channel, subfamily M, member 6Hs.2722259
TRPM7TRPM7Transient receptor potential cation channel, subfamily M, member 7Hs.51289415
TRPM8TRPM8Transient receptor potential cation channel, subfamily M, member 8Hs.3660532
TRPP1PKD2Polycystic kidney disease 2 (autosomal dominant)Hs.1812724
TRPP2PKD2L1Polycystic kidney disease 2-like 1Hs.15924110
TRPP3PKD2L2Polycystic kidney disease 2-like 2Hs.5674535
TRPML1MCOLN1Mucolipin 1Hs.63185819
TRPML2MCOLN2Mucolipin 2Hs.5914461
TRPML3MCOLN3Mucolipin 3Hs.5352391

The terminology used to describe the TRPP subfamily requires some clarification. The polycystin-1 protein encoded by PKD1 is sometimes referred to as TRPP1 (or TRPP4). This is confusing, as PKD1 encodes a protein with 11 transmembrane domains and a long extracellular NH2 terminus that does not form an ion-conducting pathway and is structurally unrelated to authentic TRP channels (390). The name “TRPP4” was also briefly used to describe the product of the PKDREJ (polycystic kidney disease and receptor for egg jelly related protein) gene in an early report. But PKDREJ is also unrelated to the mammalian TRP superfamily, and the name “TRPP4” is no longer in use. TRPP2, TRPP3, and TRPP5 have also been used to describe the products of the PKD2 (polycystic kidney disease 2) gene (gene ID: 5,311; mRNA: NM_000297), PKD2L1 gene (gene ID: 9,033; mRNA: NM_016112), and PKD2L2 gene (gene ID: 27,039; mRNA NM_014386), respectively. In this review, we use TRPP1 for PKD2, TRPP2 for PKD2L1, and TRPP3 for PKD2L2. The current International Union of Basic and Clinical Pharmacology (IUPHAR) database (http://www.guidetopharmacology.org/GRAC/FamilyDisplayForward?familyId=78) currently uses both nomenclatures, adding to confusion.

All TRP subunits are expressed as polypeptides of 553–2,022 amino acids (~64–230 kDa) containing six transmembrane domains (S1–S6) and intracellular NH2 and COOH termini of variable length (Figure 2). An ~25-amino acid conserved element called the TRP domain is present immediately distal to the S6 domain in members of the TRPC, TRPM, and TRPV subfamilies. The TRP domain contains two highly conserved, six-residue TRP boxes bordering a central region that has greater sequence variability. The function of the TRP domain is not completely understood, but it may be involved in phosphatidylinositol 4,5-bisphosphate (PIP2) binding (287) or subunit assembly (105). Multiple regulatory elements, interaction sites, and enzymatic domains are present in the intracellular termini. Numerous ankyrin repeat domains are present on the NH2 terminus of TRPC, TRPV, and TRPA subunits. These are particularly prominent in TRPA1 subunits (~15–19 repeats), where they contribute to channel regulation. The COOH terminus of TRPC channels encodes calmodulin/IP3R-binding (CIRB) domains, Ca2+-binding EF hands and, for TRPC4, a PDZ domain. A domain known as the TRPM homology region is present in the NH2 terminus of TRPM subfamily members. Functional serine-threonine kinase domains are present in the COOH terminus of TRPM6 and TRPM7, and an ADP-ribose-binding NUDIX phosphohydrolase domain is present in the COOH-terminal region of TRPM2. TRPP and TRPML channels share homology through the transmembrane region, including a large extracellular loop between the S1 and S2 domains that is not present in other subfamilies. Transcriptional splice variants, some with functionally distinct properties, have been described for almost all TRP subunits (373).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240002.jpg

Membrane topology and functional domains of TRP subunits.

Functional TRP channels arise from the assembly of four subunits, with the S5 and S6 transmembrane domains forming the ion-conducting pore. All TRP subunits, with the possible exception of TRPC1 (328), are capable of forming functional homomeric cation channels. Many cell types express multiple TRP subunits, and heteromultimeric channels composed of two or three related subunits can form and are present in vivo. Heteromultimeric channels display conduction properties and regulatory mechanisms that are distinct from those of homomeric channels. Formation of heteromultimeric channels is best characterized for TRPC channel subunits. TRPC1 subunits can form heteromultimeric channels with TRPC3, TRPC4, TRPC5, TRPC6, and TRPC7 subunits. TRPC4 and TRPC5 can heteromultimerize, as can combinations of TRPC3, TRPC6, and TRPC7 subunits (141). Other arrangements include channels composed of the closely related subunits TRPV5 and TRPV6 (138), TRPM6 and TRPM7 (186), and TRPML1 and TRPML3 (367). There is less evidence supporting interactions between channels from different subfamilies, although formation of TRPC1/TRPV4 heteromultimeric channels in the endothelium (207), TRPV1/TRPA1 channels in sensory nerves (93, 290), and other combinations (263) has been reported.

The most detailed study of TRP channel structure to date was provided by Liao et al. (192) and Cao et al. (43), who used cryoelectron microscopy to determine the structure of TRPV1 channels at high resolution (3.4 Å). These studies revealed that four TRPV1 subunits are symmetrically arranged around a central ion-permeable pore in a configuration reminiscent of previously characterized voltage-gated ion channels. Six transmembrane α-helices (S1–S6) span the lipid bilayer, with a pore-forming loop between the S5 and S6 domains. The TRP domain forms an interfacial helix distal to the S6 domain that is positioned to interact with the region preceding the S1 domain and the linker region between the S4 and S5 domains. The outer pore of the assembled channel is very wide. The selectivity filter is short and narrow in the inactive configuration, but dilates substantially when the channel is activated by capsaicin. This work confirmed the general structural features of TRP channels that had been derived using lower-resolution cryoelectron microscopy methods in studies of TRPV1 (227), TRPV4 (313), and TRPC3 (225) channels. Not surprisingly, these studies, as well as the few available crystallography analyses of the ankyrin repeat domain of TRPV2 (150, 215), revealed specific structural differences in both intracellular and extracellular domains that are clearly associated with the unique signaling partners and sensory modalities served by individual TRP channel family members.

C. Functions in Mammalian Cells

TRP channels influence numerous important biological processes by controlling the flux of cations across the plasma membrane. Several TRP family members, including TRPP2, TRPM8, TRPV1, and TRPA1, are also present and active on the membranes of intracellular organelles and are involved in protein trafficking and vesicular ionic homeostasis (see Ref. 71 for review). Cation permeability has substantial biological significance. The distribution of the monovalent cations Na+ and K+ is particularly important for the regulation of cellular membrane potential and excitability. The divalent cations Ca2+ and Mg2+ also function as intracellular second messengers and as required cofactors for enzymatic activity. Although TRP channels are often described as “nonselective cation channels,” this definition is too broad and rather imprecise. TRPV5 and TRPV6 channels are highly selective for Ca2+, with relative permeability of Ca2+ versus Na+ (PCa:PNa) of ~100:1 (253, 368, 415). TRPM4 (179, 251) and TRPM5 channels (139, 276) are selective for monovalent cations (Na+ and K+) and are essentially impermeant to Ca2+ and other divalent cations. Ion substitution experiments suggest that the remaining TRP channels are permeant to both mono- and divalent cations, but with ionic preference ranging from essentially nonselective (i.e., TRPC1; Ref. 334) to significantly selective for Ca2+ (e.g., TRPV3; PCa:PNa ~12:1; Ref. 401) or preferentially selective for Mg2+ under physiological conditions (TRPM6 and TRPM7) (186). Selectivity for particular monovalent species also varies, with certain channels favoring Na+ versus K+ (e.g., TRPC3, TRPC6) and others with the opposite bias (e.g., TRPM4, TRPV4). This property is significant, since the reversal potential of channels with high K+ permeability occurs at negative membrane potentials in physiological ionic gradients; thus activation of these channels will have only a small direct effect on membrane potential depolarization. Membrane depolarization is greater for channels with a higher Na+ conductance, which diminishes the electrochemical driving force for Ca2+ influx in nonexcitable cells. Ultimately, this leads to channels with greater K+ permeability (e.g., TRPV4) conducting mixed currents having a higher fraction of Ca2+ compared with channels with higher Na+ permeability (e.g., TRPC6). Among the channels that conduct currents with a significant (>20%) Ca2+ fraction under physiological conditions are TRPV1-4 and TRPA1; in contrast, the fractional contribution of Ca2+ to TRPC channel currents is expected to be very low. Native currents are often not apparent from patch-clamp records reported in the literature, since these experiments are usually performed using a symmetrical cationic distribution. Few studies have actually attempted to record and characterize mixed currents under physiological ionic conditions.

1. Membrane potential regulation

Asymmetric concentration gradients maintained by cellular ion transporters and pumps establish the resting membrane potential. TRP channels contribute to membrane potential regulation in both excitable cells (i.e., those expressing voltage-gated Na+ or Ca2+ channels), such as neurons, glia, myocytes, and endocrine cells, and nonexcitable cells (i.e., those lacking voltage-gated ion channels), such as epithelial and endothelial cells, by regulating the flux of cations across the plasma membrane.

TRPM4 channels have a well-documented influence on membrane potential regulation in excitable and nonexcitable cells. These channels have a unitary conductance of ~25 pS, are selective for monovalent cations and impermeant to Ca2+ ions, and are activated by high levels of intracellular Ca2+ (179). TRPM4 channels are equally permeant to Na+ and K+, resulting in a reversal potential of ~0 mV in physiological solutions. At the hyperpolarized resting membrane potentials that are typical for most cells, these channels conduct inward currents that are primarily composed of Na+ when activated by global or localized increases in [Ca2+]i, resulting in membrane depolarization. TRPM4-mediated Na+ currents influence the membrane excitability of pancreatic β-cells (53), detrusor (258, 321, 322) and cerebral artery (64, 76, 83, 104, 114, 115, 117, 118) smooth muscle cells, atrial cardiomyocytes (317), and pre-Botzinger neurons (226). In addition, membrane potential regulation in nonexcitable mast (369) and dendritic cells (26) is disrupted in TRPM4−/− mice. Membrane hyperpolarization associated with TRPM4 knockout and loss of depolarizing Na+ currents increases the driving force for Ca2+ influx, resulting in Ca2+ overload and impaired chemokine-dependent migration of dendritic cells (26).

2. Ca2+ signaling

Changes in the intracellular concentration of free Ca2+ influence muscle contraction, hormone secretion, gene expression, cell proliferation, and many other critical processes (32). The cytosolic free [Ca2+] is normally maintained at low levels (~100–300 nM) by stationary and mobile Ca2+-buffering proteins and by the activity of Ca2+-pumping ATPases located on the plasma membrane and other cellular compartments. Changes in intracellular Ca2+ occur as a result of the influx of Ca2+ from the extracellular space and/or through the release of Ca2+ from intracellular structures. The endoplasmic/sarcoplasmic reticulum (ER/SR) is the best characterized intracellular Ca2+ storage organelle, but the membranes of mitochondria, nuclei, and other structures also delimit regions with high [Ca2+]. Ca2+ signals can be global, occurring throughout the cytosol, or can be localized, creating subcellular domains where [Ca2+] is transiently elevated. Almost all TRP channels are permeable to Ca2+, and several have a strong influence on Ca2+-dependent signaling pathways (62). TRP channels have been described as direct (i.e., “ionotropic”) Ca2+ influx pathways, metabotropic (receptor-operated) Ca2+ influx channels, or store-operated Ca2+ channels, although this latter property is disputed.

a) trp channels as ionotropic ca2+ influx pathways: trp sparklets. Elementary Ca2+ signals representing Ca2+ influx through single TRPV4 channels, called TRPV4 sparklets, have been recorded from airway (423) and vascular smooth muscle cells (221), endothelial cells (337), and the intact endothelium (324) (for review, see Ref. 336). Recordings of TRPA1 sparklets have also been reported for cerebral artery endothelial cells (338). Theoretically, all TRP channels with sufficient unitary conductance and Ca2+ permeability (i.e., TRPV1-6; TRPA1) should produce detectable sparklets. TRPV4 and TRPA1 sparklets are optically recorded from cells expressing Ca2+ biosensors or loaded with Ca2+-indicator dyes using total internal reflection fluorescent (TIRF) or confocal microscopy (Figure 3). Individual TRPV4 sparklets are very large Ca2+ influx events, with the amount of Ca2+ entering estimated to be ~100 times that of a single L-type Ca2+ channel sparklet (221). The unitary amplitude of TRPA1 sparklets (338) is approximately twice that of a TRPV4 sparklet recorded under identical conditions (337), suggesting that more Ca2+ enters during the opening of a single TRPA1 channel than is the case for a single TRPV4 channel. TRPV4 sparklets generate subcellular microdomains in which local Ca2+ levels are very high, leading to activation of a variety of Ca2+-dependent signaling cascades, including endothelium-dependent vasodilation (324), negative feedback inhibition of vasoconstrictor stimuli (221), and NFATc3-dependent smooth muscle proliferation (423). TRPA1 sparklets are linked to endothelium-dependent dilation of cerebral arteries (338).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240003.jpg

TRPV4 sparklets in endothelial cells and vascular smooth muscle cells. A: time-lapse image of a typical TRPV4 sparklet recorded from a primary cerebral artery endothelial cell using TIRF microscopy. The cell was stimulated with the selective TRPV4 agonist GSK1016790A (GSK; 100 nM). B: TRPV4 sparklets recorded from a tsA-201 cell transfected with TRPV4 (top) and from a native cerebral artery myocyte (bottom). The middle and right panels demonstrate the single channel-like behavior of these events. TRPV4 sparklets were stimulated with GSK. C: TRPV4 sparklets recorded from the intact endothelium of cerebral arteries from mice expressing the Ca2+-indicator protein GCaMP2 exclusively in the endothelium using high-speed, high-resolution confocal microscopy before and after stimulation with GSK. Experiments were performed in the presence of cyclopiazonic acid (CPA) to eliminate interference from ER Ca2+-release events. Representative traces indicating single channel-like events are shown below. [A from Sullivan and Earley (336). B from Mercado et al. (221). C from Sonkusare et al. (323), with permission from American Association for the Advancement of Science.]

b) trp channels as receptor-operated channels. Activation of specific GPCRs on the cell surface causes a biphasic change in global [Ca2+]i, with an initial transient increase followed by a sustained plateau phase (Figure 4A) (for review, see Ref. 32). The transient Ca2+ signal results from release of Ca2+ from the ER/SR, whereas influx of Ca2+ through receptor-operated channels (ROCs) is responsible the sustained phase (Figure 4B). This latter mode of Ca2+ influx is called “receptor operated Ca2+ entry” (ROCE). TRPC3, TRPC6, and TRPC7 channels have been characterized as ROCs that are activated by GPCRs linked to phospholipase C (PLC) (144, 199, 266, 323). In response to receptor stimulation, PLC cleaves the membrane phospholipid PIP2 into the second messenger molecules inositol 1,4,5-trisphosphate (IP3) and diacylglycerol (DAG), the latter of which activates protein kinase C (PKC). IP3 binds to IP3 receptors (IP3Rs) present on the ER/SR to cause Ca2+ release from intracellular stores. DAG is a direct (i.e., independent of PKC) activator of TRPC3, TRPC6, and TRPC7 channels, as well as heteromultimeric channels that include these subunits (140); ROCE through these channels accounts for sustained elevations in intracellular Ca2+.

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240004.jpg

Ca2+-influx mechanisms mediated by TRP channels. A: receptor-operated Ca2+ entry (ROCE). Agonist binding to GPCRs stimulates the activity of PLC, which cleaves the membrane phospholipid PIP2 into IP3 and DAG. IP3 binds to IP3Rs on the ER/SR to cause release of Ca2+ into the cytosol. DAG directly activates Ca2+ influx through TRPC3, TRPC6, or TRPC7 channels on the plasma membrane. B: example trace of changes in intracellular Ca2+ following stimulation of a GPCR, showing the phasic increase resulting from ER/SR Ca2+ release and sustained elevation resulting from ROCE. C: store-operated Ca2+ entry (SOCE). Depletion of Ca2+ in the ER/SR causes clustering of STIM1 in the ER/SR membrane proximal to the plasma membrane. STIM1 interacts with Oria1 at the plasma membrane to promote Ca2+ influx. TRPC1, TRPC4, and/or TRPC5 channels, either independently or in association with STIM1/Oria1, may participate in SOCE. D: example of a typical SOCE experiment. Under Ca2+-free conditions, cells are treated with the SERCA inhibitor thapsigargin, causing Ca2+ to leak from the ER/SR into the cytosol. Reintroduction of Ca2+ to the bathing solution after ER/SR stores are depleted causes SOCE.

The PLC substrate PIP2 can directly influence, both positively and negatively, the activity of almost all TRP channels. For example, application of PIP2 to excised membrane patches has been shown to directly activate the ROCs discussed above (TRPC3, TRPC6, and TRPC7) in HEK cell expression systems (181). This extensive topic is the subject of several recent reviews (285, 286, 288).

c) trp channels as store-operated channels. Store-operated Ca2+ entry (SOCE) is defined as Ca2+ influx that occurs in response to depletion of intracellular Ca2+ stores, typically accomplished experimentally by prolonged stimulation of cell surface receptors or blocking SERCA activity with pharmacological agents such as thapsigargin or cyclopiazonic acid (260, 278) (Figure 4). Ca2+ directly enters the ER/SR compartment during SOCE, providing a dedicated mechanism for refilling depleted intracellular Ca2+ stores. The ion channels that conduct SOCE are called store-operated channels. A number of early studies suggested that TRPC1, TRPC4, and/or TRPC5 channels are store-operated channels (6, 97, 122, 261, 396). However, the discovery that complexes formed by stromal interacting protein 1 (STIM1) and Orai1 are critically involved in SOCE (for review, see Ref. 21) cast doubt on an obligate role for TRP channels in this response. STIM1 is a single-transmembrane-domain protein present on both the plasma membrane and the ER/SR. The NH2 terminus of STIM1 contains an EF hand Ca2+-sensing domain that resides in the ER/SR lumen and two coiled-coil domains in the cytosolic COOH terminus that mediate homomeric and heteromeric protein-protein interactions. Orai1 is a Ca2+-selective ion channel present on the plasma membrane (275). Depletion of intracellular Ca2+ stores is sensed by the EF hand domain located in the ER/SR lumen and leads to rearrangement and clustering of STIM1 proteins at the surface of the ER/SR proximal to the plasma membrane. STIM1 clusters interact with Orai1 on the plasma membrane, thereby initiating Ca2+ entry. Studies have shown that SOCE can occur in the absence of TRPC1, TRPC4, and TRPC5 channel expression and that TRPC channels can function independently of STIM1 and Orai1 (68), leading some to conclude that TRP channels are not involved in SOCE. However, other studies have reported that TRPC channels are required for SOCE in certain settings, and some investigators have noted that the structure of Orai1 is more reminiscent of a regulatory subunit than classically described ion-permeable channels (393). It has been proposed that Orai1 proteins are a regulatory component of a store-operated channel composed of TRPC1 subunits (or vice versa) (16), but this issue is currently unresolved. Experimental details are important in interpreting SOCE experiments. For example, in whole cell patch-clamp experiments, if cytosolic Ca2+ is not fully buffered, SOCE through the STIM1/Oria1 pathway can cause Ca2+-dependent activation of certain TRP channels that is independent of store depletion. In this context, it has been proposed that SOCE through Orai1 causes TRPC1 channels to traffic to the plasma membrane where they mediate a secondary Ca2+ influx event (55). The precise functional relationships between STIM1, Oria1, and TRPC channels during SOCE remain to be determined.

3. Mg2+ homeostasis

Mg2+ are central to the maintenance of nucleic acid structure, regulation of certain Ca2+ and K+ channels, and the biological activity of ATP and hundreds of enzymes (for review, see Ref. 67). TRPM6 and TRPM7 channels are important pathways for Mg2+ influx (228, 303, 305). TRPM7 has a unitary conductance of ~40–105 pS, is abundantly expressed in many types of cells, and is constitutively active in heterologous expression systems (240). TRPM7 is permeable to both monovalent and divalent cations, although some reports suggest greater permeability of divalent cations under physiological conditions. The channel is inhibited by intracellular [Mg2+] greater than 1 mM. Under physiological conditions, the channel may act as a Mg2+/Ca2+ influx pathway, with intracellular [Mg2+] providing negative feedback. The ion permeation and regulatory properties of TRPM6 are similar to those of TRPM7, but its unitary conductance is greater (~80 pS). TRPM6 can form heteromultimeric channels with TRPM7 with properties distinct from those of the homomeric channels. Significantly, mutations within TRPM6 are associated with familial hypomagnesemia with secondary hypocalcemia (303), and TRPM7 deletion causes Mg2+ deficiency, indicating that TRPM6 and TRPM7 channels are vital for Mg2+ homeostasis. TRPM6 and TRPM7 channels are also permeable to other divalent cations, including the essential elements Zn2+, Co2+, and Mn2+, and toxic metals such as Sr2+, Cd2+, Ba2+, and Ni2+, and may be involved in transporting these ions across the plasma membrane (228).

III. PHYSIOLOGICAL ROLES OF TRP CHANNELS IN THE VASCULATURE

The primary function of the vasculature is to dynamically regulate blood flow and vascular permeability to ensure appropriate matching of oxygen and nutrient supply with the metabolic demands of the tissue. Vascular resistance and blood flow are primarily regulated by the contractile state of smooth muscle in the arterial circulation, but can also be chronically altered by cellular proliferation and vascular remodeling that occur during vascular development and in association with vascular diseases. Endothelial cells have a profound influence on vascular tone and permeability, an influence that can vary between physiological and pathophysiological conditions. Perivascular neurons and astrocytes closely apposed to blood vessels can also influence smooth muscle contractility and blood flow regulation. A growing body of evidence, reviewed below, clearly demonstrates the importance of TRP channels in the initiation, modulation, and integration of regulatory pathways that govern the contributions of these diverse cell types within the vascular wall to control of vasomotor tone and tissue blood flow.

A. TRP Channels in Vascular Smooth Muscle

Luminal diameter and hence vascular resistance are determined by the contractile state of vascular smooth muscle cells, which are arranged circumferentially around the lumen of all blood vessels. Vascular smooth muscle contraction is primarily regulated by the phosphorylation state of a myosin regulatory light chain (RLC) domain in the smooth muscle myosin complex. The net balance of myosin light chain (MLC) kinase versus MLC phosphatase activity determines the phosphorylation state of the RLC. MLC kinase, a Ca2+/calmodulin-dependent enzyme, is activated by increases in cytosolic Ca2+ levels, but its phosphatase activity is regulated by intracellular signaling pathways that serve to modify the sensitivity of the contractile apparatus to intracellular Ca2+. Smooth muscle [Ca2+]i is dynamically regulated by the relative activities of Ca2+-entry, -release, and -extrusion mechanisms. The L-type voltage-dependent Ca2+ channel (VDCC) is the primary Ca2+-influx pathway in vascular smooth muscle cells. Thus membrane depolarization and subsequent influx of Ca2+ through these channels is a primary mechanism of vascular smooth muscle cell contraction (Figure 5). Increased cation permeability associated with TRP channel activation results in membrane depolarization and Ca2+ influx, which are the primary mechanisms by which these channels cause vasoconstriction. Ca2+ influx through TRP channels with significant selectivity for Ca2+ could theoretically directly activate the contractile process, but there is little experimental evidence in support of this possibility in the systemic circulation. However, some data suggest that SOCE through TRP channels can support a vasoconstrictor response in pulmonary arteries (see sect. IIIA2).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240005.jpg

TRP channels contribute to vascular smooth muscle cell contractility. Cation influx through TRP channels depolarizes the plasma membrane (ΔEm), initiating Ca2+ influx through VDCCs. Elevated Ca2+ levels increase binding of calmodulin-Ca2+ (Cd-Ca2+) complexes to MLCK, stimulating phosphorylation of the regulatory myosin light chain (MLC20). MLC20 is dephosphorylated by myosin light-chain phosphatase (MLCP). Some studies suggest that Ca2+ influx through TRP channels can directly initiate myocyte contraction.

1. TRP channels and vascular smooth muscle contractility

Nearly a dozen TRP channels have been detected in vascular smooth muscle using RT-PCR, Western blotting, and immunolabeling approaches (Table 2). Lacking selective pharmacological blockers or activators, early researchers in the field relied on antisense nucleotide-mediated knockdown of TRP channel expression or antibodies that bind to epitopes on extracellular domains of the channels to inhibit cation currents. In some cases, the specificity of these treatments was not fully demonstrated; thus the conclusions drawn from the results of these studies should be carefully considered. In addition, many studies have attempted to study smooth muscle cell contractility using cultured cells, such as the aorta-derived A7r5 transformed cell line. This approach is questionable, as A7r5 cells (and other transformed cell lines) display little or no contractility and are phenotypically distinct from native vascular smooth muscle cells. Primary smooth muscle cells are also problematic since they rapidly (within a few hours) differentiate to a proliferative, noncontractile state under standard culture conditions. Accordingly, conclusions about regulation of smooth muscle cell contractility derived from data obtained using cultured cells should be interpreted with caution. Even now, many studies continue to rely on nonselective pharmacological agents, such as trivalent cations (Gd3+, La3+), 2-aminoethoxydiphenyl borate (2-APB), ruthenium red, and SKF96365 without consideration of the broad range of targets affected by these compounds. Although these experimental tools have their place, knockout animals, small inhibitory RNA (siRNA) approaches, and selective pharmacological agents that have been developed in recent years are more definitive approaches for studying contractile responses.

Table 2.

TRP channels in vascular smooth muscle

TRPFunctionTissue
TRPC1ROCE (ET-1 constriction), SOCEBrain (124), aorta (344), lung (245247)
TRPC3GPCR-mediated vasoconstrictionBrain (4, 5, 283, 398), aorta (198), heart (266), kidney (296), lung (420)
TRPC4SOCEAorta and mesentery (196, 197)
TRPC5SOCEBrain (403)
TRPC6GPCR-mediated vasoconstriction, myogenic tonePortal vein (146), aorta (323), brain (217, 387)
TRPV1VasoconstrictionAorta, skeletal muscle (158, 201), mesentery (273), cephalic circulation (46)
TRPV2Mechanosensitive Ca2+ influxAorta (238)
TRPV3mRNA present in VSM but no known functionAorta and lung (408)
TRPV4Ca2+ influx, vasodilationBrain (79, 221), mesentery (81)
TRPA1VasodilationAorta (405)
TRPM4Depolarization, myogenic toneBrain (64, 82, 83, 104, 114, 115, 117)
TRPM8VasoconstrictionAorta (151)
TRPP2Ca2+ entry, vasoconstriction, myogenic toneMesentery (279, 308), brain (243)

ROCE, receptor-operated Ca2+ entry; SOCE, store-operated Ca2+ entry; ET-1, endothelin-1; GPCR, G protein-coupled receptor. Reference numbers are given in parentheses.

a) trpc6. Compelling evidence indicates that TRPC6 channels are functionally important in arterial myocytes. TRPC6 channel mRNA and protein have been identified in vascular smooth muscle, both arterial and venous, from many species and in many vascular beds (145). In native cells, it is often difficult to determine whether the observed ion channel activity is attributable to a specific TRPC channel isoform, perhaps because many TRPC subunits form heteromultimeric associations, creating a challenge for characterizing endogenous TRPC channels. This is in contrast to expression systems, where the composition of channel isoforms is controlled. Furthermore, TRPC channels are often activated or modulated by diverse cellular signaling mechanisms, the activity and function of which are often altered by cell isolation procedures or the specific patch-clamp configuration (i.e., on-cell, whole-cell, perforated patch) employed. Nevertheless, single-channel activity that likely is determined primarily by TRPC6 channels has been observed in a variety of vascular smooth muscle types, including portal vein (9), mesenteric artery (293), and cultured aortic smooth muscle cells (144). Single-channel conductance values ranging from 33 to 45 pS have been reported in these studies. The presence of functional ion channels in mesenteric (293) myocytes that are composed, at least in part, of TRPC6 subunit proteins has been confirmed through the use of current-blocking antibodies that target TRPC6.

In work growing out of initial observations that TRPC channel activity in expression systems is linked to activation of GPCRs (62), Inoue et al. (146) provided the first evidence that TRPC6 channels are present and functionally active in native vascular smooth muscle. These seminal studies clearly demonstrated that a nonselective cation channel with biophysical and pharmacological properties resembling those of cloned TRPC6 channels expressed in heterologous systems is present in portal vein myocytes. The activity of this portal vein smooth muscle channel was substantially elevated following α1-adrenoceptor activation, and was greatly reduced or was absent in cells treated with TRPC6-targeted antisense oligonucleotides. Subsequent studies have demonstrated that TRPC6 channels in vascular myocytes are activated by a variety of GPCR ligands. For example, nonselective cation channels with a TRPC6 biophysical and pharmacological profile (154), or that are suppressed by TRPC6 siRNA (323), are activated by vasopressin in A7r5 aortic smooth muscle-derived cells. Likewise, vasopressin-induced Ca2+ entry (323) and Ca2+ oscillations (210) in A7r5 cells are suppressed by treatment with TRPC6 siRNA. In freshly isolated rabbit mesenteric artery myocytes, low levels (1 nM) of angiotensin II (ANG II) were shown to activate a nonselective cation channel that was inhibited by TRPC6 antibodies (293). Activation of TRPC6 in these studies followed stimulation of a GPCR and subsequent signaling via PLC and generation of DAG (327). The stimulatory effect of DAG on TRPC6 channel activity appeared to be direct and was independent of PKC activation (137, 140). Furthermore, there are apparent interactions between receptor-mediated and other signaling modalities that result in amplification of vascular TRPC6 channel activity under some conditions. Specifically, transmembrane Ca2+ mobilization through TRPC6 activation is amplified through combined receptor and mechanical stimulation via signaling pathways that involve contributions by both PLC and phospholipase A2 (PLA2) (144). Other examples of vascular TRPC6 coregulation by various cellular signaling molecules have been noted. Synergistic stimulatory effects on vascular TRPC6 channels through coactivation of IP3Rs and DAG (9, 147, 152) or IP3 and arginine vasopressin receptors (213) have been described, and it has been shown that protein kinase A (PKA) activity amplifies the activating effects of ANG II on TRPC6 channels (255). In contrast to findings obtained using a HEK cell expression system (181), PIP2 was shown to inhibit TRPC6 currents in native mesenteric artery smooth muscle cells (13). TRPC6 channels may also be inhibited by interactions with other TRPC channels, including TRPC1/C5 heteromultimeric channels (312).

Functional studies have linked the activation of TRPC6 channels in vascular smooth muscle cells with pressure-induced (myogenic) vasoconstriction. Myogenic constriction, which contributes to the autoregulation of blood flow in some vascular beds (28), is mediated by depolarization of the arterial myocyte plasma membrane and Ca2+ entry through VDCCs (163). TRPC6 channels in vascular smooth muscle cells and expression systems are activated by mechanical forces such as cell swelling (387) (327) or application of suction to the plasma membrane (327). However, the exact mechanism responsible for mechanical activation of TRPC6 is a matter of controversy, with some studies indicating that TRPC6 channels are directly activated by application of force (327) and others reporting indirect activation downstream of mechanosensitive signaling pathways (217). Data from some laboratories point to GPCRs as the primary mechanosensors that activate TRPC6 channels through downstream signaling involving the formation of DAG (217, 302). A study by Welsh et al. (387) demonstrated that swelling-activated cation currents and intravascular pressure-induced smooth muscle cell depolarization and vasoconstriction were substantially reduced in cerebral arteries treated with antisense oligonucleotides targeting TRPC6, suggesting the critical involvement of this channel in the myogenic response in this vascular bed. It should be mentioned that neither agonist-evoked contraction nor myogenic tone were reduced in arteries from TRPC6−/− mice (70, 302). However, upregulation of constitutively active (12) TRPC3 channels in TRPC6−/− mice may provide a compensating depolarization and vasoconstrictor pathway in these animals.

TRPC6 appears to be an integral part of a force-sensing complex in cerebral artery smooth muscle cells that regulates myogenic vasoconstriction (118). Gonzales et al. (118) demonstrated that Ca2+ entry through TRPC6 channels, activated either by PLC-γ1-dependent generation of DAG or by direct mechanical stimulation, reinforces IP3R-dependent Ca2+ release from the SR, which then activates nearby TRPM4 channels. TRPM4 channels, in turn, mediate smooth muscle cell depolarization, promoting Ca2+ influx through VDCCs and subsequent vasoconstriction (see below for more details regarding the role of TRPM4 channels in vasoconstriction) (Figure 6). This model clearly delineates the architecture, interactions, and mechanisms of action of a variety of signaling molecules recognized as critical players in the development and modulation of arterial tone in the brain circulation.

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240006.jpg

A force-sensitive signaling network involving TRPC6 and TRPM4 channels in cerebral artery smooth muscle cells. A stretch-sensing pathway in cerebral artery myocytes that includes 1) activation of AT1R, Src tyrosine kinase, and PLC-γ1, leading to the generation of IP3; and 2) Ca2+ influx through TRPC6 channels onto IP3Rs, promoting CICR that results in activation of TRPM4 channels, which are responsible for pressure-induced depolarization of the plasma membrane. PM, plasma membrane; ΔVm, change in membrane potential. [From Gonzales et al. (118).]

TRPC6 channels are also involved in vasoconstrictor responses in the pulmonary arterial circulation (385). Small pulmonary arteries constrict in response to hypoxic challenge, a response called hypoxic pulmonary vasoconstriction (HPV). HPV improves oxygenation of the blood by matching perfusion to well-ventilated regions of the lung. Weissmann et al. (385) found that robust HPV, present in the lungs of wild-type mice, was absent in TRPC6−/− mice. Hypoxia-induced activation of Ca2+ influx as well as TRPC6-like cation currents were also absent in pulmonary artery myocytes from TRPC6−/− mice. Interestingly, in contrast to the apparent compensatory upregulation of TRPC3 expression in nonpulmonary arteries previously documented in TRPC6−/− mice (70), no change in the expression of TRPC3 or other TRPC channels (other than TRPC6) was reported in this study. A follow-up study by this group reported the clear involvement of DAG as a mediator of TRPC6 activation in pulmonary artery myocytes during hypoxia (99). Others have reported a key role for epoxyeicosatrienoic acid (EET) derivatives as mediators of enhanced TRPC6 contributions to HPV (272). Much of the ability of EETs to increase channel activity in pulmonary artery vascular smooth muscle cells is likely attributable to recruitment of TRPC6 channels to caveolae (159).

It is now abundantly clear that TRPC6 channels are fundamentally important for the onset, maintenance, and modulation of vascular tone in many different vascular beds. Additional studies are needed to further define the unique contributions of these channels in various regions within the cardiovascular system and to characterize other important modulatory inputs that may influence the activity of these channels in various physiological and pathological contexts. For instance, more detail is needed about the regulation of TRPC6 by PLC-derived molecules, the influences of related TRPC channels that form heteromultimers with TRPC6, and the contributions of TPRC6-mediated Ca2+ entry to local and global cellular signaling processes.

b) trpc3. Expression of TRPC3 channel mRNA and/or protein has been demonstrated by RT-PCR, Western blotting, and/or immunohistochemical approaches in vascular smooth muscle cells in cerebral (256, 283), cephalic (12), coronary (168, 199), renal (373), uterine (306), and pulmonary arteries (420) and in the aorta (202). However, the properties of TRPC3 currents in native vascular smooth muscle cells have been described only rarely. Albert et al. (12) recorded a constitutively active nonselective cation channel in vascular myocytes from the rabbit ear artery with pharmacological properties very similar to cloned TRPC3 channels. This channel appeared to have three distinct conductance levels (10), although none matched the single value reported for the cloned TRPC3 channel expressed in HEK cells (427). Heteromultimeric TRPC channel assembly, in particular between TRPC3 and TRPC7 channels, has been documented in coronary artery myocytes, and related diversity in channel properties has been described (266).

TRPC3 channels in vascular smooth muscle cells are involved in contractile responses, primarily evoked by receptor-ligand interactions. This function has been best described in cerebral and carotid arteries and the aorta. TRPC3 channels in vascular smooth muscle are activated by stimulation of a variety of GPCRs, including pyrimidine (283), endothelin-1 (ET-1) (266), and ANG II (199) receptors. In brain pial arteries, suppression of TRPC3 channel expression using antisense oligonucleotides was shown to greatly reduce inward currents, depolarization, and vasoconstriction induced by the purinergic receptor agonist uridine triphosphate (UTP) (283). Interestingly, in this study, intrinsic myogenic tone was unaltered by TRPC3 downregulation, suggesting a specific role for TRPC3 in receptor-mediated vasoconstriction and not in pressure-induced vasoconstriction. This concept is supported by the findings of another laboratory that showed Ca2+ entry and cerebral artery constriction induced by ET-1 was decreased following suppression of TRPC3 expression (398). ET-1 also activated nonselective cation channels in coronary artery myocytes that have properties conferred, in part, by TRPC3 channel protein (266). Liu et al. (199) reported that ANG II receptor-mediated Ca2+ entry was decreased by suppression of TRPC3 expression and increased by TRPC3 overexpression in cultured aortic vascular smooth muscle cells. In rat afferent arterioles, a heteromultimeric TRPC channel structure containing TRPC3 together with TRPC1 and TRPC6 subunits was shown to participate in a norepinephrine-stimulated, nifedipine-resistant constrictor response (296). Taken together, these studies indicate that TRPC3 channels, alone or in association with other related TRPC subunits, are important mediators of agonist-evoked vasoconstriction in many different types of arteries. An interesting aspect of TRPC3 behavior in general, and specifically in vascular smooth muscle cells, is that these channels are constitutively active (12). This basal TRPC3 activity suggests possible roles in tonic regulation of membrane potential and arterial tone (70, 306) as well as potential involvement in the enhancement of myogenic tone and/or agonist-evoked vasoconstriction (306).

The primary mechanism for activation of TRPC3 downstream of receptor stimulation in expression systems (140) and in native vascular myocytes (11) was initially thought to be through the direct actions of DAG on the channel. However, work from Kiselyov et al. (162) showed that TRPC3 channels functionally interact with IP3Rs in HEK cell expression systems, and Boulay et al. (35) showed that TRPC3 and IP3Rs could be coimmunoprecipitated. Jaggar and colleagues (35, 398) showed that IP3Rs are critical elements in the activation mechanism for TRPC3 channels in cerebral artery smooth muscle cells. This activation pathway does not involve IP3-mediated Ca2+ release from the SR, but rather reflects direct coupling between IP3Rs and TRPC3 channels mediated by IP3. Coupling of TRPC3 channels and IP3Rs in vascular smooth muscle cells leads to activation of TRPC3-mediated cation currents, followed by membrane depolarization, Ca2+ entry through VDCCs, and myocyte contraction. Close physical association of IP3Rs with TRPC3 channels is orchestrated by the integral membrane protein caveolin-1. Activation of heteromultimeric TRPC3/TRPC7 channels by ET-1 in coronary artery myocyte (266) also involves products of phosphoinositide metabolism. Phosphatidylinositol-3-phosphate, formed by the actions of specific phosphatidylinositol 3-kinase isoforms, appears to be a direct activator of this channel (311). Interestingly, in this case IP3 itself inhibited the TRPC3/C7 heteromultimeric channel, demonstrating the potential differential control of homomeric versus heteromultimeric TRPC channels. In expression systems, TRPC3 trafficking, membrane localization, and activity are regulated by PIP2, which is the primary substrate for phosphoinositide metabolism (142, 181). Whether this holds true for native vascular smooth muscle cells has not been determined.

Other kinase systems have substantial roles in the regulation of TRPC3 channels in vascular smooth muscle. In several instances, kinases have inhibitory effects on TRPC3 channel function. For instance, the serine-threonine kinase WNK4 exerts a tonic inhibitory action on TRPC3 function in rat aorta (262). Protein kinase G (PKG) also inhibits TRPC3, an action that is correlated with dilation of isolated carotid artery segments (49). This is consistent with previous observations that PKG inhibits TRPC3 channels expressed in HEK293 cells (175). In contrast, studies employing TRPC3 knockout mice found no apparent link between TRPC3 function and cGMP- or PKG-induced relaxation of the aorta or the hindlimb circulation (202), perhaps suggesting species- or tissue-specific differences in this regard. In expression systems (360) and in arterial myocytes (8), TRPC3 channels are inhibited by PKC.

The weight of evidence supports the contention that TRPC3 channels in vascular smooth muscle play important roles in regulating vascular contractility. Constitutive activity of these channels contributes to basal vascular tone, and activation of TRPC3 channels by vasoconstrictor agonists causes further membrane depolarization and Ca2+ entry through VDCCs. Future studies will provide needed details of the mechanisms by which the basal and ligand-activated states of TRPC3 channels are achieved.

c) trpm4. TRPM4 and the closely related TRPM5 are the only TRP channels that are essentially impermeant to Ca2+. However, because of their relative high Na+-versus-K+ permeability and significant inward current amplitudes in standard ionic gradients and at physiological membrane potentials, TRPM4 channels have considerable capacity to affect membrane depolarization and thereby alter cellular function. TRPM4 channels have been reported to influence Ca2+ entry in excitable (125) and nonexcitable cells (178), primarily through their depolarizing influence. In excitable cells, depolarization leads to activation of VDCCs and a global increase in Ca2+, whereas in nonexcitable cell depolarization diminishes passive Ca2+ influx because it reduces the electrical driving force.

Molecular and functional studies have shown that TRPM4 channels are present in native smooth muscle cells. TRPM4 mRNA has been reported in smooth muscle cells in the rat aorta (145, 408), as well as mesenteric (145), pulmonary (408), and cerebral arteries (83, 145). Immunohistochemical and Western blot analyses have also demonstrated the presence of TRPM4 protein in cerebral artery myocytes (64, 114). Single-channel and whole cell currents with properties very similar to those recorded from HEK cells overexpressing TRPM4 channels have been detected in cerebral artery myocytes (82, 83, 234). The activity of TRPM4 channels is primarily regulated by the intracellular [Ca2+] through interactions of the Ca2+/calmodulin complex with defined sites near the COOH terminus (252). Activation of native and exogenously expressed TRPM4 channels requires a very high (micromolar) [Ca2+] at the intracellular face of the channel (82) (251). However, global intracellular Ca2+ levels in vascular smooth muscle cells are typically in the range of 100–500 nM, suggesting that TRPM4 activation must occur within subcellular regions with locally elevated [Ca2+]. Such domains are created when Ca2+ is released through ryanodine receptors (234) or IP3Rs (114). Work by Gonzales et al. demonstrated that TRPM4 channels in native cerebral artery smooth muscle cells are activated by release of Ca2+ from IP3Rs (114) and that channel activation is dependent on intracellular Ca2+ buffering (115).

Ca2+-dependent activation of TRPM4 can be fine tuned by other signaling pathways. For example, Nilius et al. (252) demonstrated that the innate Ca2+ sensitivity of cloned TRPM4 channels expressed in HEK cells is elevated by PKC-dependent phosphorylation. Similarly, stimulation of PKC activity was shown to increase TRPM4 whole cell currents in cerebral artery myocytes and cause constriction of isolated cerebral arteries (82). This increase in Ca2+ sensitivity was found to result from PKC-δ-dependent translocation of TRPM4 channels to the plasma membrane of atrial myocytes (64, 104). PIP2 has been clearly implicated as an important regulator of TRPM4 channel activity in expression systems (250, 422). These studies demonstrated that the exogenous administration of PIP2 prevented desensitization of TRPM4 currents in response to high levels of Ca2+ (~10–100 μM) at the cytosolic face; however, the significance of this pathway in native vascular smooth muscle cells remains to be determined.

TRPM4 channels are critically important for cerebral artery function (76, 116). Suppression of TRPM4 expression in cerebral arteries was shown to result in a substantial reduction in smooth muscle membrane depolarization and vasoconstriction accompanied by increased intravascular pressure (83). This pioneering study was reinforced by a subsequent report showing that 9-phenanthrol, a highly specific pharmacological inhibitor of TRPM4 channels (119), inhibited pressure-induced depolarization and myogenic tone in isolated pial arteries (117). However, interpretation of these latter data may be called into question by a recent study suggesting that 9-phenanthrol can activate intermediate-conductance Ca2+-activated K+ channels present in the endothelium to cause vasodilation of mesenteric arteries (107). Two additional studies have confirmed the involvement of TRPM4 channels in cerebral myogenic tone, in this case in the cerebral intraparenchymal arteriolar network (37, 188). The involvement of TRPM4 in pressure-induced vasoconstriction suggests that this channel is mechanically activated. Although one study reported that application of negative pressure to the recording electrode activated single cation channels in vascular smooth muscle cells with properties much like those in HEK cells overexpressing TRPM4 channels (234), a more recent study failed to detect inherent mechanosensitivity of TRPM4 (118). Recent work points to involvement of purinergic or ANG II type 1 receptor (AT1R) GPCRs in arterial pressure-sensing. Along these lines, a compelling case has been made that TRPM4 channels are activated downstream of a mechanosensitive losartan-sensitive receptor (likely AT1R) in cerebral artery smooth muscle cells (118) (Figure 6). Mechanoactivation of P2Y4 and P2Y6 purinergic receptors and downstream activation of TRPM4 channels in cerebral intraparenchymal arteriolar myocytes have also been proposed (188).

Two studies have investigated the effect of TRPM4 activity on cardiovascular control in vivo. Using a knockdown approach, Reading et al. (282) demonstrated that TRPM4 channels in cerebral artery myocytes directly contribute to autoregulation of cerebral blood flow in intact animals, presumably through control of myogenic tone. However, another study (214) reported no difference in agonist- or pressure-induced contractile responses of the aorta or in hindlimb vascular preparations isolated from wild-type and TRPM4−/− mice. It is unclear if these discrepancies are due to differences in vascular beds or experimental preparations, or possible compensatory expression of other cation channels in the TRPM4−/− animals. Furthermore, the TRPM4−/− mice used for this study were mildly hypertensive, likely due to increased release of catecholamines from chromaffin cells. Thus it is possible that elevated catecholamine levels in TRPM4−/− mice compensate for the loss of TRPM4-induced smooth muscle contractility. Additional studies are needed to determine if vascular function in specific areas of the circulation, such as in the brain, is altered in TRPM4−/− animals, preferably using conditional and cell-specific deletion approaches.

TRPM4 activity also promotes excitability of detrusor smooth muscle in the urinary bladder (258, 321, 322), suggesting that this channel could have a broad role in the regulation of smooth muscle contractility.

d) trpp channels. The TRPP subfamily consists of TRPP1 (PKD2), TRPP2 (PKD2L1), and TRPP3 (PKD2L2). TRPP1 is a Ca2+-permeable, nonselective cation channel with a large unitary conductance (135–175 pS) that may be localized to the plasma membrane or intracellular organelles. PKD1 is sometimes referred to as TRPP1, but it is structurally distinct from the other TRPP proteins and does not form a cation-permeable ion channel. PKD1 can interact with TRPP1 and may be a mechanically sensitive regulatory subunit. PKD1 and TRPP1 are both expressed in vascular smooth muscle cells (121, 243, 354). Early reports suggested that PKD1 and TRPP1 physically interacted and might be involved in regulating the development and organization of the cytoskeletal elements of vascular myocytes (308). A report by Shaif-Naeini et al. (308) demonstrated that overexpression of TRPP1 in COS-7 cells diminished the stretch-induced currents present in native cells. Mechanosensitive currents were restored by coexpression of PKD1 and TRPP1. Selective deletion of PKD1 in smooth muscle cells (Pkd1SMdel/del) decreased stretch-activated currents in mesenteric artery smooth muscle cells and diminished myogenic constriction of isolated mesenteric arteries. Surprisingly, knockdown of TRPP1 expression in Pkd1SMdel/del mice restored stretch-activated currents in isolated smooth muscle cells and myogenic constriction of mesenteric arteries. The authors of this study proposed that, in the absence of PKD1, TRPP1 inhibits mechanosensitive ion channels involved in the vascular smooth muscle depolarization associated with myogenic tone and that these responses are restored by physical interaction of PKD1 with TRPP1.

Phenylephrine-induced aortic and mesenteric resistance artery contractions are enhanced in TRPP1+/− mice due to increased actin and myosin expression (72, 279), suggesting that TRPP1 may exert inhibitory effects on vascular tone through inhibition of contractile protein structure and function. An age-dependent vascular phenotype was observed in PKD1−/+ mice, with old mice (40 wk), but not young mice, exhibiting an increase in phenylephrine-mediated aortic contractility (133). This finding also suggests vascular bed-specific roles for TRPP isoforms, although clarification of this possibility will require additional detailed studies. It has also been proposed that the physical interaction of PKD1 with TRPP1, and related membrane association of these proteins, is an important element and causal factor in some of the vascular pathologies observed in patients with polycystic kidney disease (280).

Jaggar et al. (243) reported a different take on TRPP1 channels, concluding that TRPP1 cation currents directly activate myogenic tone in the cerebral circulation. This study showed that TRPP1 and PKD1 are present in cerebral artery myocytes and localize to the plasma membrane, and further indicated that TRPP1 mRNA is much more abundant than PKD1 mRNA. Knockdown of TRPP1 reduced swelling-activated cation currents in isolated vascular smooth muscle cells and suppressed the development of myogenic tone in isolated cerebral arteries. The reasons for the apparently divergent functions of TRPP1 channels in cerebral and mesenteric arteries in particular, and the physiological roles of TRPP proteins in vascular smooth muscle cells in general, are not clear and will require further investigation.

e) trpv1. Expression of TRPV1 channels in aortic and skeletal muscle arteriolar myocytes from rats has been reported (201, 158, 355). These studies showed that the TRPV1 activator capsaicin increases hindlimb vascular resistance in vivo and causes substantial endothelium-independent vasoconstriction of skeletal muscle arterioles in vitro. Vasoconstriction of isolated canine mesenteric arteries following direct activation of TRPV1 channels has also been reported (273). More recent work has revealed the presence of TRPV1 channels in murine arteriolar myocytes located in thermoregulatory tissues and has demonstrated a role for these channels as mediators of myocyte Ca2+ entry and contraction in intact auricular arterioles (46).

f) trpv2. TRPV2 channels are present in murine aortic, mesenteric, and basilar artery smooth muscle cells (238). Hypotonic cell swelling of aortic myocytes was shown to activate nonselective cation channels and Ca2+ entry, effects that were suppressed by treatment with TRPV2 antisense oligonucleotides. These findings led the authors of this study to propose a role for TRPV2 channels in myogenic vasoconstriction; however, functional studies investigating the role of TRPV2 channels in intact arteries have not been reported.

g) trpm8. TRPM8 channels are thought to be involved primarily in sensory nervous system signaling activated by cold and menthol. A single study showed that TRPM8 mRNA and protein are present in rat aortic, mesenteric, tail, and femoral artery myocytes (151). In most reported experiments, menthol caused a dilatory response in the aorta, as well as in mesenteric and tail arteries, that was at least partly endothelium independent. However, in some cases menthol caused small contractions of endothelium-dependent aortic rings. The exact role of TRPM8 channels in the vasculature remains undefined.

h) trpc1. TRPC1 channels are frequently linked to SOCE (30, 145), although this function has been disputed (68, 69, 274), and a few studies have investigated ROCE through TRPC1 channels in vascular smooth muscle cells. For example, ET-1 was shown to induce VDCC- and SOCE-independent Ca2+ entry in cerebral arteries (124) and in cultured aortic smooth muscle cells (344) that may be mediated by TRPC1 channels (344). A series of studies by Large, Albert, and collegues investigated receptor-operated regulation of TRPC1 and heteromultimeric channels containing TRPC1 subunits in vascular smooth muscle cells (291, 309, 310, 312). These studies suggested that TRPC1 channels are activated downstream of PLC- and PKC-dependent signaling pathways. A model proposed in another report suggests an association between TRPC1, IP3Rs, and STIM1 that could account for Ca2+ influx and contraction of vascular smooth muscle independent of SOCE (34), although supporting experimental evidence is meager. The biophysical properties of TRPC1 channels suggest that the fractional Ca2+ component of TRPC1 currents under physiological conditions is very small. This argues against involvement of direct Ca2+ influx, but is it possible that TRPC1-mediated cation influx could cause membrane depolarization and Ca2+ influx through VDCCs in excitable cells. However, vasoconstriction in response to elevated intravascular pressure (69), adrenoceptor activation, or K+-induced depolarization (304) is unaltered in TRPC1−/− mice.

2. SOCE in vascular smooth muscle cells

Many studies have investigated the contribution of TRPC-mediated SOCE in vascular smooth muscle cells to cellular differentiation, proliferation, and excitation-contraction coupling (183). In considering these reports, it is important to bear in mind that SOCE is not active in all types of native, contractile smooth muscle cells, and its presence may be an indicator of phenotypic differentiation to a proliferative state or adaptation to cell culture conditions (274). In addition, as noted above, several lines of evidence argue against the involvement of TRPC channels in SOCE in vascular smooth muscle and other cell types. (68, 274). While the issue remains controversial, there is some evidence supporting the involvement of TRPC channel activity in SOCE in smooth muscle cells, as discussed below.

a) trpc1. The contribution of TRPC1 to SOCE activity has been investigated using siRNA-mediated knockdown in A7r5 cells (40) and neutralizing antibody approaches (30, 402) in smooth muscle from several tissues, including pulmonary arteries (245247) and the portal vein (14). SOCE via TRPC1 channels is also thought to be involved in proliferation of pulmonary artery myocytes in culture systems (216, 348). However, conflicting evidence is provided by studies using TRPC1−/− mice, which found no evidence for the involvement of TRPC1 in SOCE in myocytes isolated from cerebral arteries or the thoracic aorta (69). Work by Saleh et al. (292) suggests that TRPC1 currents activated by the SERCA pump inhibitor cyclopiazonic acid (interpreted as SOCE) require PIP2 for activation.

A few studies have attempted to link SOCE through TRPC1 with smooth muscle cell contraction. In the pulmonary vasculature, SOCE has been shown to cause vasoconstriction (174), and some evidence suggests that hypoxia-induced pulmonary vasoconstriction could be mediated by SOCE through TRPC1 channels (187, 205). One report suggested that SOCE via TRPC1 leads to the activation of large-conductance, Ca2+-sensitive K+ (BK) channels and relaxation of arterial smooth muscle (176). In this study, smooth muscle cell hyperpolarization and decreased global Ca2+ in response to activation of SOCE were greatly inhibited by a TRPC1-specific antibody or by a selective inhibitor of BK channels. Coimmunoprecipitation and immunohistochemical analyses indicated that TRPC1 and BK channels are colocalized in arterial myocytes. The authors of this study proposed that TRPC1 and BK channels form a vasodilator-signaling complex in vascular smooth muscle cells, a conclusion supported by the subsequently reported interaction between TRPC1 and BK channels in aortic smooth muscle cells (166). Although it remains difficult to reconcile how the small fractional Ca2+ currents generated by TRPC1 channels (~2 pS; Ref. 177) could produce localized domains with Ca2+ levels high enough to activate BK channels, such physical interactions may provide an enabling mechanism.

b) trpc4. Two studies in cultured cells, one using aortic and one using mesenteric artery cells, suggested the involvement of TRPC4 channels in SOCE in smooth muscle cells. Cyclic stretch of these cells in culture reportedly caused a decrease in TRPC4 expression that was correlated with a decrease in SOCE (197). A second report from this group demonstrated that downregulation of TRPC4 was associated with reduced SOCE in mesenteric artery smooth muscle cells from normotensive WKY rats, but not spontaneously hypertensive rats (SHR) (196).

c) trpc5. Some evidence supporting a role for TRPC5 in SOCE in vascular smooth muscle cells, likely in heteromultimeric combinations with TRPC1, TRPC6, and/or TRPC7, has been presented. For example, an anti-TRPC5 antibody has been demonstrated to have a significant inhibitory effect on store-depletion-induced cation currents in freshly isolated cerebral pial arterioles (403). Similarly, Saleh et al. (294) showed that currents evoked by administration of the SERCA pump inhibitor cyclopiazonic acid were attenuated by anti-TRPC5, anti-TRPC1, and anti-TRPC7 antibodies in smooth muscle cells from coronary and mesenteric arteries as well as the portal vein (294).

3. TRPV4 channels in vascular smooth muscle cells and vasodilation

The presence of TRPV4 in cerebral artery smooth muscle cells has been demonstrated by immunohistochemistry (211) and RT-PCR (79). Earley et al. (79) reported the surprising finding that Ca2+ influx through TRPV4 channels in cerebral arterial myocytes caused vasodilation rather than vasoconstriction. This study showed that Ca2+ influx through TRPV4 channels stimulated by 11,12-EET activated Ca2+-induced Ca2+ release (CICR) from SR stores in the form of Ca2+ sparks. Ca2+ sparks evoked by 11,12-EET activated BK channels, leading to myocyte hyperpolarization and decreased Ca2+ entry though VDCCs, causing vasodilation. Thus, although previous reports clearly demonstrated a key role for BK channels in the mechanism of vasodilation induced by EETs (42, 143), an indirect activation pathway involving TRPV4 channels appears to be a major mechanism of EET-induced activation of BK channels. A similar role for smooth muscle TRPV4 channels in regulating mesenteric artery contractile activity in vitro and in vivo has been described (81). In this study, ~50% of 11,12 EET-induced vasodilation was due to activation of TRPV4 channels in the endothelium and 50% was due to activation of the channel in vascular smooth muscle cells. A subsequent study by Macado et al. (221) demonstrated that ANG II acts through AT1R to increase TRPV4 activity, recorded in the form of TRPV4 sparklets, in isolated cerebral artery smooth muscle cells. In intact cerebral arteries, Ca2+ influx through TRPV4 channels opposes the vasoconstrictor effects of ANG II through the Ca2+ spark-dependent mechanism described above. Using selective pharmacology and an elegant optogentic approach, these authors demonstrated that the effects of AT1R activation on TRPV4 sparklet activity are dependent on PKC. This study further demonstrated that the dynamic interactions between TRPV4 channels and PKC necessary for AT1R-dependent regulation of TRPV4 sparklets are coordinated by AKAP150 (a kinase anchoring protein 150). Collectively, these findings reveal the formation of an AKAP150-organized subcellular signaling complex involving AT1Rs, PKC, and TRPV4 channels that is capable of finely tuning local [Ca2+] and CICR to influence vascular contractility (Figure 7).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240007.jpg

Ca2+ influx through TRPV4 channels relaxes cerebral artery smooth muscle cells. Stimulation of Ca2+ influx through TRPV4 channels directly with 11,12-EET or through activation of Gq-coupled AT1Rs with ANG II, leading to generation of DAG by PLC and AKAP150-dependent PKC activity, causes an increase in the frequency of Ca2+ sparks mediated by ryanodine receptors (RyRs) on the SR through a CICR mechanism. Increased generation of Ca2+ sparks causes smooth muscle cell membrane hyperpolarization and relaxation by increasing K+ efflux through BK channels. [Modified from Earley et al. (79), with permission from Wolters Kluwer Health, and Mercado et al. (221).]

B. TRP Channels in the Vascular Endothelium

The vascular endothelium is composed of a single layer of metabolically dynamic endothelial cells that line all of the blood vessels in the body. This tissue is located between the vascular lumen and underlying smooth muscle cells, allowing the endothelium to sense blood-borne substances and shear stress resulting from fluid flow along the vessel wall. The principal functions of the endothelium are to promote smooth muscle cell relaxation and arterial dilation, control vascular permeability, exert an antithrombotic effect, and regulate angiogenesis. Endothelial dysfunction is a hallmark of common cardiovascular diseases such as hypertension, atherosclerosis, and stroke, demonstrating the importance of this tissue for vascular health. Several TRP channels, acting though regulation of the global intracellular [Ca2+] in endothelial cells and the generation of subcellular Ca2+ signals, are important for endothelial cell function (336). Approximately 20 different TRP channel family members are reported to be expressed in cultured and native endothelial cells (336), and the functional significance of nine of these channels–TRPC1 (47, 103), TRPC3 (47, 103), TRPC4 (97, 103), TRPC5 (103), TRPC6 (103), TRPV1, TRPV3 (78), TRPV4 (392), and TRPA1 (77)–has been investigated (Table 3).

Table 3.

TRP channels in the vascular endothelium

TRPFunctionTissue
TRPC1Angiogenesis, permeabilityEmbryonic tissues (411), lung (39, 218)
TRPC3Angiogenesis, endothelium-dependent dilationMesentery (198, 306), mammary artery (102)
TRPC4Angiogenesis, endothelium-dependent dilation, permeabilityUmbilical vein (19), aorta (97), lung (59, 352)
TRPC5AngiogenesisUmbilical vein (110)
TRPC6Angiogenesis, permeabilityUmbilical vein (110), epidermis (128)
TRPV1Endothelium-dependent dilationMesentery (57, 406), heart (36, 123)
TRPV3Endothelium-dependent dilationBrain (78)
TRPV4Angiogenesis, endothelium-dependent dilation, permeabilityAorta (381), mesentery (81, 295, 324, 417), skeletal muscle (169), heart (220, 371), lung (15), brain (132, 211)
TRPA1Endothelium-dependent dilationBrain (77, 281), Sullivan et al., unpublished data

Reference numbers are given in parentheses.

1. TRP channels and endothelium-dependent vasodilation

The endothelium exerts a profound relaxing effect on underlying smooth muscle cells that, balanced by tonic constrictor influences, maintains optimal vascular tone (Figure 8). Nitric oxide (NO) (100) and prostacyclin (PGI2) are well-characterized vasoactive substances produced by the endothelium that diffuse to and relax smooth muscle to cause arterial dilation. Endothelium-dependent vasodilation persists when NO and PGI2 production are blocked, a response initially attributed to an unidentified “endothelium-derived hyperpolarizing factor” (EDHF) (48). The molecular basis of EDHF is now thought to encompass multiple diffusible substances, including K+ (86), EETs (42, 94), C-type natriuretic peptide, carbon monoxide, hydrogen peroxide (H2O2), hydrogen sulfide (H2S), and others (87). In addition, vasodilatory responses originally attributed to EDHF are now known to result from direct electrotonic spread of membrane hyperpolarization from endothelial cells to smooth muscle cells through myoendothelial gap junctions (Figure 8). This factorless mechanism is referred to as endothelium-dependent hyperpolarization (EDH) to avoid confusion with EDHF-type pathways involving diffusible substances (106).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240008.jpg

Mechanisms of endothelium-dependent vasodilation. In endothelial cells (EC), Ca2+ influx stimulates the activity of eNOS to generate NO, which diffuses through the IEL to cause relaxation of underlying vascular smooth muscle cells (SMC). COX activity generates PGI2, which also relaxes SMC. A variety of other substances, collectively known as EDHF, can cause smooth muscle cell hyperpolarization and relaxation. K+ efflux hyperpolarizes the endothelial cell plasma membrane (ΔEm). Electrotonic spread of this influence through myoendothelial gap junctions (MEGJs) promotes EDH, causing SMC relaxation.

The architecture of the sites of interaction between smooth muscle and endothelial cells within the vascular wall facilitates EDH. Smooth muscle cells and endothelial cells are separated by a structure composed primarily of elastin and collagen called the internal elastic lamina (IEL). Projections of the endothelial cell plasma membrane protrude into fenestrations in the IEL, providing an interface between endothelial cell and smooth muscle cell plasma membranes. Some, but not all, of these sites, collectively termed myoendothelial projections (MEPs), contain gap junctions between smooth muscle and endothelial cells. Those sites that contain gap junctions have been referred to as myoendothelial junctions (MEJs), whereas those that lack gap junctions have been termed myoendothelial close contacts (MECCs). Connexin proteins (MEJs only), intermediate-conductance Ca2+-activated K+ (IK) channels (77, 298), IP3Rs (180), TRPA1 (77), TRPV4 (325), and other signaling proteins are present within MEPs. In mesenteric arteries, transient, localized Ca2+ signals are generated within MEPs by release of Ca2+ from the ER through IP3Rs. These signals, which occur both spontaneously and in response to activation of GPCRs, are called Ca2+ pulsars and stimulate EDH by activating SK (small-conductance) and IK Ca2+-activated K+ channels (180) (Figure 9).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240009.jpg

Ca2+ pulsars in the endothelium. A: time course of a three-dimensional Ca2+ pulsar originating from within a hole in the IEL (white circle) shown in the leftmost image. Scale bar, 5 μm. B: Ca2+ pulsars recorded from a pressurized artery (80 mmHg) expressing the Ca2+-indicator protein GCaMP2. An endothelial cell and its nucleus are outlined (dotted lines), with the initiation sites indicated by red arrows. Scale bar, 10 μm. C: model of the functional effects of endothelial Ca2+ pulsars. IP3R-dense ER stores follow portions of the endothelial cell membrane that evaginate through holes in the IEL and interface with underlying SM cell membranes. Repetitive localized Ca2+ events (pulsars) originate from these deep Ca2+ stores, which are regionally delimited to the myoendothelial junction and the base of the endothelial cell. These ongoing dynamic Ca2+ signals are driven by constitutive IP3 production and are inherently dependent on the level of endothelial stimulation. The left detail depicts a single endothelial projection through the IEL. IK (KCa3.1) channels in the plasma membranes of these endothelial projections are in very close proximity to Ca2+ pulsars, eliciting persistent Ca2+-dependent hyperpolarization of the membrane potential at the myoendothelial junctions. The right detail illustrates the endothelial influence on the smooth muscle membrane potential at the myoendothelial interaction site where Ca2+ pulsars activate KCa3.1 channels and hyperpolarize the endothelial membrane. This hyperpolarization can be transmitted to the SM through gap junction channels or by activation of SM Kir channels by K+ ions released by endothelial KCa3.1 channels. Membrane hyperpolarization promotes relaxation of SM through a decrease in VDCC open probability. [From Ledoux et al. (180).]

The Ca2+ influx pathways that stimulate endothelium-dependent vasodilation remained largely unknown until the discovery of TRP channels (248). To date, TRPV1, TRPV3, TRPV4, TRPA1, TRPC3, and TRPC4 channels have been reported to be involved in endothelium-dependent vasodilation (Table 3) (77, 78, 307, 416, 417).

a) trpv1. TRPV1 channels are activated by capsaicin, a substance found in hot peppers, and by noxious heat (45). The single-channel conductance of TRPV1 is 35 pS at −60 mV and 77 pS at +60 mV. TRPV1 channels are selective for Ca2+ versus Na+ (PCa:PNa ~9.6). Several studies have provided evidence of TRPV1 involvement in endothelium-dependent vasodilation. Poblete et al. (271) demonstrated that activation of TRPV1 channels with anandamide increased the release of endothelium-derived NO but did not cause dilation of rat mesenteric arteries. Work from Yang et al. (406) demonstrated that capsaicin stimulated Ca2+ influx and endothelial nitric oxide synthase (eNOS) phosphorylation, and elevated NO production. Capsaicin also caused endothelium-dependent relaxation of isolated mesenteric arteries, an effect that was absent in arteries from TRPV1−/− mice. In addition, this study showed that endothelial function was improved and systolic blood pressure was slightly lower in SHRs receiving dietary administration of capsaicin for 7 mo compared with animals fed a normal diet. A subsequent study by Ching et al. (57) reported that capsaicin-induced Ca2+ influx through TRPV1 channels caused TRPV1 to physically interact with eNOS through a pathway requiring phosphatidylinositol 4,5-bisphosphate 3-kinase, Akt, and calmodulin-dependent protein kinase II (CamKII), resulting in increased NO production (57). Bratz et al. (36) demonstrated that capsaicin dilated coronary arteries from pigs and showed that this response was diminished by the TRPV1 antagonist capsazepine and inhibition of NO production and was blunted in coronary arteries from obese swine. Further work from this group showed that capsaicin increased myocardial blood flow in mice in vivo, a response that was absent in TRPV1−/− animals and blunted in the db/db model of type II diabetes (123). Together, these studies support a model in which Ca2+ influx through TRPV1 channels increases the production of endothelium-derived NO to cause vasodilation. This mechanism may be disrupted during metabolic syndrome and diabetes, contributing to the endothelial dysfunction associated with these conditions.

A study by Chen et al. (50) showed that capsaicin caused an endothelium-dependent dilation of mesenteric arteries that was absent in TRPV1−/− mice. In contrast, capsaicin had no TRPV1-dependent effects on the main renal arteries, suggesting a heterogeneous distribution of TRPV1 channels in the endothelium throughout the vasculature.

b) trpv3. TRPV3 channels have a large unitary conductance (~150–200 pS) (58) and are highly permeable to Ca2+ (PCa:PNa ~12:1) (401). TRPV3 is activated by innocuous heat (401) and dietary monoterpenes such as carvacrol (derived from oregano), eugenol (clove oil), and thymol (found in thyme) (400). The channel is present in the skin and oral and nasal epithelium and is involved in chemosensation (400). Earley et al. (78) found that TRPV3 channels are also present in the endothelium of rat cerebral arteries and showed that carvacrol activated ruthenium red-sensitive cation currents and increased intracellular Ca2+ levels in native cerebral artery endothelial cells. Carvacrol administration evoked an endothelium-dependent dilation of isolated cerebral arteries that was not altered by NOS or cyclooxygenase (COX) inhibition; however, it was sensitive to block of SK, IK, and inwardly rectifying K+ (KIR) channels, and was associated with smooth muscle cell hyperpolarization. These data suggest that Ca2+ influx through TRPV3 activates EDH (78). The cardiovascular effects of TRPV3 knockout have not been reported, and the functional significance and endogenous regulation of TRPV3 in the endothelium remain unknown.

c) trpv4. TRPV4 channels are more permeable to Ca2+ than Na+ (PCa:PNa ~6:1) (332) and are activated by cell swelling (193), warm temperatures, and chemical agonists (351, 381). The unitary conductance of the TRPV4 channel is ~30–60 pS at −60 mV and ~88–100 pS at +60 mV. The involvement of TRPV4 channels in endothelium-dependent vasodilation has been extensively studied, in part because of the availability of selective pharmacological activators, including 4α-phorbol 12,13-didecanoate (4α-PDD) (381) and GSK1016790A (351), and blockers such as RN-1732 (371) and HC-067047 (89).

I) TRPV4 and EET-induced dilation. Wantanabe and co-workers discovered that TRPV4 channels are present in freshly isolated mouse aortic endothelial cells and could be activated by the phorbol compound 4α-PDD (381), heat (25–43°C) (383), and EETs (382). The EETs are a family of four regioisomers, 5,6-EET, 8,9-EET, 11,12-EET, and 14,15-EET, generated from arachidonic acid by cytochrome P-450 (CYP) epoxygenase enzymes. These compounds have significant biological activity in many tissues, and when produced by endothelial cells may act as EDHFs in some vascular beds (42, 80, 88, 94). Earley et al. (81) demonstrated that exogenous administration of 11,12-EET dilated isolated mesenteric arteries from wild-type, but not TRPV4−/−, mice. Approximately 50% of this response was lost when endothelial function was disrupted, suggesting that EETs can act on TRPV4 channels in both endothelial cells and smooth muscle cells (81). This study also showed that 11,12-EET-induced vasodilation was associated with smooth muscle cell hyperpolarization and was not affected by NOS or COX inhibition, but was blocked by inhibition of BK, IK, or SK channels, supporting an EDHF or EDH mechanism. A subsequent study by Vriens et al. (372) established that EETs generated by the CYP2C9 epoxygenase isoform could activate TRPV4 in an autocrine manner in cultured endothelial cells, providing support for the hypothesis that endogenously produced EETs activate Ca2+ influx through TRPV4 channels in an autocrine/paracrine manner to elicit dilation.

II) TRPV4 and flow-induced dilation. Increases in blood flow velocity or blood viscosity augment the level of laminar shear stress at the vascular wall. The endothelium senses this stimulus, initiating a vasodilator response known as “flow-induced dilation” (66). The molecular nature of the flow-sensing mechanism remains incompletely understood. An early study showing that TRPV4 was activated by hypotonic conditions, possibly by stretch of the plasma membrane associated with cell swelling (193), stimulated interest in this channel as a flow sensor in the endothelium. However, a subsequent study applying a cell-attached patch-clamp approach reported that TRPV4 channels were not directly activated by suction applied to the plasma membrane (332), suggesting that cell swelling activates the channel indirectly through a force-sensitive signaling cascade.

A number of studies support the possibility that TRPV4 channels are involved in flow-induced dilation. Kohler et al. (169) reported that administration of the TRPV4 agonist 4α-PDD or application of shear stress dilated rat gracilis arteries, and showed that these responses were blunted by the nonselective TRPV blocker ruthenium red. NOS inhibition attenuated flow-induced dilation but did not affect 4α-PDD-induced dilation, which was diminished by block of SK and IK channels. These results suggest that shear stress-induced Ca2+ influx through TRPV4 channels causes dilation through increased NO production, but pharmacological activation of the channel initiates EDHF or EDH pathways (169). A further study by this group demonstrated that carotid arteries from TRPV4−/− mice lack shear stress-induced dilation (132). Zhang and colleagues (220) reported that TRPV4 is critically important for flow-mediated dilation of mouse mesenteric arteries and showed that shear stress induces Ca2+ influx through TRPV4 channels in human coronary artery endothelial cells. Bubloz and co-workers showed that flow-induced dilation of human coronary arteries was blunted by ruthenium red, the selective TRPV4 blocker RN-1732 (371), and siRNA-mediated downregulation of TRPV4 expression (41), providing strong evidence for a TRPV4 influence in shear stress-induced responses in this vascular bed.

It remains unclear if TRPV4 channels are directly activated by shear stress or respond to signaling pathways activated by a flow-sensing apparatus. In support of the latter possibility, a few studies have provided evidence that TRPV4 channels are activated by metabolites produced in response to laminar shear stress. Loot et al. (204) showed that inhibition of CYP epoxygenase attenuated flow-induced dilation of mouse carotid arteries, suggesting that EETs produced in response to shear stress are responsible for TRPV4-mediated dilation. These findings are supported by reports showing that cell-swelling-induced activation of TRPV4 is blocked by selective inhibition of CYP2C9 (372) and flow-induced dilation of mouse carotid arteries is impaired by inhibition of PLA2 (132), an enzyme that liberates arachidonic acid from the plasma membrane. Together, these studies support a mechanism in which TRPV4 is indirectly activated by shear through conversion of PLA2-liberated arachidonic acid to EETs by CYP2C9. In this scheme, the molecular nature of the endothelial cell flow sensor responsible for activation of TRPV4 in response to increased shear remains unresolved. The AT1R, which is present in the endothelium and can be activated by a mechanical stimulus independently of ANG II, is a potential candidate for this function. AT1Rs are functionally coupled to TRPV4 channels in vascular smooth muscle cells (221), but this linkage has not yet been demonstrated in endothelial cells.

III) TRPV4 and agonist-induced dilation. Several studies have demonstrated involvement of TRPV4 channel activity in endothelium-dependent relaxation of systemic arteries in response to muscarinic and purinergic receptor agonists. Saliez et al. (295) reported that EDHF/EDH and NO-dependent relaxation in response to carbachol are diminished in mesenteric arteries from TRPV4−/− mice compared with arteries from wild-type animals, indicating that TRPV4 channels are involved in the response. Using pressure myography and in vivo approaches, Gutterman and colleagues (417) confirmed that acetylcholine (ACh)-induced dilation was blunted in mesenteric arteries from TRPV4−/− animals. This study also demonstrated that ACh-induced Ca2+ influx and NO production in endothelial cells isolated from TRPV4−/− mice were decreased compared with that in wild-type controls (417). Senadheera et al. (306) showed that block of TRPV4 with RN-1734 blunted ACh-induced dilation of radial uterine arteries, producing a greater effect in arteries from pregnant rats. Marrelli et al. (211) found that UTP-induced endothelial cell Ca2+ influx and dilation of cerebral arteries was blunted by ruthenium red and block of PLA2, suggesting that PLA2 may serve to generate arachidonic acid, which activates TRPV4-mediated Ca2+ influx and leads to arterial dilation through an EDHF or EDH mechanism. Chen et al. (50) showed that GSK1016790A induced a dilation of isolated renal arteries and perfused kidneys that was largely endothelium dependent and sensitive to NOS inhibitors.

TRPV4 channels are also present in the pulmonary artery endothelium. The selective TRPV4 agonist GSK1016790A relaxes pulmonary artery rings in an endothelium-dependent manner, an effect that is blocked by the selective TRPV4 antagonist HC-067047 and is largely dependent on SK and IK channel activity. These data indicate that pharmacological activation of TRPV4 causes EDH in pulmonary arteries. However, ACh-induced relaxation of pulmonary artery rings was not affected by block of TRPV4 channels with HC-067047 or RN1734 in this study, indicating that TRPV4 channels are not involved in muscarinic agonist-induced responses in isolated pulmonary arteries. Similar results have been obtained in vivo (257), indicating that TRPV4 channels are probably not involved in dilation of pulmonary arteries in response to muscarinic agonists.

IV) TRPV4 channels and subcellular Ca2+ signaling in the endothelium. An elegant study by Nelson and colleagues (324) reported that TRPV4 sparklets could be recorded from the intact endothelium of mesenteric arteries isolated from mice expressing a genetically encoded Ca2+ indicator protein exclusively in the endothelium. The amplitudes of TRPV4 sparklets stimulated by 4α-PDD, 11,12-EET, or GSK1016790A displayed a quantal distribution and exhibited cooperative gating, with simultaneous activation of two, three, or four channels occurring more frequently than predicted by binomial distribution models (324). Events with amplitudes greater than four quantal levels were not observed, suggesting that TRPV4 is present in a four-channel metastructure in endothelial cells. This coupled gating arrangement amplifies the initial Ca2+ signal created by TRPV4 channel activation and, combined with the biophysical properties of the channel, which favor Ca2+ influx, produces large-amplitude Ca2+ signals in the endothelium. Although the amplitude of individual TRPV4 sparklets is quite large, the total number of active sites per cell is surprisingly low. Both Sonkusare et al. (324) and Sullivan et al. (337) showed that only a few (~3–8 sites) TRPV4 sparklets are active in individual endothelial cells during maximal, agonist-induced dilation. TRPV4 sparklets stimulated IK and SK currents in native endothelial cells, indicating that TRPV4 sparklets activate an EDH-dependent vasodilator pathway (324). Muscarinic receptor activation was shown to increase TRPV4 sparklet activity, supporting the hypothesis that large-amplitude Ca2+ signals generated by coupled TRPV4 channels activate proximal SK and IK channels to cause vasodilation (Figure 10) (324).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240010.jpg

TRPV4 Ca2+ sparklets in endothelial cells initiate vasodilation. TRPV4 Ca2+ influx from the extracellular space through as few as 3–4 TRPV4 channels activates Ca2+-activated K+ (KCa) channels, resulting in K+ efflux from the cell. This hyperpolarizes the endothelial cell (EC) membrane, which directly hyperpolarizes smooth muscle cells (SMCs) via gap junctions located within myoendothelial junctions. IEL, internal elastic lamina; Vm, membrane potential. [From Sullivan and Earley (336).]

A report from the Dora lab (22) suggested that Ca2+ influx through TRPV4 channels could trigger CICR in the endothelium. This study showed that TRPV4 channels cluster in MEPs in mesenteric arteries and are spatially coupled with Ca2+ pulsars sites. Ca2+ pulsar activity was greater at very low intraluminal pressure (5 mmHg) compared with physiological levels (80 mmHg). The effects of low pressure on Ca2+ pulsar activity were attenuated by TRPV4 inhibition. Block of TRPV4 channels also increased myogenic tone at low levels of intraluminal pressure (20 and 40 mmHg), suggesting that the channel mediates a Ca2+-pulsar-dependent vasodilator influence at these pressures. The authors proposed that Ca2+ influx through TRPV4 channels stimulated by low intraluminal pressure is amplified by CICR from IP3Rs to generate large-amplitude Ca2+ pulsars that promote EDH and thereby suppress myogenic constriction. The significance of these data is unclear, but it is conceivable that activation of endothelial cell TRPV4 channels at low intraluminal pressure and subsequent EDH and/or increases in vascular permeability contribute to the precipitous drop in blood pressure observed during hypovolemic shock. In contrast to the findings of Bagher et al. (22), Senadheera et al. (306) did not detect enriched TRPV4 expression within IEL fenestrations of uterine arteries from control or pregnant rats, suggesting that the site of TRPV4 expression in endothelial cells may differ between vascular beds.

V) TRPV4 function in vivo. Many of the studies cited above used ex vivo preparations to demonstrate that TRPV4 channels present in the endothelium contribute to vascular regulation in response to shear stress, receptor-dependent agonists, and autocrine/paracrine factors. These data are at odds with a number of in vivo studies showing little involvement of TRPV4 channels in cardiovascular control. Two reports showed that systemic activation of TRPV4 with GSK1016790A caused a significant drop in mean arterial pressure (MAP) that was independent of changes in cardiac output (257, 389), indicating that activation of the channel lowered total peripheral resistance in whole animals. These findings suggest that loss of TRPV4 activity in vivo should increase total peripheral resistance and MAP. However, basal MAP did not differ between wild-type and TRPV4−/− mice (342) or was slightly lower in TRPV4−/− mice (254). Furthermore, acute or chronic (8 days) administration of the selective TRPV4 blocking compounds GSK2193874 and GSK2263095 had no effect on MAP or heart rate, indicating that inhibition of the channel has no significant effect on basal cardiovascular regulation (350). GSK2193974 also had no effect on decreases in MAP following intravenous injection of ACh, suggesting that TRPV4 channels are not involved in systemic muscarinic receptor-induced vasodilation in vivo (257). The only reported cardiovascular phenotype of TRPV4−/− mice is a mildly enhanced sensitivity to hypertensive stimuli (81, 254). Together, these in vivo studies suggest that TRPV4 channels in the endothelium are essentially inactive in healthy animals under basal conditions and exert only a modest effect on vascular resistance and MAP. The reason for the inconsistency between in vivo and ex vivo data is not clear, but it is possible that the influence of TRPV4 channels is confined to specific vascular beds, with other segments rapidly compensating for the loss of channel activity during pharmacological inhibition in whole animal studies.

d) trpa1. TRPA1 channels have a large unitary conductance (~98 pS) (241), are highly permeability to Ca2+ (PCa:PNa ~7.9) (157), and are activated by electrophilic compounds, including the dietary molecules allicin (208), found in garlic, and allyl isothiocyanate (AITC) (24), derived from mustard oil. The channel acts as a receptor for these pungent substances in sensory neurons and may also be involved in responses to noxious cold (329). Recent studies have demonstrated that TRPA1 contributes to vascular regulation (for review, see Ref. 75). Bautista et al. (27) reported that allicin dilated mesenteric arteries by activating TRPA1 channels in perivascular nerves to cause the release of calcitonin gene-related peptide (CGRP). Earley and co-workers subsequently found that TRPA1 channels are present in the cerebral artery endothelium of rodents and humans (77), but were not detectable in mesenteric, coronary, dermal, or renal arteries (338). This study further reported that AITC caused a robust endothelium-dependent dilation and smooth muscle cell membrane hyperpolarization in pressurized cerebral arteries that was blocked by the selective TRPA1 inhibitor HC-030031. These AITC-induced responses were insensitive to NOS and COX blockade but were diminished by inhibition of SK, IK, or KIR channels (77). This study also showed that TRPA1 channels are abundant in MEPs in the cerebral endothelium and are colocalized with IK channels in this tissue (77). Further insight into TRPA1-mediated cerebral artery dilation was provided by Qian et al. (281), who demonstrated that activation of TRPA1 channels provoked dilation by recruiting large, dynamic Ca2+ signals in the cerebral artery endothelium. These signals, generated by the release of Ca2+ from IP3Rs present on the endothelial cell ER, are distinguished from sparklets and Ca2+ pulsars by amplitude, spatial spread, and duration and appear to mediate AITC-induced vasodilation.

It is unlikely that TRPA1 channel activity in the cerebral endothelium is regulated by AITC or allicin under physiological conditions. TRPA1 channels are activated by reactive oxygen species (ROS), such as H2O2, and ROS-derived metabolites, including 4-hydroxy-2-nonenal (4-HNE), 4-oxo-nonenal (4-ONE), and 4-hydroxylhexenal (4-HHE) that are generated by peroxidation of ω6 polyunsaturated fatty acids (17, 363). Sullivan et al. (338) showed that TRPA1 channels and the ROS-generating enzyme NOX2 (NADPH oxidase isoform 2) colocalize in the endothelium of cerebral arteries. This study also showed that the amplitude of TRPA1 sparklets recorded from cerebral artery endothelial cells was approximately twice that of TRPV4 sparklets, consistent with the channel's larger unitary conductance (~98 vs. 60 pS) and higher Ca2+ selectivity (~7.9:1 vs. 6:1). Stimulation of NOX2 activity increased the frequency of TRPA1 sparklets and dilated isolated cerebral arteries. These responses were dependent on the generation of extracellular H2O2 and hydroxyl radicals replicated by administration of exogenous 4-HNE. Together, these data suggest that superoxide anions (O2) generated by NOX2 activity proximal to TRPA1 channels increase TRPA1 sparklet activity to cause vasodilation. This mechanism appears to be present only in the cerebral circulation and could provide compensatory increases in blood flow during cerebrovascular diseases associated with oxidative stress.

e) trpc3. TRPC3 channels have a unitary conductance of ~60 pS, are essentially equally permeable to Ca2+ and Na+, and are more permeable to Na+ than K+ (156). The channel is activated by GPCR pathways, is directly sensitive to DAG (and stable analogs), and may be stimulated by elevated intracellular [Ca2+]. Reports from several laboratories suggest involvement of TRPC3 in endothelium-dependent vasodilation. An early study of isolated rat mesenteric arteries showed that injection of animals with TRPC3 antisense oligonucleotides diminished flow- and bradykinin-induced vasodilation (198). A later report by Gao et al. (102) provided evidence for expression of TRPC3 in the endothelial and smooth muscle layers of the human internal mammary artery (IMA), and showed that the TRPC3 blocker Pyr3 modestly diminished muscarinic agonist-induced relaxation of precontracted IMA rings. Kochukov et al. (166) used knockout animals to demonstrate that TRPC1 and TRPC3 subunits form heteromultimeric channels involved in muscarinic agonist-induced Ca2+ influx and relaxation of aortic ring segments. A subsequent study by Sandow et al. (307) reported participation of TRPC3 channels in endothelium-dependent dilation of rat and mouse mesenteric arteries. This latter study demonstrated that TRPC3 channels are present in the mesenteric arterial endothelium, with ~70% of the channels localizing to MEPs. The selective TRPC3 blocker Pyr3 blunted acetylcholine-induced vasorelaxation and endothelial cell hyperpolarization in the presence of the NOS inhibitor l-NAME, the guanylyl cyclase inhibitor ODQ, and the COX inhibitor indomethacin. The IK channel blocker TRAM34 and the SK channel blocker apamin also attenuated vasodilatory responses. Together, these data suggest that TRPC3 channel activity stimulates EDH in mesenteric arteries following muscarinic receptor activation. A study by Kirby et al. (161) reported the presence of TRPC3 as well as SK and IK channels within IEL fenestrations of rat popliteal and first-order skeletal muscle arteries from the gastrocnemius muscle, but the function of TRPC3 in this arterial bed was not assessed. Marrelli and co-workers (165) demonstrated that TRPC3 channels contribute to EDH-mediated, SK and IK channel-dependent dilation of cerebral arteries in response to ATP. Interestingly, this study provided evidence that receptor activation leads to rapid recruitment of TRPC3 channels to the plasma membrane (165).

Taken together, the findings cited above support a mechanism whereby activation of muscarinic GPCRs stimulates TRPC3 activity, mostly likely through PLC-dependent generation of DAG. TRPC3 activity leads to increases in intracellular Ca2+ that activate SK and IK channels to cause EDH. However, the Ca2+ fraction of TRPC3-mediated mixed cation currents is predicted to be small at the endothelial cell resting membrane potential, arguing against a direct Ca2+ influx mechanism for SK and IK channel activation. The explanation for this apparent discrepancy is not clear, but it is possible that TRPC3 subunits form part of a heteromultimeric, Pyr3-sensitive channel with greater Ca2+ permeability. Another possibility, reviewed by Eder et al. (84), is that Na+ influx through TRPC3 indirectly activates Ca2+ influx through the Na+/Ca2+ exchanger (NCX) acting in reverse mode. This mechanism requires colocalization of NCX and TRPC3 to produce subcellular domains with locally elevated [Na+] capable of activating reverse-mode Ca2+ entry without causing membrane depolarization, which would diminish the electrical driving force for Ca2+ entry. Support for TRPC3-mediated NCX reverse-mode Ca2+ entry is provided by a study from Rosker et al. (289), who showed that elevated extracellular [Na+] caused Ca2+ influx in HEK cells overexpressing TRPC3 channels. Na+-induced Ca2+ influx was blocked by KB-R7943, a compound that preferentially inhibits reverse-mode NCX activity. Using a coimmunoprecipitation approach, these authors (289) also demonstrated that endogenously expressed NCX1 binds to the COOH terminus (amino acids 742–848) of heterologously expressed TRPC3. It has also been reported that TRPC3 associates with NCX1 and KB-R7943 blocks ANG II-induced Ca2+ influx in adult cardiomyocytes (85). However, KB-R7943 has off-target effect: it is a potent blocker of TRPC3, TRPC5, and TRPC6 channels (171) and a strong Ca2+-independent activator of BK channels in smooth muscle cells (191). Thus a convincing demonstration of the proposed TRPC3-NCX Ca2+-entry pathway and possible functional consequences awaits the development of more selective pharmacological agents.

f) trpc4. TRPC4 channels are equally permeable to Na+ and Ca2+ and are activated by GPCR pathways, although the specific mechanisms are not fully understood. Freichel et al. (97) reported involvement of TRPC4 channels in endothelium-dependent vasodilation. This study demonstrated that SOCE could not be evoked in endothelial cells isolated from TRPC4−/− mice, and ACh-induced Ca2+ entry in these cells was diminished compared with that in wild-type controls. Aortic ring sections from TRPC4−/− mice displayed blunted endothelium-dependent relaxation in response to ACh, suggesting that TRPC4-mediated Ca2+ influx contributes to endothelium-dependent dilation in conduit arteries (97). The authors proposed that SOCE through TRPC4 stimulates eNOS activity and NO production to induce vasodilation (98). This conclusion is tempered by subsequent reports demonstrating that TRPC4 channels are not necessary for SOCE in endothelial cells (2, 316).

2. Permeability of the vascular endothelium

The vascular endothelium forms a semi-permeable barrier that regulates passage of water, ions, and macromolecules between the blood and the interstitium (219, 222). The degree of permeability varies considerably among vascular beds and segments and is influenced by physiological and pathophysiological conditions. Substances are transferred through the vascular wall by passive diffusion and active transport mechanisms, and through clefts between neighboring endothelial cells by bulk flow. Transport across the endothelium by bulk flow is governed by mechanical coupling between adjacent endothelial cells imparted by two types of protein complexes, tight junctions and adherens junctions, linked to the actin cytoskeleton. Ca2+-dependent cytoskeletal reorganization of these interendothelial cell junctions dynamically regulates permeability of the endothelium in response to GPCR agonists such as thrombin, bradykinin, and histamine. TRPC1, TRPC4, TRPC6, and TRPV4 channels have been reported to influence endothelial permeability, as discussed below.

a) trpc1. TRPC1 channels have a unitary conductance of ~5 pS and are not selective for Ca2+ relative to Na+ in heterologous expression systems. It is unclear if homomeric TRPC1 channels form in native cells, but TRPC1 subunits have been shown to form heteromultimeric channels with TRPC5 (334), TRPC4 (334), TRPC3 (333), TRPP2 (164, 419), and TRPV6 (301). TRPC4, TRPC1, and/or heteromultimeric channels containing TRPC1 subunits have been implicated in the control of endothelial permeability in the pulmonary vasculature through regulation of SOCE. Inhibition of SERCA pump activity causes a large increase in the permeability of the pulmonary endothelium (56) that is associated with a change in the shape of pulmonary artery endothelial cells (232). Brough et al. (39) showed that SOCE in pulmonary artery endothelial cells was diminished following antisense-mediated knockdown of TRPC1 expression, suggesting involvement of the channel in this response. Mehta el al. (218) provided evidence that SOCE in cultured endothelial cells was diminished by inhibition of RhoA activity and also showed that RhoA physically interacts with IP3Rs and TRPC1 channels (218). In cells overexpressing TRPC1, stimulation with thrombin was found to result in PKC-α-mediated phosphorylation of TRPC1 channels in association with SOCE and increased transendothelial permeability (6). In human umbilical cord endothelial cells in culture, the cytokine tumor necrosis factor-α (TNF-α) was shown to increase TRPC1 expression accompanied by increased SOCE and stress fiber formation (261).

b) trpc4. Evidence supporting a role for TRPC4 channels in the regulation of pulmonary endothelial permeability was first provided by Tiruppathi et al. (352), who demonstrated that sustained increases in intracellular Ca2+ in response to thrombin receptor activation were diminished in endothelial cells from the lungs of TRPC4−/− mice compared with controls. Diminished thrombin-induced Ca2+ influx was associated with decreased actin-stress fiber formation and endothelial cell retraction. In isolated lung preparations, basal microvascular permeability did not differ between control and TRPC4−/− mice, but thrombin receptor activation-induced increases in lung permeability were significantly diminished in lungs from TRPC4−/− mice. The authors concluded that Ca2+ store depletion following thrombin receptor stimulation causes a TRPC4-mediated influx of Ca2+ that contributes to cytoskeletal rearrangement and formation of intercellular gaps in the endothelium. Cioffi et al. (59) demonstrated that the interaction between spectrin-protein 4.1 complexes and TRPC4 channels was required for SOCE in pulmonary artery endothelial cell cultures. The formation of intercellular endothelial cell adhesions promotes recruitment of TRPC4 channels to the plasma membrane. The authors of this study speculated that Ca2+ influx through TRPC4 may contribute to the maturation of junctional complexes (120).

c) trpc6. Several studies have implicated TRPC6 channels in the regulation of endothelial permeability. Singh et al. (320) reported that TRPC6 knockdown diminished thrombin-induced stress fiber formation and endothelial barrier function in pulmonary artery endothelial cells. Samapati et al. (297) provided evidence that TRPC6 channels contribute to the regulation of pulmonary vascular permeability. This study demonstrated that TRPC6 channels are enriched in pulmonary endothelial cell caveoli. The TRPC6 activator hyperforin (182, 237) elevated intracellular Ca2+ levels in primary lung endothelial cells and increased the wet weight of isolated perfused rat lungs. Furthermore, platelet-activating factor (PAF)- and sphingomyelinase (ASM)-induced increases in endothelial cell Ca2+, lung vascular filtration coefficient, and wet weight gain were absent in TRPC6−/− mice. The effects of PAF on endothelial cell Ca2+ levels and hyperforin-induced inward cation currents were diminished by administration of an NO donor or a membrane-permeant cGMP analog, and hyperforin-induced increases in the wet weight of isolated perfused lungs were augmented by NOS blockade. The authors concluded that NO (via cGMP) inhibits TRPC6 channels in endothelial cells. Stimulation of the PAF/ASM signaling pathway recruits TRPC6 channels to caveoli and silences eNOS activity, allowing TRPC6-mediated Ca2+ influx to increase vascular permeability.

d) are trpc1, trpc4, and trpc6 channels necessary for endothelial cell soce? Although many studies cited above have put forward the concept that SOCE mediated by TRPC1 and TRPC4 channels can increase the permeability of the vascular endothelium, this interpretation is called into question by work from the Trebak lab showing that endothelial cell SOCE does not require TRPC1, TRPC4 or TRPC6 expression (2, 316), despite the fact that knockdown of these TRPC channels blocks thrombin-induced increases in endothelial permeability (316). This latter study also demonstrated that STIM1 independently controls endothelial barrier function by a mechanism that does not require TRPC1 or TRPC4 channels, Orai1, or Ca2+ entry (316). These data indicate that TRPC channels contribute to the regulation of endothelial permeability by a mechanism that is independent of SOCE. A conflicting study by Cioffi et al. (60) reported that heteromultimeric channels containing TRPC4 and TRPC1 subunits are present in pulmonary artery endothelial cells and physically interact with Orai1 via TRPC4; they further showed that Oria1 activity is necessary for the formation of interendothelial cell gaps. In a similar vein, Sundivakkam et al. (341) showed that STIM1 interacts with TRPC1 and TRPC4 to form a store-operated channel in cultured endothelial cells. Although the contradictory results regarding TRPC1 and TRPC4 channels and SOCE currently remain unresolved, strong evidence supports the concept that endothelial cell SOCE can occur without TRPC1, TRPC4, and TRPC6 expression (359). The findings of the Trebak laboratory are supported by the results of unbiased high-throughput siRNA screens, which revealed the necessity for STIM1 and Orai1 in SOCE but did not demonstrate the involvement of any TRP channels (370). It has also been reported that Ca2+ release from internal stores persists in Purkinje neurons from TRPC1/C4/C6 triple-knockout mice, suggesting that these channels are not required for the maintenance of intracellular stores via SOCE (131). Furthermore, the biophysical properties of TRPC channels suggest that fractional Ca2+ currents are probably very small under physiological conditions. It should be noted that many of the studies linking SOCE with endothelial permeability have presented only data showing changes in global intracellular Ca2+ and provide little or no evidence for direct TRPC channel activation using biophysical or electrophysiological methods. For SOCE studies that have used the patch-clamp technique, the apparent involvement of TRPC channels under the conditions used could reflect PLC-mediated activation of TRPC channels downstream of store depletion (2). In addition, many of the studies supporting a role for TRPC channels in SOCE-mediated Ca2+ influx were performed using endothelial cells in culture, and in some cases, overexpression systems. A study by Bergdahl et al. (31) showed that culture conditions significantly increased TRPC1 and TRPC6 expression and SOCE in smooth muscle cells. The effects of culture conditions on endothelial cell TRPC4 expression have not been reported, but phenotypic transformations that influence SOCE and permeability regulation are possible. Additional work using intact artery preparations is needed to fully resolve these issues.

e) trpv4. TRPV4 is present in pulmonary endothelial cells in both humans and rodents (15), and studies from several laboratories have convincingly demonstrated an important role for the channel in the regulation of pulmonary vascular permeability. A study from by Townsley and colleagues (15) showed that treatment of isolated lung preparations from wild-type mice with 4α-PDD, 5,6-EET, or 14,15-EET increased microvascular permeability. That this increased microvascular permeability was mediated by TRPV channels, specifically TRPV4, was confirmed by the fact that it was attenuated by the nonselective TRPV blocker ruthenium red and was unaltered in response to 4α-PDD in lungs from TRPV4−/− mice. However, SOCE-induced permeability responses did not differ between TRPV4−/− and wild-type mice. The overall 4α-PDD and EET-induced increases in permeability were comparable to those induced by SOCE, but these treatments influenced different vascular segments. TRPV4 agonists primarily affected alveolar septal microvessels, whereas the effect of SOCE was greatest in extra-alveolar vessels. Moreover, TRPV4 agonists caused breaks in the barrier and blebbing of endothelial cells, whereas SOCE caused the formation of gaps at cellular junctions. This study clearly showed that TRPV4-mediated Ca2+ influx and SOCE have differential effects on the pulmonary endothelium.

The concept of functional specificity of Ca2+-influx pathways in the pulmonary endothelium was further explored in a study that compared the effects of endothelial cell Ca2+ entry through T-type Ca2+ channels with that of TRPV4. This study showed that Ca2+ entry via T-type Ca2+ channels, but not TRPV4-mediated Ca2+ entry, induced surface expression of the adhesion molecule P-selectin in pulmonary capillary endothelium. In contrast, TRPV4-dependent, but not T-type Ca2+ channel-dependent, Ca2+ influx increased capillary permeability (397). Inhibition of myosin light-chain kinase (MLCK) was shown to diminish cell surface levels of TRPV4 protein and decrease 4α-PDD-induced Ca2+ influx and cation currents in pulmonary vascular endothelial cells (264), suggesting that MLCK activity is involved in trafficking of TRPV4 channels to the plasma membrane in the pulmonary endothelium. Using real-time Ca2+ imaging in isolated perfused lungs, Yin et al. (410) provided evidence that TRPV4 channels mediate increases in endothelial cell [Ca2+] in response to elevations in left atrial pressure. Increases in NO production induced by increases in left atrial pressure were absent in TRPV4−/− mice, and wet-to-dry weight ratios of lungs subjected to pressure elevation were smaller in TRPV4−/− mice compared with those in wild-type animals. This study also showed that inward cation currents stimulated by 4α-PDD in pulmonary microvascular endothelial cells were attenuated by pretreatment with a cGMP analog. The authors proposed that pressure-induced activation of TRPV4 channels causes Ca2+ influx, which elevates endothelial permeability and increases NO production. NO, acting through cGMP, acts as a negative-feedback mechanism to govern TRPV4-mediated Ca2+ influx.

3. Angiogenesis

Angiogenesis, the process of new blood vessel development, occurs in a tightly regulated manner during embryonic growth and wound healing, and during pathological conditions such as diabetes, arthritis, and cancer (96). Angiogenesis is initiated by remodeling of the basement membrane followed by proliferation and migration of endothelial cells, resulting in the generation of new capillary tubes. The angiogenic process is stimulated by several growth factors, including vascular endothelial growth factor (VEGF), fibroblast growth factor (FGF), and epidermal growth factor (EGF). Blocking Ca2+ entry diminishes endothelial cell adhesion, motility, and tube formation in culture and angiogenesis in vivo (170), but the influx pathways responsible are not well characterized. However, a few studies have demonstrated the involvement of TRPC1, TRPC3, TRPC4, TRPC5, TRPC6, and TRPV4 channels.

a) trpc1. Using an in vivo zebrafish model, Yu et al. (411) demonstrated that knockdown of TRPC1 disrupted angiogenesis in developing larvae. This study showed that TRPC1 is a downstream target of VEGF-a, and its expression is necessary for extracellular signal-regulated kinase (ERK) activation. In contrast to these findings, knockdown of TRPC1 expression had no significant effect on the in vitro formation of human umbilical vein-derived endothelial tubes (19), and the vasculature was reported to develop normally in TRPC1−/− mice (304).

b) trpc3, trpc4, and trpc5. Antigny et al. (19) reported that siRNA-mediated silencing of TRPC3, TRPC4, or TRPC5 expression diminished spontaneous Ca2+ oscillations in human umbilical vein-derived endothelial cells and inhibited endothelial tube formation in vitro, suggesting a possible role for these TRPC channels in angiogenesis.

c) trpc6. Endothelial cell TRPC6 channels are activated by binding of the pro-angiogenic factor VEGF to VEGF2 receptors (54), suggesting that TRPC6-mediated cation currents could be involved in the angiogenic response. In agreement with this possibility, Hamdollah Zadeh et al. (128) reported that VEGF-induced migration and tube sprouting of microvascular endothelial cells was diminished in cultures expressing a dominant-negative form of TRPC6 (110). Further support is provided by a study by Kini et al. (160), who showed that TRPC6 associates with PTEN (phosphatase and tensin homologue) protein in human pulmonary artery endothelial cells and demonstrated that this interaction is required for surface expression of TRPC6 and thrombin-induced proliferation and tube formation. In contrast, neither TRPC6 gain-of-function mutations in humans (391) nor TRPC6 gene deletion in mice (70) was reported to have any overt effect on blood vessel structure.

d) trpv4. Several studies have examined the concept that activation of TRPV4 channels contributes to angiogenesis. Knockdown of TRPV4 expression was shown to prevent cyclic strain-induced Ca2+ influx and capillary endothelial cell reorientation, suggesting involvement of the channel in migration and sprouting associated with angiogenesis (349). Interestingly, administration of the TRPV4 agonist 4α-PDD enhanced collateral growth of cerebral blood vessels, suggesting involvement of the channel in cerebral angiogenesis (300). Total and surface TRPV4 expression and channel activity were found to be elevated in breast tumor-derived endothelial cells compared with normal dermal microvascular endothelial cells (92), and activation of TRPV4 channels with arachidonic acid promoted migration of tumor-derived, but not normal endothelial cells. Collectively, these observations suggest that TRPV4 channels may be involved in tumor angiogenesis.

C. TRP Channels in Perivascular Cells

TRP channels are present in perivascular neurons and astrocytes. Several studies have indicated that TRPV1, TRPA1, and perhaps TRPV4 channels in perivascular neurons can cause vasodilation, and one study demonstrated that TRPV4 channels present in astrocytic endfeet enveloping parenchymal cerebral arteries influence neurovascular coupling.

1. TRPV1 in perivascular sensory nerves

Neuronally evoked vasodilation is a common feature of the vasculature in many organ systems, including the brain, cutaneous tissues, and the gastrointestinal tract (33). A model describing a vasodilator role for TRPV1 channels located on perivascular sensory nerves has emerged during the past decade. Data from several research groups have demonstrated that activation of TRPV1 channels causes release of the potent vasodilator CGRP from sensory nerves, leading to vasodilation. This interaction between TRPV1 channels and CGRP activity is most frequently reported for the mesenteric resistance vasculature (335, 430, 376, 378), but also clearly applies to other vascular beds, such as the dural (7) and cerebral (430) circulations. The endogenous endocannabinoid anandamide has been strongly implicated as a potential physiological activator of the above TRPV1/CGRP vasodilator pathway, acting through direct activation of TRPV1 channels on sensory nerves rather than through activation of cannabinoid-specific receptors (7, 38). A similar vasodilatory mechanism in the coronary circulation, initiated by ethanol, has been described (109). Activation of TRPV1 channels has also been reported to inhibit sympathetic vasoconstrictor tone through the release of CGRP (412). Similar roles for TRPA1 (27, 429) and TRPV4 (101) channels in mesenteric perivascular sensory nerves have been reported (Figure 11).

An external file that holds a picture, illustration, etc.
Object name is z9j0021527240011.jpg

TRP channels in perivascular cells. A: Ca2+ influx through TRPV1 or TRPA1 channels in perivascular nerves can stimulate the release of CGRP to cause arterial dilation. B: TRPV4 channels contribute to neurovascular coupling. Stimulation of metabotropic glutamate receptors (mGluR) on parenchymal astrocytes leads to the generation of IP3, which causes the release of Ca2+ from the endoplasmic reticulum (ER), stimulating K+ efflux through BK channels. The resulting increase in [K+] in the space between the astrocytic endfoot and cerebral arteriole activates K+ efflux through KIR channels present on vascular smooth muscle cells to cause dilation. Production of EETs or prostaglandins (e.g., PGE2) may also contribute. Ca2+ influx through TRPV4 channels contributes to the propagation of Ca2+ waves throughout the endfoot.

2. TRPV4 channels and neurovascular coupling

TRPV4 activity in astrocytes has been linked to the local control of cerebral blood flow. Brain parenchymal astrocytes form close associations with neurons and with penetrating arterioles that link pial arteries on the surface of the brain with the subsurface microcirculation. Astrocytes have been proposed to serve as “middlemen” in the functional hyperemic process of neurovascular coupling, which closely matches blood perfusion with neuronal activity (318). Astrocytic endfeet envelop the parenchymal arterioles and can mediate profound vasodilation in response to elevated synaptic activity. A number of vasoactive factors, including EETs, NO, and K+, have been implicated in this response (74). Work from the Nelson laboratory has shown that K+ efflux through BK channels present on astrocytic endfeet acts through elevation of [K+] in the restricted space between the endfoot and the arteriolar wall to activate KIR channels on smooth muscle cells, causing membrane hyperpolarization and vasodilation (91) (Figure 11). This response is dependent on increased astrocytic intracellular [Ca2+] (330). Ca2+ entry through activated TRPV4 channels located in astrocytic endfeet (29) stimulates CICR from intracellular stores and enhances arterial dilation (73). TRPA1 (314, 315) and TRPC3 (331) channels are also present in astrocytes and regulate intracellular Ca2+, but their involvement in neurovascular regulation has not been reported.

IV. INVOLVEMENT OF TRP CHANNELS IN VASCULAR PATHOLOGIES

Given the clear physiological roles for TRP channels in vascular smooth muscle, the endothelium, and perivascular cells, it is not surprising that there are multiple examples where the function or dysfunction of various TRP channels contributes to cardiovascular disease (see Table 4). The literature describing the involvement of TRP channels in vascular-related diseases is discussed below.

Table 4.

TRP channels involved in vascular pathologies

TRPPathologyTissue
TRPC1Hypertension, vascular remodeling/VSM proliferationLung (113, 194, 200, 374), large veins (184, 347), brain (31)
TRPC3Hypertension, vascular remodeling/VSM proliferationMesentery (4, 51), brain (256), lung (413)
TRPC4Diminished risk of MI, vascular remodeling/VSM proliferationEndothelium (153), lung (421)
TRPC6Hypertension, I/R-induced pulmonary edema, vascular remodeling/VSM proliferationMesentery (277, 428), lung (194, 413)
TRPV1Hypertension, chronic hypoxia-induced proliferation, cardiac and cerebral I/R injury, neurogenic inflammationSystemic (185, 379), lung (380), heart (375), brain (239, 265)
TRPV3Cerebral I/R injuryBrain (235)
TRPV4Hypertension, APP-associated endothelial dysfunction, pulmonary edemaBrain (418), lung (127), mesentery (325)
TRPA1Oxidative stress response, neurogenic inflammationBrain (Sullivan et al., unpublished data), sensory nerves (18, 244, 346, 361)
TRPM2Oxidative stress responseEndothelium (135, 345)
TRPM4Cerebral ischemiaCNS microvessels (112), brain (203)
TRPM7Hypertension, neuronal death following cerebral ischemia, vascular remodeling/VSM proliferationMesentery (357, 358), brain (339), epidermis (134)

VSM, vascular smooth muscle; MI, myocardial infarction; I/R, ischemia reperfusion; CNS, central nervous system. Reference numbers are given in parentheses.

A. Hypertension

Although the causes of essential hypertension are still not established, alterations in central cardiovascular control mechanisms as well as cardiovascular, hormonal, and fluid and electrolyte adjustments typically associated with renal disease are clearly involved in the development of this disease (44). Specific features of vascular dysfunction are prevalent in virtually all situations where sustained blood pressure elevations are present. The primary manifestation of hypertension in the vasculature is an increase in peripheral resistance resulting from enhanced vascular contractility, differing degrees of vascular remodeling in the resistance vasculature, or both. Although TRP channel dysfunction may be involved in hypertension driven by neural (339), humoral (214), or renal (395) mechanisms, the following discussion will focus entirely on evidence that TRP channels in the vasculature contribute to the pathophysiological processes associated with elevated blood pressure.

1. TRPC channels

There are multiple examples of altered TRPC channel expression and function, primarily TRPC3 and TRPC6, associated with hypertension.

a) trpc3. Several studies have reported that TRPC3 expression is upregulated in arteries from hypertensive animals, a phenomenon most clearly documented in the SHR model of hypertension. Increased TRPC3 expression in vascular smooth muscle cells isolated from SHRs is correlated with enhanced ANG II-induced Ca2+ influx that is independent of VDCCs (199). A similar increase in carotid artery smooth muscle TRPC3 channel expression in SHRs accompanied by enhanced vascular contractility has also been described (256). Interestingly, this latter study found that TRPC1 expression was decreased. Others have reported increased expression of TRPC3 as well as TRPC1 and TRPC5 in mesenteric arterioles from SHRs compared with normotensive controls, with an associated increase in vasomotion (51). Enhanced ET-1-induced contractions of SHR mesenteric resistance arteries were apparently due in part to increased TRPC3 expression, which was associated with greater coupling of ET-1 receptor activity to Ca2+ entry through direct physical interactions between IP3Rs and TRPC3 channels (4). Although there are no specific reports of TRPC3 upregulation in the vascular smooth muscle cells of humans with hypertension, TRPC3 levels were found to be elevated in monocytes from patients with essential hypertension in association with enhanced migration activity of these cells (424). Also, although TRPC3 expression levels and hypertension were positively correlated in each of these studies, cause and effect analyses, for instance, examination of TRPC3 expression in prehypertensive or anti-hypertensive-treated SHRs, have not been reported. See Reference 377 for a comprehensive review of TRPC3 and hypertension.

b) trpc6. Upregulation of TRPC6 in vascular smooth muscle in association with hypertension has also been reported. In a rat model of ouabain-induced hypertension, cellular levels of TRPC6, TRPC1, the α2 subunit of the Na+ pump, and the Na+-Ca2+ exchanger are increased in mesenteric artery vascular smooth muscle cells (277). Increased and coordinated activity among these various Ca2+ transport systems is correlated with increases in [Ca2+]i, which could increase vascular contractility and contribute to elevated blood pressure. Similar changes in the expression of TRPC6 and the Na+-Ca2+ exchanger in mesenteric myocytes, and altered Ca2+ homeostasis occur in Milan hypertensive rats (195, 428). MAP is slightly elevated in TRPC6−/− mice compared with wild-type animals (70). This is unexpected, but may be due to the (apparently compensatory) upregulation of TRPC3 channels observed in TRPC6−/− mice.

Pulmonary hypertension is characterized by pulmonary arterial pressures greater than 25 mmHg and is due in part to heightened contractility of pulmonary artery smooth muscle cells as well as substantial cellular proliferation and remodeling of the pulmonary arterial wall. Both processes contribute to sustained increases in pulmonary artery resistance. Several studies have indicated that both TRPC6 and TRPC1 channels contribute to the mechanisms underlying chronic hypoxia-induced pulmonary hypertension (194, 374). For example, chronic (3–4 wk) hypoxia in rats was shown to increase TRPC6 and TRPC1 channel expression in pulmonary artery smooth muscle cells by two- to threefold (194). Chronic hypoxia also increased ROCE and SOCE; these responses were correlated with enhanced pulmonary artery contractile activity (399) and were inhibited by molecular or genetic suppression of TRPC6 and TRPC1 expression (194, 399). Others have demonstrated a central role for EETs, in particular, 11,12-EET, in the increased expression and activity of TRPC6 channels in this system, predominantly through enhanced trafficking and localization of TRPC6 channels to the plasma membrane in response to hypoxia (159, 272). A similar upregulation of TRPC6 and TRPC3 channels has been demonstrated in pulmonary artery myocytes from humans with idiopathic pulmonary hypertension (413, 414). Suppression of TRPC6 expression in pulmonary vascular smooth muscle cells cultured from these patients was shown to substantially reduce cellular proliferation. A role for TRPC6 in hypoxia-induced pulmonary hypertension is less apparent in mice, where TRPC1 may make a significant contribution (209, 385). TRPC1 channel expression and Ca2+ entry are both elevated in pulmonary arteries of rats with monocrotaline-induced pulmonary hypertension (200). Taken together, the preceding observations suggest that the common theme of upregulated TRPC6 and/or TRPC1 channel function and accompanying heightened vascular reactivity is associated with nearly all forms of hypertension.

2. TRPV channels

a) trpv1. TRPV1 channel function appears to be altered in response to hypertension and has a significant impact on the regulation of MAP. The TRPV1 channel agonist capsaicin was shown to decrease MAP in normal Wistar rats (185) and salt-sensitive Dahl rats (379) fed a high-salt diet; TRPV1 channel expression was also increased in these latter animals. Interestingly, MAP was further elevated in hypertensive rats on a high-salt diet following treatment with the TRPV1 channel antagonist capsazepine. These observations have led to the conclusion that TRPV1 receptors, likely those present predominantly on perivascular sensory nerves, are activated in animals with salt-sensitive hypertension. The associated release of CGRP from sensory nerve terminals provides a compensatory vasodilator response that opposes the elevation of blood pressure in this model of hypertension. There is no significant difference in systolic or diastolic blood pressures in TRPV1−/− compared with wild-type mice (212). This supports the hypothesis that altered TRPV1 activity occurs primarily in response to elevated blood pressure, but probably is not involved in the etiology of hypertension.

b) trpv4. A few studies have indicated the involvement of TRPV4 channels in pulmonary and systemic hypertension. In rat pulmonary arteries, chronic hypoxia-induced pulmonary hypertension was shown to upregulate TRPV4 channels (407). This increase in TRPV4 expression produced a feed-forward effect on pulmonary hypertension whereby Ca2+ entry through the expanded population of TRPV4 channels in pulmonary artery myocytes increased global [Ca2+]i, which enhanced pulmonary vascular contractility and pulmonary hypertension.

Elegant work from Nelson and colleagues (325) has shown that the coupling efficiency between endothelial cell TRPV4 channels and activation of EDH and dilation of mesenteric resistance arteries is disrupted in a mouse model of ANG II-induced chronic hypertension. Much of this effect was attributed to a profound decrease in the PKC-anchoring protein AKAP150 in the vicinity of myoendothelial gap junctions in arteries from normotensive mice, and the resulting disruption in the AKAP150-dependent clustering of TRPV4 channels that is responsible for the high degree of cooperative gating of localized TRPV4 channels. The authors of this groundbreaking study proposed that these hypertension-induced changes in TRPV4 function could contribute to the mechanisms by which high blood pressure alters local blood flow regulation.

3. TRPM7

Abundant evidence supports the hypothesis that altered regulation of Mg2+ transport contributes to the development and maintenance of hypertension (reviewed in Ref. 326). TRPM6 and TRPM7 play key roles in Mg2+ transport in numerous tissue types, including vascular smooth muscle. Recent observations point to a substantial role for TRPM channels, particularly TRPM7, and altered TRPM channel function in various forms of hypertension. TRPM7 expression and activity, as well as [Mg2+]i, are reduced in mesenteric artery smooth muscle from SHR compared with WKY rats (358). Decreased intracellular Mg2+ is associated with heightened contractility, which contributes to hypertension. Disrupted Mg2+ regulation involving altered TRPM7 activity is also correlated with changes in cellular inflammation, proliferation, and remodeling, each of which is associated with vascular disease in hypertension (357).

B. Vascular Remodeling, Intimal Hyperplasia, and Vascular Occlusive Disease

Vascular remodeling occurs in the context of a number of cardiovascular diseases, including hypertension, coronary artery disease, atherosclerosis, vascular injury, and inflammatory diseases. Increased intracellular [Ca2+] and activation of the δ isoform of Ca2+/CamKII seems to be a common mechanism associated with this type of cellular proliferative response. Signaling mechanisms downstream of CamKII activation are highly diverse, but likely involve altered expression of inhibitors and activators of cell-cycle progression, and altered gene transcription regulated by nuclear factor of activated T cells (NFAT) and cAMP response element binding protein (CREB) (319). TRP channels can facilitate increases in intracellular Ca2+ associated with activation of CamKII that are linked to vascular disease and remodeling.

1. TRPC1

Several studies provide evidence for the involvement of TRPC1 channels in disease-associated smooth muscle cell proliferation. In cultured human pulmonary artery myocytes, stimulation with serum-derived growth factors was shown to increase TRPC1 expression and Ca2+ influx in association with enhanced cellular proliferation (113). In cultured human saphenous vein myocytes, TRPC1 channel activity has been linked to smooth muscle cellular proliferation, an effect that is partly dependent on interactions with STIM1 (184). Other studies have shown that TRPC1 channel expression is upregulated in rodent models of vascular injury and myocyte proliferation in association with enhanced Ca2+ entry (31, 173). In the latter of these two studies, administration of an antibody targeting an extracellular loop of TRPC1 substantially reduced smooth muscle proliferation and Ca2+ entry in cultured human saphenous veins. Similarly, chronic stimulation of cultured human coronary artery smooth muscle cells with ANG II was shown to increase TRPC1 expression and SOCE in association with cellular hypertrophy, effects that were suppressed by siRNA-mediated knockdown of TRPC1 (347).

2. Other TRPCs

A few studies have implicated other members of the TRPC subfamily in smooth muscle cell proliferation. One report demonstrated that the proliferative response of human pulmonary artery smooth muscle to extracellular ATP, a mitogenic stimulant, occurred via CREB-mediated upregulation of TRPC4 and enhanced SOCE (421). Pulmonary artery smooth muscle from a subset of pulmonary hypertensive patients with idiopathic pulmonary arterial hypertension displayed enhanced expression of both TRPC3 and TRPC6. Proliferation of pulmonary artery myocytes cultured from these patients was greatly attenuated by suppression of TRPC6 expression (413). In addition, one group demonstrated that TRPC3 contributes to the cellular proliferation that occurs in the setting of stent-induced remodeling of large arteries (168).

3. TRPV1

One group reported that the proliferative response of cultured human pulmonary myocytes to chronic hypoxia was correlated with increased expression of TRPV1 channels and elevated SOCE (380). A role for TRPV1 in the remodeling of venous tissue exposed to altered pressure and flow has also been reported (52).

4. TRPM6 and TRPM7

Numerous studies have demonstrated that Mg2+ have a substantial impact on vascular smooth muscle and endothelial cell function (for review, see Ref. 326). In particular, changes in cytosolic Mg2+ levels may be involved in regulating vascular proliferation. TRPM6 and TRPM7 are critically important for Mg2+ transport in vascular smooth muscle and are involved in vascular diseases characterized by cellular proliferation and inflammation. Exposure to ANG II or aldosterone for 24 h was shown to substantially increase TRPM7 mRNA and protein levels in primary cultured aortic and mesenteric arterial smooth cells, an increase that was correlated with an increase in intracellular [Mg2+] (134). The effects of ANG II on TRPM7 expression, Mg2+ influx, and accompanying cellular proliferation were inhibited by siRNA-mediated TRPM7 downregulation (134). The authors of this work concluded that a TRPM7-dependent influx of Mg2+ is directly linked to the vascular proliferative response to activation of ANG II receptors. Interestingly, basal levels of TRPM7 and intracellular [Mg2+] were found to be reduced in cultured mesenteric vascular smooth muscle cells from SHRs compared with those from WKY rats (358). Furthermore, ANG II was shown to increase TRPM7 expression and intracellular [Mg2+] in WKY, but not SHR, cells. These observations suggest that changes in TRPM7 expression and Mg2+ homeostasis are better correlated with enhanced contractility than with vascular smooth muscle cell proliferation in the SHR model of hypertension. However, direct causal links between TRPM7 expression, intracellular [Mg2+], and differential effects on vascular contractility or smooth muscle proliferation are presently difficult to define given the celltype-dependent, model-related, and tissue-specific differences in these parameters that have been reported (189, 259).

An exciting recent report demonstrated a novel mechanism for the regulation of gene expression, differentiation, and development involving TRPM7 channels (172). In this study, Krapivinsky et al. (172) showed that the COOH-terminal serine/threonine kinase domain of TRPM7 is proteolytically cleaved from the cation-conducting domain and translates to the nucleus in many different types of cells. The kinase domain participates in chromatin remodeling by phosphorylating histone proteins at specific sites important for cell differentiation and development. This study also showed that TRPM7 channels control cytosolic [Zn2+], which in turn regulates binding of the cleaved kinase fragments to transcription factors containing zinc-finger domains. These data suggest that the kinase and ion-conducting domains of TRPM7 can together control epigenetic modifications and gene expression. However, it is not currently known if this system is functional in smooth muscle cells or participates in cellular proliferation.

C. Oxidative Stress and Ischemia/Reperfusion Injury

ROS generated by biological systems include H2O2, O2, and hydroxyl radicals (OH·). ROS act as signaling molecules, but when produced in excess can cause oxidative damage to membrane phospholipids, proteins, and DNA and diminish the bioavailability of endothelium-derived NO in the vasculature. Oxidative stress, defined as an imbalance between the generation of ROS by a biological system and the ability of that system to remove toxic metabolites and repair damage, is associated with numerous pathological conditions, including common cardiovascular diseases such as hypertension (65), atherosclerosis, and stroke. Reperfusion of tissue following ischemia (IR injury) is well recognized as a source of oxidative stress in the brain following stoke and in the heart following myocardial infarction. Multiple TRP channels have been implicated in responses to ROS and oxidative stress in various organ systems.

1. TRPC4

A study by Jung et al. (153) identified a single-nucleotide polymorphism (SNP) in the coding region of TRPC4 (I957V) that is associated with a diminished risk of myocardial infarction (MI) (153). Ca2+ imaging and patch-clamp analysis indicated that the mutant channel is more sensitive to muscarinic receptor stimulation compared with the wild type. The authors speculate that elevated Ca2+ influx through the I957V variant lowered the incidence of MI by enhancing endothelium-dependent relaxation in coronary arteries.

2. TRPC6

Weissmann et al. (386) demonstrated that pulmonary edema caused by I/R injury is diminished in lungs from TRPC6−/− mice compared with wild-type controls and TRPC1−/− and TRPC4−/− mice. This study further demonstrated that TRPC6 activation and I/R injury occur downstream of NOX2-generated ROS and PLC-γ-generated DAG.

3. TRPV1

A study by Wang et al. (375) demonstrated that recovery of cardiac function following I/R injury was diminished in TRPV1−/− mice or mice treated with the TRPV1 inhibitor capsazepine compared with controls. Impairment of functional recovery was worse in animals treated with capsazepine compared with controls. These data suggest that TRPV1 channel activity attenuates I/R-induced damage in the heart. An apparent protective effect of TRPV1 channel activation on I/R injury has also been demonstrated in the kidney (269, 362, 364, 365).

Several studies have examined the effects of TRPV1 activity on I/R injury in the brain. Capsaicin administered 5 min after restoration of flow was shown to diminish I/R injury-induced neuronal damage in a global cerebral ischemia model, suggesting that activation of TRPV1 in this context is neuroprotective (265). Xu et al. (404) extended these findings, showing that chronic dietary capsaicin diminished hypertrophy of the cerebral vasculature and delayed onset of stroke in stoke-prone SHRs, likely due to enhanced production of NO. In addition, a study by Muzzi et al. (239) showed that systemic administration of the TRPV1 agonist rinvanil following I/R injury caused hypothermia in mice in association with diminished neuronal damage. In contrast, Gauden et al. (108) reported that blocking TRPV1 with capsazepine reduced I/R injury-induced increases in endothelial permeability in the brain, suggesting that TRPV1 activity can have negative consequences following cerebral I/R injury (108).

4. TRPV3

A study by Moussaieff and colleagues demonstrated that incensole, a TRPV3-activating compound derived from trees of Boswellia species (235), diminished brain infarct volume and neurological deficits induced by I/R injury following carotid artery occlusion (236). The neuroprotective effects of incensole were diminished in TRPV3−/− mice and in mice pretreated with the nonselective TRPV inhibitor ruthenium red, suggesting that activation of TRPV3 channels mediates a portion of this response. The mechanism of action is unknown, but TRPV3 channels are present in the cerebral artery endothelium, and activation of the channel with carvacrol causes endothelium-dependent vasodilation (78), consistent with the possibility that incensole protects against ischemic damage by promoting cerebral blood flow.

5. TRPV4

Zhang et al. (418) showed that the selective TRPV4 activator GSK1016790A and the muscarinic receptor agonist ACh caused endothelium-dependent dilation of mouse cerebral arteries, an effect that was blocked by the selective TRPV4 antagonist HC-067047 and by inhibition of SK and IK channels. Interestingly, block of TRPV4 had little effect on ACh-induced dilation of arteries isolated from mice expressing mutant forms of the human amyloid precursor protein (APP), constitutively active tumor growth factor (TGF)-β1, or both transgenes. This effect was reversed by superoxide dismutase or catalase in arteries from APP mice, but not those from other groups. The authors concluded that endothelial TRPV4 function is impaired by the oxidative stress that is specifically associated with APP overexpression.

6. TRPA1

TRPA1 is present in the endothelium of cerebral arteries (77) and in the perivascular nerves of mesenteric arteries (27). A number of studies have demonstrated that TRPA1 channels are able to detect cellular oxygen status and oxidative stress. Work by Takahashi et al. (346) demonstrated that TRPA1 channels are activated by both hyperoxic and hypoxic conditions, suggesting that the channel is intimately involved in oxygen sensing. TRPA1 channels are also activated by substances such as 4-HNE, 4-ONE, and 4-HHE that are generated by ROS-induced peroxidation of ω6 polyunsaturated fatty acids (17, 363). A study by Sullivan et al. (338) demonstrated that endogenously generated oxidized lipids cause endothelium-dependent dilation of cerebral arteries, suggesting that TRPA1 channels may have a protective role in the cerebral vasculature during oxidative stress.

7. TRPM2

The unitary conductance of TRPM2 is ~52–60 pS at negative holding potentials and ~76 pS at positive potentials. Ionic selectivity is temperature dependent, with PCa:PNa ratios of ~0.7:1 at 20°C and ~6:1 at 35°C (267, 299, 353). TRPM2, which is broadly expressed and is present in vascular smooth muscle and endothelial cells (135), is activated by ADP-ribose (ADPR) (267, 299) and H2O2 (129, 384). Several studies have indicated that TRPM2 is critically involved in ROS-mediated signaling and responses to oxidative stress in a number of organ systems, including the vasculature (345). In cultured endothelial cells, high concentrations of H2O2 activated a cation current that was diminished by siRNA-mediated knockdown of TRPM2 channels (135). H2O2-induced Ca2+ influx and apoptosis of cultured endothelial cells is caused by PKC-α-dependent activation of TRPM2 (136, 340). The in vivo significance of these findings has not been demonstrated.

8. TRPM4

Trauma-induced expression of TRPM4 channels in the endothelium has been linked to neuronal damage. Gerzanich et al. (112) demonstrated that TRPM4 expression is increased in tissue surrounding the site of spinal cord injury, particularly in microvessels and capillaries. Downregulation of TRPM4 expression reduced secondary hemorrhage at the injury site and improved neurobehavioral performance. A study by Loh et al. (203) demonstrated that I/R injury induced by middle cerebral artery occlusion (MCAO) promoted the expression of TRPM4 in the cerebral artery endothelium. Knockdown of TRPM4 expression enhanced angiogenesis and reduced infarct volume after MCAO, effects that were accompanied by improved motor function. These studies identify TRPM4 blockade as a possible prophylactic strategy against injury- and stroke-induced neuronal damage.

9. TRPM7

A study by Aarts et al. (1) demonstrated that exposure of cultured neurons to oxygen-glucose depravation (OGD), an in vitro model of I/R injury, stimulated a cation current that was diminished by siRNA-mediated knockdown of TRPM7 expression. TRPM7 knockdown also decreased OGD-induced Ca2+ influx and reduced cell death. A follow-up study by Sun et al. (339) used targeted injection of adeno-associated virus to selectively diminish TRPM7 expression in CA1 hippocampal neurons in vivo. The resulting decrease in TRPM7 expression was associated with diminished delayed neuronal death following global cerebral ischemia and improved neurological outcomes. These data indicate an important role for TRPM7 in neuronal damage following I/R injury and suggest that block of the channel after stroke may be neuroprotective.

D. Pulmonary Edema

Pulmonary edema can result from congestive heart failure, ventilator-induced mechanical injury, prolonged exposure to high altitude, and other causes. Several studies have demonstrated that TRPV4 channels are involved in the development of pulmonary edema. A study by Hamanaka et al. (127) showed that ventilation-induced lung injury was diminished in lungs from TRPV4−/− mice compared with controls, suggesting that TRPV4 channel activity contributes to mechanically induced vascular injury in the lung. The authors proposed that stretch-induced activation of TRPV4 contributes to this process. This report also demonstrated that infusion of macrophages isolated from wild-type mice into lungs isolated from TRPV4−/− mice restored the permeability response to high ventilation pressure, suggesting that mechanical activation of TRPV4 channels present in macrophages is responsible for high pressure injury (126). Jian et al. (149) showed that elevated pulmonary endothelial permeability resulting from high vascular pressures was diminished in lungs from TRPV4−/− mice compared with controls. This response was blocked by inhibition of CYP activity, suggesting that biosynthesis of EETs during exposure to high pressures induces activation of TRPV4 channels in the pulmonary microvascular endothelium. Ventilator-induced lung edema was blocked by inhalation of nanoparticles that release the nonselective TRPV blocker ruthenium red (155), further implicating TRPV4 channels in this response. An intriguing translational study by Thorneloe et al. (350) demonstrated that novel orally active TRPV4 blockers can prevent and resolve pulmonary edema associated with heart failure in rats, suggesting that block of TRPV4 channels may have important therapeutic effects. TRPV4 blockade has also been shown to prevent lung injury in mice exposed to acid and chlorine gas (23).

E. Neurogenic Inflammation

Neurogenic inflammation is characterized by vasodilation and edema that results in redness, warmth, hypersensitivity, and swelling. This condition is initiated by substances released from small-diameter sensory neurons, such as substance P, CGRP, glutamate, and prostaglandins, that act on endothelial cells, mast cells, immune cells, and vascular smooth muscle cells (284). Neurogenic inflammation is associated with many diseases, including migraine, arthritis, asthma, inflammatory bowel disease, and chronic obstructive pulmonary disease. Several studies have implicated TRPV1 and TRPA1 channels in neurogenic inflammation (for review, see Refs. 20, 90).

1. TRPV1

Many studies have demonstrated that TRPV1 channels are involved in neuropathic pain and neurogenic inflammation (for review, see Ref. 111). Among these studies is a report by Ji et al. (148) demonstrating that TRPV1 protein levels were elevated by inflammatory stimuli in dorsal root ganglion (DRG) neurons, resulting in hypersensitivity to heat. Genetic models have been used to show that mustard oil (25) and capsaicin (356) induce significantly less inflammation in TRPV1−/− mice than in wild-type mice. The signaling pathways responsible for TRPV1 channel activation by inflammatory mediators have been extensively characterized and include sensitization mediated by PLC-, PKA-, and PKC-dependent signaling cascades (270). There is considerable interest in TRPV1 blockade and desensitization as a potential treatment for chronic pain and neurogenic inflammation (233, 343).

2. TRPA1

TRPA1 channels are often coexpressed with TRPV1 in primary sensory neurons and may play an important role in the vascular component of neurogenic inflammation. This possibility was first suggested by Trevisani et al. (361), who showed that the TRPA1 agonist 4-HNE caused edema when injected into the hind paws of rats. Additional evidence for an inflammatory role for TRPA1 was provided by a study demonstrating that, in guinea pigs, tracheal plasma extravasation resulting from cigarette smoke inhalation was attenuated by inhibition of TRPA1 with HC-030031 (18). Furthermore, increases in capillary permeability were absent in TRPA1−/− mice administered an aqueous extract of cigarette smoke (18). N-acetyl-p-benzo-quinoneimine (NAPQI), a metabolite produced following administration of large doses of acetaminophen, was shown to activate TRPA1 channels and cause an extravasation of tracheal plasma in rats and mice that was sensitive to HC-030031 and TRPA1 knockout (244). NAPQI application also induced an increase in plasma protein extravasation in the skin conjunctiva of rats and mice that was attenuated by TRPA1 blockade or gene knockout (244). These findings suggest that inhibitors of TRPA1 may be useful for treating increased capillary permeability and edema associated with certain inflammatory conditions.

V. SUMMARY AND CONCLUSIONS

TRP channels are critically involved in normal and pathophysiological responses in the vasculature. Strong evidence demonstrates that TRP channels contribute to mechanical- and agonist-induced vascular smooth muscle contractility and proliferation. TRP channels present in endothelial cells contribute to endothelium-dependent vasodilation, regulation of vascular permeability, and angiogenesis. TRP channels in other cell types, such as perivascular neurons and parenchymal astrocytes, are also important in the regulation of vascular function, primarily through the control of vascular tone. The primary regulatory paradigm for most of these functions is the control of global [Ca2+]i or the propagation of subcellular Ca2+ signaling events that modulate cellular activity. These effects can result directly from TRP channel-mediated Ca2+ influx; can be secondary to cation influx-induced membrane depolarization or activation of GPCRs; or can be caused by CICR from internal sources. Compelling evidence supports substantial roles for TRP channels in vascular disease states such as hypertension, neointimal injury, oxidative stress response, neurogenic inflammation, and pulmonary edema. These disease mechanisms involve increased expression or activation of TRP channels leading to elevated contractility, vascular hypertrophy, hyperplasia, or other forms of remodeling within the vascular wall. In light of these findings, selective pharmacological blockers of TRP channels may present exciting opportunities for the development of new therapies to prevent and treat vascular-related diseases.

Considerable work is needed to more fully understand the structural, functional, and mechanistic aspects of vascular TRP channel biology. In particular, important features of the organization and interactions between closely related, colocalized TRP channels and other signaling modalities in vascular cells remain largely unknown. Several recent studies have highlighted the importance of such signaling networks in native smooth muscle and endothelial cells, and further advances in this area will greatly enrich our understanding of vascular function. Proteomic analyses and super-resolution imaging approaches may be useful in providing a more complete understanding of the interactions between TRP channels and cellular elements, such as the cytoskeleton and membrane phospholipids. Addressing controversies surrounding the relative fractional Ca2+ composition of TRP currents in vivo will require additional efforts to record TRP currents from native vascular cells under physiological ionic gradients. Such experiments may help to resolve controversies surrounding the role of TRP channels in SOCE. Successful completion of these studies will likely require the development of a more extensive array of selective pharmacological tools for modifying TRP channel activity, such as those that are currently available for probing TRPV4 channel function. In addition, little is currently known about how TRP channel gene expression is regulated under normal conditions, during development, or during disease. Finally, advanced genetic approaches, including the development of cell-specific, conditional knockouts targeted to the vasculature, genetically encoded indicators, and the application of optogenetic techniques to modulate channel activity in specifically targeted locations, are needed to address deeper questions regarding the physiological significance of TRP channels. These new approaches will allow a more comprehensive appreciation of the molecular and cellular complexity of TRP channels in the vasculature, providing substantial insight into fundamental mechanisms of vascular physiology and pathophysiology.

GRANTS

This work was supported by National Heart, Lung, and Blood Institute Grants R01HL091905; (to S. Earley) and P01HL095488 (to J. E. Brayden).

DISCLOSURES

No conflicts of interest, financial or otherwise, are declared by the authors.

acknowledgments

We thank Drs. Thomas A. Longden and Fabrice Dabertrand, Michelle N. Sullivan, and Yao Li for critical comments on the manuscript and Dr. David Hill-Eubanks for editorial assistance.

Address for reprint requests and other correspondence: J. E. Brayden, Dept. of Pharmacology, Univ. of Vermont College of Medicine, 89 Beaumont Ave., Burlington, VT 05405 (e-mail: ude.mvu@nedyarB.eoJ).

REFERENCES

1. Aarts M, Iihara K, Wei WL, Xiong ZG, Arundine M, Cerwinski W, MacDonald JF, Tymianski M. A key role for TRPM7 channels in anoxic neuronal death. Cell 115: 863–877, 2003. [Abstract] [Google Scholar]
2. Abdullaev IF, Bisaillon JM, Potier M, Gonzalez JC, Motiani RK, Trebak M. Stim1 and Orai1 mediate CRAC currents and store-operated calcium entry important for endothelial cell proliferation. Circ Res 103: 1289–1299, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
3. Adebiyi A, Narayanan D, Jaggar JH. Caveolin-1 assembles type 1 inositol 1,4,5-trisphosphate receptors and canonical transient receptor potential 3 channels into a functional signaling complex in arterial smooth muscle cells. J Biol Chem 286: 4341–4348, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
4. Adebiyi A, Thomas-Gatewood CM, Leo MD, Kidd MW, Neeb ZP, Jaggar JH. An elevation in physical coupling of type 1 inositol 1,4,5-trisphosphate (IP3) receptors to transient receptor potential 3 (TRPC3) channels constricts mesenteric arteries in genetic hypertension. Hypertension 60: 1213–1219, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
5. Adebiyi A, Zhao G, Narayanan D, Thomas-Gatewood CM, Bannister JP, Jaggar JH. Isoform-selective physical coupling of TRPC3 channels to IP3 receptors in smooth muscle cells regulates arterial contractility. Circ Res 106: 1603–1612, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
6. Ahmmed GU, Mehta D, Vogel S, Holinstat M, Paria BC, Tiruppathi C, Malik AB. Protein kinase Calpha phosphorylates the TRPC1 channel and regulates store-operated Ca2+ entry in endothelial cells. J Biol Chem 279: 20941–20949, 2004. [Abstract] [Google Scholar]
7. Akerman S, Kaube H, Goadsby PJ. Anandamide acts as a vasodilator of dural blood vessels in vivo by activating TRPV1 receptors. Br J Pharmacol 142: 1354–1360, 2004. [Europe PMC free article] [Abstract] [Google Scholar]
8. Albert AP, Large WA. Inhibitory regulation of constitutive transient receptor potential-like cation channels in rabbit ear artery myocytes. J Physiol 560: 169–180, 2004. [Abstract] [Google Scholar]
9. Albert AP, Large WA. Synergism between inositol phosphates and diacylglycerol on native TRPC6-like channels in rabbit portal vein myocytes. J Physiol 552: 789–795, 2003. [Abstract] [Google Scholar]
10. Albert AP, Piper AS, Large WA. Properties of a constitutively active Ca2+-permeable non-selective cation channel in rabbit ear artery myocytes. J Physiol 549: 143–156, 2003. [Abstract] [Google Scholar]
11. Albert AP, Piper AS, Large WA. Role of phospholipase D and diacylglycerol in activating constitutive TRPC-like cation channels in rabbit ear artery myocytes. J Physiol 566: 769–780, 2005. [Abstract] [Google Scholar]
12. Albert AP, Pucovsky V, Prestwich SA, Large WA. TRPC3 properties of a native constitutively active Ca2+-permeable cation channel in rabbit ear artery myocytes. J Physiol 571: 361–369, 2006. [Abstract] [Google Scholar]
13. Albert AP, Saleh SN, Large WA. Inhibition of native TRPC6 channel activity by phosphatidylinositol 4,5-bisphosphate in mesenteric artery myocytes. J Physiol 586: 3087–3095, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
14. Albert AP, Saleh SN, Peppiatt-Wildman CM, Large WA. Multiple activation mechanisms of store-operated TRPC channels in smooth muscle cells. J Physiol 583: 25–36, 2007. [Abstract] [Google Scholar]
15. Alvarez DF, King JA, Weber D, Addison E, Liedtke W, Townsley MI. Transient receptor potential vanilloid 4-mediated disruption of the alveolar septal barrier: a novel mechanism of acute lung injury. Circ Res 99: 988–995, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
16. Ambudkar IS, Ong HL, Liu X, Bandyopadhyay BC, Cheng KT. TRPC1: the link between functionally distinct store-operated calcium channels. Cell Calcium 42: 213–223, 2007. [Abstract] [Google Scholar]
17. Andersson DA, Gentry C, Moss S, Bevan S. Transient receptor potential A1 is a sensory receptor for multiple products of oxidative stress. J Neurosci 28: 2485–2494, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
18. Andre E, Campi B, Materazzi S, Trevisani M, Amadesi S, Massi D, Creminon C, Vaksman N, Nassini R, Civelli M, Baraldi PG, Poole DP, Bunnett NW, Geppetti P, Patacchini R. Cigarette smoke-induced neurogenic inflammation is mediated by alpha,beta-unsaturated aldehydes and the TRPA1 receptor in rodents. J Clin Invest 118: 2574–2582, 2008. [Abstract] [Google Scholar]
19. Antigny F, Girardin N, Frieden M. Transient receptor potential canonical channels are required for in vitro endothelial tube formation. J Biol Chem 287: 5917–5927, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
20. Aubdool AA, Brain SD. Neurovascular aspects of skin neurogenic inflammation. J Invest Dermatol 15: 33–39, 2011. [Abstract] [Google Scholar]
21. Baba Y, Kurosaki T. Physiological function and molecular basis of STIM1-mediated calcium entry in immune cells. Immunol Rev 231: 174–188, 2009. [Abstract] [Google Scholar]
22. Bagher P, Beleznai T, Kansui Y, Mitchell R, Garland CJ, Dora KA. Low intravascular pressure activates endothelial cell TRPV4 channels, local Ca2+ events, and IKCa channels, reducing arteriolar tone. Proc Natl Acad Sci USA 109: 18174–18179, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
23. Balakrishna S, Song W, Achanta S, Doran SF, Liu B, Kaelberer MM, Yu Z, Sui A, Cheung M, Leishman E, Eidam HS, Ye G, Willette RN, Thorneloe KS, Bradshaw HB, Matalon S, Jordt SE. TRPV4 inhibition counteracts edema and inflammation and improves pulmonary function and oxygen saturation in chemically induced acute lung injury. Am J Physiol Lung Cell Mol Physiol 307: L158–L172, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
24. Bandell M, Story GM, Hwang SW, Viswanath V, Eid SR, Petrus MJ, Earley TJ, Patapoutian A. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron 41: 849–857, 2004. [Abstract] [Google Scholar]
25. Banvolgyi A, Pozsgai G, Brain SD, Helyes ZS, Szolcsanyi J, Ghosh M, Melegh B, Pinter E. Mustard oil induces a transient receptor potential vanilloid 1 receptor-independent neurogenic inflammation and a non-neurogenic cellular inflammatory component in mice. Neuroscience 125: 449–459, 2004. [Abstract] [Google Scholar]
26. Barbet G, Demion M, Moura IC, Serafini N, Leger T, Vrtovsnik F, Monteiro RC, Guinamard R, Kinet JP, Launay P. The calcium-activated nonselective cation channel TRPM4 is essential for the migration but not the maturation of dendritic cells. Nature Immunol 9: 1148–1156, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
27. Bautista DM, Movahed P, Hinman A, Axelsson HE, Sterner O, Hogestatt ED, Julius D, Jordt SE, Zygmunt PM. Pungent products from garlic activate the sensory ion channel TRPA1. Proc Natl Acad Sci USA 102: 12248–12252, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
28. Bayliss WM. On the local reactions of the arterial wall to changes of internal pressure. J Physiol 28: 220–231, 1902. [Abstract] [Google Scholar]
29. Benfenati V, Amiry-Moghaddam M, Caprini M, Mylonakou MN, Rapisarda C, Ottersen OP, Ferroni S. Expression and functional characterization of transient receptor potential vanilloid-related channel 4 (TRPV4) in rat cortical astrocytes. Neuroscience 148: 876–892, 2007. [Abstract] [Google Scholar]
30. Bergdahl A, Gomez MF, Dreja K, Xu SZ, Adner M, Beech DJ, Broman J, Hellstrand P, Sward K. Cholesterol depletion impairs vascular reactivity to endothelin-1 by reducing store-operated Ca2+ entry dependent on TRPC1. Circ Res 93: 839–847, 2003. [Abstract] [Google Scholar]
31. Bergdahl A, Gomez MF, Wihlborg AK, Erlinge D, Eyjolfson A, Xu SZ, Beech DJ, Dreja K, Hellstrand P. Plasticity of TRPC expression in arterial smooth muscle: correlation with store-operated Ca2+ entry. Am J Physiol Cell Physiol 288: C872–C880, 2005. [Abstract] [Google Scholar]
32. Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nature Rev Mol Cell Biol 4: 517–529, 2003. [Abstract] [Google Scholar]
33. Bevan JA, Brayden JE. Nonadrenergic neural vasodilator mechanisms. Circ Res 60: 309–326, 1987. [Abstract] [Google Scholar]
34. Bolotina VM. Orai, STIM1 and iPLA2beta: a view from a different perspective. J Physiol 586: 3035–3042, 2008. [Abstract] [Google Scholar]
35. Boulay G, Brown DM, Qin N, Jiang M, Dietrich A, Zhu MX, Chen Z, Birnbaumer M, Mikoshiba K, Birnbaumer L. Modulation of Ca2+ entry by polypeptides of the inositol 1,4,5-trisphosphate receptor (IP3R) that bind transient receptor potential (TRP): evidence for roles of TRP and IP3R in store depletion-activated Ca2+ entry. Proc Natl Acad Sci USA 96: 14955–14960, 1999. [Europe PMC free article] [Abstract] [Google Scholar]
36. Bratz IN, Dick GM, Tune JD, Edwards JM, Neeb ZP, Dincer UD, Sturek M. Impaired capsaicin-induced relaxation of coronary arteries in a porcine model of the metabolic syndrome. Am J Physiol Heart Circ Physiol 294: H2489–H2496, 2008. [Abstract] [Google Scholar]
37. Brayden JE, Li Y, Tavares MJ. Purinergic receptors regulate myogenic tone in cerebral parenchymal arterioles. J Cereb Blood Flow Metab 33: 293–299, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
38. Breyne J, Vanheel B. Methanandamide hyperpolarizes gastric arteries by stimulation of TRPV1 receptors on perivascular CGRP containing nerves. J Cardiovasc Pharmacol 47: 303–309, 2006. [Abstract] [Google Scholar]
39. Brough GH, Wu S, Cioffi D, Moore TM, Li M, Dean N, Stevens T. Contribution of endogenously expressed Trp1 to a Ca2+-selective, store-operated Ca2+ entry pathway. FASEB J 15: 1727–1738, 2001. [Abstract] [Google Scholar]
40. Brueggemann LI, Markun DR, Henderson KK, Cribbs LL, Byron KL. Pharmacological and electrophysiological characterization of store-operated currents and capacitative Ca2+ entry in vascular smooth muscle cells. J Pharmacol Exp Ther 317: 488–499, 2006. [Abstract] [Google Scholar]
41. 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 302: H634–H642, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
42. Campbell WB, Gebremedhin D, Pratt PF, Harder DR. Identification of epoxyeicosatrienoic acids as endothelium-derived hyperpolarizing factors. Circ Res 78: 415–423, 1996. [Abstract] [Google Scholar]
43. Cao E, Liao M, Cheng Y, Julius D. TRPV1 structures in distinct conformations reveal activation mechanisms. Nature 504: 113–118, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
44. Carretero OA, Oparil S. Essential hypertension. Part I: definition and etiology. Circulation 101: 329–335, 2000. [Abstract] [Google Scholar]
45. Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature 389: 816–824, 1997. [Abstract] [Google Scholar]
46. Cavanaugh DJ, Chesler AT, Jackson AC, Sigal YM, Yamanaka H, Grant R, O'Donnell D, Nicoll RA, Shah NM, Julius D, Basbaum AI. Trpv1 reporter mice reveal highly restricted brain distribution and functional expression in arteriolar smooth muscle cells. J Neurosci 31: 5067–5077, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
47. Chang AS, Chang SM, Garcia RL, Schilling WP. Concomitant and hormonally regulated expression of trp genes in bovine aortic endothelial cells. FEBS Lett 415: 335–340, 1997. [Abstract] [Google Scholar]
48. Chen G, Suzuki H, Weston AH. Acetylcholine releases endothelium-derived hyperpolarizing factor and EDRF from rat blood vessels. Br J Pharmacol 95: 1165–1174, 1988. [Europe PMC free article] [Abstract] [Google Scholar]
49. Chen J, Crossland RF, Noorani MM, Marrelli SP. Inhibition of TRPC1/TRPC3 by PKG contributes to NO-mediated vasorelaxation. Am J Physiol Heart Circ Physiol 297: H417–H424, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
50. Chen L, Kassmann M, Sendeski M, Tsvetkov D, Marko L, Michalick L, Riehle M, Liedtke WB, Kuebler WM, Harteneck C, Tepel M, Patzak A, Gollasch M. Functional transient receptor potential vanilloid 1 and transient receptor potential vanilloid 4 channels along different segments of the renal vasculature. Acta Physiol. In press. [Abstract] [Google Scholar]
51. Chen X, Yang D, Ma S, He H, Luo Z, Feng X, Cao T, Ma L, Yan Z, Liu D, Tepel M, Zhu Z. Increased rhythmicity in hypertensive arterial smooth muscle is linked to transient receptor potential canonical channels. J Cell Mol Med 14: 2483–2494, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
52. Chen YS, Lu MJ, Huang HS, Ma MC. Mechanosensitive transient receptor potential vanilloid type 1 channels contribute to vascular remodeling of rat fistula veins. J Vasc Surg 52: 1310–1320, 2010. [Abstract] [Google Scholar]
53. Cheng H, Beck A, Launay P, Gross SA, Stokes AJ, Kinet JP, Fleig A, Penner R. TRPM4 controls insulin secretion in pancreatic beta-cells. Cell Calcium 41: 51–61, 2007. [Abstract] [Google Scholar]
54. Cheng HW, James AF, Foster RR, Hancox JC, Bates DO. VEGF activates receptor-operated cation channels in human microvascular endothelial cells. Arteriosclerosis Thrombosis Vasc Biol 26: 1768–1776, 2006. [Abstract] [Google Scholar]
55. Cheng KT, Liu X, Ong HL, Swaim W, Ambudkar IS. Local Ca2+ entry via Orai1 regulates plasma membrane recruitment of TRPC1 and controls cytosolic Ca2+ signals required for specific cell functions. PLoS Biol 9: e1001025, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
56. Chetham PM, Guldemeester HA, Mons N, Brough GH, Bridges JP, Thompson WJ, Stevens T. Ca2+-inhibitable adenylyl cyclase and pulmonary microvascular permeability. Am J Physiol Lung Cell Mol Physiol 273: L22–L30, 1997. [Abstract] [Google Scholar]
57. Ching LC, Kou YR, Shyue SK, Su KH, Wei J, Cheng LC, Yu YB, Pan CC, Lee TS. Molecular mechanisms of activation of endothelial nitric oxide synthase mediated by transient receptor potential vanilloid type 1. Cardiovasc Res 91: 492–501, 2011. [Abstract] [Google Scholar]
58. Chung MK, Lee H, Mizuno A, Suzuki M, Caterina MJ. 2-Aminoethoxydiphenyl borate activates and sensitizes the heat-gated ion channel TRPV3. J Neurosci 24: 5177–5182, 2004. [Abstract] [Google Scholar]
59. Cioffi DL, Wu S, Alexeyev M, Goodman SR, Zhu MX, Stevens T. Activation of the endothelial store-operated ISOC Ca2+ channel requires interaction of protein 4.1 with TRPC4. Circ Res 97: 1164–1172, 2005. [Abstract] [Google Scholar]
60. Cioffi DL, Wu S, Chen H, Alexeyev M, St Croix CM, Pitt BR, Uhlig S, Stevens T. Orai1 determines calcium selectivity of an endogenous TRPC heterotetramer channel. Circ Res 110: 1435–1444, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
61. Clapham DE, Montell C, Schultz G, Julius D, International Union of Pharmacology. International Union of Pharmacology XLIII Compendium of voltage-gated ion channels: transient receptor potential channels. Pharmacol Rev 55: 591–596, 2003. [Abstract] [Google Scholar]
62. Clapham DE, Runnels LW, Strubing C. The TRP ion channel family. Nature Rev Neurosci 2: 387–396, 2001. [Abstract] [Google Scholar]
63. Cosens DJ, Manning A. Abnormal electroretinogram from a Drosophila mutant. Nature 224: 285–287, 1969. [Abstract] [Google Scholar]
64. Crnich R, Amberg GC, Leo MD, Gonzales AL, Tamkun MM, Jaggar JH, Earley S. Vasoconstriction resulting from dynamic membrane trafficking of TRPM4 in vascular smooth muscle cells. Am J Physiol Cell Physiol 299: C682–C694, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
65. Datla SR, Griendling KK. Reactive oxygen species, NADPH oxidases, and hypertension. Hypertension 56: 325–330, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
66. Davies PF. Flow-mediated endothelial mechanotransduction. Physiol Rev 75: 519–560, 1995. [Europe PMC free article] [Abstract] [Google Scholar]
67. De Baaij JH, Hoenderop JG, Bindels RJ. Magnesium in man: implications for health and disease. Physiol Rev 95: 1–46, 2015. [Abstract] [Google Scholar]
68. DeHaven WI, Jones BF, Petranka JG, Smyth JT, Tomita T, Bird GS, Putney JW Jr. TRPC channels function independently of STIM1 and Orai1. J Physiol 587: 2275–2298, 2009. [Abstract] [Google Scholar]
69. Dietrich A, Kalwa H, Storch U, Mederos y Schnitzler M, Salanova B, Pinkenburg O, Dubrovska G, Essin K, Gollasch M, Birnbaumer L, Gudermann T. Pressure-induced and store-operated cation influx in vascular smooth muscle cells is independent of TRPC1. Pflügers Arch 455: 465–477, 2007. [Abstract] [Google Scholar]
70. Dietrich A, Mederos YSM, Gollasch M, Gross V, Storch U, Dubrovska G, Obst M, Yildirim E, Salanova B, Kalwa H, Essin K, Pinkenburg O, Luft FC, Gudermann T, Birnbaumer L. Increased vascular smooth muscle contractility in TRPC6−/− mice. Mol Cell Biol 25: 6980–6989, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
71. Dong XP, Wang X, Xu H. TRP channels of intracellular membranes. J Neurochem 113: 313–328, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
72. Du H, Wang X, Wu J, Qian Q. Phenylephrine induces elevated RhoA activation and smooth muscle alpha-actin expression in Pkd2+/− vascular smooth muscle cells. Hypertension Res 33: 37–42, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
73. Dunn KM, Hill-Eubanks DC, Liedtke WB, Nelson MT. TRPV4 channels stimulate Ca2+-induced Ca2+ release in astrocytic endfeet and amplify neurovascular coupling responses. Proc Natl Acad Sci USA 110: 6157–6162, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
74. Dunn KM, Nelson MT. Neurovascular signaling in the brain and the pathological consequences of hypertension. Am J Physiol Heart Circ Physiol 306: H1–H14, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
75. Earley S. TRPA1 channels in the vasculature. Br J Pharmacol 167: 13–22, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
76. Earley S. TRPM4 channels in smooth muscle function. Pflügers Arch 465: 1223–1231, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
77. Earley S, Gonzales AL, Crnich R. Endothelium-dependent cerebral artery dilation mediated by TRPA1 and Ca2+-Activated K+ channels. Circ Res 104: 987–994, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
78. Earley S, Gonzales AL, Garcia ZI. A dietary agonist of transient receptor potential cation channel V3 elicits endothelium-dependent vasodilation. Mol Pharmacol 77: 612–620, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
79. Earley S, Heppner TJ, Nelson MT, Brayden JE. TRPV4 forms a novel Ca2+ signaling complex with ryanodine receptors and BKCa channels. Circ Res 97: 1270–1279, 2005. [Abstract] [Google Scholar]
80. Earley S, Pastuszyn A, Walker BR. Cytochrome p-450 epoxygenase products contribute to attenuated vasoconstriction after chronic hypoxia. Am J Physiol Heart Circ Physiol 285: H127–H136, 2003. [Abstract] [Google Scholar]
81. 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 297: H1096–H1102, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
82. Earley S, Straub SV, Brayden JE. Protein kinase C regulates vascular myogenic tone through activation of TRPM4. Am J Physiol Heart Circ Physiol 292: H2613–H2622, 2007. [Abstract] [Google Scholar]
83. Earley S, Waldron BJ, Brayden JE. Critical role for transient receptor potential channel TRPM4 in myogenic constriction of cerebral arteries. Circ Res 95: 922–929, 2004. [Abstract] [Google Scholar]
84. Eder P, Poteser M, Romanin C, Groschner K. Na+ entry and modulation of Na+/Ca2+ exchange as a key mechanism of TRPC signaling. Pflügers Arch 451: 99–104, 2005. [Abstract] [Google Scholar]
85. Eder P, Probst D, Rosker C, Poteser M, Wolinski H, Kohlwein SD, Romanin C, Groschner K. Phospholipase C-dependent control of cardiac calcium homeostasis involves a TRPC3-NCX1 signaling complex. Cardiovasc Res 73: 111–119, 2007. [Abstract] [Google Scholar]
86. Edwards G, Dora KA, Gardener MJ, Garland CJ, Weston AH. K+ is an endothelium-derived hyperpolarizing factor in rat arteries. Nature 396: 269–272, 1998. [Abstract] [Google Scholar]
87. Edwards G, Feletou M, Weston AH. Endothelium-derived hyperpolarising factors and associated pathways: a synopsis. Pflügers Arch 459: 863–879, 2010. [Abstract] [Google Scholar]
88. Ellinsworth DC, Earley S, Murphy TV, Sandow SL. Endothelial control of vasodilation: integration of myoendothelial microdomain signalling and modulation by epoxyeicosatrienoic acids. Pflügers Arch 466: 389–405, 2014. [Abstract] [Google Scholar]
89. Everaerts W, Zhen X, Ghosh D, Vriens J, Gevaert T, Gilbert JP, Hayward NJ, McNamara CR, Xue F, Moran MM, Strassmaier T, Uykal E, Owsianik G, Vennekens R, De Ridder D, Nilius B, Fanger CM, Voets T. Inhibition of the cation channel TRPV4 improves bladder function in mice and rats with cyclophosphamide-induced cystitis. Proc Natl Acad Sci USA 107: 19084–19089, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
90. Fernandes ES, Fernandes MA, Keeble JE. The functions of TRPA1 and TRPV1: moving away from sensory nerves. Br J Pharmacol 166: 510–521, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
91. Filosa JA, Bonev AD, Straub SV, Meredith AL, Wilkerson MK, Aldrich RW, Nelson MT. Local potassium signaling couples neuronal activity to vasodilation in the brain. Nature Neurosci 9: 1397–1403, 2006. [Abstract] [Google Scholar]
92. Fiorio Pla A, Ong HL, Cheng KT, Brossa A, Bussolati B, Lockwich T, Paria B, Munaron L, Ambudkar IS. TRPV4 mediates tumor-derived endothelial cell migration via arachidonic acid-activated actin remodeling. Oncogene 31: 200–212, 2012. [Abstract] [Google Scholar]
93. Fischer MJ, Balasuriya D, Jeggle P, Goetze TA, McNaughton PA, Reeh PW, Edwardson JM. Direct evidence for functional TRPV1/TRPA1 heteromers. Pflügers Arch 466: 2229–2241, 2014. [Abstract] [Google Scholar]
94. Fisslthaler B, Popp R, Kiss L, Potente M, Harder DR, Fleming I, Busse R. Cytochrome P 450 2C is an EDHF synthase in coronary arteries. Nature 401: 493–497, 1999. [Abstract] [Google Scholar]
95. Flockerzi V, Nilius B. TRPs: truly remarkable proteins. Handb Exp Pharmacol 222: 1–12, 2014. [Abstract] [Google Scholar]
96. Folkman J, Shing Y. Angiogenesis. J Biol Chem 267: 10931–10934, 1992. [Abstract] [Google Scholar]
97. Freichel M, Suh SH, Pfeifer A, Schweig U, Trost C, Weissgerber P, Biel M, Philipp S, Freise D, Droogmans G, Hofmann F, Flockerzi V, Nilius B. Lack of an endothelial store-operated Ca2+ current impairs agonist-dependent vasorelaxation in TRP4−/− mice. Nature Cell Biol 3: 121–127, 2001. [Abstract] [Google Scholar]
98. Freichel M, Vennekens R, Olausson J, Hoffmann M, Muller C, Stolz S, Scheunemann J, Weissgerber P, Flockerzi V. Functional role of TRPC proteins in vivo: lessons from TRPC-deficient mouse models. Biochem Biophys Res Commun 322: 1352–1358, 2004. [Abstract] [Google Scholar]
99. Fuchs B, Rupp M, Ghofrani HA, Schermuly RT, Seeger W, Grimminger F, Gudermann T, Dietrich A, Weissmann N. Diacylglycerol regulates acute hypoxic pulmonary vasoconstriction via TRPC6. Respir Res 12: 20, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
100. Furchgott RF, Zawadzki JV. The obligatory role of endothelial cells in the relaxation of arterial smooth muscle by acetylcholine. Nature 288: 373–376, 1980. [Abstract] [Google Scholar]
101. Gao F, Wang DH. Hypotension induced by activation of the transient receptor potential vanilloid 4 channels: role of Ca2+-activated K+ channels and sensory nerves. J Hypertens 28: 102–110, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
102. Gao G, Bai XY, Xuan C, Liu XC, Jing WB, Novakovic A, Yang Q, He GW. Role of TRPC3 channel in human internal mammary artery. Arch Med Res 43: 431–437, 2012. [Abstract] [Google Scholar]
103. Garcia RL, Schilling WP. Differential expression of mammalian TRP homologues across tissues and cell lines. Biochem Biophys Res Commun 239: 279–283, 1997. [Abstract] [Google Scholar]
104. Garcia ZI, Bruhl A, Gonzales AL, Earley S. Basal protein kinase Cdelta activity is required for membrane localization and activity of TRPM4 channels in cerebral artery smooth muscle cells. Channels 5: 210–214, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
105. Garcia-Sanz N, Fernandez-Carvajal A, Morenilla-Palao C, Planells-Cases R, Fajardo-Sanchez E, Fernandez-Ballester G, Ferrer-Montiel A. Identification of a tetramerization domain in the C terminus of the vanilloid receptor. J Neurosci 24: 5307–5314, 2004. [Abstract] [Google Scholar]
106. Garland CJ, Hiley CR, Dora KA. EDHF: spreading the influence of the endothelium. Br J Pharmacol 164: 839–852, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
107. Garland CJ, Smirnov SV, Bagher P, Lim CS, Huang CY, Mitchell R, Stanley C, Pinkney A, Dora KA. The TRPM4 inhibitor 9-phenanthrol activates endothelial cell intermediate conductance calcium-activated potassium channels in rat isolated mesenteric artery. Br J Pharmacol 172: 1114–1123, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
108. Gauden V, Hu DE, Kurokawa T, Sarker MH, Fraser PA. Novel technique for estimating cerebrovascular permeability demonstrates capsazepine protection following ischemia-reperfusion. Microcirculation 14: 767–778, 2007. [Abstract] [Google Scholar]
109. Gazzieri D, Trevisani M, Tarantini F, Bechi P, Masotti G, Gensini GF, Castellani S, Marchionni N, Geppetti P, Harrison S. Ethanol dilates coronary arteries and increases coronary flow via transient receptor potential vanilloid 1 and calcitonin gene-related peptide. Cardiovasc Res 70: 589–599, 2006. [Abstract] [Google Scholar]
110. Ge R, Tai Y, Sun Y, Zhou K, Yang S, Cheng T, Zou Q, Shen F, Wang Y. Critical role of TRPC6 channels in VEGF-mediated angiogenesis. Cancer Lett 283: 43–51, 2009. [Abstract] [Google Scholar]
111. Geppetti P, Materazzi S, Nicoletti P. The transient receptor potential vanilloid 1: role in airway inflammation and disease. Eur J Pharmacol 533: 207–214, 2006. [Abstract] [Google Scholar]
112. Gerzanich V, Woo SK, Vennekens R, Tsymbalyuk O, Ivanova S, Ivanov A, Geng Z, Chen Z, Nilius B, Flockerzi V, Freichel M, Simard JM. De novo expression of Trpm4 initiates secondary hemorrhage in spinal cord injury. Nature Med 15: 185–191, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
113. Golovina VA, Platoshyn O, Bailey CL, Wang J, Limsuwan A, Sweeney M, Rubin LJ, Yuan JX. Upregulated TRP and enhanced capacitative Ca2+ entry in human pulmonary artery myocytes during proliferation. Am J Physiol Heart Circ Physiol 280: H746–H755, 2001. [Abstract] [Google Scholar]
114. Gonzales AL, Amberg GC, Earley S. Ca2+ release from the sarcoplasmic reticulum is required for sustained TRPM4 activity in cerebral artery smooth muscle cells. Am J Physiol Cell Physiol 299: C279–C288, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
115. Gonzales AL, Earley S. Endogenous cytosolic Ca2+ buffering is necessary for TRPM4 activity in cerebral artery smooth muscle cells. Cell Calcium 51: 82–93, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
116. Gonzales AL, Earley S. Regulation of cerebral artery smooth muscle membrane potential by Ca2+-activated cation channels. Microcirculation 20: 337–347, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
117. Gonzales AL, Garcia ZI, Amberg GC, Earley S. Pharmacological inhibition of TRPM4 hyperpolarizes vascular smooth muscle. Am J Physiol Cell Physiol 299: C1195–C1202, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
118. Gonzales AL, Yang Y, Sullivan MN, Sanders L, Dabertrand F, Hill-Eubanks DC, Nelson MT, Earley S. A PLCgamma1-dependent, force-sensitive signaling network in the myogenic constriction of cerebral arteries. Science Signaling 7: ra49, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
119. Grand T, Demion M, Norez C, Mettey Y, Launay P, Becq F, Bois P, Guinamard R. 9-Phenanthrol inhibits human TRPM4 but not TRPM5 cationic channels. Br J Pharmacol 153: 1697–1705, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
120. Graziani A, Poteser M, Heupel WM, Schleifer H, Krenn M, Drenckhahn D, Romanin C, Baumgartner W, Groschner K. Cell-cell contact formation governs Ca2+ signaling by TRPC4 in the vascular endothelium: evidence for a regulatory TRPC4-beta-catenin interaction. J Biol Chem 285: 4213–4223, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
121. Griffin MD, Torres VE, Grande JP, Kumar R. Vascular expression of polycystin. J Am Soc Nephrol 8: 616–626, 1997. [Abstract] [Google Scholar]
122. Groschner K, Hingel S, Lintschinger B, Balzer M, Romanin C, Zhu X, Schreibmayer W. Trp proteins form store-operated cation channels in human vascular endothelial cells. FEBS Lett 437: 101–106, 1998. [Abstract] [Google Scholar]
123. Guarini G, Ohanyan VA, Kmetz JG, DelloStritto DJ, Thoppil RJ, Thodeti CK, Meszaros JG, Damron DS, Bratz IN. Disruption of TRPV1-mediated coupling of coronary blood flow to cardiac metabolism in diabetic mice: role of nitric oxide and BK channels. Am J Physiol Heart Circ Physiol 303: H216–H223, 2012. [Abstract] [Google Scholar]
124. Guibert C, Beech DJ. Positive and negative coupling of the endothelin ETA receptor to Ca2+-permeable channels in rabbit cerebral cortex arterioles. J Physiol 514: 843–856, 1999. [Abstract] [Google Scholar]
125. Guinamard R, Demion M, Magaud C, Potreau D, Bois P. Functional expression of the TRPM4 cationic current in ventricular cardiomyocytes from spontaneously hypertensive rats. Hypertension 48: 587–594, 2006. [Abstract] [Google Scholar]
126. Hamanaka K, Jian MY, Townsley MI, King JA, Liedtke W, Weber DS, Eyal FG, Clapp MM, Parker JC. TRPV4 channels augment macrophage activation and ventilator-induced lung injury. Am J Physiol Lung Cell Mol Physiol 299: L353–L362, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
127. Hamanaka K, Jian MY, Weber DS, Alvarez DF, Townsley MI, Al-Mehdi AB, King JA, Liedtke W, Parker JC. TRPV4 initiates the acute calcium-dependent permeability increase during ventilator-induced lung injury in isolated mouse lungs. Am J Physiol Lung Cell Mol Physiol 293: L923–L932, 2007. [Abstract] [Google Scholar]
128. Hamdollah Zadeh MA, Glass CA, Magnussen A, Hancox JC, Bates DO. VEGF-mediated elevated intracellular calcium and angiogenesis in human microvascular endothelial cells in vitro are inhibited by dominant negative TRPC6. Microcirculation 15: 605–614, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
129. 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 9: 163–173, 2002. [Abstract] [Google Scholar]
130. Hardie RC. A brief history of trp: commentary and personal perspective. Pflügers Arch 461: 493–498, 2011. [Abstract] [Google Scholar]
131. Hartmann J, Dragicevic E, Adelsberger H, Henning HA, Sumser M, Abramowitz J, Blum R, Dietrich A, Freichel M, Flockerzi V, Birnbaumer L, Konnerth A. TRPC3 channels are required for synaptic transmission and motor coordination. Neuron 59: 392–398, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
132. 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 2: e827, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
133. Hassane S, Claij N, Jodar M, Dedman A, Lauritzen I, Duprat F, Koenderman JS, van der Wal A, Breuning MH, de Heer E, Honore E, DeRuiter MC, Peters DJ. Pkd1-inactivation in vascular smooth muscle cells and adaptation to hypertension. Lab Invest 91: 24–32, 2011. [Abstract] [Google Scholar]
134. He Y, Yao G, Savoia C, Touyz RM. Transient receptor potential melastatin 7 ion channels regulate magnesium homeostasis in vascular smooth muscle cells: role of angiotensin II. Circ Res 96: 207–215, 2005. [Abstract] [Google Scholar]
135. Hecquet CM, Ahmmed GU, Vogel SM, Malik AB. Role of TRPM2 channel in mediating H2O2-induced Ca2+ entry and endothelial hyperpermeability. Circ Res 102: 347–355, 2008. [Abstract] [Google Scholar]
136. 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 114: 469–479, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
137. Helliwell RM, Large WA. Alpha 1-adrenoceptor activation of a non-selective cation current in rabbit portal vein by 1,2-diacyl-sn-glycerol. J Physiol 499: 417–428, 1997. [Abstract] [Google Scholar]
138. Hoenderop JG, Voets T, Hoefs S, Weidema F, Prenen J, Nilius B, Bindels RJ. Homo- and heterotetrameric architecture of the epithelial Ca2+ channels TRPV5 and TRPV6. EMBO J 22: 776–785, 2003. [Europe PMC free article] [Abstract] [Google Scholar]
139. Hofmann T, Chubanov V, Gudermann T, Montell C. TRPM5 is a voltage-modulated and Ca2+-activated monovalent selective cation channel. Curr Biol 13: 1153–1158, 2003. [Abstract] [Google Scholar]
140. Hofmann T, Obukhov AG, Schaefer M, Harteneck C, Gudermann T, Schultz G. Direct activation of human TRPC6 and TRPC3 channels by diacylglycerol. Nature 397: 259–263, 1999. [Abstract] [Google Scholar]
141. Hofmann T, Schaefer M, Schultz G, Gudermann T. Subunit composition of mammalian transient receptor potential channels in living cells. Proc Natl Acad Sci USA 99: 7461–7466, 2002. [Europe PMC free article] [Abstract] [Google Scholar]
142. Imai Y, Itsuki K, Okamura Y, Inoue R, Mori MX. A self-limiting regulation of vasoconstrictor-activated TRPC3/C6/C7 channels coupled to PI(4,5)P2-diacylglycerol signalling. J Physiol 590: 1101–1119, 2012. [Abstract] [Google Scholar]
143. Imig JD, Dimitropoulou C, Reddy DS, White RE, Falck JR. Afferent arteriolar dilation to 11,12-EET analogs involves PP2A activity and Ca2+-activated K+ channels. Microcirculation 15: 137–150, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
144. Inoue R, Jensen LJ, Jian Z, Shi J, Hai L, Lurie AI, Henriksen FH, Salomonsson M, Morita H, Kawarabayashi Y, Mori M, Mori Y, Ito Y. Synergistic activation of vascular TRPC6 channel by receptor and mechanical stimulation via phospholipase C/diacylglycerol and phospholipase A2/omega-hydroxylase/20-HETE pathways. Circ Res 104: 1399–1409, 2009. [Abstract] [Google Scholar]
145. Inoue R, Jensen LJ, Shi J, Morita H, Nishida M, Honda A, Ito Y. Transient receptor potential channels in cardiovascular function and disease. Circ Res 99: 119–131, 2006. [Abstract] [Google Scholar]
146. Inoue R, Okada T, Onoue H, Hara Y, Shimizu S, Naitoh S, Ito Y, Mori Y. The transient receptor potential protein homologue TRP6 is the essential component of vascular alpha(1)-adrenoceptor-activated Ca2+-permeable cation channel. Circ Res 88: 325–332, 2001. [Abstract] [Google Scholar]
147. Itsuki K, Imai Y, Hase H, Okamura Y, Inoue R, Mori MX. PLC-mediated PI(4,5)P2 hydrolysis regulates activation and inactivation of TRPC6/7 channels. J Gen Physiol 143: 183–201, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
148. Ji RR, Samad TA, Jin SX, Schmoll R, Woolf CJ. p38 MAPK activation by NGF in primary sensory neurons after inflammation increases TRPV1 levels and maintains heat hyperalgesia. Neuron 36: 57–68, 2002. [Abstract] [Google Scholar]
149. Jian MY, King JA, Al-Mehdi AB, Liedtke W, Townsley MI. High vascular pressure-induced lung injury requires P 450 epoxygenase-dependent activation of TRPV4. Am J Respir Cell Mol Biol 38: 386–392, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
150. Jin X, Touhey J, Gaudet R. Structure of the N-terminal ankyrin repeat domain of the TRPV2 ion channel. J Biol Chem 281: 25006–25010, 2006. [Abstract] [Google Scholar]
151. Johnson CD, Melanaphy D, Purse A, Stokesberry SA, Dickson P, Zholos AV. Transient receptor potential melastatin 8 channel involvement in the regulation of vascular tone. Am J Physiol Heart Circ Physiol 296: H1868–H1877, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
152. Ju M, Shi J, Saleh SN, Albert AP, Large WA. Ins(1,4,5)P3 interacts with PIP2 to regulate activation of TRPC6/C7 channels by diacylglycerol in native vascular myocytes. J Physiol 588: 1419–1433, 2010. [Abstract] [Google Scholar]
153. Jung C, Gene GG, Tomas M, Plata C, Selent J, Pastor M, Fandos C, Senti M, Lucas G, Elosua R, Valverde MA. A gain-of-function SNP in TRPC4 cation channel protects against myocardial infarction. Cardiovasc Res 91: 465–471, 2011. [Abstract] [Google Scholar]
154. Jung S, Strotmann R, Schultz G, Plant TD. TRPC6 is a candidate channel involved in receptor-stimulated cation currents in A7r5 smooth muscle cells. Am J Physiol Cell Physiol 282: C347–C359, 2002. [Abstract] [Google Scholar]
155. Jurek SC, Hirano-Kobayashi M, Chiang H, Kohane DS, Matthews BD. Prevention of ventilator-induced lung edema by inhalation of nanoparticles releasing ruthenium red. Am J Respir Cell Mol Biol 50: 1107–1117, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
156. Kamouchi M, Philipp S, Flockerzi V, Wissenbach U, Mamin A, Raeymaekers L, Eggermont J, Droogmans G, Nilius B. Properties of heterologously expressed hTRP3 channels in bovine pulmonary artery endothelial cells. J Physiol 518: 345–358, 1999. [Abstract] [Google Scholar]
157. Karashima Y, Prenen J, Talavera K, Janssens A, Voets T, Nilius B. Agonist-induced changes in Ca2+ permeation through the nociceptor cation channel TRPA1. Biophys J 98: 773–783, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
158. Kark T, Bagi Z, Lizanecz E, Pasztor ET, Erdei N, Czikora A, Papp Z, Edes I, Porszasz R, Toth A. Tissue-specific regulation of microvascular diameter: opposite functional roles of neuronal and smooth muscle located vanilloid receptor-1. Mol Pharmacol 73: 1405–1412, 2008. [Abstract] [Google Scholar]
159. Keseru B, Barbosa-Sicard E, Popp R, Fisslthaler B, Dietrich A, Gudermann T, Hammock BD, Falck JR, Weissmann N, Busse R, Fleming I. Epoxyeicosatrienoic acids and the soluble epoxide hydrolase are determinants of pulmonary artery pressure and the acute hypoxic pulmonary vasoconstrictor response. FASEB J 22: 4306–4315, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
160. Kini V, Chavez A, Mehta D. A new role for PTEN in regulating transient receptor potential canonical channel 6-mediated Ca2+ entry, endothelial permeability, and angiogenesis. J Biol Chem 285: 33082–33091, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
161. Kirby BS, Bruhl A, Sullivan MN, Francis M, Dinenno FA, Earley S. Robust internal elastic lamina fenestration in skeletal muscle arteries. PloS One 8: e54849, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
162. Kiselyov K, Xu X, Mozhayeva G, Kuo T, Pessah I, Mignery G, Zhu X, Birnbaumer L, Muallem S. Functional interaction between InsP3 receptors and store-operated Htrp3 channels. Nature 396: 478–482, 1998. [Abstract] [Google Scholar]
163. Knot HJ, Nelson MT. Regulation of arterial diameter and wall [Ca2+] in cerebral arteries of rat by membrane potential and intravascular pressure. J Physiol 508: 199–209, 1998. [Abstract] [Google Scholar]
164. Kobori T, Smith GD, Sandford R, Edwardson JM. The transient receptor potential channels TRPP2 and TRPC1 form a heterotetramer with a 2:2 stoichiometry and an alternating subunit arrangement. J Biol Chem 284: 35507–35513, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
165. Kochukov MY, Balasubramanian A, Abramowitz J, Birnbaumer L, Marrelli SP. Activation of endothelial transient receptor potential c3 channel is required for small conductance calcium-activated potassium channel activation and sustained endothelial hyperpolarization and vasodilation of cerebral artery. J Am Heart Assoc 2014, 10.1161/JAHA.114.000913. [Europe PMC free article] [Abstract] [CrossRef] [Google Scholar]
166. Kochukov MY, Balasubramanian A, Noel RC, Marrelli SP. Role of TRPC1 and TRPC3 channels in contraction and relaxation of mouse thoracic aorta. J Vasc Res 50: 11–20, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
168. Koenig S, Schernthaner M, Maechler H, Kappe CO, Glasnov TN, Hoefler G, Braune M, Wittchow E, Groschner K. A TRPC3 blocker, ethyl-1-(4-(2,3,3-trichloroacrylamide)phenyl)-5-(trifluoromethyl)-1H-pyrazole-4-c arboxylate (Pyr3), prevents stent-induced arterial remodeling. J Pharmacol Exp Ther 344: 33–40, 2013. [Abstract] [Google Scholar]
169. 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. Arteriosclerosis Thrombosis Vasc Biol 26: 1495–1502, 2006. [Abstract] [Google Scholar]
170. Kohn EC, Alessandro R, Spoonster J, Wersto RP, Liotta LA. Angiogenesis: role of calcium-mediated signal transduction. Proc Natl Acad Sci USA 92: 1307–1311, 1995. [Europe PMC free article] [Abstract] [Google Scholar]
171. Kraft R. The Na+/Ca2+ exchange inhibitor KB-R7943 potently blocks TRPC channels. Biochem Biophys Res Commun 361: 230–236, 2007. [Abstract] [Google Scholar]
172. Krapivinsky G, Krapivinsky L, Manasian Y, Clapham DE. The TRPM7 chanzyme is cleaved to release a chromatin-modifying kinase. Cell 157: 1061–1072, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
173. Kumar B, Dreja K, Shah SS, Cheong A, Xu SZ, Sukumar P, Naylor J, Forte A, Cipollaro M, McHugh D, Kingston PA, Heagerty AM, Munsch CM, Bergdahl A, Hultgardh-Nilsson A, Gomez MF, Porter KE, Hellstrand P, Beech DJ. Upregulated TRPC1 channel in vascular injury in vivo and its role in human neointimal hyperplasia. Circ Res 98: 557–563, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
174. Kunichika N, Yu Y, Remillard CV, Platoshyn O, Zhang S, Yuan JX. Overexpression of TRPC1 enhances pulmonary vasoconstriction induced by capacitative Ca2+ entry. Am J Physiol Lung Cell Mol Physiol 287: L962–L969, 2004. [Abstract] [Google Scholar]
175. Kwan HY, Huang Y, Yao X. Regulation of canonical transient receptor potential isoform 3 (TRPC3) channel by protein kinase G. Proc Natl Acad Sci USA 101: 2625–2630, 2004. [Europe PMC free article] [Abstract] [Google Scholar]
176. Kwan HY, Shen B, Ma X, Kwok YC, Huang Y, Man YB, Yu S, Yao X. TRPC1 associates with BK(Ca) channel to form a signal complex in vascular smooth muscle cells. Circ Res 104: 670–678, 2009. [Abstract] [Google Scholar]
177. Large WA, Saleh SN, Albert AP. Role of phosphoinositol 4,5-bisphosphate and diacylglycerol in regulating native TRPC channel proteins in vascular smooth muscle. Cell Calcium 45: 574–582, 2009. [Abstract] [Google Scholar]
178. Launay P, Cheng H, Srivatsan S, Penner R, Fleig A, Kinet JP. TRPM4 regulates calcium oscillations after T cell activation. Science 306: 1374–1377, 2004. [Abstract] [Google Scholar]
179. Launay P, Fleig A, Perraud AL, Scharenberg AM, Penner R, Kinet JP. TRPM4 is a Ca2+-activated nonselective cation channel mediating cell membrane depolarization. Cell 109: 397–407, 2002. [Abstract] [Google Scholar]
180. Ledoux J, Taylor MS, Bonev AD, Hannah RM, Solodushko V, Shui B, Tallini Y, Kotlikoff MI, Nelson MT. Functional architecture of inositol 1,4,5-trisphosphate signaling in restricted spaces of myoendothelial projections. Proc Natl Acad Sci USA 105: 9627–9632, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
181. Lemonnier L, Trebak M, Putney JW Jr. Complex regulation of the TRPC3, 6 and 7 channel subfamily by diacylglycerol and phosphatidylinositol-4,5-bisphosphate. Cell Calcium 43: 506–514, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
182. Leuner K, Kazanski V, Muller M, Essin K, Henke B, Gollasch M, Harteneck C, Muller WE. Hyperforin–a key constituent of St. John's wort specifically activates TRPC6 channels. FASEB J 21: 4101–4111, 2007. [Abstract] [Google Scholar]
183. Leung FP, Yung LM, Yao X, Laher I, Huang Y. Store-operated calcium entry in vascular smooth muscle. Br J Pharmacol 153: 846–857, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
184. Li J, Sukumar P, Milligan CJ, Kumar B, Ma ZY, Munsch CM, Jiang LH, Porter KE, Beech DJ. Interactions, functions, and independence of plasma membrane STIM1 and TRPC1 in vascular smooth muscle cells. Circ Res 103: e97–104, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
185. Li J, Wang DH. Function and regulation of the vanilloid receptor in rats fed a high salt diet. J Hypertens 21: 1525–1530, 2003. [Abstract] [Google Scholar]
186. Li M, Jiang J, Yue L. Functional characterization of homo- and heteromeric channel kinases TRPM6 and TRPM7. J Gen Physiol 127: 525–537, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
187. Li S, Ran Y, Zheng X, Pang X, Wang Z, Zhang R, Zhu D. 15-HETE mediates sub-acute hypoxia-induced TRPC1 expression and enhanced capacitative calcium entry in rat distal pulmonary arterial myocytes. Prostaglandins Other Lipid Mediators 93: 60–74, 2010. [Abstract] [Google Scholar]
188. Li Y, Baylie RL, Tavares MJ, Brayden JE. TRPM4 channels couple purinergic receptor mechanoactivation and myogenic tone development in cerebral parenchymal arterioles. J Cereb Blood Flow Metab 34: 1706–1714, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
189. Li Y, Jiang H, Ruan C, Zhong J, Gao P, Zhu D, Niu W, Guo S. The interaction of transient receptor potential melastatin 7 with macrophages promotes vascular adventitial remodeling in transverse aortic constriction rats. Hypertension Res 37: 35–42, 2014. [Abstract] [Google Scholar]
191. Liang GH, Kim JA, Seol GH, Choi S, Suh SH. The Na+/Ca2+ exchanger inhibitor KB-R7943 activates large-conductance Ca2+-activated K+ channels in endothelial and vascular smooth muscle cells. Eur J Pharmacol 582: 35–41, 2008. [Abstract] [Google Scholar]
192. Liao M, Cao E, Julius D, Cheng Y. Structure of the TRPV1 ion channel determined by electron cryo-microscopy. Nature 504: 107–112, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
193. 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 103: 525–535, 2000. [Europe PMC free article] [Abstract] [Google Scholar]
194. Lin MJ, Leung GP, Zhang WM, Yang XR, Yip KP, Tse CM, Sham JS. Chronic hypoxia-induced upregulation of store-operated and receptor-operated Ca2+ channels in pulmonary arterial smooth muscle cells: a novel mechanism of hypoxic pulmonary hypertension. Circ Res 95: 496–505, 2004. [Abstract] [Google Scholar]
195. Linde CI, Karashima E, Raina H, Zulian A, Wier WG, Hamlyn JM, Ferrari P, Blaustein MP, Golovina VA. Increased arterial smooth muscle Ca2+ signaling, vasoconstriction, and myogenic reactivity in Milan hypertensive rats. Am J Physiol Heart Circ Physiol 302: H611–H620, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
196. Lindsey SH, Songu-Mize E. Stretch-induced TRPC4 downregulation is accompanied by reduced capacitative Ca2+ entry in WKY but not SHR mesenteric smooth muscle cells. Clin Exp Hypertens 32: 288–292, 2010. [Abstract] [Google Scholar]
197. Lindsey SH, Tribe RM, Songu-Mize E. Cyclic stretch decreases TRPC4 protein and capacitative calcium entry in rat vascular smooth muscle cells. Life Sci 83: 29–34, 2008. [Abstract] [Google Scholar]
198. Liu CL, Huang Y, Ngai CY, Leung YK, Yao XQ. TRPC3 is involved in flow- and bradykinin-induced vasodilation in rat small mesenteric arteries. Acta Pharmacol Sin 27: 981–990, 2006. [Abstract] [Google Scholar]
199. Liu D, Yang D, He H, Chen X, Cao T, Feng X, Ma L, Luo Z, Wang L, Yan Z, Zhu Z, Tepel M. Increased transient receptor potential canonical type 3 channels in vasculature from hypertensive rats. Hypertension 53: 70–76, 2009. [Abstract] [Google Scholar]
200. Liu XR, Zhang MF, Yang N, Liu Q, Wang RX, Cao YN, Yang XR, Sham JS, Lin MJ. Enhanced store-operated Ca2+ entry and TRPC channel expression in pulmonary arteries of monocrotaline-induced pulmonary hypertensive rats. Am J Physiol Cell Physiol 302: C77–C87, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
201. Lizanecz E, Bagi Z, Pasztor ET, Papp Z, Edes I, Kedei N, Blumberg PM, Toth A. Phosphorylation-dependent desensitization by anandamide of vanilloid receptor-1 (TRPV1) function in rat skeletal muscle arterioles and in Chinese hamster ovary cells expressing TRPV1. Mol Pharmacol 69: 1015–1023, 2006. [Abstract] [Google Scholar]
202. Loga F, Domes K, Freichel M, Flockerzi V, Dietrich A, Birnbaumer L, Hofmann F, Wegener JW. The role of cGMP/cGKI signalling and Trpc channels in regulation of vascular tone. Cardiovasc Res 100: 280–287, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
203. Loh KP, Ng G, Yu CY, Fhu CK, Yu D, Vennekens R, Nilius B, Soong TW, Liao P. TRPM4 inhibition promotes angiogenesis after ischemic stroke. Pflügers Arch 466: 563–576, 2014. [Abstract] [Google Scholar]
204. Loot AE, Popp R, Fisslthaler B, Vriens J, Nilius B, Fleming I. Role of cytochrome P 450-dependent transient receptor potential V4 activation in flow-induced vasodilatation. Cardiovasc Res 80: 445–452, 2008. [Abstract] [Google Scholar]
205. Lu W, Wang J, Shimoda LA, Sylvester JT. Differences in STIM1 and TRPC expression in proximal and distal pulmonary arterial smooth muscle are associated with differences in Ca2+ responses to hypoxia. Am J Physiol Lung Cell Mol Physiol 295: L104–L113, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
206. Lucas P, Ukhanov K, Leinders-Zufall T, Zufall F. A diacylglycerol-gated cation channel in vomeronasal neuron dendrites is impaired in TRPC2 mutant mice: mechanism of pheromone transduction. Neuron 40: 551–561, 2003. [Abstract] [Google Scholar]
207. Ma X, Cheng KT, Wong CO, O'Neil RG, Birnbaumer L, Ambudkar IS, Yao X. Heteromeric TRPV4–C1 channels contribute to store-operated Ca2+ entry in vascular endothelial cells. Cell Calcium 50: 502–509, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
208. Macpherson LJ, Geierstanger BH, Viswanath V, Bandell M, Eid SR, Hwang S, Patapoutian A. The pungency of garlic: activation of TRPA1 and TRPV1 in response to allicin. Curr Biol 15: 929–934, 2005. [Abstract] [Google Scholar]
209. Malczyk M, Veith C, Fuchs B, Hofmann K, Storch U, Schermuly RT, Witzenrath M, Ahlbrecht K, Fecher-Trost C, Flockerzi V, Ghofrani HA, Grimminger F, Seeger W, Gudermann T, Dietrich A, Weissmann N. Classical transient receptor potential channel 1 in hypoxia-induced pulmonary hypertension. Am J Respir Crit Care Med 188: 1451–1459, 2013. [Abstract] [Google Scholar]
210. Mani BK, Brueggemann LI, Cribbs LL, Byron KL. Opposite regulation of KCNQ5 and TRPC6 channels contributes to vasopressin-stimulated calcium spiking responses in A7r5 vascular smooth muscle cells. Cell Calcium 45: 400–411, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
211. Marrelli SP, O'Neil RG, Brown RC, Bryan RM Jr. PLA2 and TRPV4 channels regulate endothelial calcium in cerebral arteries. Am J Physiol Heart Circ Physiol 292: H1390–H1397, 2007. [Abstract] [Google Scholar]
212. Marshall NJ, Liang L, Bodkin J, Dessapt-Baradez C, Nandi M, Collot-Teixeira S, Smillie SJ, Lalgi K, Fernandes ES, Gnudi L, Brain SD. A role for TRPV1 in influencing the onset of cardiovascular disease in obesity. Hypertension 61: 246–252, 2013. [Abstract] [Google Scholar]
213. Maruyama Y, Nakanishi Y, Walsh EJ, Wilson DP, Welsh DG, Cole WC. Heteromultimeric TRPC6-TRPC7 channels contribute to arginine vasopressin-induced cation current of A7r5 vascular smooth muscle cells. Circ Res 98: 1520–1527, 2006. [Abstract] [Google Scholar]
214. Mathar I, Vennekens R, Meissner M, Kees F, Van der Mieren G, Camacho Londono JE, Uhl S, Voets T, Hummel B, van den Bergh A, Herijgers P, Nilius B, Flockerzi V, Schweda F, Freichel M. Increased catecholamine secretion contributes to hypertension in TRPM4-deficient mice. J Clin Invest 120: 3267–3279, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
215. McCleverty CJ, Koesema E, Patapoutian A, Lesley SA, Kreusch A. Crystal structure of the human TRPV2 channel ankyrin repeat domain. Protein Sci 15: 2201–2206, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
216. McElroy SP, Gurney AM, Drummond RM. Pharmacological profile of store-operated Ca2+ entry in intrapulmonary artery smooth muscle cells. Eur J Pharmacol 584: 10–20, 2008. [Abstract] [Google Scholar]
217. Mederos y Schnitzler M, Storch U, Meibers S, Nurwakagari P, Breit A, Essin K, Gollasch M, Gudermann T. Gq-coupled receptors as mechanosensors mediating myogenic vasoconstriction. EMBO J 27: 3092–3103, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
218. Mehta D, Ahmmed GU, Paria BC, Holinstat M, Voyno-Yasenetskaya T, Tiruppathi C, Minshall RD, Malik AB. RhoA interaction with inositol 1,4,5-trisphosphate receptor and transient receptor potential channel-1 regulates Ca2+ entry. Role in signaling increased endothelial permeability. J Biol Chem 278: 33492–33500, 2003. [Abstract] [Google Scholar]
219. Mehta D, Malik AB. Signaling mechanisms regulating endothelial permeability. Physiol Rev 86: 279–367, 2006. [Abstract] [Google Scholar]
220. Mendoza SA, Fang J, Gutterman DD, Wilcox DA, Bubolz AH, Li R, Suzuki M, Zhang DX. TRPV4-mediated endothelial Ca2+ influx and vasodilation in response to shear stress. Am J Physiol Heart Circ Physiol 298: H466–H476, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
221. 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. J Gen Physiol 143: 559–575, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
222. Michel CC, Curry FE. Microvascular permeability. Physiol Rev 79: 703–761, 1999. [Abstract] [Google Scholar]
223. Minke B. The history of the Drosophila TRP channel: the birth of a new channel superfamily. J Neurogenet 24: 216–233, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
224. Minke B, Wu C, Pak WL. Induction of photoreceptor voltage noise in the dark in Drosophila mutant. Nature 258: 84–87, 1975. [Abstract] [Google Scholar]
225. Mio K, Ogura T, Kiyonaka S, Hiroaki Y, Tanimura Y, Fujiyoshi Y, Mori Y, Sato C. The TRPC3 channel has a large internal chamber surrounded by signal sensing antennas. J Mol Biol 367: 373–383, 2007. [Abstract] [Google Scholar]
226. Mironov SL. Metabotropic glutamate receptors activate dendritic calcium waves and TRPM channels which drive rhythmic respiratory patterns in mice. J Physiol 586: 2277–2291, 2008. [Abstract] [Google Scholar]
227. Moiseenkova-Bell VY, Stanciu LA, Serysheva II, Tobe BJ, Wensel TG. Structure of TRPV1 channel revealed by electron cryomicroscopy. Proc Natl Acad Sci USA 105: 7451–7455, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
228. Monteilh-Zoller MK, Hermosura MC, Nadler MJ, Scharenberg AM, Penner R, Fleig A. TRPM7 provides an ion channel mechanism for cellular entry of trace metal ions. J Gen Physiol 121: 49–60, 2003. [Europe PMC free article] [Abstract] [Google Scholar]
229. Montell C. The history of TRP channels, a commentary and reflection. Pflügers Arch 461: 499–506, 2011. [Abstract] [Google Scholar]
230. Montell C, Jones K, Hafen E, Rubin G. Rescue of the Drosophila phototransduction mutation trp by germline transformation. Science 230: 1040–1043, 1985. [Abstract] [Google Scholar]
231. Montell C, Rubin GM. Molecular characterization of the Drosophila trp locus: a putative integral membrane protein required for phototransduction. Neuron 2: 1313–1323, 1989. [Abstract] [Google Scholar]
232. Moore TM, Brough GH, Babal P, Kelly JJ, Li M, Stevens T. Store-operated calcium entry promotes shape change in pulmonary endothelial cells expressing Trp1. Am J Physiol Lung Cell Mol Physiol 275: L574–L582, 1998. [Abstract] [Google Scholar]
233. Morales-Lazaro SL, Simon SA, Rosenbaum T. The role of endogenous molecules in modulating pain through transient receptor potential vanilloid 1 (TRPV1). J Physiol 591: 3109–3121, 2013. [Abstract] [Google Scholar]
234. Morita H, Honda A, Inoue R, Ito Y, Abe K, Nelson MT, Brayden JE. Membrane stretch-induced activation of a TRPM4-like nonselective cation channel in cerebral artery myocytes. J Pharmacol Sci 103: 417–426, 2007. [Abstract] [Google Scholar]
235. Moussaieff A, Rimmerman N, Bregman T, Straiker A, Felder CC, Shoham S, Kashman Y, Huang SM, Lee H, Shohami E, Mackie K, Caterina MJ, Walker JM, Fride E, Mechoulam R. Incensole acetate, an incense component, elicits psychoactivity by activating TRPV3 channels in the brain. FASEB J 22: 3024–3034, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
236. Moussaieff A, Yu J, Zhu H, Gattoni-Celli S, Shohami E, Kindy MS. Protective effects of incensole acetate on cerebral ischemic injury. Brain Res 1443: 89–97, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
237. Muller M, Essin K, Hill K, Beschmann H, Rubant S, Schempp CM, Gollasch M, Boehncke WH, Harteneck C, Muller WE, Leuner K. Specific TRPC6 channel activation, a novel approach to stimulate keratinocyte differentiation. J Biol Chem 283: 33942–33954, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
238. Muraki K, Iwata Y, Katanosaka Y, Ito T, Ohya S, Shigekawa M, Imaizumi Y. TRPV2 is a component of osmotically sensitive cation channels in murine aortic myocytes. Circ Res 93: 829–838, 2003. [Abstract] [Google Scholar]
239. Muzzi M, Felici R, Cavone L, Gerace E, Minassi A, Appendino G, Moroni F, Chiarugi A. Ischemic neuroprotection by TRPV1 receptor-induced hypothermia. J Cerebral Blood Flow Metab 32: 978–982, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
240. Nadler MJ, Hermosura MC, Inabe K, Perraud AL, Zhu Q, Stokes AJ, Kurosaki T, Kinet JP, Penner R, Scharenberg AM, Fleig A. LTRPC7 is a Mg.ATP-regulated divalent cation channel required for cell viability. Nature 411: 590–595, 2001. [Abstract] [Google Scholar]
241. Nagata K, Duggan A, Kumar G, Garcia-Anoveros J. Nociceptor and hair cell transducer properties of TRPA1, a channel for pain and hearing. J Neurosci 25: 4052–4061, 2005. [Abstract] [Google Scholar]
243. Narayanan D, Bulley S, Leo MD, Burris SK, Gabrick KS, Boop FA, Jaggar JH. Smooth muscle cell transient receptor potential polycystin-2 (TRPP2) channels contribute to the myogenic response in cerebral arteries. J Physiol 591: 5031–5046, 2013. [Abstract] [Google Scholar]
244. Nassini R, Materazzi S, Andre E, Sartiani L, Aldini G, Trevisani M, Carnini C, Massi D, Pedretti P, Carini M, Cerbai E, Preti D, Villetti G, Civelli M, Trevisan G, Azzari C, Stokesberry S, Sadofsky L, McGarvey L, Patacchini R, Geppetti P. Acetaminophen, via its reactive metabolite N-acetyl-p-benzo-quinoneimine and transient receptor potential ankyrin-1 stimulation, causes neurogenic inflammation in the airways and other tissues in rodents. FASEB J 24: 4904–4916, 2010. [Abstract] [Google Scholar]
245. Ng LC, Airey JA, Hume JR. The contribution of TRPC1 and STIM1 to capacitative Ca2+ entry in pulmonary artery. Adv Exp Med Biol 661: 123–135, 2010. [Abstract] [Google Scholar]
246. Ng LC, McCormack MD, Airey JA, Singer CA, Keller PS, Shen XM, Hume JR. TRPC1 and STIM1 mediate capacitative Ca2+ entry in mouse pulmonary arterial smooth muscle cells. J Physiol 587: 2429–2442, 2009. [Abstract] [Google Scholar]
247. Ng LC, O'Neill KG, French D, Airey JA, Singer CA, Tian H, Shen XM, Hume JR. TRPC1 and Orai1 interact with STIM1 and mediate capacitative Ca2+ entry caused by acute hypoxia in mouse pulmonary arterial smooth muscle cells. Am J Physiol Cell Physiol 303: C1156–C1172, 2012. [Abstract] [Google Scholar]
248. Nilius B, Droogmans G, Wondergem R. Transient receptor potential channels in endothelium: solving the calcium entry puzzle? Endothelium 10: 5–15, 2003. [Abstract] [Google Scholar]
249. Nilius B, Flockerzi V. What do we really know and what do we need to know: some controversies, perspectives, and surprises. Handb Exp Pharmacol 223: 1239–1280, 2014. [Abstract] [Google Scholar]
250. Nilius B, Mahieu F, Prenen J, Janssens A, Owsianik G, Vennekens R, Voets T. The Ca2+-activated cation channel TRPM4 is regulated by phosphatidylinositol 4,5-biphosphate. EMBO J 25: 467–478, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
251. Nilius B, Prenen J, Droogmans G, Voets T, Vennekens R, Freichel M, Wissenbach U, Flockerzi V. Voltage dependence of the Ca2+-activated cation channel TRPM4. J Biol Chem 278: 30813–30820, 2003. [Abstract] [Google Scholar]
252. Nilius B, Prenen J, Tang J, Wang C, Owsianik G, Janssens A, Voets T, Zhu MX. Regulation of the Ca2+ sensitivity of the nonselective cation channel TRPM4. J Biol Chem 280: 6423–6433, 2005. [Abstract] [Google Scholar]
253. Nilius B, Vennekens R, Prenen J, Hoenderop JG, Bindels RJ, Droogmans G. Whole-cell and single channel monovalent cation currents through the novel rabbit epithelial Ca2+ channel ECaC. J Physiol 527: 239–248, 2000. [Abstract] [Google Scholar]
254. Nishijima Y, Zheng X, Lund H, Suzuki M, Mattson DL, Zhang DX. Characterization of blood pressure and endothelial function in TRPV4-deficient mice with l-NAME- and angiotensin II-induced hypertension. Physiol Reports 2: e00199, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
255. Nishioka K, Nishida M, Ariyoshi M, Jian Z, Saiki S, Hirano M, Nakaya M, Sato Y, Kita S, Iwamoto T, Hirano K, Inoue R, Kurose H. Cilostazol suppresses angiotensin II-induced vasoconstriction via protein kinase A-mediated phosphorylation of the transient receptor potential canonical 6 channel. Arteriosclerosis Thrombosis Vasc Biol 31: 2278–2286, 2011. [Abstract] [Google Scholar]
256. Noorani MM, Noel RC, Marrelli SP. Upregulated TRPC3 and downregulated TRPC1 channel expression during hypertension is associated with increased vascular contractility in rat. Front Physiol 2: 42, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
257. Pankey EA, Kassan M, Choi SK, Matrougui K, Nossaman BD, Hyman AL, Kadowitz PJ. Vasodilator responses to acetylcholine are not mediated by the activation of soluble guanylate cyclase or Trpv4 channels in the rat. Am J Physiol Heart Circ Physiol 306: H1495–H1506, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
258. Parajuli SP, Hristov KL, Sullivan MN, Xin W, Smith AC, Earley S, Malysz J, Petkov GV. Control of urinary bladder smooth muscle excitability by the TRPM4 channel modulator 9-phenanthrol. Channels 7: 537–540, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
259. Paravicini TM, Yogi A, Mazur A, Touyz RM. Dysregulation of vascular TRPM7 and annexin-1 is associated with endothelial dysfunction in inherited hypomagnesemia. Hypertension 53: 423–429, 2009. [Abstract] [Google Scholar]
260. Parekh AB, Putney JW Jr. Store-operated calcium channels. Physiol Rev 85: 757–810, 2005. [Abstract] [Google Scholar]
261. Paria BC, Vogel SM, Ahmmed GU, Alamgir S, Shroff J, Malik AB, Tiruppathi C. Tumor necrosis factor-alpha-induced TRPC1 expression amplifies store-operated Ca2+ influx and endothelial permeability. Am J Physiol Lung Cell Mol Physiol 287: L1303–L1313, 2004. [Abstract] [Google Scholar]
262. Park HW, Kim JY, Choi SK, Lee YH, Zeng W, Kim KH, Muallem S, Lee MG. Serine-threonine kinase with-no-lysine 4 (WNK4) controls blood pressure via transient receptor potential canonical 3 (TRPC3) in the vasculature. Proc Natl Acad Sci USA 108: 10750–10755, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
263. Park JY, Hwang EM, Yarishkin O, Seo JH, Kim E, Yoo J, Yi GS, Kim DG, Park N, Ha CM, La JH, Kang D, Han J, Oh U, Hong SG. TRPM4b channel suppresses store-operated Ca2+ entry by a novel protein-protein interaction with the TRPC3 channel. Biochem Biophys Res Commun 368: 677–683, 2008. [Abstract] [Google Scholar]
264. Parker JC, Hashizumi M, Kelly SV, Francis M, Mouner M, Meyer AL, Townsley MI, Wu S, Cioffi DL, Taylor MS. TRPV4 calcium entry and surface expression attenuated by inhibition of myosin light chain kinase in rat pulmonary microvascular endothelial cells. Physiol Reports 1: e00121, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
265. Pegorini S, Braida D, Verzoni C, Guerini-Rocco C, Consalez GG, Croci L, Sala M. Capsaicin exhibits neuroprotective effects in a model of transient global cerebral ischemia in Mongolian gerbils. Br J Pharmacol 144: 727–735, 2005. [Europe PMC free article] [Abstract] [Google Scholar]
266. Peppiatt-Wildman CM, Albert AP, Saleh SN, Large WA. Endothelin-1 activates a Ca2+-permeable cation channel with TRPC3 and TRPC7 properties in rabbit coronary artery myocytes. J Physiol 580: 755–764, 2007. [Abstract] [Google Scholar]
267. 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 411: 595–599, 2001. [Abstract] [Google Scholar]
268. Petersen CC, Berridge MJ, Borgese MF, Bennett DL. Putative capacitative calcium entry channels: expression of Drosophila trp and evidence for the existence of vertebrate homologues. Biochem J 311: 41–44, 1995. [Europe PMC free article] [Abstract] [Google Scholar]
269. Pires PW, Earley S. TRPs in the kidney–location, location, location. Acta Physiol 213: 296–297, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
270. Planells-Cases R, Garcia-Sanz N, Morenilla-Palao C, Ferrer-Montiel A. Functional aspects and mechanisms of TRPV1 involvement in neurogenic inflammation that leads to thermal hyperalgesia. Pflügers Arch 451: 151–159, 2005. [Abstract] [Google Scholar]
271. Poblete IM, Orliac ML, Briones R, Adler-Graschinsky E, Huidobro-Toro JP. Anandamide elicits an acute release of nitric oxide through endothelial TRPV1 receptor activation in the rat arterial mesenteric bed. J Physiol 568: 539–551, 2005. [Abstract] [Google Scholar]
272. Pokreisz P, Fleming I, Kiss L, Barbosa-Sicard E, Fisslthaler B, Falck JR, Hammock BD, Kim IH, Szelid Z, Vermeersch P, Gillijns H, Pellens M, Grimminger F, van Zonneveld AJ, Collen D, Busse R, Janssens S. Cytochrome P 450 epoxygenase gene function in hypoxic pulmonary vasoconstriction and pulmonary vascular remodeling. Hypertension 47: 762–770, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
273. Porszasz R, Porkolab A, Ferencz A, Pataki T, Szilvassy Z, Szolcsanyi J. Capsaicin-induced nonneural vasoconstriction in canine mesenteric arteries. Eur J Pharmacol 441: 173–175, 2002. [Abstract] [Google Scholar]
274. Potier M, Gonzalez JC, Motiani RK, Abdullaev IF, Bisaillon JM, Singer HA, Trebak M. Evidence for STIM1- and Orai1-dependent store-operated calcium influx through ICRAC in vascular smooth muscle cells: role in proliferation and migration. FASEB J 23: 2425–2437, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
275. Prakriya M, Feske S, Gwack Y, Srikanth S, Rao A, Hogan PG. Orai1 is an essential pore subunit of the CRAC channel. Nature 443: 230–233, 2006. [Abstract] [Google Scholar]
276. Prawitt D, Monteilh-Zoller MK, Brixel L, Spangenberg C, Zabel B, Fleig A, Penner R. TRPM5 is a transient Ca2+-activated cation channel responding to rapid changes in [Ca2+]i. Proc Natl Acad Sci USA 100: 15166–15171, 2003. [Europe PMC free article] [Abstract] [Google Scholar]
277. Pulina MV, Zulian A, Berra-Romani R, Beskina O, Mazzocco-Spezzia A, Baryshnikov SG, Papparella I, Hamlyn JM, Blaustein MP, Golovina VA. Upregulation of Na+ and Ca2+ transporters in arterial smooth muscle from ouabain-induced hypertensive rats. Am J Physiol Heart Circ Physiol 298: H263–H274, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
278. Putney JW., Jr A model for receptor-regulated calcium entry. Cell Calcium 7: 1–12, 1986. [Abstract] [Google Scholar]
279. Qian Q, Hunter LW, Du H, Ren Q, Han Y, Sieck GC. Pkd2+/− vascular smooth muscles develop exaggerated vasocontraction in response to phenylephrine stimulation. J Am Soc Nephrol 18: 485–493, 2007. [Abstract] [Google Scholar]
280. Qian Q, Li M, Cai Y, Ward CJ, Somlo S, Harris PC, Torres VE. Analysis of the polycystins in aortic vascular smooth muscle cells. J Am Soc Nephrol 14: 2280–2287, 2003. [Abstract] [Google Scholar]
281. Qian X, Francis M, Solodushko V, Earley S, Taylor MS. Recruitment of dynamic endothelial Ca2+ signals by the TRPA1 channel activator AITC in rat cerebral arteries. Microcirculation 20: 138–148, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
282. Reading SA, Brayden JE. Central role of TRPM4 channels in cerebral blood flow regulation. Stroke 38: 2322–2328, 2007. [Abstract] [Google Scholar]
283. Reading SA, Earley S, Waldron BJ, Welsh DG, Brayden JE. TRPC3 mediates pyrimidine receptor-induced depolarization of cerebral arteries. Am J Physiol Heart Circ Physiol 288: H2055–H2061, 2005. [Abstract] [Google Scholar]
284. Richardson JD, Vasko MR. Cellular mechanisms of neurogenic inflammation. J Pharmacol Exp Ther 302: 839–845, 2002. [Abstract] [Google Scholar]
285. Rohacs T. Phosphoinositide regulation of TRP channels. Handb Exp Pharmacol 223: 1143–1176, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
286. Rohacs T. Regulation of TRP channels by PIP2. Pflügers Arch 453: 753–762, 2007. [Abstract] [Google Scholar]
287. Rohacs T, Lopes CM, Michailidis I, Logothetis DE. PI(4,5)P2 regulates the activation and desensitization of TRPM8 channels through the TRP domain. Nature Neurosci 8: 626–634, 2005. [Abstract] [Google Scholar]
288. Rohacs T, Nilius B. Regulation of transient receptor potential (TRP) channels by phosphoinositides. Pflügers Arch 455: 157–168, 2007. [Abstract] [Google Scholar]
289. Rosker C, Graziani A, Lukas M, Eder P, Zhu MX, Romanin C, Groschner K. Ca2+ signaling by TRPC3 involves Na+ entry and local coupling to the Na+/Ca2+ exchanger. J Biol Chem 279: 13696–13704, 2004. [Abstract] [Google Scholar]
290. Salas MM, Hargreaves KM, Akopian AN. TRPA1-mediated responses in trigeminal sensory neurons: interaction between TRPA1 and TRPV1. Eur J Neurosci 29: 1568–1578, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
291. Saleh SN, Albert AP, Large WA. Activation of native TRPC1/C5/C6 channels by endothelin-1 is mediated by both PIP3 and PIP2 in rabbit coronary artery myocytes. J Physiol 587: 5361–5375, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
292. Saleh SN, Albert AP, Large WA. Obligatory role for phosphatidylinositol 4,5-bisphosphate in activation of native TRPC1 store-operated channels in vascular myocytes. J Physiol 587: 531–540, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
293. Saleh SN, Albert AP, Peppiatt CM, Large WA. Angiotensin II activates two cation conductances with distinct TRPC1 and TRPC6 channel properties in rabbit mesenteric artery myocytes. J Physiol 577: 479–495, 2006. [Abstract] [Google Scholar]
294. Saleh SN, Albert AP, Peppiatt-Wildman CM, Large WA. Diverse properties of store-operated TRPC channels activated by protein kinase C in vascular myocytes. J Physiol 586: 2463–2476, 2008. [Abstract] [Google Scholar]
295. Saliez J, Bouzin C, Rath G, Ghisdal P, Desjardins F, Rezzani R, Rodella LF, Vriens J, Nilius B, Feron O, Balligand JL, Dessy C. Role of caveolar compartmentation in endothelium-derived hyperpolarizing factor-mediated relaxation: Ca2+ signals and gap junction function are regulated by caveolin in endothelial cells. Circulation 117: 1065–1074, 2008. [Abstract] [Google Scholar]
296. Salomonsson M, Braunstein TH, Holstein-Rathlou NH, Jensen LJ. Na+-independent, nifedipine-resistant rat afferent arteriolar Ca2+ responses to noradrenaline: possible role of TRPC channels. Acta Physiol 200: 265–278, 2010. [Abstract] [Google Scholar]
297. Samapati R, Yang Y, Yin J, Stoerger C, Arenz C, Dietrich A, Gudermann T, Adam D, Wu S, Freichel M, Flockerzi V, Uhlig S, Kuebler WM. Lung endothelial Ca2+ and permeability response to platelet-activating factor is mediated by acid sphingomyelinase and transient receptor potential classical 6. Am J Respir Crit Care Med 185: 160–170, 2012. [Abstract] [Google Scholar]
298. Sandow SL, Neylon CB, Chen MX, Garland CJ. Spatial separation of endothelial small- and intermediate-conductance calcium-activated potassium channels [K(Ca)] and connexins: possible relationship to vasodilator function? J Anat 209: 689–698, 2006. [Abstract] [Google Scholar]
299. Sano Y, Inamura K, Miyake A, Mochizuki S, Yokoi H, Matsushime H, Furuichi K. Immunocyte Ca2+ influx system mediated by LTRPC2. Science 293: 1327–1330, 2001. [Abstract] [Google Scholar]
300. Schierling W, Troidl K, Apfelbeck H, Troidl C, Kasprzak PM, Schaper W, Schmitz-Rixen T. Cerebral arteriogenesis is enhanced by pharmacological as well as fluid-shear-stress activation of the Trpv4 calcium channel. Eur J Vasc Endovasc Surg 41: 589–596, 2011. [Abstract] [Google Scholar]
301. Schindl R, Fritsch R, Jardin I, Frischauf I, Kahr H, Muik M, Riedl MC, Groschner K, Romanin C. Canonical transient receptor potential (TRPC) 1 acts as a negative regulator for vanilloid TRPV6-mediated Ca2+ influx. J Biol Chem 287: 35612–35620, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
302. Schleifenbaum J, Kassmann M, Szijarto IA, Hercule HC, Tano JY, Weinert S, Heidenreich M, Pathan AR, Anistan YM, Alenina N, Rusch NJ, Bader M, Jentsch TJ, Gollasch M. Stretch-activation of angiotensin II type 1a receptors contributes to the myogenic response of mouse mesenteric and renal arteries. Circ Res 115: 263–272, 2014. [Abstract] [Google Scholar]
303. Schlingmann KP, Weber S, Peters M, Niemann Nejsum L, Vitzthum H, Klingel K, Kratz M, Haddad E, Ristoff E, Dinour D, Syrrou M, Nielsen S, Sassen M, Waldegger S, Seyberth HW, Konrad M. Hypomagnesemia with secondary hypocalcemia is caused by mutations in TRPM6, a new member of the TRPM gene family. Nature Genet 31: 166–170, 2002. [Abstract] [Google Scholar]
304. Schmidt K, Dubrovska G, Nielsen G, Fesus G, Uhrenholt TR, Hansen PB, Gudermann T, Dietrich A, Gollasch M, de Wit C, Kohler R. Amplification of EDHF-type vasodilatations in TRPC1-deficient mice. Br J Pharmacol 161: 1722–1733, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
305. Schmitz C, Perraud AL, Johnson CO, Inabe K, Smith MK, Penner R, Kurosaki T, Fleig A, Scharenberg AM. Regulation of vertebrate cellular Mg2+ homeostasis by TRPM7. Cell 114: 191–200, 2003. [Abstract] [Google Scholar]
306. Senadheera S, Bertrand PP, Grayson TH, Leader L, Tare M, Murphy TV, Sandow SL. Enhanced contractility in pregnancy is associated with augmented TRPC3, L-type, and T-type voltage-dependent calcium channel function in rat uterine radial artery. Am J Physiol Regul Integr Comp Physiol 305: R917–R926, 2013. [Abstract] [Google Scholar]
307. Senadheera S, Kim Y, Grayson TH, Toemoe S, Kochukov MY, Abramowitz J, Housley GD, Bertrand RL, Chadha PS, Bertrand PP, Murphy TV, Tare M, Birnbaumer L, Marrelli SP, Sandow SL. Transient receptor potential canonical type 3 channels facilitate endothelium-derived hyperpolarization-mediated resistance artery vasodilator activity. Cardiovasc Res 95: 439–447, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
308. Sharif-Naeini R, Folgering JH, Bichet D, Duprat F, Lauritzen I, Arhatte M, Jodar M, Dedman A, Chatelain FC, Schulte U, Retailleau K, Loufrani L, Patel A, Sachs F, Delmas P, Peters DJ, Honore E. Polycystin-1 and -2 dosage regulates pressure sensing. Cell 139: 587–596, 2009. [Abstract] [Google Scholar]
309. Shi J, Birnbaumer L, Large WA, Albert AP. Myristoylated alanine-rich C kinase substrate coordinates native TRPC1 channel activation by phosphatidylinositol 4,5-bisphosphate and protein kinase C in vascular smooth muscle. FASEB J 28: 244–255, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
310. Shi J, Ju M, Abramowitz J, Large WA, Birnbaumer L, Albert AP. TRPC1 proteins confer PKC and phosphoinositol activation on native heteromeric TRPC1/C5 channels in vascular smooth muscle: comparative study of wild-type and TRPC1−/− mice. FASEB J 26: 409–419, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
311. Shi J, Ju M, Large W, Albert A. Pharmacological profile of phosphatidylinositol 3-kinases and related phosphatidylinositols mediating endothelin(A) receptor-operated native TRPC channels in rabbit coronary artery myocytes. Br J Pharmacol 166: 2161–2175, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
312. Shi J, Ju M, Saleh SN, Albert AP, Large WA. TRPC6 channels stimulated by angiotensin II are inhibited by TRPC1/C5 channel activity through a Ca2+- and PKC-dependent mechanism in native vascular myocytes. J Physiol 588: 3671–3682, 2010. [Abstract] [Google Scholar]
313. Shigematsu H, Sokabe T, Danev R, Tominaga M, Nagayama K. A 3.5-nm structure of rat TRPV4 cation channel revealed by Zernike phase-contrast cryoelectron microscopy. J Biol Chem 285: 11210–11218, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
314. Shigetomi E, Jackson-Weaver O, Huckstepp RT, O'Dell TJ, Khakh BS. TRPA1 channels are regulators of astrocyte basal calcium levels and long-term potentiation via constitutive d-serine release. J Neurosci 33: 10143–10153, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
315. Shigetomi E, Tong X, Kwan KY, Corey DP, Khakh BS. TRPA1 channels regulate astrocyte resting calcium and inhibitory synapse efficacy through GAT-3. Nature Neurosci 15: 70–80, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
316. Shinde AV, Motiani RK, Zhang X, Abdullaev IF, Adam AP, Gonzalez-Cobos JC, Zhang W, Matrougui K, Vincent PA, Trebak M. STIM1 controls endothelial barrier function independently of Orai1 and Ca2+ entry. Sci Signal 6: ra18, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
317. Simard C, Hof T, Keddache Z, Launay P, Guinamard R. The TRPM4 non-selective cation channel contributes to the mammalian atrial action potential. J Mol Cell Cardiol 59: 11–19, 2013. [Abstract] [Google Scholar]
318. Simard M, Arcuino G, Takano T, Liu QS, Nedergaard M. Signaling at the gliovascular interface. J Neurosci 23: 9254–9262, 2003. [Abstract] [Google Scholar]
319. Singer HA. Ca2+/calmodulin-dependent protein kinase II function in vascular remodelling. J Physiol 590: 1349–1356, 2012. [Abstract] [Google Scholar]
320. Singh I, Knezevic N, Ahmmed GU, Kini V, Malik AB, Mehta D. Galphaq-TRPC6-mediated Ca2+ entry induces RhoA activation and resultant endothelial cell shape change in response to thrombin. J Biol Chem 282: 7833–7843, 2007. [Abstract] [Google Scholar]
321. Smith AC, Hristov KL, Cheng Q, Xin W, Parajuli SP, Earley S, Malysz J, Petkov GV. Novel role for the transient potential receptor melastatin 4 channel in guinea pig detrusor smooth muscle physiology. Am J Physiol Cell Physiol 304: C467–C477, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
322. Smith AC, Parajuli SP, Hristov KL, Cheng Q, Soder RP, Afeli SA, Earley S, Xin W, Malysz J, Petkov GV. TRPM4 channel: a new player in urinary bladder smooth muscle function in rats. Am J Physiol Renal Physiol 304: F918–F929, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
323. Soboloff J, Spassova M, Xu W, He LP, Cuesta N, Gill DL. Role of endogenous TRPC6 channels in Ca2+ signal generation in A7r5 smooth muscle cells. J Biol Chem 280: 39786–39794, 2005. [Abstract] [Google Scholar]
324. 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 336: 597–601, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
325. Sonkusare SK, Dalsgaard T, Bonev AD, Hill-Eubanks DC, Kotlikoff MI, Scott JD, Santana LF, Nelson MT. AKAP150-dependent cooperative TRPV4 channel gating is central to endothelium-dependent vasodilation and is disrupted in hypertension. Sci Signal 7: ra66, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
326. Sontia B, Touyz RM. Role of magnesium in hypertension. Arch Biochem Biophys 458: 33–39, 2007. [Abstract] [Google Scholar]
327. Spassova MA, Hewavitharana T, Xu W, Soboloff J, Gill DL. A common mechanism underlies stretch activation and receptor activation of TRPC6 channels. Proc Natl Acad Sci USA 103: 16586–16591, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
328. Storch U, Forst AL, Philipp M, Gudermann T, Mederos y Schnitzler M. Transient receptor potential channel 1 (TRPC1) reduces calcium permeability in heteromeric channel complexes. J Biol Chem 287: 3530–3540, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
329. Story GM, Peier AM, Reeve AJ, Eid SR, Mosbacher J, Hricik TR, Earley TJ, Hergarden AC, Andersson DA, Hwang SW, McIntyre P, Jegla T, Bevan S, Patapoutian A. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell 112: 819–829, 2003. [Abstract] [Google Scholar]
330. Straub SV, Bonev AD, Wilkerson MK, Nelson MT. Dynamic inositol trisphosphate-mediated calcium signals within astrocytic endfeet underlie vasodilation of cerebral arterioles. J Gen Physiol 128: 659–669, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
331. Streifel KM, Gonzales AL, De Miranda B, Mouneimne R, Earley S, Tjalkens R. Dopaminergic neurotoxicants cause biphasic inhibition of purinergic calcium signaling in astrocytes. PloS One 9: e110996, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
332. Strotmann R, Harteneck C, Nunnenmacher K, Schultz G, Plant TD. OTRPC4, a nonselective cation channel that confers sensitivity to extracellular osmolarity. Nature Cell Biol 2: 695–702, 2000. [Abstract] [Google Scholar]
333. Strubing C, Krapivinsky G, Krapivinsky L, Clapham DE. Formation of novel TRPC channels by complex subunit interactions in embryonic brain. J Biol Chem 278: 39014–39019, 2003. [Abstract] [Google Scholar]
334. Strubing C, Krapivinsky G, Krapivinsky L, Clapham DE. TRPC1 and TRPC5 form a novel cation channel in mammalian brain. Neuron 29: 645–655, 2001. [Abstract] [Google Scholar]
335. Sudhahar V, Shaw S, Imig JD. Mechanisms involved in oleamide-induced vasorelaxation in rat mesenteric resistance arteries. Eur J Pharmacol 607: 143–150, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
336. Sullivan MN, Earley S. TRP channel Ca2+ sparklets: fundamental signals underlying endothelium-dependent hyperpolarization. Am J Physiol Cell Physiol 305: C999–C1008, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
337. Sullivan MN, Francis M, Pitts NL, Taylor MS, Earley S. Optical recording reveals novel properties of GSK1016790A-induced vanilloid transient receptor potential channel TRPV4 activity in primary human endothelial cells. Mol Pharmacol 82: 464–472, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
338. Sullivan MN, Gonzales AL, Pires PW, Bruhl A, Leo MD, Li W, Oulidi A, Boop FA, Feng Y, Jaggar JH, Welsh DG, Earley S. Localized TRPA1 channel Ca2+ signals stimulated by reactive oxygen species promote cerebral artery dilation. Sci Signal 8: ra2, 2015. [Europe PMC free article] [Abstract] [Google Scholar]
339. Sun HS, Jackson MF, Martin LJ, Jansen K, Teves L, Cui H, Kiyonaka S, Mori Y, Jones M, Forder JP, Golde TE, Orser BA, Macdonald JF, Tymianski M. Suppression of hippocampal TRPM7 protein prevents delayed neuronal death in brain ischemia. Nature Neurosci 12: 1300–1307, 2009. [Abstract] [Google Scholar]
340. Sun L, Yau HY, Wong WY, Li RA, Huang Y, Yao X. Role of TRPM2 in H2O2-induced cell apoptosis in endothelial cells. PloS One 7: e43186, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
341. Sundivakkam PC, Freichel M, Singh V, Yuan JP, Vogel SM, Flockerzi V, Malik AB, Tiruppathi C. The Ca2+ sensor stromal interaction molecule 1 (STIM1) is necessary and sufficient for the store-operated Ca2+ entry function of transient receptor potential canonical (TRPC) 1 and 4 channels in endothelial cells. Mol Pharmacol 81: 510–526, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
342. Suzuki M, Mizuno A, Kodaira K, Imai M. Impaired pressure sensation in mice lacking TRPV4. J Biol Chem 278: 22664–22668, 2003. [Abstract] [Google Scholar]
343. Szallasi A, Sheta M. Targeting TRPV1 for pain relief: limits, losers and laurels. Expert Opin Invest Drugs 21: 1351–1369, 2012. [Abstract] [Google Scholar]
344. Tai K, Hamaide MC, Debaix H, Gailly P, Wibo M, Morel N. Agonist-evoked calcium entry in vascular smooth muscle cells requires IP3 receptor-mediated activation of TRPC1. Eur J Pharmacol 583: 135–147, 2008. [Abstract] [Google Scholar]
345. Takahashi N, Kozai D, Kobayashi R, Ebert M, Mori Y. Roles of TRPM2 in oxidative stress. Cell Calcium 50: 279–287, 2011. [Abstract] [Google Scholar]
346. 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 Chem Biol 7: 701–711, 2011. [Abstract] [Google Scholar]
347. Takahashi Y, Watanabe H, Murakami M, Ohba T, Radovanovic M, Ono K, Iijima T, Ito H. Involvement of transient receptor potential canonical 1 (TRPC1) in angiotensin II-induced vascular smooth muscle cell hypertrophy. Atherosclerosis 195: 287–296, 2007. [Abstract] [Google Scholar]
348. Takahashi Y, Watanabe H, Murakami M, Ono K, Munehisa Y, Koyama T, Nobori K, Iijima T, Ito H. Functional role of stromal interaction molecule 1 (STIM1) in vascular smooth muscle cells. Biochem Biophys Res Commun 361: 934–940, 2007. [Abstract] [Google Scholar]
349. Thodeti CK, Matthews B, Ravi A, Mammoto A, Ghosh K, Bracha AL, Ingber DE. TRPV4 channels mediate cyclic strain-induced endothelial cell reorientation through integrin-to-integrin signaling. Circ Res 104: 1123–1130, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
350. Thorneloe KS, Cheung M, Bao W, Alsaid H, Lenhard S, Jian MY, Costell M, Maniscalco-Hauk K, Krawiec JA, Olzinski A, Gordon E, Lozinskaya I, Elefante L, Qin P, Matasic DS, James C, Tunstead J, Donovan B, Kallal L, Waszkiewicz A, Vaidya K, Davenport EA, Larkin J, Burgert M, Casillas LN, Marquis RW, Ye G, Eidam HS, Goodman KB, Toomey JR, Roethke TJ, Jucker BM, Schnackenberg CG, Townsley MI, Lepore JJ, Willette RN. An orally active TRPV4 channel blocker prevents and resolves pulmonary edema induced by heart failure. Science Translational Med 4: 159ra148, 2012. [Abstract] [Google Scholar]
351. Thorneloe KS, Sulpizio AC, Lin Z, Figueroa DJ, Clouse AK, McCafferty GP, Chendrimada TP, Lashinger ES, Gordon E, Evans L, Misajet BA, Demarini DJ, Nation JH, Casillas LN, Marquis RW, Votta BJ, Sheardown SA, Xu X, Brooks DP, Laping NJ, Westfall TD. N-((1S)-1-{[4-((2S)-2-{[(2,4-dichlorophenyl)sulfonyl]amino}-3-hydroxypropanoyl)-1 -piperazinyl]carbonyl}-3-methylbutyl)-1-benzothiophene-2-carboxamide (GSK1016790A), a novel and potent transient receptor potential vanilloid 4 channel agonist induces urinary bladder contraction and hyperactivity: Part I. J Pharmacol Exp Ther 326: 432–442, 2008. [Abstract] [Google Scholar]
352. Tiruppathi C, Freichel M, Vogel SM, Paria BC, Mehta D, Flockerzi V, Malik AB. Impairment of store-operated Ca2+ entry in TRPC4(−/−) mice interferes with increase in lung microvascular permeability. Circ Res 91: 70–76, 2002. [Abstract] [Google Scholar]
353. 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. EMBO J 25: 1804–1815, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
354. Torres VE, Cai Y, Chen X, Wu GQ, Geng L, Cleghorn KA, Johnson CM, Somlo S. Vascular expression of polycystin-2. J Am Soc Nephrol 12: 1–9, 2001. [Abstract] [Google Scholar]
355. Toth A, Czikora A, Pasztor ET, Dienes B, Bai P, Csernoch L, Rutkai I, Csato V, Manyine IS, Porszasz R, Edes I, Papp Z, Boczan J. Vanilloid receptor-1 (TRPV1) expression and function in the vasculature of the rat. J Histochem Cytochem 62: 129–144, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
356. Toth DM, Szoke E, Bolcskei K, Kvell K, Bender B, Bosze Z, Szolcsanyi J, Sandor Z. Nociception, neurogenic inflammation and thermoregulation in TRPV1 knockdown transgenic mice. Cell Mol Life Sci 68: 2589–2601, 2011. [Abstract] [Google Scholar]
357. Touyz RM. Transient receptor potential melastatin 6 and 7 channels, magnesium transport, and vascular biology: implications in hypertension. Am J Physiol Heart Circ Physiol 294: H1103–H1118, 2008. [Abstract] [Google Scholar]
358. Touyz RM, He Y, Montezano AC, Yao G, Chubanov V, Gudermann T, Callera GE. Differential regulation of transient receptor potential melastatin 6 and 7 cation channels by ANG II in vascular smooth muscle cells from spontaneously hypertensive rats. Am J Physiol Regul Integr Comp Physiol 290: R73–R78, 2006. [Abstract] [Google Scholar]
359. Trebak M. STIM1/Orai1, ICRAC, and endothelial SOC. Circ Res 104: e56–57, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
360. Trebak M, Hempel N, Wedel BJ, Smyth JT, Bird GS, Putney JW Jr. Negative regulation of TRPC3 channels by protein kinase C-mediated phosphorylation of serine 712. Mol Pharmacol 67: 558–563, 2005. [Abstract] [Google Scholar]
361. 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 USA 104: 13519–13524, 2007. [Europe PMC free article] [Abstract] [Google Scholar]
362. Tsagogiorgas C, Wedel J, Hottenrott M, Schneider MO, Binzen U, Greffrath W, Treede RD, Theisinger B, Theisinger S, Waldherr R, Kramer BK, Thiel M, Schnuelle P, Yard BA, Hoeger S. N-octanoyl-dopamine is an agonist at the capsaicin receptor TRPV1 and mitigates ischemia-induced [corrected] acute kidney injury in rat. PloS One 7: e43525, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
363. Uchida K. 4-Hydroxy-2-nonenal: a product and mediator of oxidative stress. Prog Lipid Res 42: 318–343, 2003. [Abstract] [Google Scholar]
364. Ueda K, Tsuji F, Hirata T, Takaoka M, Matsumura Y. Preventive effect of TRPV1 agonists capsaicin and resiniferatoxin on ischemia/reperfusion-induced renal injury in rats. J Cardiovasc Pharmacol 51: 513–520, 2008. [Abstract] [Google Scholar]
365. Ueda K, Tsuji F, Hirata T, Ueda K, Murai M, Aono H, Takaoka M, Matsumura Y. Preventive effect of SA13353 [1-[2-(1-adamantyl)ethyl]-1-pentyl-3-[3-(4-pyridyl)propyl]urea], a novel transient receptor potential vanilloid 1 agonist, on ischemia/reperfusion-induced renal injury in rats. J Pharmacol Exp Ther 329: 202–209, 2009. [Abstract] [Google Scholar]
366. Vaca L, Sinkins WG, Hu Y, Kunze DL, Schilling WP. Activation of recombinant trp by thapsigargin in Sf9 insect cells. Am J Physiol Cell Physiol 267: C1501–C1505, 1994. [Abstract] [Google Scholar]
367. Venkatachalam K, Hofmann T, Montell C. Lysosomal localization of TRPML3 depends on TRPML2 and the mucolipidosis-associated protein TRPML1. J Biol Chem 281: 17517–17527, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
368. Vennekens R, Hoenderop JG, Prenen J, Stuiver M, Willems PH, Droogmans G, Nilius B, Bindels RJ. Permeation and gating properties of the novel epithelial Ca2+ channel. J Biol Chem 275: 3963–3969, 2000. [Abstract] [Google Scholar]
369. Vennekens R, Olausson J, Meissner M, Bloch W, Mathar I, Philipp SE, Schmitz F, Weissgerber P, Nilius B, Flockerzi V, Freichel M. Increased IgE-dependent mast cell activation and anaphylactic responses in mice lacking the calcium-activated nonselective cation channel TRPM4. Nature Immunol 8: 312–320, 2007. [Abstract] [Google Scholar]
370. Vig M, Peinelt C, Beck A, Koomoa DL, Rabah D, Koblan-Huberson M, Kraft S, Turner H, Fleig A, Penner R, Kinet JP. CRACM1 is a plasma membrane protein essential for store-operated Ca2+ entry. Science 312: 1220–1223, 2006. [Abstract] [Google Scholar]
371. Vincent F, Acevedo A, Nguyen MT, Dourado M, DeFalco J, Gustafson A, Spiro P, Emerling DE, Kelly MG, Duncton MA. Identification and characterization of novel TRPV4 modulators. Biochem Biophys Res Commun 389: 490–494, 2009. [Abstract] [Google Scholar]
372. Vriens J, Owsianik G, Fisslthaler B, Suzuki M, Janssens A, Voets T, Morisseau C, Hammock BD, Fleming I, Busse R, Nilius B. Modulation of the Ca2+ permeable cation channel TRPV4 by cytochrome P 450 epoxygenases in vascular endothelium. Circ Res 97: 908–915, 2005. [Abstract] [Google Scholar]
373. Walker RL, Hume JR, Horowitz B. Differential expression and alternative splicing of TRP channel genes in smooth muscles. Am J Physiol Cell Physiol 280: C1184–C1192, 2001. [Abstract] [Google Scholar]
374. Wang J, Weigand L, Lu W, Sylvester JT, Semenza GL, Shimoda LA. Hypoxia inducible factor 1 mediates hypoxia-induced TRPC expression and elevated intracellular Ca2+ in pulmonary arterial smooth muscle cells. Circ Res 98: 1528–1537, 2006. [Abstract] [Google Scholar]
375. Wang L, Wang DH. TRPV1 gene knockout impairs postischemic recovery in isolated perfused heart in mice. Circulation 112: 3617–3623, 2005. [Abstract] [Google Scholar]
376. Wang LH, Luo M, Wang Y, Galligan JJ, Wang DH. Impaired vasodilation in response to perivascular nerve stimulation in mesenteric arteries of TRPV1-null mutant mice. J Hypertens 24: 2399–2408, 2006. [Abstract] [Google Scholar]
377. Wang P, Liu D, Tepel M, Zhu Z. Transient receptor potential canonical type 3 channels–their evolving role in hypertension and its related complications. J Cardiovasc Pharmacol 61: 455–460, 2013. [Abstract] [Google Scholar]
378. Wang Y, Kaminski NE, Wang DH. VR1-mediated depressor effects during high-salt intake: role of anandamide. Hypertension 46: 986–991, 2005. [Abstract] [Google Scholar]
379. Wang Y, Wang DH. A novel mechanism contributing to development of Dahl salt-sensitive hypertension: role of the transient receptor potential vanilloid type 1. Hypertension 47: 609–614, 2006. [Abstract] [Google Scholar]
380. Wang YX, Wang J, Wang C, Liu J, Shi LP, Xu M, Wang C. Functional expression of transient receptor potential vanilloid-related channels in chronically hypoxic human pulmonary arterial smooth muscle cells. J Membr Biol 223: 151–159, 2008. [Abstract] [Google Scholar]
381. 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 277: 13569–13577, 2002. [Abstract] [Google Scholar]
382. Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature 424: 434–438, 2003. [Abstract] [Google Scholar]
383. 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 277: 47044–47051, 2002. [Abstract] [Google Scholar]
384. Wehage E, Eisfeld J, Heiner I, Jungling E, Zitt C, Luckhoff A. Activation of the cation channel long transient receptor potential channel 2 (LTRPC2) by hydrogen peroxide. A splice variant reveals a mode of activation independent of ADP-ribose. J Biol Chem 277: 23150–23156, 2002. [Abstract] [Google Scholar]
385. Weissmann N, Dietrich A, Fuchs B, Kalwa H, Ay M, Dumitrascu R, Olschewski A, Storch U, Mederos y Schnitzler M, Ghofrani HA, Schermuly RT, Pinkenburg O, Seeger W, Grimminger F, Gudermann T. Classical transient receptor potential channel 6 (TRPC6) is essential for hypoxic pulmonary vasoconstriction and alveolar gas exchange. Proc Natl Acad Sci USA 103: 19093–19098, 2006. [Europe PMC free article] [Abstract] [Google Scholar]
386. Weissmann N, Sydykov A, Kalwa H, Storch U, Fuchs B, Mederos y Schnitzler M, Brandes RP, Grimminger F, Meissner M, Freichel M, Offermanns S, Veit F, Pak O, Krause KH, Schermuly RT, Brewer AC, Schmidt HH, Seeger W, Shah AM, Gudermann T, Ghofrani HA, Dietrich A. Activation of TRPC6 channels is essential for lung ischaemia-reperfusion induced oedema in mice. Nature Commun 3: 649, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
387. Welsh DG, Morielli AD, Nelson MT, Brayden JE. Transient receptor potential channels regulate myogenic tone of resistance arteries. Circ Res 90: 248–250, 2002. [Abstract] [Google Scholar]
388. Wes PD, Chevesich J, Jeromin A, Rosenberg C, Stetten G, Montell C. TRPC1, a human homolog of a Drosophila store-operated channel. Proc Natl Acad Sci USA 92: 9652–9656, 1995. [Europe PMC free article] [Abstract] [Google Scholar]
389. Willette RN, Bao W, Nerurkar S, Yue TL, Doe CP, Stankus G, Turner GH, Ju H, Thomas H, Fishman CE, Sulpizio A, Behm DJ, Hoffman S, Lin Z, Lozinskaya I, Casillas LN, Lin M, Trout RE, Votta BJ, Thorneloe K, Lashinger ES, Figueroa DJ, Marquis R, Xu X. Systemic activation of the transient receptor potential vanilloid subtype 4 channel causes endothelial failure and circulatory collapse: Part 2. J Pharmacol Exp Ther 326: 443–452, 2008. [Abstract] [Google Scholar]
390. Wilson PD. Polycystic kidney disease: new understanding in the pathogenesis. Int J Biochem Cell Biol 36: 1868–1873, 2004. [Abstract] [Google Scholar]
391. Winn MP, Conlon PJ, Lynn KL, Farrington MK, Creazzo T, Hawkins AF, Daskalakis N, Kwan SY, Ebersviller S, Burchette JL, Pericak-Vance MA, Howell DN, Vance JM, Rosenberg PB. A mutation in the TRPC6 cation channel causes familial focal segmental glomerulosclerosis. Science 308: 1801–1804, 2005. [Abstract] [Google Scholar]
392. Wissenbach U, Bodding M, Freichel M, Flockerzi V. Trp12, a novel Trp related protein from kidney. FEBS Lett 485: 127–134, 2000. [Abstract] [Google Scholar]
393. Wissenbach U, Philipp SE, Gross SA, Cavalie A, Flockerzi V. Primary structure, chromosomal localization and expression in immune cells of the murine ORAI and STIM genes. Cell Calcium 42: 439–446, 2007. [Abstract] [Google Scholar]
394. Wong F, Schaefer EL, Roop BC, LaMendola JN, Johnson-Seaton D, Shao D. Proper function of the Drosophila trp gene product during pupal development is important for normal visual transduction in the adult. Neuron 3: 81–94, 1989. [Abstract] [Google Scholar]
395. Woudenberg-Vrenken TE, Bindels RJ, Hoenderop JG. The role of transient receptor potential channels in kidney disease. Nature Rev Nephrol 5: 441–449, 2009. [Abstract] [Google Scholar]
396. Wu S, Cioffi EA, Alvarez D, Sayner SL, Chen H, Cioffi DL, King J, Creighton JR, Townsley M, Goodman SR, Stevens T. Essential role of a Ca2+-selective, store-operated current (ISOC) in endothelial cell permeability: determinants of the vascular leak site. Circ Res 96: 856–863, 2005. [Abstract] [Google Scholar]
397. Wu S, Jian MY, Xu YC, Zhou C, Al-Mehdi AB, Liedtke W, Shin HS, Townsley MI. Ca2+ entry via alpha1G and TRPV4 channels differentially regulates surface expression of P-selectin and barrier integrity in pulmonary capillary endothelium. Am J Physiol Lung Cell Mol Physiol 297: L650–L657, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
398. Xi Q, Adebiyi A, Zhao G, Chapman KE, Waters CM, Hassid A, Jaggar JH. IP3 constricts cerebral arteries via IP3 receptor-mediated TRPC3 channel activation and independently of sarcoplasmic reticulum Ca2+ release. Circ Res 102: 1118–1126, 2008. [Europe PMC free article] [Abstract] [Google Scholar]
399. Xia Y, Yang XR, Fu Z, Paudel O, Abramowitz J, Birnbaumer L, Sham JS. Classical transient receptor potential 1 and 6 contribute to hypoxic pulmonary hypertension through differential regulation of pulmonary vascular functions. Hypertension 63: 173–180, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
400. Xu H, Delling M, Jun JC, Clapham DE. Oregano, thyme and clove-derived flavors and skin sensitizers activate specific TRP channels. Nature Neurosci 9: 628–635, 2006. [Abstract] [Google Scholar]
401. Xu H, Ramsey IS, Kotecha SA, Moran MM, Chong JA, Lawson D, Ge P, Lilly J, Silos-Santiago I, Xie Y, DiStefano PS, Curtis R, Clapham DE. TRPV3 is a calcium-permeable temperature-sensitive cation channel. Nature 418: 181–186, 2002. [Abstract] [Google Scholar]
402. Xu SZ, Beech DJ. TrpC1 is a membrane-spanning subunit of store-operated Ca2+ channels in native vascular smooth muscle cells. Circ Res 88: 84–87, 2001. [Abstract] [Google Scholar]
403. Xu SZ, Boulay G, Flemming R, Beech DJ. E3-targeted anti-TRPC5 antibody inhibits store-operated calcium entry in freshly isolated pial arterioles. Am J Physiol Heart Circ Physiol 291: H2653–H2659, 2006. [Abstract] [Google Scholar]
404. Xu X, Wang P, Zhao Z, Cao T, He H, Luo Z, Zhong J, Gao F, Zhu Z, Li L, Yan Z, Chen J, Ni Y, Liu D. Activation of transient receptor potential vanilloid 1 by dietary capsaicin delays the onset of stroke in stroke-prone spontaneously hypertensive rats. Stroke 42: 3245–3251, 2011. [Abstract] [Google Scholar]
405. Yanaga A, Goto H, Nakagawa T, Hikiami H, Shibahara N, Shimada Y. Cinnamaldehyde induces endothelium-dependent and -independent vasorelaxant action on isolated rat aorta. Biol Pharm Bull 29: 2415–2418, 2006. [Abstract] [Google Scholar]
406. 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 Metab 12: 130–141, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
407. Yang XR, Lin AH, Hughes JM, Flavahan NA, Cao YN, Liedtke W, Sham JS. Upregulation of osmo-mechanosensitive TRPV4 channel facilitates chronic hypoxia-induced myogenic tone and pulmonary hypertension. Am J Physiol Lung Cell Mol Physiol 302: L555–L568, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
408. Yang XR, Lin MJ, McIntosh LS, Sham JS. Functional expression of transient receptor potential melastatin- and vanilloid-related channels in pulmonary arterial and aortic smooth muscle. Am J Physiol Lung Cell Mol Physiol 290: L1267–L1276, 2006. [Abstract] [Google Scholar]
409. Yildirim E, Dietrich A, Birnbaumer L. The mouse C-type transient receptor potential 2 (TRPC2) channel: alternative splicing and calmodulin binding to its N terminus. Proc Natl Acad Sci USA 100: 2220–2225, 2003. [Europe PMC free article] [Abstract] [Google Scholar]
410. Yin J, Hoffmann J, Kaestle SM, Neye N, Wang L, Baeurle J, Liedtke W, Wu S, Kuppe H, Pries AR, Kuebler WM. Negative-feedback loop attenuates hydrostatic lung edema via a cGMP-dependent regulation of transient receptor potential vanilloid 4. Circ Res 102: 966–974, 2008. [Abstract] [Google Scholar]
411. Yu PC, Gu SY, Bu JW, Du JL. TRPC1 is essential for in vivo angiogenesis in zebrafish. Circ Res 106: 1221–1232, 2010. [Abstract] [Google Scholar]
412. Yu SQ, Wang DH. Intrathecal injection of TRPV1 shRNA leads to increases in blood pressure in rats. Acta Physiol 203: 139–147, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
413. Yu Y, Fantozzi I, Remillard CV, Landsberg JW, Kunichika N, Platoshyn O, Tigno DD, Thistlethwaite PA, Rubin LJ, Yuan JX. Enhanced expression of transient receptor potential channels in idiopathic pulmonary arterial hypertension. Proc Natl Acad Sci USA 101: 13861–13866, 2004. [Europe PMC free article] [Abstract] [Google Scholar]
414. Yu Y, Keller SH, Remillard CV, Safrina O, Nicholson A, Zhang SL, Jiang W, Vangala N, Landsberg JW, Wang JY, Thistlethwaite PA, Channick RN, Robbins IM, Loyd JE, Ghofrani HA, Grimminger F, Schermuly RT, Cahalan MD, Rubin LJ, Yuan JX. A functional single-nucleotide polymorphism in the TRPC6 gene promoter associated with idiopathic pulmonary arterial hypertension. Circulation 119: 2313–2322, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
415. Yue L, Peng JB, Hediger MA, Clapham DE. CaT1 manifests the pore properties of the calcium-release-activated calcium channel. Nature 410: 705–709, 2001. [Abstract] [Google Scholar]
416. Zhang DX, Gutterman DD. Transient receptor potential channel activation and endothelium-dependent dilation in the systemic circulation. J Cardiovasc Pharmacol 57: 133–139, 2011. [Europe PMC free article] [Abstract] [Google Scholar]
417. Zhang DX, Mendoza SA, Bubolz AH, Mizuno A, Ge ZD, Li R, Warltier DC, Suzuki M, Gutterman DD. Transient receptor potential vanilloid type 4-deficient mice exhibit impaired endothelium-dependent relaxation induced by acetylcholine in vitro and in vivo. Hypertension 53: 532–538, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
418. Zhang L, Papadopoulos P, Hamel E. Endothelial TRPV4 channels mediate dilation of cerebral arteries: impairment and recovery in cerebrovascular pathologies related to Alzheimer's disease. Br J Pharmacol 170: 661–670, 2013. [Europe PMC free article] [Abstract] [Google Scholar]
419. Zhang P, Luo Y, Chasan B, Gonzalez-Perrett S, Montalbetti N, Timpanaro GA, Cantero Mdel R, Ramos AJ, Goldmann WH, Zhou J, Cantiello HF. The multimeric structure of polycystin-2 (TRPP2): structural-functional correlates of homo- and hetero-multimers with TRPC1. Hum Mol Genet 18: 1238–1251, 2009. [Europe PMC free article] [Abstract] [Google Scholar]
420. Zhang S, Patel HH, Murray F, Remillard CV, Schach C, Thistlethwaite PA, Insel PA, Yuan JX. Pulmonary artery smooth muscle cells from normal subjects and IPAH patients show divergent cAMP-mediated effects on TRPC expression and capacitative Ca2+ entry. Am J Physiol Lung Cell Mol Physiol 292: L1202–L1210, 2007. [Abstract] [Google Scholar]
421. Zhang S, Remillard CV, Fantozzi I, Yuan JX. ATP-induced mitogenesis is mediated by cyclic AMP response element-binding protein-enhanced TRPC4 expression and activity in human pulmonary artery smooth muscle cells. Am J Physiol Cell Physiol 287: C1192–C1201, 2004. [Abstract] [Google Scholar]
422. Zhang Z, Okawa H, Wang Y, Liman ER. Phosphatidylinositol 4,5-bisphosphate rescues TRPM4 channels from desensitization. J Biol Chem 280: 39185–39192, 2005. [Abstract] [Google Scholar]
423. Zhao L, Sullivan MN, Chase M, Gonzales AL, Earley S. Calcineurin/NFAT-coupled TRPV4 Ca sparklets stimulate airway smooth muscle cell proliferation. Am J Respir Cell Mol Biol 50: 1064–1075, 2014. [Europe PMC free article] [Abstract] [Google Scholar]
424. Zhao Z, Ni Y, Chen J, Zhong J, Yu H, Xu X, He H, Yan Z, Scholze A, Liu D, Zhu Z, Tepel M. Increased migration of monocytes in essential hypertension is associated with increased transient receptor potential channel canonical type 3 channels. PloS One 7: e32628, 2012. [Europe PMC free article] [Abstract] [Google Scholar]
425. Zhu X, Chu PB, Peyton M, Birnbaumer L. Molecular cloning of a widely expressed human homologue for the Drosophila trp gene. FEBS Lett 373: 193–198, 1995. [Abstract] [Google Scholar]
426. Zhu X, Jiang M, Peyton M, Boulay G, Hurst R, Stefani E, Birnbaumer L. trp, a novel mammalian gene family essential for agonist-activated capacitative Ca2+ entry. Cell 85: 661–671, 1996. [Abstract] [Google Scholar]
427. Zitt C, Obukhov AG, Strubing C, Zobel A, Kalkbrenner F, Luckhoff A, Schultz G. Expression of TRPC3 in Chinese hamster ovary cells results in calcium-activated cation currents not related to store depletion. J Cell Biol 138: 1333–1341, 1997. [Europe PMC free article] [Abstract] [Google Scholar]
428. Zulian A, Baryshnikov SG, Linde CI, Hamlyn JM, Ferrari P, Golovina VA. Upregulation of Na+/Ca2+ exchanger and TRPC6 contributes to abnormal Ca2+ homeostasis in arterial smooth muscle cells from Milan hypertensive rats. Am J Physiol Heart Circ Physiol 299: H624–H633, 2010. [Europe PMC free article] [Abstract] [Google Scholar]
429. Zygmunt PM, Andersson DA, Hogestatt ED. Delta 9-tetrahydrocannabinol and cannabinol activate capsaicin-sensitive sensory nerves via a CB1 and CB2 cannabinoid receptor-independent mechanism. J Neurosci 22: 4720–4727, 2002. [Abstract] [Google Scholar]
430. Zygmunt PM, Petersson J, Andersson DA, Chuang H, Sorgard M, Di Marzo V, Julius D, Hogestatt ED. Vanilloid receptors on sensory nerves mediate the vasodilator action of anandamide. Nature 400: 452–457, 1999. [Abstract] [Google Scholar]

Articles from Physiological Reviews are provided here courtesy of American Physiological Society

Citations & impact 


Impact metrics

Jump to Citations

Citations of article over time

Alternative metrics

Altmetric item for https://www.altmetric.com/details/4032497
Altmetric
Discover the attention surrounding your research
https://www.altmetric.com/details/4032497

Smart citations by scite.ai
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by EuropePMC if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
Explore citation contexts and check if this article has been supported or disputed.
https://scite.ai/reports/10.1152/physrev.00026.2014

Supporting
Mentioning
Contrasting
7
374
0

Article citations


Go to all (221) article citations

Funding 


Funders who supported this work.

NHLBI NIH HHS (3)