Review Open Access
Copyright ©2012 Baishideng Publishing Group Co., Limited. All rights reserved.
World J Biol Chem. Jul 26, 2012; 3(7): 127-158
Published online Jul 26, 2012. doi: 10.4331/wjbc.v3.i7.127
Update on vascular endothelial Ca2+ signalling: A tale of ion channels, pumps and transporters
Francesco Moccia, Franco Tanzi, Department of Biology and Biotechnologies “Lazzaro Spallanzani”, Laboratory of Physiology, University of Pavia, Via Forlanini 6, 27100 Pavia, Italy
Roberto Berra-Romani, Department of Biomedicine, School of Medicine, Benemérita Universidad Autónoma de Puebla, 3 Sur 2702, Colonia Volcanes Puebla 72410, México
Author contributions: Moccia F wrote the manuscript; Berra-Romani R revised the manuscript, Tanzi F wrote the manuscript.
Correspondence to: Dr. Francesco Moccia, Department of Biology and Biotechnologies “Lazzaro Spallanzani”, Laboratory of Physiology, University of Pavia, Via Forlanini 6, 27100 Pavia, Italy. francesco.moccia@unipv.it
Telephone: +39-382-987169 Fax: +39-382-987527
Received: May 17, 2012
Revised: July 4, 2012
Accepted: July 11, 2012
Published online: July 26, 2012

Abstract

A monolayer of endothelial cells (ECs) lines the lumen of blood vessels and forms a multifunctional transducing organ that mediates a plethora of cardiovascular processes. The activation of ECs from as state of quiescence is, therefore, regarded among the early events leading to the onset and progression of potentially lethal diseases, such as hypertension, myocardial infarction, brain stroke, and tumor. Intracellular Ca2+ signals have long been know to play a central role in the complex network of signaling pathways regulating the endothelial functions. Notably, recent work has outlined how any change in the pattern of expression of endothelial channels, transporters and pumps involved in the modulation of intracellular Ca2+ levels may dramatically affect whole body homeostasis. Vascular ECs may react to both mechanical and chemical stimuli by generating a variety of intracellular Ca2+ signals, ranging from brief, localized Ca2+ pulses to prolonged Ca2+ oscillations engulfing the whole cytoplasm. The well-defined spatiotemporal profile of the subcellular Ca2+ signals elicited in ECs by specific extracellular inputs depends on the interaction between Ca2+ releasing channels, which are located both on the plasma membrane and in a number of intracellular organelles, and Ca2+ removing systems. The present article aims to summarize both the past and recent literature in the field to provide a clear-cut picture of our current knowledge on the molecular nature and the role played by the components of the Ca2+ machinery in vascular ECs under both physiological and pathological conditions.

Key Words: Endothelial cells, Ca2+ signalling, Plasma membrane, Endoplasmic reticulum, Intracellular Ca2+ release, Ca2+ entry, Ca2+ removal, Ca2+ oscillations



INTRODUCTION

It has long been known that an increase in intracellular Ca2+ concentration ([Ca2+]i) plays a key role in the intricate network of signal transduction pathways exploited by endothelial cells (ECs) to maintain cardiovascular homeostasis[1]. Due to its strategic location at the interface between the vascular wall and bloodstream, the endothelium is exposed to a myriad of transmitters (released by autonomic and sensory nerves or platelets), circulating hormones, autacoids, cytokines, growth factors, and drugs, as well as to mechanical stimuli, such as pulsatile stretch, shear stress, and changes in the local osmotic pressure[2,3]. Moreover, vascular ECs might serve as thermosensors and modulate peripheral vasoconstriction or vasodilation depending on the environmental temperature[4]. A remarkable blend of membrane receptors, transporters, and ion channels, which are located both on the plasma membrane and within the intracellular Ca2+ reservoirs (namely, endoplasmic reticulum, lysosomes, and Golgi), may be utilized by ECs to detect and react selectively to the incoming stimulus with the most suitable Ca2+ signal[1,4-8]. A number of excellent reviews have recently outlined features and activation mechanisms of Ca2+-permeable channels in vascular endothelium[1,7,8]. The present article will survey the Ca2+ signaling toolkit whereby ECs shape their Ca2+ responses to chemical and mechanical stimuli by summarizing the most recent developments in the field. Recent work has shown that endothelial progenitor cells (EPCs), a subpopulation of mononuclear cells that are mobilized from bone marrow upon an ischemic insult, impinge on Ca2+ signals to proliferate, traffic to the injured tissue, and form capillary-like structures. However, we refer to a recent review from our laboratory for an exhaustive description of our current knowledge about the functional role and the molecular underpinnings of Ca2+ signalling in human EPCs[9].

Ca2+ SIGNALLING IN VASCULAR ECs

In resting, i.e., non-stimulated cells, [Ca2+]i is very low (around 100 nmol/L) due to the activity of various active mechanisms that remove Ca2+ out of the cytosol[10]. More specifically, the plasma membrane Ca2+-ATPase (PMCA) and the sarco-endoplasmic reticulum Ca2+-ATPase (SERCA) extrude Ca2+ by direct ATP hydrolysis, while the Na+/Ca2+ exchanger (NCX) clears cytosolic Ca2+ by exploiting the Na+ gradient across the plasma membrane[10]. An increase in [Ca2+]i up to 1 μmol/L is the key signal to activate vascular ECs following recruitment of either tyrosine-kinase linked receptors (TRKs) or G-protein coupled receptors (GPCRs) by growth factor and vasoactive agents, respectively[1,6,7,9,11,12]. The endothelial Ca2+ response may comprise an initial Ca2+ spike, shaped by Ca2+ mobilization from the intracellular Ca2+ reservoir, followed by an intermediate plateau level due to Ca2+ entry across the plasma membrane[6,7,9]. However, at low-dose agonist stimulation, the Ca2+ signal may adopt an oscillatory pattern driven by the interplay between intracellular Ca2+ release and Ca2+ influx from the extracellular space[1,6,13]. Responding cells with an intermediate behavior between the extremes described above can also be found. Accordingly, in most microvascular ECs, chemical stimulation does not elicit a remarkable Ca2+ inflow and results in either a transient Ca2+ spike or localized, high frequency Ca2+ oscillations, which are attributable to Ca2+ mobilization from the intracellular store[14-16]. Conversely, when the Ca2+ signal is uniquely modelled by Ca2+ entry, a sustained and monotonic increase in [Ca2+]i occurs[17,18]. Mechanical stimulation may also affect vascular endothelium by causing an elevation in [Ca2+]i. Indeed, exposure of ECs to fluid shear stress, pulsatile stretch, and cell swelling may evoke both Ca2+ discharge from intracellular Ca2+ pools and Ca2+ inflow through mechanosensitive channels on the plasmalemma[4,7,19]. Recent studies have argued that ion channels serve as direct sensor of mechanical stimuli, but rather as a relay that is engaged by a yet to identify upstream receptor[19]. Unlike the Ca2+ response to extracellular ligands, mechanosensitive Ca2+ signals normally lack the biphasic kinetics and exhibit either a transient[20,21] or an oscillatory pattern[21-25]. In a few cases, a dose-dependent stepwise elevation in [Ca2+]i was observed in cultured ECs exposed to laminar shear stress[24,26,27]. The contact between air microbubbles and cultured ECs may also result in an intracellular Ca2+ wave[28]. Finally, an increase in [Ca2+]i arises in vascular ECs upon contact with circulating leukocytes[29] or metastatizing tumor cells[30,31].

Ca2+ RELEASE FROM INTRACELLULAR STORES

The endoplasmic reticulum (ER) houses the largest intracellular Ca2+ reservoir in ECs, amounting to approximately 75% of the total cellular storage capacity[32]. A recent ultrastructural analysis has revealed that the endothelial ER consists of tubulovesicular structures that are interconnected at various sites to form a widespread network within the ECs. In the periphery of the cells, they lie immediately adjacent to the plasma membrane and the caveolae at the apical, lateral, and basal surfaces of the cells and occasionally run closely parallel to the plasma membrane itself[33-35]. Intriguingly, the ER-plasmalemma coupling is tighter (around 8 nm) in macrovascular than in microvascular (around 87 nm) ECs[33]. As Ca2+ entry may be governed by the physical interaction between an ER Ca2+-sensor and the plasmalemmal Ca2+-permeable channels (see below), this feature might help explain why microvascular ECs normally lack the plateau phase[33]. In the middle of the cells, the tubulovesicular structures may be in close contact with the mitochondria, sometimes enveloping them[34-36], and are in continuity with the nuclear membrane[34]. Ca2+ is sequestered into the ER lumen, where it attains a concentration of about 100-500 μmol/L, by high-affinity (Km around 1 μmol/L) SERCA pumps, which can be inhibited by pharmacological agents, such as the marine toxin thapsigargin and cyclopiazonic acid (CPA)[37,38]. Inhibition of the SERCA pump with either drug leads to Ca2+ store depletion, which indicates that ER membrane is constitutively leaky to Ca2+. Candidates to mediate the leak channel in a variety of cell types include polycystin-2, presenilins, and the ribosome-translocon complex[37]. No study has, however, hitherto addressed the molecular underpinnings of passive Ca2+ leak from ER in ECs[1]. Calreticulin represents the major Ca2+ buffer within the ER lumen in all non-muscle cells[38], as well as in vascular endothelium[34,35]. Maintenance of a constant luminal level of Ca2+ in ECs is also fulfilled by a number of additional Ca2+-binding proteins, including calnexin, glucose-regulated protein (GRP)-78, and GRP-94[38]. All these Ca2+ buffers also serve as molecular chaperones and their protein folding activity is regulated by ER Ca2+ concentration ([Ca2+]ER)[38]. Monitoring of intraluminal Ca2+ levels in ECs transiently transfected with a cameleon-based Ca2+ probe targeted to the ER, D1ER, has shown that, in the presence of extracellular Ca2+, extracellular autacoids do not induce a massive decline in [Ca2+]ER[39-42]. Consequently, Ca2+-dependent protein folding is largely maintained during physiological cell stimulation and does not activate any stress signaling pathway[38,42]. Conversely, a rapid drop in ER Ca2+ content occurs when ECs are stimulated in the absence of extracellular Ca2+ entry[43]. Owing to the large electrochemical gradient existing between ER lumen and the cytosol, rapid release of Ca2+ from the intracellular organelle occurs upon the opening of two largely homologous receptor types on its membrane, which are named for their affinity for either the cell metabolite inositol-1,4,5-trisphosphate (InsP3) or the plant alkaloid ryanodine[1,6,11]. Ryanodine receptors (RyRs), in particular, may be engaged by either a local increase in intracellular Ca2+ levels trough a process called Ca2+-induced Ca2+ release (CICR) or the second messenger, cyclic ADPribose (cADPr)[44-48]. The remaining 25% of the total stored Ca2+ in ECs is loaded in non-ER intracellular pools, mainly mitochondria, which contribute to agonists-induced intracellular Ca2+ signaling[32,36,40,41]. Ca2+ enters the organelles through a uniporter, present in the inner membrane, and can be exported via the stimulation of a mitochondrial NCX (NCXmito)[36,41], thus creating a Ca2+ cycling system for the control of mitochondrial Ca2+ concentration that, under resting conditions, is equal to around 200 nmol/L in ECs[49]. Mitochondria give rise to an impressive network in ECs, which is in tight apposition with either Ca2+ release channels on the ER or Ca2+-permable pathways, mainly store-operated Ca2+ channels (SOCs), on the plasma membrane[34-36]. Acidic, lysosome-like vesicles may also serve as Ca2+ storage organelles in vascular endothelium[50,51]. Unlike the large network that is the ER, lysosomes constitute relatively small, discrete and mobile vesicles with a limited Ca2+ content, which can be discharged by NAADP[50,51]. An attempt to quantify the relative lysosomal Ca2+ content suggested the latter to amount to around 79% of the ER Ca2+ pool[52]. Such a value, which is at odds with a previous analysis of the relative capacities of intracellular Ca2+ storage organelles[32], does not truly reflect lysosomal Ca2+ levels. Indeed, unlike the large network that is the ER, lysosomes generally constitute relatively small, discrete and mobile vesicles with a limited Ca2+ content[53]. However, such acidic stores may be situated very close to the Ca2+-releasing channels on ER. As a consequence, NAADP induces scattered, local Ca2+ signals that are capable of generating a global Ca2+ wave only by triggering further Ca2+ mobilization from InsP3Rs and/or RyRs via the CICR mechanism[53] (see below). Along with InsP3, cADPr, and NAADP, ECs may liberate intracellular Ca2+ in response to other second messengers, such as arachidonid acid (AA)[54], whose intracellular target remains to be elucidated, and reactive oxygen species (ROS), such as H2O2[55,56] and superoxide anion (O2.-), which directly evoke Ca2+ release from InsP3Rs[57]. As to the sphingolipid-derived messenger, sphingosine-1-phosphate (S1P), that may elicit intracellular Ca2+ release independently from InsP3Rs and RyRs[10], it has been shown that, when endogenously synthesized in autacoids-stimulated ECs, it modulates SOCs, but does not stimulate Ca2+ mobilization[58]. In blood vessels, however, plasma S1P may be produced and released by erythrocytes and platelets and induce Ca2+ mobilization in ECs upon binding to the Gi protein-coupled S1P1 receptor[59]. The latter, in turn, recruits phospholipase C (PLC) to generate InsP3 and, thus, stimulates InsP3Rs and activates SOCs[60,61].

Phospholipase C, InsP3, and InsP3-receptors

The intracellular second messenger InsP3 liberates Ca2+ ions stored within the ER and, in ECs, is synthesized following activation of either GPCRs or TRKs[1,6,11,62]. InsP3 is a highly diffusible hydrophilic messenger that is produced by the cleavage of phosphatidylinositol-4,5-bisphosphate (PIP2). This reaction is catalyzed by a family of PLC isozymes, which, depending on the subtype, can be activated by G-proteins (Gq/11), phosphorylation or Ca2+ itself[63,64]. The other product of PIP2 hydrolysis is diacylglycerol (DAG), which stays within the plasma membrane where it either activates protein kinase C (PKC) or gates members of the transient receptor potential (TRP) canonical (TRPC) subfamily of Ca2+-permeable channels[10]. The 13 mammalian PLC isozymes (excluding alternatively spliced forms) identified to date are all single polypeptides and can be divided into five types: PLCβ (β1-4), PLCδ (δ1-4), PLCγ (γ1 and γ2), PLCε, PLCζ, and PLCη (η1 and η2)[63,64]. The amino acid sequences of PLC isozymes are relatively nonconserved with the exception of two regions, known as the X and Y domains, that form the catalytic core and are associated to various combinations of regulatory domains that are common to many other signalling proteins. In more detail, in γ-type isozymes, this region contains two additional Src homology 2 (SH2) domains and one SH3 domain, which bind phosphorylated tyrosine residues and proline-rich sequences, respectively[63,64]. All PLC isoforms, with the exception of PLCε, are endowed with at least one potential polyphosphoinositides binding site, the pleckstrin homology (PH) domain, which is located in the NH2-terminal region preceding the X domain[63,64]. Moreover, the COOH-terminal region possesses a C2 domain which confers a strong Ca2+ sensitivity to the enzymatic activity of PLC. The C2 domain is, however, absent in the γ-type[63,64]. In this view, four EF domains are located between the PH and X domains of all PLC subclasses but their functional significance is yet to be fully elucidated[63,64]. PLCε, finally, exhibits a NH2-terminal Ras guanine nucleotide exchange factor (RasGEF)-like domain and at least one COOH-terminal Ras binding (RA) domains that are involved in its activation[63,64]. ECs express a variety of PLC isoforms. PLCβ1 is enlisted following GPCR binding to their specific agonists, which include a plethora of circulating autacoids: acetylcholine, ATP, ADP, thrombin, bradykinin, serotonin, angiotensin II, and anandamide[1,7,65]. Receptor occupancy results in the activation of the pertuxis toxin (PTX)-insensitive heterotrimeric Gq/11 protein family. As a consequence, the GTP-bound Gαq/11 subunit dissociates from the Gβγ dimer and activates the enzyme[1,7,63,64]. A remarkable exception is provided by the proteinase-activated receptor 1, which utilizes both Gαq[66,67] and Gαo[68] to trigger Ca2+ signals in thrombin-stimulated ECs[69], albeit only the former leads to InsP3 production[70]. The Gαo-dependent activation is mediated by the Gβγ dimer[68], which may recruit all PLC-β subtypes with the exception of the β4 isoform[63,64]. This mechanism is also involved in PLCβ engagement by the PTX-sensitive family of Gi proteins[63,64], as observed in ECs exposed to S1P[59-61]. PLCγ1 is the main γ-type expressed in ECs[9], with PLCγ2 being primarily restricted to cells of the hematopoietic lineage[71]. PLCγ1 is the isozyme mediating the increase in [Ca2+]i elicited in vascular endothelium by growth factors, such as epidermal growth factor (EGF)[72], vascular endothelial growth factor (VEGF)[73-75], platelet-derived growth factor (PDGF)[76], and basic fibroblast growth factor (bFGF)[73,77]. Upon growth factor binding to its cognate receptor, PLC-γ1 is recruited to the activated receptors via SH2-domain-mediated phosphotyrosine interaction and then subjected to phosphorylation by receptor tyrosine kinase[63,64]. It has, however, been reported that PLC-γ1 may be also stimulated by GPCR in ECs. For instance, it may associate with the COOH-terminal intracellular domain of the bradykinin BR receptor (B2), which results in a transient phosphorylation and InsP3 production[78,79]. On the other hand, VEGF has been recently shown to elicit PLCβ3 lipase activity upon binding to VEGFR-2 (KDR), which in turn physically interacts with Gq/11 proteins to induce Gβγ dissociation from Gα subunits[80]. Gβγ may thus stimulate PLCβ3 to generate InsP3 and trigger an intracellular Ca2+ wave[81]. PLCδ1 has been so far detected in a limited number of vascular beds, although its ability to evoke InsP3-dependent Ca2+ mobilization has been probed only in human umbilical vein ECs (HUVECs) transfected with PLCδ1-containing plasmids[82]. While there is no report of PLCη expression in ECs and PLCζ is present almost exclusively in mammal sperms[6], PLCε may promote angiogenesis in a mouse model of adenocarcinoma, but its role in Ca2+ signaling has not been explored[83]. It is, however, remarkable that β2-adrenergic excitation results to InsP3-dependent Ca2+ discharge upon recruitment of the exchange protein activated by cAMP-1 (Epac-1)[84], which has long been known to stimulate PLCε[64]. In agreement with PLC activation by both GPCRs and TKRs, InsP3 production has been evaluated in a variety of stimulated ECs[1]. In [3H]myoinositol-labeled cultured cells, both thrombin and histamine-induced increases in InsP3 occurred in less than 15 s, and were temporally correlated with the Ca2+ signals[70,85-87]. Bradykinin and ATP also evoked an elevation in InsP3 concentration, which attained a peak value at 15 s after agonist application and then decayed to a plateau level with kinetics similar to those of the ensuing Ca2+ signals[70,88,89]. The same methodology revealed that platelet activating factor (PAF)-induced InsP3 formation is maximal after 1 min of stimulation[90]. Similarly, a radioreceptor assay of InsP3 accumulation in ECs revealed that VEGF induced a rapid increase in InsP3 levels, which achieved a peak after 4 min of exposure[91]. Due to the objective difficulties entailed by these pioneering measurements, it is unlikely that they truly reflect the dynamics of agonists-induced InsP3 accumulation in vascular ECs. A more recent investigation, based on an innovative analytical method, revealed that InsP3 levels peak within < 1 s from ATP stimulation and then rapidly subside due to degradation to lower order inositol phosphates[92]. Besides agonist stimulation, InsP3 production in vascular ECs may be stimulated by reactive oxygen species (ROS), such as O2.-[93] and H2O2[94].

In addition to extracellular agonists, mechanical stimuli may induce InsP3 formation in vascular endothelium. Cyclic strain was found to elicit a transient spike of immunoreactive InsP3 in bovine aortic ECs (BAECs): peak levels occurred 10 s after the initiation of strain and decayed to control values 25 s later[95]. Moreover, BAEC exposure to shear stress induced a biphasic increase in intracellular InsP3 concentration, whereas a transient InsP3 peak arising within the first 15 s was followed by a major peak observed at 15 min after the onset of the stimulus. InsP3 levels returned to pre-shear stress level within the following 15 min[96]. Recent studies have demonstrated that the vascular system is endowed with mechanosensitive GPCRs translating mechanical stimuli into an InsP3-dependent Ca2+ mobilization[97]. Consistently, shear stress induces bradykinin B2 receptor to undergo a conformational transition that triggers the downstream signaling cascades, i.e., InsP3 and DAG production, even in the absence of the agonist[97,98]. The kinetics of InsP3 synthesis under these conditions, however, has not been determined.

Once produced by PLC isozymes, InsP3 diffuses into the cytosol and evokes a short-lasting Ca2+ efflux from ER[99]. InsP3-induced Ca2+ mobilization from ER has been characterized both in saponin-permeabilized ECs loaded with Ca2+ and intact ECs perfused with caged InsP3via a patch pipette[45,99-101]. The release of Ca2+ by this second messenger displayed an EC50 around 1 μmol/L and was independent of InsP3 metabolism[99,101]. InsP3 acts on a tetrameric receptor characterized by a very long cytosolic NH2-terminal tail that contains the InsP3-binding site, which has been crystallized recently[102]. Three InsP3 receptor subtypes have been identified that may assemble to form either homo- or hetero-tetrameric Ca2+-releasing channels. InsP3 binds to the very long cytosolic NH2-terminal tail, while the proper Ca2+ channel is located in the COOH-terminal portion of the molecule, which contains the sixth transmembrane™ domain motif and the loop connecting helices 5 and 6 folding back into the membrane[102]. The molecular mass of each InsP3R isoform is around 260 kDa[103,104], which indicates that the native protein is around 1 MDa[102]. The Hill coefficient for InsP3-elicited Ca2+ release is around 3.8 and suggests that InsP3 binding to all subunits is required to induce channel opening[99]. All InsP3R subtypes have been detected in ECs (Table 1), both under culture conditions[104-106] and in intact vessels[107,108]. The expression pattern of InsP3R isoforms in both freshly isolated and cultured ECs is: InsP3R-3 > InsP3R-2 > InsP3R-1[106]. This feature becomes even more remarkable in proliferating cells, which suggests that distinct isoforms may impact on different aspects of EC physiology[106]. In this view, recent work has revealed that InsP3Rs forms mostly, albeit not exclusively, homotetramers in vascular ECs[109]. The subcellular localization of InsP3R subtypes may vary with cell isolation and culture. In cultured ECs, InsP3R-1 is mainly localized to the ER, InsP3R-2 resides within the nucleus, and InsP3R-3 is detected at the perinuclear region[104], while, in the intact vascular wall, all isoforms are tightly associated to F-actin fibers and InsP3R-1 and InsP3R-2 lie along with InsP3R-3 near the edge of the nucleus[107]. None of the three subtypes has been detected in close association with the plasma membrane[104,107], although some studies have reported a protein with high homology to the InsP3R-1 in endothelial caveolae via binding to F-actin[34,110]. Such plasmalemmal InsP3 might mediate the InsP3-gated Ca2+ entry originally described by Vaca and Kunze[111]. Moreover, InsP3R-1 may be associated to the EC side of myoendothelial junctions in vivo and thus plays a key role in transmitting intracellular Ca2+ waves from vascular endothelium to the adjoining smooth muscle cell layer[108]. InsP3R expression may be altered under pathological conditions, such as hypertension, whereas ECs isolated from spontaneously hypertensive rats display an increase in InsP3R-2 transcripts and a decrease in InsP3R-1 mRNA[112]. Moreover, InsP3R levels augment as a consequence of ethanol metabolism and oxidative stress[113]. InsP3R gating is regulated by both InsP3 and Ca2+. InsP3 is able to activate InsP3Rs when the surrounding Ca2+ is between 50 nmol/L and 200 nmol/L, whereas higher Ca2+ levels inhibit Ca2+ release even at saturating InsP3 concentrations[102]. Such a biphasic dependence on ambient Ca2+ is due to two independent Ca2+-binding sites that mediate channel activation or closure, respectively. More specifically, InsP3 tunes InsP3R sensitivity to cytosolic Ca2+: it favors Ca2+ binding to the stimulatory site, which is the trigger for pore opening, and inhibits Ca2+ binding to the inhibitory one[102,114]. Consistent with this model, InsP3-induced Ca2+ mobilization is enhanced at lower (200 nmol/L) than at higher (2 μmol/L) Ca2+ concentrations in saponin-permeabilized BAECs[99,115]. The large modulatory domain between the InsP3-binding site and the channel-forming region presents consensus sequences for numerous protein kinases, including PKC, protein kinase A (PKA), cGMP-dependent protein kinase (PKG), Akt kinase, calmodulin-dependent protein kinase II (CaMKII), tyrosine kinases, and cyclin-dependent kinase 1/cyclin B (cdc2/CyB) complex[102]. Additionally, InsP3R channel activity is sensitive to a variety of modulators, including ATP, redox balance, and luminal Ca2+, and interacting proteins, such as calmodulin, Ca2+-binding proteins (CaBPs), chromogranins, InsP3R binding protein released with inositol-1,4,5-trisphosphate (IRBIT), RACK1 and Gβγ, ERp44, FKBP12, cytochrome C, Bcl-2 proteins, huntingtin, adapter proteins (ankyrin, homer, protein 4.1N), and proteases (caspase-3 and calpain)[102]. As to vascular ECs, PKC and CaMKII block InsP3-dependent Ca2+ release via ATP-dependent phosphorylation of InsP3Rs[116-118]. Conversely, PKA and mammalian target of rapamycin (mTOR) phosphorylate InsP3Rs to render them more sensitive to ligand binding and, thereby, increase the velocity of InsP3-dependent Ca2+ waves[118]. In addition, an InsP3R-ankyrin complex has been detected in cells exposed to hyaluronic acid (HA). Such stimulation favors the ankyrin-mediated recruitment of ER InsP3Rs into plasmalemmal caveolae; an event that triggers Ca2+ release independently on agonist binding[119]. InsP3R modulation in ECs may be further accomplished by the gasotransmitters, NO and H2S. While the former induces Ca2+ release likely via receptor S-nitrosylation[120,121], the latter inhibits InsP3-induced Ca2+ discharge probably through receptor S-sulfhydration[122,123]. An additional post-translational modification of endothelial InsP3Rs is provided by S-glutathionylation, which might sensitize the receptor to the basal level of InsP3[124]. In this view, the InsP3R may be directly regulated by ROS, which may oxidize the two highly conserved thiol groups located in the COOH-tail of the protein[102]. It has, indeed, been shown that Ca2+ release from endothelial InsP3Rs may be evoked by H2O2[55-57], which may increase the sensitivity of the intracellular Ca2+ pool to InsP3[56].

Table 1 Pattern of expression of InsP3Rs, RyRs and TPCs in vascular endothelial cells as detected by reverse transcription/polymerase chain reaction, immunostaining for cultured cells, immunohistochemistry, Western blotting, and immunoprecipitation.
InsP3R1InsP3R 2InsP3R 3RyR1RyR2RyR3TPCs1-2
Human brain microvascular ECs[113]+(RT-PCR, WB, IC)
Human umbilical vein ECs[106]+(RT-PCR)+(RT-PCR)+(RT-PCR)
Human microvascular dermal ECs[213]+(IP)
Human mesenteric artery ECs[127]+(RT-PCR)-(RT-PCR)-(RT-PCR)+(RT-PCR)
EA.hy926 cells[51,127]+(RT-PCR)-(RT-PCR)-(RT-PCR)+(RT-PCR)+(WB)
Mouse cremasteric ECs[108]+(WB, IC)+(WB, IC)+(WB, IC)
Rat hippocampus ECs[413]-(IHC)+(IHC)-(IHC)
Rat aortic ECs[106,107]+(RT-PCR, IC, IHC)+(RT-PCR, IC, IHC)+(RT-PCR, IC, IHC)
Rat mesenteric artery ECs[107]+(RT-PCR, IC, IHC)+(RT-PCR, IC, IHC)+(RT-PCR, IC, IHC)
Rat basilar artery ECs[107]+(RT-PCR, IC, IHC)+(RT-PCR, IC, IHC)+(RT-PCR, IC, IHC)
Rat adrenal medulla microvascular ECs[106]+(RT-PCR)+(RT-PCR)+(RT-PCR)
Rat splenic sinus ECs[35]-(IC)-(IC)+(IC)
Bovine aortic ECs[104,109]+(WB, IC)+(WB, IC)+(WB, IC)
Ovine uterine artery ECs[414]+(IP)
RyRs and cADPr

Although InsP3R is the predominant ER Ca2+-releasing channels in ECs, RyR-dependent signalling has also been reported in vascular endothelium. Three RyR isoforms have been identified (RYR-1, RyR-2, RyR-3) which assemble as homo-tetrameric channels consisting of four 560-kDa polypeptides[125]. They possess a very large cytosolic NH2-domain comprising around 90% of the receptor amino acid sequence. This region is part of a macromolecular complex and serves as a scaffold for numerous regulatory or modulatory proteins, such as calmodulin, calstabin, sorcin, and presents binding sites for cytosolic ions and modulators, including Ca2+, Mg2+, and adenine nucleotides. RyRs may also be modulated PKA- and CamKII-mediated phosphorylation, may undergo oxidative/nitrosative modifications, and are subject to pH modulation[126]. In addition, RyRs are modulated by SR luminal Ca2+via a signaling pathway that includes calsequestrin (CSQ), triadin and junction[126]. Comparative studies carried out by either single-cell reverse transcription/polymerase chain reaction (RT-PCR) or immunocytochemistry have revealed that ECs mainly express RyR-3[34,127], but not RyR-1 and RyR-2 (Table 1) . Immunogold labeling with anti-RyR indicates that RyR-3 is located in the nuclear membranes, the ER tubulovesicular system, and the cristae membrane of mitochondria[34]. In the light of the ability of RyRs to modulate sub-plasmalemmal Ca2+ signaling in ECs (see below), it is remarkable that RyR-3 is largely expressed in the ER structures facing the plasma membrane[34]. Ca2+ release from RyRs in ECs, as well as in both excitable and non-excitable cells[126], is triggered by CICR[44,45], that occurs when 1-10 μmol/L Ca2+ nearby the receptor and is prevented when Ca2+ is 1-10 mm[126]. RyRs may be pharmacologically activated by the methylxanthine derivative, caffeine, which sensitizes the receptor to resting Ca2+ levels, and by ryanodine, which binds to and locks the receptor in an open state[128]. In a number of ECs from different vascular districts, caffeine/ryanodine may evoke either fast Ca2+ transients[46,129,130] or slow Ca2+ release from ER[131,132]. Other studies, however, failed to report any increase in bulk [Ca2+]i in ECs exposed to either caffeine[72,130] or ryanodine[132,133]. Similarly, RyR contribution to agonists-induced Ca2+ signals has been detected in some[44,47,132,134,135], but not all ECs[72,136,137]. RyR-dependent signaling might be masked by SERCA activity, so that Ca2+ ions liberated into the cytosol by RyR opening are immediately sequestered into the InsP3-sensitive region of ER lumen by its Ca2+-ATPase[130]. In addition, monitoring of subplasmalemmal and perinuclear Ca2+ levels revealed that RyR-sensitive Ca2+ mobilization may selectively occur beneath the plasma membrane, but not in the bulk cytosol[133]. Indeed, RyRs may be part of a subplasmalemmal Ca2+ control unit (SCCU), as confirmed by their localization in the ER membrane facing the cytosolic leaflet of the plasma membrane (see above). The SCCU enables a functional coupling between RyRs expressed on superficial ER and plasmalemmal NCX, so that Ca2+ released from RyRs is vectorially pumped out by NCX without affecting [Ca2+]i. Under these conditions, RyR contribution to endothelial Ca2+ signaling may be appreciated only upon NCX inhibition[133,138,139] or by monitoring the local activation of large conductance Ca2+-dependent K+ channels[140,141]. Besides CICR, RyRs may be stimulated by cADPr[10], a second messenger that can be generated in response to either agonist stimulation[142] or oxidative stress through the synthesis of H2O2[143,144]. The ADP ribosyl cyclase, CD38, which is located both on the cell surface and on intracellular membranes, catalyzes nicotinamide adenine dinucleotide (NAD) conversion into cADPr by a cyclization reaction occurring at neutral pH[142]. cADPr, in turn, acts by displacing the FK506 binding protein 12 (FKBP12 or calstabin1) or its related isoform FKBP12.6 (FKBP12.6 or calstabin2) from RyRs[48], which increases their open probability and leads to Ca2+ release[125]. A recent study has demonstrated that bradykinin induces Ca2+ release by stimulating cADPr synthesis and RyR activation in human coronary artery ECs[47]. Moreover, genetic ablation of FKBP12/12.6 from endothelial RyRs caused an intracellular Ca2+ leak associated with the impairment of NO production and endothelial dysfunction[145,146]. It is, however, yet to demonstrate that cADPr triggers Ca2+ release by dissociating FKBP12/12.6 from RyRs in ECs. It should also be noticed that a number of studies have failed to find direct activation of RyRs by cADPr, as discussed previously[48]. For instance, cADPr has been reported to promote the activity of the SERCA pump, thereby increasing the rate of ER refilling and Ca2+ release via the luminal excitation of RyRs[142]. Mechanical stimulation may also lead to RyR activation in vascular ECs. More specifically, shear stress may induce oscillations in [Ca2+]i by stimulating the interplay between mechano-sensitive Ca2+ influx and RyRs through the CICR mechanism[25].

NAADP and two-pore channels

NAADP is the newest and most efficacious member of the family of Ca2+-mobilizing messengers, being active already at nanomolar concentrations[53]. NAADP is synthesized from NADP by CD38 in a base-exchange reaction occurring at acidic pH and targets acidic organelles, such as the endolysosomal Ca2+ storage compartment[53,142], although other signaling pathways may control NAADP production[53]. Ca2+ is sequestered into lysosomal lumen by a putative Ca2+/H+ exchanger, which is driven by a vacuolar-H+ ATPase[53]. NAADP-gated Ca2+ release is mediated by TPC2, whose primary sequence places them in the superfamily of voltage-gated cation channels. Hydropathy analysis has revealed that the full-length protein contains 12 putative TM (TM) α helices that are organized into two homologous domains of six TM segments, each with a membrane re-entrant pore loop between the fifth and sixth segments. Although three TPCs have been identified (TRPC1-3), only TPC2 is endowed with both high (5 nmol/L) and low (10 μmol/L) affinity binding sites for NAADP and is selectively targeted to lysosomes[53]. Due to the limited amount of Ca2+ in the endolysosomal compartment, which consists of small vesicles (0.2-1 μm in diameter) moving throughout the cytosol, NAADP-induced Ca2+ signals per se appear as scattered, discrete events[53]. The relatively small quantity of Ca2+ mobilized by endolysosomes may, however, be amplified into a global Ca2+ wave by recruitment of InsP3Rs and RyRs via CICR[53,147]. This feature requires the physical proximity between the NAADP-sensitive Ca2+ pool and the Ca2+-releasing channels on ER, that has been documented in a number of cell types[53]. NAADP has recently been involved in endothelial Ca2+ signaling (Table 1). A cell-permeable NAADP analog induced an increase in [Ca2+]i in human aortic ECs, which was associated with membrane hyperpolarization and NO production. Importantly, NAADP-dependent Ca2+ release triggered the Ca2+ response to acetylcholine and thrombin, but not ATP[50]. Importantly, NAADP-dependent Ca2+ release triggered the Ca2+ response to acetylcholine and thrombin, but not ATP[50]. Consistent with these data, it was subsequently shown that histamine binding to H1 receptors caused an increase in intracellular NAADP levels in EA.hy926 cells, which evoked a TPC2-mediated Ca2+ signal and secretion of von Willebrand factor[51]. Unlike human aortic ECs, however, the Ca2+ response to thrombin was insensitive to the inhibition of NAADP-dependent signalling[51].

Ca2+ ENTRY FROM THE EXTRACELLULAR SPACE

Ca2+ influx across the plasma membrane sustains the Ca2+ response to both chemical and mechanical stimuli in ECs, as well as most other nonexcitable cells. In addition to refilling the intracellular Ca2+ stores, Ca2+ inflow is responsible for the activation of Ca2+-sensitive effectors which are strategically located near the inner mouth of endothelial Ca2+ channels[6,9,11,62]. Ca2+ entry in vascular endothelium may be mediated by voltage-operated Ca2+ channels (VOCs), agonist-operated Ca2+ channels (AOCs), and mechanosensitive channels (MSCs)[7,11,13,62]. AOCs, in turn, comprise at least three types of Ca2+-permeable membrane pathways, which may be gated by distinct signaling mechanisms: (1) receptor-operated channels (ROCs), which are ionotropic receptors activated by direct binding of the ligand; (2) second-messenger-operated channels (SMOCs), which are stimulated by intracellularly generated mediators, such as DAG, AA and its derivatives (eicosanoids and anandamide), ADPr, and cyclic nucleotides; and (3) SOCs, which open in response to a depletion of the ER Ca2+ pool. Most, but not all, endothelial SMOCs, as well as MSCs, belong to the TRP family of nonselective cation channels[4,8,19,148-150], whereas the molecular nature of SOCs is still debated[151,152]. The inrush of Ca2+ ions consequent to the opening of the Ca2+-permeable membrane route is governed by the following equation:

JCa = 2 ×F×N×p×γCa× (VM - ECa)

where JCa is Ca2+ influx, VM is the membrane potential, ECa the is equilibrium potential for Ca2+, F is the Faraday constant, N is the number of channels, p and γCa, respectively, are their open probability and conductance[7]. Ca2+ entry may, therefore, be modulated by any change in VM according to the following general rule: hyperpolarization increases the driving-force pushing Ca2+ into the cytosol, while depolarization reduces Ca2+ influx across the plasmalemma[7,153]. This rule, however, may not apply to microvascular ECs, which may be endowed with VOCs, as depicted below. We refer the reader interested in the ion channels controlling endothelial VM to a number of exhaustive and relatively recent reviews[7,153]. The description of the Ca2+-permeable membrane pathways in ECs will be addressed by focussing on the following topics: (1) VOCs; (2) ROCs; (3) cyclic-nucleotide-gated channels, which belong to SMOCs; (4) TRP channels, which comprise SMOCs, SOCs, and SACs; and (5) Orai1 and Stim1, which mediate SOCs.

Voltage-gated Ca2+ influx in ECs

Although classically considered as non-excitable cells, ECs from capillaries, but not from large vessels, may express both L- and T-type VOCs[154-159]. VOCs have been detected in microvascular ECs from bovine adrenal glands (BAMCECs)[154-156], rat and pig brain[160,161], rat and mice lungs[157,162-165]. The following features suggest that the voltage-gated currents measured in BAMCECs belong to the L- and T-type described in endocrine secretory cells: their kinetic and gating properties, current-voltage relationships, sensitivity to BAY K 8644, nifedipine, and amiloride, and their different selectivity for Ba2+ and Ca2+[156]. The T-type transient current displayed a conductance of 8 pS and may contribute to depolarisation-induced Ca2+ influx[155,166], while the L-type large current manifests a single-channel conductance of 20 pS and is affected by dihydropyridines[154]. Besides T- and L-type VOC, BAMCECs express a voltage-dependent Ca2+ channel (termed SB) that shows a single-channel conductance of 2.8 pS in 100 mmol/L Ba2+ and is sensitive to BAY K 8644, but not nicardipine[154]. Such a pathway is operative at negative resting potentials and might contribute to low-threshold Ca2+ inflow[154]. Pulmonary microvascular vascular ECs (PMVECs) exhibit a Cav3.1 (α1G) voltage-gated T-type Ca2+ current that is transiently activated at ≥ -60 mV and peaks at around -10 mV[158,159]. Although α1G may mediate T-type voltage-dependent Ca2+ currents without any interaction with its ancillary proteins (β-, γ-, and α2δ subunits), α2δ2-, β1- and β3-, but not β2- and β4-subunits are expressed in PMVECs[158]. The window current of T-type Cav 3.1 channels occurs at membrane potentials ranging from -60 mV to -30 mV, a voltage interval that may be achieved in ECs challenged with extracellular agonists when their VM is close to the K+ equilibrium potential, i.e., -70/-80 mV[7]. Thrombin, in particular, utilize SOCs to depolarize PMVECs and exploit T-type Cav3.1-mediated Ca2+ entry to drive the exocytosis of von Willebrand factor[162]. Exposure of PMVECs to shear stress may upregulate Cav3.1 (α1G) expression, so that flow cessation results in an abrupt Ca2+ entry through T-type Ca2+ channels[157]. Ca2+ entry through T-type Cav 3.1 channels has also been observed in mouse pulmonary capillary endothelium, where it selectively drives P-selectin expression[163]. An R-type VOC has been, finally, described in canine aortic ECs, that may contribute to PAF-induced Ca2+ entry[167]. This channel does not inactivate during long-lasting depolarizations, therefore, it is well suited to mediate prolonged Ca2+ inflow[7]. It is, finally, noticeable that angiotensin II causes α1G though AT1 receptors, Ras and MEK[168]. This study, however, did not assess the appearance of voltage-gated Ca2+ signals, although it showed that mibefradil, a T-type Ca2+ channel inhibitor, prevented the regeneration of an injured monolayer[168].

ROCs in ECs

Vascular ECs may express at least two distinct ROCs, namely the nicotinic acetylcholine (ACh) receptor (nAChR) and the purinergic P2X receptor, which are gated by ACh and ATP, respectively. These pathways are also permeable to K+ and Na+ and gate an inward current at physiological VM. Therefore, in capillary ECs endowed with VOCs, ROCs may also affect Ca2+ signaling by depolarizing the cells and inducing a voltage-dependent Ca2+ inrush.

Neuronal nAChRs: nAChRs are ionotropic receptors that mediate fast cationic currents in response to their selective agonist, ACh. They are named after nicotine, a psychoactive ingredient of tobacco that can mimic the action of ACh in opening these channels. The neuronal nAChRs are pentameric channels comprising either combinations of two different types of subunit (α and β) or five copies of the same α subunit symmetrically arranged in a barrel-like configuration[169]. Each subunit consists of four α-helical spanning domains (TM1-TM4), with TM2 of each subunit lining the receptor pore[169,170]. Ten α isoforms (α1-α10) and four β subtypes [β1-β4], which may assemble in either homomeric or heteromeric configurations, have been identified[170,171]. The diverse stoichiometric combinations may result in ion channels displaying different pharmacological [ligand affinity and α-bungarotoxin (BTX) sensitivity] and biophysical (mean open time and Ca2+ permeability) features[172]. The neuronal α-BTX-sensitive nAChRs are formed by α7-α9 homopentamers, are inhibited by α-BTX, and display the highest Ca2+:Na+ permeability ratio (PCa/PNa), while the non-α-BTX nAChRs are heteropentamers containing at least one α (out of α2-α6) and one β (out of β2-β4) subunits and exhibit a much lower PCa/PNa[172]. As reviewed by Arias et al[173] and Gahring and Rogers[170], ECs from both macro- and microvascular origin present a pattern of α [α2-α5, α7] and β [β2 and β4] isoforms consistent with the expression of both nAChR7 and nAChR3. Activation of the latter does not elicit any detectable Ca2+ influx in vascular endothelium[174], while the former triggers a proangiogenic Ca2+ entry that might be exploited to enhance tissue revascularization[173,175]. Moreover, nAChRs-mediated membrane depolarization induces VOCs-dependent Ca2+ inrush in BAMCECs[166].

P2X receptors: The ionotropic P2X receptors are channels that respond to extracellular ATP to induce membrane depolarization and an influx of Ca2+. Seven P2X receptor subtypes (P2X1-7) have been identified: each of them possesses two TM domains (TM1 and TM2) connected by a large extracellular loop and with the NH2- and COOH-terminal tails extending into the cytoplasm[176]. The extracellular loop contains the ATP binding site, while TM1 and TM2 are thought to line the channel pore[176]. The P2X receptors assemble as homo- or heterotrimers with a range of biophysical features and functions, including different PCa/PNa values (which may range from 1 to 71) and rates of inactivation[177]. For instance, P2X3-mediated currents rapidly decay during prolonged exposition to ATP, whereas P2X7 remains open for several minutes[176]. Vascular ECs express all P2X receptor isoforms in both arteries and veins[178,179], although P2X4 levels are higher in the latter[179]. P2X4 expression in the arterial endothelial layer undergoes a dramatic increase following balloon injury, which hints at this subunit as a key player in EC regeneration[179]. The electrophysiological analysis showed that channels with the electrophysiological and pharmacological properties of both P2X4 (fast inactivation, inward rectification, conductance of around 36 pS, insensitivity to reactive blue) and P2X7 (slow inactivation, conductance of 9 pS, linear current-to-voltage relationship, sensitivity to reactive blue) are functionally present on EC plasma membranes[180]. P2X5 subunits may assemble with P2X4 subtypes and form a heterotrimer associated to a current displaying the characteristics of a P2X4 current[180]. Ca2+ imaging recordings demonstrated that P2X4 is the main mediator of ATP- and flow-induced Ca2+ inflow in vascular endothelium[181,182]. Shear stress induces ATP release from ECs through different mechanisms, including exocytosis or ATP binding of cassette transporters[183,184], or it may stimulate a cell surface ATP-synthase[185]. ATP, in turn, causes a step-like elevation in [Ca2+]i[181,182], which leads to NO production and vasorelaxation in vivo[186]. ATP-dependent Ca2+ entry is linearly correlated with the amplitude of the mechanical stimulation[187].

Cyclic nucleotide-gated Ca2+ channels

Cyclic nucleotide-gated channels (CNGs) are nonselective cation channels gated by direct binding of the cyclic nucleotides, cAMP and cGMP[188]. There are six members of the CNGC family: four α subunits (CNGA1-4) and two β subunits (CNGB1 and CNGB3). They all share structural homology with voltage-gated channels and have six TM domains with both the NH2- and COOH-termini facing the cytoplasm. A pore region lies between TM5 and TM6. The CNGA1-3 subunits form channels whereas CNGA4 and the two CNGB isotypes have a modulatory function. The cyclic nucleotide binding site is located at the COOH-terminal domain of each subunit[188]. All CNGs conduct Ca2+ better than Na+, a feature which enables these conductances to affect significantly intracellular Ca2+ signaling[189]. Both CNGA1, which is selectively gated by cGMP, and CNGA2, which opens in response to cAMP, are widely expressed by ECs from several vascular beds, although cultured cells undergo downregulation of both subunits[190,191]. Both cAMP and cGMP levels may increase in ECs challenged with β-receptor agonists, such as epinephrine, noradrenaline and isoprenaline[192,193]. Moreover, intracellular levels of both cyclic nucleotides may be raised when ECs are stimulated with adenosine and ATP[192-194], which are coupled, respectively, to A2B and P2Y1 receptors[194-196], and VEGF, which acts through VEGFR-2[196]. Finally, SOCs-mediated Ca2+ entry may lead to the intracellular accumulation of cGMP, which in turn leads to CNG opening and membrane depolarization, thus establishing a negative feedback responsible for SOC closure[197]. CNGA2 is the main contributor to cyclic-nucleotide-induced Ca2+ inflow in vascular endothelium stimulated with epinephrine[198], ATP[195] or adenosine[194]. CNGA2-mediated Ca2+ influx, in turn, results in NO production and vasorelaxation[194]. Moreover, NO may upregulate CNGA2 levels in ECs[199].

TRP channels

The super-family of TRP channels comprises 28 members, which have been subdivided in six categories based on the homology of their amino-acid sequence: the TRPC (canonical; TRPC1-TRPC7), TRPV (vanilloid; TRPV1-TRPV6), TRPM (melastatin; TRPM1-TRPM8), TRPP (polycystin; TRPP1-TRPP3), TRPML (mucolipin; TRPML1-TRPML3), and the TRPA (ankyrin; TRPA1) families[200]. They all are nonselective cation channels, with the exception of TRPM4 and TRPM5, which are only permeable to Na+ and K+ (PCa/PNa < 0.05), and TRPV5 and TRPV6, which are Ca2+ selective (PCa/PNa > 100)[200]. All the family members have the same basic structure consisting of six TM domains (TM1-TM6) with the NH2- and COOH-termini located in the cytoplasm. Between TM5 and TM6, which are highly conserved among all the homologs, there is a re-entrant pore loop constituted by a short hydrophobic stretch lining the channel pore. This molecular architecture resembles that of the archetypal 6-TM voltage-gated channels, although TM4 lacks the conserved block of arginine residues forming the voltage sensor of transmembrane electrical potential[201]. Most of the differences between the three families are found within the NH2- and COOH-terminal tails. The TRPC and TRPV family members contain three or four ankyrin repeats at their NH2 terminus, while TRPA1 contains 14 ankyrin repeats. All mammalian TRP families present a homologous sequence of around 25 amino acids immediately COOH-terminal to TM6, which has been termed TRP domain and consists of a highly conserved six-amino-acid TRP box (EWKFAR) and a proline-rich motif. The TRP domain is, however, lacking in TRPA and TRPP[201]. TRPC proteins contain a number of calmodulin/InsP3R binding domains (CIRB) at both their NH2- and COOH-tails, while a predicted coil:coil region may be present in both locations of TRPC, TRPM, and TRPP[19]. The TRPM family is characterized by having very long N and C termini. The latter contain enzymatic activity in some of the channels, e.g., TRPM2 has an ADP-ribose pyrophosphatase, whereas TRPM7 contains an atypical protein kinase[200]. TRP subunits within one subfamily may assemble in either homo- or heterotetramers and this gives rise to large variability in their biophysical properties and activating mechanisms[19,200,201]. Vascular ECs have been shown to express TRP channels from all sub-families[8,150]. More specifically, all the TRPC, all TRPV, all TRPM except TRPM5, TRPP1, TRPP2 and TRPA1 are present in ECs (Tables 2, 3 and 4). Endothelial TRP channels serve as polymodal cell sensors, as each TRP subunit may be engaged by more than one signaling pathway. The selective coupling to distinct downstream Ca2+-sensitive decoders enables TRP-mediated Ca2+ influx to govern a plethora of processes in ECs[4,19,148,149]. For sake of clarity, we will separately address the activating mechanisms and the main functions fulfilled by each TRP isoform in vascular endothelium.

Table 2 Pattern of expression of transient receptor potential canonical channels in vascular endothelial cells as detected by reverse transcription/polymerase chain reaction, immunostaining for cultured cells, immunohistochemistry, Western blotting, northern blots, and in situ hybridization.
TRPC1TRPC2TRPC3TRPC4TRPC5TRPC6TRPC7
PCa/PNa and conductance (pS)[202]Non-selective, 162.7, 421.6, 60-661.1, 30-429, 47-665, 28-370.5-5.4, 25-50
Human pulmonary artery ECs[210,211,219,222,249,261,415,416]+(RT-PCR, NB, WB, IC)+/-(RT-PCR)+(RT-PCR, WB, IC)-(RT-PCR)+/-(RT-PCR) +(WB)
Human coronary artery ECs[234,237]+(RT-PCR, WB, ISH, IHC)+(RT-PCR, WB, ISH, IHC)+(RT-PCR, WB, ISH, IHC)+(RT-PCR, WB, ISH, IHC)+(RT-PCR, WB, ISH, IHC)+(RT-PCR, WB, ISH)
Human cerebral artery ECs+(RT-PCR, ISH, IHC)+(RT-PCR, ISH, IHC)+(RT-PCR, ISH, IHC)+(RT-PCR, ISH, IHC)+(RT-PCR, ISH, IHC)+(RT-PCR, ISH)
Human intact mesenteric artery[127]+(RT-PCR)+(RT-PCR)-(RT-PCR)-(RT-PCR)
Human glomerular ECs+(WB, IC)
Human preglomerular ECs[244,245]+(RT-PCR, WB, IHC)+(IHC)
Human dermal microvascular ECs[18,212,221,222,264]+(RT-PCR, WB)-(RT-PCR)+(RT-PCR)+(RT-PCR, WB)-(RT-PCR)+/-(RT-PCR) +(WB)+/-(RT-PCR)
HUVECs[127,205,212,215,218,220-223,227,231]+(RT-PCR, WB)-(RT-PCR)+/-(RT-PCR)+/-(RT-PCR)+/-(RT-PCR)+/-(RT-PCR)+(RT-PCR)
+(WB)
EA.hy926 cells[127,243,269]+(RT-PCR)+/-(RT-PCR) +(WB)+/-(RT-PCR)+(RT-PCR, WB)+/-(RT-PCR)
Mouse aortic ECs[216,228,248,258]+(RT-PCR, WB, IC)+(WB)+(RT-PCR, WB)+(WB)+(WB)
Mouse pulmonary artery ECs[218,251,259,262]+(RT-PCR, WB)-(RT-PCR)+/-(RT-PCR)+(RT-PCR, WB)-(RT-PCR)+(RT-PCR, WB)-(RT-PCR)
Mouse brain capillary ECs+(RT-PCR)+(RT-PCR)+(RT-PCR)+(RT-PCR)+(RT-PCR)+(RT-PCR)+(RT-PCR)
Bovine pulmonary artery ECs[232]+(RT-PCR)-(RT-PCR)+(RT-PCR)
Bovine aortic ECs[207-209,248,267,268]+(RT-PCR, WB, IC)+/-(RT-PCR) +(WB)+(RT-PCR, WB)+(RT-PCR, WB)+(RT-PCR, WB)
Bovine brain capillary ECs+(RT-PCR)-(RT-PCR)+(RT-PCR)-(RT-PCR)+(RT-PCR)-(RT-PCR)-(RT-PCR)
Bovine adrenal capillary ECs[292]
Rat aortic ECs[233,239]-(WB)+(WB)-(WB)+(WB)+(WB)
Rat pulmonary artery ECs[260]+(RT-PCR)+(RT-PCR)+/-(RT-PCR)+(WB, IC)-(RT-PCR)-(RT-PCR)
-(RT-PCR)
Rat coronary arterioles ECs+(RT-PCR, IHC)
Rat cerebral artery ECs[311]+(RT-PCR)
Rat mesenteric artery ECs[227,312]+(RT-PCR, WB, IHC)+(WB, IHC)+(WB, IHC)+(WB, IHC)
Rat descending vasa recta ECs[254]+(RT-PCR, WB, IC)-(RT-PCR)
Rat splenic sinus ECs[35]+(IHC)
Porcine aortic ECs[241,242]+(RT-PCR, WB)+(WB)
Porcine coronary artery ECs[235]+(RT-PCR, WB)
Zebrafish aortic ECs[225]+(IHC, ISH)+(ISH)
Ovine uterine artery ECs[414]+(WB)+(WB)+(WB)+(WB)+(WB)
Table 3 Pattern of expression of transient receptor potential vanilloid channels in vascular endothelial cells as detected by reverse transcription/polymerase chain reaction, immunostaining for cultured cells, immunohistochemistry, Western blotting, and in situ hybridization.
TRPV1TRPV2TRPV3TRPV4TRPV5TRPV6
PCa/PNa and conductance (pS)[202]10, 35-801-3, NM12, 1726, 90> 100, 75> 100, 40-70
Human coronary artery ECs[300](+RT-PCR, WB, IHC)
Human pulmonary artery ECs[415]+(RT-PCR)+(RT-PCR)+(RT-PCR)
Human pulmonary microvascular ECs+(RT-PCR)
Human cerebral microvascular ECs[272]+(RT-PCR, IC)
Human cerebral arterioles ECs[305]+(RT-PCR, IHC)
Human dermal microvascular ECs[319]+(RT-PCR, WB)
Breast cancer derived microvascular ECs[319]+(RT-PCR, WB)
Human umbilical vein ECs[227,277,296,301]+(RT-PCR)+(WB, IHC)
Mouse aortic ECs[280,285,299,306]+(WB)+(RT-PCR, WB, NM, IHC)
Mouse pulmonary artery ECs[313]+(WB, IHC)
Mouse mesenteric artery ECs[27,280,302]+(RT-PCR, WB, IC)+(RT-PCR, IC)
Mouse cerebral microvascular ECs+(RT-PCR)+(RT-PCR, WB)+(RT-PCR)
Mouse dermal microvascular ECs[307]+(RT-PCR, WB)
Mouse carotid artery ECs[290,296]+(IHC)
Rat mesenteric artery ECs[226,275,302]+(WB)+(WB, IC, IHC)
Rat femoral artery ECs[312]+(RT-PCR, IC)
Rat pulmonary artery ECs[303]+(WB, IHC)
Rat renal artery ECs[303]+(IHC)
Rat cardiac microvascular ECs[303,316]+(IHC)
Rat cerebral artery ECs[287,297,311]+(RT-PCR)+(RT-PCR, IC)+(RT-PCR, IC)
Rat carotid artery ECs[291]+(RT-PCR)
Porcine aortic ECs[312,317]+(RT-PCR, IC, WB)
Porcine coronary artery ECs[273]+(IHC)
Bovine aortic ECs[271]+(WB)
Bovine adrenal capillary ECs[292]+(RT-PCR, WB)
Zebrafish ventricular ECs+(ISH)
Table 4 Pattern of expression of transient receptor potential melastatin channels in vascular endothelial cells as detected by reverse transcription/polymerase chain reaction, Western blotting, and in situ hybridization.
TRPM1TRPM2TRPM3TRPM4TRPM5TRPM6TRPM7TRPM8
PCa/PNa and conductance (pS)[202]NM, NM0.5-1.6, 52-800.1-10, 65 (Ca2+)-130Not permeable, 25Not permeable, 16-25PCa/PNa~6, NM3, 40-1051-3, 83
Human Pulmonary artery ECs[320]+(RT-PCR)+/-(RT-PCR) +(WB)+(RT-PCR)+(RT-PCR)-(RT-PCR)+(RT-PCR)+(RT-PCR)+(RT-PCR)
Human umbilical vein ECs[325,327,328]+(RT-PCR, WB)-(RT-PCR)+(RT-PCR, WB)
Mouse aortic ECs[329]+(RT-PCR, WB)+(RT-PCR)
Human dermal microvascular ECs[327]+(RT-PCR, WB)
Mouse cerebral microvascular ECs-(RT-PCR)+(RT-PCR)+(RT-PCR)+(RT-PCR)-(RT-PCR)-(RT-PCR)+(RT-PCR)-(RT-PCR)
Mouse cardiac microvascular ECs[8]+(RT-PCR, WB)+(RT-PCR)
Rat cerebral artery ECs[311]+(RT-PCR)+(RT-PCR)+(RT-PCR)
Rat spinal cord capillary ECs[324]+(ISH)
Rat mesenteric artery ECs[312]+(RT-PCR)
TRPC channels

The TRPC family may be subdivided in four subfamilies according to their amino acid homology: TRPC1, TRPC2, TRPC3/6/7 (around 80% homology), and TRPC4/5 (around 60% homology). TRPC2 is a pseudogene in humans and does not fulfil any functional role in vascular ECs. TRPC channels are nonselective cation channels whose PCa/PNa may vary from 0.9 to 9[202]. The TRPC subunits may assemble as homo- and heterotetramers. In the latter case, they may combine with members of their own subfamily or TRPC1 may form heteromers with TRPC1, TRPC4/5 and TRPC3[200]. Endothelial TRPC channels are activated by GPCRs and TKRs coupled to PIP2 hydrolysis by PLC activation and, depending on their stimulating mechanism, may be part of either SMOCs or SOCs.

TRPC1: TRPC1 is widespread in both cultured and native ECs (Table 2), where it mainly serves as SOC and controls transendothelial permeability[203,204], although recent studies have argued against this conclusion[205,206]. TRPC1 is engaged by a number of inflammatory agonists, such as ATP, thrombin, tumor necrosis factor (ΤΝF)-α, and angiotensin II (AngII), and by growth factors, such as VEGF and bFGF[5,12,148,207-209]. TRPC1 may act as a SOC due to its ability to interact with InsP3Rs through several signaling cascades. The first model implies RhoA-dependent actin rearrangement, which leads InsP3Rs to couple to and gate TRPC1[210]. RhoA activation, in turn, is induced by TRPC6 via PKCα-mediated phosphorylation[211]. At the same time, PKCα also phosphorylates TRPC1 to enhance RhoA-dependent stimulation[212]. Alternatively, thrombin utilizes caveolin-1 (Cav-1) to trigger TRPC1-mediated Ca2+ entry: Cav-1 scaffolding domain (CSD) enables the physical and functional interaction of TRPC1 and of InsP3R3 with Cav-1 protein within the caveolae[213]. In this view, a recent work has demonstrated that polyunsaturated fatty acids, such as docosahexaenoic acid, may reduce store-operated Ca2+ entry (SOCE) by displacing TRPC1 from membrane caveolae lipid rafts[214]. Finally, VEGF recruits TRPC1 via direct coupling to InsP3R2 in an angiopoietin-1-regulated fashion[215]. The circulating protein, klotho, may bind to both VEGFR-2 and TRPC1 to promote their co-internalization, reduce Ca2+ influx and maintain endothelial integrity[216]. A number of studies have, however, reported that TRPC1 must combine with TRPC4 to be operated by intracellular Ca2+ stores[217,218]. This issue will be addressed below. TRPC1-mediated Ca2+ entry enlists calmodulin to activate the myosin light chain kinase (MLCK) and induce EC retraction, thus disassembling intercellular junctions and increasing endothelial permeability[204]. Moreover, TRPC1-gated Ca2+ entry is capable of stimulating nuclear factor (NF)-κB via AMP-activated protein kinase and PKCδ, so that NF-κB may initiate the transcriptional programme involved in the host defense to inflammatory stimuli and in EC resistance to apoptosis[219,220]. NF-κB, in turn, increases TRPC1 expression in ΤΝF-α-stimulated cells, thus establishing a feed-forward loop that amplifies agonist-elicited SOCE and the increase in transendothelial permeability[221,222]. Similarly, AngII upregulates TRPC1 levels in an NF-κB-dependent manner in ECs[223]. Finally, TRPC1 may be involved in endothelial tube formation both in vitro[224] and in vivo[225]. The expression of TRPC1 in vascular ECs may be increased by high glucose levels, with a consequent increase in the amplitude of agonist-induced Ca2+ influx and a dramatic impact on endothelial function in diabetic patients[209]. TRPC1 not only acts as an SOC in vascular endothelium, but may also serve as an MSC by assembling with TRPV4 to mediate flow-induced Ca2+ entry in vascular endothelium[226]. A drop in luminal Ca2+ favors the translocation of the TRPC1-TRPV4 heteromeric channel from the ER to plasmalemmal caveolae[227], where the complex amplifies shear-stress-induced Ca2+ influx and augments the ensuing vascular dilation[227]. In addition, such an enhanced vesicular trafficking of TRPC1-TRPV4 channels may contribute to endothelial SOCs[228]. Similarly, TRPC1 may mediate stretch-induced Ca2+ entry after traumatic injury in microvessel endothelial cells, thereby resulting in NO synthesis and actin stress fiber formation[229].

TRPC3/6/7: TRPC3, TRPC6 and TRPC7 are Ca2+-permeable channels that may be activated by DAG in a PKC-independent manner and may thus be regarded as SMOCs[200,201]. TRPC3 and TRPC6 are widely expressed both in cultured ECs and in situ endothelium, while TRPC7 is far less abundant (Table 2). DAG stimulates TRPC3- and TRPC6-gated Ca2+ entry into ECs exposed to a number of agonists, including VEGF[18,230,231], ATP[232-234], and bradykinin[232,235]. TRPC3- and TRPC6-induced Ca2+ inflow underpin the stimulatory effect of VEGF on endothelial proliferation, migration and permeability[18,231,236], while ATP exploits TRPC3 to activate NF-κB and increase vascular cell adhesion molecule-1 expression[237]. TRPC3-driven NO synthesis and vasorelaxation may be attenuated by the hypoxia-reoxygenation injury of vascular ECs, which prevents TRPC3 trafficking to the plasma membrane[235]. Moreover, TRPC3 supports both spontaneous and histamine-evoked Ca2+ oscillations arising, respectively, during and after EC tubule formation[224]. Finally, TRPC3-mediated Ca2+ inflow may stimulate flow-dependent endothelium-dependent vasodilation[238]. Endothelial TRPC3 and TRPC6 may be inhibited by NO in a protein-kinase-G-dependent fashion[239,240], while TRPC3 may also be inhibited by a PKC-mediated phosphorylation on Ser-712[239]. Both TRPC3 and TRPC6 may be recruited by signaling mechanisms other than DAG in vascular ECs. TRPC3 may provide a leakage Ca2+ entry pathway in unstimulated ECs[234] or behave as an SOC owing to its ability to physically interact with InsP3R2 in uterine artery ECs[105]. Connexin-43-mediated intercellular communication between adjacent ECs favors TRPC3/InsP3R2 association, although via a yet to be identified mechanism[105]. Moreover, endothelial TRPC3 may assemble with its distant relative, TRPC4, and form a heteromeric channel sensitive to oxidative stress and with biophysical properties distinct from both TRPC3 and TRPC4 homomers[241,242]. Finally, TRPC3 may be activated by thapsigargin-induced ER Ca2+ store depletion following PLC-dependent DAG synthesis and PKCη activation. PKCη, in turn, engages Src tyrosine kinase to phosphorylate and gate TRPC3[243]. TRPC3 levels in vascular ECs may be downregulated by VEGF in a PI3K-sensitive manner, and upregulated by an increase in blood pressure[244,245]. TRPC6, on the other hand, may be translocated from intracellular sites to the caveolae compartment of the plasma membrane by cytochrome P450 (CYP)-derived epoxyeicosanotrienoic acids (EETs). This process enhances bradykinin-induced Ca2+ entry and thus prolongs the activation of small and intermediate Ca2+-activated K+ channels responsible for endothelial hyperpolarization[246]. The functional consequence of TRPC6 insertion into the plasmalemma is related to the ability of ECs to transmit their negative shift in VM to the adjacent smooth muscle cells and trigger vasorelaxation (the so-called endothelial-dependent hyperpolarizing factor, EDHF)[247]. TRPC6 externalization may also be induced by lysophosphatidylcholine (LPC): the following Ca2+ influx induces TRPC5 translocation to the plasma membrane, thus allowing sustained Ca2+ entry into ECs[248]. TRPC5-mediated Ca2+ inflow, in turn, is essential to restrict endothelial movement and might be relevant in impairing the healing process of the intimal layer in atherosclerotic arteries, where LPC is abundant[248]. TRPC6 externalization requires the phosphatase and tensine homolog (PTEN), whose tail 394-403 residues associate with the channel and favour its plasmalemmal localization[249]. It has yet to be elucidated whether EETs and LPC utilize PTEN to induce the intracellular movements of TRPC6 in ECs. Recent studies have shown that TRPC6 plays a key role in the onset of the lung ischemia-reperfusion edema (LIRE) as a consequence of endothelial NOX2-dependent ROS, which, in turns, leads to PLCγ activation, DAG kinase inhibition and an increase in submembranal DAG levels[250]. Consistently, PAF-induced lung edema depends upon the activation of acid sphingomyelinase (ASM), which engages TRPC6 to caveolae, thereby resulting in Ca2+ influx and following increase in endothelial permeability[251]. This process is inhibited by NO donors[251].

TRPC4/5: TRPC4 and TRPC5 form a subfamily with the closest homology with TRPC1[201]. Both channels open in response to the Gq/11 family GPCRs and TKRs, although neither InsP3 nor DAG alone are able to activate them[200,201,252]. It has, however, been demonstrated that TRPC4 opening is favored by PIP2 hydrolysis, although it also requires Gi recruitment and intracellular Ca2+[253]. The scenario is further complicated by the necessity for TRPC4 to interact with the adaptor NHERF, ezrin/radixin/moesin ERM proteins, and cortical actin via its COOH-terminal PDZ-binding domain to undergo PIP2 modulation[253,254]. Similarly, TRPC5 activation may be potentiated by intracellular Ca2+ release and is modulated by the scaffolding ERM-binding phosphoprotein-50 (EBP50)[255,256]. TRPC4 and TRPC5 may assemble either as homo- or heterotetramers via their first ankyrin repeat in the NH2-terminal[257] and are abundantly expressed in both cultured cells and naive tissues. TRPC4 has long been considered as the key component of SOC in vascular endothelium[203], because TRPC4-deficient mice lack SOCE and exhibit impaired NO production and increased vascular permeability[258,259]. According to this model, the extent of store depletion is conveyed to PM by protein 4.1N, which links the cytoskeletal protein, spectrin, to TRPC4[260]. The latter, in turn, contributes pore-forming subunits to the store-dependent channel along with TRPC1 according to the following stoichiometry, 1 TRPC1 and 3 TRPC4[203,217,260,261]. The association between the two subunits is favoured by Cav-1, which also recruits InsP3R1 into the membranal signalplex[262]. InsP3R1, in turn, may interact with protein 4.1N, a feature which would explain the dependence of the TRPC1-TRPC4 complex on intracellular Ca2+ stores[263]. An alternative model proposes that this heterotetramer is activated by a protein termed stromal interaction molecule-1 (Stim1)[218], which senses intraluminal ER Ca2+ levels, as more widely illustrated below. A recent study, however, questioned TRPC1 and TRPC4 involvement in endothelial SOCE[205,206]. The controversy was explained by the finding that TRPC4 is externalized to the plasma membrane upon binding to β-catenin only in subconfluent, migrating ECs, but not in quiescent, barrier-forming cells, which is the typical condition for Ca2+ signaling studies[264]. Therefore, TRPC4 might serve as an agonist-recruited SOC during endothelial cell-cycle transitions from a proliferating to a quiescent phenotype[265]. Whether TRPC4 contributes to the repetitive Ca2+ spikes arising in HUVECs both before (spontaneous) and after (histamine-evoked) tubule formation in a store-dependent or independent manner has yet to be elucidated[224]. A single nucleotide polymorphism in the TRPC4 gene has been associated with protection from myocardial infarction and results in a larger current density when the mis-sense channel is expressed in heterologous cell systems[266]. This feature is likely due to a less bulky Val-957, which favors TRPC4 insertion into the plasma membrane and its interaction with the catalytic site of the tyrosine kinase that phosphorylates the channel at Tyr-959[266]. Less information is available on the role played by TRPC5 in ECs, where it may be constitutively open and also participate to ATP-induced Ca2+ entry[267]. TRPCs have been suggested to be activated also by NO via direct S-nitrosylation[186], but this mode of regulation is absent in vascular endothelium, where TRPC5 is actually inhibited by NO[267]. Conversely, the endothelial TRPC5 channel may be potentiated by the plant-derived isoflavone, genistein, independently on tyrosine kinases[268] and is upregulated by erythropoietin[269].

TRPV channels

The TRPV family consists of four subfamilies on the basis of their structure and function: TRPV1/2, TRPV3, TRPV4, and TRPV5/6[201]. The latter are highly Ca2+-selective pathways, while TRPV1-4 are nonselective cation channels whose PCa/PNa ranges between 1 and 15 pS[202]. All TRPV channels are endowed with 3-6 NH2-terminal ankyrin repeats, which enable the protein to traffic to the plasma membrane, are required for tetramerization, and control protein-protein interactions[270]. TRPV channels may assemble either as homo- or heterotetramers, although their exact stoichiometry in situ is still unknown[270]. TRPV1-4, which are the only isoforms expressed in vascular ECs, may be gated by a number of stimuli, including Gq signaling, endocannabinoids, dietary agonists, changes in temperature and osmolality, pulsatile stretch and shear stress[270]. TRPV channels are, therefore, a remarkable example of polymodal channels and may serve as both MSCs and SMOCs.

TRPV1: TRPV1 is a nonselective cation channel with a preference for calcium that can be expressed in both cultured and native ECs (Table 3). Endothelial TRPV1 channels may be activated by the dietary agonist, capsaicin, and the endogenous cannabinoids, anandamide and 2-arachidonoyl-glycerol (2-AG)[149,270-272]. TRPV1 may, therefore, serve as an SMOC, although the physiological ligands that are capable of recruiting this channel in ECs are yet to be elucidated. Pharmacological stimulation of TRPV1, either by capsaicin or anandamide, causes vascular relaxation upon release of several vasodilators, including NO[273-275], calcitonin gene-related peptide (CGRP)[276,277], and EDHF[273]. CGRP may also afford a protective effect against LPC- and lipopolysaccharide (LPS)-dependent injury[276,277]. Furthermore, TRPV1 mediates the vasorelaxing effects of the synthetic, cannabinoid-like compound, VSN16, albeit it is still unknown whether TRPV1 is engaged by VSN16 directly or via the “abnormal-cannabidiol receptor”[278]. The signaling cascade responsible for endothelial NO synthase (eNOS) activation by TRPV1 has been unravelled. TRPV1 may trigger Ca2+-dependent PI3K/Akt/CaMKII signaling, which leads to TRPV1 phosphorylation, increases TRPV1-eNOS complex formation, eNOS stimulation, and NO synthesis[271]. This signaling pathway might involve the AMP-activated protein kinase (AMPK), a multifunctional regulator of energy homeostasis that is phosphorylated upon TRPV1-mediated Ca2+ influx[279]. TRPV1-induced vasodilation is impaired in obese Ossabaw swine with the metabolic syndrome and thus could be a mechanism involved in endothelial dysfunction and development of coronary disease[273]. Chronic stimulation of TRPV1 by capsaicin enhances PKA and eNOS phosphorylation, which, in turn, results in vasorelaxation, reduces blood pressure and delays the onset of brain stroke in spontaneously hypertensive rats[280]. Moreover, dietary capsaicin reduces the high-salt-diet-induced endothelial dysfunction and nocturnal hypertension by preventing the generation of O2.- anions and NO reduction through vascular TRPV1 activation[281]. Baseline mean arterial pressures are, however, not statistically different in wild-type and TRPV1-/- mice[282,283]. Besides chemical agonists, TRPV1 may be gated by temperature higher than 40 °C in ECs[277,284]. This feature gives support to the hypothesis that TRPV1 contribute to peripheral vascular dilatation and high temperatures[8]. Conversely, a recent report described TRPV1-induced endothelin production by cultured ECs, which might cause the activation of vascular ETA receptors and increase the mean arterial pressure in vivo[285]. Finally, capsaicin opposes angiogenesis by causing G1 arrest of ECs via downregulation of cyclin D1 and inhibited VEGF-triggered signaling pathways[286]. Whether this effect depends on TRPV1 stimulation is yet to be demonstrated.

TRPV2: TRPV2 has been mainly located in vascular smooth muscle cells (VSMCs), where it may be activated by cell swelling and act as an MSC[270]. Only a few studies have reported TRPV2 expression in ECs (Table 3) and, therefore, its function in vascular endothelium is still unclear. Due to its ability to mediate Ca2+ inflow in ECs exposed to moderate heat, however, it might aid TRPV1 to stimulate NO synthesis and peripheral vasorelaxation at > 43 °C[8,284].

TRPV3: TRPV3 is a polymodal channel that may be gated by either natural compounds, such as carvacrol, eugenol, and camphor, or innocuous warm temperatures (> 30 °C)[200]. TRPV3 expression has recently been shown in both in situ and cultured ECs (Table 3). In intact tissue, TRPV3 stimulation by carvacrol induces VSMC hyperpolarization via the Ca2+-dependent opening of intermediate- (IKCa) and small-conductance (SKCa) K+ channels. The consequent K+ efflux activates smooth muscle inwardly rectifying potassium channels (KIR) and amplifies the negative shift in VSMC VM, thereby promoting arterial relaxation[287]. In human corneal ECs, TRPV3-induced Ca2+ entry is evoked by raising bath temperature from 25 to > 44 °C[284]. If present also in vascular endothelium, this feature would render the channel capable of sensing high temperatures in peripheral circulation along with TRPV1 and TRPV2.

TRPV4: TRPV4 is the most polymodal TRP channel, being a molecular integrator of both chemical and physical stimuli[200,270], and may be expressed in both native and cultured ECs (Table 3). More specifically, endothelial TRPV4 may be activated by non-noxious temperatures (> 24 °C)[288], hypo-osmotic cell swelling[289], shear stress[290,291], cyclic strain[292,293], and pharmacologically, by the non-PKC-activating phorbol ester, 4α-phorbol-12,13-didecanoate (4αPDD)[294] and by the plant-derived flavone, apigenin[295]. Cell swelling and shear stress do not gate TRPV4 directly, but via the PLA2-dependent synthesis of AA, and its subsequent metabolization to 5’,6’-epoxyeicosatrienoic acid (EET) through a CYP epoxygenase-dependent pathway[290,291,296-299]. A unique mechanism has been described in human coronary arterioles isolated from patients suffering from coronary artery disease, where the endothelial TRPV4 stimulate mitochondria to release ROS, which in turn cause VSMCs to relax[300]. Conversely, TRPV4 signaling induced by mechanical strain is mediated by direct force transfer from β1 integrins to the integrin-associated transmembrane CD98 protein within focal adhesions[293]. This feature renders TRPV4 a pure MSC that can be activated within only 4 ms by external force application[293]. TRPV4-mediated Ca2+ responses to heat and 4αPDD require an aromatic amino acid at the NH2-terminal of the third transmembrane domain[289]. Recent work proposed that Cav-1 governs TRPV4 activity and location to the plasma membrane in ECs[301]. NO, PGI2, and IKCa/SKCa/MaxiK-mediated EDHF are vasodilatory pathways that are activated following by pharmacological activation of TRPV4[27,291,302-304]. An alternative mechanism has been unravelled in human coronary arterioles from patients with coronary artery disease, where flow-induced TRPV4 activation causes ECs to release mitochondrial ROS, thereby relaxing the adjoining VSMCs[300]. Under physiological conditions, TRPV4-mediated vasodilation, mainly via NO and EDHF production, is induced by both shear stress[27,290,291,296] and extracellular autacoids, such as Ach and UTP[297,301,305,306]. Although the cellular mechanisms linking membrane receptors to TRPV4 opening are still unclear, UTP utilizes PLA2 to gate the channel[297], while PKCα mediates Ach-dependent activation[307]. Despite the evidence for TRPV4-evoked vasodilation, the baseline blood pressure is not affected in TRPV4-/- mice[305,306]. Nevertheless, systemic activation of the channel may induce vasorelaxation in vivo[302,303,308,309] and, eventually, lead to circulatory collapse[303]. The vasorelaxing effect of TRPV4 is more evident in rodent models of Nω-nitro-l-arginine (L-NNA)- and salt-induced hypertension[305,308,309]: this feature suggests that its main function might be to adverse excessive vasoconstricting stimuli and protect the vasculature during times of extreme stress[270]. It is, therefore, intriguing that TRPV4 is downregulated in a genetic model of salt-susceptible hypertensive rats exposed to a salt-enriched diet, whereas its function and expression are enhanced in salt-resistant rat strains[308]. In this view, apigenin might afford a protective effect on the development of cardiovascular diseases by stimulating TRPV4 to release EDHF[295]. An additional role for TRPV4-dependent Ca2+ signals consists in driving the shear stress-induced growth of collateral vessels (arteriogenesis) which occurs as a consequence of arterial occlusion[310-312]. Ca2+ entry may, indeed, recruit several Ca2+-sensitive transcription factors, including nuclear factor of activated T cells, cytoplasmic, calcineurin-dependent 1 (NFATc1), Kv channel interacting protein 3, calsenilin (KCNIP3/CSEN/DREAM), and myocyte enhancer factor 2C (MEF2C)[310]. On the other hand, TRPV4-mediated Ca2+ influx elicited by cyclic strain triggers cytoskeletal remodeling and movement necessary for cell realignment in the vasculature[292]. TRPV4 channels are extensively expressed in alveolar septal capillaries[163,313], where they control lung endothelial permeability and barrier integrity in response to intravascular and airway pressures[314-316]. More specifically, TRPV4-gated Ca2+ entry increases vascular permeability to such an extent that barrier disruption and alveolar flooding may occur[313]. Finally, TRPV4 may enable ECs to sense body temperature, because the threshold for the heat-dependent activation for this channel is set at 25-27 °C and it is activated by moderate heat in ECs[4,288,317,318]. The results provided by knocked out mice, however, suggest that the steady state release of NO from endothelium is not governed by such a mechanism. A recent study implicated TRPV4 in tumor angiogenesis, as Florio Pla and coworkers reported an increase in TRPV4 levels in breast tumor ECs (b-TECs) as compared to control cells belonging to a human dermal microvascular endothelial cell line (HMVEC). In particular, AA-dependent TRPV4-mediated Ca2+ influx selectively drives cell migration via remodeling of the actin cytoskeleton in b-TECs, but not in HMVECs[319].

TRPM channels

The melastatin-related TRP subfamily was named based on the first discovered member, melastatin 1 (TRPM1), the gene of which was identified from melanomas. Members of the TRPM family are divided into four groups: TRPM1/3, TRPM2/8, TRPM4/5, and TRPM6/7[200,201]. TRPM channels exhibit highly varying cation permeability, from Ca2+ impermeable (TRPM4/5) to highly Ca2+ and Mg2+ permeable (TRPM6 and TRPM7)[202]. All the TRPM isoforms except TRPM5 are expressed in ECs (Table 4), although only the role of TRPM2 and TRPM7 has been extensively elucidated in these cells[270]. These channels are regarded as chanzymes, as they result from the fusion of an ion-channel pore-forming domain with an enzymatic domain[200].

TRPM2: TRPM2 is a non-elective cation channel whose predominant feature is the so-called Nudix box, a consensus region for pyrophosphatases that is localized in the cytoplasmic COOH-terminal tail of the channel protein and confers a unique activation mechanism, i.e., gating by ADPr, on the channel[144]. ADPr arises from breakdown of β-NAD, CD38, or other enzymes acting on cADPr and hydrolysis of ADP polymers by poly-ADP ribose glycohydrolase (PARP). TRPM2 is also activated by oxidative or nitrosative stress (e.g., H2O2)[4,144], perhaps mediated by mitochondrial ADPr. Therefore, TRPM2 serves as a sensor for the intracellular redox status by mediating oxidative stress-induced Ca2+ entry and triggering the subsequent Ca2+-dependent elevation in endothelial permeability[4,200,320]. This chain of reactions is suppressed by a truncated TRPM2 isoform (TRPM2-S), generated by alternative splicing of the full-length protein (TRPM2-L)[144]. Moreover, TRPM2 may underlie neutrophil-elicited increase in endothelial [Ca2+]i, which is a prerequisite for transendothelial migration and, consequently, for sepsis-induced increase in lung vascular permeability[321]. The NH2-terminal domain possesses a calmodulin-binding site which renders the channel sensitive to changes in sub-plasmalemmal Ca2+, which shifts the dose-response curve to ADPr to the left[144]. In addition, the NH2-terminal cytosolic tail of TRPM2 also contains a high affinity binding site for PKCα, which rapidly colocalizes with TRPM2-S in ECs challenged with H2O2[4,321]. PKCα might, thus, increase TRPM2-mediated Ca2+ influx and increase in endothelial permeability by phosphorylating TRPM2-S[4,321].

TRPM4: TRPM4, which is expressed as two splice variants, TRPM4a (non functional) and TRPM4b (functional, traditionally referred to as TRPM4), which is selective for monovalent cations, with an order of permeability Na+ > K+ > Cs+ > Li+, but not for Ca2+[200,202]. TRPM4 is present in both intact vessels and cultured ECs (Table 4) and is gated by cytosolic Ca2+ with a KD ranging from 0.4 to 9.8 μmol/L[322]. Therefore, TRPM4 represents the long-sought Ca2+-activated nonselective cation channels[200]. A short stretch of six acidic amino acids in the pore loop determines its monovalent selectivity[202]. We refer the reader to other recent and comprehensive reviews for the molecular details of TRPM4 modulation in cell types other than ECs[322,323]. Briefly, the Ca2+-dependent activation is regulated by ATP, Ca2+-calmodulin, and PKC. The current-voltage exhibits a large outward rectification, indicative of a voltage-dependent behavior which is further enhanced by increasing the temperature between 15 and 35 °C. PIP2 also modulates channel activity by influencing its Ca2+ and voltage sensitivity[322]. The main function attributed to TRPM4 in ECs is that, when activated, it protects against Ca2+ overload by depolarizing the cell, thereby decreasing the driving force for Ca2+ entry. It has, however, recently been shown that TRPM4 plays a crucial role in promoting EC death. TRPM4 is upregulated in capillary ECs of injured spinal cord, which leads to microvessel fragmentation, formation of satellite (petecchial) hemorrhagic lesions, and secondary hemorrhages[324]. In addition, TRPM4 mediates LPS-induced EC death by a yet to be unveiled mechanism[325].

TRPM7: TRPM7 is a nonselective cation channel that exhibits moderate permeability to Mg2+ (albeit at very low conductances) and is, therefore, involved in intracellular Mg2+ homeostasis[200]. In addition, TRPM7 is permeable to Zn2+, Co2+, and Mn2+, thereby providing a membrane pathway for metal ions[323]. TRPM7 might form either homo- or heteromultimers with its close homolog TRPM6, and has been found both in in situ and in vitro ECs (Table 4). Similar to TRPM2, TRPM7 serves as a chanzyme, its primary structure resulting from the fusion of a cation-selective ion channel pore with a COOH-terminal region possessing homology to a unique family of serine/threonine kinase domains, named the α-kinase family[200,326]. The enzymatic activity of the α-kinase domain is not required for pore opening, although evidence exists that it may have an alternative means of influencing channel gating[323,326]. The mechanism underlying the constitutive activation of the channel is, therefore, still unresolved[323]. The phosphotransferase activity of the kinase domain is, nevertheless, able to respond to local changes in free Mg2+ occurring as the result of Mg2+ flux through the channel. TRPM7 might, therefore, act as “Mg2+-sensor”[323,326]. Consistently, silencing TRPM7 mimics the effect of lowering extracellular Mg2+ in human microvascular ECs by impairing cell migration and arresting the cells in G0/G1 and G2/M phases[327]. Similarly, TRPM7 may inhibit EC proliferation via an extracellular signal-regulated kinase (ERK)-dependent reduction in eNOS levels[328]. This finding has been supported by in vivo observations performed in mice with inherited hypomagnesemia, TRPM7 upregulation is associated with endothelial dysfunction consequent to a decrease in eNOS levels and an increase in systolic blood pressure, to an augmented expression of vascular cell adhesion molecule-1 and plasminogen activator inhibitor-1, and to the downregulation of annexin-1, a downstream target of TRPM7[329]. Unlike VSMCs, the endothelial TRPM7 is insensitive to laminar shear stress[330].

TRPP channels or polycystins

The TRPP family consists of eight members, only two of which, i.e., polycystic kidney disease 2 (PKD2 or TRPP2) and PKD2-like (or TRPP3), form Ca2+-permable channels[200,323]. TRPP2 may, however, assemble with the non-channel protein, TRPP1 (or PKD1), in order to form a heterocomplex able to transduce laminar shear stress in intracellular [Ca2+]i signals[200,323]. Both TRPP1 and TRPP2 have been found in cultured and in situ vascular ECs (Table 5). TRPP1 is an integral membrane glycoprotein with 11 transmembrane domains, an extensive NH2-terminal extracellular tail involved in cell-cell, cell-matrix interaction and signaling pathways, and a cytosolic COOH-terminal endowed with a coiled-coil domain[331]. TRPP1 interacts with TRPP2, a six-transmembrane domain protein of the TRP ion channel family[200,323]. Polycystins also interact with multiple partners, including the TRP channel subunits TRPC1 and TRPV4, as well as several elements of the cytoskeleton[331]. By the coiled-coil COOH-terminal domain, TRPP2 and TRPP1 can form a functional polycystin complex in the apical cilia of both cultured and native ECs[332,333]. Due to its large extracellular NH2-terminal tail, TRPP1 has been suggested to be a mechanical sensor regulating the opening of the associated calcium-permeable channel TRPP2[323]. Shear-stress-induced and TRPP1/TRPP2-dependent Ca2+ entry, in turn, triggers NO production by recruiting PKC, Ca2+-calmodulin and the PI3K/Akt pathway[332,333]. Moreover, TRPP2 may aid TRPC1 in mediating Ca2+ influx, NO production, and actin reorganization in bEnd3 cells exposed to mechanical stress[229], as aforementioned.

Table 5 Pattern of expression of transient receptor potential ankyrin and polycystic channels in vascular endothelial cells as detected by reverse transcription/polymerase chain reaction, immunohistochemistry, and Western blotting.
TRPA1TRPP1TRPP2
PCa/PNa and conductance (pS)[202]0.8-1.4, NM1-5, 40-1774, 137
Human renal artery ECs[332]+(RT-PCR, WB, IC)
Human umbilical vein ECs[332]+(RT-PCR, WB)
Mouse aortic ECs[332,333]+(RT-PCR, WB, IC)+(RT-PCR, WB)
Mouse femoral artery ECs[332]+(RT-PCR, WB)
Rat cerebral arteries[287,304,311]+(RT-PCR, IC)+(RT-PCR)
TRPA1 channels

TRPA1 is the sole member of the TRPA family, whose structural hallmark is represented by the 14 ankyrin repeats at its NH2 terminus[200,323]. These repeats might be required for the channel to interact with cytoskeletal components or to bind to specific agonists[323]. In addition, the NH2 terminus possesses a Ca2+-binding EF-domain[323]. TRPA1 homotetramers assemble to form a nonselective cation channel displaying a relatively high Ca2+ permeability when compared with other TRP channels, including TRPV1 (Table 5). TRPA1 may be activated by electrophilic compounds such as acrolein (an active component of tear gas), allicin (found in garlic), and allyl isothiocyanate (AITC) (derived from mustard oil). α, β-Unsaturated aldehydes produced endogenously in response to oxidative stress, such as 4-hydroxy-2-nonenal, 4-oxo-nonenal, and 4-hydroxyhexenal, also activate the channel[323]. Most identified TRPA1 agonists act via covalent modification of cysteines in the NH2 termini, with specific binding residues identified[323]. In addition, TRPA1 might open in response to noxious cold (< 17 °C), but the thermosensitivity of the channel is still debated[323]. TRPA1 has been detected in native ECs, localized to endothelial cell membrane projections through the internal elastic lamina in the direction of VSMCs. TRPA1-mediated Ca2+ influx causes endothelium-dependent SMC hyperpolarization and relaxation of cerebral arteries by activating IKCa and SKCa[304]. IKCa- and SKCa-gated K+ efflux, in turn, activates KIR in SMCs, thus amplifying myocyte relaxation and dilation[304].

Stim1 and Orai1

The physical interaction between the ER Ca2+-sensor, Stim1, and the plasmalemmal, Ca2+-permeable channel, Orai1, has recently been shown to mediate SOCE in a number of vascular ECs[205,206,243,334]. Stim1 is an ER-resident type I single transmembrane protein with an NH2 terminus located either in the ER lumen or facing the extracellular space. It contains two NH2-terminal EF hand domains (a canonical and a hidden one) followed by a sterile α-motif (SAM) domain, the transmembrane region, three predicted coiled-coil domains comprising an ezrin-radixin-moesin (ERM) motif and at the COOH terminus, a proline-rich and a lysine-rich domain[335,336]. The negatively charged NH2-terminal EF hand domain is the Ca2+ sensing subunit of the protein. In the resting state, Stim1 seems to be present as a dimer in the ER membrane. The dimerization happens through the COOH-terminal coiled-coil domains. In case of store release, and subsequent Ca2+ dissociation from the canonical EF hand domain, further oligomerization of Stim1 through the SAM domain occurs. These Stim1 oligomers migrate towards sites of the ER that are in close vicinity to the plasma membrane, where they assembly into punctuate clusters and activate SOCE by directly binding to Orai1[335,336]. Accordingly, the latter has been found to be an essential component of SOCE by forming the pore unit of the channel in a variety of cell types[335,337], including vascular ECs[205,206,338]. Orai1 is a four-transmembrane protein with intracellular NH2 and COOH termini. In the resting state, it is present as a dimer in the plasma membrane, however, the functional SOC channel seems to be formed by Orai1 tetramers. The interaction between Stim1 and Orai1 is mediated by the CRAC activation domain (CAD) located at the COOH terminus of Stim1 and the cytosolic NH2 and COOH tails of Orai1[335,336]. Accordingly, the genetic suppression of either Orai1 or Stim1 suppresses SOCE induced by either pharmacological (CPA and thapsigargin) and physiological (VEGF, thrombin) stimulation in HUVECs, EA.hy296 cells, and porcine aortic ECs[205,206,218,243,334]. This, in turn, leads to the blockade of several key steps of the angiogenic process, such as proliferation and tubulogenesis[205,206,338]. These data are supported by the finding that, under resting conditions, the ER Ca2+ sensor is homogeneously distributed throughout the cytosol, whereas emptying of the InsP3-sensitive Ca2+ reservoir in the absence of extracellular Ca2+ induces Stim1 homo-oligomerization and redistribution into sub-plasmalemmal puncta[206,218,339]. Both processes are recovered upon Ca2+ restoration to the bathing solution, as Ca2+ entry recharges the intracellular Ca2+ pool and inactivates Stim1[340]. Stim1 puncta in vascular EC reassembly at the same loci during repetitive cycles of ER Ca2+ depletion and refilling suggest that this process is driven by accessory proteins, such as the microtubule-plus-end tracking protein, EB1[335,336,340]. The engagement of Stim1 is, however, limited when ECs are stimulated in the presence of extracellular Ca2+, which is consistent with modest reduction in the ER Ca2+ load observed under these conditions[40,41,340]. An important aspect of Stim1 dynamics in ECs regards its modulation by cytosolic Ca2+: an increase in submembranal Ca2+ levels prevents Stim1 clustering, thereby impeding prolonged SOC activity and intracellular Ca2+ overload[340]. This feature might contribute to explain the long-known inhibition of SOCE by Ca2+ microdomains arising at the mouth of the channel[341]. SOCE mediated by Stim1 and Orai1 is gated by an inwardly rectifying, Ca2+-selective current[205], whose amplitude may fall below the resolution limit of a whole-cell patch-clamp recording system[206]. Orai1 has two homologs in mammals, namely Orai2 and Orai3, which do not contribute to SOCE in ECs[205,218,338]. In contrast to Stim1, Stim2 has no major effect on SOCE, but might be involved in the regulation and stabilization of basal cytosolic and ER Ca2+ levels in ECs[342]. The involvement of Stim1 and Orai1 to endothelial SOCE is, however, still controversial and might depend on both the vascular bed and the cell type. For instance, recent work ruled out any detectable role for Orai1 in thrombin- and thapsigargin-induced Ca2+ inflow in mouse lung ECs[218], where SOCE activity requires functional TRPC1 and TRPC4[218]. The same study reported that, in these cells, Stim1 is mobilized into sub-plasmalemmal puncta in a TRPC4-dependent manner and, in turn, it signals TRPC1 and TRPC4 about the ER Ca2+ content.

Ca2+ EXTRUSION FROM THE CYTOSOL

A variety of pumps and exchangers fulfil the function to remove Ca2+ from the cytosol following EC stimulation, namely SERCA, NCX, PMCA, and the mitochondrial uniporter[36,40,41,139,343-349]. SERCA, NCX, PMCA, mitochondria may be aligned either in series or in parallel in ECs in order to remove Ca2+ from the cytosol properly, to achieve successful ER Ca2+ refilling, and facilitate SOCE by buffering sub-plasmalemmal Ca2+. These functions are facilitated by their optimal coupling to either InsP3Rs or RyRs.

Expression of SERCA, NCX and PMCA in ECs

SERCA is encoded by three different genes (ATP2A1, ATP2A2, and ATP2A3) with numerous alternative splice isoforms at the 3′ ends of the mRNA. More specifically, ATP2A1 encodes for SERCA1a-1b, ATP2A2 for SERCA2a-2c, and ATP2A3 for SERCA3a-3f[350]. The SERCA isoforms differ mainly by their affinity for Ca2+ (2b > 2a = 1 >> 3) and their Ca2+ transport turnover rates, SERCA2b having the lowest transport capacity of all SERCAs[350]. Both SERCA2a and SERCA3 have been found in isolated and in situ ECs[106,351-356], albeit SERCA3 transcript levels decrease during EC proliferation in culture and in hypertension[106,112]. SERCA3 expression is regulated by Ca2+ itself in a calcineurin/NFAT-dependent manner[357] and its genetic deletion leads to a decrease in agonist-induced elevation in [Ca2+]i and NO synthesis[358]. The Ca2+-ATPase activity in vascular endothelium may be enhanced by NO, which considerably increases Ca2+ loading into ER lumen[359]. Moreover, ECs may be endowed with phospholamban (PLB), a 24- 27-kDa phosphoprotein that is closely associated with the cardiac sarcoplasmic reticulum, where it modulates SERCA function[360]. More specifically, PLB inhibits Ca2+ uptake by SERCA in Ach-stimulated cells and dampens the endothelial-dependent vasorelaxation[360]. In addition, SERCA activity is more resistant to peroxynitrite and peroxide in ECs as compared to VSMCs[361,362]; a feature that may confer a protective effect during the massive production of free radicals observed under pathological conditions, such as ischemia-reperfusion. Conversely, selective targeting of thiol groups (i.e., Cys residues) by myeloperoxidase-derived oxidants may inhibit the Ca2+-ATPase activity and lead to an unbalanced increase in [Ca2+]i, which might contribute to the endothelial dysfunction occurring during several inflammatory diseases[363]. In this view, oxidized and glycated low-density lipoprotein causes SERCA oxidation, aberrant ER stress and atherosclerosis in vivo, although this detrimental chain of events is suppressed by AMPK[364,365].

The Na+/Ca2+ exchanger family comprises three members, namely NCX1, NCX2, and NCX3, which act by expelling one Ca2+ ion for the uptake of three Na+ ions (the so-called “forward-mode” of NCX). The electrochemical gradient for Na+ provides the energy necessary to drive Ca2+ out of the cell against its large gradient and with a stoichiometric ratio of 3 Na+ in: 1 Ca2+ out[366]. Cultured ECs may express all of the three NCX subtypes[355,367,368], while the only evidence about NCX presence in situ has been reported in rat brain[369]. NCX expression may be upregulated by an increase in [Ca2+]i[370]. In addition to extruding Ca2+ out of the cytosol, NCX may contribute to Ca2+ entry when operating in the so-called “reverse-mode” (3 Na+ out: 1 Ca2+ in). More specifically, the close proximity between NCX and the TRPC channels, which may either physically interact[371] or be clustered in limited membrane nanodomains[372], causes NCX to switch into the reverse-mode upon Na+ accumulation beneath the plasma membrane[371,372]. Intriguingly, NCX-mediated Ca2+ entry in ECs may be stimulated by Ach[373], histamine[43], and VEGF[374], but not by ATP[375]. In particular, NCX-driven Ca2+ influx underpins PKCα translocation to the plasma membrane, where it stimulates ERK1/2 to trigger a number of steps (proliferation, migration, and tubulogenesis) involved in VEGF-dependent angiogenesis[374]. It is, therefore, conceivable that NCX role in the regulation of Ca2+-dependent processes in vascular endothelium may be governed by the incoming stimuli.

PMCA is encoded by four genes, termed PMCA1-4. Alternative splicing of their primary transcripts yields a large variety of PMCA proteins that differ in their kinetics and regulation[350]. The pump operates with high Ca2+ affinity and low transport capacity, with a 1:1 Ca2+/ATP stoichiometry[350]. PMCA1 is the main isoform found in vascular endothelium, although lower levels of PMCA2 and PMCA4 may also be present[355,376]. PMCA-dependent Ca2+ clearance is triggered by local elevation in intracellular Ca2+, which acts by promoting the association between CaM and a specific CaM-binding domain located at the COOH terminus of the pump[348,377]. PMCA activity and expression is negatively regulated by endothelin-1, which may thus lead to an increase in [Ca2+]i by reducing Ca2+ clearance from the cytosol[378]. PMCA is preferentially located at caveolae on the luminal surface of ECs, and its activity is reduced by phosphatidylserine externalization[379,380]. In addition to extruding Ca2+ across the plasma membrane, PMCA suppresses both eNOS activity and the subsequent NO production by promoting Thr-495 phosphorylation in resting and activated ECs[381]. PMCA activity in vascular ECs is, in turn, inhibited by S-glutathionylation, which leads to a progressive raise in [Ca2+]i[124]. The elucidation of the role served by PMCA in endothelial signaling might benefit from the recent discovery of caloxin 1b3, a specific PMCA1 inhibitor[382].

Mitochondrial Ca2+ uptake

Mitochondria contribute to the dynamics of Ca2+ signaling by aiding SERCA, NCX, and PMCA in removing Ca2+ from the cytoplasm during the recovery phase. Mitochondrial Ca2+ release mechanisms, then, return Ca2+ back into the cytosol, where it can be sequestered by SERCA. Indeed, there is a close functional relationship between mitochondria and ER[10]. This interaction is reinforced by the chaperon glucose-regulated protein 75 (grp75), which tethers ER InsP3Rs to the voltage-dependent anion channels (VDACs) in the outer mitochondrial membrane. VDACs enable the transfer of cytosolic Ca2+ into the mitochondrial intermembrane space, from where the electrochemical gradient (the inner mitochondrial membrane potential is equal to about -180 mV) drives Ca2+ into the mitochondrial matrix mainly by two transport mechanisms[383]. The mitochondrial Ca2+ uniporter (MCU), which is located in the inner mitochondrial membrane, operates at the micromolar Ca2+ concentrations that are only transiently reached in cells near Ca2+ release channels, and might be contributed by uncoupling proteins 2 and 3 (UCP2 and UCP3) in ECs[384]. However, UCP2/3 buffer intracellularly mobilized Ca2+, but not that entering the cells via SOCs[385]. Mitochondrial Ca2+ uptake from SOCE in ECs is accomplished by a second type of MCU, which mediates Ca2+ uptake into mitochondria at nanomolar cytosolic Ca2+ concentrations and has been termed Letm1[383]. This is a leucine-zipper-EF hand-containing a TM region which catalyzes the 1:1 electrogenic exchange of Ca2+ for H+[386]. Conversely, the contribution of the so-called mitochondrial calcium uptake 1 (MICU1), which has recently been discovered by working on HeLa and HEK-293 cells[387], to endothelial Ca2+ signaling has not been ascertained yet The mitochondrial NCX (mNCX) is the main mechanism responsible for the outward transport of Ca2+ towards the cytosol with a stoichiometry of 3Na+:1Ca2+ (as for the plasmalemmal NCX)[36,41]: the Na+ that enters down the electrochemical gradient is exchanged for Ca2+[383].

Microarchitecture of the Ca2+ transporting systems in ECs

The microstructural physical localization of Ca2+ pumps, transporters and channels in the ER and plasma membrane relative to each other determines the pattern of Ca2+ clearance from the endothelial cytosol[36,40,41,138,139,345,388]. This organization of the Ca2+ transporting systems creates Ca2+ gradients between the sub-plasmalemmal region and bulk cytoplasm, which, in turn, enables the same second messenger, i.e., Ca2+, to regulate diverse cellular functions depending on its proximity to specific Ca2+-sensitive effectors[10]. The sequence of events leading to removal of cytosolic Ca2+ may be different in ECs from different vascular beds and three main Ca2+-buffering pathways have hitherto been identified. (1) In-series arrangement of RyRs and NCX and of SERCA and RyRs[139,345]: according to this model, that has been described in rabbit aortic ECs, cytosolic Ca2+ is captured by SERCA, released from the ER via RyRs, and extruded across the plasma membrane via the NCX. PMCA acts in parallel to this pathway to bind to and deliver Ca2+ to the extracellular space; (2) Vectorial Ca2+ release from sub-plasmalemmal InsP3Rs to the extracellular space via NCX[389]: this system operates following either a moderate or a slow increase in InsP3 levels in porcine coronary artery ECs, prevents subplasmalemmal Ca2+ from reaching the deeper cytosol by triggering CICR on RyRs, and has, therefore, been termed SCCU. The SCCU has also been found in EA.hy926 cells, where it subserves a signaling role by activating the endothelial Ca2+-dependent large conductance K+ channels (BKCa) after cell stimulation by low concentrations of histamine[140,141]. An increase in the agonist dose overcomes the moderate RyRs-dependent subplasmalemmal Ca2+ release due to the higher InsP3-dependent Ca2+ mobilization; and (3) Mitochondrion-mediated ER refilling: this pathway redirects entering Ca2+ through SOCs to the ER in the presence of an InsP3-generating agonist[36,41,388]. SOCE is buffered by sub-plasmalemmal mitochondria and vectorially transferred via mNCX to SERCA. When SOCE occurs in the absence of InsP3, the superficial domains of the ER mimic mitochondrial Ca2+ buffering by immediately capturing entering Ca2+. This Ca2+-buffering mode has been mainly characterized in EA.hy926 cells, where it coexists with the SCCU system described above. The former is responsible for the extrusion of Ca2+ entering into the cells via SOCE, whereas the latter removes from the cytosol the intracellularly-released calcium ions. The mitochondria-mediated ER recharging might also be present in other vascular beds, but this mechanism has not been thoroughly investigated in mature endothelium.

The endothelial Ca2+-buffering system may, however, operate in parallel, as depicted by a number of studies conducted on microvascular endothelial cells. In rat brain capillary ECs, for instance, the SERCA provides the major pathway responsible for clearing cytosolic Ca2+, whereas NCX extrudes across the PM a minor percentage (around 30%) of the calcium ions mobilized from ER[343]. Conversely, in rat cardiac microvascular cells, PMCA and SERCA have been shown to restore endothelial Ca2+ to prestimulation levels during exposure to the agonist. However, upon interruption of the stimulation, SERCA quickly recaptures Ca2+ back into the ER lumen[349].

SPATIOTEMPORAL ORGANIZATION OF INTRACELLULAR Ca2+ SIGNALS IN ECS

The remarkable versatility of Ca2+ is due to the multitude of Ca2+ signals that can be generated by extracellular stimuli within a single cell. Such heterogeneity depends on the recruitment of specific components of the Ca2+ toolkit which lead to the generation of intracellular signals varying in amplitude, spatial dimension and frequency[10,390,391]. The opening of Ca2+-permeable pathways either on the ER or plasma membrane may give rise to local or elementary Ca2+ signals, which are spatially restricted within hundreds of nanometers to micrometers from the mouth of the channel pore[392]. The local nature of these events is attributable to the concerted action of Ca2+ buffering and reuptake mechanisms that limit the diffusion of the Ca2+ signal[37]. The most simple Ca2+ release events are the Ca2+ blips and the Ca2+ quarks, which arise from the opening of single InsP3Rs and RyRs, respectively[10,392]. Endothelial Ca2+ blips have average amplitude of 23 nmol/L, are 1- to 3-μm wide, and last < 100 ms[393]. Conversely, there is no report of endothelial Ca2+ quarks. The concerted opening of a fixed cluster of InsP3Rs and RyRs give rise, respectively, to Ca2+ puffs and Ca2+ sparks[10,392]. Spontaneous Ca2+ puffs have been reported in native ECs[394-396], which can be excited by adjoining SMCs through gap junctional communication[396-398]. Isolated ECs display only agonist-induced Ca2+ puffs, which reach a peak of about 50 nmol/L and spread for no longer than 30 μm[393]. Elementary Ca2+ release events mediated by InsP3Rs, but different from the classic Ca2+ puffs, may occur in restricted spaces of myoendothelial projections and have been termed Ca2+ pulsars[16,399]. The endothelial Ca2+ pulsars colocalize with the intermediate conductance, Ca2+-sensitive K+ channels (KCa3.1) so that local Ca2+ signals may be translated into a negative shift in VM which is rapidly transmitted to the adjacent VSMCs in terms of EDHF[399]. Consistently, sympathetic nerve stimulation activates α-adrenergic receptors on SMCs to recruit new pulsar sites in the adjoining ECs in order to oppose vasoconstriction[400]. The spatiotemporal summation of adjacent Ca2+ puffs leads to a regenerative intracellular Ca2+ wave that can spread throughout the cytosol at an average speed of 5-60 μm/s[118,393,401]. The regenerative propagation of the Ca2+ elevation is, indeed, faster in the perinuclear region and at the initiation site[118]. The endothelial Ca2+ sweep tends to originate from multiple peripheral foci localized in the caveolae compartments of the plasma membrane[402]. These triggering sites remain constant during repetitive cell stimulation[393,401,402]. Indeed, when the information embedded within the [Ca2+]i increase must be conveyed over a long time, both the elementary events and the global Ca2+ sweeps can adopt an oscillatory pattern[16,72,339,396]. However, the physiological outcome of local Ca2+ spikes depends on the Ca2+-sensitive decoders located within a few nanometers from the channel pore[395], and the frequency of repetitive Ca2+ waves encodes for the selective engagement of specific downstream targets[403-405]. The interspike interval of the Ca2+ transients is inversely correlated to the strength of the stimulus until, at least at high agonist doses, they apparently fuse to generate a plateau level[16,72,406,407]. Endothelial Ca2+ oscillations may, indeed, be elicited by mechanical forces, extracellular autacoids or oxidative stress. Shear stress and pressure may induce ECs to undergo repetitive Ca2+ transients by triggering Ca2+ entry through Gd3+-sensitive MSCs followed by CICR from RyRs[22,24,25]. The opening of the MSC generates a transient local Ca2+ event that has been termed Ca2+ hotspot[24]. Unlike the oscillatory Ca2+ response to extracellular ligands[72,406,407], the amplitude of the Ca2+ spikes increases with the intensity of the mechanical stimulation[22-25]. Agonist-evoked Ca2+ oscillations require PLC activation by either GPCRs or TKRs[72,339,404,405,408]. The following InsP3-dependent Ca2+ mobilization may be amplified by adjacent RyRs via CICR[44,134] and requires Ca2+ inflow to supply releasable Ca2+ over time and maintain InsP3R sensitivity to the ligand[72,407,409,410]. Ca2+ entering through either SOCE or the reverse-mode of NCX is sequestered by SERCA into ER lumen before being released again into the cytosol through the open InsP3Rs[72,134,339,404,405]. Some[16], but not all[72], microvascular ECs may be independent on Ca2+ inflow. In addition to bringing about the downstroke of each Ca2+ spike, SERCA shapes both the rate of the initial Ca2+ release and the amplitude of the regenerative upstroke, whereas it does not affect the Ca2+ wave speed[339,411]. As a consequence of the rhythmic interplay between InsP3-induced Ca2+ release and SERCA-mediated Ca2+ reuptake, SOCE is not constant during the Ca2+ train. Accordingly, store-sensitive Ca2+ inflow is activated by the partial depletion of the Ca2+ pool occurring during each Ca2+ spike and is then turn off by store replenishment[410,412]. Finally, ECs may display repetitive Ca2+ oscillations in response to the NAD(P)H oxidase-derived H2O2 that is endogenously produced during the ischemia/reperfusion injury[56,413]. H2O2-elicited Ca2+ spikes are supported by the periodic opening of InsP3Rs, but rapidly run down in absence of extracellular Ca2+[56,413]. It would, therefore, be interesting to assess whether the H2O2-sensitive plasmalemmal channel, TRPM2, is involved in the signal transduction pathways recruited under these conditions[320].

CONCLUSION

Once regarded as an inert barrier between circulation and vascular tissue, ECs constitute the largest signal transduction platform of the organism, which is essential for the proper functioning of the cardiovascular system. Any alteration in the endothelial capability to detect and decode both physical and humoural stimuli into the most suitable vessel wall phenotype may lead to an astonishing variety of severe diseases, including myocardial infarction, brain stroke, atherosclerosis, peripheral artery disease, and tumor. As recently outlined by Munaron and Florio Pla[11], it is imperative to understand the molecular mechanisms that enable vascular ECs to fulfil their multifaceted role in order to either prevent or reduce the adverse consequences of so-called endothelial dysfunction. The Ca2+ signalling toolkit plays a key role in the control of a wide array of endothelial processes and, therefore, any change in the pattern of expression of each of its components is somehow involved in the pathogenesis of an increasing number of pathologies. Despite the most recent advances in the comprehension of the molecular underpinnings of the Ca2+ responses to both physical and chemical stimulation, a huge amount of work is standing in front of us. First, the blend of endothelial ion channels, transporters and receptors may vary with species, age, sex, hormonal status, vascular bed, and blood vessel diameter. This issue has been further stressed by the recent finding that Stim1 and Orai1 control in vitro tubulogenesis in HUVECs[206], but not in the HUV-derived EC line, EA.hy926[224]. As a consequence, the information on the Ca2+ machinery acquired on animal or cell line models cannot undergo a straightforward clinical translation, but must be confirmed in humans and in the vascular region implicated in the disease under investigation. This task may not be easy due to the difficulties of obtaining human samples from either healthy or pathological vessels. Second, the controversy on the molecular nature of SOCs, which are regarded as the most important pathway for Ca2+ influx into vascular endothelium, is far from being fully resolved. As described above, a recent study suggested that Orai1 and Stim1, whose physical interaction mediates SOCE in the immune system, underlie this Ca2+ entry route only in HUVECs, whereas TRPC1 and TRPC4 are activated by the emptying of the Ca2+ pool in other vascular beds. The same study, however, contributed to boost such controversy by demonstrating that the genetic suppression of TRPC1 dampens SOCE in HUVECs, a finding which is not consistent with those previously reported on the same preparation. This conflict is not surprising when considering the abundance of discrepant data about TRPC1 modulation by Stim1. The possibility that the molecular structure of SOCE (i.e., either the ER Ca2+ sensor and the Ca2+-permeable pore) might differ in distinct species and vascular beds, as put forward by Sundivakkam et al[218], should be also considered. Moreover, the possibility that the components of SOCE may be different in distinct phases of the cell cycle should be taken in account. Yet, it is evident that further studies will have to be devoted to clarify this issue in vascular ECs. A recent study attempted to solve this controversy by investigating the association between Orai1 and TRPC1, which has been described in a growing number of nonexcitable cells, in rat pulmonary artery ECs. In these cells, SOCE has long been associated with TRPC1 and TRPC4, as widely described above[217]. Surprisingly, Cioffi et al[414] have found that Orai1 constitutively interacts with TRPC4 and drives the activation of the TRPC1-TRPC complex upon intracellular Ca2+ release. It is not clear, however, whether and how protein 4.1 and Stim1 independently confer store-sensitivity to the ternary complex on the plasma membrane. Third, whereas the intracellular target of InsP3 has long been recognized, the selective receptor of the two newly discovered second messengers, cADPr and NAADP, has recently been questioned. Indeed, a couple of recent investigations demonstrated that the genetic suppression of TPC1-2 does not prevent cell labeling by NAADP[415,416]. Conversely, the same studies suggested that a 22- and 23-kDa pair of accessory proteins might serve as NAADP receptors and regulate the activity of its target channels, including TPC1-2 in ECs and RyRs in the immune system. The same controversy holds true also for the molecular effector of cADPr-dependent signalling in mature ECs. As mentioned already, the mechanistic link between cADPr and RyRs-mediated Ca2+ is yet to be elucidated. Fourth, ECs establish either transient or permanent intimate contacts with a variety of cell types, such as circulating blood cells or tumoral cells, fibroblasts, and SMCs. Only a few studies have hitherto addressed whether and how these tight relationships may affect endothelial Ca2+ homeostasis. In addition, ECs may undergo a rearrangement of their Ca2+ signaling proteins when detached from the vessel walls and cultured in vitro[7,153]. It follows that future studies should be conducted both in vitro and in vivo, by exploiting the genetic Ca2+-sensors that have been described in the literature, to explore this fascinating aspect of Ca2+ signaling. Finally, intracellular Ca2+ waves have recently been shown to drive EPC proliferation, homing, and tubulogenesis. Only a few investigations have so far been devoted to unmask the players of the Ca2+ machinery involved in these tasks. In the light of the promising therapeutic potential of the Ca2+-permeable channels in these cells, either as a target or facilitators of clinical interventions, we are confident that this field will gain the interest of researchers worldwide. Therefore, it appears that the way towards full comprehension of the intricate network of membrane receptors, ion channels, transporters, pumps, and cytosolic decoders that convert an extracellular stimulus into a meaningful Ca2+ signal still has a long way to go and overcome many obstacles, but a multidisciplinary approach will certainly be helpful to reach this goal.

ACKNOWLEDGMENTS

The authors are grateful to all their collaborators for the ongoing and fruitful discussions on Ca2+ signalling in endothelial cells and endothelial progenitor cells. They are also delighted to acknowledge Mrs Simona Laudadia for her excellent bibliographic assistance.

Footnotes

Peer reviewers: Saobo Lei, Associate Professor, Department of Pharmacology, Physiology and Therapeutics, University of North Dakota, 501 N Columbia Road, Grand Forks, ND 58203, United States; Claudio F Perez, PhD, Department of Anesthesia, Brigham and Women’s Hospital, GW Thorn Building, TH-726B, 20 Shattuck Street, Boston, MA 02115, United States

S- Editor Cheng JX L- Editor Kerr C E- Editor Zhang DN

References
1.  Tran QK, Watanabe H. Calcium signalling in the endothelium. Handb Exp Pharmacol. 2006;145-187.  [PubMed]  [DOI]  [Cited in This Article: ]
2.  Félétou M, Vanhoutte PM. Endothelial dysfunction: a multifaceted disorder (The Wiggers Award Lecture). Am J Physiol Heart Circ Physiol. 2006;291:H985-1002.  [PubMed]  [DOI]  [Cited in This Article: ]
3.  Khazaei M, Moien-Afshari F, Laher I. Vascular endothelial function in health and diseases. Pathophysiology. 2008;15:49-67.  [PubMed]  [DOI]  [Cited in This Article: ]
4.  Di A, Malik AB. TRP channels and the control of vascular function. Curr Opin Pharmacol. 2010;10:127-132.  [PubMed]  [DOI]  [Cited in This Article: ]
5.  Fiorio Pla A, Avanzato D, Munaron L, Ambudkar IS. Ion channels and transporters in cancer. 6. Vascularizing the tumor: TRP channels as molecular targets. Am J Physiol Cell Physiol. 2012;302:C9-C15.  [PubMed]  [DOI]  [Cited in This Article: ]
6.  Santella L, Lim D, Moccia F. Calcium and fertilization: the beginning of life. Trends Biochem Sci. 2004;29:400-408.  [PubMed]  [DOI]  [Cited in This Article: ]
7.  Nilius B, Droogmans G. Ion channels and their functional role in vascular endothelium. Physiol Rev. 2001;81:1415-1459.  [PubMed]  [DOI]  [Cited in This Article: ]
8.  Yao X, Garland CJ. Recent developments in vascular endothelial cell transient receptor potential channels. Circ Res. 2005;97:853-863.  [PubMed]  [DOI]  [Cited in This Article: ]
9.  Mancardi D, Pla AF, Moccia F, Tanzi F, Munaron L. Old and new gasotransmitters in the cardiovascular system: focus on the role of nitric oxide and hydrogen sulfide in endothelial cells and cardiomyocytes. Curr Pharm Biotechnol. 2011;12:1406-1415.  [PubMed]  [DOI]  [Cited in This Article: ]
10.  Berridge MJ, Bootman MD, Roderick HL. Calcium signalling: dynamics, homeostasis and remodelling. Nat Rev Mol Cell Biol. 2003;4:517-529.  [PubMed]  [DOI]  [Cited in This Article: ]
11.  Munaron L, Fiorio Pla A. Endothelial calcium machinery and angiogenesis: understanding physiology to interfere with pathology. Curr Med Chem. 2009;16:4691-4703.  [PubMed]  [DOI]  [Cited in This Article: ]
12.  Tiruppathi C, Ahmmed GU, Vogel SM, Malik AB. Ca2+ signaling, TRP channels, and endothelial permeability. Microcirculation. 2006;13:693-708.  [PubMed]  [DOI]  [Cited in This Article: ]
13.  Nilius B, Viana F, Droogmans G. Ion channels in vascular endothelium. Annu Rev Physiol. 1997;59:145-170.  [PubMed]  [DOI]  [Cited in This Article: ]
14.  Chetham PM, Babál P, Bridges JP, Moore TM, Stevens T. Segmental regulation of pulmonary vascular permeability by store-operated Ca2+ entry. Am J Physiol. 1999;276:L41-L50.  [PubMed]  [DOI]  [Cited in This Article: ]
15.  Moccia F, Baruffi S, Spaggiari S, Coltrini D, Berra-Romani R, Signorelli S, Castelli L, Taglietti V, Tanzi F. P2y1 and P2y2 receptor-operated Ca2+ signals in primary cultures of cardiac microvascular endothelial cells. Microvasc Res. 2001;61:240-252.  [PubMed]  [DOI]  [Cited in This Article: ]
16.  Borisova L, Wray S, Eisner DA, Burdyga T. How structure, Ca signals, and cellular communications underlie function in precapillary arterioles. Circ Res. 2009;105:803-810.  [PubMed]  [DOI]  [Cited in This Article: ]
17.  Pupo E, Pla AF, Avanzato D, Moccia F, Cruz JE, Tanzi F, Merlino A, Mancardi D, Munaron L. Hydrogen sulfide promotes calcium signals and migration in tumor-derived endothelial cells. Free Radic Biol Med. 2011;51:1765-1773.  [PubMed]  [DOI]  [Cited in This Article: ]
18.  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. 2008;15:605-614.  [PubMed]  [DOI]  [Cited in This Article: ]
19.  Yin J, Kuebler WM. Mechanotransduction by TRP channels: general concepts and specific role in the vasculature. Cell Biochem Biophys. 2010;56:1-18.  [PubMed]  [DOI]  [Cited in This Article: ]
20.  Schwarz G, Callewaert G, Droogmans G, Nilius B. Shear stress-induced calcium transients in endothelial cells from human umbilical cord veins. J Physiol. 1992;458:527-538.  [PubMed]  [DOI]  [Cited in This Article: ]
21.  Shen J, Luscinskas FW, Connolly A, Dewey CF, Gimbrone MA. Fluid shear stress modulates cytosolic free calcium in vascular endothelial cells. Am J Physiol. 1992;262:C384-C390.  [PubMed]  [DOI]  [Cited in This Article: ]
22.  Kuebler WM, Ying X, Bhattacharya J. Pressure-induced endothelial Ca(2+) oscillations in lung capillaries. Am J Physiol Lung Cell Mol Physiol. 2002;282:L917-L923.  [PubMed]  [DOI]  [Cited in This Article: ]
23.  Ohata H, Ikeuchi T, Kamada A, Yamamoto M, Momose K. Lysophosphatidic acid positively regulates the fluid flow-induced local Ca(2+) influx in bovine aortic endothelial cells. Circ Res. 2001;88:925-932.  [PubMed]  [DOI]  [Cited in This Article: ]
24.  Ohata H, Yamada H, Momose K. Lysophosphatidic acid induces shear stress-dependent Ca2+ influx in mouse aortic endothelial cells in situ. Exp Physiol. 2011;96:468-475.  [PubMed]  [DOI]  [Cited in This Article: ]
25.  Hoyer J, Köhler R, Distler A. Mechanosensitive Ca2+ oscillations and STOC activation in endothelial cells. FASEB J. 1998;12:359-366.  [PubMed]  [DOI]  [Cited in This Article: ]
26.  Ando J, Yamamoto K. Vascular mechanobiology: endothelial cell responses to fluid shear stress. Circ J. 2009;73:1983-1992.  [PubMed]  [DOI]  [Cited in This Article: ]
27.  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. 2010;298:H466-H476.  [PubMed]  [DOI]  [Cited in This Article: ]
28.  Sobolewski P, Kandel J, Klinger AL, Eckmann DM. Air bubble contact with endothelial cells in vitro induces calcium influx and IP3-dependent release of calcium stores. Am J Physiol Cell Physiol. 2011;301:C679-C686.  [PubMed]  [DOI]  [Cited in This Article: ]
29.  Hawkins BJ, Solt LA, Chowdhury I, Kazi AS, Abid MR, Aird WC, May MJ, Foskett JK, Madesh M. G protein-coupled receptor Ca2+-linked mitochondrial reactive oxygen species are essential for endothelial/leukocyte adherence. Mol Cell Biol. 2007;27:7582-7593.  [PubMed]  [DOI]  [Cited in This Article: ]
30.  Lewalle JM, Cataldo D, Bajou K, Lambert CA, Foidart JM. Endothelial cell intracellular Ca2+ concentration is increased upon breast tumor cell contact and mediates tumor cell transendothelial migration. Clin Exp Metastasis. 1998;16:21-29.  [PubMed]  [DOI]  [Cited in This Article: ]
31.  Nejime N, Tanaka N, Yoshihara R, Kagota S, Yoshikawa N, Nakamura K, Kunitomo M, Hashimoto M, Shinozuka K. Effect of P2 receptor on the intracellular calcium increase by cancer cells in human umbilical vein endothelial cells. Naunyn Schmiedebergs Arch Pharmacol. 2008;377:429-436.  [PubMed]  [DOI]  [Cited in This Article: ]
32.  Wood PG, Gillespie JI. Evidence for mitochondrial Ca(2+)-induced Ca2+ release in permeabilised endothelial cells. Biochem Biophys Res Commun. 1998;246:543-548.  [PubMed]  [DOI]  [Cited in This Article: ]
33.  King J, Hamil T, Creighton J, Wu S, Bhat P, McDonald F, Stevens T. Structural and functional characteristics of lung macro- and microvascular endothelial cell phenotypes. Microvasc Res. 2004;67:139-151.  [PubMed]  [DOI]  [Cited in This Article: ]
34.  Uehara K, Onoue H, Jeyakumar LH, Fleischer S, Uehara A. Localization of ryanodine receptor 3 in the sinus endothelial cells of the rat spleen. Cell Tissue Res. 2004;317:137-145.  [PubMed]  [DOI]  [Cited in This Article: ]
35.  Uehara K. Localization of TRPC1 channel in the sinus endothelial cells of rat spleen. Histochem Cell Biol. 2005;123:347-356.  [PubMed]  [DOI]  [Cited in This Article: ]
36.  Malli R, Frieden M, Osibow K, Zoratti C, Mayer M, Demaurex N, Graier WF. Sustained Ca2+ transfer across mitochondria is Essential for mitochondrial Ca2+ buffering, sore-operated Ca2+ entry, and Ca2+ store refilling. J Biol Chem. 2003;278:44769-44779.  [PubMed]  [DOI]  [Cited in This Article: ]
37.  Laude AJ, Simpson AW. Compartmentalized signalling: Ca2+ compartments, microdomains and the many facets of Ca2+ signalling. FEBS J. 2009;276:1800-1816.  [PubMed]  [DOI]  [Cited in This Article: ]
38.  Berridge MJ. The endoplasmic reticulum: a multifunctional signaling organelle. Cell Calcium. 2002;32:235-249.  [PubMed]  [DOI]  [Cited in This Article: ]
39.  Jousset H, Malli R, Girardin N, Graier WF, Demaurex N, Frieden M. Evidence for a receptor-activated Ca2+ entry pathway independent from Ca2) store depletion in endothelial cells. Cell Calcium. 2008;43:83-94.  [PubMed]  [DOI]  [Cited in This Article: ]
40.  Malli R, Frieden M, Osibow K, Graier WF. Mitochondria efficiently buffer subplasmalemmal Ca2+ elevation during agonist stimulation. J Biol Chem. 2003;278:10807-10815.  [PubMed]  [DOI]  [Cited in This Article: ]
41.  Malli R, Frieden M, Trenker M, Graier WF. The role of mitochondria for Ca2+ refilling of the endoplasmic reticulum. J Biol Chem. 2005;280:12114-12122.  [PubMed]  [DOI]  [Cited in This Article: ]
42.  Osibow K, Frank S, Malli R, Zechner R, Graier WF. Mitochondria maintain maturation and secretion of lipoprotein lipase in the endoplasmic reticulum. Biochem J. 2006;396:173-182.  [PubMed]  [DOI]  [Cited in This Article: ]
43.  Girardin NC, Antigny F, Frieden M. Electrophysiological characterization of store-operated and agonist-induced Ca2+ entry pathways in endothelial cells. Pflugers Arch. 2010;460:109-120.  [PubMed]  [DOI]  [Cited in This Article: ]
44.  Mozhayeva MG, Mozhayeva GN. Evidence for the existence of inositol (1,4,5)-trisphosphate- and ryanodine-sensitive pools in bovine endothelial cells. Ca2+ releases in cells with different basal level of intracellular Ca2+. Pflugers Arch. 1996;432:614-622.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 4]  [Cited by in F6Publishing: 6]  [Article Influence: 0.2]  [Reference Citation Analysis (0)]
45.  Otun H, Aidulis DM, Yang JM, Gillespie JI. Interactions between inositol trisphosphate and Ca2+ dependent Ca2+ release mechanisms on the endoplasmic reticulum of permeabilised bovine aortic endothelial cells. Cell Calcium. 1996;19:315-325.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 10]  [Cited by in F6Publishing: 11]  [Article Influence: 0.4]  [Reference Citation Analysis (0)]
46.  Rusko J, Wang X, van Breemen C. Regenerative caffeine-induced responses in native rabbit aortic endothelial cells. Br J Pharmacol. 1995;115:811-821.  [PubMed]  [DOI]  [Cited in This Article: ]
47.  Zhang G, Teggatz EG, Zhang AY, Koeberl MJ, Yi F, Chen L, Li PL. Cyclic ADP ribose-mediated Ca2+ signaling in mediating endothelial nitric oxide production in bovine coronary arteries. Am J Physiol Heart Circ Physiol. 2006;290:H1172-H1181.  [PubMed]  [DOI]  [Cited in This Article: ]
48.  Zhang AY, Li PL. Vascular physiology of a Ca2+ mobilizing second messenger - cyclic ADP-ribose. J Cell Mol Med. 2006;10:407-422.  [PubMed]  [DOI]  [Cited in This Article: ]
49.  Lawrie AM, Rizzuto R, Pozzan T, Simpson AW. A role for calcium influx in the regulation of mitochondrial calcium in endothelial cells. J Biol Chem. 1996;271:10753-10759.  [PubMed]  [DOI]  [Cited in This Article: ]
50.  Brailoiu GC, Gurzu B, Gao X, Parkesh R, Aley PK, Trifa DI, Galione A, Dun NJ, Madesh M, Patel S. Acidic NAADP-sensitive calcium stores in the endothelium: agonist-specific recruitment and role in regulating blood pressure. J Biol Chem. 2010;285:37133-37137.  [PubMed]  [DOI]  [Cited in This Article: ]
51.  Esposito B, Gambara G, Lewis AM, Palombi F, D'Alessio A, Taylor LX, Genazzani AA, Ziparo E, Galione A, Churchill GC. NAADP links histamine H1 receptors to secretion of von Willebrand factor in human endothelial cells. Blood. 2011;117:4968-4977.  [PubMed]  [DOI]  [Cited in This Article: ]
52.  Srinivas SP, Ong A, Goon L, Goon L, Bonanno JA. Lysosomal Ca(2+) stores in bovine corneal endothelium. Invest Ophthalmol Vis Sci. 2002;43:2341-2350.  [PubMed]  [DOI]  [Cited in This Article: ]
53.  Zhu MX, Ma J, Parrington J, Galione A, Evans AM. TPCs: Endolysosomal channels for Ca2+ mobilization from acidic organelles triggered by NAADP. FEBS Lett. 2010;584:1966-1974.  [PubMed]  [DOI]  [Cited in This Article: ]
54.  Fiorio Pla A, Munaron L. Calcium influx, arachidonic acid,and control of endothelial cell proliferation. Cell Calcium. 2001;30:235-244.  [PubMed]  [DOI]  [Cited in This Article: ]
55.  Edwards DH, Li Y, Griffith TM. Hydrogen peroxide potentiates the EDHF phenomenon by promoting endothelial Ca2+ mobilization. Arterioscler Thromb Vasc Biol. 2008;28:1774-1781.  [PubMed]  [DOI]  [Cited in This Article: ]
56.  Hu Q, Corda S, Zweier JL, Capogrossi MC, Ziegelstein RC. Hydrogen peroxide induces intracellular calcium oscillations in human aortic endothelial cells. Circulation. 1998;97:268-275.  [PubMed]  [DOI]  [Cited in This Article: ]
57.  Zheng Y, Shen X. H2O2 directly activates inositol 1,4,5-trisphosphate receptors in endothelial cells. Redox Rep. 2005;10:29-36.  [PubMed]  [DOI]  [Cited in This Article: ]
58.  Itagaki K, Yun JK, Hengst JA, Yatani A, Hauser CJ, Spolarics Z, Deitch EA. Sphingosine 1-phosphate has dual functions in the regulation of endothelial cell permeability and Ca2+ metabolism. J Pharmacol Exp Ther. 2007;323:186-191.  [PubMed]  [DOI]  [Cited in This Article: ]
59.  Lucke S, Levkau B. Endothelial functions of sphingosine-1-phosphate. Cell Physiol Biochem. 2010;26:87-96.  [PubMed]  [DOI]  [Cited in This Article: ]
60.  Mehta D, Konstantoulaki M, Ahmmed GU, Malik AB. Sphingosine 1-phosphate-induced mobilization of intracellular Ca2+ mediates rac activation and adherens junction assembly in endothelial cells. J Biol Chem. 2005;280:17320-17328.  [PubMed]  [DOI]  [Cited in This Article: ]
61.  Seol GH, Kim MY, Liang GH, Kim JA, Kim YJ, Oh S, Suh SH. Sphingosine-1-phosphate-induced intracellular Ca2+ mobilization in human endothelial cells. Endothelium. 2005;12:263-269.  [PubMed]  [DOI]  [Cited in This Article: ]
62.  Munaron L. Intracellular calcium, endothelial cells and angiogenesis. Recent Pat Anticancer Drug Discov. 2006;1:105-119.  [PubMed]  [DOI]  [Cited in This Article: ]
63.  Rhee SG. Regulation of phosphoinositide-specific phospholipase C. Annu Rev Biochem. 2001;70:281-312.  [PubMed]  [DOI]  [Cited in This Article: ]
64.  Suh PG, Park JI, Manzoli L, Cocco L, Peak JC, Katan M, Fukami K, Kataoka T, Yun S, Ryu SH. Multiple roles of phosphoinositide-specific phospholipase C isozymes. BMB Rep. 2008;41:415-434.  [PubMed]  [DOI]  [Cited in This Article: ]
65.  Zoratti C, Kipmen-Korgun D, Osibow K, Malli R, Graier WF. Anandamide initiates Ca(2+) signaling via CB2 receptor linked to phospholipase C in calf pulmonary endothelial cells. Br J Pharmacol. 2003;140:1351-1362.  [PubMed]  [DOI]  [Cited in This Article: ]
66.  Gilchrist A, Vanhauwe JF, Li A, Thomas TO, Voyno-Yasenetskaya T, Hamm HE. G alpha minigenes expressing C-terminal peptides serve as specific inhibitors of thrombin-mediated endothelial activation. J Biol Chem. 2001;276:25672-25679.  [PubMed]  [DOI]  [Cited in This Article: ]
67.  Hirano K, Nomoto N, Hirano M, Momota F, Hanada A, Kanaide H. Distinct Ca2+ requirement for NO production between proteinase-activated receptor 1 and 4 (PAR1 and PAR4) in vascular endothelial cells. J Pharmacol Exp Ther. 2007;322:668-677.  [PubMed]  [DOI]  [Cited in This Article: ]
68.  Vanhauwe JF, Thomas TO, Minshall RD, Tiruppathi C, Li A, Gilchrist A, Yoon EJ, Malik AB, Hamm HE. Thrombin receptors activate G(o) proteins in endothelial cells to regulate intracellular calcium and cell shape changes. J Biol Chem. 2002;277:34143-34149.  [PubMed]  [DOI]  [Cited in This Article: ]
69.  Liao JK, Homcy CJ. The G proteins of the G alpha i and G alpha q family couple the bradykinin receptor to the release of endothelium-derived relaxing factor. J Clin Invest. 1993;92:2168-2172.  [PubMed]  [DOI]  [Cited in This Article: ]
70.  Lambert TL, Kent RS, Whorton AR. Bradykinin stimulation of inositol polyphosphate production in porcine aortic endothelial cells. J Biol Chem. 1986;261:15288-15293.  [PubMed]  [DOI]  [Cited in This Article: ]
71.  Wilde JI, Watson SP. Regulation of phospholipase C gamma isoforms in haematopoietic cells: why one, not the other? Cell Signal. 2001;13:691-701.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 115]  [Cited by in F6Publishing: 123]  [Article Influence: 5.3]  [Reference Citation Analysis (0)]
72.  Moccia F, Berra-Romani R, Tritto S, Signorelli S, Taglietti V, Tanzi F. Epidermal growth factor induces intracellular Ca2+ oscillations in microvascular endothelial cells. J Cell Physiol. 2003;194:139-150.  [PubMed]  [DOI]  [Cited in This Article: ]
73.  McLaughlin AP, De Vries GW. Role of PLCgamma and Ca(2+) in VEGF- and FGF-induced choroidal endothelial cell proliferation. Am J Physiol Cell Physiol. 2001;281:C1448-C1456.  [PubMed]  [DOI]  [Cited in This Article: ]
74.  Meyer RD, Latz C, Rahimi N. Recruitment and activation of phospholipase Cgamma1 by vascular endothelial growth factor receptor-2 are required for tubulogenesis and differentiation of endothelial cells. J Biol Chem. 2003;278:16347-16355.  [PubMed]  [DOI]  [Cited in This Article: ]
75.  Xiong Y, Huo Y, Chen C, Zeng H, Lu X, Wei C, Ruan C, Zhang X, Hu Z, Shibuya M. Vascular endothelial growth factor (VEGF) receptor-2 tyrosine 1175 signaling controls VEGF-induced von Willebrand factor release from endothelial cells via phospholipase C-gamma 1- and protein kinase A-dependent pathways. J Biol Chem. 2009;284:23217-23224.  [PubMed]  [DOI]  [Cited in This Article: ]
76.  Ridefelt P, Yokote K, Claesson-Welsh L, Siegbahn A. PDGF-BB triggered cytoplasmic calcium responses in cells with endogenous or stably transfected PDGF beta-receptors. Growth Factors. 1995;12:191-201.  [PubMed]  [DOI]  [Cited in This Article: ]
77.  Maffucci T, Raimondi C, Abu-Hayyeh S, Dominguez V, Sala G, Zachary I, Falasca M. A phosphoinositide 3-kinase/phospholipase Cgamma1 pathway regulates fibroblast growth factor-induced capillary tube formation. PLoS One. 2009;4:e8285.  [PubMed]  [DOI]  [Cited in This Article: ]
78.  Fleming I, Busse R. Tyrosine phosphorylation and bradykinin-induced signaling in endothelial cells. Am J Cardiol. 1997;80:102A-109A.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 27]  [Cited by in F6Publishing: 28]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
79.  Venema VJ, Ju H, Sun J, Eaton DC, Marrero MB, Venema RC. Bradykinin stimulates the tyrosine phosphorylation and bradykinin B2 receptor association of phospholipase C gamma 1 in vascular endothelial cells. Biochem Biophys Res Commun. 1998;246:70-75.  [PubMed]  [DOI]  [Cited in This Article: ]
80.  Zeng H, Zhao D, Yang S, Datta K, Mukhopadhyay D. Heterotrimeric G alpha q/G alpha 11 proteins function upstream of vascular endothelial growth factor (VEGF) receptor-2 (KDR) phosphorylation in vascular permeability factor/VEGF signaling. J Biol Chem. 2003;278:20738-20745.  [PubMed]  [DOI]  [Cited in This Article: ]
81.  Bhattacharya R, Kwon J, Li X, Wang E, Patra S, Bida JP, Bajzer Z, Claesson-Welsh L, Mukhopadhyay D. Distinct role of PLCbeta3 in VEGF-mediated directional migration and vascular sprouting. J Cell Sci. 2009;122:1025-1034.  [PubMed]  [DOI]  [Cited in This Article: ]
82.  Mi LY, Ettenson DS, Edelman ER. Phospholipase C-delta extends intercellular signalling range and responses to injury-released growth factors in non-excitable cells. Cell Prolif. 2008;41:671-690.  [PubMed]  [DOI]  [Cited in This Article: ]
83.  Li M, Edamatsu H, Kitazawa R, Kitazawa S, Kataoka T. Phospholipase Cepsilon promotes intestinal tumorigenesis of Apc(Min/+) mice through augmentation of inflammation and angiogenesis. Carcinogenesis. 2009;30:1424-1432.  [PubMed]  [DOI]  [Cited in This Article: ]
84.  Mayati A, Levoin N, Paris H, N'Diaye M, Courtois A, Uriac P, Lagadic-Gossmann D, Fardel O, Le Ferrec E. Induction of intracellular calcium concentration by environmental benzo(a)pyrene involves a β2-adrenergic receptor/adenylyl cyclase/Epac-1/inositol 1,4,5-trisphosphate pathway in endothelial cells. J Biol Chem. 2012;287:4041-4052.  [PubMed]  [DOI]  [Cited in This Article: ]
85.  Jaffe EA, Grulich J, Weksler BB, Hampel G, Watanabe K. Correlation between thrombin-induced prostacyclin production and inositol trisphosphate and cytosolic free calcium levels in cultured human endothelial cells. J Biol Chem. 1987;262:8557-8565.  [PubMed]  [DOI]  [Cited in This Article: ]
86.  Moscat J, Moreno F, Herrero C, López C, García-Barreno P. Endothelial cell growth factor and ionophore A23187 stimulation of production of inositol phosphates in porcine aorta endothelial cells. Proc Natl Acad Sci USA. 1988;85:659-663.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 23]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
87.  Pollock WK, Wreggett KA, Irvine RF. Inositol phosphate production and Ca2+ mobilization in human umbilical-vein endothelial cells stimulated by thrombin and histamine. Biochem J. 1988;256:371-376.  [PubMed]  [DOI]  [Cited in This Article: ]
88.  Brock TA, Dennis PA, Griendling KK, Diehl TS, Davies PF. GTP gamma S loading of endothelial cells stimulates phospholipase C and uncouples ATP receptors. Am J Physiol. 1988;255:C667-C673.  [PubMed]  [DOI]  [Cited in This Article: ]
89.  Pirotton S, Verjans B, Boeynaems JM, Erneux C. Metabolism of inositol phosphates in ATP-stimulated vascular endothelial cells. Biochem J. 1991;277:103-110.  [PubMed]  [DOI]  [Cited in This Article: ]
90.  Lin AY, Rui YC. Platelet-activating factor induced calcium mobilization and phosphoinositide metabolism in cultured bovine cerebral microvascular endothelial cells. Biochim Biophys Acta. 1994;1224:323-328.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 26]  [Cited by in F6Publishing: 28]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
91.  Faehling M, Kroll J, Föhr KJ, Fellbrich G, Mayr U, Trischler G, Waltenberger J. Essential role of calcium in vascular endothelial growth factor A-induced signaling: mechanism of the antiangiogenic effect of carboxyamidotriazole. FASEB J. 2002;16:1805-1807.  [PubMed]  [DOI]  [Cited in This Article: ]
92.  Buxton IL, Anzinger JJ. Agonist-specific regulation of inositol phosphate metabolism in cardiac endothelial cells. Proc West Pharmacol Soc. 2008;51:23-26.  [PubMed]  [DOI]  [Cited in This Article: ]
93.  Graier WF, Hoebel BG, Paltauf-Doburzynska J, Kostner GM. Effects of superoxide anions on endothelial Ca2+ signaling pathways. Arterioscler Thromb Vasc Biol. 1998;18:1470-1479.  [PubMed]  [DOI]  [Cited in This Article: ]
94.  Sun L, Yau HY, Lau OC, Huang Y, Yao X. Effect of hydrogen peroxide and superoxide anions on cytosolic Ca2+: comparison of endothelial cells from large-sized and small-sized arteries. PLoS One. 2011;6:e25432.  [PubMed]  [DOI]  [Cited in This Article: ]
95.  Rosales OR, Sumpio BE. Changes in cyclic strain increase inositol trisphosphate and diacylglycerol in endothelial cells. Am J Physiol. 1992;262:C956-C962.  [PubMed]  [DOI]  [Cited in This Article: ]
96.  Prasad AR, Logan SA, Nerem RM, Schwartz CJ, Sprague EA. Flow-related responses of intracellular inositol phosphate levels in cultured aortic endothelial cells. Circ Res. 1993;72:827-836.  [PubMed]  [DOI]  [Cited in This Article: ]
97.  Patel A, Sharif-Naeini R, Folgering JR, Bichet D, Duprat F, Honoré E. Canonical TRP channels and mechanotransduction: from physiology to disease states. Pflugers Arch. 2010;460:571-581.  [PubMed]  [DOI]  [Cited in This Article: ]
98.  Chachisvilis M, Zhang YL, Frangos JA. G protein-coupled receptors sense fluid shear stress in endothelial cells. Proc Natl Acad Sci USA. 2006;103:15463-15468.  [PubMed]  [DOI]  [Cited in This Article: ]
99.  Carter TD, Ogden D. Kinetics of Ca2+ release by InsP3 in pig single aortic endothelial cells: evidence for an inhibitory role of cytosolic Ca2+ in regulating hormonally evoked Ca2+ spikes. J Physiol. 1997;504:17-33.  [PubMed]  [DOI]  [Cited in This Article: ]
100.  Braet K, Mabilde C, Cabooter L, Rapp G, Leybaert L. Electroporation loading and photoactivation of caged InsP3: tools to investigate the relation between cellular ATP release in response to intracellular InsP3 elevation. J Neurosci Methods. 2004;132:81-89.  [PubMed]  [DOI]  [Cited in This Article: ]
101.  Freay A, Johns A, Adams DJ, Ryan US, Van Breemen C. Bradykinin and inositol 1,4,5-trisphosphate-stimulated calcium release from intracellular stores in cultured bovine endothelial cells. Pflugers Arch. 1989;414:377-384.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 70]  [Cited by in F6Publishing: 79]  [Article Influence: 2.3]  [Reference Citation Analysis (0)]
102.  Foskett JK, White C, Cheung KH, Mak DO. Inositol trisphosphate receptor Ca2+ release channels. Physiol Rev. 2007;87:593-658.  [PubMed]  [DOI]  [Cited in This Article: ]
103.  Bourguignon LY, Iida N, Sobrin L, Bourguignon GJ. Identification of an IP3 receptor in endothelial cells. J Cell Physiol. 1994;159:29-34.  [PubMed]  [DOI]  [Cited in This Article: ]
104.  Laflamme K, Domingue O, Guillemette BI, Guillemette G. Immunohistochemical localization of type 2 inositol 1,4,5-trisphosphate receptor to the nucleus of different mammalian cells. J Cell Biochem. 2002;85:219-228.  [PubMed]  [DOI]  [Cited in This Article: ]
105.  Gifford SM, Yi FX, Bird IM. Pregnancy-enhanced Ca2+ responses to ATP in uterine artery endothelial cells is due to greater capacitative Ca2+ entry rather than altered receptor coupling. J Endocrinol. 2006;190:373-384.  [PubMed]  [DOI]  [Cited in This Article: ]
106.  Mountian I, Manolopoulos VG, De Smedt H, Parys JB, Missiaen L, Wuytack F. Expression patterns of sarco/endoplasmic reticulum Ca(2+)-ATPase and inositol 1,4,5-trisphosphate receptor isoforms in vascular endothelial cells. Cell Calcium. 1999;25:371-380.  [PubMed]  [DOI]  [Cited in This Article: ]
107.  Grayson TH, Haddock RE, Murray TP, Wojcikiewicz RJ, Hill CE. Inositol 1,4,5-trisphosphate receptor subtypes are differentially distributed between smooth muscle and endothelial layers of rat arteries. Cell Calcium. 2004;36:447-458.  [PubMed]  [DOI]  [Cited in This Article: ]
108.  Isakson BE. Localized expression of an Ins(1,4,5)P3 receptor at the myoendothelial junction selectively regulates heterocellular Ca2+ communication. J Cell Sci. 2008;121:3664-3673.  [PubMed]  [DOI]  [Cited in This Article: ]
109.  Béliveau E, Guillemette G. Microfilament and microtubule assembly is required for the propagation of inositol trisphosphate receptor-induced Ca2+ waves in bovine aortic endothelial cells. J Cell Biochem. 2009;106:344-352.  [PubMed]  [DOI]  [Cited in This Article: ]
110.  Fujimoto T, Nakade S, Miyawaki A, Mikoshiba K, Ogawa K. Localization of inositol 1,4,5-trisphosphate receptor-like protein in plasmalemmal caveolae. J Cell Biol. 1992;119:1507-1513.  [PubMed]  [DOI]  [Cited in This Article: ]
111.  Vaca L, Kunze DL. IP3-activated Ca2+ channels in the plasma membrane of cultured vascular endothelial cells. Am J Physiol. 1995;269:C733-C738.  [PubMed]  [DOI]  [Cited in This Article: ]
112.  Mountian F, Jonas JC, Humbert De Smedt F, Parys JB. Expression of Ca(2+) Transport Genes in Platelets and Endothelial Cells in Hypertension. Hypertension. 2001;37:135-141.  [PubMed]  [DOI]  [Cited in This Article: ]
113.  Haorah J, Knipe B, Gorantla S, Zheng J, Persidsky Y. Alcohol-induced blood-brain barrier dysfunction is mediated via inositol 1,4,5-triphosphate receptor (IP3R)-gated intracellular calcium release. J Neurochem. 2007;100:324-336.  [PubMed]  [DOI]  [Cited in This Article: ]
114.  Taylor CW, Tovey SC. IP(3) receptors: toward understanding their activation. Cold Spring Harb Perspect Biol. 2010;2:a004010.  [PubMed]  [DOI]  [Cited in This Article: ]
115.  Adams DJ, Barakeh J, Laskey R, Van Breemen C. Ion channels and regulation of intracellular calcium in vascular endothelial cells. FASEB J. 1989;3:2389-2400.  [PubMed]  [DOI]  [Cited in This Article: ]
116.  Aromolaran AA, Blatter LA. Modulation of intracellular Ca2+ release and capacitative Ca2+ entry by CaMKII inhibitors in bovine vascular endothelial cells. Am J Physiol Cell Physiol. 2005;289:C1426-C1436.  [PubMed]  [DOI]  [Cited in This Article: ]
117.  Aromolaran AS, Zima AV, Blatter LA. Role of glycolytically generated ATP for CaMKII-mediated regulation of intracellular Ca2+ signaling in bovine vascular endothelial cells. Am J Physiol Cell Physiol. 2007;293:C106-C118.  [PubMed]  [DOI]  [Cited in This Article: ]
118.  Béliveau È, Lapointe F, Guillemette G. The activation state of the inositol 1,4,5-trisphosphate receptor regulates the velocity of intracellular Ca2+ waves in bovine aortic endothelial cells. J Cell Biochem. 2011;112:3722-3731.  [PubMed]  [DOI]  [Cited in This Article: ]
119.  Bourguignon LY, Singleton PA, Diedrich F. Hyaluronan-CD44 interaction with Rac1-dependent protein kinase N-gamma promotes phospholipase Cgamma1 activation, Ca(2+) signaling, and cortactin-cytoskeleton function leading to keratinocyte adhesion and differentiation. J Biol Chem. 2004;279:29654-29669.  [PubMed]  [DOI]  [Cited in This Article: ]
120.  Pan L, Wu X, Zhao D, Hessari NM, Lee I, Zhang X, Xu J. Sulfhydryl modification induces calcium entry through IP₃-sensitive store-operated pathway in activation-dependent human neutrophils. PLoS One. 2011;6:e25262.  [PubMed]  [DOI]  [Cited in This Article: ]
121.  Volk T, Mäding K, Hensel M, Kox WJ. Nitric oxide induces transient Ca2+ changes in endothelial cells independent of cGMP. J Cell Physiol. 1997;172:296-305.  [PubMed]  [DOI]  [Cited in This Article: ]
122.  Berra-Romani R, Raqeeb A, Avelino-Cruz JE, Moccia F, Oldani A, Speroni F, Taglietti V, Tanzi F. Ca2+ signaling in injured in situ endothelium of rat aorta. Cell Calcium. 2008;44:298-309.  [PubMed]  [DOI]  [Cited in This Article: ]
123.  Moccia F, Bertoni G, Pla AF, Dragoni S, Pupo E, Merlino A, Mancardi D, Munaron L, Tanzi F. Hydrogen sulfide regulates intracellular Ca2+ concentration in endothelial cells from excised rat aorta. Curr Pharm Biotechnol. 2011;12:1416-1426.  [PubMed]  [DOI]  [Cited in This Article: ]
124.  Lock JT, Sinkins WG, Schilling WP. Effect of protein S-glutathionylation on Ca2+ homeostasis in cultured aortic endothelial cells. Am J Physiol Heart Circ Physiol. 2011;300:H493-H506.  [PubMed]  [DOI]  [Cited in This Article: ]
125.  Kushnir A, Marks AR. The ryanodine receptor in cardiac physiology and disease. Adv Pharmacol. 2010;59:1-30.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 79]  [Cited by in F6Publishing: 35]  [Article Influence: 2.5]  [Reference Citation Analysis (0)]
126.  Fill M, Copello JA. Ryanodine receptor calcium release channels. Physiol Rev. 2002;82:893-922.  [PubMed]  [DOI]  [Cited in This Article: ]
127.  Köhler R, Brakemeier S, Kühn M, Degenhardt C, Buhr H, Pries A, Hoyer J. Expression of ryanodine receptor type 3 and TRP channels in endothelial cells: comparison of in situ and cultured human endothelial cells. Cardiovasc Res. 2001;51:160-168.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 63]  [Cited by in F6Publishing: 66]  [Article Influence: 2.9]  [Reference Citation Analysis (0)]
128.  West DJ, Williams AJ. Pharmacological regulators of intracellular calcium release channels. Curr Pharm Des. 2007;13:2428-2442.  [PubMed]  [DOI]  [Cited in This Article: ]
129.  Fransen P, Katnik C, Adams DJ. ACh- and caffeine-induced Ca2+ mobilization and current activation in rabbit arterial endothelial cells. Am J Physiol. 1998;275:H1748-H1758.  [PubMed]  [DOI]  [Cited in This Article: ]
130.  Graier WF, Simecek S, Bowles DK, Sturek M. Heterogeneity of caffeine- and bradykinin-sensitive Ca2+ stores in vascular endothelial cells. Biochem J. 1994;300:637-641.  [PubMed]  [DOI]  [Cited in This Article: ]
131.  Wang X, Lau F, Li L, Yoshikawa A, van Breemen C. Acetylcholine-sensitive intracellular Ca2+ store in fresh endothelial cells and evidence for ryanodine receptors. Circ Res. 1995;77:37-42.  [PubMed]  [DOI]  [Cited in This Article: ]
132.  Ziegelstein RC, Spurgeon HA, Pili R, Passaniti A, Cheng L, Corda S, Lakatta EG, Capogrossi MC. A functional ryanodine-sensitive intracellular Ca2+ store is present in vascular endothelial cells. Circ Res. 1994;74:151-156.  [PubMed]  [DOI]  [Cited in This Article: ]
133.  Paltauf-Doburzynska J, Posch K, Paltauf G, Graier WF. Stealth ryanodine-sensitive Ca2+ release contributes to activity of capacitative Ca2+ entry and nitric oxide synthase in bovine endothelial cells. J Physiol. 1998;513:369-379.  [PubMed]  [DOI]  [Cited in This Article: ]
134.  Paltauf-Doburzynska J, Frieden M, Spitaler M, Graier WF. Histamine-induced Ca2+ oscillations in a human endothelial cell line depend on transmembrane ion flux, ryanodine receptors and endoplasmic reticulum Ca2+-ATPase. J Physiol. 2000;524 Pt 3:701-713.  [PubMed]  [DOI]  [Cited in This Article: ]
135.  Ullmer C, Boddeke HG, Schmuck K, Lübbert H. 5-HT2B receptor-mediated calcium release from ryanodine-sensitive intracellular stores in human pulmonary artery endothelial cells. Br J Pharmacol. 1996;117:1081-1088.  [PubMed]  [DOI]  [Cited in This Article: ]
136.  Mauban JR, Wilkinson K, Schach C, Yuan JX. Histamine-mediated increases in cytosolic [Ca2+] involve different mechanisms in human pulmonary artery smooth muscle and endothelial cells. Am J Physiol Cell Physiol. 2006;290:C325-C336.  [PubMed]  [DOI]  [Cited in This Article: ]
137.  Sacks RS, Firth AL, Remillard CV, Agange N, Yau J, Ko EA, Yuan JX. Thrombin-mediated increases in cytosolic [Ca2+] involve different mechanisms in human pulmonary artery smooth muscle and endothelial cells. Am J Physiol Lung Cell Mol Physiol. 2008;295:L1048-L1055.  [PubMed]  [DOI]  [Cited in This Article: ]
138.  Graier WF, Paltauf-Doburzynska J, Posch K. Stealth Ca2-induced Ca2 release in the subplasmalemmal area contributes to activation of nitric oxide synthase upon autacoids in human endothelial cells. Circulation. 1998;98:667-667.  [PubMed]  [DOI]  [Cited in This Article: ]
139.  Liang W, Buluc M, van Breemen C, Wang X. Vectorial Ca2+ release via ryanodine receptors contributes to Ca2+ extrusion from freshly isolated rabbit aortic endothelial cells. Cell Calcium. 2004;36:431-443.  [PubMed]  [DOI]  [Cited in This Article: ]
140.  Frieden M, Graier WF. Subplasmalemmal ryanodine-sensitive Ca2+ release contributes to Ca2+-dependent K+ channel activation in a human umbilical vein endothelial cell line. J Physiol. 2000;524 Pt 3:715-724.  [PubMed]  [DOI]  [Cited in This Article: ]
141.  Frieden M, Malli R, Samardzija M, Demaurex N, Graier WF. Subplasmalemmal endoplasmic reticulum controls K(Ca) channel activity upon stimulation with a moderate histamine concentration in a human umbilical vein endothelial cell line. J Physiol. 2002;540:73-84.  [PubMed]  [DOI]  [Cited in This Article: ]
142.  Fliegert R, Gasser A, Guse AH. Regulation of calcium signalling by adenine-based second messengers. Biochem Soc Trans. 2007;35:109-114.  [PubMed]  [DOI]  [Cited in This Article: ]
143.  Aley PK, Porter KE, Boyle JP, Kemp PJ, Peers C. Hypoxic modulation of Ca2+ signaling in human venous endothelial cells. Multiple roles for reactive oxygen species. J Biol Chem. 2005;280:13349-13354.  [PubMed]  [DOI]  [Cited in This Article: ]
144.  Hecquet CM, Malik AB. Role of H(2)O(2)-activated TRPM2 calcium channel in oxidant-induced endothelial injury. Thromb Haemost. 2009;101:619-625.  [PubMed]  [DOI]  [Cited in This Article: ]
145.  Long C, Cook LG, Hamilton SL, Wu GY, Mitchell BM. FK506 binding protein 12/12.6 depletion increases endothelial nitric oxide synthase threonine 495 phosphorylation and blood pressure. Hypertension. 2007;49:569-576.  [PubMed]  [DOI]  [Cited in This Article: ]
146.  Long C, Cook LG, Wu GY, Mitchell BM. Removal of FKBP12/12.6 from endothelial ryanodine receptors leads to an intracellular calcium leak and endothelial dysfunction. Arterioscler Thromb Vasc Biol. 2007;27:1580-1586.  [PubMed]  [DOI]  [Cited in This Article: ]
147.  Patel S, Marchant JS, Brailoiu E. Two-pore channels: Regulation by NAADP and customized roles in triggering calcium signals. Cell Calcium. 2010;47:480-490.  [PubMed]  [DOI]  [Cited in This Article: ]
148.  Dietrich A, Kalwa H, Gudermann T. TRPC channels in vascular cell function. Thromb Haemost. 2010;103:262-270.  [PubMed]  [DOI]  [Cited in This Article: ]
149.  Earley S, Brayden JE. Transient receptor potential channels and vascular function. Clin Sci (. Lond). 2010;119:19-36.  [PubMed]  [DOI]  [Cited in This Article: ]
150.  Kwan HY, Huang Y, Yao X. TRP channels in endothelial function and dysfunction. Biochim Biophys Acta. 2007;1772:907-914.  [PubMed]  [DOI]  [Cited in This Article: ]
151.  Beech DJ. Harmony and discord in endothelial calcium entry. Circ Res. 2009;104:e22-e23.  [PubMed]  [DOI]  [Cited in This Article: ]
152.  Trebak M. STIM1/Orai1, ICRAC, and endothelial SOC. Circ Res. 2009;104:e56-e57.  [PubMed]  [DOI]  [Cited in This Article: ]
153.  Adams DJ, Hill MA. Potassium channels and membrane potential in the modulation of intracellular calcium in vascular endothelial cells. J Cardiovasc Electrophysiol. 2004;15:598-610.  [PubMed]  [DOI]  [Cited in This Article: ]
154.  Bossu JL, Feltz A, Rodeau JL, Tanzi F. Voltage-dependent transient calcium currents in freshly dissociated capillary endothelial cells. FEBS Lett. 1989;255:377-380.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 64]  [Cited by in F6Publishing: 68]  [Article Influence: 1.9]  [Reference Citation Analysis (0)]
155.  Bossu JL, Elhamdani A, Feltz A. Voltage-dependent calcium entry in confluent bovine capillary endothelial cells. FEBS Lett. 1992;299:239-242.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 38]  [Cited by in F6Publishing: 42]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
156.  Vinet R, Vargas FF. L- and T-type voltage-gated Ca2+ currents in adrenal medulla endothelial cells. Am J Physiol. 1999;276:H1313-H1322.  [PubMed]  [DOI]  [Cited in This Article: ]
157.  Wei Z, Manevich Y, Al-Mehdi AB, Chatterjee S, Fisher AB. Ca2+ flux through voltage-gated channels with flow cessation in pulmonary microvascular endothelial cells. Microcirculation. 2004;11:517-526.  [PubMed]  [DOI]  [Cited in This Article: ]
158.  Zhou C, Wu S. T-type calcium channels in pulmonary vascular endothelium. Microcirculation. 2006;13:645-656.  [PubMed]  [DOI]  [Cited in This Article: ]
159.  Townsley MI, King JA, Alvarez DF. Ca2+ channels and pulmonary endothelial permeability: insights from study of intact lung and chronic pulmonary hypertension. Microcirculation. 2006;13:725-739.  [PubMed]  [DOI]  [Cited in This Article: ]
160.  Delpiano MA, Altura BM. Modulatory effect of extracellular Mg2+ ions on K+ and Ca2+ currents of capillary endothelial cells from rat brain. FEBS Lett. 1996;394:335-339.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 25]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
161.  Yakubu MA, Leffler CW. L-type voltage-dependent Ca2+ channels in cerebral microvascular endothelial cells and ET-1 biosynthesis. Am J Physiol Cell Physiol. 2002;283:C1687-C1695.  [PubMed]  [DOI]  [Cited in This Article: ]
162.  Wu S, Haynes J, Taylor JT, Obiako BO, Stubbs JR, Li M, Stevens T. Cav3.1 (alpha1G) T-type Ca2+ channels mediate vaso-occlusion of sickled erythrocytes in lung microcirculation. Circ Res. 2003;93:346-353.  [PubMed]  [DOI]  [Cited in This Article: ]
163.  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. 2009;297:L650-L657.  [PubMed]  [DOI]  [Cited in This Article: ]
164.  Zhou C, Chen H, Lu F, Sellak H, Daigle JA, Alexeyev MF, Xi Y, Ju J, van Mourik JA, Wu S. Cav3.1 (alpha1G) controls von Willebrand factor secretion in rat pulmonary microvascular endothelial cells. Am J Physiol Lung Cell Mol Physiol. 2007;292:L833-L844.  [PubMed]  [DOI]  [Cited in This Article: ]
165.  Zhou C, Chen H, King JA, Sellak H, Kuebler WM, Yin J, Townsley MI, Shin HS, Wu S. Alpha1G T-type calcium channel selectively regulates P-selectin surface expression in pulmonary capillary endothelium. Am J Physiol Lung Cell Mol Physiol. 2010;299:L86-L97.  [PubMed]  [DOI]  [Cited in This Article: ]
166.  Vinet R, Cortés M, Luxoro M, Delpiano MA. Nicotine-evoked cytosolic Ca(2+) increase and cell depolarization in capillary endothelial cells of the bovine adrenal medulla. Biol Res. 2009;42:111-119.  [PubMed]  [DOI]  [Cited in This Article: ]
167.  Bkaily G, d'Orléans-Juste P, Naik R, Pérodin J, Stankova J, Abdulnour E, Rola-Pleszczynski M. PAF activation of a voltage-gated R-type Ca2+ channel in human and canine aortic endothelial cells. Br J Pharmacol. 1993;110:519-520.  [PubMed]  [DOI]  [Cited in This Article: ]
168.  Wang D, Hirase T, Inoue T, Node K. Atorvastatin inhibits angiotensin II-induced T-type Ca2+ channel expression in endothelial cells. Biochem Biophys Res Commun. 2006;347:394-400.  [PubMed]  [DOI]  [Cited in This Article: ]
169.  Cooke JP, Ghebremariam YT. Endothelial nicotinic acetylcholine receptors and angiogenesis. Trends Cardiovasc Med. 2008;18:247-253.  [PubMed]  [DOI]  [Cited in This Article: ]
170.  Gahring LC, Rogers SW. Neuronal nicotinic acetylcholine receptor expression and function on nonneuronal cells. AAPS J. 2005;7:E885-E894.  [PubMed]  [DOI]  [Cited in This Article: ]
171.  Cooke JP. NO and angiogenesis. Atheroscler Suppl. 2003;4:53-60.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 189]  [Cited by in F6Publishing: 176]  [Article Influence: 8.8]  [Reference Citation Analysis (0)]
172.  Fucile S. Ca2+ permeability of nicotinic acetylcholine receptors. Cell Calcium. 2004;35:1-8.  [PubMed]  [DOI]  [Cited in This Article: ]
173.  Arias HR, Richards VE, Ng D, Ghafoori ME, Le V, Mousa SA. Role of non-neuronal nicotinic acetylcholine receptors in angiogenesis. Int J Biochem Cell Biol. 2009;41:1441-1451.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 68]  [Cited by in F6Publishing: 75]  [Article Influence: 5.0]  [Reference Citation Analysis (0)]
174.  Moccia F, Frost C, Berra-Romani R, Tanzi F, Adams DJ. Expression and function of neuronal nicotinic ACh receptors in rat microvascular endothelial cells. Am J Physiol Heart Circ Physiol. 2004;286:H486-H491.  [PubMed]  [DOI]  [Cited in This Article: ]
175.  Li XW, Wang H. Non-neuronal nicotinic alpha 7 receptor, a new endothelial target for revascularization. Life Sci. 2006;78:1863-1870.  [PubMed]  [DOI]  [Cited in This Article: ]
176.  Roberts JA, Vial C, Digby HR, Agboh KC, Wen H, Atterbury-Thomas A, Evans RJ. Molecular properties of P2X receptors. Pflugers Arch. 2006;452:486-500.  [PubMed]  [DOI]  [Cited in This Article: ]
177.  Egan TM, Samways DS, Li Z. Biophysics of P2X receptors. Pflugers Arch. 2006;452:501-512.  [PubMed]  [DOI]  [Cited in This Article: ]
178.  Pulvirenti TJ, Yin JL, Chaufour X, McLachlan C, Hambly BD, Bennett MR, Barden JA. P2X (purinergic) receptor redistribution in rabbit aorta following injury to endothelial cells and cholesterol feeding. J Neurocytol. 2000;29:623-631.  [PubMed]  [DOI]  [Cited in This Article: ]
179.  Ray FR, Huang W, Slater M, Barden JA. Purinergic receptor distribution in endothelial cells in blood vessels: a basis for selection of coronary artery grafts. Atherosclerosis. 2002;162:55-61.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 57]  [Article Influence: 2.6]  [Reference Citation Analysis (0)]
180.  Ramirez AN, Kunze DL. P2X purinergic receptor channel expression and function in bovine aortic endothelium. Am J Physiol Heart Circ Physiol. 2002;282:H2106-H2116.  [PubMed]  [DOI]  [Cited in This Article: ]
181.  Yamamoto K, Korenaga R, Kamiya A, Ando J. Fluid shear stress activates Ca(2+) influx into human endothelial cells via P2X4 purinoceptors. Circ Res. 2000;87:385-391.  [PubMed]  [DOI]  [Cited in This Article: ]
182.  Yamamoto K, Korenaga R, Kamiya A, Qi Z, Sokabe M, Ando J. P2X(4) receptors mediate ATP-induced calcium influx in human vascular endothelial cells. Am J Physiol Heart Circ Physiol. 2000;279:H285-H292.  [PubMed]  [DOI]  [Cited in This Article: ]
183.  Bodin P, Burnstock G. Purinergic signalling: ATP release. Neurochem Res. 2001;26:959-969.  [PubMed]  [DOI]  [Cited in This Article: ]
184.  Reigada D, Mitchell CH. Release of ATP from retinal pigment epithelial cells involves both CFTR and vesicular transport. Am J Physiol Cell Physiol. 2005;288:C132-C140.  [PubMed]  [DOI]  [Cited in This Article: ]
185.  Yamamoto K, Shimizu N, Obi S, Kumagaya S, Taketani Y, Kamiya A, Ando J. Involvement of cell surface ATP synthase in flow-induced ATP release by vascular endothelial cells. Am J Physiol Heart Circ Physiol. 2007;293:H1646-H1653.  [PubMed]  [DOI]  [Cited in This Article: ]
186.  Yoshida T, Inoue R, Morii T, Takahashi N, Yamamoto S, Hara Y, Tominaga M, Shimizu S, Sato Y, Mori Y. Nitric oxide activates TRP channels by cysteine S-nitrosylation. Nat Chem Biol. 2006;2:596-607.  [PubMed]  [DOI]  [Cited in This Article: ]
187.  Yamamoto K, Sokabe T, Ohura N, Nakatsuka H, Kamiya A, Ando J. Endogenously released ATP mediates shear stress-induced Ca2+ influx into pulmonary artery endothelial cells. Am J Physiol Heart Circ Physiol. 2003;285:H793-H803.  [PubMed]  [DOI]  [Cited in This Article: ]
188.  Bradley J, Reisert J, Frings S. Regulation of cyclic nucleotide-gated channels. Curr Opin Neurobiol. 2005;15:343-349.  [PubMed]  [DOI]  [Cited in This Article: ]
189.  Kaupp UB, Seifert R. Cyclic nucleotide-gated ion channels. Physiol Rev. 2002;82:769-824.  [PubMed]  [DOI]  [Cited in This Article: ]
190.  Cheng KT, Chan FL, Huang Y, Chan WY, Yao X. Expression of olfactory-type cyclic nucleotide-gated channel (CNGA2) in vascular tissues. Histochem Cell Biol. 2003;120:475-481.  [PubMed]  [DOI]  [Cited in This Article: ]
191.  Yao X, Leung PS, Kwan HY, Wong TP, Fong MW. Rod-type cyclic nucleotide-gated cation channel is expressed in vascular endothelium and vascular smooth muscle cells. Cardiovasc Res. 1999;41:282-290.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 22]  [Cited by in F6Publishing: 24]  [Article Influence: 1.0]  [Reference Citation Analysis (0)]
192.  Kwan HY, Huang Y, Yao XQ, Leung FP. Role of cyclic nucleotides in the control of cytosolic Ca levels in vascular endothelial cells. Clin Exp Pharmacol Physiol. 2009;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]
193.  Maurice DH. Subcellular signaling in the endothelium: cyclic nucleotides take their place. Curr Opin Pharmacol. 2011;11:656-664.  [PubMed]  [DOI]  [Cited in This Article: ]
194.  Cheng KT, Leung YK, Shen B, Kwok YC, Wong CO, Kwan HY, Man YB, Ma X, Huang Y, Yao X. CNGA2 channels mediate adenosine-induced Ca2+ influx in vascular endothelial cells. Arterioscler Thromb Vasc Biol. 2008;28:913-918.  [PubMed]  [DOI]  [Cited in This Article: ]
195.  Kwan HY, Cheng KT, Ma Y, Huang Y, Tang NL, Yu S, Yao X. CNGA2 contributes to ATP-induced noncapacitative Ca2+ influx in vascular endothelial cells. J Vasc Res. 2010;47:148-156.  [PubMed]  [DOI]  [Cited in This Article: ]
196.  Yamamizu K, Yamashita JK. Roles of cyclic adenosine monophosphate signaling in endothelial cell differentiation and arterial-venous specification during vascular development. Circ J. 2011;75:253-260.  [PubMed]  [DOI]  [Cited in This Article: ]
197.  Wu S, Moore TM, Brough GH, Whitt SR, Chinkers M, Li M, Stevens T. Cyclic nucleotide-gated channels mediate membrane depolarization following activation of store-operated calcium entry in endothelial cells. J Biol Chem. 2000;275:18887-18896.  [PubMed]  [DOI]  [Cited in This Article: ]
198.  Shen B, Cheng KT, Leung YK, Kwok YC, Kwan HY, Wong CO, Chen ZY, Huang Y, Yao X. Epinephrine-induced Ca2+ influx in vascular endothelial cells is mediated by CNGA2 channels. J Mol Cell Cardiol. 2008;45:437-445.  [PubMed]  [DOI]  [Cited in This Article: ]
199.  Zhang J, Xia SL, Block ER, Patel JM. NO upregulation of a cyclic nucleotide-gated channel contributes to calcium elevation in endothelial cells. Am J Physiol Cell Physiol. 2002;283:C1080-C1089.  [PubMed]  [DOI]  [Cited in This Article: ]
200.  Gees M, Colsoul B, Nilius B. The role of transient receptor potential cation channels in Ca2+ signaling. Cold Spring Harb Perspect Biol. 2010;2:a003962.  [PubMed]  [DOI]  [Cited in This Article: ]
201.  Ramsey IS, Delling M, Clapham DE. An introduction to TRP channels. Annu Rev Physiol. 2006;68:619-647.  [PubMed]  [DOI]  [Cited in This Article: ]
202.  Owsianik G, Talavera K, Voets T, Nilius B. Permeation and selectivity of TRP channels. Annu Rev Physiol. 2006;68:685-717.  [PubMed]  [DOI]  [Cited in This Article: ]
203.  Cioffi DL, Lowe K, Alvarez DF, Barry C, Stevens T. TRPing on the lung endothelium: calcium channels that regulate barrier function. Antioxid Redox Signal. 2009;11:765-776.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 55]  [Cited by in F6Publishing: 56]  [Article Influence: 3.7]  [Reference Citation Analysis (0)]
204.  Vandenbroucke E, Mehta D, Minshall R, Malik AB. Regulation of endothelial junctional permeability. Ann N Y Acad Sci. 2008;1123:134-145.  [PubMed]  [DOI]  [Cited in This Article: ]
205.  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. 2008;103:1289-1299.  [PubMed]  [DOI]  [Cited in This Article: ]
206.  Li J, Cubbon RM, Wilson LA, Amer MS, McKeown L, Hou B, Majeed Y, Tumova S, Seymour VA, Taylor H. Orai1 and CRAC channel dependence of VEGF-activated Ca2+ entry and endothelial tube formation. Circ Res. 2011;108:1190-1198.  [PubMed]  [DOI]  [Cited in This Article: ]
207.  Antoniotti S, Lovisolo D, Fiorio Pla A, Munaron L. Expression and functional role of bTRPC1 channels in native endothelial cells. FEBS Lett. 2002;510:189-195.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 56]  [Cited by in F6Publishing: 62]  [Article Influence: 2.8]  [Reference Citation Analysis (0)]
208.  Antoniotti S, Fiorio Pla A, Barral S, Scalabrino O, Munaron L, Lovisolo D. Interaction between TRPC channel subunits in endothelial cells. J Recept Signal Transduct Res. 2006;26:225-240.  [PubMed]  [DOI]  [Cited in This Article: ]
209.  Bishara NB, Ding H. Glucose enhances expression of TRPC1 and calcium entry in endothelial cells. Am J Physiol Heart Circ Physiol. 2010;298:H171-H178.  [PubMed]  [DOI]  [Cited in This Article: ]
210.  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. 2003;278:33492-33500.  [PubMed]  [DOI]  [Cited in This Article: ]
211.  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. 2007;282:7833-7843.  [PubMed]  [DOI]  [Cited in This Article: ]
212.  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. 2004;279:20941-20949.  [PubMed]  [DOI]  [Cited in This Article: ]
213.  Sundivakkam PC, Kwiatek AM, Sharma TT, Minshall RD, Malik AB, Tiruppathi C. Caveolin-1 scaffold domain interacts with TRPC1 and IP3R3 to regulate Ca2+ store release-induced Ca2+ entry in endothelial cells. Am J Physiol Cell Physiol. 2009;296:C403-C413.  [PubMed]  [DOI]  [Cited in This Article: ]
214.  Ye S, Tan L, Ma J, Shi Q, Li J. Polyunsaturated docosahexaenoic acid suppresses oxidative stress induced endothelial cell calcium influx by altering lipid composition in membrane caveolar rafts. Prostaglandins Leukot Essent Fatty Acids. 2010;83:37-43.  [PubMed]  [DOI]  [Cited in This Article: ]
215.  Jho D, Mehta D, Ahmmed G, Gao XP, Tiruppathi C, Broman M, Malik AB. Angiopoietin-1 opposes VEGF-induced increase in endothelial permeability by inhibiting TRPC1-dependent Ca2 influx. Circ Res. 2005;96:1282-1290.  [PubMed]  [DOI]  [Cited in This Article: ]
216.  Kusaba T, Okigaki M, Matui A, Murakami M, Ishikawa K, Kimura T, Sonomura K, Adachi Y, Shibuya M, Shirayama T. Klotho is associated with VEGF receptor-2 and the transient receptor potential canonical-1 Ca2+ channel to maintain endothelial integrity. Proc Natl Acad Sci USA. 2010;107:19308-19313.  [PubMed]  [DOI]  [Cited in This Article: ]
217.  Cioffi DL, Barry C, Stevens T. Store-operated calcium entry channels in pulmonary endothelium: the emerging story of TRPCS and Orai1. Adv Exp Med Biol. 2010;661:137-154.  [PubMed]  [DOI]  [Cited in This Article: ]
218.  Sundivakkam PC, Freichel M, Singh V, Yuan JP, Vogel SM, Flockerzi V, Malik AB, Tiruppathi C. The Ca(2+) sensor stromal interaction molecule 1 (STIM1) is necessary and sufficient for the store-operated Ca(2+) entry function of transient receptor potential canonical (TRPC) 1 and 4 channels in endothelial cells. Mol Pharmacol. 2012;81:510-526.  [PubMed]  [DOI]  [Cited in This Article: ]
219.  Bair AM, Thippegowda PB, Freichel M, Cheng N, Ye RD, Vogel SM, Yu Y, Flockerzi V, Malik AB, Tiruppathi C. Ca2+ entry via TRPC channels is necessary for thrombin-induced NF-kappaB activation in endothelial cells through AMP-activated protein kinase and protein kinase Cdelta. J Biol Chem. 2009;284:563-574.  [PubMed]  [DOI]  [Cited in This Article: ]
220.  Thippegowda PB, Singh V, Sundivakkam PC, Xue J, Malik AB, Tiruppathi C. Ca2+ influx via TRPC channels induces NF-kappaB-dependent A20 expression to prevent thrombin-induced apoptosis in endothelial cells. Am J Physiol Cell Physiol. 2010;298:C656-C664.  [PubMed]  [DOI]  [Cited in This Article: ]
221.  Paria BC, Malik AB, Kwiatek AM, Rahman A, May MJ, Ghosh S, Tiruppathi C. Tumor necrosis factor-alpha induces nuclear factor-kappaB-dependent TRPC1 expression in endothelial cells. J Biol Chem. 2003;278:37195-37203.  [PubMed]  [DOI]  [Cited in This Article: ]
222.  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. 2004;287:L1303-L1313.  [PubMed]  [DOI]  [Cited in This Article: ]
223.  Yang LX, Guo RW, Liu B, Wang XM, Qi F, Guo CM, Shi YK, Wang H. Role of TRPC1 and NF-kappaB in mediating angiotensin II-induced Ca2+ entry and endothelial hyperpermeability. Peptides. 2009;30:1368-1373.  [PubMed]  [DOI]  [Cited in This Article: ]
224.  Antigny F, Girardin N, Frieden M. Transient receptor potential canonical channels are required for in vitro endothelial tube formation. J Biol Chem. 2012;287:5917-5927.  [PubMed]  [DOI]  [Cited in This Article: ]
225.  Yu PC, Gu SY, Bu JW, Du JL. TRPC1 is essential for in vivo angiogenesis in zebrafish. Circ Res. 2010;106:1221-1232.  [PubMed]  [DOI]  [Cited in This Article: ]
226.  Ma X, Qiu S, Luo J, Ma Y, Ngai CY, Shen B, Wong CO, Huang Y, Yao X. Functional role of vanilloid transient receptor potential 4-canonical transient receptor potential 1 complex in flow-induced Ca2+ influx. Arterioscler Thromb Vasc Biol. 2010;30:851-858.  [PubMed]  [DOI]  [Cited in This Article: ]
227.  Ma X, Cao J, Luo J, Nilius B, Huang Y, Ambudkar IS, Yao X. Depletion of intracellular Ca2+ stores stimulates the translocation of vanilloid transient receptor potential 4-c1 heteromeric channels to the plasma membrane. Arterioscler Thromb Vasc Biol. 2010;30:2249-2255.  [PubMed]  [DOI]  [Cited in This Article: ]
228.  Ma X, Cheng KT, Wong CO, O'Neil RG, Birnbaumer L, Ambudkar IS, Yao X. Heteromeric TRPV4-C1 channels contribute to store-operated Ca(2+) entry in vascular endothelial cells. Cell Calcium. 2011;50:502-509.  [PubMed]  [DOI]  [Cited in This Article: ]
229.  Berrout J, Jin M, O'Neil RG. Critical role of TRPP2 and TRPC1 channels in stretch-induced injury of blood-brain barrier endothelial cells. Brain Res. 2012;1436:1-12.  [PubMed]  [DOI]  [Cited in This Article: ]
230.  Cheng HW, James AF, Foster RR, Hancox JC, Bates DO. VEGF activates receptor-operated cation channels in human microvascular endothelial cells. Arterioscler Thromb Vasc Biol. 2006;26:1768-1776.  [PubMed]  [DOI]  [Cited in This Article: ]
231.  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. 2009;283:43-51.  [PubMed]  [DOI]  [Cited in This Article: ]
232.  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. 1999;518 Pt 2:345-358.  [PubMed]  [DOI]  [Cited in This Article: ]
233.  Leung PC, Cheng KT, Liu C, Cheung WT, Kwan HY, Lau KL, Huang Y, Yao X. Mechanism of non-capacitative Ca2+ influx in response to bradykinin in vascular endothelial cells. J Vasc Res. 2006;43:367-376.  [PubMed]  [DOI]  [Cited in This Article: ]
234.  Smedlund K, Vazquez G. Involvement of native TRPC3 proteins in ATP-dependent expression of VCAM-1 and monocyte adherence in coronary artery endothelial cells. Arterioscler Thromb Vasc Biol. 2008;28:2049-2055.  [PubMed]  [DOI]  [Cited in This Article: ]
235.  Huang JH, He GW, Xue HM, Yao XQ, Liu XC, Underwood MJ, Yang Q. TRPC3 channel contributes to nitric oxide release: significance during normoxia and hypoxia-reoxygenation. Cardiovasc Res. 2011;91:472-482.  [PubMed]  [DOI]  [Cited in This Article: ]
236.  Pocock TM, Foster RR, Bates DO. Evidence of a role for TRPC channels in VEGF-mediated increased vascular permeability in vivo. Am J Physiol Heart Circ Physiol. 2004;286:H1015-H1026.  [PubMed]  [DOI]  [Cited in This Article: ]
237.  Smedlund K, Tano JY, Vazquez G. The constitutive function of native TRPC3 channels modulates vascular cell adhesion molecule-1 expression in coronary endothelial cells through nuclear factor kappaB signaling. Circ Res. 2010;106:1479-1488.  [PubMed]  [DOI]  [Cited in This Article: ]
238.  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. 2006;27:981-990.  [PubMed]  [DOI]  [Cited in This Article: ]
239.  Kwan HY, Huang Y, Yao X. Protein kinase C can inhibit TRPC3 channels indirectly via stimulating protein kinase G. J Cell Physiol. 2006;207:315-321.  [PubMed]  [DOI]  [Cited in This Article: ]
240.  Takahashi S, Lin H, Geshi N, Mori Y, Kawarabayashi Y, Takami N, Mori MX, Honda A, Inoue R. Nitric oxide-cGMP-protein kinase G pathway negatively regulates vascular transient receptor potential channel TRPC6. J Physiol. 2008;586:4209-4223.  [PubMed]  [DOI]  [Cited in This Article: ]
241.  Balzer M, Lintschinger B, Groschner K. Evidence for a role of Trp proteins in the oxidative stress-induced membrane conductances of porcine aortic endothelial cells. Cardiovasc Res. 1999;42:543-549.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 118]  [Cited by in F6Publishing: 122]  [Article Influence: 4.9]  [Reference Citation Analysis (0)]
242.  Poteser M, Graziani A, Rosker C, Eder P, Derler I, Kahr H, Zhu MX, Romanin C, Groschner K. TRPC3 and TRPC4 associate to form a redox-sensitive cation channel. Evidence for expression of native TRPC3-TRPC4 heteromeric channels in endothelial cells. J Biol Chem. 2006;281:13588-13595.  [PubMed]  [DOI]  [Cited in This Article: ]
243.  Antigny F, Jousset H, König S, Frieden M. Thapsigargin activates Ca²+ entry both by store-dependent, STIM1/Orai1-mediated, and store-independent, TRPC3/PLC/PKC-mediated pathways in human endothelial cells. Cell Calcium. 2011;49:115-127.  [PubMed]  [DOI]  [Cited in This Article: ]
244.  Thilo F, Loddenkemper C, Berg E, Zidek W, Tepel M. Increased TRPC3 expression in vascular endothelium of patients with malignant hypertension. Mod Pathol. 2009;22:426-430.  [PubMed]  [DOI]  [Cited in This Article: ]
245.  Thilo F, Baumunk D, Krause H, Schrader M, Miller K, Loddenkemper C, Zakrzewicz A, Krueger K, Zidek W, Tepel M. Transient receptor potential canonical type 3 channels and blood pressure in humans. J Hypertens. 2009;27:1217-1223.  [PubMed]  [DOI]  [Cited in This Article: ]
246.  Fleming I, Rueben A, Popp R, Fisslthaler B, Schrodt S, Sander A, Haendeler J, Falck JR, Morisseau C, Hammock BD. Epoxyeicosatrienoic acids regulate Trp channel dependent Ca2+ signaling and hyperpolarization in endothelial cells. Arterioscler Thromb Vasc Biol. 2007;27:2612-2618.  [PubMed]  [DOI]  [Cited in This Article: ]
247.  Dora KA. Coordination of vasomotor responses by the endothelium. Circ J. 2010;74:226-232.  [PubMed]  [DOI]  [Cited in This Article: ]
248.  Chaudhuri P, Colles SM, Bhat M, Van Wagoner DR, Birnbaumer L, Graham LM. Elucidation of a TRPC6-TRPC5 channel cascade that restricts endothelial cell movement. Mol Biol Cell. 2008;19:3203-3211.  [PubMed]  [DOI]  [Cited in This Article: ]
249.  Kini V, Chavez A, Mehta D. A new role for PTEN in regulating transient receptor potential canonical channel 6-mediated Ca2+ entry, endothelial permeability, and angiogenesis. J Biol Chem. 2010;285:33082-33091.  [PubMed]  [DOI]  [Cited in This Article: ]
250.  Weissmann N, Sydykov A, Kalwa H, Storch U, Fuchs B, Mederos y Schnitzler M, Brandes RP, Grimminger F, Meissner M, Freichel M. Activation of TRPC6 channels is essential for lung ischaemia-reperfusion induced oedema in mice. Nat Commun. 2012;3:649.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 140]  [Cited by in F6Publishing: 145]  [Article Influence: 12.1]  [Reference Citation Analysis (0)]
251.  Samapati R, Yang Y, Yin J, Stoerger C, Arenz C, Dietrich A, Gudermann T, Adam D, Wu S, Freichel M. 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. 2012;185:160-170.  [PubMed]  [DOI]  [Cited in This Article: ]
252.  Plant TD, Schaefer M. Receptor-operated cation channels formed by TRPC4 and TRPC5. Naunyn Schmiedebergs Arch Pharmacol. 2005;371:266-276.  [PubMed]  [DOI]  [Cited in This Article: ]
253.  Otsuguro K, Tang J, Tang Y, Xiao R, Freichel M, Tsvilovskyy V, Ito S, Flockerzi V, Zhu MX, Zholos AV. Isoform-specific inhibition of TRPC4 channel by phosphatidylinositol 4,5-bisphosphate. J Biol Chem. 2008;283:10026-10036.  [PubMed]  [DOI]  [Cited in This Article: ]
254.  Lee-Kwon W, Wade JB, Zhang Z, Pallone TL, Weinman EJ. Expression of TRPC4 channel protein that interacts with NHERF-2 in rat descending vasa recta. Am J Physiol Cell Physiol. 2005;288:C942-C949.  [PubMed]  [DOI]  [Cited in This Article: ]
255.  Blair NT, Kaczmarek JS, Clapham DE. Intracellular calcium strongly potentiates agonist-activated TRPC5 channels. J Gen Physiol. 2009;133:525-546.  [PubMed]  [DOI]  [Cited in This Article: ]
256.  Obukhov AG, Nowycky MC. TRPC5 activation kinetics are modulated by the scaffolding protein ezrin/radixin/moesin-binding phosphoprotein-50 (EBP50). J Cell Physiol. 2004;201:227-235.  [PubMed]  [DOI]  [Cited in This Article: ]
257.  Schindl R, Frischauf I, Kahr H, Fritsch R, Krenn M, Derndl A, Vales E, Muik M, Derler I, Groschner K. The first ankyrin-like repeat is the minimum indispensable key structure for functional assembly of homo- and heteromeric TRPC4/TRPC5 channels. Cell Calcium. 2008;43:260-269.  [PubMed]  [DOI]  [Cited in This Article: ]
258.  Freichel M, Suh SH, Pfeifer A, Schweig U, Trost C, Weissgerber P, Biel M, Philipp S, Freise D, Droogmans G. Lack of an endothelial store-operated Ca2+ current impairs agonist-dependent vasorelaxation in TRP4-/- mice. Nat Cell Biol. 2001;3:121-127.  [PubMed]  [DOI]  [Cited in This Article: ]
259.  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. 2002;91:70-76.  [PubMed]  [DOI]  [Cited in This Article: ]
260.  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. 2005;97:1164-1172.  [PubMed]  [DOI]  [Cited in This Article: ]
261.  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. 2001;15:1727-1738.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 139]  [Cited by in F6Publishing: 142]  [Article Influence: 6.2]  [Reference Citation Analysis (0)]
262.  Murata T, Lin MI, Stan RV, Bauer PM, Yu J, Sessa WC. Genetic evidence supporting caveolae microdomain regulation of calcium entry in endothelial cells. J Biol Chem. 2007;282:16631-16643.  [PubMed]  [DOI]  [Cited in This Article: ]
263.  Cioffi DL, Stevens T. Regulation of endothelial cell barrier function by store-operated calcium entry. Microcirculation. 2006;13:709-723.  [PubMed]  [DOI]  [Cited in This Article: ]
264.  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. 2010;285:4213-4223.  [PubMed]  [DOI]  [Cited in This Article: ]
265.  Groschner K. Polymodal TRPC signaling: Emerging role in phenotype switching and tissue remodeling. Commun Integr Biol. 2010;3:393-395.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 3]  [Cited by in F6Publishing: 4]  [Article Influence: 0.3]  [Reference Citation Analysis (0)]
266.  Jung C, Gené GG, Tomás M, Plata C, Selent J, Pastor M, Fandos C, Senti M, Lucas G, Elosua R. A gain-of-function SNP in TRPC4 cation channel protects against myocardial infarction. Cardiovasc Res. 2011;91:465-471.  [PubMed]  [DOI]  [Cited in This Article: ]
267.  Wong CO, Sukumar P, Beech DJ, Yao X. Nitric oxide lacks direct effect on TRPC5 channels but suppresses endogenous TRPC5-containing channels in endothelial cells. Pflugers Arch. 2010;460:121-130.  [PubMed]  [DOI]  [Cited in This Article: ]
268.  Wong CO, Huang Y, Yao X. Genistein potentiates activity of the cation channel TRPC5 independently of tyrosine kinases. Br J Pharmacol. 2010;159:1486-1496.  [PubMed]  [DOI]  [Cited in This Article: ]
269.  Liu Y, Xu Y, Thilo F, Friis UG, Jensen BL, Scholze A, Zheng J, Tepel M. Erythropoietin increases expression and function of transient receptor potential canonical 5 channels. Hypertension. 2011;58:317-324.  [PubMed]  [DOI]  [Cited in This Article: ]
270.  Baylie RL, Brayden JE. TRPV channels and vascular function. Acta Physiol (. Oxf). 2011;203:99-116.  [PubMed]  [DOI]  [Cited in This Article: ]
271.  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. 2011;91:492-501.  [PubMed]  [DOI]  [Cited in This Article: ]
272.  Golech SA, McCarron RM, Chen Y, Bembry J, Lenz F, Mechoulam R, Shohami E, Spatz M. Human brain endothelium: coexpression and function of vanilloid and endocannabinoid receptors. Brain Res Mol Brain Res. 2004;132:87-92.  [PubMed]  [DOI]  [Cited in This Article: ]
273.  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. 2008;294:H2489-H2496.  [PubMed]  [DOI]  [Cited in This Article: ]
274.  Kark T, Bagi Z, Lizanecz E, Pásztor ET, Erdei N, Czikora A, Papp Z, Edes I, Pórszász R, Tóth A. Tissue-specific regulation of microvascular diameter: opposite functional roles of neuronal and smooth muscle located vanilloid receptor-1. Mol Pharmacol. 2008;73:1405-1412.  [PubMed]  [DOI]  [Cited in This Article: ]
275.  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. 2005;568:539-551.  [PubMed]  [DOI]  [Cited in This Article: ]
276.  Ye F, Deng PY, Li D, Luo D, Li NS, Deng S, Deng HW, Li YJ. Involvement of endothelial cell-derived CGRP in heat stress-induced protection of endothelial function. Vascul Pharmacol. 2007;46:238-246.  [PubMed]  [DOI]  [Cited in This Article: ]
277.  Luo D, Zhang YW, Peng WJ, Peng J, Chen QQ, Li D, Deng HW, Li YJ. Transient receptor potential vanilloid 1-mediated expression and secretion of endothelial cell-derived calcitonin gene-related peptide. Regul Pept. 2008;150:66-72.  [PubMed]  [DOI]  [Cited in This Article: ]
278.  Hoi PM, Visintin C, Okuyama M, Gardiner SM, Kaup SS, Bennett T, Baker D, Selwood DL, Hiley CR. Vascular pharmacology of a novel cannabinoid-like compound, 3-(5-dimethylcarbamoyl-pent-1-enyl)-N-(2-hydroxy-1-methyl-ethyl)benzamide (VSN16) in the rat. Br J Pharmacol. 2007;152:751-764.  [PubMed]  [DOI]  [Cited in This Article: ]
279.  Ching LC, Chen CY, Su KH, Hou HH, Shyue SK, Kou YR, Lee TS. Implication of AMP-activated protein kinase in transient receptor potential vanilloid type 1-mediated activation of endothelial nitric oxide synthase. Mol Med. 2012;Epub ahead of print.  [PubMed]  [DOI]  [Cited in This Article: ]
280.  Yang D, Luo Z, Ma S, Wong WT, Ma L, Zhong J, He H, Zhao Z, Cao T, Yan Z. Activation of TRPV1 by dietary capsaicin improves endothelium-dependent vasorelaxation and prevents hypertension. Cell Metab. 2010;12:130-141.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 242]  [Cited by in F6Publishing: 243]  [Article Influence: 17.4]  [Reference Citation Analysis (0)]
281.  Hao X, Chen J, Luo Z, He H, Yu H, Ma L, Ma S, Zhu T, Liu D, Zhu Z. TRPV1 activation prevents high-salt diet-induced nocturnal hypertension in mice. Pflugers Arch. 2011;461:345-353.  [PubMed]  [DOI]  [Cited in This Article: ]
282.  Pacher P, Bátkai S, Kunos G. Haemodynamic profile and responsiveness to anandamide of TRPV1 receptor knock-out mice. J Physiol. 2004;558:647-657.  [PubMed]  [DOI]  [Cited in This Article: ]
283.  Wang Y, Wang DH. Aggravated renal inflammatory responses in TRPV1 gene knockout mice subjected to DOCA-salt hypertension. Am J Physiol Renal Physiol. 2009;297:F1550-F1559.  [PubMed]  [DOI]  [Cited in This Article: ]
284.  Mergler S, Valtink M, Coulson-Thomas VJ, Lindemann D, Reinach PS, Engelmann K, Pleyer U. TRPV channels mediate temperature-sensing in human corneal endothelial cells. Exp Eye Res. 2010;90:758-770.  [PubMed]  [DOI]  [Cited in This Article: ]
285.  Ohanyan VA, Guarini G, Thodeti CK, Talasila PK, Raman P, Haney RM, Meszaros JG, Damron DS, Bratz IN. Endothelin-mediated in vivo pressor responses following TRPV1 activation. Am J Physiol Heart Circ Physiol. 2011;301:H1135-H1142.  [PubMed]  [DOI]  [Cited in This Article: ]
286.  Min JK, Han KY, Kim EC, Kim YM, Lee SW, Kim OH, Kim KW, Gho YS, Kwon YG. Capsaicin inhibits in vitro and in vivo angiogenesis. Cancer Res. 2004;64:644-651.  [PubMed]  [DOI]  [Cited in This Article: ]
287.  Earley S, Gonzales AL, Garcia ZI. A dietary agonist of transient receptor potential cation channel V3 elicits endothelium-dependent vasodilation. Mol Pharmacol. 2010;77:612-620.  [PubMed]  [DOI]  [Cited in This Article: ]
288.  Watanabe H, Vriens J, Suh SH, Benham CD, Droogmans G, Nilius B. Heat-evoked activation of TRPV4 channels in a HEK293 cell expression system and in native mouse aorta endothelial cells. J Biol Chem. 2002;277:47044-47051.  [PubMed]  [DOI]  [Cited in This Article: ]
289.  Vriens J, Watanabe H, Janssens A, Droogmans G, Voets T, Nilius B. Cell swelling, heat, and chemical agonists use distinct pathways for the activation of the cation channel TRPV4. Proc Natl Acad Sci USA. 2004;101:396-401.  [PubMed]  [DOI]  [Cited in This Article: ]
290.  Hartmannsgruber V, Heyken WT, Kacik M, Kaistha A, Grgic I, Harteneck C, Liedtke W, Hoyer J, Köhler R. Arterial response to shear stress critically depends on endothelial TRPV4 expression. PLoS One. 2007;2:e827.  [PubMed]  [DOI]  [Cited in This Article: ]
291.  Köhler R, Heyken WT, Heinau P, Schubert R, Si H, Kacik M, Busch C, Grgic I, Maier T, Hoyer J. Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation. Arterioscler Thromb Vasc Biol. 2006;26:1495-1502.  [PubMed]  [DOI]  [Cited in This Article: ]
292.  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. 2009;104:1123-1130.  [PubMed]  [DOI]  [Cited in This Article: ]
293.  Matthews BD, Thodeti CK, Tytell JD, Mammoto A, Overby DR, Ingber DE. Ultra-rapid activation of TRPV4 ion channels by mechanical forces applied to cell surface beta1 integrins. Integr Biol (. Camb). 2010;2:435-442.  [PubMed]  [DOI]  [Cited in This Article: ]
294.  Watanabe H, Davis JB, Smart D, Jerman JC, Smith GD, Hayes P, Vriens J, Cairns W, Wissenbach U, Prenen J. Activation of TRPV4 channels (hVRL-2/mTRP12) by phorbol derivatives. J Biol Chem. 2002;277:13569-13577.  [PubMed]  [DOI]  [Cited in This Article: ]
295.  Ma X, He D, Ru X, Chen Y, Cai Y, Bruce IC, Xia Q, Yao X, Jin J. Apigenin, a plant-derived flavone, activates transient receptor potential vanilloid 4 cation channel. Br J Pharmacol. 2012;166:349-358.  [PubMed]  [DOI]  [Cited in This Article: ]
296.  Loot AE, Popp R, Fisslthaler B, Vriens J, Nilius B, Fleming I. Role of cytochrome P450-dependent transient receptor potential V4 activation in flow-induced vasodilatation. Cardiovasc Res. 2008;80:445-452.  [PubMed]  [DOI]  [Cited in This Article: ]
297.  Marrelli SP, O'neil RG, Brown RC, Bryan RM. PLA2 and TRPV4 channels regulate endothelial calcium in cerebral arteries. Am J Physiol Heart Circ Physiol. 2007;292:H1390-H1397.  [PubMed]  [DOI]  [Cited in This Article: ]
298.  Watanabe H, Vriens J, Prenen J, Droogmans G, Voets T, Nilius B. Anandamide and arachidonic acid use epoxyeicosatrienoic acids to activate TRPV4 channels. Nature. 2003;424:434-438.  [PubMed]  [DOI]  [Cited in This Article: ]
299.  Vriens J, Owsianik G, Fisslthaler B, Suzuki M, Janssens A, Voets T, Morisseau C, Hammock BD, Fleming I, Busse R. Modulation of the Ca2 permeable cation channel TRPV4 by cytochrome P450 epoxygenases in vascular endothelium. Circ Res. 2005;97:908-915.  [PubMed]  [DOI]  [Cited in This Article: ]
300.  Bubolz AH, Mendoza SA, Zheng X, Zinkevich NS, Li R, Gutterman DD, Zhang DX. Activation of endothelial TRPV4 channels mediates flow-induced dilation in human coronary arterioles: role of Ca2+ entry and mitochondrial ROS signaling. Am J Physiol Heart Circ Physiol. 2012;302:H634-H642.  [PubMed]  [DOI]  [Cited in This Article: ]
301.  Saliez J, Bouzin C, Rath G, Ghisdal P, Desjardins F, Rezzani R, Rodella LF, Vriens J, Nilius B, Feron O. 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. 2008;117:1065-1074.  [PubMed]  [DOI]  [Cited in This Article: ]
302.  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. 2010;28:102-110.  [PubMed]  [DOI]  [Cited in This Article: ]
303.  Willette RN, Bao W, Nerurkar S, Yue TL, Doe CP, Stankus G, Turner GH, Ju H, Thomas H, Fishman CE. Systemic activation of the transient receptor potential vanilloid subtype 4 channel causes endothelial failure and circulatory collapse: Part 2. J Pharmacol Exp Ther. 2008;326:443-452.  [PubMed]  [DOI]  [Cited in This Article: ]
304.  Earley S, Gonzales AL, Crnich R. Endothelium-dependent cerebral artery dilation mediated by TRPA1 and Ca2+-Activated K+ channels. Circ Res. 2009;104:987-994.  [PubMed]  [DOI]  [Cited in This Article: ]
305.  Earley S, Pauyo T, Drapp R, Tavares MJ, Liedtke W, Brayden JE. TRPV4-dependent dilation of peripheral resistance arteries influences arterial pressure. Am J Physiol Heart Circ Physiol. 2009;297:H1096-H1102.  [PubMed]  [DOI]  [Cited in This Article: ]
306.  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. 2009;53:532-538.  [PubMed]  [DOI]  [Cited in This Article: ]
307.  Adapala RK, Talasila PK, Bratz IN, Zhang DX, Suzuki M, Meszaros JG, Thodeti CK. PKCα mediates acetylcholine-induced activation of TRPV4-dependent calcium influx in endothelial cells. Am J Physiol Heart Circ Physiol. 2011;301:H757-H765.  [PubMed]  [DOI]  [Cited in This Article: ]
308.  Gao F, Wang DH. Impairment in function and expression of transient receptor potential vanilloid type 4 in Dahl salt-sensitive rats: significance and mechanism. Hypertension. 2010;55:1018-1025.  [PubMed]  [DOI]  [Cited in This Article: ]
309.  Gao F, Sui D, Garavito RM, Worden RM, Wang DH. Salt intake augments hypotensive effects of transient receptor potential vanilloid 4: functional significance and implication. Hypertension. 2009;53:228-235.  [PubMed]  [DOI]  [Cited in This Article: ]
310.  Troidl C, Nef H, Voss S, Schilp A, Kostin S, Troidl K, Szardien S, Rolf A, Schmitz-Rixen T, Schaper W. Calcium-dependent signalling is essential during collateral growth in the pig hind limb-ischemia model. J Mol Cell Cardiol. 2010;49:142-151.  [PubMed]  [DOI]  [Cited in This Article: ]
311.  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. 2011;41:589-596.  [PubMed]  [DOI]  [Cited in This Article: ]
312.  Troidl C, Troidl K, Schierling W, Cai WJ, Nef H, Möllmann H, Kostin S, Schimanski S, Hammer L, Elsässer A. Trpv4 induces collateral vessel growth during regeneration of the arterial circulation. J Cell Mol Med. 2009;13:2613-2621.  [PubMed]  [DOI]  [Cited in This Article: ]
313.  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. 2006;99:988-995.  [PubMed]  [DOI]  [Cited in This Article: ]
314.  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. 2007;293:L923-L932.  [PubMed]  [DOI]  [Cited in This Article: ]
315.  Jian MY, King JA, Al-Mehdi AB, Liedtke W, Townsley MI. High vascular pressure-induced lung injury requires P450 epoxygenase-dependent activation of TRPV4. Am J Respir Cell Mol Biol. 2008;38:386-392.  [PubMed]  [DOI]  [Cited in This Article: ]
316.  Yin J, Hoffmann J, Kaestle SM, Neye N, Wang L, Baeurle J, Liedtke W, Wu S, Kuppe H, Pries AR. Negative-feedback loop attenuates hydrostatic lung edema via a cGMP-dependent regulation of transient receptor potential vanilloid 4. Circ Res. 2008;102:966-974.  [PubMed]  [DOI]  [Cited in This Article: ]
317.  Fian R, Grasser E, Treiber F, Schmidt R, Niederl P, Rosker C. The contribution of TRPV4-mediated calcium signaling to calcium homeostasis in endothelial cells. J Recept Signal Transduct Res. 2007;27:113-124.  [PubMed]  [DOI]  [Cited in This Article: ]
318.  Mergler S, Valtink M, Taetz K, Sahlmüller M, Fels G, Reinach PS, Engelmann K, Pleyer U. Characterization of transient receptor potential vanilloid channel 4 (TRPV4) in human corneal endothelial cells. Exp Eye Res. 2011;93:710-719.  [PubMed]  [DOI]  [Cited in This Article: ]
319.  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. 2012;31:200-212.  [PubMed]  [DOI]  [Cited in This Article: ]
320.  Hecquet CM, Ahmmed GU, Vogel SM, Malik AB. Role of TRPM2 channel in mediating H2O2-induced Ca2+ entry and endothelial hyperpermeability. Circ Res. 2008;102:347-355.  [PubMed]  [DOI]  [Cited in This Article: ]
321.  Hecquet CM, Ahmmed GU, Malik AB. TRPM2 channel regulates endothelial barrier function. Adv Exp Med Biol. 2010;661:155-167.  [PubMed]  [DOI]  [Cited in This Article: ]
322.  Guinamard R, Demion M, Launay P. Physiological roles of the TRPM4 channel extracted from background currents. Physiology (. Bethesda). 2010;25:155-164.  [PubMed]  [DOI]  [Cited in This Article: ]
323.  Wu LJ, Sweet TB, Clapham DE. International Union of Basic and Clinical Pharmacology. LXXVI. Current progress in the mammalian TRP ion channel family. Pharmacol Rev. 2010;62:381-404.  [PubMed]  [DOI]  [Cited in This Article: ]
324.  Gerzanich V, Woo SK, Vennekens R, Tsymbalyuk O, Ivanova S, Ivanov A, Geng Z, Chen Z, Nilius B, Flockerzi V. De novo expression of Trpm4 initiates secondary hemorrhage in spinal cord injury. Nat Med. 2009;15:185-191.  [PubMed]  [DOI]  [Cited in This Article: ]
325.  Becerra A, Echeverría C, Varela D, Sarmiento D, Armisén R, Nuñez-Villena F, Montecinos M, Simon F. Transient receptor potential melastatin 4 inhibition prevents lipopolysaccharide-induced endothelial cell death. Cardiovasc Res. 2011;91:677-684.  [PubMed]  [DOI]  [Cited in This Article: ]
326.  Yogi A, Callera GE, Antunes TT, Tostes RC, Touyz RM. Vascular biology of magnesium and its transporters in hypertension. Magnes Res. 2010;23:S207-S215.  [PubMed]  [DOI]  [Cited in This Article: ]
327.  Baldoli E, Maier JA. Silencing TRPM7 mimics the effects of magnesium deficiency in human microvascular endothelial cells. Angiogenesis. 2012;15:47-57.  [PubMed]  [DOI]  [Cited in This Article: ]
328.  Inoue K, Xiong ZG. Silencing TRPM7 promotes growth/proliferation and nitric oxide production of vascular endothelial cells via the ERK pathway. Cardiovasc Res. 2009;83:547-557.  [PubMed]  [DOI]  [Cited in This Article: ]
329.  Paravicini TM, Yogi A, Mazur A, Touyz RM. Dysregulation of vascular TRPM7 and annexin-1 is associated with endothelial dysfunction in inherited hypomagnesemia. Hypertension. 2009;53:423-429.  [PubMed]  [DOI]  [Cited in This Article: ]
330.  Oancea E, Wolfe JT, Clapham DE. Functional TRPM7 channels accumulate at the plasma membrane in response to fluid flow. Circ Res. 2006;98:245-253.  [PubMed]  [DOI]  [Cited in This Article: ]
331.  Wilson PD. Polycystic kidney disease: new understanding in the pathogenesis. Int J Biochem Cell Biol. 2004;36:1868-1873.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 101]  [Cited by in F6Publishing: 83]  [Article Influence: 4.4]  [Reference Citation Analysis (0)]
332.  AbouAlaiwi WA, Takahashi M, Mell BR, Jones TJ, Ratnam S, Kolb RJ, Nauli SM. Ciliary polycystin-2 is a mechanosensitive calcium channel involved in nitric oxide signaling cascades. Circ Res. 2009;104:860-869.  [PubMed]  [DOI]  [Cited in This Article: ]
333.  Nauli SM, Kawanabe Y, Kaminski JJ, Pearce WJ, Ingber DE, Zhou J. Endothelial cilia are fluid shear sensors that regulate calcium signaling and nitric oxide production through polycystin-1. Circulation. 2008;117:1161-1171.  [PubMed]  [DOI]  [Cited in This Article: ]
334.  Hirano K, Hirano M, Hanada A. Involvement of STIM1 in the proteinase-activated receptor 1-mediated Ca2+ influx in vascular endothelial cells. J Cell Biochem. 2009;108:499-507.  [PubMed]  [DOI]  [Cited in This Article: ]
335.  Feske S. ORAI1 and STIM1 deficiency in human and mice: roles of store-operated Ca2+ entry in the immune system and beyond. Immunol Rev. 2009;231:189-209.  [PubMed]  [DOI]  [Cited in This Article: ]
336.  Hogan PG, Lewis RS, Rao A. Molecular basis of calcium signaling in lymphocytes: STIM and ORAI. Annu Rev Immunol. 2010;28:491-533.  [PubMed]  [DOI]  [Cited in This Article: ]
337.  Giachini FR, Lima VV, Hannan JL, Carneiro FS, Webb RC, Tostes RC. STIM1/Orai1-mediated store-operated Ca2+ entry: the tip of the iceberg. Braz J Med Biol Res. 2011;44:1080-1087.  [PubMed]  [DOI]  [Cited in This Article: ]
338.  Beech DJ. Orai1 calcium channels in the vasculature. Pflugers Arch. 2012;463:635-647.  [PubMed]  [DOI]  [Cited in This Article: ]
339.  Mumtaz S, Burdyga G, Borisova L, Wray S, Burdyga T. The mechanism of agonist induced Ca2+ signalling in intact endothelial cells studied confocally in in situ arteries. Cell Calcium. 2011;49:66-77.  [PubMed]  [DOI]  [Cited in This Article: ]
340.  Malli R, Naghdi S, Romanin C, Graier WF. Cytosolic Ca2+ prevents the subplasmalemmal clustering of STIM1: an intrinsic mechanism to avoid Ca2+ overload. J Cell Sci. 2008;121:3133-3139.  [PubMed]  [DOI]  [Cited in This Article: ]
341.  Madge L, Marshall IC, Taylor CW. Delayed autoregulation of the Ca2+ signals resulting from capacitative Ca2+ entry in bovine pulmonary artery endothelial cells. J Physiol. 1997;498:351-369.  [PubMed]  [DOI]  [Cited in This Article: ]
342.  Brandman O, Liou J, Park WS, Meyer T. STIM2 is a feedback regulator that stabilizes basal cytosolic and endoplasmic reticulum Ca2+ levels. Cell. 2007;131:1327-1339.  [PubMed]  [DOI]  [Cited in This Article: ]
343.  Dömötör E, Abbott NJ, Adam-Vizi V. Na+-Ca2+ exchange and its implications for calcium homeostasis in primary cultured rat brain microvascular endothelial cells. J Physiol. 1999;515:147-155.  [PubMed]  [DOI]  [Cited in This Article: ]
344.  Goto Y, Miura M, Iijima T. Extrusion mechanisms of intracellular Ca2+ in human aortic endothelial cells. Eur J Pharmacol. 1996;314:185-192.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 23]  [Cited by in F6Publishing: 24]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
345.  Wang X, Reznick S, Li P, Liang W, van Breemen C. Ca(2+) removal mechanisms in freshly isolated rabbit aortic endothelial cells. Cell Calcium. 2002;31:265-277.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 28]  [Cited by in F6Publishing: 29]  [Article Influence: 1.3]  [Reference Citation Analysis (0)]
346.  Li L, van Breemen C. Na(+)-Ca2+ exchange in intact endothelium of rabbit cardiac valve. Circ Res. 1995;76:396-404.  [PubMed]  [DOI]  [Cited in This Article: ]
347.  Paltauf-Doburzynska J, Frieden M, Graier WF. Mechanisms of Ca2+ store depletion in single endothelial cells in a Ca(2+)-free environment. Cell Calcium. 1999;25:345-353.  [PubMed]  [DOI]  [Cited in This Article: ]
348.  Sedova M, Blatter LA. Dynamic regulation of [Ca2+]i by plasma membrane Ca(2+)-ATPase and Na+/Ca2+ exchange during capacitative Ca2+ entry in bovine vascular endothelial cells. Cell Calcium. 1999;25:333-343.  [PubMed]  [DOI]  [Cited in This Article: ]
349.  Moccia F, Berra-Romani R, Baruffi S, Spaggiari S, Signorelli S, Castelli L, Magistretti J, Taglietti V, Tanzi F. Ca2+ uptake by the endoplasmic reticulum Ca2+-ATPase in rat microvascular endothelial cells. Biochem J. 2002;364:235-244.  [PubMed]  [DOI]  [Cited in This Article: ]
350.  Brini M, Carafoli E. Calcium pumps in health and disease. Physiol Rev. 2009;89:1341-1378.  [PubMed]  [DOI]  [Cited in This Article: ]
351.  Anger M, Samuel JL, Marotte F, Wuytack F, Rappaport L, Lompré AM. The sarco(endo)plasmic reticulum Ca(2+)-ATPase mRNA isoform, SERCA 3, is expressed in endothelial and epithelial cells in various organs. FEBS Lett. 1993;334:45-48.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 52]  [Cited by in F6Publishing: 57]  [Article Influence: 1.8]  [Reference Citation Analysis (0)]
352.  Hadri L, Ozog A, Soncin F, Lompré AM. Basal transcription of the mouse sarco(endo)plasmic reticulum Ca2+-ATPase type 3 gene in endothelial cells is controlled by Ets-1 and Sp1. J Biol Chem. 2002;277:36471-36478.  [PubMed]  [DOI]  [Cited in This Article: ]
353.  Khan I, Sandhu V, Misquitta CM, Grover AK. SERCA pump isoform expression in endothelium of veins and arteries: every endothelium is not the same. Mol Cell Biochem. 2000;203:11-15.  [PubMed]  [DOI]  [Cited in This Article: ]
354.  Ozog A, Pouzet B, Bobe R, Lompré AM. Characterization of the 3' end of the mouse SERCA 3 gene and tissue distribution of mRNA spliced variants. FEBS Lett. 1998;427:349-352.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 20]  [Article Influence: 0.8]  [Reference Citation Analysis (0)]
355.  Szewczyk MM, Davis KA, Samson SE, Simpson F, Rangachari PK, Grover AK. Ca2+-pumps and Na2+-Ca2+-exchangers in coronary artery endothelium versus smooth muscle. J Cell Mol Med. 2007;11:129-138.  [PubMed]  [DOI]  [Cited in This Article: ]
356.  Wu KD, Lee WS, Wey J, Bungard D, Lytton J. Localization and quantification of endoplasmic reticulum Ca(2+)-ATPase isoform transcripts. Am J Physiol. 1995;269:C775-C784.  [PubMed]  [DOI]  [Cited in This Article: ]
357.  Hadri L, Pavoine C, Lipskaia L, Yacoubi S, Lompré AM. Transcription of the sarcoplasmic/endoplasmic reticulum Ca2+-ATPase type 3 gene, ATP2A3, is regulated by the calcineurin/NFAT pathway in endothelial cells. Biochem J. 2006;394:27-33.  [PubMed]  [DOI]  [Cited in This Article: ]
358.  Liu LH, Paul RJ, Sutliff RL, Miller ML, Lorenz JN, Pun RY, Duffy JJ, Doetschman T, Kimura Y, MacLennan DH. Defective endothelium-dependent relaxation of vascular smooth muscle and endothelial cell Ca2+ signaling in mice lacking sarco(endo)plasmic reticulum Ca2+-ATPase isoform 3. J Biol Chem. 1997;272:30538-30545.  [PubMed]  [DOI]  [Cited in This Article: ]
359.  Dedkova EN, Blatter LA. Nitric oxide inhibits capacitative Ca2+ entry and enhances endoplasmic reticulum Ca2+ uptake in bovine vascular endothelial cells. J Physiol. 2002;539:77-91.  [PubMed]  [DOI]  [Cited in This Article: ]
360.  Sutliff RL, Hoying JB, Kadambi VJ, Kranias EG, Paul RJ. Phospholamban is present in endothelial cells and modulates endothelium-dependent relaxation. Evidence from phospholamban gene-ablated mice. Circ Res. 1999;84:360-364.  [PubMed]  [DOI]  [Cited in This Article: ]
361.  Grover AK, Samson SE. Peroxide resistance of ER Ca2+ pump in endothelium: implications to coronary artery function. Am J Physiol. 1997;273:C1250-C1258.  [PubMed]  [DOI]  [Cited in This Article: ]
362.  Schmidt T, Zaib F, Samson SE, Kwan CY, Grover AK. Peroxynitrite resistance of sarco/endoplasmic reticulum Ca2+ pump in pig coronary artery endothelium and smooth muscle. Cell Calcium. 2004;36:77-82.  [PubMed]  [DOI]  [Cited in This Article: ]
363.  Cook NL, Viola HM, Sharov VS, Hool LC, Schöneich C, Davies MJ. Myeloperoxidase-derived oxidants inhibit sarco/endoplasmic reticulum Ca2+-ATPase activity and perturb Ca2+ homeostasis in human coronary artery endothelial cells. Free Radic Biol Med. 2012;52:951-961.  [PubMed]  [DOI]  [Cited in This Article: ]
364.  Dong Y, Zhang M, Wang S, Liang B, Zhao Z, Liu C, Wu M, Choi HC, Lyons TJ, Zou MH. Activation of AMP-activated protein kinase inhibits oxidized LDL-triggered endoplasmic reticulum stress in vivo. Diabetes. 2010;59:1386-1396.  [PubMed]  [DOI]  [Cited in This Article: ]
365.  Dong Y, Zhang M, Liang B, Xie Z, Zhao Z, Asfa S, Choi HC, Zou MH. Reduction of AMP-activated protein kinase alpha2 increases endoplasmic reticulum stress and atherosclerosis in vivo. Circulation. 2010;121:792-803.  [PubMed]  [DOI]  [Cited in This Article: ]
366.  Lytton J. Na+/Ca2+ exchangers: three mammalian gene families control Ca2+ transport. Biochem J. 2007;406:365-382.  [PubMed]  [DOI]  [Cited in This Article: ]
367.  Minelli A, Castaldo P, Gobbi P, Salucci S, Magi S, Amoroso S. Cellular and subcellular localization of Na+-Ca2+ exchanger protein isoforms, NCX1, NCX2, and NCX3 in cerebral cortex and hippocampus of adult rat. Cell Calcium. 2007;41:221-234.  [PubMed]  [DOI]  [Cited in This Article: ]
368.  Yamaji R, Fujita K, Takahashi S, Yoneda H, Nagao K, Masuda W, Naito M, Tsuruo T, Miyatake K, Inui H. Hypoxia up-regulates glyceraldehyde-3-phosphate dehydrogenase in mouse brain capillary endothelial cells: involvement of Na+/Ca2+ exchanger. Biochim Biophys Acta. 2003;1593:269-276.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 76]  [Cited by in F6Publishing: 81]  [Article Influence: 3.9]  [Reference Citation Analysis (0)]
369.  Gobbi P, Castaldo P, Minelli A, Salucci S, Magi S, Corcione E, Amoroso S. Mitochondrial localization of Na+/Ca2+ exchangers NCX1-3 in neurons and astrocytes of adult rat brain in situ. Pharmacol Res. 2007;56:556-565.  [PubMed]  [DOI]  [Cited in This Article: ]
370.  Kaye DM, Kelly RA. Expression and regulation of the sodium-calcium exchanger in cardiac microvascular endothelial cells. Clin Exp Pharmacol Physiol. 1999;26:651-655.  [PubMed]  [DOI]  [Cited in This Article: ]
371.  Eder P, Poteser M, Romanin C, Groschner K. Na(+) entry and modulation of Na(+)/Ca(2+) exchange as a key mechanism of TRPC signaling. Pflugers Arch. 2005;451:99-104.  [PubMed]  [DOI]  [Cited in This Article: ]
372.  Poburko D, Fameli N, Kuo KH, van Breemen C. Ca2+ signaling in smooth muscle: TRPC6, NCX and LNats in nanodomains. Channels (. Austin). 2008;2:10-12.  [PubMed]  [DOI]  [Cited in This Article: ]
373.  Bondarenko A. Sodium-calcium exchanger contributes to membrane hyperpolarization of intact endothelial cells from rat aorta during acetylcholine stimulation. Br J Pharmacol. 2004;143:9-18.  [PubMed]  [DOI]  [Cited in This Article: ]
374.  Andrikopoulos P, Baba A, Matsuda T, Djamgoz MB, Yaqoob MM, Eccles SA. Ca2+ influx through reverse mode Na+/Ca2+ exchange is critical for vascular endothelial growth factor-mediated extracellular signal-regulated kinase (ERK) 1/2 activation and angiogenic functions of human endothelial cells. J Biol Chem. 2011;286:37919-37931.  [PubMed]  [DOI]  [Cited in This Article: ]
375.  Berra-Romani R, Raqeeb A, Guzman-Silva A, Torres-Jácome J, Tanzi F, Moccia F. Na+-Ca2+ exchanger contributes to Ca2+ extrusion in ATP-stimulated endothelium of intact rat aorta. Biochem Biophys Res Commun. 2010;395:126-130.  [PubMed]  [DOI]  [Cited in This Article: ]
376.  Pande J, Mallhi KK, Sawh A, Szewczyk MM, Simpson F, Grover AK. Aortic smooth muscle and endothelial plasma membrane Ca2+ pump isoforms are inhibited differently by the extracellular inhibitor caloxin 1b1. Am J Physiol Cell Physiol. 2006;290:C1341-C1349.  [PubMed]  [DOI]  [Cited in This Article: ]
377.  Di Leva F, Domi T, Fedrizzi L, Lim D, Carafoli E. The plasma membrane Ca2+ ATPase of animal cells: structure, function and regulation. Arch Biochem Biophys. 2008;476:65-74.  [PubMed]  [DOI]  [Cited in This Article: ]
378.  Yokomori H, Oda M, Ogi M, Yoshimura K, Nomura M, Fujimaki K, Kamegaya Y, Tsukada N, Ishii H. Endothelin-1 suppresses plasma membrane Ca++-ATPase, concomitant with contraction of hepatic sinusoidal endothelial fenestrae. Am J Pathol. 2003;162:557-566.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 11]  [Cited by in F6Publishing: 11]  [Article Influence: 0.5]  [Reference Citation Analysis (0)]
379.  Manoonkitiwongsa PS, Whitter EF, Wareesangtip W, McMillan PJ, Nava PB, Schultz RL. Calcium-dependent ATPase unlike ecto-ATPase is located primarily on the luminal surface of brain endothelial cells. Histochem J. 2000;32:313-324.  [PubMed]  [DOI]  [Cited in This Article: ]
380.  Zhang J, Xiao P, Zhang X. Phosphatidylserine externalization in caveolae inhibits Ca2+ efflux through plasma membrane Ca2+-ATPase in ECV304. Cell Calcium. 2009;45:177-184.  [PubMed]  [DOI]  [Cited in This Article: ]
381.  Holton M, Mohamed TM, Oceandy D, Wang W, Lamas S, Emerson M, Neyses L, Armesilla AL. Endothelial nitric oxide synthase activity is inhibited by the plasma membrane calcium ATPase in human endothelial cells. Cardiovasc Res. 2010;87:440-448.  [PubMed]  [DOI]  [Cited in This Article: ]
382.  Szewczyk MM, Pande J, Akolkar G, Grover AK. Caloxin 1b3: a novel plasma membrane Ca(2+)-pump isoform 1 selective inhibitor that increases cytosolic Ca(2+) in endothelial cells. Cell Calcium. 2010;48:352-357.  [PubMed]  [DOI]  [Cited in This Article: ]
383.  Malli R, Graier WF. Mitochondrial Ca2+ channels: Great unknowns with important functions. FEBS Lett. 2010;584:1942-1947.  [PubMed]  [DOI]  [Cited in This Article: ]
384.  Trenker M, Malli R, Fertschai I, Levak-Frank S, Graier WF. Uncoupling proteins 2 and 3 are fundamental for mitochondrial Ca2+ uniport. Nat Cell Biol. 2007;9:445-452.  [PubMed]  [DOI]  [Cited in This Article: ]
385.  Waldeck-Weiermair M, Malli R, Naghdi S, Trenker M, Kahn MJ, Graier WF. The contribution of UCP2 and UCP3 to mitochondrial Ca(2+) uptake is differentially determined by the source of supplied Ca(2+). Cell Calcium. 2010;47:433-440.  [PubMed]  [DOI]  [Cited in This Article: ]
386.  Waldeck-Weiermair M, Jean-Quartier C, Rost R, Khan MJ, Vishnu N, Bondarenko AI, Imamura H, Malli R, Graier WF. Leucine zipper EF hand-containing transmembrane protein 1 (Letm1) and uncoupling proteins 2 and 3 (UCP2/3) contribute to two distinct mitochondrial Ca2+ uptake pathways. J Biol Chem. 2011;286:28444-28455.  [PubMed]  [DOI]  [Cited in This Article: ]
387.  Drago I, Pizzo P, Pozzan T. After half a century mitochondrial calcium in- and efflux machineries reveal themselves. EMBO J. 2011;30:4119-4125.  [PubMed]  [DOI]  [Cited in This Article: ]
388.  Malli R, Frieden M, Hunkova M, Trenker M, Graier WF. Ca2+ refilling of the endoplasmic reticulum is largely preserved albeit reduced Ca2+ entry in endothelial cells. Cell Calcium. 2007;41:63-76.  [PubMed]  [DOI]  [Cited in This Article: ]
389.  Graier WF, Paltauf-Doburzynska J, Hill BJ, Fleischhacker E, Hoebel BG, Kostner GM, Sturek M. Submaximal stimulation of porcine endothelial cells causes focal Ca2+ elevation beneath the cell membrane. J Physiol. 1998;506:109-125.  [PubMed]  [DOI]  [Cited in This Article: ]
390.  Bootman MD, Berridge MJ, Lipp P. Cooking with calcium: the recipes for composing global signals from elementary events. Cell. 1997;91:367-373.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 288]  [Cited by in F6Publishing: 305]  [Article Influence: 11.3]  [Reference Citation Analysis (0)]
391.  Clapham DE. Calcium signaling. Cell. 2007;131:1047-1058.  [PubMed]  [DOI]  [Cited in This Article: ]
392.  Rizzuto R, Pozzan T. Microdomains of intracellular Ca2+: molecular determinants and functional consequences. Physiol Rev. 2006;86:369-408.  [PubMed]  [DOI]  [Cited in This Article: ]
393.  Hüser J, Blatter LA. Elementary events of agonist-induced Ca2+ release in vascular endothelial cells. Am J Physiol. 1997;273:C1775-C1782.  [PubMed]  [DOI]  [Cited in This Article: ]
394.  Burdyga T, Shmygol A, Eisner DA, Wray S. A new technique for simultaneous and in situ measurements of Ca2+ signals in arteriolar smooth muscle and endothelial cells. Cell Calcium. 2003;34:27-33.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 40]  [Cited by in F6Publishing: 42]  [Article Influence: 2.0]  [Reference Citation Analysis (0)]
395.  Duza T, Sarelius IH. Localized transient increases in endothelial cell Ca2+ in arterioles in situ: implications for coordination of vascular function. Am J Physiol Heart Circ Physiol. 2004;286:H2322-H2331.  [PubMed]  [DOI]  [Cited in This Article: ]
396.  Kansui Y, Garland CJ, Dora KA. Enhanced spontaneous Ca2+ events in endothelial cells reflect signalling through myoendothelial gap junctions in pressurized mesenteric arteries. Cell Calcium. 2008;44:135-146.  [PubMed]  [DOI]  [Cited in This Article: ]
397.  Duza T, Sarelius IH. Increase in endothelial cell Ca(2+) in response to mouse cremaster muscle contraction. J Physiol. 2004;555:459-469.  [PubMed]  [DOI]  [Cited in This Article: ]
398.  Isakson BE, Ramos SI, Duling BR. Ca2+ and inositol 1,4,5-trisphosphate-mediated signaling across the myoendothelial junction. Circ Res. 2007;100:246-254.  [PubMed]  [DOI]  [Cited in This Article: ]
399.  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. 2008;105:9627-9632.  [PubMed]  [DOI]  [Cited in This Article: ]
400.  Nausch LW, Bonev AD, Heppner TJ, Tallini Y, Kotlikoff MI, Nelson MT. Sympathetic nerve stimulation induces local endothelial Ca2+ signals to oppose vasoconstriction of mouse mesenteric arteries. Am J Physiol Heart Circ Physiol. 2012;302:H594-H602.  [PubMed]  [DOI]  [Cited in This Article: ]
401.  Missiaen L, Lemaire FX, Parys JB, De Smedt H, Sienaert I, Casteels R. Initiation sites for Ca2+ signals in endothelial cells. Pflugers Arch. 1996;431:318-324.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 16]  [Cited by in F6Publishing: 16]  [Article Influence: 0.6]  [Reference Citation Analysis (0)]
402.  Isshiki M, Ando J, Korenaga R, Kogo H, Fujimoto T, Fujita T, Kamiya A. Endothelial Ca2+ waves preferentially originate at specific loci in caveolin-rich cell edges. Proc Natl Acad Sci USA. 1998;95:5009-5014.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 110]  [Cited by in F6Publishing: 109]  [Article Influence: 4.2]  [Reference Citation Analysis (0)]
403.  Hu Q, Deshpande S, Irani K, Ziegelstein RC. [Ca(2+)](i) oscillation frequency regulates agonist-stimulated NF-kappaB transcriptional activity. J Biol Chem. 1999;274:33995-33998.  [PubMed]  [DOI]  [Cited in This Article: ]
404.  Scharbrodt W, Abdallah Y, Kasseckert SA, Gligorievski D, Piper HM, Böker DK, Deinsberger W, Oertel MF. Cytosolic Ca2+ oscillations in human cerebrovascular endothelial cells after subarachnoid hemorrhage. J Cereb Blood Flow Metab. 2009;29:57-65.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 21]  [Cited by in F6Publishing: 22]  [Article Influence: 1.5]  [Reference Citation Analysis (0)]
405.  Zhu L, Luo Y, Chen T, Chen F, Wang T, Hu Q. Ca2+ oscillation frequency regulates agonist-stimulated gene expression in vascular endothelial cells. J Cell Sci. 2008;121:2511-2518.  [PubMed]  [DOI]  [Cited in This Article: ]
406.  Carter TD, Bogle RG, Bjaaland T. Spiking of intracellular calcium ion concentration in single cultured pig aortic endothelial cells stimulated with ATP or bradykinin. Biochem J. 1991;278:697-704.  [PubMed]  [DOI]  [Cited in This Article: ]
407.  Hallam TJ, Jacob R, Merritt JE, Rink TJ. cytoplasmic calcium oscillations in cultured human umbilical vein endothelial-cells stimulated by histamine. J Physiol. 1988;403:93.  [PubMed]  [DOI]  [Cited in This Article: ]
408.  Jacob R, Merritt JE, Hallam TJ, Rink TJ. Repetitive spikes in cytoplasmic calcium evoked by histamine in human endothelial cells. Nature. 1988;335:40-45.  [PubMed]  [DOI]  [Cited in This Article: ]
409.  Morgan AJ, Jacob R. Ca2+ influx does more than provide releasable Ca2+ to maintain repetitive spiking in human umbilical vein endothelial cells. Biochem J. 1996;320:505-517.  [PubMed]  [DOI]  [Cited in This Article: ]
410.  Sedova M, Klishin A, Huser J, Blatter LA. Capacitative Ca2+ entry is graded with degree of intracellular Ca2+ store depletion in bovine vascular endothelial cells. J Physiol. 2000;523 Pt 3:549-559.  [PubMed]  [DOI]  [Cited in This Article: ]
411.  Morgan AJ, Jacob R. Differential modulation of the phases of a Ca2+ spike by the store Ca2+-ATPase in human umbilical vein endothelial cells. J Physiol. 1998;513:83-101.  [PubMed]  [DOI]  [Cited in This Article: ]
412.  Koyama T, Kimura C, Park SJ, Oike M, Ito Y. Functional implications of Ca2+ mobilizing properties for nitric oxide production in aortic endothelium. Life Sci. 2002;72:511-520.  [PubMed]  [DOI]  [Cited in This Article: ]  [Cited by in Crossref: 17]  [Cited by in F6Publishing: 19]  [Article Influence: 0.9]  [Reference Citation Analysis (0)]
413.  Hu Q, Zheng G, Zweier JL, Deshpande S, Irani K, Ziegelstein RC. NADPH oxidase activation increases the sensitivity of intracellular Ca2+ stores to inositol 1,4,5-trisphosphate in human endothelial cells. J Biol Chem. 2000;275:15749-15757.  [PubMed]  [DOI]  [Cited in This Article: ]
414.  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. 2012;110:1435-1444.  [PubMed]  [DOI]  [Cited in This Article: ]
415.  Walseth TF, Lin-Moshier Y, Jain P, Ruas M, Parrington J, Galione A, Marchant JS, Slama JT. Photoaffinity labeling of high affinity nicotinic acid adenine dinucleotide phosphate (NAADP)-binding proteins in sea urchin egg. J Biol Chem. 2012;287:2308-2315.  [PubMed]  [DOI]  [Cited in This Article: ]
416.  Lin-Moshier Y, Walseth TF, Churamani D, Davidson SM, Slama JT, Hooper R, Brailoiu E, Patel S, Marchant JS. Photoaffinity labeling of nicotinic acid adenine dinucleotide phosphate (NAADP) targets in mammalian cells. J Biol Chem. 2012;287:2296-2307.  [PubMed]  [DOI]  [Cited in This Article: ]