Open access peer-reviewed chapter

The Endothelium: The Vascular Information Exchange

Written By

Ran Wei, Stephanie E. Lunn, Stephen L. Gust, Paul M. Kerr and Frances Plane

Submitted: 16 February 2018 Reviewed: 03 July 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.79897

From the Edited Volume

Calcium and Signal Transduction

Edited by John N. Buchholz and Erik J. Behringer

Chapter metrics overview

1,336 Chapter Downloads

View Full Metrics

Abstract

Maintenance of adequate blood flow to tissues and organs requires that endothelial cells dynamically respond in a stimulus-specific manner to elicit appropriate changes in smooth muscle contractility and thus, arterial diameter. Endothelial cells can be stimulated directly by increases in blood flow and by humoral factors acting on surface receptors, as well as through flux of second messengers from smooth muscle cells activated by release of neurotransmitters from perivascular nerves. The ability of endothelial cells to generate stimulus-specific responses to these diverse inputs is facilitated by organization of ion channels and signaling proteins into microdomains that permit finely-tuned, spatially-restricted Ca2+ events to differentially activate key effectors such as nitric oxide (NO) synthase and Ca2+-activated K+ (KCa) channels. NO is a diffusible mediator which acts locally to cause vasodilation. Opening of KCa channels causes hyperpolarization of the endothelial membrane potential which spreads to surrounding smooth muscle cells to also cause local vasodilation. However, once initiated, hyperpolarization also spreads longitudinally through the endothelium to effect coordinated changes in blood flow within multiple arterial segments. Thus, the signaling pathways activated by a particular stimulus determine whether it’s effects on arterial diameter are localized or can impact blood flow at the level of the vascular bed.

Keywords

  • endothelium
  • calcium
  • nitric oxide
  • microdomain
  • potassium channels

1. Introduction

Appropriate local control of blood flow through resistance arteries is critical to the functioning of tissues and organs, and to regulation of blood pressure. Lying at the interface between the blood and smooth muscle cells of the vessel wall, the endothelium plays a vital role in this dynamic process by transducing diverse chemical and mechanical stimuli into coordinated changes in arterial diameter. Endothelial cells respond to vasodilatory stimuli by releasing diffusible mediators such as nitric oxide (NO) and prostacyclin (PGI2) and by initiating membrane hyperpolarization that spreads to surrounding smooth muscle cells via myoendothelial gap junctions (MEGJs) to inhibit contractility, a mechanism termed endothelium-dependent hyperpolarization (EDH) [1, 2, 3]. NO and PGI2 are local mediators that diffuse to surrounding smooth muscle cells to cause relaxation. Once initiated, EDH spreads to surrounding smooth muscle cells to affect relaxation but conduction of hyperpolarization longitudinally through the endothelial layer means that EDH also contributes to coordination of changes in blood flow in multiple arterial segments within a vascular bed [4]. Thus, the ability of a stimulus to engage diffusible mediators versus EDH determines whether it’s effects on arterial diameter and thus blood flow, are restricted to the local area or can impact blood flow at the level of the vascular bed.

Increases in endothelial intracellular Ca2+ concentration ([Ca2+]i) drive these vasodilator pathways; NO is synthesized from L-arginine by the Ca2+-calmodulin-dependent enzyme NO synthase (NOS) [3], PGI2 is generated by the actions of cyclooxygenase on arachidonic acid released by the actions of Ca2+-dependent phospholipase A2 on membrane phospholipids [5, 6], and opening of Ca2+-activated K+ (KCa) channels causes hyperpolarization [3, 7]. Global changes in endothelial [Ca2+]i have been widely studied [8, 9], but development of new technologies such as high-speed, high-resolution confocal Ca2+ imaging and generation of transgenic mice expressing genetically encoded Ca2+ indicators has allowed resolution of a wide range of transient Ca2+ events within endothelial cells of intact arteries to provide a growing body of support for the concept of stimulus-specific engagement of effectors underpinned by spatially and temporally discrete Ca2+ signaling patterns that occur independently of changes in bulk endothelial [Ca2+]i [10, 11, 12, 13].

Pulsars [10] and wavelets [11] are spontaneous, short-lived, (<0.03 s duration) spatially fixed Ca2+ events originating from distinct clusters of inositol 1,4,5-trisphosphate (InsP3) receptors on the membrane of endoplasmic reticulum (ER). First identified in mouse mesenteric artery and hamster skeletal muscle arteriolar endothelial cells, these events predominantly occur close to endothelial projections that abut or form MEGJs with smooth muscle cells [10, 11] and exert a basal vasodilator influence through activation of intermediate conductance (IKCa) Ca2+-activated K+ channels. Their dependence on InsP3 provides a mechanism by which pulsars are linked to and regulated by G protein-coupled receptor (GPCR) signaling. Elevation of InsP3 by endothelium-dependent vasodilators [10] or by flux of InsP3 from smooth muscle cells following stimulation of α1-adrenoceptors [11] increases pulsar size and/or frequency through recruitment of new sites and a reduction in the interval between pulsars at a given site. In porcine coronary arteries, InsP3-dependent Ca2+ events similar to pulsars propagate into longer lasting Ca2+ waves (>8 s duration) facilitated by the longitudinal arrangement of ER/InsP3 receptors to promote directional Ca2+-induced Ca2+ release along the endothelial cell axis and are associated with activation of both NOS and KCa channels [12].

Sparklets are generated by spatially restricted Ca2+ influx through members of the transient receptor potential (TRP) ion channel family [14, 15]. Sparklets were first identified in mouse mesenteric arteries under experimental conditions in which InsP3-mediated pulsars were eliminated [14]. Exposure of the endothelium to TRPV4 agonists and/or acetylcholine increased the activity of these discrete Ca2+ signals which were linked to activation of both IKCa and small conductance (SKCa) Ca2+-activated K+ channels, effects which were absent in arteries from mice lacking TRPV4 [14]. In rat cremaster arterioles, clustering of TRPV4-mediated sparklets in endothelial projections was linked exclusively to activation of IKCa channels [16] and in mouse small pulmonary arteries, shear stress-stimulated TRPV4 activity was linked to NO production [17]. Larger endothelial sparklets mediated by simultaneous opening of two TRPA1 and leading to activation of IKCa channels were shown to underlie dilation to reactive oxygen species in rat cerebral arteries [18]. We will now discuss how grouping of Ca2+ signaling and effector proteins into microdomains allows dynamic, stimulus-specific Ca2+ events which determine the recruitment of effectors thus the degree to which blood flow is impacted.

Advertisement

2. Stimulus-specific endothelial Ca2+ signaling

2.1. Shear stress

In vivo, endothelial sensing of laminar shear stress, the tangential frictional force exerted by blood flowing across the cell surface, plays a dominant role in acute modulation of vascular tone and therefore, tissue perfusion [19, 20, 21]. In the majority of resistance arteries, increases in blood flow stimulate endothelium-dependent relaxation of surrounding smooth muscle cells and increase arterial diameter, a response termed flow-mediated dilatation. Flow also influences gene expression and structural remodeling with areas of disturbed flow and reduced shear stress is associated with development of atherosclerotic plaques [22]. Measurement of acute responses to increases in shear stress is the most widely used clinical index of endothelial function and vascular health with attenuation of flow-induced dilation associated with increased risk of cardiovascular diseases [23, 24]. Indeed, reductions in shear stress are a likely mechanism by which endothelial function is altered with inactivity, an effect which can be overcome by exercise interventions [25, 26].

In animals and humans, acute shear stress-induced vasodilation can be mediated by both NO and EDH [27, 28, 29, 30, 31]. Although endothelial cells express both SKCa and IKCa channels, data from isolated arteries indicate that it is SKCa channels that play a predominate role in mediating the EDH component of this response. Deletion of SKCa but not IKCa channels impaired both NO and EDH-mediated dilation to shear stress stimulation in mouse isolated carotid arteries [32]. In rat isolated perfused mesenteric beds, shear stress-induced modulation of sympathetic vasoconstriction was prevented by both the NOS inhibitor L-NG-nitroarginine methyl ester (L-NAME) and apamin, a selective blocker of SKCa channels, but not by the IKCa channel inhibitor TRAM-34 [33]. Similarly, shear stress-evoked dilation of mouse isolated coronary arteries was inhibited by apamin [34] and L-NAME [35].

Mechanotransduction, the conversion of increases in shear stress into changes in arterial diameter, is reliant upon rises in endothelial [Ca2+]i mediated by Ca2+ entry. In vitro studies have identified multiple candidates as potential endothelial mechanosensors including integrins [36], tyrosine kinase receptors [37], intercellular junction proteins [38], and P2X4 receptors which are cation channels activated by adenosine triphosphate (ATP) [39]. Work with transgenic mice has implicated GPR68, a proton-sensing rhodopsin-like GPCR [40], inwardly rectifying K+ channels [41] and PIEZO1, a Ca2+-permeable, non-selective cation channel [42, 43]. However, recently, it is the role of TRPV4 in endothelial responses to increases in shear stress that has received particular attention. This Ca2+ permeable channel can be directly activated by shear stress [34, 44] via membrane deformation or through a lever-like action involving cytoskeletal linkages to molecules embedded in the glycocalyx [45, 46], the layer of proteoglycans and glycoproteins that covers the luminal surface of the endothelium, or indirectly through upstream production of arachidonic acid metabolites [47]. Genetic deletion of TRPV4 results in blunted flow-mediated dilation of mouse carotid and mesenteric arteries [31, 47, 48] and pharmacological inhibition of these channels blocked flow-evoked increases in endothelial [Ca2+]i in isolated mouse mesenteric, human coronary, and rat carotid and gracilis arteries [31, 49, 50, 51]. In bovine coronary endothelial cells block of TRPV4 inhibited both shear stress-evoked increases in [Ca2+]i and activation of SKCa channels [34], indicating that there may be a direct link between TRPV4-mediated Ca2+ influx and SKCa channel activity. This idea is supported by the demonstration that in rat pulmonary arteries, vasodilation to the TRPV4 agonist GSK1016790A was mediated by activation of SKCa channels [52]. In the same vessels, and in mouse small pulmonary arteries, shear stress-stimulated TRPV4 activity was also linked to NO production [17] suggesting a further link between TRPV4 and NOS.

Building on these findings, several lines of evidence now support the notion that acute increases in shear stress cause Ca2+ influx through TRPV4 channels to selectively activate both SKCa channels and NOS, and that this pathway is enabled by organization of TRPV4, NOS, SKCa channels and the caveolae scaffold protein caveolin-1 into microdomains within caveolae, flask-shaped structures on the endothelial cell surface rich in signaling proteins [34, 53, 54]. SKCa channels are localized to the luminal membrane of endothelial cells in rat mesenteric arteries [55] and SKCa channel protein was co-immunoprecipitated with caveolin-1 from endothelial cells of the same arteries and from porcine coronary arteries [56]. It is well established that endothelial NOS is localized to caveolae where it is negatively regulated through its interaction with caveolin-1 [57]. Increases in [Ca2+]i promote recruitment of Ca2+-calmodulin to displace caveolin-1 from NOS thereby activating it [53]. In bovine coronary and human microvascular endothelial cells, SKCa channels were co-localized with caveolin-1, NOS and TRPV4 channels within microdomains at the luminal endothelial cell surface [34, 57, 58]. Furthermore, in mesenteric arteries from mice lacking caveolin-1, endothelial TRPV4 channel activity was impaired indicating that a direct interaction between TRPV4 and caveolin-1 may be functionally important for Ca2+ entry in response to shear stress [57]. Caveolin-1 has been shown to initiate downstream signaling in response to increases in shear stress [59] leading to the suggestion that caveolae act as mechanosensors to elicit a cascade of events that promote vasodilation. In line with this proposal, shear stress-induced dilation is defective and endothelial SKCa current reduced in coronary and carotid arteries of mice lacking caveolin-1, an effect rescued by re-introduction of endothelial specific caveolin-1 [58, 60].

Together these findings, gathered using a range of approaches and from a number of different arteries, support an elegant model in which shear stress-evoked Ca2+ influx through TRPV4 channels on the luminal surface of endothelial cells leads to spatially-restricted Ca2+ sparklets within a signaling microdomain to selectively activate SKCa channels and NOS to elicit vasodilation (Figure 1). However, a number of questions remain to be addressed. Shear stress increases PGI2 production in bovine coronary and human umbilical vein endothelial cells [34, 61] and rabbit isolated femoral arteries [62], and hydrogen peroxide (H2O2) contributes to flow-mediated dilation in coronary arterioles from patients with heart disease [63] but the functional role of these factors in acute flow-mediated vasodilation has not fully been explored. A significant component of flow-induced dilation remained in isolated mesenteric arteries of mice lacking TRPV4 [31] which could indicate that, as suggested in earlier reports, Ca2+-independent processes may also contribute to this response [64] or the involvement of another route for Ca2+ influx. The possibility that flow-induced increases in endothelial cell [Ca2+]i are stimulated by localized release endothelium-derived paracrine mediators such as ATP, substance P or acetylcholine, first proposed over 30 years ago [65, 66], has recently received renewed support with the demonstration that endothelial organic cation transporters release acetylcholine in response to increases in shear stress in rat isolated carotid arteries [67]. The same study suggests that InsP3-mediated Ca2+ release from ER stores and Ca2+ entry through TRPC but not TRPV4 contributes to flow-induced endothelial Ca2+ signaling in these vessels. This finding highlights the fact that further work is required to elucidate the differential signaling networks underlying endothelial responses to acute increases in shear stress in different arteries.

Figure 1.

Model of localized endothelial Ca2+ signaling evoked by increases in shear stress. Shear stress-evoked Ca2+ influx through TRPV4 channels on the luminal surface of endothelial cells leads to spatially-restricted Ca2+ sparklets within a signaling microdomain to selectively activate SKca channels and endothelial NOS (eNOS).

2.2. Agonists at endothelial GPCRs

Many endogenous and exogenous chemicals bind to GPCRs leading to stimulation of EDH and production of NO, PGI2 and other diffusible mediators such as epoxyeicosatrienoic acids and H2O2, to cause vasodilation [1, 2, 68]. Measurements of bulk endothelial [Ca2+]i established the role of InsP3-mediated Ca2+ release and subsequently store-operated Ca2+ entry (SOCE) in this process [9]. The mechanism underlying endothelial SOCE has been controversial but recent evidence supports a model in which Ca2+ store depletion allows spatial reorganization of Ca2+ sensor protein stromal interaction molecules (STIMs) so that they can aggregate into clusters that physically interact with Ca2+-selective Orai channels at the ER-plasma membrane junction [69, 70, 71, 72]. TRPC and TRPV4 can also interact with STIMs [73] and studies of knock-out mice have provided evidence for a role for TRPC4 in acetylcholine-evoked SOCE in aortae [74] and for TRPV4 in SOCE in mesenteric [75] and carotid arteries [76]. A receptor-operated Ca2+ entry mechanism can also be mediated by DAG-induced activation of TRPC and TRPV channels. For example, in human umbilical vein endothelial cells, bradykinin stimulated both translocation of DAG-sensitive TRPC6 to the cell membrane and Ca2+ influx [77, 78]. Expression of mRNA for another ER Ca2+ release channel, the ryanodine receptor (RyR), has been detected in endothelial cells of human mesenteric arteries [79], and RyRs have been suggested to mediate Ca2+ oscillations in cultured bovine aortic and human umbilical vein endothelial cells [80] but to date, ryanodine has been shown to have no effect on endothelial Ca2+ signaling or vasodilation [12, 81]. There is significant variation in the reported contribution of EDH, NO and other mediators to agonist-evoked dilation, both in terms of differences between agonists and arteries. Thus, for the purposes of this chapter we will limit our discussion to three agents commonly used to stimulate endothelium-dependent vasodilation in experimental studies, acetylcholine, ATP and substance P.

The first evidence that differential endothelial Ca2+ signaling underlies agonist-evoked EDH and NO came from a study of rat isolated middle cerebral arteries in which EDH-dependent vasodilation to purinergic agonists required a larger increase in [Ca2+]i than for NO [82]. Measurements of global [Ca2+]i indicated that different sources of Ca2+ contributed to agonist-stimulated production of NO and EDH; NO production is associated with SOCE [83] whereas EDH is linked to both InsP3-mediated Ca2+ release and SOCE [84]. Similarly, both agonist evoked SOCE and NO production are suppressed in isolated aortae from mice lacking STIM1, the primary endothelial STIM [85]. Building on these findings, data accrued over the past 15 years from functional, histological, Ca2+ imaging and immunohistochemical studies of intact arteries and endothelial-smooth muscle co-cultures support agonist-evoked EDH and NO release being mediated by distinct Ca2+ signaling within specialized domains.

2.2.1. EDH

EDH is mediated by opening of both IKCa and SKCa channels but their relative contribution to agonist-evoked vasodilation, based on the effects of selective pharmacological inhibitors, displays significant variation between agonists and arteries. Simultaneous block of both IKCa and SKCa channels is required to inhibit acetylcholine-evoked EDH in mesenteric arteries from rats and mice, and guinea-pig coronary arteries [86, 87, 88] whereas in rat cerebral and human mesenteric arteries the same response is largely reliant on IKCa channels [89, 90].

IKCa and SKCa channels display a differential spatial distribution within endothelial cells and their contribution to agonist-evoked EDH appears to be mediated by different signaling pathways. In mesenteric arteries from rats, mice and humans, IKCa channels are localized within regions associated with MEGJs in which ER membrane, InsP3 receptors, gap junction connexins, TRPC3 and TRPV4 have also been identified [10, 11, 14, 55, 90, 91, 92, 93]. In mouse and rat mesenteric arteries, acetylcholine increased the frequency and number of InsP3-dependent Ca2+ pulsars in this region which were linked to activation of IKCa channels to evoke EDH [10, 93] (Figure 2). TRPC3 may support this process by providing Ca2+ entry for refilling of InsP3-sensitive ER stores, and/or direct activation of both SKCa and IKCa channels [91, 94]. However, in other reports acetylcholine exclusively stimulates TRPV4 in the vicinity of MEGJs to generate Ca2+ sparklets and which in turn activate IKCa channels in mouse mesenteric arteries [14, 95]. This occurs via a mechanism dependent on the anchoring protein AKAP, and is consistent with deletion of TRPV4 resulting in blunting of acetylcholine-evoked increases in endothelial [Ca2+]i and loss of EDH in mouse mesenteric arteries [75, 76]. TRPV3 [96] and TRPA1 [97] are also expressed in endothelial cells and activators of these channels can certainly initiate increases in Ca2+ signaling and EDH in cerebral arteries, but a role for these channels in agonist-stimulated EDH has yet to be demonstrated.

Figure 2.

Schematic showing discrete Ca2+ signaling events elicited by agonists at endothelial GPCRs. InsP3-dependent Ca2+ pulsars are linked to activation of IKCa channels to evoke EDH whereas Ca2+ influx through DAG-activated TRPC channels is the primary source of Ca2+ for agonist stimulation of endothelial SKCa channels and eNOS.

In contrast to IKCa channels, SKCa channels are associated with caveolae and are diffusely distributed across the cell membrane with a higher level expression at endothelial-endothelial cell borders [12, 55, 90, 93]. Also, unlike IKCa channels, evidence is lacking for a direct link between agonist-evoked, InsP3-mediated Ca2+ events and SKCa channel activity. Instead, it appears that Ca2+ influx through TRP channels is the primary source of Ca2+ for agonist stimulation of endothelial SKCa channels [94, 98]. In mouse cerebral artery, ATP caused rapid trafficking of TRPC3 to the plasma membrane to provide Ca2+ influx to selectively activate SKCa channels to cause EDH [98] (Figure 2). As described earlier, TRPV4 are also associated with caveolae and are a source of Ca2+ for SKCa channel activation in response to increases in shear stress but whether this relationship accounts for engagement of SKCa channels by agonists has not been explored.

2.2.2. NO

In contrast to EDH, the role of localized Ca2+ signaling in agonist-evoked NO release has received little attention. NOS, TRPV4 and TRPC3 are located in caveolae microdomains, and deletion of either channel blunts acetylcholine-evoked NO release and NO-mediated relaxation in mouse mesenteric and carotid arteries [75, 99] suggesting they may provide a source of Ca2+ for agonist-driven NOS activation. Heteromultimers of TRPV4-TRPC1 channels mediate vasorelaxation of rabbit mesenteric arteries in response to stimulation of the Ca2+-sensing receptor through NO production [100] but the underlying Ca2+ dynamics were not assessed. A recent study has shown that TRPV4-mediated sparklets underlie ATP driven activation of endothelial NOS in mouse small pulmonary arteries. The resulting NO initiates vasodilation and also guanylyl cyclase-protein kinase G signaling in the endothelium that limits TRPV4 channel cooperativity and serves as a negative feedback signal to regulate TRPV4 channel function [17]. This description of ATP-evoked, spatially distinct TRPV4 sparklets and localized TRPV4-NOS signaling support a novel paradigm that NOS can be activated by spatially restricted Ca2+ signals, and identifies TRPV4 channels as a key regulator of NOS activity in the pulmonary microcirculation.

In contrast, in porcine isolated coronary arteries, substance P increased the occurrence of discrete InsP3-dependent endothelial Ca2+ events in a concentration-dependent manner; low concentrations primarily increased the number of Ca2+ events and at higher concentrations the number of Ca2+ events saturated while the magnitude of individual events increased [12]. This pattern correlated with a greater role for NO in vasorelaxation at lower concentrations suggesting subtle Ca2+ signal expansion at low stimulation levels may preferentially target NOS. A key finding of this study was that idiosyncratic Ca2+ signal expansion corresponded with coronary artery vasorelaxation whereas global changes in [Ca2+]i did not highlighting that frequency modulation of discrete Ca2+ signals is the primary driver of this functional response and that measurement of changes in bulk [Ca2+]i do not adequately describe the Ca2+ signaling pathways that underlie endothelium-dependent vasodilation.

2.2.3. Membrane potential and Ca2+ microdomain signaling

Production of NO and stimulation of EDH have long been regarded as separate mechanisms for agonist-evoked vasodilation but several lines of evidence indicate that there may be a facilitatory relationship between endothelial SKCa and IKCa channel activity and NO. SKCa channel activity has been linked to NO-mediated vasodilation to agonists with deletion of these channels causing impaired NO-mediated dilation to acetylcholine in mouse carotid arteries and increased expression enhancing NO-mediated dilation of cremaster arterioles [32]. In rat mesenteric arteries, block of SKCa and IKCa channels reduces agonist-evoked, NO-mediated vasorelaxation and NO release [101]. Conversely, activators of endothelial KCa channels can enhance NO release from cultured endothelial cells, enhance ATP-induced increases in cytosolic Ca2+ concentration and NO synthesis in rat cremaster arterioles, and elicit NO-mediated relaxation in mesenteric arteries [102, 103, 104].

Lacking voltage-operated Ca2+ channels, endothelial Ca2+ influx is mediated by TRP channels and so membrane hyperpolarization may be required to maintain an appropriate electrochemical driving force for agonist-induced Ca2+ influx and also to prevent channel inactivation and/or reduction in unitary conductance [105, 106]. Membrane depolarization does inhibit both agonist-induced increases in [Ca2+]i and NO release in cultured endothelial cells [107, 108], and in rat isolated basilar arteries, endothelial depolarization was associated with a reduction in NO-mediated relaxation to acetylcholine [109]. Nonetheless, the ability of hyperpolarization to regulate Ca2+ entry by increasing the electrical driving force is controversial. The large concentration gradient (∼20,000-fold for extracellular versus intracellular) [110] and driving force for Ca2+ entry raising the question of whether a small amplitude hyperpolarization will be insufficient to modulate Ca2+ entry. In rat mesenteric and cerebral arteries, that certainly appeared to be the case as changes in global endothelial [Ca2+]i were independent of changes membrane potential [89, 111]. However, more recent work with endothelial cell tubes isolated from resistance arteries has provided renewed support for hyperpolarization enhancing acetylcholine-evoked Ca2+ influx through TRPV4 [112] and indicate that pharmacological activation of SKCa and IKCa channel may not only enhance Ca2+ entry to further amplify KCa channel activity, but also boost NO production [113]. In mouse mesenteric arteries, acetylcholine-evoked TRPV4-dependent Ca2+ signaling was inhibited in arteries from mice lacking IKCa channels indicating that in these arteries, endothelial stimulation drives sufficient IKCa-dependent Ca2+ entry through TRPV4 to enhance dynamics [13]. IKCa channel activity modestly augmented Ca2+ event amplitude but the most notable impact was in recruiting new Ca2+ firing sites as well as increasing firing frequencies at pre-existing sites. In the same study, increasing or decreasing SKCa expression had little additional effect on the occurrence of Ca2+ events but did promote increased amplitudes and durations indicating that SKCa channels may play a role in positive feedback Ca2+ regulation by shaping the size and time course of individual events. In porcine coronary arteries stimulation of NOS by InsP3-dependent, large amplitude-low frequency Ca2+ waves [12], exactly the types of events which were lost in mesenteric arteries from mice with an endothelial specific knockout of SKCa channels [114], suggests that SKCa channels are required for their development. As mentioned above, deletion of SKCa channels impaired NO-mediated dilation to acetylcholine [32] and together, these findings support the notion that their role in protraction of Ca2+ events may be important in stimulation NOS.

2.3. Myoendothelial feedback

The sympathetic nervous system regulates total peripheral resistance and is a key modulator of resistance artery diameter through release of noradrenaline and co-transmitters such as ATP and neuropeptide Y [115]. Noradrenaline causes vasoconstriction through activating α1-adrenoceptors on vascular smooth muscle cells, a process which is limited by engagement of endothelial mechanisms through myoendothelial feedback. The current model of myoendothelial feedback involves flux of InsP3 from smooth muscle to endothelial cells to elicit localized increases in Ca2+, activation of IKCa channels and possibly NOS, to limit smooth muscle contractility [11, 91, 116]. This model is supported by ultrastructural and histochemical studies showing that in rat mesenteric and basilar, and hamster retractor feed arteries, MEGJ connexins and IKCa channels are in close spatial association with ER and InsP3 receptors within endothelial projections that extend through the internal elastic lamina to make contact with smooth muscle cells [11, 55, 91, 94]. In hamster retractor feed arteries, myoendothelial feedback is fully accounted for by EDH. The α1-adrenoceptor agonist phenylephrine induced localized, InsP3-mediated Ca2+ signaling events within endothelial projections and block of endothelial IKCa channels enhanced smooth muscle depolarization and vasoconstriction [11]. In rat basilar arteries in which NO makes a major contribution to myoendothelial feedback, smooth muscle depolarization to 5-HT was accompanied by IKCa channel-mediated endothelial hyperpolarization. Inhibition of IKCa channels, gap junctional communication, TRPC3 or NOS potentiated smooth muscle depolarization to 5-HT in a non-additive manner indicating that rather being distinct pathways, NO and endothelial IKCa channel activity are part of an integrated mechanism for the regulation of agonist-induced vasoconstriction [91]. In the latter study, Ca2+ signaling was not investigated and the link between IKCa channel activity and NO production was not defined. However, NOS has now been localized close to MEGJs [117] and in co-cultures stimulation of smooth muscle cells with phenylephrine leads to MEGJ specific NOS phosphorylation within endothelial cells to increase NO [118]. Also, in mouse mesenteric vessels, phenylephrine stimulated endothelial TRPV4 sparklets in an InsP3-dependent manner, to engage SKCa and IKCa channels as well as, to a lesser extent, NOS [17]. Thus, given the ability of IKCa channels to modulate endothelial Ca2+ dynamics [12, 113, 114], it may be proposed that activation of IKCa channels at MEGJs following stimulation of smooth muscle cells by GPCR agonists, may amplify dynamic Ca2+ signals to enhance NO production.

Advertisement

3. Local versus conducted responses

The majority of studies described in this chapter have been conducted on isolated resistance arteries which in in vivo would be part of branching network of resistance vessels supplied by feed arteries in which effective control of blood flow requires coordinated behaviour amongst arterial segments [119]. As described above, diffusible mediators such as NO act locally to increase arterial diameter. In contrast, KCa channel-mediated hyperpolarization leads to both local dilation and conduction of the response through the endothelium for distances of several millimeters. This conduction allows for coordination of changes in arterial diameter in multiple vessel segments and so optimizes blood flow [4, 119, 120]. That is not to say that diffusible mediators do not play a role in global blood flow regulation within vascular beds. A recent study of the vascular bed of the mouse gluteus maximus muscle revealed that NO and EDH provide complementary endothelial pathways for ascending vasodilatation to optimize oxygen delivery to the muscle. EDH of downstream arterioles conducts along the endothelium into proximal feed arteries to cause dilation, and NO is released in response to luminal shear stress which increases secondary to downstream dilatation [120].

Advertisement

4. Conclusion

It has become apparent over the past 15 years that endothelial Ca2+ signaling patterns underlie the engagement of effectors such as NOS and/or KCa channels. The physiological significance of these stimulus-specific signaling pathways is not just that they determine the mediator of vasodilation, but also the scope of the impact of each stimulus on blood flow. Stimuli which predominantly elicit release of diffusible mediators will elicit local vasodilation whereas those that initiate EDH have the potential to dilate multiple arterial segments and so affect tissue perfusion. Further work is required to determine if the patterns of Ca2+ signaling described here have widespread applicability, and how they are impacted by age, sex and cardiovascular risk factors. Investigation of how changes in the components of signaling microdomains contribute to the etiology of endothelial dysfunction in conditions such as diabetes and hypertension may lead to the identification of new therapeutic targets.

Advertisement

Acknowledgments

Work in the author’s lab is supported by a grant-in-aid from the Heart and Stroke Foundation of Alberta, Nunavut and the North West Territories to F.P. (G130003076). P.M.K. received a RASCAF award from MacEwan University. Stipend support for graduate students was provided by the University of Alberta (QEII scholarships to S.E.L. and S.L.G., Walter H Johns Graduate Fellowship to S.E.L., 75th Anniversary award to R.W., Faculty of Medicine Recruitment Scholarship to S.L.G.), CIHR (Banting and Best Canada Graduate Scholarship, Masters Award to S.E.L.) and AIHS (Doctoral Studentship to S.E.L.).

Advertisement

Conflict of interest

The authors have declared no conflict of interest.

References

  1. 1. Vanhoutte PM, Shimokawa H, Feletou M, Tang EH. Endothelial dysfunction and vascular disease—A 30th anniversary update. Acta Physiologica (Oxford, England). 2017;219(1):22-96
  2. 2. Bryan RM, You J, Golding EM, Marrelli SP. Endothelium-derived hyperpolarizing factor: A cousin to nitric oxide and prostacyclin. Anesthesiology. 2005;102:1261-1277
  3. 3. Sessa WC. Regulation of endothelial derived nitric oxide in health and disease. Memórias do Instituto Oswaldo Cruz. 2005;100:15-18
  4. 4. Emerson GG, Segal SS. Endothelial cell pathway for conduction of hyperpolarization and vasodilation along hamster feed artery. Circulation Research. 2000;86(1):94-100
  5. 5. Kudo I, Murakami M. Phospholipase A2 enzymes. Prostaglandins & Other Lipid Mediators. 2002;68-69:3-58
  6. 6. Smith WL, Marnett LJ. Prostaglandin endoperoxide synthase: Structure and catalysis. Biochimica et Biophysica Acta. 1991;1083(1):1-17
  7. 7. Sandow SL, Hill CE. Incidence of myoendothelial gap junctions in the proximal and distal mesenteric arteries of the rat is suggestive of a role in endothelium-derived hyperpolarizing factor-mediated responses. Circulation Research. 2000;86:341-346
  8. 8. Laskey RE, Adams DJ, Cannell M, van Breemen C. Calcium entry-dependent oscillations of cytoplasmic calcium concentration in cultured endothelial cell monolayers. Proceedings of the National Academy of Sciences of the United States of America. 1992;89(5):1690-1694
  9. 9. Blatter LA. Tissue specificity: SOCE: Implications for Ca2+ handling in endothelial cells. Advances in Experimental Medicine and Biology. 2017;993:343-361
  10. 10. 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. Proceedings of the National Academy of Sciences of the United States of America. 2008;105(28):9627-9632
  11. 11. Tran CHT, Taylor MS, Plane F, Nagaraja S, Tsoukias NM, Solodushko V, Vigmond EJ, Furstenhaupt T, Brigdan M, Welsh DG. Endothelial Ca2+ wavelets and the induction of myoendothelial feedback. The American Journal of Physiology. 2012;302(8):C1226-C1242
  12. 12. Francis M, Waldrup JR, Qian X, Solodushko V, Meriwether J, Taylor MS. Functional tuning of intrinsic endothelial Ca2+ dynamics in swine coronary arteries. Circulation Research. 2016;118(7):1078-1090
  13. 13. Qian X, Francis M, Köhler R, Solodushko V, Lin M, Taylor MS. Positive feedback regulation of agonist-stimulated endothelial Ca2+ dynamics by KCa3.1 channels in mouse mesenteric arteries. The American Journal of Physiology. 2016;310(9):H1151-H1163
  14. 14. Sonkusare SK, Bonev AD, Ledoux J, Liedtke W, Kotlikoff MI, Heppner TJ, Hill-Eubanks DC, Nelson MT. Elementary Ca2+ signals through endothelial TRPV4 channels regulate vascular function. Science. 2012;336(6081):597-601
  15. 15. Sullivan MN, Earley S. TRP channel Ca2+ sparklets: Fundamental signals underlying endothelium-dependent hyperpolarization. The American Journal of Physiology. 2013;305(10):C999-C1008
  16. 16. Bagher P, Beleznai T, Kansui Y, Mitchell R, Garland CJ, Dora KA. Low intravascular pressure activates endothelial cell TRPV4 channels, local Ca2+ events, and IKCa channels, reducing arteriolar tone. Proceedings of the National Academy of Sciences of the United States of America. 2012;109(44):18174-18179
  17. 17. Marziano C, Hong K, Cope EL, Kotlikoff MI, Isakson BE, Sonkusare SK. Nitric oxide-dependent feedback loop regulates transient receptor potential vanilloid 4 (TRPV4) channel cooperativity and endothelial function in small pulmonary arteries. Journal of the American Heart Association. 2017;6(12): pii: e007157
  18. 18. Sullivan MN, Gonzales AL, Pires PW, Bruhl A, Leo MD, Li W, Oulidi A, Boop FA, Feng Y, Jaggar JH, Welsh DG, Earley S. Localized TRPA1 channel Ca2+ signals stimulated by reactive oxygen species promote cerebral artery dilation. Science Signaling. 2015;8(358):ra2
  19. 19. Baeyens N, Bandyopadhyay C, Coon BG, Yun S, Schwartz MA. Endothelial fluid shear stress sensing in vascular health and disease. The Journal of Clinical Investigation. 2016;126(3):821-828
  20. 20. Chistiakov DA, Orekhov AN, Bobryshev YV. Effects of shear stress on endothelial cells: Go with the flow. Acta Physiologica. 2017;219(2):382-408
  21. 21. Thosar SS, Johnson BD, Johnston JD, Wallace JP. Sitting and endothelial dysfunction: The role of shear stress. Medical Science Monitor. 2012;18(12):RA173-RA180
  22. 22. Zhou J, Li YS, Chien S. Shear stress-initiated signaling and its regulation of endothelial function. Arteriosclerosis, Thrombosis, and Vascular Biology. 2014;34(10):2191-2198
  23. 23. Ghiadoni L, Salvetti M, Muiesan ML, Taddei S. Evaluation of endothelial function by flow mediated dilation: Methodological issues and clinical importance. High Blood Pressure & Cardiovascular Prevention. 2015;22(1):17-22
  24. 24. Yeboah J, Crouse JR, Hsu F-C, Burke GL, Herrington DM. Brachial flow-mediated dilation predicts incident cardiovascular events in older adults: The cardiovascular health study. Circulation. 2007;115(18):2390-2397
  25. 25. Phillips SA, Mahmoud AM, Brown MD, Haus JM. Exercise interventions and peripheral arterial function: Implications for cardio-metabolic disease. Progress in Cardiovascular Diseases. 2015;57(5):521-534
  26. 26. Son Y, Kim K, Jeon S, Kang M, Lee S, Park Y. Effect of exercise intervention on flow-mediated dilation in overweight and obese adults: Meta-analysis. International Journal of Vascular Medicine. 2017;2017:7532702
  27. 27. Green DJ, Dawson EA, Groenewoud HM, Jones H, Thijssen DH. Is flow-mediated dilation nitric oxide mediated?: A meta-analysis. Hypertension. 2014;63(2):376-382
  28. 28. Stoner L, Erickson ML, Young JM, Fryer S, Sabatier MJ, Faulkner J, Lambrick DM, McCully KK. There’s more to flow-mediated dilation than nitric oxide. Journal of Atherosclerosis and Thrombosis. 2012;19(7):589-600
  29. 29. Uematsu M, Kitabatake A, Tanouchi J, Doi Y, Masuyama T, Fujii K, Yoshida Y, Ito H, Ishihara K, Hori M. Reduction of endothelial microfilament bundles in the low-shear region of the canine aorta. Association with intimal plaque formation in hypercholesterolemia. Arteriosclerosis and Thrombosis. 1991;11(1):107-115
  30. 30. Gündüz F, Koçer G, Ulker S, Meiselman HJ, Başkurt OK, Sentürk UK. Exercise training enhances flow-mediated dilation in spontaneously hypertensive rats. Physiological Research. 2011;60(4):589-597
  31. 31. 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. The American Journal of Physiology. 2010;298(2):H466-H476
  32. 32. Brähler S, Kaistha A, Schmidt VJ, Wölfle SE, Busch C, Kaistha BP, Kacik M, Hasenau A-L, Grgic I, Si H, Bond CT, Adelman JP, Wulff H, de Wit C, Hoyer J, Köhler R. Genetic deficit of SK3 and IK1 channels disrupts the endothelium-derived hyperpolarizing factor vasodilator pathway and causes hypertension. Circulation. 2009;119(17):2323-2332
  33. 33. Wei R, Lunn SE, Tam R, Gust SL, Classen B, Kerr PM, Plane F. Vasoconstrictor stimulus determines the functional contribution of myoendothelial feedback to mesenteric arterial tone. The Journal of Physiology. 2018;596(7):1181-1197
  34. 34. Lu T, Wang XL, Chai Q, Sun X, Sieck GC, Katusic ZS, Lee HC. Role of the endothelial caveolae microdomain in shear stress-mediated coronary vasorelaxation. The Journal of Biological Chemistry. 2017;292(46):19013-19023
  35. 35. Chai Q, Wang X-L, Zeldin DC, Lee H-C. Role of caveolae in shear stress-mediated endothelium-dependent dilation in coronary arteries. Cardiovascular Research. 2013;100:151-159
  36. 36. Jalali S, del Pozo MA, Chen K, Miao H, Li Y, Schwartz MA, Shyy JY, Chien S. Integrin-mediated mechanotransduction requires its dynamic interaction with specific extracellular matrix (ECM) ligands. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:1042-1046
  37. 37. Wang Y, Miao H, Li S, Chen KD, Li YS, Yuan S, Shyy JY, Chien S. Interplay between integrins and FLK-1 in shear stress-induced signaling. The American Journal of Physiology. 2002;283:C1540-C1547
  38. 38. Tzima E, Irani-Tehrani M, Kiosses WB, Dejana E, Schultz DA, Engelhardt B, Cao G, DeLisser H, Schwartz MA. A mechanosensory complex that mediates the endothelial cell response to fluid shear stress. Nature. 2005;437:426-431
  39. 39. Yamamoto K, Sokabe T, Matsumoto T, Yoshimura K, Shibata M, Ohura N, Fukuda T, Sato T, Sekine K, Kato S, Isshiki M, Fujita T, Kobayashi M, Kawamura K, Masuda H, Kamiya A, Ando J. Impaired flow-dependent control of vascular tone and remodeling in P2X4-deficient mice. Nature Medicine. 2006;12:133-137
  40. 40. Xu J, Mathur J, Vessières E, Hammack S, Nonomura K, Favre J, Grimaud L, Petrus M, Francisco A, Li J, Lee V, Xiang FL, Mainquist JK, Cahalan SM, Orth AP, Walker JR, Ma S, Lukacs V, Bordone L, Bandell M, Laffitte B, Xu Y, Chien S, Henrion D, Patapoutian A. GPR68 senses flow and is essential for vascular physiology. Cell. 2018;173(3):762-775
  41. 41. Ahn SJ, Fancher IS, Bian JT, Zhang CX, Schwab S, Gaffin R, Phillips SA, Levitan I. Inwardly rectifying K+ channels are major contributors to flow-induced vasodilatation in resistance arteries. The Journal of Physiology. 2017;595(7):2339-2364
  42. 42. Zhang T, Chi S, Jiang F, Zhao Q, Xiao B. A protein interaction mechanism for suppressing the mechanosensitive Piezo channels. Nature Communications. 2017;8(1):1797
  43. 43. Li J, Hou B, Tumova S, Muraki K, Bruns A, Ludlow MJ, Sedo A, Hyman AJ, McKeown L, Young RS, Yuldasheva NY, Majeed Y, Wilson LA, Rode B, Bailey MA, Kim HR, Fu Z, Carter DA, Bilton J, Imrie H, Ajuh P, Dear TN, Cubbon RM, Kearney MT, Prasad RK, Evans PC, Ainscough JF, Beech DJ. Piezo1 integration of vascular architecture with physiological force. Nature. 2014;515(7526):279-282
  44. 44. Gao X, Wu L, O’Neil RG. Temperature-modulated diversity of TRPV4 channel gating: Activation by physical stresses and phorbol ester derivatives through protein kinase C-dependent and -independent pathways. The Journal of Biological Chemistry. 2003;278(29):27129-27137
  45. 45. Moccia F, Berra-Romani R, Tanzi F. Update on vascular endothelial Ca2+ signalling: A tale of ion channels, pumps and transporters. World Journal of Biological Chemistry. 2012;3:127-158
  46. 46. Reitsma S, Slaaf DW, oude Egbrink MG. The endothelial glycocalyx: Composition, functions, and visualization. Pflügers Archiv. 2007;454:345-359
  47. 47. 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. Cardiovascular Research. 2008;80(3):445-452
  48. 48. 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(9):e827
  49. 49. Filosa JA, Yao X, Rath G. TRPV4 and the regulation of vascular tone. Journal of Cardiovascular Pharmacology. 2013;61(2):113-119
  50. 50. 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. The American Journal of Physiology. 2012;302(3):H634-H642
  51. 51. 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. Arteriosclerosis, Thrombosis, and Vascular Biology. 2006;26(7):1495-1502
  52. 52. Addison MP, Singh TU, Parida S, Choudhury S, Kasa JK, Sukumaran SV, Darzi SA, Kandasamy K, Singh V, Kumar D, Mishra SKNO. Synthase inhibition attenuates EDHF-mediated relaxation induced by TRPV4 channel agonist GSK1016790A in the rat pulmonary artery: Role of TxA2. Pharmacological Reports. 2016;68(3):620-626
  53. 53. Gratton JP, Bernatchez P, Sessa WC. Caveolae and caveolins in the cardiovascular system. Circulation Research. 2004;94:1408-1417
  54. 54. Patel HH, Murray F, Insel PA. Caveolae as organizers of pharmacologically relevant signal transduction molecules. Annual Review of Pharmacology and Toxicology. 2008;48:359-391
  55. 55. Sandow SL, Neylon CB, Chen MX, Garland CJ. Spatial separation of endothelial small- and intermediate-conductance calcium-activated potassium channels (KCa) and connexins: Possible relationship to vasodilator function? Journal of Anatomy. 2006;209(5):689-698
  56. 56. Absi M, Burnham MP, Weston AH, Harno E, Rogers M, Edwards G. Effects of methyl β-cyclodextrin on EDHF responses in pig and rat arteries; association between SKCa channels and caveolin-rich domains. British Journal of Pharmacology. 2009;151(3):332-340
  57. 57. Saliez J, Bouzin C, Rath G, Ghisdal P, Desjardins F, Rezzani R, Rodella LF, Vriens J, Nilius B, Feron O, Balligand JL, Dessy C. Role of caveolar compartmentation in endothelium-derived hyperpolarizing factor-mediated relaxation-Ca2+ signals and gap junction function are regulated by caveolin in endothelial cells. Circulation. 2008;117(8):1065-1074
  58. 58. Goedicke-Fritz S, Kaistha A, Kacik M, Markert S, Hofmeister A, Busch C, Bänfer S, Jacob R, Grgic I, Hoyer J. Evidence for functional and dynamic microcompartmentation of Cav-1/TRPV4/KCa in caveolae of endothelial cells. European Journal of Cell Biology. 2015;94(7–9):391-400
  59. 59. Albinsson S, Nordström I, Swärd K, Hellstrand P. Differential dependence of stretch and shear stress signaling on caveolin-1 in the vascular wall. The American Journal of Physiology. 2008;294(1):C271-C279
  60. 60. Yu J, Bergaya S, Murata T, Alp IF, Bauer MP, Lin MI, Drab M, Kurzchalia TV, Stan RV, Sessa WC. Direct evidence for the role of caveolin-1 and caveolae in mechanotransduction and remodeling of blood vessels. The Journal of Clinical Investigation. 2006;116(5):1284-1291
  61. 61. Frangos JA, Eskin SG, McIntire LV, Ives CL. Flow effects on prostacyclin production by cultured human endothelial cells. Science. 1985;227:1477-1479
  62. 62. Hecker M, Mülsch A, Bassenge E, Busse R. Vasoconstriction and increased flow: Two principal mechanisms of shear stress-dependent endothelial autacoid release. The American Journal of Physiology. 1993;265(3 Pt 2):H828-H833
  63. 63. Miura H, Bosnjak JJ, Ning G, Saito T, Miura M, Gutterman DD. Role for hydrogen peroxide in flow-induced dilation of human coronary arterioles. Circulation Research. 2003;92(2):e31-e40
  64. 64. Muller JM, Davis MJ, Kuo L, Chilian WM. Changes in coronary endothelial cell Ca2+ concentration during shear stress- and agonist-induced vasodilation. The American Journal of Physiology. 1999;276(5 Pt 2):H1706-H1714
  65. 65. Parnavelas JG, Kelly W, Burnstock G. Ultrastructural localization of choline acetyltransferase in vascular endothelial cells in rat brain. Nature. 1985;316:724-725
  66. 66. 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. The American Journal of Physiology. 2003;285:H793-H803
  67. 67. Wilson C, Lee MD, McCarron JG. Acetylcholine released by endothelial cells facilitates flow-mediated dilatation. The Journal of Physiology. 2016;594(24):7267-7307
  68. 68. Ellinsworth DC, Sandow SL, Shukla N, Liu Y, Jeremy JY, Gutterman DD. Endothelium-derived hyperpolarization and coronary vasodilation: Diverse and integrated roles of epoxyeicosatrienoic acids, hydrogen peroxide, and gap junctions. Microcirculation. 2016;23(1):15-32
  69. 69. Roos J, DiGregorio PJ, Yeromin AV, Ohlsen K, Lioudyno M, Zhang S, Safrina O, Kozak JA, Wagner SL, Cahalan MD, Velicelebi G, Stauderman KA. STIM1, an essential and conserved component of store-operated Ca2+ channel function. The Journal of Cell Biology. 2005;169(3):435-445
  70. 70. Liou J, Kim ML, Heo WD, Jones JT, Myers JW, Ferrell JE Jr, Meyer T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Current Biology. 2005;15(13):1235-1241
  71. 71. Prakriya M, Feske S, Gwack Y, Srikanth S, Rao A, Hogan PG. Orai1 is an essential pore subunit of the CRAC channel. Nature. 2006;443(7108):230-233
  72. 72. Feske S, Gwack Y, Prakriya M, Srikanth S, Puppel SH, Tanasa B, Hogan PG, Lewis RS, Daly M, Rao A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature. 2006;441(7090):179-185
  73. 73. Choi S, Maleth J, Jha A, Lee KP, Kim MS, So I, Ahuja M, Muallem S. The TRPCs-STIM1-Orai interaction. Handbook of Experimental Pharmacology. 2014;223:1035-1054
  74. 74. Freichel M, Suh SH, Pfeifer A, Schweig U, Trost C, Weissgerber P, Biel M, Philipp S, Freise D, Droogmans G, Hofmann F, Flockerzi V, Nilius B. Lack of an endothelial store-operated Ca2+ current impairs agonist-dependent vasorelaxation in TRP4−/− mice. Nature Cell Biology. 2001;3(2):121-127
  75. 75. 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(3):532-538
  76. 76. Earley S, Pauyo T, Drapp R, Tavares MJ, Liedtke W, Brayden JE. TRPV4-dependent dilation of peripheral resistance arteries influences arterial pressure. The American Journal of Physiology. 2009;297(3):H1096-H1102
  77. 77. Fleming I, Rueben A, Popp R, Fisslthaler B, Schrodt S, Sander A, Haendeler J, Falck JR, Morisseau C, Hammock BD, Busse R. Epoxyeicosatrienoic acids regulate Trp channel dependent Ca2+ signaling and hyperpolarization in endothelial cells. Arteriosclerosis, Thrombosis, and Vascular Biology. 2007;27:2612-2618
  78. 78. Hofmann T, Obukhov AG, Schaefer M, Harteneck C, Gudermann T, Schultz G. Direct activation of human TRPC6 and TRPC3 channels by diacylglycerol. Nature. 1999;397(6716):259-263
  79. 79. 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. Cardiovascular Research. 2001;51(1):160-168
  80. 80. Jacob R, Merritt JE, Hallam TJ, Rink TJ. Repetitive spikes in cytoplasmic calcium evoked by histamine in human endothelial cells. Nature. 1988;335(6185):40-45
  81. 81. 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(2):135-146
  82. 82. Marrelli SP. Mechanisms of endothelial P2Y(1)- and P2Y(2)-mediated vasodilatation involve differential [Ca2+]i responses. The American Journal of Physiology. 2001;281(4):H1759-H1766
  83. 83. Dedkova EN, Blatter LA. Nitric oxide inhibits capacitative Ca2+ entry and enhances endoplasmic reticulum Ca2+ uptake in bovine vascular endothelial cells. The Journal of Physiology. 2002;539(Pt 1):77-91
  84. 84. Fukao M, Hattori Y, Kanno M, Sakuma I, Kitabatake A. Sources of Ca2+ in relation to generation of acetylcholine-induced endothelium-dependent hyperpolarization in rat mesenteric artery. British Journal of Pharmacology. 1997;120(7):1328-1334
  85. 85. Nishimoto M, Mizuno R, Fujita T, Isshiki M. Stromal interaction molecule 1 modulates blood pressure via NO production in vascular endothelial cells. Hypertension Research. 2018 Apr 25 [Epub ahead of print]
  86. 86. Doughty JM, Plane F, Langton PD. Charybdotoxin and apamin block EDHF in rat mesenteric artery if selectively applied to the endothelium. The American Journal of Physiology. 1999;276(3 Pt 2):H1107-H1112
  87. 87. Pannirselvam M, Ding H, Anderson TJ, Triggle CR. Pharmacological characteristics of endothelium-derived hyperpolarizing factor-mediated relaxation of small mesenteric arteries from db/db mice. European Journal of Pharmacology. 2006;551(1–3):98-107
  88. 88. Yamanaka A, Ishikawa T, Goto K. Characterization of endothelium-dependent relaxation independent of NO and prostaglandins in Guinea pig coronary artery. The Journal of Pharmacology and Experimental Therapeutics. 1998;285(2):480-489
  89. 89. Marrelli SP, Eckmann MS, Hunte MS. Role of endothelial intermediate conductance KCa channels in cerebral EDHF-mediated dilations. The American Journal of Physiology. 2003;285(4):H1590-H1599
  90. 90. Chadha PS, Liu L, Rikard-Bell M, Senadheera S, Howitt L, Bertrand RL, Grayson TH, Murphy TV, Sandow SL. Endothelium-dependent vasodilation in human mesenteric artery is primarily mediated by myoendothelial gap junctions intermediate conductance calcium-activated K+ channel and nitric oxide. The Journal of Pharmacology and Experimental Therapeutics. 2011;336(3):701-708
  91. 91. Sandow SL, Haddock RE, Hill CE, Chadha PS, Kerr PM, Welsh DG, Plane F. What’s where and why at a vascular myoendothelial microdomain signalling complex. Clinical and Experimental Pharmacology & Physiology. 2009;36(1):67-76
  92. 92. Kerr PM, Wei R, Tam R, Sandow SL, Murphy TV, Ondrusova K, Lunn SE, Tran CHT, Welsh DG, Plane F. Activation of endothelial IKCa channels underlies NO-dependent myoendothelial feedback. Vascular Pharmacology. 2015;74:130-138
  93. 93. Dora KA, Gallagher NT, McNeish A, Garland CJ. Modulation of endothelial cell KCa3.1 channels during endothelium-derived hyperpolarizing factor signaling in mesenteric resistance arteries. Circulation Research. 2008;102(10):1247-1255
  94. 94. Senadheera S, Kim Y, Grayson TH, Toemoe S, Kochukov MY, Abramowitz J, Housley GD, Bertrand RL, Chadha PS, Bertrand PP, Murphy TV, Tare M, Birnbaumer L, Marrelli SP, Sandow SL. Transient receptor potential canonical type 3 channels facilitate endothelium-derived hyperpolarization-mediated resistance artery vasodilator activity. Cardiovascular Research. 2012;95(4):439-447
  95. 95. Sonkusare SK, Dalsgaard T, Bonev AD, Hill-Eubanks DC, Kotlikoff MI, Scott JD, Santana LF, Nelson MT. AKAP150-dependent cooperative TRPV4 channel gating is central to endothelium-dependent vasodilation and is disrupted in hypertension. Science Signaling. 2014;7(333):ra66
  96. 96. Earley S, Gonzales AL, Garcia ZI. A dietary agonist of transient receptor potential cation channel V3 elicits endothelium-dependent vasodilation. Molecular Pharmacology. 2010;77(4):612-620
  97. 97. Earley S, Gonzales AL, Crnich R. Endothelium-dependent cerebral artery dilation mediated by TRPA1 and Ca2+-activated K+ channels. Circulation Research. 2009;104(8):987-994
  98. 98. Kochukov MY, Balasubramanian A, Abramowitz J, Birnbaumer L, Marrelli SP. Activation of endothelial transient receptor potential C3 channel is required for small conductance calcium-activated potassium channel activation and sustained endothelial hyperpolarization and vasodilation of cerebral artery. Journal of the American Heart Association. 2014;3(4): pii: e000913
  99. 99. Yeon SI, Kim JY, Yeon DS, Abramowitz J, Birnbaumer L, Muallem S, Lee YH. Transient receptor potential canonical type 3 channels control the vascular contractility of mouse mesenteric arteries. PLoS One. 2014;9(10):e110413
  100. 100. Greenberg HZE, Carlton-Carew SRE, Khan DM, Zargaran AK, Jahan KS, Ho WS, Albert AP. Heteromeric TRPV4/TRPC1 channels mediate calcium-sensing receptor-induced nitric oxide production and vasorelaxation in rabbit mesenteric arteries. Vascular Pharmacology. 2017;96-98:53-62
  101. 101. Stankevicius E, Lopez-Valverde V, Rivera L, Hughes A, Mulvany MJ, Simonsen U. Combination of Ca2+-activated K+ channel blockers inhibits acetylcholine-evoked nitric oxide release in rat superior mesenteric artery. British Journal of Pharmacology. 2006;149(5):560-572
  102. 102. Sheng J, Braun AP. Small- and intermediate-conductance Ca2+-activated K+ channels directly control agonist-evoked nitric oxide synthesis in human vascular endothelial cells. The American Journal of Physiology. 2007;293(1):C458-C467
  103. 103. Stankevicus E, Dalsgaard T, Kroigaard C, Beck L, Boedtkjer E, Misfeldt M, Nielsen G, Schjorring O, Hughes A. Opening of small and intermediate calcium-activated potassium channels induces relaxation mainly mediated by nitric-oxide release in large arteries and endothelium-derived hyperpolarizing factor in small arteries from rat. The Journal of Pharmacology and Experimental Therapeutics. 2011;339(3):842-850
  104. 104. Sheng JZ, Ella S, Davis MJ, Hill MA, Braun AP. Openers of SKCa and IKCa channels enhance agonist-evoked endothelial nitric oxide synthesis and arteriolar vasodilation. The FASEB Journal. 2009;23(4):1138-1145
  105. 105. Oike M, Gericke M, Droogmans G, Nilius B. Calcium entry activated by store depletion in human umbilical vein endothelial cells. Cell Calcium. 1994;16:367-376
  106. 106. Wang X, Van Breeman C. Depolarization-mediated inhibition of Ca2+ entry in endothelial cells. The American Journal of Physiology. 1999;277:H1498-H1504
  107. 107. Adams DJ, Barakeh J, Laskey R, Van Breemen C. Ion channels and regulation of intracellular calcium in vascular endothelial cells. The FASEB Journal. 1989;3(12):2389-2400
  108. 108. Lückhoff A, Busse R. Calcium influx into endothelial cells and formation of endothelium-derived relaxing factor is controlled by the membrane potential. Pflügers Archiv. 1990;416(3):305-311
  109. 109. Allen T, Iftinca M, Cole WC, Plane F. Smooth muscle membrane potential modulates endothelium-dependent relaxation of rat basilar artery via myoendothelial gap junctions. The Journal of Physiology. 2002;545(Pt 3:975-986
  110. 110. Clapham DE. Calcium signaling. Cell. 2007;131(6):1047-1058
  111. 111. McSherry IN, Spitaler MM, Takano H, Dora KA. Endothelial cell Ca2+ increases are independent of membrane potential in pressurized rat mesenteric arteries. Cell Calcium. 2005;38(1):23-33
  112. 112. Behringer EJ, Segal SS. Membrane potential governs calcium influx into microvascular endothelium: Integral role for muscarinic receptor activation. The Journal of Physiology. 2015;593(20):4531-4548
  113. 113. Behringer EJ. Calcium and electrical signaling in arterial endothelial tubes: New insights into cellular physiology and cardiovascular function. Microcirculation. 2017;24(3). DOI: 10.1111/micc.12328
  114. 114. Yap FC, Weber DS, Taylor MS, Townsley MI, Comer BS, Maylie J, Adelman JP, Lin MT. Endothelial SK3 channel-associated Ca2+ microdomains modulate blood pressure. The American Journal of Physiology. 2016;310(9):H1151-H1163
  115. 115. Westcott EB, Segal SS. Perivascular innervation: A multiplicity of roles in vasomotor control and myoendothelial signaling. Microcirculation. 2013;20(3):217-238
  116. 116. Kerr PM, Tam R, Ondrusova K, Mittal R, Narang D, Tran CH, Welsh DG, Plane F. Endothelial feedback and the myoendothelial projection. Microcirculation. 2012;19(5):416-422
  117. 117. Mutchler SM, Straub AC. Compartmentalized nitric oxide signaling in the resistance vasculature. Nitric Oxide. 2015;49:8-15
  118. 118. Biwer LA, Taddeo EP, Kenwood BM, Hoehn KL, Straub AC, Isakson BE. Two functionally distinct pools of eNOS in endothelium are facilitated by myoendothelial junction lipid composition. Biochimica et Biophysica Acta. 2016;1861(7):671-679
  119. 119. Segal SS, Jacobs TL. Role for endothelial cell conduction in ascending vasodilatation and exercise hyperaemia in hamster skeletal muscle. The Journal of Physiology. 2001;536(Pt 3):937-946
  120. 120. Sinkler SY, Segal SS. Rapid versus slow ascending vasodilatation: Intercellular conduction versus flow-mediated signalling with tetanic versus rhythmic muscle contractions. The Journal of Physiology. 2017;595:7149-7165

Written By

Ran Wei, Stephanie E. Lunn, Stephen L. Gust, Paul M. Kerr and Frances Plane

Submitted: 16 February 2018 Reviewed: 03 July 2018 Published: 05 November 2018