2,647
Views
19
CrossRef citations to date
0
Altmetric
Review

Elementary calcium signaling in arterial smooth muscle

, , &
Pages 505-519 | Received 24 Jul 2019, Accepted 23 Oct 2019, Published online: 04 Dec 2019

ABSTRACT

Vascular smooth muscle cells (VSMCs) of small peripheral arteries contribute to blood pressure control by adapting their contractile state. These adaptations depend on the VSMC cytosolic Ca2+ concentration, regulated by complex local elementary Ca2+ signaling pathways. Ca2+ sparks represent local, transient, rapid calcium release events from a cluster of ryanodine receptors (RyRs) in the sarcoplasmic reticulum. In arterial SMCs, Ca2+ sparks activate nearby calcium-dependent potassium channels, cause membrane hyperpolarization and thus decrease the global intracellular [Ca2+] to oppose vasoconstriction. Arterial SMC Cav1.2 L-type channels regulate intracellular calcium stores content, which in turn modulates calcium efflux through RyRs. Cav3.2 T-type channels contribute to a minor extend to Ca2+ spark generation in certain types of arteries. Their localization within cell membrane caveolae is essential. We summarize present data on local elementary calcium signaling (Ca2+ sparks) in arterial SMCs with focus on RyR isoforms, large-conductance calcium-dependent potassium (BKCa) channels, and cell membrane-bound calcium channels (Cav1.2 and Cav3.2), particularly in caveolar microdomains.

Introduction

Hypertension is the major risk factor for cardiovascular disease (stroke, myocardial infarction, and others) – the most common cause of death worldwide. A main contributor to systemic blood pressure regulation is the resistance of small arteries and arterioles in the peripheral circulation. These vessels are capable of adapting their diameter independently in response to pressure and flow-associated shear stress. Intravasal pressure of about 60 mmHg (e.g. [Citation1]) or higher triggers the development of myogenic tone by the contraction of vascular smooth muscle cells (VSMCs) that are wound around the lumen of small vessels [Citation2]. In all muscle cells, this contraction state depends, among other factors, on the cytosolic Ca2+ concentration.

Ca2+ sparks and hyperpolarizing currents in vascular smooth muscle cells

In vascular smooth muscle cells (VSMCs), the global cytosolic Ca2+ concentration is regulated by membrane potential-dependent Ca2+ influx pathways, with a major role of Cav1.2 L-type Ca-channels [Citation3Citation7]. Ca2+ sparks are elementary Ca2+-release events generated by a single Ca2+-release unit (CRU) composed of a cluster of ryanodine receptors (RyRs) in the sarcoplasmic reticulum (SR) [Citation8Citation10]. Ca2+ sparks can be visualized by the use of fluorescent, calcium-sensitive dyes (e.g. Fluo-3/-4/AM) in all types of muscle, including striated and smooth muscle cells (SMCs). The basic properties of Ca2+ sparks in human SMCs are similar to arterial SMCs of non-human species [Citation11] . An example of a Ca2+ spark recorded in a mouse mesenteric artery SMC is shown in .

Figure 1. (A) Isolated mesenteric artery smooth muscle cell (VSMC). (B) Ca2+ fluorescence image of a Fluo-4-AM–loaded mesenteric VSMC. (C) Ca2+ fluorescence image of the same cell as in B during the occurrence of an elementary Ca2+ release event (Ca2+ spark). The Ca2+ fluorescence images were recorded with an inverted microscope (Nikon Eclipse Ti, oil immersion objective x40, numerical aperture 1.3). Details are as described in [Citation6]. Magnification in A (10x ocular, 20x objective, numerical aperture 0.45). Bar, 2 µm.

Figure 1. (A) Isolated mesenteric artery smooth muscle cell (VSMC). (B) Ca2+ fluorescence image of a Fluo-4-AM–loaded mesenteric VSMC. (C) Ca2+ fluorescence image of the same cell as in B during the occurrence of an elementary Ca2+ release event (Ca2+ spark). The Ca2+ fluorescence images were recorded with an inverted microscope (Nikon Eclipse Ti, oil immersion objective x40, numerical aperture 1.3). Details are as described in [Citation6]. Magnification in A (10x ocular, 20x objective, numerical aperture 0.45). Bar, 2 µm.

Ca2+ sparks are involved in both positive- and negative-feedback regulation of global cytosolic [Ca2+] to regulate the contraction of smooth muscle cells. Unlike Ca2+ influx via voltage-dependent Ca2+ channels (VDCCs), Ca2+ release from the SR in the form of Ca2+ sparks paradoxically causes vasodilation in peripheral resistance arteries [Citation8,Citation12]. There are two reasons for this counterintuitive effect of Ca2+ sparks in arterial vascular smooth muscle cells (VSMCs). First, a single spark produces a remarkably high (10–100 μmol/L) local (∼1% of the cell volume) increase in [Ca2+]i [Citation13,Citation14], but increases global [Ca2+]i by less than 2 nmol/L [Citation8,Citation10,Citation15]. For triggering Ca2+ sparks in VSMCs RyRs need to form clusters [Citation16,Citation17]. At least 10 RyRs open for one Ca2+ spark [Citation15,Citation18]. Second, Ca2+ sparks occur in close proximity to the cell membrane, where every Ca2+ spark activates numerous, approximately 15 [Citation15], large-conductance Ca2+-sensitive K+ (BKCa) channels, causing K+ efflux [Citation8,Citation13,Citation15,Citation19Citation22]. The resultant “spontaneous transient outward currents” (STOCs) hyperpolarize VSMCs, thereby decreasing Ca2+ entry through voltage-dependent Ca2+ channels via a deactivation process [Citation15]. The net result of Ca2+ spark–BKCa channel coupling is decreased global [Ca2+]i in VSMCs resulting in vasodilation [Citation8,Citation16,Citation19Citation21,Citation23]. The hallmarks of BKCa channel biophysics and its relationship with auxiliary subunit expression have been recently reviewed [Citation24]. In some smooth muscle, Ca2+ sparks may exert a depolarizing effect through opening Cl channels [Citation15]. In this regard, TMEM16A Cl channels have been proposed to comprise an important pathologic mechanism underlying the vasoconstriction and remodeling in pulmonary arteries [Citation25]. TMEM16A is required for peripheral blood vessel contractility and plays a general role in arteriolar and capillary blood flow [Citation26]. TMEM16A is also a key regulator of coronary blood flow and is implicated in the altered contractility of coronary arteries in hypertension [Citation27] (for review see also [Citation28]).

Excitation-contraction coupling in muscle

In different muscle types, there are unique mechanisms of excitation-contraction coupling to regulate the cytosolic Ca2+ concentration [Citation29] (). In skeletal muscle, a direct physical interaction between the L-type channel Cav1.1 and predominantly type 1 ryanodine receptors (RyR1) forms the basis of excitation contraction (E-C) coupling [Citation29]. In this respect, opening of Cav1.1 channel upon membrane depolarization acts as a voltage sensor with conformational changes of the channel leading to a physical interaction with RyR1, promoting its opening, and triggering the release of Ca2+ from internal stores ()) [Citation29Citation31]. Thus, in skeletal muscle, Ca2+ entry is not required for RyR mediated calcium release [Citation32]. In cardiac muscle, physical coupling between Cav1.2 L-type calcium channels and RyRs is not involved in E-C coupling. Instead, membrane depolarization of cardiac myocytes leads to the opening of the L-type channels with Ca2+ influx stimulating the opening of the RyR (predominantly RyR2) and subsequent Ca2+ release from SR2 [Citation9,Citation33Citation35] ()). The close proximity between SR and plasma membrane (t-tubuli) is the morphological requisite for effective calcium-induced calcium release (CICR) in cardiac cells [Citation29].

Figure 2. Control of calcium sparks by voltage-dependent calcium channels. (A) Ca2+ entry from extracellular space is not required for calcium release in skeletal muscle. Charge movement within L-type channel Cav1.1 activates RyR1 channel via a direct physical interaction. Ca2+ efflux from SR through the opened RyR1 channel activates nearby RyR (RyR1 and RyR3) channels via calcium-induced calcium release (CICR). (B) In cardiomyocytes Cav1.2 mediates influx of extracellular Ca2+ into cytosol. Ca2+ then binds to and activates RyR2 channels via CICR. (C) Triggering of Ca2+ sparks is not controlled by rapid, direct cross-talk between Cav1.2 channels and RyR2 in arterial smooth muscle in contrast to cardiac and skeletal muscle cells. Instead, Cav1.2 channels contribute to global cytosolic [Ca2+], which in turn influences luminal SR calcium and thus calcium sparks. SERCA: sarco/endoplasmic reticulum calcium ATPase, BK: big conductance calcium-activated K+ channel. Modified from [Citation29].

Figure 2. Control of calcium sparks by voltage-dependent calcium channels. (A) Ca2+ entry from extracellular space is not required for calcium release in skeletal muscle. Charge movement within L-type channel Cav1.1 activates RyR1 channel via a direct physical interaction. Ca2+ efflux from SR through the opened RyR1 channel activates nearby RyR (RyR1 and RyR3) channels via calcium-induced calcium release (CICR). (B) In cardiomyocytes Cav1.2 mediates influx of extracellular Ca2+ into cytosol. Ca2+ then binds to and activates RyR2 channels via CICR. (C) Triggering of Ca2+ sparks is not controlled by rapid, direct cross-talk between Cav1.2 channels and RyR2 in arterial smooth muscle in contrast to cardiac and skeletal muscle cells. Instead, Cav1.2 channels contribute to global cytosolic [Ca2+], which in turn influences luminal SR calcium and thus calcium sparks. SERCA: sarco/endoplasmic reticulum calcium ATPase, BK: big conductance calcium-activated K+ channel. Modified from [Citation29].

In arterial smooth muscle, a prerequisite for generation of Ca2+ sparks is a close proximity of RyRs, BKCa channels, and VDCCs, forming a functional unit ()) [Citation15]. Parts of the smooth muscle SR membrane are less than 20 nm distant from the plasma membrane [Citation36]. RyR channels have been localized to both the peripheral and central SR [Citation16,Citation36,Citation37]. VDCC and RyR channels colocalize in smooth muscle [Citation16,Citation38].

The ratio of RyRs/L-type channels in skeletal muscle is about 0.5–0.9 [Citation39]. The RyRs/L-type channel ratio in cardiomyocytes is higher and is 3.7 in rabbits, >10 in ferrets and 4 to 7.3 in rats [Citation39,Citation40]. The amount of RyRs forming individual calcium release units (CRUs) in cardiomyocytes seems to vary (up to >100) with a distribution described by exponential equations [Citation41]. The RyRs/L-type channel ratio in bladder SMCs was determined as 2:1 [Citation42].

Ryanodine receptors

RyRs are major Ca2+-release channels in the SR membrane of myocytes that contribute to the regulation of contractility. They are large proteins (~565 kDa), regulated by a multitude of factors and conditions like pH, phosphorylation, oxidation, inorganic ions like Ca2+ and Mg2+, the name giving ryanodine, caffeine, nucleotides, neomycin, tetracaine, halothane, isoflurane, dantrolene, imperatoxin, arachidonic acid and many others (for review see, e.g. [Citation43Citation45]). Boittin et al. [Citation46] found RyRs in SR sections close to the cell membrane. Fritz et al. [Citation47] confirmed that finding for RyR1 in urinary bladder SMCs. However, they detected RyR2 and RyR3 mainly in the deep SR. Although already expressed at birth, RyRs form clusters during ontogeny of arterial smooth muscle cells. These CRUs are a prerequisite to produce Ca2+ sparks as elementary Ca2+ release events in VSMCs to produce vasodilation in resistance arteries [Citation16].

The ability of drugs to regulate RyRs to generate Ca2+-sparks in VSMCs (portal vein) was demonstrated for caffeine (1 mmol/L) or low doses of ryanodine (1 µmol/L) [Citation48]. VSMCs exhibit all three known RyR isoforms (RyR1, −2, −3). The impact of individual RyR isoforms on elementary Ca2+ signaling and adaptive vascular responses has been clarified in the last years with different approaches including silencing single isoforms by antisense (as) oligonucleotides (e.g.Citation[49]) and an inducible smooth muscle-specific RyR2 knock-out mouse model [Citation6].

Coussin et al. reported that treatment with antisense oligonucleotides against RyR1, the dominant RyR subform in skeletal muscle [Citation50], and treatment with antisense oligonucleotides against RyR2, the dominant RyR subform in heart [Citation50,Citation51], abolished Ca2+ sparks in rat portal vein myocytes [Citation49]. The authors propose the existence of mixed Ca2+ channel CRUs with both receptor subtypes necessary to function. Responses to caffeine were weakened, but not completely abolished, with RyR2 or RyR1 silenced [Citation49]. Essin & Gollasch [Citation29] observed that in vivo suppression of RyR2 expression by antisense oligonucleotides increases the SR Ca2+ content without effects on spontaneous calcium sparks in rat cerebral arteries. Noteworthy, the frequency and amplitude of Ca2+ sparks in cerebral arteries was reported to depend on the size of RyR2 clusters [Citation52], which supports the idea the RyR2 could play a dominant role in Ca2+ spark generation in arterial smooth muscle. Very recent data demonstrated that conditional smooth muscle-specific deficiency of RyR2 in mice leads to complete loss of SR Ca2+ release and contractile reactivity following caffeine application of various arteries [Citation6]. RyR2 has an essential function in systemic blood pressure control and vascular adaptive responses to pressure. In the conditional SM-specific RyR2 knock-out mouse model, the ability of the VSMCs to produce Ca2+ sparks was completely abolished by lack of RyR2 (). Furthermore, the negative feedback control of myogenic tone was demonstrated to rely on functional RyR2 [Citation6]. Functional RyR2 also limited the pressure increase in the chronic phase of hypoxic pulmonary vasoconstriction [Citation6,Citation53]. In contrast, RyR2 do not play a role in the adaptive vascular response to blood flow after a femoral artery ligation [Citation6,Citation54].

Figure 3. Ca2+ sparks in wild-type and smooth muscle myosin heavy chain (SMMHC)–Ryr2−/- tibial artery smooth muscle cells (SMCs). (A) Ca2+ fluorescence image of a Fluo-4-AM–loaded control SMC. (B) Ca2+ fluorescence image of the same cell as in A during the occurrence of a Ca2+ spark. Two-dimensional images were recorded at a rate of 5/s using an inverted microscope (Nikon Eclipse Ti, oil immersion objective x40, numerical aperture 1.3;) as described in [Citation6]. (C) Time course of Ca2+ fluorescence changes in cellular regions of interest (ROIs) without sparks (ROI a, left) and with sparks (ROI b, right). (D) Time course of Ca2+ fluorescence changes in an ROI (similar size as that in A) of a SMMHC-Ryr2−/- SMC. (E) Percentage of SMCs with Ca2+ sparks. (F) Ca2+ spark frequencies in SMCs from control and SMMHC-Ryr2−/- mice. F/F0, fluorescence/background fluorescence. Bar, 2 µm. *P < 0.05. Modified from [Citation6].

Figure 3. Ca2+ sparks in wild-type and smooth muscle myosin heavy chain (SMMHC)–Ryr2−/- tibial artery smooth muscle cells (SMCs). (A) Ca2+ fluorescence image of a Fluo-4-AM–loaded control SMC. (B) Ca2+ fluorescence image of the same cell as in A during the occurrence of a Ca2+ spark. Two-dimensional images were recorded at a rate of 5/s using an inverted microscope (Nikon Eclipse Ti, oil immersion objective x40, numerical aperture 1.3;) as described in [Citation6]. (C) Time course of Ca2+ fluorescence changes in cellular regions of interest (ROIs) without sparks (ROI a, left) and with sparks (ROI b, right). (D) Time course of Ca2+ fluorescence changes in an ROI (similar size as that in A) of a SMMHC-Ryr2−/- SMC. (E) Percentage of SMCs with Ca2+ sparks. (F) Ca2+ spark frequencies in SMCs from control and SMMHC-Ryr2−/- mice. F/F0, fluorescence/background fluorescence. Bar, 2 µm. *P < 0.05. Modified from [Citation6].

RyR3 are widely expressed throughout different organs of the body, including VSMCs and other smooth muscle types (e.g. in digestive tract and uterus) [Citation50,Citation51]. The physiological data concerning Ca2+ signaling are diverse and in part contradictory. Silencing RyR3 with antisense oligonucleotides in rat portal vein myocytes reduced the Ca2+ responses to caffeine or phenylephrine in a solution containing 10 mmol/L Ca2+; thus, RyR3 might be activatable by local high [Ca2+][Citation55]. However, spontaneous Ca2+ sparks were not affected in this study. In mouse cerebral artery VSMCs of Ryr3−/- mice [Citation56], the frequencies of Ca2+ sparks and STOCs were increased [Citation57], indicating a possible influence of RyR3 on Cav1.2/RyR2 coupling in arterial smooth muscle. A possible reason could be a lower sensitivity of RyR3 to Ca2+ dependent inactivation than that of RyR2 [Citation58]. The RyR3-mediated prolonged Ca2+ release would inactivate RyR2 channels [Citation29]. A specific role of RyR3 in calcium spark signaling in non-spiking smooth muscle compared to spiking smooth muscle may be caused by different expression levels [Citation59] and/or splicing variants of RyR3 [Citation60,Citation61]. A dominant negative (DN) variant has been proposed to function as predominant RyR3 variant in mesenteric smooth artery muscle [Citation62]. The authors showed that RyR3 (especially DN-RYR3) suppressed the Ca2+ release by forming complexes with RyR2.

BKCa channels activated by Ca2+ sparks appeared to hyperpolarize and dilate pressurized myogenic arteries because ryanodine and thapsigargin depolarized and constricted these arteries to an extent similar to that produced by blockers of BKCa channels [Citation8,Citation12]. Pressure-induced myogenic tone is involved in autoregulation of local blood flow and confers protection against excessive pressure levels in small peripheral arteries and arterioles. In a number of arteries, the development of myogenic tone relies on ligand-independent mechanoactivation of angiotensin II type 1a receptors [Citation1], a Gq/11-protein coupled receptor type [Citation63]. Interestingly, angiotensin II has been found to stimulate Ca2+ sparks via AT1 receptors and L-type Ca2+ channels [Citation64]. This pathway may cause increased Ca2+ spark–BKCa channel coupling during ligand-dependent vasoconstriction to limit excess pressure levels. However, the molecular mechanism is not entirely clear since activators of protein kinase C have been found to decrease Ca2+ spark frequency in smooth muscle cells from cerebral arteries [Citation65]. In addition to RyRs, inositol 1,4,5-trisphosphate (IP3) receptors may have a function in regulating the cytosolic Ca2+ concentration, since unitary Ca2+ release may also occur due to inositol trisphosphate receptor (IP3R) activation in the SR at the same intracellular Ca2+ reservoir as ryanodine receptors [Citation46,Citation66Citation68]. IP3 dynamically regulates assembly of IP3 receptors into clusters that underlie the hierarchical recruitment of elementary IP3 receptor-mediated Ca2+ release events (Ca2+ puffs) [Citation69]. Interestingly, opening of both RyR1 and IP3 receptors can be allosterically coupled to a displacement of the N-terminal domain from the following two domains. This displacement can be affected by glutathionylation of a highly reactive cysteine residue, or ligand binding [Citation70]. In human coronary artery VSMCs, Ca2+ entry into the cell through reverse-mode Na+/Ca2+ exchange determines calcium store refilling, which in turn regulates STOC generation [Citation71,Citation72].

RyRs change their physiological function during the aging process. Although there are no studies on the effects of age on Ca2+ sparks in arterial smooth muscle, age-dependent alterations in RyR function influence the node cycle length and atrio-His interval in hearts of rats [Citation73]. RyRs may modulate gap junction formation during stress and aging. The pan RyR activator caffeine decreased the expression of gap junction forming proteins, the connexins Cx40, Cx43, and Cx45, and reduced the contractility of cardiomyocytes [Citation74]. These effects have been attributed to long-lasting activated ryanodine receptors by caffeine. With aging, expression of Cx40, SERCA type 2a, and the L-type Ca2+ channel Cav1.3 increases while the expression of Cx43 and RyR2 decreases in certain parts of the heart, such as the inferior nodal extension and proximal penetrating bundle [Citation73]. Gap junction impairment could also disturb cell-cell communications between VSMCs necessary for vascular adaptation to pressure and local blood flow control [Citation75].

Ca2+ channels in vascular smooth muscle cells

Voltage-dependent Ca2+ channels in the plasma membrane of VSMCs are regulated by membrane potential. They dominate the extracellular Ca2+ influx into VSMCs in response to transmural pressure to elevate the global cytosolic [Ca2+] to cause myogenic vasoconstriction [Citation10,Citation76]. Ca2+ influx through Ca2+ channels contributes to global cytosolic [Ca2+], which in turn also influences luminal SR calcium and triggers Ca2+ sparks to activate BKCa channels. The resultant BKCa channel mediated plasma membrane hyperpolarization, which opposes myogenic vasoconstriction, i.e. Ca2+ sparks cause paradoxically arterial vasodilatation [Citation8,Citation13Citation15,Citation29]. In addition to VDCCs, transient receptor potential (TRP) channels and ionotropic purinergic (P2X) receptors, which are nonselective cation channels, also contribute to extracellular Ca2+ influx [Citation77Citation79].

VDCCs are multi-subunit complexes comprising a pore-forming α1 subunit and regulatory β, α2δ, and γ subunits [Citation80]. Electrophysiological studies performed with vertebrates VDCC subunits have demonstrated that the α1 subunit largely determines both the pharmacological and biophysical properties of Ca2+ currents [Citation7,Citation81]. These properties have defined different current subtypes, according to their Low or High Voltage for Activation (LVA or HVA), their Long-lasting or Transient kinetics of inactivation (L or T-type), and their distinct sensitivities to Dihydropyridines (DHP) and different toxins. VDCCs can be grouped in three major families based on their α1 subunit amino acid sequence similarity: Cav1, Cav2, and Cav3. VDCCs have been differentiated into T-(transient), N-(originally termed as neither T nor L [Citation82]) and L-(long lasting) channels. N-channels are apparently present only in neuronal cells whereas T-type and L-type channels have been identified in most cells, including vascular smooth muscle cells (see ).

Table 1. Voltage-dependent Ca2+ channels (VDCCs) expressed in vascular smooth muscle.

L-type Ca2+ channels in vasculature

Voltage-dependent L-type Ca2+ channels play a critical role in excitation-contraction coupling through calcium signaling in cardiac, skeletal, and smooth muscle. They have a large single-channel conductance compared to T-type channels, are sensitive to dihydropyridine blockers, and encoded by the Cav1 family of α1 subunits[Citation85]. Four subtypes of L-type Ca2+ channels (Cav1.1–1.4) have been identified in cardiac, smooth muscle and neurons. They are of particular functional importance and essential for cardiac contraction and for neuronal function [Citation90,Citation91]. L-type Cav1.2 channels provide an important pathway for influx of extracellular Ca2+ and their graded activation plays a central role in setting cytosolic [Ca2+] in smooth muscle contraction [Citation15,Citation92].

Cav1.2 channels have been explored in vascular smooth muscle by pharmacological or genetic tools, and they exhibit a critical role in the regulation of Ca2+ sparks, STOCs, and myogenic tone in arteries [Citation3,Citation5,Citation7]. Deletion of the smooth muscle-specific Cav1.2 channel gene in (SMAKO) mice decreased Ca2+ spark frequency by ~70-80%, and the STOCs frequency by ~50% [Citation3,Citation5]. Resistance arteries from SMAKO mice show very weak myogenic constriction in response to increased intravascular pressure [Citation7,Citation93].

Cav1.3, encoded by the CACNA1D gene, is expressed together with Cav1.2 in many tissues, including neurons, endocrine cells, as well as smooth muscle [Citation94]. Although Cav1.3 can mediate Ca2+ sparklets in vascular smooth muscle, Cav1.2 channels underlie predominant Ca2+ currents, which were similar to those recorded from cells lacking Cav1.3 channels [Citation95].

T-type Ca2+ channels in vasculature

Cav3.1 and Cav3.2 T-type channels have been found in rat and human renal microvessels (preglomerular vessels, juxtamedullary efferent arterioles, and vasa recta) to contribute to control of blood pressure, glomerular filtration rate, and salt and water homeostasis [Citation86,Citation96,Citation97]. Cav3.1 and Cav3.2 channels have been identified by RT-PCR, Western blot, immunolocalization or functional studies in cerebral arteries [Citation84,Citation98,Citation99], aorta and mesenteric arteries of rats [Citation100,Citation101], in human subcutaneous vessels [Citation101] and in mouse cremaster arterioles [Citation102]. T-type channels could contribute to alterations of rat cerebral network perfusion (20-50% of the total response) as biophysical modeling revealed based on data obtained in arterial diameter measurements [Citation84].

Role of VDCC-RyR signaling on Ca2+ spark generation in vascular smooth muscle

Similar to the heart, L-type Cav1.2 Ca2+ channels are presumably the predominant pathway by which extracellular Ca2+ triggers Ca2+ sparks in arterial smooth muscle. In cardiac muscle cells, tight coupling between Cav1.2/RyR2 seems to underlie CICR to generate Ca2+ sparks. Wang et al. [Citation103] determined the coupling fidelity of L-type Ca2+ channels and RyR in rat ventricular myocytes with 0.71, i.e. 71% opening events of L-type channels trigger Ca2+ sparks. In contrast, RyRs activation is not tightly linked to the opening of L-type Ca2+ channels to trigger Ca2+ sparks in smooth muscle [Citation3,Citation104]. The analysis of latencies of calcium sparks appearing after L-type Cav1.2 channel openings at different membrane potentials [Citation3,Citation104] revealed a rather loose coupling. Nevertheless, Cav1.2 channels could be shown to contribute to global cytosolic [Ca2+], which in turn influences luminal SR calcium and thus Ca2+ sparks in arterial smooth muscle [Citation3,Citation105]. The loose coupling between L-type Cav1.2 channel and RyRs is presumably due to an increase in the effective distance between both channels, resulting in an uncoupling of the obligate relationship that exists in striated muscle between the action potential and calcium release [Citation104].

The close proximity between Cav channels and RyRs for E-C coupling in striated muscle is ensured by placing the L-type Cav1 channels inside t-tubules precisely aligned to parts of the SR in skeletal muscle [Citation106-Citation108] and cardiomyocytes [Citation90,Citation109,Citation110]. Also in VSMCs a precise delivery of the Ca2+ ions supplied by VDCCs with short diffusion ways to the SR is essential [Citation111] to trigger Ca2+ sparks for the indirect regulation of the VSMC contraction state. However, smooth muscle cells do not have t-tubuli. Instead, they possess abundant cell membrane caveolae [Citation112]. Löhn et al. were the first to propose that caveolae could represent additional structural elements necessary for the generation of Ca2+ sparks in cardiomyocytes and arterial smooth muscle [Citation111]. This concept is supported by findings of Boittin et al. [Citation46], who found RyRs in vicinity of caveolae. Impairments of the caveolae structure, e.g. via cholesterol depletion with methyl-beta-cyclodextrin or by genetically knocking out essential proteins like EHD2 or caveolin 1, reduce the frequency of Ca2+ sparks [Citation4,Citation5] (Fan et al. 2019 (submitted)). We recently also found that BKCa channels are not the downstream mediators of perivascular adipose tissue-dependent arterial vasorelaxation [Citation113Citation115].

Caveolae microdomains and Ca2+ sparks

L-type Cav1.2 is typically thought to play a predominant role in the signaling pathway of VDCCs-RyRs-BKCa in vascular smooth muscle [Citation3]. Nevertheless, pharmacological or genetic Cav1.2 channel ablation does not fully eliminate Ca2+ spark events and STOCs, RyR2-mediated Ca2+ spark-dependent regulation of arterial myogenic tone [Citation3,Citation98,Citation103]. L-type Cav1.2 channels are not the only voltage-dependent calcium channels expressed in vascular tissue (see ). T-type Cav3.2 channels reside inside caveolae [Citation111,Citation112]. In mouse mesenteric arteries, but not in tibial or cerebral arteries, they represent additional Ca2+ influx channels to trigger Ca2+ sparks [Citation4,Citation5,Citation112] ().

Figure 4. Proposed model of the role of Cav1.2 and Cav3.2 channels in arterial smooth muscle Ca2+ spark generation. Ca2+ sparks are produced by opening of clustered ryanodine receptors (RyR2) in the SR, which produces a negative-feedback effect on vasoconstriction. This vasodilatory effect is mediated by activation of large-conductance Ca2+-activated K+ (BKCa) channels, which results in hyperpolarization of VSMCs and reduced global cytosolic [Ca2+]. The majority (~70-80%) of Ca2+ sparks are triggered by Cav1.2 channels contributing to global cytosolic [Ca2+], which in turn influences luminal SR calcium via SERCA [Citation3,Citation104]. Cav3.2 T-type channels contribute to a minor extent [Citation5]. SERCA, calcium pump; SR, sarcoplasmic reticulum; VSMC, mesenteric artery vascular smooth muscle cell.

Figure 4. Proposed model of the role of Cav1.2 and Cav3.2 channels in arterial smooth muscle Ca2+ spark generation. Ca2+ sparks are produced by opening of clustered ryanodine receptors (RyR2) in the SR, which produces a negative-feedback effect on vasoconstriction. This vasodilatory effect is mediated by activation of large-conductance Ca2+-activated K+ (BKCa) channels, which results in hyperpolarization of VSMCs and reduced global cytosolic [Ca2+]. The majority (~70-80%) of Ca2+ sparks are triggered by Cav1.2 channels contributing to global cytosolic [Ca2+], which in turn influences luminal SR calcium via SERCA [Citation3,Citation104]. Cav3.2 T-type channels contribute to a minor extent [Citation5]. SERCA, calcium pump; SR, sarcoplasmic reticulum; VSMC, mesenteric artery vascular smooth muscle cell.

T-type calcium channels were originally described as low-voltage-activated (LVA) channels because they can be activated at more negative membrane potentials compared to high-voltage-activated channels, such as Cav1.2. They are also characterized by their tiny and transient inward Ca2+ currents (called T-type currents) resulting from small depolarization [Citation116,Citation117]. No specific antagonists are available to separate each subtype of T-type channel. Mibefradil, the most commonly used T-type channel blocker, has not only equal function on each of the three T-type channel subtypes [Citation118], but also has been shown to block other ion channels [Citation119Citation121]. Low concentrations of nickel (with an IC50 value of about 13 µmol/L) have been demonstrated successfully in Cav3.2 channel blocking, whereas ~20-fold of this IC50 concentration was necessary to inhibit Cav3.1 channels or L-type channels, with IC50 values of 250 µmol/L and 324 µmol/L, respectively [Citation122,Citation123]. Experiments using Cav3.1 knockout mice indicate that Cav3.1 channels contribute to the sinoatrial node pacemaker activity and atrioventricular conduction but do not control systemic blood pressure [Citation124]. Also, Cav3.1 channels appear to be necessary for proper progression of the cell cycle in human pulmonary artery smooth muscle cells [Citation125]. However, there are no reports on vascular abnormalities in Cav3.1 knock-out animals.

Research based on the pharmacological (Ni2+) or/and genetic tools clearly ascertained the contribution of Cav3.2 in VSMCs. Findings reveal that Ca2+ influx through Cav3.2 channel triggers Ca2+ spark generation and then activates BKCa channels to hyperpolarize and relax certain artery types, such as cerebral and mesenteric arteries [Citation112,Citation126].

Although knockout mice provide an important and reliable tool for investigation of the roles of Cav3.1 and Cav3.2 channels, the potential of compensatory and masking effects (or interference) caused by L-type channels remains a general challenge in specific T-type deletion studies. In this state-of-affairs, smooth muscle-specific Cav1.2-/- (SMAKO) mice [Citation3] have been successfully used to explore the role of T-type Ca2+ channels in vascular function. Intriguingly, the Cav3.2 blocker nickel fully abolished Ca2+ sparks after silencing Cav1.2 channels in mouse mesenteric smooth muscle cells. However, nickel failed to inhibit Ca2+ sparks in mouse tibial and cerebral arteries. Thus, T-type channels provide sufficient Ca2+ influx to trigger Ca2+ sparks (responsible for about 20%~30%) at the resting membrane potential in certain artery types, when the interference from L-type Cav1.2 channels is excluded () [Citation5]. T-type Cav3 channel activity on vascular tone regulation is most prominent at voltages close to the VSMC resting potential [Citation112] and at low intravasal pressures (40–80 mmHg) [Citation127,Citation128].

Based on these data, the question was raised whether direct vs. indirect Cav3.2 channel-RyR2 coupling underlies Ca2+ spark generation, for example in specialized microdomains, such as caveolae. Recent data show that there is close apposition of Cav3.2 channels in caveolae to RyRs, which is crucial for T-type Cav3.2 to generate Ca2+ sparks. The effective distance between Cav3.2 and RyRs in a restricted microdomain is ~15 nm between caveolae in the plasma membrane and the sarcoplasmic reticulum [Citation112]. Unfolding the pits structure of caveolae, pharmacologically or genetically, blocked the Ni2+-sensitive Ca2+ sparks and Cav3.2/BKCa-mediated feedback in vascular smooth muscle [Citation4,Citation5,Citation112].

Interestingly, a novel junctophilin–caveolin interaction has been identified in mesenteric arteries, which enables efficient coupling between RyRs and BKCa channels in the VSMC Ca2+ microdomain. The data indicate that junctophilin-2, interacting with the caveolae-forming protein caveolin-1, generates bridges between the plasma membrane and the SR and thereby contributes to the formation of microdomains essential for Ca2+ sparks. Downregulation of junctophilin-2 diminished the amplitude of STOCs. The frequency of Ca2+ sparks was increased but not the STOCs frequency [Citation129]. The authors conclude that junctophilin-2/Cav1 interaction provides a structural and functional basis for the Ca2+ microdomain at plasma membrane-SR junctions and mediates cross-talk between RyRs and BKCa channels, converts local Ca2+ sparks into membrane hyperpolarization, and contributes to stabilizing resting tone in VSMCs.

Perspectives

In the past decade, a number of new molecular mechanisms have been identified which control Ca2+ spark–BKCa channel coupling and relaxation in arterial smooth muscle. As such, very recent data identified major roles of RyR2, Cav3.2 T-type Ca2+ channels and caveolae in these processes. These molecular components and associated pathways may represent novel promising molecular targets for the development of new drugs to treat high blood pressure and regional blood flow. They could be potentially important for understanding BKCa channel-mediated pathways in human baroreflex control [Citation130]. Further work is necessary to study the changes these molecules undergo during aging. Recent evidence also supports the reverse role for BKCa channels, in which they facilitate Ca2+ influx into the cell interior through open nonselective cation (e.g., transient receptor potential; TRP) channels in accord with robust electrical (hyperpolarization) and concentration (~20,000-fold) transmembrane gradients for Ca2+ [Citation131]. Such an arrangement supports a feed-forward activation of membrane hyperpolarization while potentially affecting Ca2+ spark–BKCa channel coupling. It is tempting to speculate that TRP channels might represent promising new targets in chronic, age-related cardiovascular disease to affect relaxation of arterial smooth muscle by Ca2+ sparks. Further work is also required to ascertain the precise relationship between Cav3.2 mediated Ca2+ influx and Ca2+ sparks, and the contribution of Cav3.2 channels on luminal calcium content. We proposed experimental study designs for distinguishing between biological and technical replicates in such studies using isolated arteries [Citation132]. Future work is also required to study the role of these pathways in aging and cardiovascular disease states. In light of the global obesity epidemic, it will be important to study the impact of adipokins, such as ADRF, released from perivascular adipose tissue [Citation133] on arterial Ca2+ spark–BKCa channel coupling and relaxation and their role in the treatment and prevention of chronic, age-related cardiovascular and metabolic diseases.

Author contributions

MK drafted the article, and all authors, contributed to its completion.

Acknowledgments

We would like to express our thanks to the Deutsche Forschungsgemeinschaft (DFG) for continuous support, which enabled us to pursue this exciting research direction. Furthermore, this study was supported by a grant from the Deutsche Akademische Austauschdienst (DAAD).

Disclosure statement

No potential conflict of interest was reported by the authors.

References

  • Schleifenbaum J, Kassmann M, Szijarto IA, et al. Stretch-activation of angiotensin II type 1a receptors contributes to the myogenic response of mouse mesenteric and renal arteries. Circ Res. 2014;115:263–272.
  • Bayliss WM. On the local reactions of the arterial wall to changes of internal pressure. J Physiol. 1902;28:220–231.
  • Essin K, Welling A, Hofmann F, et al. Indirect coupling between Cav1.2 channels and ryanodine receptors to generate Ca2+ sparks in murine arterial smooth muscle cells. J Physiol. 2007;584:205–219.
  • Hashad AM, Harraz OF, Brett SE, et al. Caveolae link CaV3. 2 channels to BKCa-mediated feedback in vascular smooth muscle. Arterioscl Throm Vas Biol. 2018;38:2371–2381.
  • Fan G, Kaßmann M, Hashad AM, et al. Differential targeting and signalling of voltage‐gated T‐type Cav3. 2 and L‐type Cav1. 2 channels to ryanodine receptors in mesenteric arteries. J Physiol. 2018;596:4863–4877.
  • Kassmann M, Szijarto IA, García‐Prieto CF, et al. Role of ryanodine type 2 receptors in elementary Ca2+ signaling in arteries and vascular adaptive responses. J Am Heart Assoc. 2019;8:e010090.
  • Moosmang S, Schulla V, Welling A, et al. Dominant role of smooth muscle L-type calcium channel Cav1.2 for blood pressure regulation. Embo J. 2003;22:6027–6034.
  • Nelson MT, Cheng H, Rubart M, et al. Relaxation of arterial smooth muscle by calcium sparks. Science. 1995;270:633–637.
  • Wang S-Q, Wei C, Zhao G, et al. Imaging microdomain Ca2+ in muscle cells. Circ Res. 2019;6:1011–1022.
  • Jaggar JH, Wellman GC, Heppner TJ, et al. Ca2+ channels, ryanodine receptors and Ca(2+)-activated K+ channels: a functional unit for regulating arterial tone. Acta Physiol Scand. 1998;164:577–587.
  • Furstenau M, Lohn M, Ried C, et al. Calcium sparks in human coronary artery smooth muscle cells resolved by confocal imaging. J Hypertens. 2000;18:1215–1222.
  • Knot HJ, Standen NB, Nelson MT. Ryanodine receptors regulate arterial diameter and wall [Ca2+] in cerebral arteries of rat via Ca2+-dependent K+ channels. J Physiol. 1998;508(Pt 1):211–221.
  • Perez GJ, Bonev AD, Patlak JB, et al. Functional coupling of ryanodine receptors to KCa channels in smooth muscle cells from rat cerebral arteries. J Gen Physiol. 1999;113:229–238.
  • Perez GJ, Bonev AD, Nelson MT. Micromolar Ca(2+) from sparks activates Ca(2+)-sensitive K(+) channels in rat cerebral artery smooth muscle. Am J Physiol Cell Physiol. 2001;281:C1769–75.
  • Jaggar JH, Porter VA, Lederer WJ, et al. Calcium sparks in smooth muscle. Am J Physiol Cell Physiol. 2000;278:C235–56.
  • Gollasch M, Wellman GC, Knot HJ, et al. Ontogeny of local sarcoplasmic reticulum Ca2+ signals in cerebral arteries: Ca2+ sparks as elementary physiological events. Circ Res. 1998;83:1104–1114.
  • Cannell MB, Soeller C. Numerical analysis of ryanodine receptor activation by L-type channel activity in the cardiac muscle diad. Biophys J. 1997;73:112–122.
  • Mejia-Alvarez R, Kettlun C, Rios E, et al. Unitary Ca2+ current through cardiac ryanodine receptor channels under quasi-physiological ionic conditions. J Gen Physiol. 1999;113:177–186.
  • Brenner R, Perez GJ, Bonev AD, et al. Vasoregulation by the beta1 subunit of the calcium-activated potassium channel. Nature. 2000;407:870–876.
  • Pluger S, Faulhaber J, Furstenau M, et al. Mice with disrupted BK channel beta1 subunit gene feature abnormal Ca(2+) spark/STOC coupling and elevated blood pressure. Circ Res. 2000;87:E53–60.
  • Sausbier M, Arntz C, Bucurenciu I, et al. Elevated blood pressure linked to primary hyperaldosteronism and impaired vasodilation in BK channel-deficient mice. Circulation. 2005;112:60–68.
  • Benham CD, Bolton TB. Spontaneous transient outward currents in single visceral and vascular smooth muscle cells of the rabbit. J Physiol. 1986;381:385–406.
  • Filosa JA, Bonev AD, Straub SV, et al. Local potassium signaling couples neuronal activity to vasodilation in the brain. Nat Neurosci. 2006;9:1397–1403.
  • Contreras GF, Castillo K, Enrique N, et al. A BK (Slo1) channel journey from molecule to physiology. Channels (Austin). 2013;7:442–458.
  • Papp R, Nagaraj C, Zabini D, et al. Targeting TMEM16A to reverse vasoconstriction and remodelling in idiopathic pulmonary arterial hypertension. Eur Respir J. 2019;53: 1800965.
  • Heinze C, Seniuk A, Sokolov MV, et al. Disruption of vascular Ca2+-activated chloride currents lowers blood pressure. J Clin Invest. 2014;124:675–686.
  • Askew Page HR, Dalsgaard T, Baldwin SN, et al. TMEM16A is implicated in the regulation of coronary flow and is altered in hypertension. Br J Pharmacol. 2019;176:1635–1648.
  • Dam VS, Boedtkjer DM, Aalkjaer C, et al. The bestrophin- and TMEM16A-associated Ca(2+)- activated Cl(-) channels in vascular smooth muscles. Channels (Austin). 2014;8:361–369.
  • Essin K, Gollasch M. Role of ryanodine receptor subtypes in initiation and formation of calcium sparks in arterial smooth muscle: comparison with striated muscle. J Biomed Biotechnol. 2009;2009:135249.
  • Rios E, Brum G. Involvement of dihydropyridine receptors in excitation-contraction coupling in skeletal muscle. Nature. 1987;325:717–720.
  • Klein MG, Schneider MF. Ca2+ sparks in skeletal muscle. Prog Biophys Mol Biol. 2006;92:308–332.
  • Armstrong CM, Bezanilla FM, Horowicz P. Twitches in the presence of ethylene glycol bis(-aminoethyl ether)-N,N’-tetracetic acid. Biochim Biophys Acta. 1972;267:605–608.
  • Cheng H, Lederer WJ. Calcium sparks. Physiol Rev. 2008;88:1491–1545.
  • Fabiato A. Calcium-induced release of calcium from the cardiac sarcoplasmic reticulum. A J Physiol. 1983;245:C1–14.
  • Domeier TL, Blatter LA, Zima AV. Changes in intra-luminal calcium during spontaneous calcium waves following sensitization of ryanodine receptor channels. Channels (Austin). 2010;4:87–92.
  • Devine CE, Somlyo AV, Somlyo AP. Sarcoplasmic reticulum and excitation-contraction coupling in mammalian smooth muscles. J Cell Biol. 1972;52:690–718.
  • Lesh RE, Nixon GF, Fleischer S, et al. Localization of ryanodine receptors in smooth muscle. Circ Res. 1998;82:175–185.
  • Carrington WA, Lynch RM, Moore ED, et al. Superresolution three-dimensional images of fluorescence in cells with minimal light exposure. Science. 1995;268:1483–1487.
  • Bers DM, Stiffel VM. Ratio of ryanodine to dihydropyridine receptors in cardiac and skeletal muscle and implications for E-C coupling. A J Physiol. 1993;264:C1587–93.
  • Cannell MB, Cheng H, Lederer WJ. Spatial non-uniformities in [Ca2+]i during excitation-contraction coupling in cardiac myocytes. Biophys J. 1994;67:1942–1956.
  • Soeller C, Baddeley D. Super-resolution imaging of EC coupling protein distribution in the heart. J Mol Cell Cardiol. 2013;58:32–40.
  • Moore ED, Voigt T, Kobayashi YM, et al. Organization of Ca2+ release units in excitable smooth muscle of the guinea-pig urinary bladder. Biophys J. 2004;87:1836–1847.
  • Coronado R, Morrissette J, Sukhareva M, et al. Structure and function of ryanodine receptors. Am J Physiol Cell Physiol. 1994;266:C1485–504.
  • Franzini-Armstrong C, Protasi F. Ryanodine receptors of striated muscles: a complex channel capable of multiple interactions. Physiol Rev. 1997;77:699–729.
  • Mohler PJ, Wehrens XHT. Mechanisms of human arrhythmia syndromes: abnormal cardiac macromolecular interactions. Physiology. 2007;22:342–350.
  • Boittin F-X, Macrez N, Halet G, et al. Norepinephrine-induced Ca2+ waves depend on InsP3 and ryanodine receptor activation in vascular myocytes. Am J Physiol Cell Physiol. 1999;277:C139–51.
  • Fritz N, Morel J-L, Jeyakumar LH, et al. RyR1-specific requirement for depolarization-induced Ca2+ sparks in urinary bladder smooth muscle. J Cell Sci. 2007;120:3784–3791.
  • Mironneau J, Arnaudeau S, Macrez-Lepretre N, et al. Ca2+ sparks and Ca2+ waves activate different Ca2+-dependent ion channels in single myocytes from rat portal vein. Cell Calcium. 1996;20:153–160.
  • Coussin F, Macrez N, Morel J-L, et al. Requirement of ryanodine receptor subtypes 1 and 2 for Ca2+ -induced Ca2+ release in vascular myocytes. J Biol Chem. 2000;275:9596–9603.
  • Hakamata Y, Nakai J, Takeshima H, et al. Primary structure and distribution of a novel ryanodine receptor/calcium release channel from rabbit brain. FEBS Lett. 1992;312:229–235.
  • Giannini G. The ryanodine receptor/calcium channel genes are widely and differentially expressed in murine brain and peripheral tissues. J Cell Biol. 1995;128:893–904.
  • Pritchard HAT, Pires PW, Yamasaki E, et al. Nanoscale remodeling of ryanodine receptor cluster size underlies cerebral microvascular dysfunction in Duchenne muscular dystrophy. Proc Natl Acad Sci U S A. 2018;115:E9745–E52.
  • Wang L, Yin J, Nickles HT, et al. Hypoxic pulmonary vasoconstriction requires connexin 40-mediated endothelial signal conduction. J Clin Invest. 2012;122:4218–4230.
  • Buschmann I, Pries A, Styp-Rekowska B, et al. Pulsatile shear and Gja5 modulate arterial identity and remodeling events during flow-driven arteriogenesis. Development. 2010;137:2187–2196.
  • Mironneau J, Coussin F, Jeyakumar LH, et al. Contribution of ryanodine receptor subtype 3 to ca2+ responses in Ca2+-overloaded cultured rat portal vein myocytes. J Biol Chem. 2001;276:11257–11264.
  • Bertocchini F, Ovitt CE, Conti A, et al. Requirement for the ryanodine receptor type 3 for efficient contraction in neonatal skeletal muscles. Embo J. 1997;16:6956–6963.
  • Löhn M, Jessner W, Fürstenau M, et al. Regulation of calcium sparks and spontaneous transient outward currents by RyR3 in arterial vascular smooth muscle cells. Circ Res. 2001;89:1051–1057.
  • Sonnleitner A. Functional properties of the ryanodine receptor type 3 (RyR3) Ca2+ release channel. Embo J. 1998;17:2790–2798.
  • Perez CF, Lopez JR, Allen PD. Expression levels of RyR1 and RyR3 control resting free Ca2+ in skeletal muscle. Am J Physiol Cell Physiol. 2005;288:C640–9.
  • Miyatake R, Furukawa A, Matsushita M, et al. Tissue-specific alternative splicing of mouse brain type ryanodine receptor/calcium release channel mRNA. FEBS Lett. 1996;395:123–126.
  • Jiang D, Xiao B, Li X, et al. Smooth muscle tissues express a major dominant negative splice variant of the type 3 Ca2+ release channel (ryanodine receptor). J Biol Chem. 2003;278:4763–4769.
  • Matsuki K, Kato D, Takemoto M, et al. Negative regulation of cellular Ca2+ mobilization by ryanodine receptor type 3 in mouse mesenteric artery smooth muscle. Am J Physiol Cell Physiol. 2018;315:C1–C9.
  • Mederos y Schnitzler M, Storch U, Meibers S, et al. Gq-coupled receptors as mechanosensors mediating myogenic vasoconstriction. Embo J. 2008;27:3092–3103.
  • Arnaudeau S, Macrez-Lepretre N, Mironneau J. Activation of calcium sparks by angiotensin II in vascular myocytes. Biochem Biophys Res Commun. 1996;222:809–815.
  • Bonev AD, Jaggar JH, Rubart M, et al. Activators of protein kinase C decrease Ca2+ spark frequency in smooth muscle cells from cerebral arteries. Am J Physiol Cell Physiol. 1997;273:C2090–5.
  • Bootman M, Niggli E, Berridge M, et al. Imaging the hierarchical Ca2+ signalling system in HeLa cells. J Physiol. 1997;499(2):307–314.
  • Yao Y, Choi J, Parker I. Quantal puffs of intracellular Ca2+ evoked by inositol trisphosphate in Xenopus oocytes. J Physiol. 1995;482(3):533–553.
  • Rainbow RD, Macmillan D, McCarron JG. The sarcoplasmic reticulum Ca2+ store arrangement in vascular smooth muscle. Cell Calcium. 2009;46:313–322.
  • Rahman T, Taylor CW. Dynamic regulation of IP3 receptor clustering and activity by IP3. Channels (Austin). 2009;3:226–232.
  • Yuchi Z, Van Petegem F. Common allosteric mechanisms between ryanodine and inositol-1,4,5-trisphosphate receptors. Channels (Austin). 2011;5:120–123.
  • Bychkov R, Gollasch M, Ried C, et al. Regulation of spontaneous transient outward potassium currents in human coronary arteries. Circulation. 1997;95:503–510.
  • Gollasch M, Ried C, Bychkov R, et al. K+ currents in human coronary artery vascular smooth muscle cells. Circ Res. 1996;78:676–688.
  • Saeed Y, Temple IP, Borbas Z, et al. Structural and functional remodeling of the atrioventricular node with aging in rats: the role of hyperpolarization-activated cyclic nucleotide-gated and ryanodine 2 channels. Heart Rhythm. 2018;15:752–760.
  • Ahir BK, Pratten MK. The impact of caffeine on connexin expression in the embryonic chick cardiomyocyte micromass culture system. J Appl Toxicol. 2016;36:903–913.
  • Welsh DG, Segal SS. Endothelial and smooth muscle cell conduction in arterioles controlling blood flow. A J Physiol. 1998;274:H178–86.
  • Xiong Z, Sperelakis N. Regulation of L-type calcium channels of vascular smooth muscle cells. J Mol Cell Cardiol. 1995;27:75–91.
  • Jung S, Strotmann R, Schultz G, et al. TRPC6 is a candidate channel involved in receptor-stimulated cation currents in A7r5 smooth muscle cells. Am J Physiol Cell Physiol. 2002;282:C347–59.
  • Dietrich A,Y, Schnitzler MM, Gollasch M, et al. Increased vascular smooth muscle contractility in TRPC6-/- mice. Mol Cell Biol. 2005;25:6980–6989.
  • Guan Z, Osmond DA, Inscho EW. P2X receptors as regulators of the renal microvasculature. Trends Pharmacol Sci. 2007;28:646–652.
  • Catterall WA. Voltage-gated calcium channels. Cold Spring Harb Perspect Biol. 2011;3:a003947.
  • Sather WA, Tanabe T, Zhang J-F, et al. Distinctive biophysical and pharmacological properties of class A (BI) calcium channel α1 subunits. Neuron. 1993;11:291–303.
  • Nowycky MC, Fox AP, Tsien RW. Three types of neuronal calcium channel with different calcium agonist sensitivity. Nature. 1985;316:440–443.
  • Casamassima F, Hay AC, Benedetti A, et al. L‐type calcium channels and psychiatric disorders: a brief review. Am J Med Genet B Neuropsychiatr Genet. 2010;153B:1373–1390.
  • Abd El-Rahman RRA, Harraz OF, Brett SE, et al. Identification of L- and T-type Ca2+ channels in rat cerebral arteries: role in myogenic tone development. Am J Physiol Heart Circ Physiol. 2013;304:H58–71.
  • Catterall WA, Perez-Reyes E, Snutch TP, et al. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev. 2005;57:411–425.
  • Hansen PB, Jensen BL, Andreasen D, et al. Differential expression of T- and L-type voltage-dependent calcium channels in renal resistance vessels. Circ Res. 2001;89:630–638.
  • Liao P, Yong TF, Liang MC, et al. Splicing for alternative structures of Ca1.2 Ca channels in cardiac and smooth muscles. Cardiovasc Res. 2005;68:197–203.
  • Lopez-Lopez JR, Hell JW. Increased phosphorylation of the neuronal L-type Ca(2+) channel Ca(v)1.2 during aging. Proc Natl Acad Sci U S A. 2003;100:16018–16023.
  • Takimoto K, Li D, Nerbonne JM, et al. Distribution, splicing and glucocorticoid-induced expression of Cardiacα1Candα1DVoltage-gated Ca2+Channel mRNAs. J Mol Cell Cardiol. 1997;29:3035–3042.
  • Lopez-Lopez JR, Shacklock PS, Balke CW, et al. Local calcium transients triggered by single L-type calcium channel currents in cardiac cells. Science. 1995;268:1042–1045.
  • Augustine GJ, Santamaria F, Tanaka K. Local calcium signaling in neurons. Neuron. 2003;40:331–346.
  • Wray S. Calcium signaling in smooth muscle. Handbook of cell signaling. Elsevier; 2010. p. 1009–1025.
  • Zhang J, Berra-Romani R, Sinnegger-Brauns MJ, et al. Role of Cav1. 2 L-type Ca2+ channels in vascular tone: effects of nifedipine and Mg2+. Am J Physiol Heart Circ Physiol. 2007;292:H415–H25.
  • Striessnig J, Bolz HJ, Koschak A. Channelopathies in Cav1. 1, Cav1. 3, and Cav1. 4 voltage-gated L-type Ca2+ channels. Pflüg Arch Eur J Phy. 2010;460:361–374.
  • Navedo MF, Amberg GC, Westenbroek RE, et al. Cav1. 3 channels produce persistent calcium sparklets, but Cav1. 2 channels are responsible for sparklets in mouse arterial smooth muscle. Am J Physiol Heart Circ Physiol. 2007;293:H1359–H70.
  • Poulsen CB, Al-Mashhadi RH, Cribbs LL, et al. T-type voltage-gated calcium channels regulate the tone of mouse efferent arterioles. Kidney Int. 2011;79:443–451.
  • Hansen PB. Functional and pharmacological consequences of the distribution of voltage-gated calcium channels in the renal blood vessels. Acta Physiol. 2013;207:690–699.
  • Kuo IY, Ellis A, Seymour VA, et al. Dihydropyridine-insensitive calcium currents contribute to function of small cerebral arteries. J Cereb Blood Flow Metab. 2010;30:1226–1239.
  • McNeish AJ, Jimenez Altayo F, Garland CJ. Evidence both L-type and non-L-type voltage-dependent calcium channels contribute to cerebral artery vasospasm following loss of NO in the rat. Vascul Pharmacol. 2010;53:151–159.
  • Braunstein TH, Inoue R, Cribbs L, et al. The role of L- and T-type calcium channels in local and remote calcium responses in rat mesenteric terminal arterioles. J Vasc Res. 2009;46:138–151.
  • Ball CJ, Wilson DP, Turner SP, et al. Heterogeneity of L- and T-channels in the vasculature: rationale for the efficacy of combined L- and T-blockade. Hypertension. 2009;53:654–660.
  • Howitt L, Kuo IY, Ellis A, et al. Chronic deficit in nitric oxide elicits oxidative stress and augments T-type calcium-channel contribution to vascular tone of rodent arteries and arterioles. Cardiovasc Res. 2013;98:449–457.
  • Wang SQ, Song LS, Lakatta EG, et al. Ca2+ signalling between single L-type Ca2+ channels and ryanodine receptors in heart cells. Nature. 2001;410:592–596.
  • Collier ML, Ji G, Wang YX, et al. Calcium-induced calcium release in smooth muscle: loose coupling between the action potential and calcium release. J Gen Physiol. 2000;115:653–662.
  • Wu C, Sui G, Fry CH. The role of the L‐type Ca2+ channel in refilling functional intracellular Ca2+ stores in guinea‐pig detrusor smooth muscle. J Physiol. 2002;538:357–369.
  • Fosset M, Jaimovich E, Delpont E, et al. [3H]nitrendipine receptors in skeletal muscle. J Biol Chem. 1983;258:6086–6092.
  • Block BA, Imagawa T, Campbell KP, et al. Structural evidence for direct interaction between the molecular components of the transverse tubule/sarcoplasmic reticulum junction in skeletal muscle. J Cell Biol. 1988;107:2587–2600.
  • Franzini-Armstrong C, Protasi F, Ramesh V. Shape, size, and distribution of Ca(2+) release units and couplons in skeletal and cardiac muscles. Biophys J. 1999;77:1528–1539.
  • Carl SL, Felix K, Caswell AH, et al. Immunolocalization of sarcolemmal dihydropyridine receptor and sarcoplasmic reticular triadin and ryanodine receptor in rabbit ventricle and atrium. J Cell Biol. 1995;129:673–682.
  • Cheng H, Lederer MR, Xiao RP, et al. Excitation-contraction coupling in heart: new insights from Ca2+ sparks. Cell Calcium. 1996;20:129–140.
  • Lohn M, Furstenau M, Sagach V, et al. Ignition of calcium sparks in arterial and cardiac muscle through caveolae. Circ Res. 2000;87:1034–1039.
  • Harraz OF, Abd El-Rahman RR, Bigdely-Shamloo K, et al. Ca(V)3.2 channels and the induction of negative feedback in cerebral arteries. Circ Res. 2014;115:650–661.
  • Zavaritskaya O, Zhuravleva N, Schleifenbaum J, et al. Role of KCNQ channels in skeletal muscle arteries and periadventitial vascular dysfunction. Hypertension. 2013;61:151–159.
  • Tsvetkov D, Tano JY, Kassmann M, et al. The role of DPO-1 and XE991-sensitive potassium channels in perivascular adipose tissue-mediated regulation of vascular tone. Front Physiol. 2016;7:335.
  • Löhn M, Dubrovska G, Lauterbach B, et al. Periadventitial fat releases a vascular relaxing factor. Faseb J. 2002;16:1057–1063.
  • Perez-Reyes E. Molecular physiology of low-voltage-activated t-type calcium channels. Physiol Rev. 2003;83:117–161.
  • Soong TW, Stea A, Hodson CD, et al. Structure and functional expression of a member of the low voltage-activated calcium channel family. Science. 1993;260:1133–1136.
  • Martin RL, Lee J-H, Cribbs LL, et al. Mibefradil block of cloned T-type calcium channels. J Pharmacol Exp Ther. 2000;295:302–308.
  • Yoo HY, Zheng H, Nam JH, et al. Facilitation of Ca2+-activated K+ channels (IKCa1) by mibefradil in B lymphocytes. Pflüg Arch Eur J Phy. 2008;456:549–560.
  • Viana F, Van Den Bosch L, Missiaen L, et al. Mibefradil (Ro 40m5967) blocks multiple types of voltage-gated calcium channels in cultured rat spinal motoneurones. Cell Calcium. 1997;22:299–311.
  • Strege PR, Bernard CE, Ou Y, et al. Effect of mibefradil on sodium and calcium currents. Am J Physiol Gastrointest Liver Physiol. 2005;289:G249–G53.
  • Lee J-H, Gomora JC, Cribbs LL, et al. Nickel block of three cloned T-type calcium channels: low concentrations selectively block α1H. Biophys J. 1999;77:3034–3042.
  • Kang H-W, Park J-Y, Jeong S-W, et al. A molecular determinant of nickel inhibition in Cav3. 2 T-type calcium channels. J Biol Chem. 2006;281:4823–4830.
  • Mangoni ME, Traboulsie A, Leoni A-L, et al. Bradycardia and slowing of the atrioventricular conduction in mice lacking CaV3. 1/α1G T-type calcium channels. Circ Res. 2006;98:1422–1430.
  • Pluteanu F, Cribbs LL. Regulation and function of Cav3. 1 T-type calcium channels in IGF-1 stimulated pulmonary artery smooth muscle cells. Am J Physiol Cell Physiol. 2010;300: C517–C525.
  • Harraz OF, Brett SE, Zechariah A, et al. Genetic ablation of CaV3.2 channels enhances the arterial myogenic response by modulating the RyR-BKCa axis. Arterioscler Thromb Vasc Biol. 2015;35:1843–1851.
  • Bjorling K, Morita H, Olsen MF, et al. Myogenic tone is impaired at low arterial pressure in mice deficient in the low-voltage-activated CaV 3.1 T-type Ca(2+) channel. Acta Physiol. 2013;207:709–720.
  • Jensen LJ, Nielsen MS, Salomonsson M, et al. T-type Ca(2+) channels and autoregulation of local blood flow. Channels (Austin). 2017;11:183–195.
  • Saeki T, Suzuki Y, Yamamura H, et al. A junctophilin-caveolin interaction enables efficient coupling between ryanodine receptors and BKCa channels in the Ca(2+) microdomain of vascular smooth muscle. J Biol Chem. 2019;294:13093–13105.
  • Gollasch M, Tank J, Luft FC, et al. The BK channel beta1 subunit gene is associated with human baroreflex and blood pressure regulation. J Hypertens. 2002;20:927–933.
  • Behringer EJ, Hakim MA. Functional interaction among KCa and TRP channels for cardiovascular physiology: modern perspectives on aging and chronic disease. Int J Mol Sci. 2019;20:1380.
  • Tsvetkov D, Kolpakov E, Kassmann M, et al. Distinguishing between biological and technical replicates in hypertension research on isolated arteries. Front Med (Lausanne). 2019;6:126.
  • Gollasch M. Adipose-vascular coupling and potential therapeutics. Annu Rev Pharmacol Toxicol. 2017;57:417–436.