Next Article in Journal
Toxic and Teratogenic Effects of Prenatal Alcohol Exposure on Fetal Development, Adolescence, and Adulthood
Next Article in Special Issue
Negative Influence by the Force: Mechanically Induced Hyperpolarization via K2P Background Potassium Channels
Previous Article in Journal
Anti-Cancer and Electrochemical Properties of Thiogenistein—New Biologically Active Compound
Previous Article in Special Issue
Selective Chemical Activation of Piezo1 in Leukemia Cell Membrane: Single Channel Analysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Cross-Talk between Mechanosensitive Ion Channels and Calcium Regulatory Proteins in Cardiovascular Health and Disease

1
Chongqing Key Laboratory of Natural Product Synthesis and Drug Research, School of Pharmaceutical Sciences, Chongqing University, Chongqing 401331, China
2
Leeds Institute of Cardiovascular and Metabolic Medicine, School of Medicine, University of Leeds, Leeds LS2 9JT, UK
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(16), 8782; https://doi.org/10.3390/ijms22168782
Submission received: 28 July 2021 / Revised: 13 August 2021 / Accepted: 14 August 2021 / Published: 16 August 2021
(This article belongs to the Special Issue Mechanosensitive Ion Channels in Health and Disease)

Abstract

:
Mechanosensitive ion channels are widely expressed in the cardiovascular system. They translate mechanical forces including shear stress and stretch into biological signals. The most prominent biological signal through which the cardiovascular physiological activity is initiated or maintained are intracellular calcium ions (Ca2+). Growing evidence show that the Ca2+ entry mediated by mechanosensitive ion channels is also precisely regulated by a variety of key proteins which are distributed in the cell membrane or endoplasmic reticulum. Recent studies have revealed that mechanosensitive ion channels can even physically interact with Ca2+ regulatory proteins and these interactions have wide implications for physiology and pathophysiology. Therefore, this paper reviews the cross-talk between mechanosensitive ion channels and some key Ca2+ regulatory proteins in the maintenance of calcium homeostasis and its relevance to cardiovascular health and disease.

1. Introduction

Despite decades of efforts, cardiovascular disease is still the number one killer in the world. The latest data show that one in six elderly people dies of cardiovascular disease. In 2019, ischemic heart disease and stroke were reported to be the leading causes for disability in the age groups of 50–74 and 75 years or above [1]. As a dominant second messenger, Ca2+ plays an important role in the cardiovascular health and diseases. For example, dietary calcium supplement and its retention reduce cardiovascular response to sodium stress in Black people [2]. There are still some controversies over the cardiovascular effects of high Ca2+ intake in the diet, and whether the relationship between Ca2+ intake and cardiovascular disease risk is J- or U-shape [3]. Ca2+ is pivotal in maintaining the functions of endothelial cells, smooth muscle cells (SMCs) and cardiomyocytes. It controls the contraction or relaxation of arteries and heart and regulates blood pressure and cardiac functions [4,5,6,7].
The cell membrane controls the balance between intracellular and extracellular Ca2+ via various proteins, and thus maintains Ca2+ homeostasis. Plasma membrane Ca2+ ATPase (PMCA), voltage-gated calcium channel (VGCC), Na+/Ca2+ exchanger (NCX), and Orai have been identified as the main calcium regulatory proteins on the cell membrane [8,9,10,11]. Endoplasmic reticulum (ER) as an important Ca2+ reservoir in the cell also contains some calcium regulatory proteins. Such proteins in the ER include stromal interaction molecules (STIM), inositol 1,4,5-trisphosphate receptor (IP3R), and sarco/endoplasmic reticulum calcium-ATPase (SERCA). All these proteins are critical in controlling cell functions, such as growth, migration, apoptosis, and metabolism [12,13,14]. As a general feedback mechanism, these key Ca2+ regulatory proteins are also regulated by intracellular Ca2+ level.
Mechanical forces are crucial for cardiovascular functions and therefore the discoveries of mechanosensitive ion channels represent a major breakthrough in understanding cardiovascular mechanobiology. In particular, the endothelium in the cardiovascular system is subjected to regular mechanical stimulus evoked by blood flow. The discovered mechanosensitive ion channels including Piezo channels and transient receptor potential (TRP) channels can conduct the entry of cationic ions, particularly Ca2+, in response to the stimulus from shear stress of blood flow [15,16]. Both Piezo and TRP channels are closely linked to the development of cardiovascular disease. In some cardiovascular diseases, such as hypertension, atherosclerosis, or aneurysmal plaques, altered mechanical stress which can directly activate mechanosensitive ion channels has been reported [17].
Ca2+ regulatory proteins sensitize any subtle change of intracellular Ca2+. It has emerged that these proteins can cross-talk to mechanosensitive ion channels (Figure 1) and such cross-talk can even take place with direct and physical interactions between them, such as Piezo and SERCA [18]. This paper reviews the cross-talk between mechanosensitive ion channels and some key Ca2+ regulatory proteins in the regulation of Ca2+ homeostasis from the perspective of cardiovascular health and disease. In-depth understanding of the structure, function, and cross-talk of these proteins will help us to unravel the pathogenesis of cardiovascular disease and further develop novel therapeutic strategies. It is noted that the cross-talk or interaction between mechanical ion channels, PMCA, VGCC in cardiovascular system awaits further investigation and hence is not reviewed in this paper.

2. Ca2+ Regulatory Proteins in Cardiovascular System

2.1. NCX

NCX mainly works in the forward mode, which uses the electrochemical gradient driven by Na+ to expel Ca2+ from cells in order to maintain the concentration of Ca2+ required for physiological activities, and the ion stoichiometric ratio is 3Na+:1Ca2+ [19]. There are three types of NCX: NCX1, NCX2, and NCX3 [20]. Structural studies revealed that eukaryotic NCX protein consists of 10 transmembrane helixes. There is a large cytoplasmic regulatory loop between transmembrane helix 5 and 6. This loop includes two regulatory (Ca2+)-binding domains 1 and 2 [21,22], which adjust the rate of from cells to adapt to the dynamic Ca2+ oscillation [21,23,24]. The Ca2+ extrusion usually leads to smooth muscle relaxation, so that the vascular tension is reduced [25]. However, under some extreme conditions, such as high concentration of intracellular Na+ or high positive membrane potential, NCX works in a reverse mode to evoke Ca2+ influx instead of the typical extrusion [23], which can then lead to contraction of SMCs and artery [26,27]. Interestingly, the forward mode of NCX can be changed to its reverse mode by the increase of cytosolic Na+ and membrane potential [28], suggesting a feedback regulatory mechanism may exist [19].
NCX regulates many essential physiological events, such as muscle excitation-contraction or blood pressure regulation [29,30]. The deletion of NCX causes the loss of NCX function in myocardium and consequentially results in embryonic death [25]. The altered expression and regulation of NCXs could disrupt Ca2+ homeostasis and initiate molecular and cellular remodeling in various tissues, which is related to hypertension and heart failure. Inhibitors of NCX can improve myocardial function in patients with heart failure and bring the Ca2+ hyper-responsiveness back to normal in vascular SMCs from hypertensive patients [26,31]. NCX1 in smooth muscle and endothelium could play opposite roles in regulating blood pressure. Arterial blood pressure is correlated with the expression level of NCX1 in vascular SMCs [26]. The increased expression of vascular NCX1 is associated with the vasoconstriction in several animal models of salt-dependent hypertension [26]. Furthermore, reduced arterial myogenic tone and low blood pressure were observed in vascular smooth muscle NCX1 conditional knockout mice. However, for the mice with NCX1 overexpression in vascular SMCs, high blood pressure and vasoconstriction even accompanied with increased expression of transient receptor potential canonical channel (TRPC) 6 were reported [26], which suggests that NCX could control arterial constriction and regulate blood pressure by cross-talking to TRPC6 channels [26]. In the mesentery constricted by phenylephrine, antagonists of NCX reverse mode eliminate acetylcholine-evoked nitric oxide production in intact mesenteric arteries and inhibit acetylcholine- or ATP-induced increase of intracellular Ca2+ in cultured endothelial cells, indicating that the activation of NCX reverse mode can play an important role in mediating the acetylcholine-induced vasodilation in resistance arterial endothelial cells [32].

2.2. Orai

Orai is a highly (Ca2+)-selective ion channel in the plasma membrane formed by four transmembrane domains. Orai family includes Orai1, Orai2, and Orai3 [8]. The calcium release-activated calcium (CRAC) channels are composed of Orai and STIM, representing a typical voltage independent store-operated Ca2+ entry (SOCE) [33,34]. Store-operated Ca2+ channels fine tune Ca2+ entry in both cardiomyocytes and SMCs, and they are activated once the Ca2+ store in ER or sarcoplasmic reticulum (SR) is depleted or the level of cytosolic Ca2+ is lowered, thereby facilitating agonist-induced Ca2+ influx. It has also been suggested that STIM1, Orai and TRPC channels could form the molecular basis of SOCE in some types of cells and their intricate interactions control the entry of Ca2+ into cells to regulate numerous physiological processes [35]. Orai1 in plasma membrane and STIM1 in ER conduct Orai-STIM signaling at the membrane junction between ER and plasma membrane, and they are the bona fide molecular components of SOCE and CRAC [8,36]. Once the ER Ca2+ store is depleted, STIM1 protein can move to the plasma membrane and activate Orai and TRPC channels, allowing extracellular Ca2+ to enter the cytoplasm [8,36]. Orai2 and Orai3 channels have been discovered to be key players in regenerative Ca2+ oscillations induced by physiological receptor activation, while Orai1 is not necessarily involved in this process. However, the binding of Orai2 and Orai3 to Orai1 could expand the sensitivity range of receptor-activated Ca2+ signals [37].
Orai plays a critical role in regulating cardiovascular function in both health and disease [35,38,39]. Orai1 protein deficiency leads to heart failure in zebrafish [40]. The knockout of Orai3 in cardiomyocyte causes dilated cardiomyopathy and heart failure in mice [41]. Both Orai1 and Orai3 are the phenotype modulators of vascular SMCs. Orai1 is upregulated in SMCs during vascular injury. The downregulation of Orai1 inhibits SMC proliferation and reduces neointima formation following balloon injury of rat carotids [42]. Orai3 is also upregulated in neointimal SMCs in rat balloon injured carotid artery, and the knockdown of Orai3 inhibits neointima formation [43]. The transformation of vascular SMC phenotypes is one of the pathological characteristics in chronic hypertension, and the synergistic action between Orai and STIM mediates Ca2+ entry and drives the fibroproliferative gene program [44]. Orai facilitates Ca2+ entry and is a potential therapeutic target for the treatment of hypertension [35]. Most cardiovascular diseases are closely associated with cellular remodeling, and Ca2+ signaling pathways have emerged as important regulators of smooth muscle, endothelial, epithelial, platelet, and immune cell remodeling [45]. Ca2+-permeable Orai channel is also important for endothelial cell proliferation and angiogenesis [46,47]. In terms of vascular physiology and functional regulation, Orai1 appears to trigger the increase of vascular permeability, which is an early marker of atherogenesis. Knockdown of Orai1 reduces the histone 1-induced hyperpermeability in endothelial cells [48]. ApoE knockout mice are a common model for atherosclerosis. A high fat diet can upregulate the expression of Orai1 mRNA and protein in aortic tissue. SiRNA knockdown of Orai1 can reduce the size of atherosclerotic plaque [49]. The migration of neutrophil is another hallmark in atherosclerosis, and during this process Orai1 is required for neutrophil migration to the inflammatory endothelium [45]. All these experimental evidence show that Orai1 expression is associated with development of atherogenesis. Moreover, Orai often acts in conjunction with STIM to form CRAC, which can be responsible for many physiological functions or the development of various cardiovascular diseases [35,39,50].

2.3. STIM

STIM is a single pass transmembrane protein residing in the ER. It contains two homologous proteins, STIM1 and STIM2 [51,52]. The function of STIM is to sense the concentration of Ca2+ in ER and makes appropriate response through conformational change to regulate Ca2+ homeostasis [39,53]. STIM1 stays at a closed state when ER lumen is filled with Ca2+ and transitions to an open state when Ca2+ in the ER lumen is decreased [54]. The Ca2+-sensitive domain in the STIM N-terminal senses Ca2+ level of ER ranging from 100 to 400 µM [52,55], and the C-terminal of STIM interacts with Orai to form CRAC channel to induce Ca2+ influx [52,56]. As mentioned above, the interactions between STIM1 and Orai1 regulate physiological and pathological functions [35,39,50].
STIM is involved in both the cardiac physiological functions and the cardiac disease development. STIM is essential for the maintenance of myocardial contractility, and its knockout leads to a reduction in left ventricular contractility. STIM1 is expressed more abundantly in early cardiomyocytes than in somatic cells. Cardiomyocyte-STIM1-specific knockout mice exhibit dilated cardiomyopathy and cardiac fibrosis with increased stress biomarkers and altered organelle morphology in the heart, suggesting that STIM1 can regulate myocardial development and heart function [39]. However, the spatially differential distribution of STIM1-triggered Ca2+ signaling generates the Ca2+ microdomain that regulates myofilament remodeling and activates pro-hypertrophic factors locally, and as a consequence, pathological cardiac hypertrophy is induced [57]. The STIM1-guided Ca2+ signaling is also involved in thrombosis. The aggregation of platelets at the site of thrombosis requires the increase of intracellular Ca2+ concentration. STIM1 is involved in this process through maintaining a high Ca2+ concentration. In addition, STIM1 stabilizes the thrombus by promoting the expression of phosphatidylserine in plasma membrane [39,58]. The upregulation of STIM induces fibroproliferative gene expression and vascular SMC remodeling, which eventually leads to chronic hypertension [44]. The cell proliferation and migration promoted by STIM1 are also involved in atherosclerosis [59,60]. Oxidized low-density lipoprotein (ox-LDL) can increase the expression of STIM1, and then promote cell proliferation and migration in mouse aortic SMCs. Silencing STIM1 inhibits ox-LDL-induced cell proliferation and migration and hence suppresses atherosclerosis [59,60]. The role of STIM1 in the pathogenesis of these diseases suggests that specific inhibition of STIM1 may contribute to the treatment of these diseases. STIM2 is another important protein for health. Studies on STIM2 deficient mice show that they gradually die from 4 to 8 weeks [61].STIM2 has similar functional effects to STIM1 in some aspects. Both proteins can promote vascular remodeling by inducing the transformation of phenotypes in pulmonary artery SMCs [62].

2.4. IP3R

IP3R is a tetrameric channel consisting of four glycoproteins in the ER or SR. So far, three types of IP3R channels have been identified: IP3R1, IP3R2, and IP3R3 [63]. IP3R has four structural regions: IP3 binding region, central regulatory region, transmembrane domain, and C-terminal region. It can be activated by the selective ligand inositol 1,4,5-triphosphate (IP3) and is permeable to Ca2+ [64]. All these isoforms of IP3R can be expressed in vascular SMCs. They are important for the physiological functioning of the cardiovascular system [65]. IP3R is one of the major sources for intracellular Ca2+ release. The overexpression of IP3R enhances ER Ca2+ depletion, which reduces ER intraluminal Ca2+ concentration in the vicinity of STIM1 and then activates Orai ion channels [66]. In response to increased IP3 or decreased Ca2+ in ER, IP3Rs empty Ca2+ stored in the ER and activate Ca2+ inward flow [67]. IP3R also functions on the membrane contact sites between ER and mitochondria to transport Ca2+ from ER to mitochondria. Each isoform of IP3R can mediate this contact and Ca2+ transport, but IP3R2 is the most efficient one in delivering Ca2+ to mitochondria from ER [68,69]. The voltage-dependent anion channel on the outer mitochondrial membrane can also enhance Ca2+ accumulation through physical interaction with IP3R [70], which is vital for the maintenance of mitochondrial function.
Under physiological conditions, IP3R signal controls the contraction, migration, and proliferation of vascular SMCs. However, under the pathological conditions, IP3R is involved in the development of atherosclerosis and hypertension [71]. IP3R is activated following the stimulation of G-protein coupled receptors and binds to STIM1 upon Ca2+ depletion in ER. The association of IP3R-STIM1 increases Ca2+ inward flow [66]. When IP3 binds with IP3Rs, vasoconstriction and hypertension can be induced as a consequence to the increased concentration of cytoplasmic Ca2+ released from the ER. The deletion of IP3Rs reduces the contractile response to vasoconstrictors and even reverses the pathological states [65]. In the heart, IP3R-mediated Ca2+ release ensures the integrity of cardiac excitation-contraction coupling, which forms the basis of the heartbeat [72]. The dysfunction of IP3R in cardiomyocytes leads to the disturbance of local Ca2+ homeostasis, which is closely related to congenital diseases, increased risk of arrhythmia, decreased contractility, or heart failure related arrhythmias [73]. The expression of IP3R is upregulated in atrial fibrillation, and inhibition of IP3R can significantly reduce the occurrence and duration of atrial fibrillation. Therefore, IP3R may emerge as a new target for the treatment of atrial fibrillation [74].

2.5. SERCA

SERCA is a Ca2+ transporter located on SR/ER and is mainly responsible for the transport of cytoplasmic Ca2+ back to SR/ER. SERCA isoform is encoded by SERCA1, SERCA2, or SERCA3 genes. Each isoform may have differential roles in different tissues or cells [75]. There are four functional domains (M, N, P, and A) and a polypeptide chain in SERCA protein. The M domain contains transmembrane components and Ca2+ binding sites, while N, P and A located in the sarcoplasm are responsible for ATP hydrolysis [76]. Each ATP hydrolysis can transport 2 Ca2+ to the ER lumen in exchange for 1 H+ [77]. SERCA2a is the major isoform of cardiac SERCA, while SERCA2b is the major one of vascular SERCA.
The influx of Ca2+ into SR/ER mediated by SERCA2 is necessary for the relaxation of cardiomyocytes and blood vessels. The disruption of SERCA2 activity leads to ER stress and cardiovascular disease [76]. Hormones, phospholamban and sarcolipin are the common regulators of SERCA. Especially, phospholamban plays a major role in regulating SERCA, and its interaction with SERCA2a reduces the binding affinity of SERCA2a to Ca2+ at low cytoplasmic Ca2+ concentration [78]. The downregulation of SERCA2a is found in failing heart and atherosclerotic vessels [79]. The decreased protein level of SERCA2a and p16-phospholamban leads to left ventricular diastolic dysfunction and elevated arterial blood pressure [80]. Activation of SERCA can accelerate the reuptake of Ca2+ by SR, which would improve the diastolic dysfunction of myocardium, and result in strong antiarrhythmic effect [81,82]. Our groups found that the S-glutathiolation of the amino acid residue Cys674 (C674) is key to the increase of the activity in SERCA2 under physiological conditions [83,84], but this post-translational protein modification is prevented by the irreversible oxidation of C674 thiol in pathology hallmarked by high level of ROS, including atherosclerosis, aortic aneurysms, aging and hypertension [83,85,86,87]. The substitution of the SERCA2 C674 by serine causes impaired angiogenesis following hindlimb ischemia by interrupting the physiological functions of endothelial cells and macrophage [88,89], increases blood pressure by inducing sodium retention and ER stress in the kidney [87], exacerbates angiotensin II-induced aortic aneurysm by switching the phenotypes in aortic SMCs [90], aggravates high fat diet-induced aortic atherosclerosis by evoking inflammatory response in endothelial cells and macrophage (our unpublished data), promotes pulmonary vascular remodeling, and protects against left ventricular dilation caused by chronic ascending aortic constriction [91]. All these data imply that the redox state of C674 and the function of SERCA2 are critical to the maintenance of cardiovascular homeostasis.
Currently, there are some ongoing clinical trials for the drugs specifically targeting these Ca2+ regulatory proteins in the cardiovascular system, as shown in Table 1. These trials provide promising opportunities for the treatment of cardiovascular diseases.

3. Mechanosensitive Ion Channels in Cardiovascular System

3.1. Piezo Channels

Piezo channels are cation channels activated by mechanical force. They have two subtypes: Piezo1 and Piezo2. Piezo1 is the major one distributed in cardiovascular system. Upon mechanical stimulus, Piezo channels convert the mechanical signals into biological signals to participate in cellular physiological events. When Piezo1 channels sense shear stress in cardiovascular system, they conduct the inward flow of cationic ions, especially Ca2+ [99]. Piezo1 in the plasma membrane of endothelial cells can cause depolarization in response to the blood flow [17,100]. Since the discovery of Piezo channels in 2010, more and more evidence have revealed its importance in physiological activities and disease development [99,101,102]. Cryo-electron microscopy studies show that mouse Piezo1 is a three-bladed, propeller-shaped homologous trimeric protein consisting of a central cap, three peripheral blade-like structures on the outside of the cell, three long beams on the inside of the cell connecting the blades to the cap, and a transmembrane region between these features [103,104,105,106].
The systemic knockout or specific endothelial disruption of Piezo1 in mice severely disrupts angiogenesis and results in embryonic death within days of cardiac pulsation [107]. Endothelial Piezo1 mediates atheroprotective or atheroprone signaling depending on the flow pattern [108]. Disturbed flow leads to Piezo1-mediated inflammation, and Piezo1 deficiency in endothelium reduces atherosclerosis [108]. Piezo1 can also regulate vascular tone and arterial blood pressure through increasing intracellular cationic ions and modulating AKT and eNOS phosphorylation [109]. Especially, Piezo1 has an interesting dichotomy in endothelial cells [100,101]. When the channel opens, it causes intracellular elevation Ca2+ and Na+. The latter can depolarize endothelial cells. In this case, depolarization in the endothelium can be transduced to the vascular SMCs in which voltage-gated Ca2+ channels could be activated to drive vasoconstriction [110]. Such vasoconstriction is beneficial in increasing physical performance by re-distributing blood supply in the body to support sustainability of physical activities [110]. The important roles Piezo1 play in the health and disease indicate that the channel would be potentially key therapeutic target for the treatment of some cardiovascular diseases.

3.2. TRP Channels

Based on amino acid sequence, TRP channels can be classified into seven subfamilies: TRPC, TRPV, TRPM, TRPA, TRPP, TRPML, and TRPN. They share some common structural features including six transmembrane domains and a pore lining between the fifth and sixth transmembrane domains [111,112]. Most TRP channels are Ca2+-permeable, but their Ca2+ permeability varies [113]. For example, the permeability of TRPV4 to Ca2+ is much greater than that of TRPC6 [114], while TRPM4 channel is almost impermeable to Ca2+ [115]. As mechanosensitive ion channels, TRP proteins can regulate contraction, relaxation, proliferation, differentiation, and apoptosis in cardiovascular physiology or pathophysiology [116]. Meanwhile, they can also be stimulated by the depletion of SR/ER Ca2+ stores [113,117]. Recent studies have shown that TRPC1 channels can form store-operated Ca2+ channels with a contractile phenotype independently of Orai1 in vascular SMCs [118].
TRP channels control acute hemodynamics and regulate cardiac remodeling through the modulation of Ca2+ signaling, which is relevant to physiological and pathological functions in the heart [119]. Shear stress is almost universally associated with vasodilation in the peripheral vasculature, and there is considerable evidence that TRPV4 channel-mediated Ca2+ entry following stimulation of shear stress mediates the vasodilatory response in endothelial cells [120,121]. TRPV2 channel is mainly expressed in the sarcomeres and ER to maintain appropriate mechano-electric coupling in cardiomyocyte. Under pathological conditions, TRPV2 is translocated to the sarcolemma to mediate an abnormal Ca2+ entry [122]. TRPA1 channel activated by pressure overload increases Ca2+ inward flow and activates Ca2+-dependent pathways, which would lead to diastolic-mediated heart failure in cardiomyocytes [123,124]. TRPA1 expression is increased in failing hearts. Inhibition of TRPA1 channel improves myocardial hypertrophy and heart failure [125]. TRPA1 channels in cardiomyocytes contribute nearly 40% of the increase of intracellular Ca2+ concentration under physiological conditions, which then induces nitric oxide release and vasodilation [124,126]. TRPA1 has a similar role to TRPV4 in lowering blood pressure through such vasodilation [127,128]. From these perspectives, there are some similarities between TRP channels and the above-mentioned Piezo channels in regulating blood pressure. Both types of mechanosensitive ion channels exert their effects through the ions they conduct in response to mechanical force and the downstream signaling pathways in vasculature.
Clinical trials for drugs specifically targeting these mechanosensitive ion channels are being run in the cardiovascular system, as shown in Table 2.

4. Cross-Talk between Mechanosensitive Ion Channels and Ca2+ Regulatory Proteins in Cardiovascular System

There are many similarities between mechanosensitive ion channels and Ca2+ regulatory proteins in the regulation of intracellular Ca2+. To date, many scientific reports have shown that they both can cross-talk to each other so that intracellular Ca2+ signaling can be precisely coordinated in cardiovascular health and diseases (as listed in Table 3). For some physiological activities, both types of proteins can even physically interact with each other to regulate Ca2+ concentration.

4.1. Piezo and SERCA

Pizeo and SERCA have been found to cross-talk to each other through physical interaction in endothelial cells. SERCA2 inhibits Piezo1-dependent endothelial cell migration through a 14-residue linker region [18]. Mutations in this linker impair the interaction between Piezo1 and SERCA2. As a consequence, the mechanical stimulation of Piezo1 channel activity is significantly weakened. In addition, a synthesized linker peptide also disrupts the regulatory role of SERCA2 in Piezo1 channel, providing further evidence that the linker mediates the physical interaction between SERCA2 and Piezo1 [18]. In Piezo1-deficient elegans, the interference of ER Ca2+ stores by SERCA RNAi results in severe reproductive defects [151], reiterating that the importance of cross-talk between Piezo1 and SERCA in the maintenance of physiological functions. When overexpressed human Piezo1 were stimulated by shear stress with or without the presence of SERCA inhibitor thapsigargin in HEK293 cells, Ca2+ measurement experiments showed that nearly 72% of the increase in Ca2+ is from the mechanically activated Piezo1 while 28% comes from the Ca2+ store depletion due to SERCA inactivation [152], suggesting that SERCA inactivation can upregulate Piezo1 activity. Our unpublished data in pulmonary artery SMCs indicates that the substitution of the SERCA2 C674 reactive thiol by serine increases the activity of Piezo as well. This further supports the regulatory role of SERCA2 in mechanically-activated Piezo1 activity, though the exact molecular mechanism remains to be investigated. So far, the interactions between Piezo and SERCA have not been sufficiently reported in the development of cardiovascular disease. Considering the overlapping roles of Piezo and SERCA at the lesion site in some cardiovascular disease such as arterial hypertension and aneurysms, we postulate that both proteins also cross-talk to regulate Ca2+ signaling, albeit rather differently, in these diseases.

4.2. TRP and NCX

NCX is involved in the regulation of intracellular Ca2+ by triggering the Ca2+ entry mediated by TRP-formed SOC channels. Pulmonary hypertension, as an example, is closely related to both NCX and TRP proteins. Dysfunction of either protein disrupts the Ca2+ homeostasis, which changes contraction, proliferation, and migration in SMCs of pulmonary artery. Gradually the hypertension is developed [136,137].
The cross-talk between TRP channels and NCX can regulate contractile dysfunction and spontaneous ectopic [119]. Growing evidence indicates that NCX cannot function properly without the involvement of TRP channels. TRP channels directly interact with NCX to regulate the Ca2+ homeostasis. Co-immunoprecipitation experiments show the co-localization of TRPC3 and NCX1, suggesting that TRPC3 cross-talks to NCX1 through physical interaction in the Ca2+ regulation [119]. Glutathione s-transferase pull-down assays consistently replicate the natural TRPC3/NCX1 complex in cardiomyocytes. The NCX-mediated Ca2+ signaling is significantly reduced in cells with negative TRPC3 expression [153]. In addition to physical interactions, there is also functional coupling underlying the cross-talk between these two types of proteins. Mechanical stimulation activates TRPV1, TRPV2 or TRPV4 and then Ca2+ entry is induced. The increased intracellular Ca2+ concentration in turn activates NCX to extrude the cytoplasmic Ca2+ [154]. NCX and TRP can also be synergistically involved in the Ca2+ entry by cooperatively mediating the hypoxia-induced increase in intracellular Ca2+ and then causing vasoconstriction [137]. However, the NCX inhibitor KB-R7943 can only be effective in inhibiting TRPC3, TRPC5 or TRPC6. Also, the coupling of TRPM4 with the reverse mode of NCX can induce atrial arrhythmia [138]. However, it should be noted that not all types of TRP channels are functionally coupled to NCX [155].

4.3. TRP and Orai

TRP and Orai are both located in the cell membrane [139]. The interaction between TRP and Orai is often inseparable from STIM1. A summary of TRP, Orai1 and STIM revealed that these three proteins can also interact with each other or together. STIM can activate both Orai1 and TRP, and then either TRPC1-STIM1 or Orai1-STIM1-TRPC complex is formed to mediate SOCE [139]. Some experimental evidence have suggested that the function of TRPC1 is not only dependent on STIM1, but also Orai. The knockdown of Orai protein disrupts the normal function of TRPC channels [36]. As SOCE has a profound effect on the development of pulmonary hypertension, inhibition of any such protein can be used as an effective treatment of pulmonary hypertension [140]. In addition, SOCE is also involved in thrombosis thrombin production [141].

4.4. TRP and STIM

TRP can cross-talk to STIM1 through physical interaction between them [139]. The interaction between TRP and STIM is often triggered by intracellular Ca2+ change. Upon depletion of Ca2+ stores in the ER, STIM1 can form a dimer and then activate TRPC channels. However, the TRPC channel function is not always dependent on STIM1 and therefore the interaction between STIM1 and TRPC may be absent in some cells [36]. When they form a complex together, they can coordinate Ca2+ signal to regulate another protein. It has been observed that TRPC1, TRPC2, or TRPC4 can interact directly with the ezrin/radixin/moesin domain of STIM1 [156]. Interaction between TRPC1 protein and STIM1 could stimulate phospholipase C and then induce TRPC1 channel activation in vascular SMCs [157]. Fluorescence resonance energy transfer (FRET), immunoprecipitation and total internal reflection fluorescent microscope (TIRFM) experiments have been extensively carried out to demonstrate that TRPC1 strongly co-localizes and interacts with STIM1 after depletion of ER Ca2+ stores. In addition, the Orai1-mediated Ca2+ entry into the cytosol can also trigger TRPC1 interaction with STIM1. In this case, each TRPC1 tetramer binds to two STIM1s and the 639DD640 of TRPC1 C-terminus is then directly interacted by the 684KK689 region of STIM1 C-terminus [142]. Caveolin-1 as a key scaffold in the plasma membrane where TRP proteins reside has been reported to underlie the formation of TRPC-STIM1 complex. Its knockdown prevents TRPC binding to STIM1 [158]. Though the interaction of TRPC1 with STIM1 may be required for SOCE activation, the knockdown of TRPC1 decreases SOCE activity by approximately 60%, while the deletion of STIM1 completely abolished SOCE activity [142]. The interaction between TRP and STIM is not always dependent on Ca2+ store depletion. When the EF-hand of STIM1 is mutated, it spontaneously aggregates with TRPC1 and then activates TRPC1 [159]. The interaction between TRP and STIM is implicated in the progression of many diseases, particularly some cardiovascular diseases such as arterial hypertension, atherosclerosis, or thrombosis [141,143,144,145,146].

4.5. TRP and IP3R

The IP3R-initiated Ca2+ depletion is followed by activation of TRP channels to promote Ca2+ influx, which supports the notion that IP3R cross-talks to TRP channels to regulate intracellular Ca2+ [67]. This is further confirmed by the interaction of IP3R with TRP [71]. The interaction between TRP and IP3R has been studied in more details with structural approaches. It has been found that TRPV4 or other member of TRP family can interact with IP3Rs. The IP3R interacts with TRP protein in a Ca2+-dependent manner. Also, IP3R has differential affinity to various TRP member. This affinity ranges from 10 nm for TRP2 to 290 nm for TRP6 [160]. Prior studies showed that the interaction between IP3R and TRP could not be achieved without Homer1. Homer1 is a scaffolding protein that enhances the physical interaction between TRPC1 and IP3R [161]. Co-immunoprecipitation and co-localization both demonstrate the direct interactions between Homer1b/c isoforms, IP3R and TRPC2 [162]. The coupling of IP3R to TRP forms the mechanism underlying the physiological vasoconstriction or the endothelin-1-induced hypertensive vasoconstriction [147,148]. Therefore, such coupling and cross-talk not only contribute to vascular SMCs contraction, but also is associated with hypertensive resistance [163,164].

4.6. TRP and SERCA

Though the direct physical interaction between TRP channels and SERCA has not yet reported, numerous studies have shown that they both can cross-talk to each other. SERCA pump inhibitor thapsigargin blocks the activation of TRPC7 channel by diacylglycerol or the activation of TRPC3 channel [165]. The SERCA modulator 2-aminoethyl diphenylborinate also has a significant inhibitory effect on TRPV6 activity [166]. The Ca2+ signaling compensatory mechanism in which the transcription of TRPC4 and TRPC5 are increased after SERCA silencing has been proposed to be responsible for cardiac development and hypertrophy [150]. A similar compensatory mechanism also occurs under hypoxia. Hypoxia inhibits SERCA activity and increases the Ca2+ influx through TRPC6 [167]. The expression of TRPC1 is inversely correlated with SERCA2 [168]. SERCA also regulates the expression and activity of TRP channels. Silencing SERCA2 increases the transcription and activity of TRPC4 and TRPC5 in cardiomyocytes, suggesting that intracellular Ca2+ can be compensated by TRP channels when the Ca2+ in the store is insufficient [150]. However, it remains unclear whether down-regulation of TRPC4/5 inversely up-regulates the expression of SERCA2. The cross-talk between the TRP channel and SERCA contributes to the development and function of heart. It has been evidenced that the rapid cyclic change of intracellular Ca2+ concentration during cardiac pulsation is a consequential result of temporal cross-talk between SERCA and TRP. Thus these proteins could be promising targets in treating cardiac disease [149].

5. Conclusions

The multifaceted effects of the Ca2+ signaling pathway in both physiology and pathology require the dynamic cross-talk between mechanosensitive ion channels and Ca2+ regulatory proteins. Ca2+ regulatory proteins are diverse and widely distributed in the cells, mostly on the cell membrane or in ER. Notably, certain pathogenic factors are largely activated by changes in intracellular Ca2+ concentrations, and pharmacological modulation of these proteins has emerged as an attractive strategy to prevent cellular dysfunction and tissue remodeling, and eventually to treat cardiovascular disease. Mechanosensitive ion channels play a dominant role in endothelium-dependent vascular development and blood pressure regulation, vascular tension control, angiogenesis and the atherosclerotic process. When either mechanosensitive ion channel or Ca2+ regulatory protein is activated, a Ca2+-related pathway is often triggered and then the cross-talk between both proteins is frequently initiated. In addition, it is interesting to observe that in some cardiovascular diseases, such as hypertension, atherosclerosis or aneurysms, mechanical stress or shear stress is altered, and inherent activation of the mechanosensitive ion channels are reported. Accordingly, Ca2+ regulatory proteins are communicated by the channels on a temporal and spatial basis. In some cells, when mechanosensitive ion channel expression is defective or disrupted, the inability to sense the mechanical stimuli directly affects the regulation of downstream Ca2+-related pathways, which leads to defects in some physiological functions. The cross-talk between these mechanosensitive ion channels and Ca2+ regulatory proteins in the cardiovascular system is therefore not only relevant to the maintenance of health but is also involved in the development of diseases.
The summary of recent articles on the role of mechanosensitive ion channels and Ca2+ regulatory proteins in cardiovascular health and disease through the regulation of Ca2+ related pathways reveals an apparent cross-talk between both types of proteins. The cross-talk occurs with either physical interaction or functional coupling. Especially, the physical interaction in such cross-talk has emerged to be important for the cardiovascular physiology or pathology. We have summarized some interactions between mechanosensitive ion channels and Ca2+ regulatory proteins. The study of protein–protein interaction depends on FRET, TIRFM, and immunoprecipitation techniques. With the comprehensive application of protein-protein interaction methods and the development of electron microscopy technology, we expect that more interactions between mechanosensitive ion channels and Ca2+ regulatory proteins will be discovered. In addition, the functional coupling in the cross-talk between the channels and Ca2+ regulatory proteins will also be unraveled with functional assays combined with appropriate animal models. These studies will certainly be useful in identifying novel therapeutic approaches for treatment of cardiovascular diseases.

Author Contributions

Conceptualization, Y.W., J.S. and X.T.; writing—original draft preparation, Y.W.; writing—review and editing, J.S. and X.T.; supervision, J.S. and X.T.; project administration, J.S. and X.T.; funding acquisition, J.S. and X.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Natural Science Foundation of China (31571172, 81870343, Tong X) and the British Heart Foundation grants (FS/17/2/32559, PG/21/10595, Shi J).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vos, T.; Lim, S.S.; Abbafati, C.; Abbas, K.M.; Abbasi, M.; Abbasifard, M.; Abbasi-Kangevari, M.; Abbastabar, H.; Abd-Allah, F.; Abdelalim, A.; et al. Global burden of 369 diseases and injuries in 204 countries and territories, 1990–2019: A systematic analysis for the Global Burden of Disease Study 2019. Lancet 2020, 396, 1204–1222. [Google Scholar] [CrossRef]
  2. Ernst, F.A.; Enwonwu, C.O.; Francis, R.A. Calcium attenuates cardiovascular reactivity to sodium and stress in blacks. Am. J. Hypertens. 1990, 3 Pt 1, 451–457. [Google Scholar] [CrossRef]
  3. Mohammadifard, N.; Gotay, C.; Humphries, K.H.; Ignaszewski, A.; Esmaillzadeh, A.; Sarrafzadegan, N. Electrolyte minerals intake and cardiovascular health. Crit. Rev. Food Sci. Nutr. 2019, 59, 2375–2385. [Google Scholar] [CrossRef]
  4. Ottolini, M.; Hong, K.; Sonkusare, S.K. Calcium signals that determine vascular resistance. Wiley Interdiscip. Rev. Syst. Biol. Med. 2019, 11, e1448. [Google Scholar] [CrossRef] [PubMed]
  5. Wilson, C.; Zhang, X.; Buckley, C.; Heathcote, H.R.; Lee, M.D.; McCarron, J.G. Increased vascular contractility in hypertension results from impaired endothelial calcium signaling. Hypertension 2019, 74, 1200–1214. [Google Scholar] [CrossRef] [PubMed]
  6. Gusev, K.O.; Vigont, V.V.; Grekhnev, D.A.; Shalygin, A.V.; Glushankova, L.N.; Kaznacheeva, E.V. Store-operated calcium entry in mouse cardiomyocytes. Bull. Exp. Biol. Med. 2019, 167, 311–314. [Google Scholar] [CrossRef] [PubMed]
  7. Gorski, P.A.; Kho, C.; Oh, J.G. Measuring cardiomyocyte contractility and calcium handling in vitro. Methods Mol. Biol. 2018, 1816, 93–104. [Google Scholar]
  8. Trebak, M.; Putney, J.W., Jr. ORAI calcium channels. Physiology 2017, 32, 332–342. [Google Scholar] [CrossRef] [PubMed]
  9. Lariccia, V.; Piccirillo, S.; Preziuso, A.; Amoroso, S.; Magi, S. Cracking the code of sodium/calcium exchanger (NCX) gating: Old and new complexities surfacing from the deep web of secondary regulations. Cell Calcium 2020, 87, 102169. [Google Scholar] [CrossRef]
  10. Ferreira-Gomes, M.S.; Mangialavori, I.C.; Ontiveros, M.Q.; Rinaldi, D.E.; Martiarena, J.; Verstraeten, S.V.; Rossi, J. Selectivity of plasma membrane calcium ATPase (PMCA)-mediated extrusion of toxic divalent cations in vitro and in cultured cells. Arch. Toxicol. 2018, 92, 273–288. [Google Scholar] [CrossRef] [PubMed]
  11. Gilbert, G.; Courtois, A.; Dubois, M.; Cussac, L.A.; Ducret, T.; Lory, P.; Marthan, R.; Savineau, J.P.; Quignard, J.F. T-type voltage gated calcium channels are involved in endothelium-dependent relaxation of mice pulmonary artery. Biochem. Pharmacol. 2017, 138, 61–72. [Google Scholar] [CrossRef]
  12. Krebs, J.; Agellon, L.B.; Michalak, M. Ca2+ homeostasis and endoplasmic reticulum (ER) stress: An integrated view of calcium signaling. Biochem. Biophys. Res. Commun. 2015, 460, 114–121. [Google Scholar] [CrossRef] [PubMed]
  13. Marchi, S.; Patergnani, S.; Missiroli, S.; Morciano, G.; Rimessi, A.; Wieckowski, M.R.; Giorgi, C.; Pinton, P. Mitochondrial and endoplasmic reticulum calcium homeostasis and cell death. Cell Calcium 2018, 69, 62–72. [Google Scholar] [CrossRef]
  14. Schachter, M. Vascular smooth muscle cell migration, atherosclerosis, and calcium channel blockers. Int. J. Cardiol. 1997, 62 (Suppl. 2), S85–S90. [Google Scholar] [CrossRef]
  15. Chubinskiy-Nadezhdin, V.I.; Vasileva, V.Y.; Pugovkina, N.A.; Vassilieva, I.O.; Morachevskaya, E.A.; Nikolsky, N.N.; Negulyaev, Y.A. Local calcium signalling is mediated by mechanosensitive ion channels in mesenchymal stem cells. Biochem. Biophys. Res. Commun. 2017, 482, 563–568. [Google Scholar] [CrossRef] [PubMed]
  16. Ilkan, Z.; Wright, J.R.; Goodall, A.H.; Gibbins, J.M.; Jones, C.I.; Mahaut-Smith, M.P. Evidence for shear-mediated Ca2+ entry through mechanosensitive cation channels in human platelets and a megakaryocytic cell line. J. Biol. Chem. 2017, 292, 9204–9217. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Beech, D.J.; Kalli, A.C. Force sensing by Piezo channels in cardiovascular health and disease. Arterioscler. Thromb. Vasc. Biol. 2019, 39, 2228–2239. [Google Scholar] [CrossRef]
  18. Zhang, T.; Chi, S.; Jiang, F.; Zhao, Q.; Xiao, B. A protein interaction mechanism for suppressing the mechanosensitive Piezo channels. Nat. Commun. 2017, 8, 1797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  19. Giladi, M.; Tal, I.; Khananshvili, D. Structural features of ion transport and allosteric regulation in Sodium-Calcium Exchanger (NCX) proteins. Front. Physiol. 2016, 7, 30. [Google Scholar] [CrossRef] [Green Version]
  20. Héja, L.; Kardos, J. NCX activity generates spontaneous Ca2+ oscillations in the astrocytic leaflet microdomain. Cell Calcium 2020, 86, 102137. [Google Scholar] [CrossRef]
  21. Hilge, M.; Aelen, J.; Vuister, G.W. Ca2+ regulation in the Na+/Ca2+ exchanger involves two markedly different Ca2+ sensors. Mol. Cell 2006, 22, 15–25. [Google Scholar] [CrossRef]
  22. Liao, J.; Li, H.; Zeng, W.; Sauer, D.B.; Belmares, R.; Jiang, Y. Structural insight into the ion-exchange mechanism of the sodium/calcium exchanger. Science 2012, 335, 686–690. [Google Scholar] [CrossRef] [PubMed]
  23. Blaustein, M.P.; Lederer, W.J. Sodium/calcium exchange: Its physiological implications. Physiol. Rev. 1999, 79, 763–854. [Google Scholar] [CrossRef]
  24. Philipson, K.D.; Nicoll, D.A. Sodium-calcium exchange: A molecular perspective. Annu. Rev. Physiol. 2000, 62, 111–133. [Google Scholar] [CrossRef]
  25. Nishimura, J. Topics on the Na+/Ca2+ exchanger: Involvement of Na+/Ca2+ exchanger in the vasodilator-induced vasorelaxation. J. Pharmacol. Sci. 2006, 102, 27–31. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Zhang, J. New insights into the contribution of arterial NCX to the regulation of myogenic tone and blood pressure. Adv. Exp. Med. Biol. 2013, 961, 329–343. [Google Scholar] [PubMed]
  27. Li, M.; Shang, Y.X. Neurokinin-1 receptor antagonist decreases [Ca2+]i in airway smooth muscle cells by reducing the reverse-mode Na+/Ca2+ exchanger current. Peptides 2019, 115, 69–74. [Google Scholar] [CrossRef]
  28. Alves-Lopes, R.; Neves, K.B.; Anagnostopoulou, A.; Rios, F.J.; Lacchini, S.; Montezano, A.C.; Touyz, R.M. Crosstalk between vascular redox and calcium signaling in hypertension involves TRPM2 (Transient Receptor Potential Melastatin 2) cation channel. Hypertension 2020, 75, 139–149. [Google Scholar] [CrossRef] [PubMed]
  29. Khananshvili, D. Sodium-calcium exchangers (NCX): Molecular hallmarks underlying the tissue-specific and systemic functions. Pflug. Arch. Eur. J. Physiol. 2014, 466, 43–60. [Google Scholar] [CrossRef]
  30. Filadi, R.; Pozzan, T. Generation and functions of second messengers microdomains. Cell Calcium 2015, 58, 405–414. [Google Scholar] [CrossRef]
  31. Primessnig, U.; Bracic, T.; Levijoki, J.; Otsomaa, L.; Pollesello, P.; Falcke, M.; Pieske, B.; Heinzel, F.R. Long-term effects of Na+ /Ca2+ exchanger inhibition with ORM-11035 improves cardiac function and remodelling without lowering blood pressure in a model of heart failure with preserved ejection fraction. Eur. J. Heart Fail. 2019, 21, 1543–1552. [Google Scholar] [CrossRef]
  32. Lillo, M.A.; Gaete, P.S.; Puebla, M.; Ardiles, N.M.; Poblete, I.; Becerra, A.; Simon, F.; Figueroa, X.F. Critical contribution of Na+-Ca2+ exchanger to the Ca2+-mediated vasodilation activated in endothelial cells of resistance arteries. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2018, 32, 2137–2147. [Google Scholar]
  33. Gudlur, A.; Hogan, P.G. The STIM-Orai pathway: Orai, the pore-forming subunit of the CRAC channel. Adv. Exp. Med. Biol. 2017, 993, 39–57. [Google Scholar]
  34. Nguyen, N.T.; Han, W.; Cao, W.M.; Wang, Y.; Wen, S.; Huang, Y.; Li, M.; Du, L.; Zhou, Y. Store-operated calcium entry mediated by ORAI and STIM. Compr. Physiol. 2018, 8, 981–1002. [Google Scholar] [PubMed]
  35. Bhullar, S.K.; Shah, A.K.; Dhalla, N.S. Store-operated calcium channels: Potential target for the therapy of hypertension. Rev. Cardiovasc. Med. 2019, 20, 139–151. [Google Scholar]
  36. Choi, S.; Maleth, J.; Jha, A.; Lee, K.P.; Kim, M.S.; So, I.; Ahuja, M.; Muallem, S. The TRPCs-STIM1-Orai interaction. Handb. Exp. Pharmacol. 2014, 223, 1035–1054. [Google Scholar] [PubMed]
  37. Yoast, R.E.; Emrich, S.M.; Zhang, X.; Xin, P.; Johnson, M.T.; Fike, A.J.; Walter, V.; Hempel, N.; Yule, D.I.; Sneyd, J.; et al. The native ORAI channel trio underlies the diversity of Ca2+ signaling events. Nat. Commun. 2020, 11, 2444. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, W.; Trebak, M. STIM1 and Orai1: Novel targets for vascular diseases? Sci. China Life Sci. 2011, 54, 780–785. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Tanwar, J.; Trebak, M.; Motiani, R.K. Cardiovascular and hemostatic disorders: Role of STIM and Orai proteins in vascular disorders. Adv. Exp. Med. Biol. 2017, 993, 425–452. [Google Scholar]
  40. Völkers, M.; Dolatabadi, N.; Gude, N.; Most, P.; Sussman, M.A.; Hassel, D. Orai1 deficiency leads to heart failure and skeletal myopathy in zebrafish. J. Cell Sci. 2012, 125 (Pt 2), 287–294. [Google Scholar] [CrossRef] [Green Version]
  41. Gammons, J.; Trebak, M.; Mancarella, S. Cardiac-specific deletion of Orai3 leads to severe dilated cardiomyopathy and heart failure in mice. J. Am. Heart Assoc. 2021, 10, e019486. [Google Scholar] [CrossRef]
  42. Zhang, W.; Halligan, K.E.; Zhang, X.; Bisaillon, J.M.; Gonzalez-Cobos, J.C.; Motiani, R.K.; Hu, G.; Vincent, P.A.; Zhou, J.; Barroso, M.; et al. Orai1-mediated I (CRAC) is essential for neointima formation after vascular injury. Circ. Res. 2011, 109, 534–542. [Google Scholar] [CrossRef] [Green Version]
  43. González-Cobos, J.C.; Zhang, X.; Zhang, W.; Ruhle, B.; Motiani, R.K.; Schindl, R.; Muik, M.; Spinelli, A.M.; Bisaillon, J.M.; Shinde, A.V.; et al. Store-independent Orai1/3 channels activated by intracrine leukotriene C4: Role in neointimal hyperplasia. Circ. Res. 2013, 112, 1013–1025. [Google Scholar] [CrossRef] [Green Version]
  44. Johnson, M.T.; Gudlur, A.; Zhang, X.; Xin, P.; Emrich, S.M.; Yoast, R.E.; Courjaret, R.; Nwokonko, R.M.; Li, W.; Hempel, N.; et al. L-type Ca2+ channel blockers promote vascular remodeling through activation of STIM proteins. Proc. Natl. Acad. Sci. USA 2020, 117, 17369–17380. [Google Scholar] [CrossRef] [PubMed]
  45. Johnson, M.; Trebak, M. ORAI channels in cellular remodeling of cardiorespiratory disease. Cell Calcium 2019, 79, 1–10. [Google Scholar] [CrossRef] [PubMed]
  46. Bai, S.; Wei, Y.; Hou, W.; Yao, Y.; Zhu, J.; Hu, X.; Chen, W.; Du, Y.; He, W.; Shen, B.; et al. Orai-IGFBP3 signaling complex regulates high-glucose exposure-induced increased proliferation, permeability, and migration of human coronary artery endothelial cells. BMJ Open Diabetes Res. Care 2020, 8, e001400. [Google Scholar] [CrossRef] [PubMed]
  47. Li, J.; Cubbon, R.M.; Wilson, L.A.; Amer, M.S.; McKeown, L.; Hou, B.; Majeed, Y.; Tumova, S.; Seymour, V.A.; Taylor, H.; et al. Orai1 and CRAC channel dependence of VEGF-activated Ca2+ entry and endothelial tube formation. Circ. Res. 2011, 108, 1190–1198. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Zou, M.; Dong, H.; Meng, X.; Cai, C.; Li, C.; Cai, S.; Xue, Y. Store-operated Ca2+ entry plays a role in HMGB1-induced vascular endothelial cell hyperpermeability. PLoS ONE 2015, 10, e0123432. [Google Scholar] [CrossRef] [Green Version]
  49. Liang, S.J.; Zeng, D.Y.; Mai, X.Y.; Shang, J.Y.; Wu, Q.Q.; Yuan, J.N.; Yu, B.X.; Zhou, P.; Zhang, F.R.; Liu, Y.Y.; et al. Inhibition of Orai1 store-operated calcium channel prevents foam cell formation and atherosclerosis. Arterioscler. Thromb. Vasc. Biol. 2016, 36, 618–628. [Google Scholar] [CrossRef] [Green Version]
  50. Samanta, K.; Parekh, A.B. Spatial Ca2+ profiling: Decrypting the universal cytosolic Ca2+ oscillation. J. Physiol. 2017, 595, 3053–3062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Feldman, C.H.; Grotegut, C.A.; Rosenberg, P.B. The role of STIM1 and SOCE in smooth muscle contractility. Cell Calcium 2017, 63, 60–65. [Google Scholar] [CrossRef]
  52. Fahrner, M.; Grabmayr, H.; Romanin, C. Mechanism of STIM activation. Curr. Opin. Physiol. 2020, 17, 74–79. [Google Scholar] [CrossRef]
  53. Soboloff, J.; Rothberg, B.S.; Madesh, M.; Gill, D.L. STIM proteins: Dynamic calcium signal transducers. Nat. Rev. Mol. Cell Biol. 2012, 13, 549–565. [Google Scholar] [CrossRef] [Green Version]
  54. Yu, F.; Sun, L.; Hubrack, S.; Selvaraj, S.; Machaca, K. Intramolecular shielding maintains the ER Ca²⁺ sensor STIM1 in an inactive conformation. J. Cell Sci. 2013, 126 Pt 11, 2401–2410. [Google Scholar]
  55. Brandman, O.; Liou, J.; Park, W.S.; Meyer, T. STIM2 is a feedback regulator that stabilizes basal cytosolic and endoplasmic reticulum Ca2+ levels. Cell 2007, 131, 1327–1339. [Google Scholar] [CrossRef] [Green Version]
  56. Grabmayr, H.; Romanin, C.; Fahrner, M. STIM proteins: An ever-expanding family. Int. J. Mol. Sci. 2020, 22, 378. [Google Scholar] [CrossRef] [PubMed]
  57. Parks, C.; Alam, M.A.; Sullivan, R.; Mancarella, S. STIM1-dependent Ca2+ microdomains are required for myofilament remodeling and signaling in the heart. Sci. Rep. 2016, 6, 25372. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Gilio, K.; van Kruchten, R.; Braun, A.; Berna-Erro, A.; Feijge, M.A.; Stegner, D.; van der Meijden, P.E.; Kuijpers, M.J.; Varga-Szabo, D.; Heemskerk, J.W.; et al. Roles of platelet STIM1 and Orai1 in glycoprotein VI- and thrombin-dependent procoagulant activity and thrombus formation. J. Biol. Chem. 2010, 285, 23629–23638. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Fang, M.; Li, Y.; Wu, Y.; Ning, Z.; Wang, X.; Li, X. miR-185 silencing promotes the progression of atherosclerosis via targeting stromal interaction molecule 1. Cell Cycle 2019, 18, 682–695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  60. Mao, Y.Y.; Wang, J.Q.; Guo, X.X.; Bi, Y.; Wang, C.X. Circ-SATB2 upregulates STIM1 expression and regulates vascular smooth muscle cell proliferation and differentiation through miR-939. Biochem. Biophys. Res. Commun. 2018, 505, 119–125. [Google Scholar] [CrossRef] [PubMed]
  61. Berna-Erro, A.; Jardin, I.; Salido, G.M.; Rosado, J.A. Role of STIM2 in cell function and physiopathology. J. Physiol. 2017, 595, 3111–3128. [Google Scholar] [CrossRef] [Green Version]
  62. Fernandez, R.A.; Wan, J.; Song, S.; Smith, K.A.; Gu, Y.; Tauseef, M.; Tang, H.; Makino, A.; Mehta, D.; Yuan, J.X. Upregulated expression of STIM2, TRPC6, and Orai2 contributes to the transition of pulmonary arterial smooth muscle cells from a contractile to proliferative phenotype. Am. J. Physiol. Cell Physiol. 2015, 308, C581–C593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Zhang, X.; Huang, R.; Zhou, Y.; Zhou, W.; Zeng, X. IP3R channels in male reproduction. Int. J. Mol. Sci. 2020, 21, 9179. [Google Scholar] [CrossRef]
  64. Serysheva, I.I. Toward a high-resolution structure of IP₃R channel. Cell Calcium 2014, 56, 125–132. [Google Scholar] [CrossRef] [Green Version]
  65. Lin, Q.; Zhao, G.; Fang, X.; Peng, X.; Tang, H.; Wang, H.; Jing, R.; Liu, J.; Lederer, W.J.; Chen, J.; et al. IP(3) receptors regulate vascular smooth muscle contractility and hypertension. JCI Insight 2016, 1, e89402. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Sampieri, A.; Santoyo, K.; Asanov, A.; Vaca, L. Association of the IP3R to STIM1 provides a reduced intraluminal calcium microenvironment, resulting in enhanced store-operated calcium entry. Sci. Rep. 2018, 8, 13252. [Google Scholar] [CrossRef]
  67. Boulay, G.; Brown, D.M.; Qin, N.; Jiang, M.; Dietrich, A.; Zhu, M.X.; Chen, Z.; Birnbaumer, M.; Mikoshiba, K.; Birnbaumer, L. Modulation of Ca2+ entry by polypeptides of the inositol 1,4, 5-trisphosphate receptor (IP3R) that bind transient receptor potential (TRP): Evidence for roles of TRP and IP3R in store depletion-activated Ca2+ entry. Proc. Natl. Acad. Sci. USA 1999, 96, 14955–14960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  68. Hamada, K.; Mikoshiba, K. IP(3) receptor plasticity underlying diverse functions. Annu. Rev. Physiol. 2020, 82, 151–176. [Google Scholar] [CrossRef] [Green Version]
  69. Bartok, A.; Weaver, D.; Golenár, T.; Nichtova, Z.; Katona, M.; Bánsághi, S.; Alzayady, K.J.; Thomas, V.K.; Ando, H.; Mikoshiba, K.; et al. IP(3) receptor isoforms differently regulate ER-mitochondrial contacts and local calcium transfer. Nat. Commun. 2019, 10, 3726. [Google Scholar] [CrossRef] [Green Version]
  70. Szabadkai, G.; Bianchi, K.; Várnai, P.; De Stefani, D.; Wieckowski, M.R.; Cavagna, D.; Nagy, A.I.; Balla, T.; Rizzuto, R. Chaperone-mediated coupling of endoplasmic reticulum and mitochondrial Ca2+ channels. J. Cell Biol. 2006, 175, 901–911. [Google Scholar] [CrossRef] [Green Version]
  71. Narayanan, D.; Adebiyi, A.; Jaggar, J.H. Inositol trisphosphate receptors in smooth muscle cells. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H2190–H2210. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Luo, X.; Li, W.; Künzel, K.; Henze, S.; Cyganek, L.; Strano, A.; Poetsch, M.S.; Schubert, M.; Guan, K. IP3R-mediated compensatory mechanism for calcium handling in human induced pluripotent stem cell-derived cardiomyocytes with cardiac ryanodine receptor deficiency. Front. Cell Dev. Biol. 2020, 8, 772. [Google Scholar] [CrossRef] [PubMed]
  73. Stokke, M.K.; Rivelsrud, F.; Sjaastad, I.; Sejersted, O.M.; Swift, F. From global to local: A new understanding of cardiac electromechanical coupling. Tidsskr. Nor. Laegeforen. Tidsskr. Prakt. Med. Raekke 2012, 132, 1457–1460. [Google Scholar] [CrossRef] [Green Version]
  74. Xiao, J.; Liang, D.; Zhao, H.; Liu, Y.; Zhang, H.; Lu, X.; Liu, Y.; Li, J.; Peng, L.; Chen, Y.H. 2-Aminoethoxydiphenyl borate, a inositol 1,4,5-triphosphate receptor inhibitor, prevents atrial fibrillation. Exp. Biol. Med. 2010, 235, 862–868. [Google Scholar] [CrossRef]
  75. Periasamy, M.; Kalyanasundaram, A. SERCA pump isoforms: Their role in calcium transport and disease. Muscle Nerve 2007, 35, 430–442. [Google Scholar] [CrossRef]
  76. Rahate, K.; Bhatt, L.K.; Prabhavalkar, K.S. SERCA stimulation: A potential approach in therapeutics. Chem. Biol. Drug Des. 2020, 95, 5–15. [Google Scholar] [CrossRef]
  77. Shaikh, S.A.; Sahoo, S.K.; Periasamy, M. Phospholamban and sarcolipin: Are they functionally redundant or distinct regulators of the Sarco(Endo)Plasmic Reticulum Calcium ATPase? J. Mol. Cell. Cardiol. 2016, 91, 81–91. [Google Scholar] [CrossRef] [Green Version]
  78. Periasamy, M.; Bhupathy, P.; Babu, G.J. Regulation of sarcoplasmic reticulum Ca2+ ATPase pump expression and its relevance to cardiac muscle physiology and pathology. Cardiovasc. Res. 2008, 77, 265–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Lipskaia, L.; Keuylian, Z.; Blirando, K.; Mougenot, N.; Jacquet, A.; Rouxel, C.; Sghairi, H.; Elaib, Z.; Blaise, R.; Adnot, S.; et al. Expression of sarco (endo) plasmic reticulum calcium ATPase (SERCA) system in normal mouse cardiovascular tissues, heart failure and atherosclerosis. Biochim. Biophys. Acta 2014, 1843, 2705–2718. [Google Scholar] [CrossRef]
  80. Dupont, S.; Maizel, J.; Mentaverri, R.; Chillon, J.M.; Six, I.; Giummelly, P.; Brazier, M.; Choukroun, G.; Tribouilloy, C.; Massy, Z.A.; et al. The onset of left ventricular diastolic dysfunction in SHR rats is not related to hypertrophy or hypertension. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H1524–H1532. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Fernandez-Tenorio, M.; Niggli, E. Stabilization of Ca2+ signaling in cardiac muscle by stimulation of SERCA. J. Mol. Cell. Cardiol. 2018, 119, 87–95. [Google Scholar] [CrossRef] [PubMed]
  82. Torre, E.; Lodrini, A.; Barassi, P.; Ferrandi, M.; Boz, E.; Bussadori, C.; Ferrari, P.; Bianchi, G.; Rocchetti, M. Istaroxime improves diabetic diastolic dysfunction through SERCA stimulation. Arch. Cardiovasc. Dis. Suppl. 2019, 11, 234–235. [Google Scholar] [CrossRef]
  83. Adachi, T.; Weisbrod, R.M.; Pimentel, D.R.; Ying, J.; Sharov, V.S.; Schöneich, C.; Cohen, R.A. S-Glutathiolation by peroxynitrite activates SERCA during arterial relaxation by nitric oxide. Nat. Med. 2004, 10, 1200–1207. [Google Scholar] [CrossRef] [PubMed]
  84. Tong, X.; Ying, J.; Pimentel, D.R.; Trucillo, M.; Adachi, T.; Cohen, R.A. High glucose oxidizes SERCA cysteine-674 and prevents inhibition by nitric oxide of smooth muscle cell migration. J. Mol. Cell. Cardiol. 2008, 44, 361–369. [Google Scholar] [CrossRef] [Green Version]
  85. Qin, F.; Siwik, D.A.; Lancel, S.; Zhang, J.; Kuster, G.M.; Luptak, I.; Wang, L.; Tong, X.; Kang, Y.J.; Cohen, R.A.; et al. Hydrogen peroxide-mediated SERCA cysteine 674 oxidation contributes to impaired cardiac myocyte relaxation in senescent mouse heart. J. Am. Heart Assoc. 2013, 2, e000184. [Google Scholar] [CrossRef] [Green Version]
  86. Ying, J.; Sharov, V.; Xu, S.; Jiang, B.; Gerrity, R.; Schoneich, C.; Cohen, R.A. Cysteine-674 oxidation and degradation of sarcoplasmic reticulum Ca2+ ATPase in diabetic pig aorta. Free Radic. Biol. Med. 2008, 45, 756–762. [Google Scholar] [CrossRef] [Green Version]
  87. Liu, G.; Wu, F.; Jiang, X.; Que, Y.; Qin, Z.; Hu, P.; Lee, K.S.S.; Yang, J.; Zeng, C.; Hammock, B.D.; et al. Inactivation of Cys(674) in SERCA2 increases BP by inducing endoplasmic reticulum stress and soluble epoxide hydrolase. Br. J. Pharmacol. 2020, 177, 1793–1805. [Google Scholar] [CrossRef]
  88. Thompson, M.D.; Mei, Y.; Weisbrod, R.M.; Silver, M.; Shukla, P.C.; Bolotina, V.M.; Cohen, R.A.; Tong, X. Glutathione adducts on sarcoplasmic/endoplasmic reticulum Ca2+ ATPase Cys-674 regulate endothelial cell calcium stores and angiogenic function as well as promote ischemic blood flow recovery. J. Biol. Chem. 2014, 289, 19907–19916. [Google Scholar] [CrossRef] [Green Version]
  89. Mei, Y.; Thompson, M.D.; Shiraishi, Y.; Cohen, R.A.; Tong, X. Sarcoplasmic/endoplasmic reticulum Ca2+ ATPase C674 promotes ischemia- and hypoxia-induced angiogenesis via coordinated endothelial cell and macrophage function. J. Mol. Cell. Cardiol. 2014, 76, 275–282. [Google Scholar] [CrossRef] [Green Version]
  90. Que, Y.; Shu, X.; Wang, L.; Hu, P.; Wang, S.; Xiong, R.; Liu, J.; Chen, H.; Tong, X. Inactivation of cysteine 674 in the SERCA2 accelerates experimental aortic aneurysm. J. Mol. Cell. Cardiol. 2020, 139, 213–224. [Google Scholar] [CrossRef]
  91. Goodman, J.B.; Qin, F.; Morgan, R.J.; Chambers, J.M.; Croteau, D.; Siwik, D.A.; Hobai, I.; Panagia, M.; Luptak, I.; Bachschmid, M.; et al. Redox-resistant SERCA [Sarco(endo)plasmic Reticulum Calcium ATPase] attenuates oxidant-stimulated mitochondrial calcium and apoptosis in cardiac myocytes and pressure overload-induced myocardial failure in mice. Circulation 2020, 142, 2459–2469. [Google Scholar] [CrossRef]
  92. SERCA. Available online: https://clinicaltrials.gov/ct2/show/NCT01643330?term=SERCA&draw=2&rank=5 (accessed on 12 August 2021).
  93. SERCA. Available online: https://clinicaltrials.gov/ct2/show/NCT01966887?term=SERCA&draw=2&rank=2 (accessed on 12 August 2021).
  94. SERCA. Available online: https://clinicaltrials.gov/ct2/show/NCT00534703?term=SERCA&draw=2&rank=1 (accessed on 12 August 2021).
  95. SERCA. Available online: https://clinicaltrials.gov/ct2/show/NCT04703842?term=SERCA&draw=2&rank=4 (accessed on 12 August 2021).
  96. SERCA. Available online: https://clinicaltrials.gov/ct2/show/NCT02772068?term=SERCA&draw=2&rank=3 (accessed on 12 August 2021).
  97. George, M.; Rajaram, M.; Shanmugam, E.; VijayaKumar, T.M. Novel drug targets in clinical development for heart failure. Eur. J. Clin. Pharmacol. 2014, 70, 765–774. [Google Scholar] [CrossRef] [PubMed]
  98. Na+/Ca2+ Exchanger. Available online: https://clinicaltrials.gov/ct2/show/NCT00534703?term=Na%2B%2FCa2%2B+exchanger&draw=2&rank=1 (accessed on 12 August 2021).
  99. Parpaite, T.; Coste, B. Piezo channels. Curr. Biol. 2017, 27, R250–R252. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  100. Coste, B.; Murthy, S.E.; Mathur, J.; Schmidt, M.; Mechioukhi, Y.; Delmas, P.; Patapoutian, A. Piezo1 ion channel pore properties are dictated by C-terminal region. Nat. Commun. 2015, 6, 7223. [Google Scholar] [CrossRef] [PubMed]
  101. Coste, B.; Mathur, J.; Schmidt, M.; Earley, T.J.; Ranade, S.; Petrus, M.J.; Dubin, A.E.; Patapoutian, A. Piezo1 and Piezo2 are essential components of distinct mechanically activated cation channels. Science 2010, 330, 55–60. [Google Scholar] [CrossRef] [Green Version]
  102. Steinecker-Frohnwieser, B.; Kullich, W.; Kratschmann, C.; Cezanne, M.; Toegel, S.; Weigl, L.J.O. Cartilage, Activation of the mechanosensitive ion channel PIEZO1/2 by YODA1 modulates cellular functions of human oa chondrocytes. Osteoarthr. Cartil. 2020, 28, S101. [Google Scholar] [CrossRef]
  103. Zhao, Q.; Zhou, H.; Chi, S.; Wang, Y.; Wang, J.; Geng, J.; Wu, K.; Liu, W.; Zhang, T.; Dong, M.Q.; et al. Structure and mechanogating mechanism of the Piezo1 channel. Nature 2018, 554, 487–492. [Google Scholar] [CrossRef]
  104. Guo, Y.R.; MacKinnon, R. Structure-based membrane dome mechanism for Piezo mechanosensitivity. Elife 2017, 6, e33660. [Google Scholar] [CrossRef]
  105. Wang, L.; Zhou, H.; Zhang, M.; Liu, W.; Deng, T.; Zhao, Q.; Li, Y.; Lei, J.; Li, X.; Xiao, B. Structure and mechanogating of the mammalian tactile channel PIEZO2. Nature 2019, 573, 225–229. [Google Scholar] [CrossRef]
  106. Zhao, Q.; Zhou, H.; Li, X.; Xiao, B. The mechanosensitive Piezo1 channel: A three-bladed propeller-like structure and a lever-like mechanogating mechanism. FEBS J. 2019, 286, 2461–2470. [Google Scholar] [CrossRef] [Green Version]
  107. Li, J.; Hou, B.; Tumova, S.; Muraki, K.; Bruns, A.; Ludlow, M.J.; Sedo, A.; Hyman, A.J.; McKeown, L.; Young, R.S.; et al. Piezo1 integration of vascular architecture with physiological force. Nature 2014, 515, 279–282. [Google Scholar] [CrossRef] [PubMed]
  108. Albarrán-Juárez, J.; Iring, A.; Wang, S.; Joseph, S.; Grimm, M.; Strilic, B.; Wettschureck, N.; Althoff, T.F.; Offermanns, S. Piezo1 and G(q)/G(11) promote endothelial inflammation depending on flow pattern and integrin activation. J. Exp. Med. 2018, 215, 2655–2672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Wang, S.; Chennupati, R.; Kaur, H.; Iring, A.; Wettschureck, N.; Offermanns, S. Endothelial cation channel PIEZO1 controls blood pressure by mediating flow-induced ATP release. J. Clin. Investig. 2016, 126, 4527–4536. [Google Scholar] [CrossRef]
  110. Rode, B.; Shi, J.; Endesh, N.; Drinkhill, M.J.; Webster, P.J.; Lotteau, S.J.; Bailey, M.A.; Yuldasheva, N.Y.; Ludlow, M.J.; Cubbon, R.M.; et al. Piezo1 channels sense whole body physical activity to reset cardiovascular homeostasis and enhance performance. Nat. Commun. 2017, 8, 350. [Google Scholar] [CrossRef]
  111. Ramsey, I.S.; Delling, M.; Clapham, D.E. An introduction to TRP channels. Annu. Rev. Physiol. 2006, 68, 619–647. [Google Scholar] [CrossRef] [Green Version]
  112. Li, H. TRP channel classification. Adv. Exp. Med. Biol. 2017, 976, 1–8. [Google Scholar] [PubMed]
  113. Mulier, M.; Vriens, J.; Voets, T. TRP channel pores and local calcium signals. Cell Calcium 2017, 66, 19–24. [Google Scholar] [CrossRef]
  114. Hill-Eubanks, D.C.; Gonzales, A.L.; Sonkusare, S.K.; Nelson, M.T. Vascular TRP channels: Performing under pressure and going with the flow. Physiology 2014, 29, 343–360. [Google Scholar] [CrossRef] [Green Version]
  115. Nilius, B.; Prenen, J.; Droogmans, G.; Voets, T.; Vennekens, R.; Freichel, M.; Wissenbach, U.; Flockerzi, V. Voltage dependence of the Ca2+-activated cation channel TRPM4. J. Biol. Chem. 2003, 278, 30813–30820. [Google Scholar] [CrossRef] [Green Version]
  116. Yue, Z.; Xie, J.; Yu, A.S.; Stock, J.; Du, J.; Yue, L. Role of TRP channels in the cardiovascular system. Am. J. Physiol. Heart Circ. Physiol. 2015, 308, H157–H182. [Google Scholar] [CrossRef] [Green Version]
  117. Shi, J.; Miralles, F.; Birnbaumer, L.; Large, W.A.; Albert, A.P. Store depletion induces Gαq-mediated PLCβ1 activity to stimulate TRPC1 channels in vascular smooth muscle cells. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2016, 30, 702–715. [Google Scholar] [CrossRef] [PubMed]
  118. Shi, J.; Miralles, F.; Kinet, J.P.; Birnbaumer, L.; Large, W.A.; Albert, A.P. Evidence that Orai1 does not contribute to store-operated TRPC1 channels in vascular smooth muscle cells. Channels 2017, 11, 329–339. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Ezeani, M. TRP channels mediated pathological Ca2+-handling and spontaneous ectopy. Front. Cardiovasc. Med. 2019, 6, 83. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Bubolz, A.H.; Mendoza, S.A.; Zheng, X.; Zinkevich, N.S.; Li, R.; Gutterman, D.D.; Zhang, D.X. Activation of endothelial TRPV4 channels mediates flow-induced dilation in human coronary arterioles: Role of Ca2+ entry and mitochondrial ROS signaling. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H634–H642. [Google Scholar] [CrossRef] [Green Version]
  121. Köhler, R.; Heyken, W.T.; Heinau, P.; Schubert, R.; Si, H.; Kacik, M.; Busch, C.; Grgic, I.; Maier, T.; Hoyer, J. Evidence for a functional role of endothelial transient receptor potential V4 in shear stress-induced vasodilatation. Arterioscler. Thromb. Vasc. Biol. 2006, 26, 1495–1502. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Entin-Meer, M.; Keren, G. Potential roles in cardiac physiology and pathology of the cation channel TRPV2 expressed in cardiac cells and cardiac macrophages: A mini-review. Am. J. Physiol. Heart Circ. Physiol. 2020, 318, H181–H188. [Google Scholar] [CrossRef]
  123. Wang, Z.; Xu, Y.; Wang, M.; Ye, J.; Liu, J.; Jiang, H.; Ye, D.; Wan, J. TRPA1 inhibition ameliorates pressure overload-induced cardiac hypertrophy and fibrosis in mice. EBioMedicine 2018, 36, 54–62. [Google Scholar] [CrossRef] [Green Version]
  124. Wang, Z.; Ye, D.; Ye, J.; Wang, M.; Liu, J.; Jiang, H.; Xu, Y.; Zhang, J.; Chen, J.; Wan, J. The TRPA1 channel in the cardiovascular system: Promising features and challenges. Front. Pharmacol. 2019, 10, 1253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  125. Andrei, S.R.; Sinharoy, P.; Bratz, I.N.; Damron, D.S. TRPA1 is functionally co-expressed with TRPV1 in cardiac muscle: Co-localization at z-discs, costameres and intercalated discs. Channels 2016, 10, 395–409. [Google Scholar] [CrossRef]
  126. Shang, S.; Zhu, F.; Liu, B.; Chai, Z.; Wu, Q.; Hu, M.; Wang, Y.; Huang, R.; Zhang, X.; Wu, X.; et al. Intracellular TRPA1 mediates Ca2+ release from lysosomes in dorsal root ganglion neurons. J. Cell Biol. 2016, 215, 369–381. [Google Scholar] [CrossRef] [Green Version]
  127. Earley, S.; Pauyo, T.; Drapp, R.; Tavares, M.J.; Liedtke, W.; Brayden, J.E. TRPV4-dependent dilation of peripheral resistance arteries influences arterial pressure. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1096–H1102. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Baratchi, S.; Almazi, J.G.; Darby, W.; Tovar-Lopez, F.J.; Mitchell, A.; McIntyre, P. Shear stress mediates exocytosis of functional TRPV4 channels in endothelial cells. Cell. Mol. Life Sci. CMLS 2016, 73, 649–666. [Google Scholar] [CrossRef]
  129. Piezo. Available online: https://clinicaltrials.gov/ct2/show/NCT04372498?term=Senicapoc+%2CPiezo&draw=2&rank=1 (accessed on 12 August 2021).
  130. Machado, R.F.; Gladwin, M.T. Chronic sickle cell lung disease: New insights into the diagnosis, pathogenesis and treatment of pulmonary hypertension. Br. J. Haematol. 2005, 129, 449–464. [Google Scholar] [CrossRef] [PubMed]
  131. Transient Receptor Potential. Available online: https://clinicaltrials.gov/ct2/show/NCT02533076?term=Cystinosis%2CTRPV1&draw=2&rank=1 (accessed on 12 August 2021).
  132. Kasimer, R.N.; Langman, C.B. Adult complications of nephropathic cystinosis: A systematic review. Pediatric Nephrol. 2021, 36, 223–236. [Google Scholar] [CrossRef] [PubMed]
  133. Transient Receptor Potential. Available online: https://clinicaltrials.gov/ct2/show/NCT02497937?term=GSK2798745%2CTRPV4&draw=2&rank=1 (accessed on 12 August 2021).
  134. Transient Receptor Potential. Available online: https://clinicaltrials.gov/ct2/show/NCT01408446?term=Transient+Receptor+Potential%2CMenthol&draw=2&rank=3 (accessed on 12 August 2021).
  135. Silva, H. Current knowledge on the vascular effects of menthol. Front. Physiol. 2020, 11, 298. [Google Scholar] [CrossRef] [PubMed]
  136. Zhang, S.; Yuan, J.X.; Barrett, K.E.; Dong, H. Role of Na+/Ca2+ exchange in regulating cytosolic Ca2+ in cultured human pulmonary artery smooth muscle cells. Am. J. Physiol. Cell Physiol. 2005, 288, C245–C252. [Google Scholar] [CrossRef] [PubMed]
  137. Meng, F.; To, W.K.; Gu, Y. Inhibition effect of arachidonic acid on hypoxia-induced [Ca2+](i) elevation in PC12 cells and human pulmonary artery smooth muscle cells. Respir. Physiol. Neurobiol. 2008, 162, 18–23. [Google Scholar] [CrossRef] [PubMed]
  138. Watanabe, H.; Murakami, M.; Ohba, T.; Ono, K.; Ito, H. The pathological role of transient receptor potential channels in heart disease. Circ. J. Off. J. Jpn. Circ. Soc. 2009, 73, 419–427. [Google Scholar] [CrossRef] [Green Version]
  139. Bodnar, D.; Chung, W.Y.; Yang, D.; Hong, J.H.; Jha, A.; Muallem, S. STIM-TRP pathways and microdomain organization: Ca2+ influx channels: The Orai-STIM1-TRPC complexes. Adv. Exp. Med. Biol. 2017, 993, 139–157. [Google Scholar]
  140. Reyes, R.V.; Castillo-Galán, S.; Hernandez, I.; Herrera, E.A.; Ebensperger, G.; Llanos, A.J. Revisiting the role of TRP, Orai, and ASIC channels in the pulmonary arterial response to hypoxia. Front. Physiol. 2018, 9, 486. [Google Scholar] [CrossRef] [Green Version]
  141. Berna-Erro, A.; Jardín, I.; Smani, T.; Rosado, J.A. Regulation of platelet function by Orai, STIM and TRP. Adv. Exp. Med. Biol. 2016, 898, 157–181. [Google Scholar] [PubMed]
  142. Ong, H.L.; Ambudkar, I.S. STIM-TRP pathways and microdomain organization: Contribution of TRPC1 in store-operated Ca2+ entry: Impact on Ca2+ signaling and cell function. Adv. Exp. Med. Biol. 2017, 993, 159–188. [Google Scholar] [PubMed]
  143. Cussac, L.A.; Cardouat, G.; Tiruchellvam Pillai, N.; Campagnac, M.; Robillard, P.; Montillaud, A.; Guibert, C.; Gailly, P.; Marthan, R.; Quignard, J.F.; et al. TRPV4 channel mediates adventitial fibroblast activation and adventitial remodeling in pulmonary hypertension. Am. J. Physiol. Lung Cell. Mol. Physiol. 2020, 318, L135–L146. [Google Scholar] [CrossRef]
  144. Hong, K.S.; Lee, M.G. Endothelial Ca2+ signaling-dependent vasodilation through transient receptor potential channels. Korean J. Physiol. Pharmacol. Off. J. Korean Physiol. Soc. Korean Soc. Pharmacol. 2020, 24, 287–298. [Google Scholar] [CrossRef]
  145. Castillo-Galán, S.; Arenas, G.A.; Reyes, R.V.; Krause, B.J.; Iturriaga, R. Stim-activated TRPC-ORAI channels in pulmonary hypertension induced by chronic intermittent hypoxia. Pulm. Circ. 2020, 10 (Suppl. 1), 13–22. [Google Scholar] [CrossRef]
  146. Mammadova-Bach, E.; Nagy, M.; Heemskerk, J.W.M.; Nieswandt, B.; Braun, A. Store-operated calcium entry in thrombosis and thrombo-inflammation. Cell Calcium 2019, 77, 39–48. [Google Scholar] [CrossRef]
  147. Birnbaumer, L.; Boulay, G.; Brown, D.; Jiang, M.; Dietrich, A.; Mikoshiba, K.; Zhu, X.; Qin, N. Mechanism of capacitative Ca2+ entry (CCE): Interaction between IP3 receptor and TRP links the internal calcium storage compartment to plasma membrane CCE channels. Recent Prog. Horm. Res. 2000, 55, 127–161. [Google Scholar]
  148. Adebiyi, A.; Thomas-Gatewood, C.M.; Leo, M.D.; Kidd, M.W.; Neeb, Z.P.; Jaggar, J.H. An elevation in physical coupling of type 1 inositol 1,4,5-trisphosphate (IP3) receptors to transient receptor potential 3 (TRPC3) channels constricts mesenteric arteries in genetic hypertension. Hypertension 2012, 60, 1213–1219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. Freichel, M.; Berlin, M.; Schürger, A.; Mathar, I.; Bacmeister, L.; Medert, R.; Frede, W.; Marx, A.; Segin, S.; Londoño, J.E.C. Frontiers in neuroscience TRP channels in the heart. In Neurobiology of TRP Channels; Emir, T.L.R., Ed.; CRC Press/Taylor & Francis © 2018 by Taylor & Francis Group, LLC.: Boca Raton, FL, USA, 2017; pp. 149–185. [Google Scholar]
  150. Seth, M.; Sumbilla, C.; Mullen, S.P.; Lewis, D.; Klein, M.G.; Hussain, A.; Soboloff, J.; Gill, D.L.; Inesi, G. Sarco(endo)plasmic reticulum Ca2+ ATPase (SERCA) gene silencing and remodeling of the Ca2+ signaling mechanism in cardiac myocytes. Proc. Natl. Acad. Sci. USA 2004, 101, 16683–16688. [Google Scholar] [CrossRef] [Green Version]
  151. Bai, X.; Bouffard, J.; Lord, A.; Brugman, K.; Sternberg, P.W.; Cram, E.J.; Golden, A. Caenorhabditis elegans PIEZO channel coordinates multiple reproductive tissues to govern ovulation. Elife 2020, 9, e53603. [Google Scholar] [CrossRef] [PubMed]
  152. Maneshi, M.M.; Gottlieb, P.A.; Hua, S.Z. A Microfluidic approach for studying Piezo channels. Curr. Top. Membr. 2017, 79, 309–334. [Google Scholar]
  153. Eder, P.; Probst, D.; Rosker, C.; Poteser, M.; Wolinski, H.; Kohlwein, S.D.; Romanin, C.; Groschner, K. Phospholipase C-dependent control of cardiac calcium homeostasis involves a TRPC3-NCX1 signaling complex. Cardiovasc. Res. 2007, 73, 111–119. [Google Scholar] [CrossRef] [Green Version]
  154. Sato, M.; Sobhan, U.; Tsumura, M.; Kuroda, H.; Soya, M.; Masamura, A.; Nishiyama, A.; Katakura, A.; Ichinohe, T.; Tazaki, M.; et al. Hypotonic-induced stretching of plasma membrane activates transient receptor potential vanilloid channels and sodium-calcium exchangers in mouse odontoblasts. J. Endod. 2013, 39, 779–787. [Google Scholar] [CrossRef] [PubMed]
  155. Kraft, R. The Na+/Ca2+ exchange inhibitor KB-R7943 potently blocks TRPC channels. Biochem. Biophys. Res. Commun. 2007, 361, 230–236. [Google Scholar] [CrossRef] [PubMed]
  156. Huang, G.N.; Zeng, W.; Kim, J.Y.; Yuan, J.P.; Han, L.; Muallem, S.; Worley, P.F. STIM1 carboxyl-terminus activates native SOC, I(crac) and TRPC1 channels. Nat. Cell Biol. 2006, 8, 1003–1010. [Google Scholar] [CrossRef]
  157. Shi, J.; Miralles, F.; Birnbaumer, L.; Large, W.A.; Albert, A.P. Store-operated interactions between plasmalemmal STIM1 and TRPC1 proteins stimulate PLCβ1 to induce TRPC1 channel activation in vascular smooth muscle cells. J. Physiol. 2017, 595, 1039–1058. [Google Scholar] [CrossRef] [PubMed]
  158. Pani, B.; Ong, H.L.; Brazer, S.C.; Liu, X.; Rauser, K.; Singh, B.B.; Ambudkar, I.S. Activation of TRPC1 by STIM1 in ER-PM microdomains involves release of the channel from its scaffold caveolin-1. Proc. Natl. Acad. Sci. USA 2009, 106, 20087–20092. [Google Scholar] [CrossRef] [Green Version]
  159. Pani, B.; Ong, H.L.; Liu, X.; Rauser, K.; Ambudkar, I.S.; Singh, B.B. Lipid rafts determine clustering of STIM1 in endoplasmic reticulum-plasma membrane junctions and regulation of store-operated Ca2+ entry (SOCE). J. Biol. Chem. 2008, 283, 17333–17340. [Google Scholar] [CrossRef] [Green Version]
  160. Tang, J.; Lin, Y.; Zhang, Z.; Tikunova, S.; Birnbaumer, L.; Zhu, M.X. Identification of common binding sites for calmodulin and inositol 1,4,5-trisphosphate receptors on the carboxyl termini of trp channels. J. Biol. Chem. 2001, 276, 21303–21310. [Google Scholar] [CrossRef] [Green Version]
  161. Yuan, J.P.; Kiselyov, K.; Shin, D.M.; Chen, J.; Shcheynikov, N.; Kang, S.H.; Dehoff, M.H.; Schwarz, M.K.; Seeburg, P.H.; Muallem, S.; et al. Homer binds TRPC family channels and is required for gating of TRPC1 by IP3 receptors. Cell 2003, 114, 777–789. [Google Scholar] [CrossRef] [Green Version]
  162. Mast, T.G.; Brann, J.H.; Fadool, D.A. The TRPC2 channel forms protein-protein interactions with Homer and RTP in the rat vomeronasal organ. BMC Neurosci. 2010, 11, 61. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Carlström, M.; Wilcox, C.S.; Arendshorst, W.J. Renal autoregulation in health and disease. Physiol. Rev. 2015, 95, 405–511. [Google Scholar] [CrossRef] [Green Version]
  164. Eid, A.H.; El-Yazbi, A.F.; Zouein, F.; Arredouani, A.; Ouhtit, A.; Rahman, M.M.; Zayed, H.; Pintus, G.; Abou-Saleh, H. Inositol 1,4,5-trisphosphate receptors in hypertension. Front. Physiol. 2018, 9, 1018. [Google Scholar] [CrossRef]
  165. Lemonnier, L.; Trebak, M.; Lievremont, J.P.; Bird, G.S.; Putney, J.W., Jr. Protection of TRPC7 cation channels from calcium inhibition by closely associated SERCA pumps. FASEB J. Off. Publ. Fed. Am. Soc. Exp. Biol. 2006, 20, 503–505. [Google Scholar] [CrossRef]
  166. Hofer, A.; Kovacs, G.; Zappatini, A.; Leuenberger, M.; Hediger, M.A.; Lochner, M. Design, synthesis and pharmacological characterization of analogs of 2-aminoethyl diphenylborinate (2-APB), a known store-operated calcium channel blocker, for inhibition of TRPV6-mediated calcium transport. Bioorg. Med. Chem. 2013, 21, 3202–3213. [Google Scholar] [CrossRef] [PubMed]
  167. Liu, B.; Wang, D.; Luo, E.; Hou, J.; Qiao, Y.; Yan, G.; Wang, Q.; Tang, C. Role of TG2-mediated SERCA2 serotonylation on hypoxic pulmonary vein remodeling. Front. Pharmacol. 2019, 10, 1611. [Google Scholar] [CrossRef] [PubMed]
  168. Pani, B.; Cornatzer, E.; Cornatzer, W.; Shin, D.M.; Pittelkow, M.R.; Hovnanian, A.; Ambudkar, I.S.; Singh, B.B. Up-regulation of transient receptor potential canonical 1 (TRPC1) following sarco(endo)plasmic reticulum Ca2+ ATPase 2 gene silencing promotes cell survival: A potential role for TRPC1 in Darier’s disease. Mol. Biol. Cell 2006, 17, 4446–4458. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Ca2+ links Ca2+ regulatory proteins with mechanosensitive ion channels to regulate cardiovascular health and diseases. The mechanical stress generated in either cardiovascular health or disease directly stimulates mechanosensitive ion channels and induces Ca2+ entry. Then the Ca2+ regulatory proteins sensing the changes of Ca2+ cross-talk to or interact with mechanosensitive ion channels, which contributes to the development of cardiovascular disease or health.
Figure 1. Ca2+ links Ca2+ regulatory proteins with mechanosensitive ion channels to regulate cardiovascular health and diseases. The mechanical stress generated in either cardiovascular health or disease directly stimulates mechanosensitive ion channels and induces Ca2+ entry. Then the Ca2+ regulatory proteins sensing the changes of Ca2+ cross-talk to or interact with mechanosensitive ion channels, which contributes to the development of cardiovascular disease or health.
Ijms 22 08782 g001
Table 1. Current clinical trials for drugs targeting these Ca2+ regulatory proteins in the cardiovascular system according to the ClinicalTrials.gov website (available online and accessed on 12 August 2021).
Table 1. Current clinical trials for drugs targeting these Ca2+ regulatory proteins in the cardiovascular system according to the ClinicalTrials.gov website (available online and accessed on 12 August 2021).
Ca2+ Regulatory ProteinsTreatmentCardiovascular DiseasePhase
SERCAAAV1/SERCA2a (MYDICAR) [92]Ischemic cardiomyopathy; non-ischemic cardiomyopathy; heart failure; cardiomyopathiesPhase 2
MYDICAR-single intracoronary infusion [93]Heart failure, congestive; ischemic cardiomyopathy; non-ischemic cardiomyopathyPhase 2
MYDICAR [94]Chronic heart failurePhase 2
SRD-001 [95]Congestive and systolic heart failurePhase 1/Phase 2
Istaroxime [96]Heart failure [97]Early phase 1
NCXMYDICAR [98]Chronic heart failurePhase 2
OraiNo resourceNo resourceNo resource
STIMNo resourceNo resourceNo resource
IP3RNo resourceNo resourceNo resource
Table 2. Current clinical trials for drug specifically targeting these mechanosensitive ion channels are run in the cardiovascular system, according to the ClinicalTrials.gov website (available online and accessed on 12 August 2021).
Table 2. Current clinical trials for drug specifically targeting these mechanosensitive ion channels are run in the cardiovascular system, according to the ClinicalTrials.gov website (available online and accessed on 12 August 2021).
Mechanosensitive Ion ChannelsTreatmentCardiovascular DiseasePhase
PiezoSenicapoc (synonyms: ICA-17043; 2,2-bis-(4-fluorophenyl)-2-phenylacetamide) [129]Dehydrated hereditary stomatocytosis (related to pulmonary hypertension [130])Phase 1/Phase 2
TRPV1Capsaicin and mechanical stimulation with Von Frey filaments [131]Cystinosis (related to portal hypertension [132])Early phase 1
TRPV4GSK2798745 [133]Heart failurePhase 2
TRPM8Menthol [134]Hypertension [135]; PrehypertensionPhase 2/Phase 3
Table 3. Summary of the cross-talk between mechanosensitive ion channels and Ca2+ regulatory proteins in cardiovascular health and disease.
Table 3. Summary of the cross-talk between mechanosensitive ion channels and Ca2+ regulatory proteins in cardiovascular health and disease.
Mechanosensitive Ion ChannelsCa2+ Regulatory ProteinsCardiovascular HealthCardiovascular Disease
PiezoSERCAInhibition of piezo-dependent endothelial cell migration [18]No resource
TRPNCXCell contraction, proliferation and migration [136]; vasoconstriction [119,137]Pulmonary hypertension [136,137]; arrhythmia [138]
OraiParticipation in SOCE [139,140]Pulmonary hypertension [140]; thrombosis [141]
STIMSOCE activation [142]Hypertension and atherosclerosis [143,144,145]; thrombosis [141,146]
IP3RVasoconstriction; regulating the VGCC function [147,148]Hypertension [148]
SERCAHeartbeat and heart development [149]Cardiac hypertrophy [150]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, Y.; Shi, J.; Tong, X. Cross-Talk between Mechanosensitive Ion Channels and Calcium Regulatory Proteins in Cardiovascular Health and Disease. Int. J. Mol. Sci. 2021, 22, 8782. https://doi.org/10.3390/ijms22168782

AMA Style

Wang Y, Shi J, Tong X. Cross-Talk between Mechanosensitive Ion Channels and Calcium Regulatory Proteins in Cardiovascular Health and Disease. International Journal of Molecular Sciences. 2021; 22(16):8782. https://doi.org/10.3390/ijms22168782

Chicago/Turabian Style

Wang, Yaping, Jian Shi, and Xiaoyong Tong. 2021. "Cross-Talk between Mechanosensitive Ion Channels and Calcium Regulatory Proteins in Cardiovascular Health and Disease" International Journal of Molecular Sciences 22, no. 16: 8782. https://doi.org/10.3390/ijms22168782

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop