Next Article in Journal
Transcription Factors Active in the Anterior Blastema of Schmidtea mediterranea
Next Article in Special Issue
Channelopathy Genes in Pulmonary Arterial Hypertension
Previous Article in Journal
Quantitative Trait Locus Mapping Combined with RNA Sequencing Reveals the Molecular Basis of Seed Germination in Oilseed Rape
Previous Article in Special Issue
Revisiting the Large-Conductance Calcium-Activated Potassium (BKCa) Channels in the Pulmonary Circulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Role of Store-Operated Ca2+ Entry in the Pulmonary Vascular Remodeling Occurring in Pulmonary Arterial Hypertension

1
Faculté de Médecine, School of Medicine, Université Paris-Saclay, 94276 Le Kremlin-Bicêtre, France
2
INSERM UMR_S 999 Pulmonary Hypertension: Pathophysiology and Novel Therapies, Groupe Hospitalier Paris Saint-Joseph, Hôpital Marie Lannelongue, 92350 Le Plessis-Robinson, France
3
Assistance Publique—Hôpitaux de Paris (AP-HP), Department of Respiratory and Intensive Care Medicine, Pulmonary Hypertension National Referral Center, Hôpital Bicêtre, 94276 Le Kremlin-Bicêtre, France
4
Research and Innovation Unit, Groupe Hospitalier Paris Saint-Joseph, Hôpital Marie Lannelongue, 92350 Le Plessis-Robinson, France
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(12), 1781; https://doi.org/10.3390/biom11121781
Submission received: 5 November 2021 / Revised: 19 November 2021 / Accepted: 23 November 2021 / Published: 27 November 2021

Abstract

:
Pulmonary arterial hypertension (PAH) is a severe and multifactorial disease. PAH pathogenesis mostly involves pulmonary arterial endothelial and pulmonary arterial smooth muscle cell (PASMC) dysfunction, leading to alterations in pulmonary arterial tone and distal pulmonary vessel obstruction and remodeling. Unfortunately, current PAH therapies are not curative, and therapeutic approaches mostly target endothelial dysfunction, while PASMC dysfunction is under investigation. In PAH, modifications in intracellular Ca2+ homoeostasis could partly explain PASMC dysfunction. One of the most crucial actors regulating Ca2+ homeostasis is store-operated Ca2+ channels, which mediate store-operated Ca2+ entry (SOCE). This review focuses on the main actors of SOCE in human and experimental PASMC, their contribution to PAH pathogenesis, and their therapeutic potential in PAH.

1. Overview of SOCE

In mammals, the modulation of intracellular Ca2+ concentration ([Ca2+]i) constitutes a signal transduction mechanism for all cell types, regulating a large range of cellular functions, including contraction, proliferation, migration, gene transcription, metabolism, death, and apoptosis [1,2,3]. Intracellular Ca2+ signals are generated by Ca2+ release from the internal stores and/or by Ca2+ entry from the extracellular compartment through Ca2+ channels.
One of the major Ca2+ channels located in the plasma membrane (PM) is store-operated Ca2+ channels (SOCs), which produce store-operated Ca2+ entry (SOCE). The principle of SOCE was first described by Putney in 1986 [4] in non-excitable cells, where he found that extracellular Ca2+ influx was coupled to endoplasmic reticulum (ER) Ca2+ depletion, and then SOCE could be pumped into the ER by sarcoplasmic/endoplasmic reticulum Ca2+−ATPase (SERCA) to restore ER Ca2+ homoeostasis. A few years later, in mast cells, Hoth and Penner provided proof that SOCE is partly mediated by Ca2+ currents called Ca2+ release-activated Ca2+ (CRAC) currents (ICRAC) [5]. ICRAC is a non-voltage-activated Ca2+ inward current with strong inward rectification, a positive reverse potential, and a high selectivity for Ca2+ ions [5]. ICRAC is regulated through negative feedback that prevents excessive Ca2+ influx called Ca2+-dependent inactivation (CDI) [5,6]. CDI is composed of fast CDI and slow CDI.
Over the past several years, evidence from a large variety of cell types has demonstrated the existence of SOCE, and its involvement in cellular processes makes it a potential target for many diseases [7]. For two decades, the molecular identity of SOCs has been unknown. However, the advent of molecular biology techniques has made it possible to identify the molecules responsible for SOCE. In 2005–2006, two major ubiquitous SOCE actors were identified: the stromal interacting molecule (STIM) [8,9] and the Ca2+ channel Orai activated by STIM [10]. Additionally, in several cell types, in addition to Orai Ca2+ channel activation, STIM molecules activate transient receptor potential canonical channels (TRPCs), which are nonselective cation channels permeable to Ca2+.

1.1. STIM1

The link between ER Ca2+ depletion and Ca2+ entry was resolved with the discovery of STIM1 [8,9]. STIM1, a single pass transmembrane protein predominantly located in the ER, is able to sense luminal [Ca2+] and activate Orai and TRPC channels.
STIM1 is constituted by a luminal N-terminal part, which notably contains a Ca2+-binding EF hand domain that is able to sense the variation in the ER Ca2+ concentration ([Ca2+]ER) [11,12,13,14] (Figure 1). This EF hand domain forms with the sterile alpha motif (SAM) of the EF–SAM domain, which oligomerizes after ER Ca2+ depletion [12,15,16].
This luminal part is followed by a transmembrane segment (TM), potentially involved in the conformational change in STIM upon ER Ca2+ depletion [17]. The cytosolic part of STIM1 is composed of three coiled-coil parts (CC1, CC2, and CC3) (Figure 1). The structural interplay between these three CC domains is necessary for the elongation of STIM1 during its activation [18,19,20]. CC2 and CC3 domains are part of the STIM1–Orai1-activating region (SOAR) [21,22,23], also called CAD for CRAC-activating domain, and are necessary for coupling with Orai1 protein and for Orai1 activation [18,24]. SOAR is followed by a positively charged polylysine domain, which allows STIM1–TRPC electrostatic interaction [25]. The COOH cytosolic part ends in a polybasic domain (PBD), a domain required for the accumulation of STIM1 into puncta at the ER–PM junction by interaction of PBD with negatively charged phospholipids in the PM [22,26,27]. The STIM family is also composed of STIM2, which has high sequence homology with STIM1 with some differences at the N- and C-terminal parts [11]. STIM2 is more sensitive to [Ca2+]ER changes due to its lower affinity for Ca2+ and needs a lower level of store depletion than STIM1 for its activation [28,29]. Here, we only described the mechanism of action of SOCE by STIM1 because it is better characterized than STIM2.
Over the years, few additional variants have been found for STIM1 and STIM2. STIM1 long (STIM1L) was described as the product of alternative splicing of exon 11. STIM1L forms permanent clusters that colocalize with Orai1 through an interaction between STIM1L and actin fibers via the additional actin-binding domain (not present in classical STIM1). Most likely, due to its preactivated form STIM1L, it has faster activation than STIM1 associated with rapid activation of SOCE, as demonstrated in skeletal muscle cells [30]. A neuron-specific splice variant called STIM1B was also found, which is specifically localized in the presynaptic ER, explained by a targeting motif added through splice insertion in the cytosolic region of STIM1. STIM1B slowly activates SOCE and reduces Ca2+-dependent inactivation of ICRAC [31]. For STIM2 variants, there are STIM2.1 (or STIM2β) and STIM2.3, considering that the classical STIM2 (known to activate SOCE) is called STIM2.2 or STIM2α. STIM2.1 is known to negatively regulate SOCE, probably through a sequence-specific allosteric interaction with Orai1 [32,33].

1.2. Orai Channels

In 2006, Feske et al. described the absence of ICRAC on T cells isolated from patients with severe combined immune deficiency (SCID) syndrome [34]. This loss of ICRAC was independent of Ca2+ stores or the expression of several genes implicated in the control of Ca2+ entry in lymphocytes (KCNA3/Kv1.3, KCNN4/IKCa1, TRPC1, TRPC3, TRPV6, STIM1). By sequencing genomic DNA from SCID patients, they identified loss-of-function mutations in ORAI1, identifying Orai1 as an essential archetype of the CRAC current [10]. At the same time, by genome-wide RNA interference screening in Drosophila S2 cells, Zhang et al. confirmed that olf186-F (named Orai1) was the main channel of Icrac [35], which was subsequently confirmed by further studies [36,37,38].
The Orai1 channel presents a hexameric structure, as defined by the crystal structure of Drosphilia melanogaster Orai [39]. This architecture is formed by six Orai1 subunits coming from three dimers (Figure 2). Each of them contains an assembly of four transmembrane α-helices (TM1-TM4). This arrangement is made around a central point: the ion pore, with the 6 M1 subunits lining the ionic pore. The N- and C-termini are in the cytosol and contain the binding sites for the STIM protein, notably with the extended transmembrane Orai1 N-terminal (ETON) region. The ETON region mediates the interaction with the SOAR of STIM1 [40].
The Orai family consists of three members: Orai1, Orai2, and Orai3. All three members are widely expressed in humans [41] and can constitute functional CRAC channels [14,42,43]. Co-expression of STIM1 and Orai2 or Orai3 induced Ca2+ currents similar to those of Orai1-induced ICRAC, with small differences in selectivity and inactivation. Regarding selectivity, monovalent cations, such as Na+, permeate better in Orai3 than in Orai1 and Orai2 [44]. Regarding Orai inactivation, Orai2, and especially Orai3, have higher levels of fast CDI, and Orai2 does not have slow CDI [44,45].
In humans, Orai1 has two isoforms: the longer-form Orai1α, of approximately 33 kDa, and the shorter-form Orai1β, of approximately 23 kDa, resulting from an alternative translation initiation site [46]. Both isoforms can support SOCE [47].

1.3. TRPC Channels

TRPC channels are part of the transient receptor potential (TRP) channels, a family of cationic nonselective channels (Ca2+, Na+, K+). This group of channels is subdivided into seven distinct subfamilies according to their sequence homology [48,49]: TRPC (canonical), TRPV (vanilloid), TRPM (melastin), TRPA (ankyrin), TRPP (polycystin), TRPML (mucolipin), and TRPN (drosophilia no mechanoreceptor potential C (NOMPC)). Except for TRPM4 and TRPM5, all TRP channels are cationic channels permeable to Ca2+ [50]. In the present review, we only focused on TRPC channels because evidence from the literature indicated that TRPC channels are involved in SOCE.
The TRPC family can be divided into four subgroups: TRPC1; TRPC3, -C6, -C7; TRPC4, -C5; and TRPC2, which is a pseudogene in humans [51]. TRPC channels are mostly Ca2+- and Na+-selective, except TRPC2, which is cation-nonselective [52].
Structurally, TRPC channels are relatively analogous [53,54]. They can form heterotetramers or homotetramers. Each TRPC channel consists of at least six transmembrane domains with an ionic pore formed between TM5 and TM6 (Figure 3). On the N-terminal cytosolic domain, there are four to five ankyrin repeats, leading to protein–protein interactions [55,56]. On the C-terminal side, the TRP domain, a well-conserved region in all TRPCs, is essential for channel gating [54], followed by the calmodulin (CaM)/inositol 1,4,5-trisphosphate receptor-binding (CIRB) site, with CaM binding in a Ca2+-dependent manner [51,57,58]. Interestingly, incorrect trafficking of TRPC3 was observed when the CIRB region was deleted [59].
TRPC channels are reported to contribute to both SOCE and receptor-activated Ca2+ entry pathways (ROCE). As described below, several studies have demonstrated TRPC channel activation following ER Ca2+ depletion.
TRPC1 was first reported to have a role in SOCE from studies using the expression of the splice variant of TRPC1, so-called TRPC1A, in heterologous cell system expression [60] or the expression of a full-length cDNA of human TRPC1 in COS (fibroblast-like cell lines derived from monkey kidney tissue) cells [61]. The role of TRPC1 was confirmed through studies using different approaches (RNA interference, dominant-negative overexpression) in human submandibular gland cell lines [62], endothelial cells [63], human myoblasts and myotubes [64,65,66], and platelets [67]. Further evidence was provided by studies on trpc1−/− mice that showed reduced SOCE in different tissues (salivar gland, pancreatic acinar cells, and aortic endothelial cells) [68,69,70].
TRPC3 was found to be activated by depletion of Ca2+ stores in DT40 chicken B lymphocytes [71]. Selective suppression of TRPC3 protein by small interfering RNA (siRNA) showed TRPC3-mediated SOCE in H19-7 hippocampal neuronal cells [72] and human umbilical vein endothelial cells [73]. When TRPC3 is expressed at low levels in HEK293 cells, the TRPC3 current is inhibited in cells treated with siRNA against STIM1, showing the regulation of TRPC3 by STIM1 [74]. Salivary glands and pancreatic cells isolated from trpc3−/− mice and cells treated with a TRPC3 inhibitor exhibit reduced SOCE [75,76]. The regulatory mechanism of TRPC3 appears to be determined by its expression level [71].
TRPC4 has been shown to be activated by Ca2+ store depletion in HEK293 cells [74] and Xenopus laevis oocytes [77]. Other evidence has been obtained notably by siRNA strategy in human myoblasts and myotubes [64,65], human adrenal cells [78], human corneal endothelial cells [79], human gingival keratinocytes [80], neonatal rat ventricular cardiomyocytes [81] or human and mouse endothelial cells [82]. TRPC4 implication in SOCE has also been shown in neonatal rat ventricular cardiomyocytes by a dominant-negative strategy [81]. In addition, aortic and lung endothelial cells from trpc4−/− mice showed reduced SOCE [83,84].
Few studies report the implication of TRPC5 in SOCE, although it was found in smooth muscle cells isolated from rabbit pial arterioles [85] and in neonatal rat ventricular myocytes [81,86].
Regarding TRPC6, several studies have shown that it is also able to drive SOCE. This was shown by siRNA in a leukemic granulocyte cell line [87], antibody inhibition in human platelets [88,89,90], or overexpression and knockdown by short hairpin RNA (shRNA) in hepatoma cells [91]. On the other hand, in HEK293 cells co-transfected with STIM1 and TRPC6, STIM1 does not influence the TRPC6-induced current [92], except when TRPC6 is expressed at low levels [74].
TRPC7 is not regulated by STIM1 or Orai1, and it was proposed that TRPC7 is activated by a mechanism independent of store depletion [93].
It is also known that some TRPC channels can be directly activated by the phospholipase C (PLC) pathway, notably with the PI(4,5)P2 degradation product diacylglycerol (DAG), by the G protein coupled receptor, protein kinase C (PKC) pathway, or even membrane stretching [73,94,95], resulting in receptor-operated Ca2+ entry (ROCE) instead of SOCE.
However, due to their electrophysiological characteristics (Table 1), TRPC channels are not able to produce ICRAC. Indeed, TRPC channels form nonselective Ca2+-permeable channels that do not exhibit the electrophysiological identity of ICRAC (Figure 4A). However, almost all TRPC channels are activated in response to ER–SR Ca2+ depletion, with the formation of a nonselective current with a reverse potential close to 0 mV, called ISOC (Figure 4B).

1.4. SOCE Mechanism

The activation mechanisms leading to SOCE at the PM by STIM proteins are now better understood: one of the most physiological mechanisms inducing ER–SR Ca2+ store depletion is the activation of the PLC via G-protein coupled receptors, with inositol 1,4,5-triphosphate (IP3) generation and IP3 receptor (IP3R) activation [107,108]. In cells with furnished Ca2+ stores, STIM1 maintains an inactivated conformation and localizes throughout the ER–SR membrane. In this conformation, STIM1 is able to bind Ca2+ with his EF hand–SAM domain. This inactivated conformation of STIM1 prevents spontaneous activation since the CC1 domain is associated with CAD–SOAR [17,109,110].
After ER–SR Ca2+ depletion, the luminal side of STIM1 rearranges by oligomerization of the EF–SAM domain [12,15,16] (Figure 5). This association induces the modification of the TM domain, leading to the liberation of CAD–SOAR [20]. This conformational change leads to rapid oligomerization of STIM1 through the CC3 domain and, thus, translocation to the ER–PM junction [18,26,111]. After this oligomerization and translocation, STIM1 forms puncta in this region where Orai is also located. It was shown that one to two STIM1 channels could trap Orai1, but eight STIM1 channels for each Orai1 channel were required to optimally activate these channels [112]. STIM1 interacts and activates with the ETON domain of Orai1 via CAD–SOAR97. The Ca2+ stores of the ER–SR will then fill up through the activity of the SERCA pump located on the ER–SR membrane. Icrac is then regulated through the CDI.
It has been shown that STIM1 can regulate all TRPC channels, except TRPC7 [74]. STIM1 can bind TRPC1, -C4, and -C5 [74,113] via electrostatic interactions that exist between two positively charged lysins (684KK685) on STIM1 and two negatively charged conserved aspartates (639DD640) on TRPC1 [25]. Such electrostatic interactions also seem to occur with TRPC4 and TRPC5 [114]. Alternatively, STIM1 can regulate TRPC3 and TRPC6 indirectly through heteromultimerisation of TRPC1 with TRPC3 and TRPC4 with TRPC6 [74]. Furthermore, it has been demonstrated that local Ca2+ entry via Orai1 channels can regulate plasma membrane recruitment of TRPC1 [115]. TRPC1-containing vesicles can be detected in the region of the subplasma membrane in the immediate vicinity of the ER–PM region where Orai1 and STIM1 aggregate upon Ca2+ depletion. Recruitment of TRPC1 for clustering with STIM1 in the ER–PM junction occurs by trafficking via fast recycling endosomes [116]. Despite these studies and putative models, the precise mechanism of TRPC channel contributions in SOCE remains to be elucidated.
In addition to SOCE, activation of PLC-coupled receptors can activate the store-independent Ca2+ entry (SICE) pathway. The SICE pathway is performed through arachidonic acid-regulated Ca2+-selective (ARC) [83,117] and leukotriene C4 (LTC4)-regulated Ca2+ (LRC) [118]. These channels are composed of both Orai1 and Orai3 subunits [119,120] and can be regulated by a pool of STIM1 at the PM [119,121], notably through the interaction of STIM1 with Orai3 [122].

2. SOCE in PAH

2.1. Pulmonary Arterial Hypertension

Pulmonary arterial hypertension (PAH) is a progressive and devastating disorder due to a progressive obstruction of distal pulmonary arteries (PA) (<500 μm in diameter), resulting in an elevation of pulmonary arterial pressure and progressive right heart dysfunction [123]. Recently, PAH was proposed to be defined by a mean pulmonary arterial pressure (mPAP) higher than 20 mmHg (compared to a normal PAPm, a rest of 14 ± 3 mmHg), a pulmonary vascular resistance higher than 3 woods units and a pulmonary arterial wedge pressure lower than 15 mmHg [124,125].
PAH is clinically classified, not according to clinical manifestations but according to etiology. One of the most commonly diagnosed groups is idiopathic PAH (iPAH), which includes patients with no identified causative factor. PAH can also be heritable, induced by drugs and toxins, or associated with other diseases [125]. The classification also includes a subgroup for long-term responders to Ca2+ channel blockers, another one for PAH with overt features of venous capillary (pulmonary veno-occlusive disease or pulmonary capillary haemangiomatosis) involvement and a last one for persistent PH of the new-born syndrome [124].
PAH is characterized by an increase in pulmonary vascular resistance explained by endothelial dysfunction, excessive proliferation of pulmonary arterial smooth muscle cell (PASMC) and endothelial cell (PAEC), PA vasoconstriction, and in situ thrombosis [123]. PAEC dysfunction in PAH is mostly targeted by the existing three current therapies. The three main treatments target the prostacyclin pathway with prostacyclin analogues, the endothelin-1 (ET-1) pathway with ET-1 receptor antagonist and the nitric oxide (NO) pathway with PDE5 inhibitors or guanylate cyclase stimulators [126]. However, these treatments cannot cure PAH, and lung transplantation is still necessary for the most severe patients. As the prognosis remains poor with a 5-year survival rate of approximately 60%, there is a need for innovative treatments to target pulmonary vascular remodeling [127,128].
SOCs are ubiquitous, and their role in PAEC dysfunction and right ventricular (RV) failure occurring in PAH should be studied with interest. Indeed, SOCs contribute to systemic PAEC proliferation and migration [73,129] and right ventricular Ca2+ handling remodeling in experimental PH models [130,131].
Moreover, PASMC dysfunction is not directly targeted by current PAH therapies. Therefore, we focused this review on the role of SOCE in PASMC dysfunction. PASMC dysfunction is characterized by excessive proliferation, apoptosis resistance, abnormal migration, and vasoconstriction [123,132], which strongly contribute to pulmonary artery remodeling. Since an increase in PASMC [Ca2+]i is a key trigger of these biological processes, resulting in pulmonary vascular remodeling. SOC modification could act on pulmonary remodeling in PAH and, thus, emerge as a potential additional therapeutic target in PAH.
The understanding of PAH pathogenesis mainly emanates from an experimental model of PAH. Although they do not perfectly reproduce all the clinical parameters of PAH, they are considered PH animal models, mostly performed in rodents, with three main experimental procedures: monocrotaline (MCT), chronic hypoxia (CH), and Sugen 5416 hypoxia (SU/Hx).

2.2. Physiological Implication of SOCE in Control PASMC

The expression of Orai1/2/3 and STIM1/2 was detected in mouse, rat, and human PASMC (hPASMC) [133,134,135]. In hPASMC, Orai1 and STIM1/2 promotes SOCE [136,137]. In rodent PASMC, Orai1/2/3 and STIM1/2 promote SOCE [134,135,138,139,140]. Only TRPC1, -C3, -C4, and -C6 were expressed at the mRNA and protein levels in primary hPASMC cultures [141,142,143,144,145,146], while TRPC1, -C3, -C4, -C5, and -C6 were expressed in rodent PASMC (Table 2).
TRPC1 is overexpressed in proliferative control hPASMC in association with increased SOCE, indicating that TRPC1 contributes to SOCE in control hPASMC [143,157]. In systemic vascular smooth muscle cell (VSMC), caveolin-1 knockdown by siRNA induces a decrease in TRPC1 expression accompanied by a reduction in SOCE [158]. Furthermore, in mouse PASMC, STIM1 and TRPC1 are functionally coupled and mediate SOCE [159].
The inhibition of TRPC1 by a specific antibody or by the nonselective SOCE inhibitor 2-aminoethyldiphenylbonirate (2-APB) significantly reduced ex vivo neointimal growth in human veins as well as SOCE and proliferation of cultured PASMC. These results highlight the involvement of TRPC1 in vascular remodeling occurring in arteriosclerosis and the protective potential of TRPC1 inhibitors against vascular disease [160].
STIM1 mediates the proliferation of primary human coronary artery SMC [161]. Moreover, the knockdown of STIM1 or Orai1 reduced SOCE and the migration of human saphenous VSMC. Alternatively, the overexpression of the negative form of TRPC5 reduced cell proliferation and cell migration [162]. In addition, an Orai1 blocker, Synta66 (S66), reduced PDGF-evoked human saphenous VSMC migration without a significant change in cell proliferation or viability.
In cultured rat aortic VSMC, Orai1 and STIM1 are overexpressed compared to freshly isolated VSMC, suggesting that these proteins enhance the VSMC proliferative phenotype [163]. STIM1, Orai1, and TRPC1 expression/function are crucial for angiotensin II- or urotensin II-induced rat VSMC proliferation [164,165].
Moreover, PASMCs have two phenotypes: a contractile phenotype, accessible in freshly isolated cells, and a proliferative and migratory phenotype, generally obtained in culture or pathological status. Treatment of rat proliferative PASMC with factors (transforming growth factor β (TGF-β) and heparin), favoring the differentiation of PASMC into a contractile phenotype, enhanced SOCE. Indeed, TGF-β decreased Orai1/2, STIM2, and TRPC6 expression, and heparin decreased Orai1/2/3, STIM1/2, and TRPC6 expression [135].
In rat aortic SMC, STIM1 or Orai1 knockdown reduces SOCE and migration, while the knockdown of Orai2 or STIM2 has no effect on SOCE, proliferation, and migration [163]. In cultured rat aortic SMC, STIM1–Orai1-mediated Ca2+ influx plays a key role in platelet-derived growth factor (PDGF)-induced migration, since siRNA knockdown inhibits PDGF-induced migration [166]. Moreover, enhancement of Orai1 function with IA65, a pharmacological activator, results in an increase in rat VSMC migration [167].
These results indicate that Orai1-induced SOCE contributes significantly to in vitro systemic VSMC migration and proliferation. We hypothesize that the SOCE archetype Orai1 could contribute to aberrant PASMC migration and proliferation phenotypes in PAH.
In rat coronary arteries, silencing of Orai1 or STIM1 prevents urotensin-II-induced vasoconstriction [165]. In mouse aortae, VSMC-specific knockout of the stim1 gene impaired SOCE and reduced the phenylephrine-induced vasoconstrictive response [168]. Moreover, the overexpression of TRPC1 enhances rat PA vasoconstriction [142]. These results indicate that SOCE could contribute to VSMC contraction. Similar to systemic vessels, we hypothesize that SOCs contribute to PA contraction as well as PA vasoconstriction in PAH. However, the contribution of SOCs to the regulation of pulmonary arterial tone has yet to be understudied under physiological conditions and in the physiopathological condition of PAH. In systemic VSMC, SOCE seems to be constituted by STIM–Orai and the TRPC channel complex [169,170]. In addition to their role in SOCE, the Orai–TRPC complex has been demonstrated to also regulate Ca2+-sensitive K+ channels (KCa), contributing indirectly to the regulation of resting membrane potential. In isolated rat aortic SMC, TRPC1 forms a complex with the large conductance KCa (BKCa) that promotes Ca2+ entry through and then membrane hyperpolarization. This membrane hyperpolarization mediated by TRPC1–BKCa coupling could prevent excessive contraction of VSMC by reducing agonist-induced membrane depolarization [171].

2.3. Physiopathological Implication of SOCE in PAH PASMC

Several studies found an increase in SOCE in hPASMC isolated from iPAH patients, resulting in an increase in [Ca2+]i [157,172]. Control hPASMC exposed to hypoxia (3% O2 for 72 h) saw an increase in [Ca2+]i associated with an increase in Orai1/2, STIM1/2, and TRPC6 protein levels [173]. The knockdown of HIF-1α in rat PASMC or HIF-1α deficiency in mice (hif-1α−/−) reduced the overexpression of Orai2 induced by hypoxia exposure (10% O2, 21 days for mice or 4% O2, 60 h for rat PASMC) [134]. As the increased expression of HIF-1α plays a key role in the pathogenesis of PAH [174,175,176], we hypothesized that HIF-1α overexpression in PAH promotes an increase in SOCE in iPAH PASMC.
An increase in STIM2 protein levels has been demonstrated in iPAH hPASMC compared with control PASMCs, while the STIM1 protein level remained unchanged. Using a siRNA strategy, they also found that the knockdown of STIM2 reduces SOCE and proliferation of iPAH hPASMC without any effect on control hPASMC [136]. In an additional study, they demonstrated that the transition of rat PASMCs from a contractile to proliferative phenotype was associated with enhanced SOCE due to increased expression of Orai2, STIM2, and TRPC6 [135]. Moreover, they demonstrated that nicotinamide phosphoribosyl transferase promotes pulmonary vascular remodeling in PAH and experimental PH, partly due to an overexpression of Orai2 and STIM2 protein levels leading to an increase in SOCE and in hPASMC proliferation [177].
SOCE increased in rat distal PASMCs exposed ex vivo to acute hypoxia (4% O2) compared to rat proximal PASMC, in association with an increased expression of TRPC1 and STIM1 [178]. Another study also found increased expression of STIM1 in the distal PA of experimental PH in rats exposed to hypobaric chronic hypoxia CH (for 21 days) [140]. Additionally, in rat PASMC, it was demonstrated that the knockdown of STIM1 inhibited hypoxia-induced nuclear factor of activated T-cells (NFAT) cytoplasmic 3 (NFATc3) nuclear translocation and, thus, excessive PASMC proliferation [140]. In contrast to these studies, STIM1 expression was unchanged in rat distal PA and PASMC exposed to CH (10% O2 for 21 days) [134]. In this study, the authors found increased expression of Orai1/2, while Orai3 remained stable. It was also reported in mouse PASMC exposed to acute hypoxia (95% N2 and 5% CO2) that there was an increase in TRPC1, Orai1, and STIM1 interaction to mediate increased SOCE [151].
STIM1 and Orai1 expression and function were increased in aortas from spontaneously hypertensive rats [179]. In an experimental obstructive sleep apnea model induced in rats by chronic intermittent hypoxia (CIH) (5% inspired O2 for 20 s, followed by 280 s of room air, 12 times per hour for 8 h a day for 14 to 28 days), Castillo-Galán et al. showed the increased protein expression of Orai1, STIM1, and TRPC1, -C4, -C6 in the lungs of CIH rats [180]. Treatment of these rats with 2-APB decreased right ventricular systolic pressure (RVSP) and pulmonary vessel remodeling, which increased following CIH exposure [181].
In control hPASMC, platelet-derived growth factor (PDGF) application stimulated STIM1 and Orai1 overexpression through the AKT/mTOR signaling pathway, which led to an increase in SOCE [137]. As it is well recognized that PDGF is overproduced in PAH [182,183,184], this overproduction could explain why SOCE increased in iPAH hPASMC.
In experimental PH induced by CH exposure in mice, trpc1−/− + trpc6−/− double knockout protected mice against PH. In this experimental PH model, the deletion of trpc1 and trpc6 prevented pulmonary vessel remodeling after CH [152], demonstrating that TRPC1 and TRPC6 are essential for the development of PH under CH exposure in mice. In addition, Sun et al. showed that, in a CH-induced PH murine model (11% O2 for 28 days), intratracheal in vivo administration of siRNA against TRPC1 attenuates pulmonary and RV arterial remodeling, with a significant reduction in right ventricular systolic pressure [185]. In vivo TRPC1-siRNA delivery also counteracts the increase in inflammatory biomarkers (TNF-α and MMP-9) in the lung and apoptotic biomarkers in the RV myocardium that was observed in a murine model of CH-induced PAH.
In PH rats induced by CH exposure, TRPC1 and -C6 expression was increased in isolated PA, which was accompanied by an increase in SOCE. In addition, nonselective pharmacological inhibition of SOCE by lanthanum (La3+) and SKF-96365 normalizes elevated [Ca2+]i in PASMC and vascular tone in PAs of CH–PH rats [145].
Another study found that TRPC1 mRNA levels were increased in murine PASMC under hypoxia exposure (1% O2 for 72 h). In this study, hypoxia exposure enhanced the proliferation of PASMC, which was reduced by the knockdown of TRPC1 (siRNA approach). The proliferation of PASMC isolated from trpc1−/− mice was also reduced compared to the proliferation of PASMC isolated from trpc1+/+ mice. Furthermore, trpc1−/− mice developed partial protection against PH in response to CH compared to trpc1+/+ mice exposed to CH. Both trpc1+/+ and trpc1−/− mice exposed to CH present similar RV hypertrophy, while pulmonary vascular remodeling is attenuated in trpc1−/− mice [153].
Increased TRPC1 and TRPC4 mRNA expression and protein levels in PAs isolated from MCT-PH rats were also found to be associated with increased SOCE and PA vasoconstriction [186].
Additionally, an increase in the expression of TRPC3 was described in iPAH hPASMC, which could partly explain the exacerbated SOCE in iPAH hPASMC173.
Using trpc4−/− rats exposed to the sugen–hypoxia protocol (single subcutaneous injection of 20 mg/kg SU5416 followed by 10% O2 for 3 weeks), Alzoubi et al. found a reduction in the occlusion of small PA and plexiform lesions in trpc4−/− rats compared with wild-type (WT) rats. However, trpc4−/− rats have similar hemodynamic parameters to WT rats, but TRPC4 deficiency provided a significant survival benefit, with preservation of cardiac output [187].
Regarding the role of TRPC6 in PAH, single nucleotide polymorphisms (SNPs) were found in the promoter of the TRPC6 gene in a few iPAH patients, facilitating the expression of the TRPC6 protein and increasing its function in iPAH [188,189]. Therefore, this SNP could predispose individuals who have this mutation to develop PAH.
TRPC6 expression (mRNA and protein) was found to be increased in lung and hPASMC from PAH patients, and the knockdown of TRPC6 by siRNA reduced the hyperproliferative phenotype of PAH hPASMC [146]. The role of TRPC6 was also predicted in PASMCs from Milan hypertensive strain (MHS) rats, a genetic model of systemic hypertension. Increased expression of TRPC6, SERCA2, and Na+/Ca2+ exchanger-1 (NCX1) was reported to likely contribute to the abnormal Ca2+ homoeostasis and subsequent vasoconstriction and increase blood pressure observed in MHS rats [190].
Recently, Jain et al. demonstrated that oral gavage with TRPC6 blocker BI-749237 reduced the development of PH induced by CH exposure in mice [191], suggesting that TRPC6 could be a potential target in PAH.
The majority of patients with heritable PAH have mutations in bone morphogenetic protein type 2 receptor (BMPR2) gene [192,193]. Reduction in BMPR-II signaling is a hallmark of PAH [194,195,196]. BMPR2 is activated by BMP ligands, including BMP2 and BMP4 [197]. BMP2 is protective against CH-induced PH, while BMP4 contributes to the development of PH [198]. BMP2 application in rat distal PASMCs induced a decrease in TRPC1, -C4, and -C6 expression (mRNA and protein levels), and therefore a reduction in SOCE, a reduction in basal [Ca2+]i, and a reduction in PASMC proliferation and migration [199]. Moreover, BMP4 application promotes pulmonary vascular remodeling in CH-PH mice, and bmp4+/− mice exposed to CH were protected against the development of PH [198]. In rat distal PASMC, in contrast to BMP2 application, BMP4 application increases the expression of TRPC1, -C4, and -C6 (mRNA and protein levels); increases SOCE and [Ca2+]i; enhances PASMC proliferation and migration; and protects PASMC mice from apoptosis [154,155].
The resistance to apoptosis of hPASMC is also a hallmark of PAH. Currently, there are no studies on PAH hPASMC linking apoptosis resistance to SOCE. However, several studies have shown the role of Orai in the regulation of apoptosis in other cell types and diseases. For example, reduced Orai1 and STIM1 activity increases apoptosis of pancreatic cancer cells, and the in vivo silencing of Orai1 in mice protects macrophages from apoptosis [200,201]. On the other hand, it was demonstrated in human lymphoma B cell lines that pharmacological inhibition or silencing of Orai1 leads to an enhancement of apoptosis [202]. Thus, the contribution of SOCE and SOCs to the regulation of apoptosis resistance in iPAH PASMC needs to be further investigated.
The involvement of SOCs in hypoxic pulmonary vasoconstriction was also documented in rats using nonselective pharmacological approaches [203,204]. The real contribution of SOCs to the regulation of pulmonary arterial tone needs to be investigated using more selective pharmacology or genetic tools. The physiopathological implications of SOCs in PAH PASMC are summarized in Table 3.
Since SOCE contribute to the regulation of SR Ca2+ content and since SOCE is increased in PAH hPASMC, we can hypothesize that SOCE-induced increase in SR Ca2+ content result in an increase in the activity SR Ca2+ channels such as IP3R and ryanodine receptors (RyR). VSMC including PASMC exhibit three different RyR isoforms (RyR1, RyR2, and RyR3) and three different IP3R isoforms (IP3R1, IP3R2, and IP3R3) which could contribute to the regulation of PASMC functions. It has been demonstrated that IP3R2−/− mice developed more severe PH than WT mice under CH-exposure [205]. Using isolated PASMC, Shibata et al. found that SOCE is strongly increased in IP3R2−/− PASMC in normoxic condition as well as in the CH condition. The authors demonstrated that IP3R2 negatively regulated SOCE in mouse PASMC, indicating the presence of IP3R2 to protect against the development of experimental PH by regulating SOCE [205]. The contribution of the IP3R channel has not been studied in the aberrant phenotypes of PAH PASMC (proliferation, resistance to apoptosis, migration, contraction). IP3R was previously found to be functionally coupled with TRPC by a direct binding with the N-terminus of the IP3R and the C-terminal CIRB motif of the TRPC subtypes [57,206]. This direct binding with IP3R and TRPC has been demonstrated for all IP3R isoforms and all TRPC subtypes [58,207]. However, the mechanism by which IP3R isoforms regulate of TRPC function has not been yet demonstrated in PASMC from PAH patients.
In ryr2−/−mice, it has been demonstrated that an RyR2-mediated Ca2+ release contributes to pathological ROS generation induced by CH exposure [208]. SMC-specific ryr2−/− mice have similar PAP in response to hypoxia compared to WT; however, RyR2 was suggested to have a role in the sustained phase of hypoxic pulmonary vasoconstriction [209]. In PASMC isolated from CH-PH and MCT-PH rats, important changes in the cellular localization of RyR isoforms and Ca2+ stores have been demonstrated, contributing to the pathogenesis of experimental PH [210]. The expression of TRPV4 and RyR2 proteins increased in isolated PA from CH-PH rats. In this study, the authors demonstrated that RyR2 regulated TRPV4 function in PASMC from CH-PH rats, contributing to excessive PA constriction [211].

3. Targeted SOCE in PAH: A Novel Therapeutic Option?

In PAH, the only ion channels targeted in the clinic are L-type voltage-gated Ca2+ channels. However, L-type voltage-gated Ca2+ channel blockers (nifedipine, diltiazem, or amlodipine) are only effective for less than 10% of PAH patients, named “the responder”, and defined by a fall of 10–40 mmHg in PAPm during inhalation of NO. Voltage-gated Ca2+ channel blockers act on the reduction in excessive pulmonary arterial vasoconstriction [212,213].
As SOCs constitute another important source of Ca2+ mostly involved in pulmonary artery remodeling, and L-type voltage-dependent Ca2+ channels are mainly involved in pulmonary artery vasoconstriction, targeting SOCs may be the next challenge to overcome PAH phenotypes. Indeed, PAH hPASMC are characterized by excessive proliferation, excessive migration, and vasoconstrictive phenotypes resulting from Ca2+ homoeostasis remodeling (particularly due to SOC activities).
Drugs acting directly and selectively on different SOCs could be attractive for improving PASMC remodeling in the context of PAH. As represented in Table 4, several compounds are available to differentially inhibit almost all SOCs.
Concerning inhibitors of Orai1, there is a well-known family of pyrazole derivatives, with the widely used BTP2 (YM-58483) [214,215,216,217,218] and GSK compounds (GSK7975A and GSK-5503A) [220], acting on the selectivity filter of Orai1, Synta-66, which binds directly to Orai1 [221,222,223,224] and its suitable in vivo analogue JPIII [225]. The lanthanide family (Gd3+ or La3+) has been known as an Orai blocker for many years [219,226]. The less-used AnCoA4 [227] and 5J4 [228] also exist. ML-9 is used to inhibit STIM1 puncta formation [229]. Since 2006 and the identification of Orai1 as the archetype of SOCE, Orai1 has been a potential target in several pathological processes, including psoriasis, pancreatitis, asthma, and COVID-19 pneumonia.
Auxora (or CM2489), an Orai1 channel inhibitor, was tested in a clinical trial to treat moderate to severe plaque psoriasis. However, it showed only a limited clinical effect but was safe and well tolerated by patients. The compound CM2489 had some chemical improvements and design to treat acute pancreatitis. The preclinical trials were successful in mice, and the phase I clinical trials were successful, allowing entry into phase II with a good safety profile and improvement in patient outcomes [245,246]. In addition, a clinical study, although limited by its design, has been conducted for the treatment of severe or critical COVID-19 pneumonia with this compound [247]. Another Orai1 inhibitor (RP3128) has reached clinical trials (Phase I/IIa) for asthma [248]. Finally, a last candidate, the Orai1 inhibitor PRCL-02, has achieved a phase IIa clinical trial for the treatment of moderate to severe chronic plaque psoriasis [249].
Concerning TRPC channel inhibitors used in experimental research, very few of them are selective because of the structural similarity between all TRPC channels. Xanthine-based inhibitors are known to inhibit TRPC1, -C5 and -C6 at nanomolar concentrations [232]. Among the xanthine-based inhibitors, HC-608 (also called Pico145), described to be one of the most potent inhibitors of TRPC1, -C4, and -C5 [234], and HC-070, is widely used to inhibit TRPC4 and -C5 [233]. The benzamidiole ML-204 is used for TRPC4 [231] and AC1903 for TRPC5 [235], but they both also inhibit TRPC4 and -C5. Alternatively, GFB-8438, produced from pyridazinone 1 (Pyr1), is a potent TRPC4 and TRPC4/C5 inhibitor [236,237]. Pyr3 and Pyr10 are known to inhibit TRPC3 [230]. Another sub micromolar potent inhibitor of TRPC3 and -C6 is GSK2833503, which is known to be more than 100-fold more selective against TRPC3 and TRPC6 than other Ca2+ channels [240,241]. The BI 749327 inhibitor was developed to specifically inhibit TRPC6. It is an orally bioavailable inhibitor that does not inhibit TRPC5 or other TRP channels [242]. The (+)-larixol derivative SH045 is also used to inhibit TRPC6, with good selectivity [243]. TRPC6 can also be inhibited by a nanomolar inhibitor and orally bioavailable SAR7334 [238] or by its analogue AM-1473 [244].
Some studies are ongoing, such as the TRPC5 channel inhibitor GFB887, tested in Phase I on healthy volunteers, and now in Phase II for patients with diabetic nephropathy or with focal segmental glomerulosclerosis [250]. However, research on TRPC inhibitors is at an early stage but should be accelerated by recent cryo-EM structure discovery.
As described in this part, the selectivity of these molecules is not enough to discriminate each of the TRPC isoforms. In vitro, these inhibitors may be used at specific concentrations to ensure selectivity, but in vivo the use of these molecules is challenging due to their lack of selectivity. Additionally, the mechanism of action or channel specificity remains unclear for some of these pharmacological agents. Moreover, some compounds are not stable in vivo, which prevents their use in preclinical or clinical research. Progress in the identification of SOCs involved in the pathogenesis of PAH is essential to develop specific therapeutic tools. The recent advance in the determination of cryo-EM structures of TRPC channels with sufficient resolution to allow the identification of specific bounding site of small molecules [251,252] should help to develop more potent selective inhibitors or activators, since the need of selective molecules is crucial for the development of innovative therapy in various diseases, including PAH. Despite important progress in the selectivity of the molecules described in this review, the development of more selective pharmacological tools to selectively inhibit each SOC isoform could facilitate the emergence of innovative therapies concerning multiple diseases, including PAH.

4. Conclusions

PAH represents a major human and social burden because this orphan disease is associated with a poor prognosis, affects children and young adults, and has no cure treatment except lung transplantation for eligible patients. The data presented in the present review support evidence of the involvement of SOCs in the pathogenesis of PAH, and each of these channels could be considered a potential new therapeutic target for PAH. Nevertheless, these channels are ubiquitous. Consequently, specific pharmacological tools towards a target organ should be carefully considered in preclinical experiments to identify eventual side effects.

Author Contributions

Conceptualization, B.M., V.C. and F.A.; drawing figures, B.M. and F.A.; writing-original draft preparation, B.M., V.C. and F.A.; Writing—Review & Editing, B.M., V.C., M.H., D.M. and F.A. supervision, V.C. and F.A. All authors have read and agreed to the published version of the manuscript.

Funding

B.M. was supported by the Therapeutic Innovation Doctoral School (ED569).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

M.H. and D.M. have relationships with drug companies, including Actelion, Bayer, GSK, Novartis, and Pfizer. In addition to being investigators in trials involving these companies, other relationships include consultancy services and memberships to scientific advisory boards. The other authors have no conflict of interest to declare.

References

  1. Bootman, M.D.; Collins, T.J.; Peppiatt, C.M.; Prothero, L.S.; MacKenzie, L.; De Smet, P.; Travers, M.; Tovey, S.C.; Seo, J.T.; Berridge, M.J.; et al. Calcium Signalling—an Overview. Semin. Cell Dev. Biol. 2001, 12, 3–10. [Google Scholar] [CrossRef]
  2. Berridge, M.J.; Bootman, M.D.; Roderick, H.L. Calcium Signalling: Dynamics, Homeostasis and Remodelling. Nat. Rev. Mol. Cell Biol. 2003, 4, 517–529. [Google Scholar] [CrossRef] [Green Version]
  3. Clapham, D.E. Calcium Signaling. Cell 2007, 131, 1047–1058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Putney, J.W. A Model for Receptor-Regulated Calcium Entry. Cell Calcium 1986, 7, 1–12. [Google Scholar] [CrossRef]
  5. Hoth, M.; Penner, R. Depletion of Intracellular Calcium Stores Activates a Calcium Current in Mast Cells. Nature 1992, 355, 353–356. [Google Scholar] [CrossRef] [PubMed]
  6. Zweifach, A.; Lewis, R.S. Slow Calcium-Dependent Inactivation of Depletion-Activated Calcium Current. Store-Dependent and -Independent Mechanisms. J. Biol. Chem. 1995, 270, 14445–14451. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Parekh, A.B.; Penner, R. Store Depletion and Calcium Influx. Physiol. Rev. 1997, 77, 901–930. [Google Scholar] [CrossRef]
  8. Roos, J.; DiGregorio, P.J.; Yeromin, A.V.; Ohlsen, K.; Lioudyno, M.; Zhang, S.; Safrina, O.; Kozak, J.A.; Wagner, S.L.; Cahalan, M.D.; et al. STIM1, an Essential and Conserved Component of Store-Operated Ca2+ Channel Function. J. Cell Biol. 2005, 169, 435–445. [Google Scholar] [CrossRef] [Green Version]
  9. Liou, J.; Kim, M.L.; Heo, W.D.; Jones, J.T.; Myers, J.W.; Ferrell, J.E.; Meyer, T. STIM Is a Ca2+ Sensor Essential for Ca2+-Store-Depletion-Triggered Ca2+ Influx. Curr. Biol. 2005, 15, 1235–1241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Feske, S.; Gwack, Y.; Prakriya, M.; Srikanth, S.; Puppel, S.-H.; Tanasa, B.; Hogan, P.G.; Lewis, R.S.; Daly, M.; Rao, A. A Mutation in Orai1 Causes Immune Deficiency by Abrogating CRAC Channel Function. Nature 2006, 441, 179–185. [Google Scholar] [CrossRef]
  11. Williams, R.T.; Manji, S.S.; Parker, N.J.; Hancock, M.S.; Van Stekelenburg, L.; Eid, J.P.; Senior, P.V.; Kazenwadel, J.S.; Shandala, T.; Saint, R.; et al. Identification and Characterization of the STIM (Stromal Interaction Molecule) Gene Family: Coding for a Novel Class of Transmembrane Proteins. Biochem. J. 2001, 357, 673–685. [Google Scholar] [CrossRef] [PubMed]
  12. Stathopulos, P.B.; Li, G.-Y.; Plevin, M.J.; Ames, J.B.; Ikura, M. Stored Ca2+ Depletion-Induced Oligomerization of Stromal Interaction Molecule 1 (STIM1) via the EF-SAM Region. J. Biol. Chem. 2006, 281, 35855–35862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Spassova, M.A.; Soboloff, J.; He, L.-P.; Xu, W.; Dziadek, M.A.; Gill, D.L. STIM1 Has a Plasma Membrane Role in the Activation of Store-Operated Ca2+ Channels. Proc. Natl. Acad. Sci. USA 2006, 103, 4040–4045. [Google Scholar] [CrossRef] [Green Version]
  14. Mercer, J.C.; DeHaven, W.I.; Smyth, J.T.; Wedel, B.; Boyles, R.R.; Bird, G.S.; Putney, J.W. Large store-operated calcium-selective currents due to co-expression of orai1 or orai2 with the intracellular calcium sensor, stim1. J. Biol. Chem. 2006, 281, 24979–24990. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Stathopulos, P.B.; Zheng, L.; Ikura, M. Stromal Interaction Molecule (STIM) 1 and STIM2 Calcium Sensing Regions Exhibit Distinct Unfolding and Oligomerization Kinetics. J. Biol. Chem. 2009, 284, 728–732. [Google Scholar] [CrossRef] [Green Version]
  16. Kim, C.A.; Bowie, J.U. SAM Domains: Uniform Structure, Diversity of Function. Trends Biochem. Sci. 2003, 28, 625–628. [Google Scholar] [CrossRef]
  17. Ma, G.; Wei, M.; He, L.; Liu, C.; Wu, B.; Zhang, S.L.; Jing, J.; Liang, X.; Senes, A.; Tan, P.; et al. Inside-out Ca2+ Signalling Prompted by STIM1 Conformational Switch. Nat. Commun. 2015, 6, 7826. [Google Scholar] [CrossRef] [Green Version]
  18. Stathopulos, P.B.; Schindl, R.; Fahrner, M.; Zheng, L.; Gasmi-Seabrook, G.M.; Muik, M.; Romanin, C.; Ikura, M. STIM1/Orai1 Coiled-Coil Interplay in the Regulation of Store-Operated Calcium Entry. Nat. Commun. 2013, 4, 2963. [Google Scholar] [CrossRef]
  19. Fahrner, M.; Muik, M.; Schindl, R.; Butorac, C.; Stathopulos, P.; Zheng, L.; Jardin, I.; Ikura, M.; Romanin, C. A Coiled-Coil Clamp Controls Both Conformation and Clustering of Stromal Interaction Molecule 1 (STIM1). J. Biol. Chem. 2014, 289, 33231–33244. [Google Scholar] [CrossRef] [Green Version]
  20. Muik, M.; Fahrner, M.; Schindl, R.; Stathopulos, P.; Frischauf, I.; Derler, I.; Plenk, P.; Lackner, B.; Groschner, K.; Ikura, M.; et al. STIM1 Couples to ORAI1 via an Intramolecular Transition into an Extended Conformation. EMBO J. 2011, 30, 1678–1689. [Google Scholar] [CrossRef]
  21. Kim, M.S.; Zeng, W.; Yuan, J.P.; Shin, D.M.; Worley, P.F.; Muallem, S. Native Store-Operated Ca2+ Influx Requires the Channel Function of Orai1 and TRPC1. J. Biol. Chem. 2009, 284, 9733–9741. [Google Scholar] [CrossRef] [Green Version]
  22. Yuan, J.P.; Zeng, W.; Dorwart, M.R.; Choi, Y.-J.; Worley, P.F.; Muallem, S. SOAR and the Polybasic STIM1 Domains Gate and Regulate Orai Channels. Nat. Cell Biol. 2009, 11, 337–343. [Google Scholar] [CrossRef]
  23. Park, C.Y.; Shcheglovitov, A.; Dolmetsch, R. The CRAC Channel Activator STIM1 Binds and Inhibits L-Type Voltage-Gated Calcium Channels. Science 2010, 330, 101–105. [Google Scholar] [CrossRef] [PubMed]
  24. Frischauf, I.; Muik, M.; Derler, I.; Bergsmann, J.; Fahrner, M.; Schindl, R.; Groschner, K.; Romanin, C. Molecular Determinants of the Coupling between STIM1 and Orai Channels: Differential Activation Of Orai1–3 Channels by a Stim1 Coiled-Coil Mutant. J. Biol. Chem. 2009, 284, 21696–21706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zeng, W.; Yuan, J.P.; Kim, M.S.; Choi, Y.J.; Huang, G.N.; Worley, P.F.; Muallem, S. STIM1 Gates TRPC Channels but Not Orai1 by Electrostatic Interaction. Mol. Cell 2008, 32, 439–448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Liou, J.; Fivaz, M.; Inoue, T.; Meyer, T. Live-Cell Imaging Reveals Sequential Oligomerization and Local Plasma Membrane Targeting of Stromal Interaction Molecule 1 after Ca2+ Store Depletion. Proc. Natl. Acad. Sci. USA 2007, 104, 9301–9306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Park, C.Y.; Hoover, P.J.; Mullins, F.M.; Bachhawat, P.; Covington, E.D.; Raunser, S.; Walz, T.; Garcia, K.C.; Dolmetsch, R.E.; Lewis, R.S. STIM1 Clusters and Activates CRAC Channels via Direct Binding of a Cytosolic Domain to Orai1. Cell 2009, 136, 876–890. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  28. 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]
  29. Zhou, Y.; Mancarella, S.; Wang, Y.; Yue, C.; Ritchie, M.; Gill, D.L.; Soboloff, J. The Short N-Terminal Domains of STIM1 and STIM2 Control the Activation Kinetics of Orai1 Channels. J. Biol. Chem. 2009, 284, 19164–19168. [Google Scholar] [CrossRef] [Green Version]
  30. Darbellay, B.; Arnaudeau, S.; Bader, C.R.; Konig, S.; Bernheim, L. STIM1L Is a New Actin-Binding Splice Variant Involved in Fast Repetitive Ca2+ Release. J. Cell Biol. 2011, 194, 335–346. [Google Scholar] [CrossRef] [Green Version]
  31. Ramesh, G.; Jarzembowski, L.; Schwarz, Y.; Poth, V.; Konrad, M.; Knapp, M.L.; Schwär, G.; Lauer, A.A.; Grimm, M.O.W.; Alansary, D.; et al. A Short Isoform of STIM1 Confers Frequency-Dependent Synaptic Enhancement. Cell Rep. 2021, 34, 108844. [Google Scholar] [CrossRef]
  32. Miederer, A.-M.; Alansary, D.; Schwär, G.; Lee, P.-H.; Jung, M.; Helms, V.; Niemeyer, B.A. A STIM2 Splice Variant Negatively Regulates Store-Operated Calcium Entry. Nat. Commun. 2015, 6, 6899. [Google Scholar] [CrossRef] [Green Version]
  33. Rana, A.; Yen, M.; Sadaghiani, A.M.; Malmersjö, S.; Park, C.Y.; Dolmetsch, R.E.; Lewis, R.S. Alternative Splicing Converts STIM2 from an Activator to an Inhibitor of Store-Operated Calcium Channels. J. Cell Biol. 2015, 209, 653–670. [Google Scholar] [CrossRef] [Green Version]
  34. Feske, S.; Prakriya, M.; Rao, A.; Lewis, R.S. A Severe Defect in CRAC Ca2+ Channel Activation and Altered K+ Channel Gating in T Cells from Immunodeficient Patients. J. Exp. Med. 2005, 202, 651–662. [Google Scholar] [CrossRef]
  35. Zhang, S.L.; Yeromin, A.V.; Zhang, X.H.-F.; Yu, Y.; Safrina, O.; Penna, A.; Roos, J.; Stauderman, K.A.; Cahalan, M.D. Genome-Wide RNAi Screen of Ca2+ Influx Identifies Genes That Regulate Ca2+ Release-Activated Ca2+ Channel Activity. Proc. Natl. Acad. Sci. USA 2006, 103, 9357–9362. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Yeromin, A.V.; Zhang, S.L.; Jiang, W.; Yu, Y.; Safrina, O.; Cahalan, M.D. Molecular Identification of the CRAC Channel by Altered Ion Selectivity in a Mutant of Orai. Nature 2006, 443, 226–229. [Google Scholar] [CrossRef]
  37. Vig, M.; Peinelt, C.; Beck, A.; Koomoa, D.L.; Rabah, D.; Koblan-Huberson, M.; Kraft, S.; Turner, H.; Fleig, A.; Penner, R.; et al. CRACM1 Is a Plasma Membrane Protein Essential for Store-Operated Ca2+ Entry. Science 2006, 312, 1220–1223. [Google Scholar] [CrossRef] [Green Version]
  38. Prakriya, M.; Feske, S.; Gwack, Y.; Srikanth, S.; Rao, A.; Hogan, P.G. Orai1 Is an Essential Pore Subunit of the CRAC Channel. Nature 2006, 443, 230–233. [Google Scholar] [CrossRef] [PubMed]
  39. Hou, X.; Pedi, L.; Diver, M.M.; Long, S.B. Crystal Structure of the Calcium Release-Activated Calcium Channel Orai. Science 2012, 338, 1308–1313. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Derler, I.; Plenk, P.; Fahrner, M.; Muik, M.; Jardin, I.; Schindl, R.; Gruber, H.J.; Groschner, K.; Romanin, C. The Extended Transmembrane Orai1 N-Terminal (ETON) Region Combines Binding Interface and Gate for Orai1 Activation by STIM1. J. Biol. Chem. 2013, 288, 29025–29034. [Google Scholar] [CrossRef] [Green Version]
  41. Gwack, Y.; Srikanth, S.; Feske, S.; Cruz-Guilloty, F.; Oh-hora, M.; Neems, D.S.; Hogan, P.G.; Rao, A. Biochemical and Functional Characterization of Orai Proteins. J. Biol. Chem. 2007, 282, 16232–16243. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Gross, S.A.; Wissenbach, U.; Philipp, S.E.; Freichel, M.; Cavalié, A.; Flockerzi, V. Murine ORAI2 Splice Variants Form Functional Ca2+ Release-Activated Ca2+ (CRAC) Channels. J. Biol. Chem. 2007, 282, 19375–19384. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Motiani, R.K.; Abdullaev, I.F.; Trebak, M. A Novel Native Store-Operated Calcium Channel Encoded by Orai3: Selective requirement of orai3 versus orai1 in estrogen receptor-positive versus estrogen receptor-negative breast cancer cells. J. Biol. Chem. 2010, 285, 19173–19183. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Lis, A.; Peinelt, C.; Beck, A.; Parvez, S.; Monteilh-Zoller, M.; Fleig, A.; Penner, R. CRACM1, CRACM2, and CRACM3 Are Store-Operated Ca2+ Channels with Distinct Functional Properties. Curr. Biol. 2007, 17, 794–800. [Google Scholar] [CrossRef] [Green Version]
  45. Hoth, M.; Niemeyer, B.A. Chapter Ten—The Neglected CRAC Proteins: Orai2, Orai3, and STIM2. In Current Topics in Membranes; Prakriya, M., Ed.; Store-Operated Calcium Channels; Academic Press: Cambridge, MA, USA, 2013; Volume 71, pp. 237–271. [Google Scholar]
  46. Fukushima, M.; Tomita, T.; Janoshazi, A.; Putney, J.W. Alternative Translation Initiation Gives Rise to Two Isoforms of Orai1 with Distinct Plasma Membrane Mobilities. J. Cell Sci. 2012, 125, 4354–4361. [Google Scholar] [CrossRef] [Green Version]
  47. Desai, P.N.; Zhang, X.; Wu, S.; Janoshazi, A.; Bolimuntha, S.; Putney, J.W.; Trebak, M. Multiple Types of Calcium Channels Arising from Alternative Translation Initiation of the Orai1 Message. Sci. Signal. 2015, 8, ra74. [Google Scholar] [CrossRef] [Green Version]
  48. Montell, C. The History of TRP Channels, a Commentary and Reflection. Pflug. Arch.—Eur. J. Physiol. 2011, 461, 499–506. [Google Scholar] [CrossRef]
  49. Li, H. TRP Channel Classification. In Transient Receptor Potential Canonical Channels and Brain Diseases; Wang, Y., Ed.; Advances in Experimental Medicine and Biology; Springer: Dordrecht, The Netherlands, 2017; pp. 1–8. ISBN 978-94-024-1088-4. [Google Scholar]
  50. Owsianik, G.; Talavera, K.; Voets, T.; Nilius, B. Permeation and Selectivity of TRP Channels. Annu. Rev. Physiol. 2006, 68, 685–717. [Google Scholar] [CrossRef]
  51. Vazquez, G.; Wedel, B.J.; Aziz, O.; Trebak, M.; Putney, J.W. The Mammalian TRPC Cation Channels. Biochim. Et Biophys. Acta (BBA)—Mol. Cell Res. 2004, 1742, 21–36. [Google Scholar] [CrossRef] [Green Version]
  52. Salido, G.M.; Sage, S.O.; Rosado, J.A. TRPC Channels and Store-Operated Ca2+ Entry. Biochim. Et Biophys. Acta (BBA)—Mol. Cell Res. 2009, 1793, 223–230. [Google Scholar] [CrossRef] [Green Version]
  53. Tang, Q.; Guo, W.; Zheng, L.; Wu, J.-X.; Liu, M.; Zhou, X.; Zhang, X.; Chen, L. Structure of the Receptor-Activated Human TRPC6 and TRPC3 Ion Channels. Cell Res. 2018, 28, 746–755. [Google Scholar] [CrossRef] [Green Version]
  54. Wang, H.; Cheng, X.; Tian, J.; Xiao, Y.; Tian, T.; Xu, F.; Hong, X.; Zhu, M.X. TRPC Channels: Structure, Function, Regulation and Recent Advances in Small Molecular Probes. Pharmacol. Ther. 2020, 209, 107497. [Google Scholar] [CrossRef] [PubMed]
  55. Lepage, P.K.; Lussier, M.P.; Barajas-Martinez, H.; Bousquet, S.M.; Blanchard, A.P.; Francoeur, N.; Dumaine, R.; Boulay, G. Identification of Two Domains Involved in the Assembly of Transient Receptor Potential Canonical Channels. J. Biol. Chem. 2006, 281, 30356–30364. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Gaudet, R. A Primer on Ankyrin Repeat Function in TRP Channels and Beyond. Mol. Biosyst. 2008, 4, 372–379. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. 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]
  58. Zhang, Z.; Tang, J.; Tikunova, S.; Johnson, J.D.; Chen, Z.; Qin, N.; Dietrich, A.; Stefani, E.; Birnbaumer, L.; Zhu, M.X. Activation of Trp3 by Inositol 1,4,5-Trisphosphate Receptors through Displacement of Inhibitory Calmodulin from a Common Binding Domain. Proc. Natl. Acad. Sci. USA 2001, 98, 3168–3173. [Google Scholar] [CrossRef] [Green Version]
  59. Wedel, B.J.; Vazquez, G.; McKay, R.R.; Bird, G.S.T.; Putney, J.W. A Calmodulin/Inositol 1,4,5-Trisphosphate (IP3) Receptor-Binding Region Targets TRPC3 to the Plasma Membrane in a Calmodulin/IP3 Receptor-Independent Process. J. Biol. Chem. 2003, 278, 25758–25765. [Google Scholar] [CrossRef] [Green Version]
  60. Zitt, C.; Zobel, A.; Obukhov, A.G.; Harteneck, C.; Kalkbrenner, F.; Lückhoff, A.; Schultz, G. Cloning and Functional Expression of a Human Ca2+-Permeable Cation Channel Activated by Calcium Store Depletion. Neuron 1996, 16, 1189–1196. [Google Scholar] [CrossRef] [Green Version]
  61. Zhu, X.; Jiang, M.; Peyton, M.; Boulay, G.; Hurst, R.; Stefani, E.; Birnbaumer, L. Trp, a Novel Mammalian Gene Family Essential for Agonist-Activated Capacitative Ca2+ Entry. Cell 1996, 85, 661–671. [Google Scholar] [CrossRef] [Green Version]
  62. Liu, X.; Wang, W.; Singh, B.B.; Lockwich, T.; Jadlowiec, J.; O’Connell, B.; Wellner, R.; Zhu, M.X.; Ambudkar, I.S. Trp1, a Candidate Protein for the Store-Operated Ca2+ Influx Mechanism in Salivary Gland Cells. J. Biol. Chem. 2000, 275, 3403–3411. [Google Scholar] [CrossRef] [Green Version]
  63. Mehta, D.; Ahmmed, G.U.; Paria, B.C.; Holinstat, M.; Voyno-Yasenetskaya, T.; Tiruppathi, C.; Minshall, R.D.; Malik, A.B. RhoA Interaction with Inositol 1,4,5-Trisphosphate Receptor and Transient Receptor Potential Channel-1 Regulates Ca2+ Entry: Role in signaling increased endothelial permeability. J. Biol. Chem. 2003, 278, 33492–33500. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Antigny, F.; Koenig, S.; Bernheim, L.; Frieden, M. During Post-Natal Human Myogenesis, Normal Myotube Size Requires TRPC1- and TRPC4-Mediated Ca2+ Entry. J. Cell Sci. 2013, 126, 2525–2533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Sabourin, J.; Lamiche, C.; Vandebrouck, A.; Magaud, C.; Rivet, J.; Cognard, C.; Bourmeyster, N.; Constantin, B. Regulation of TRPC1 and TRPC4 Cation Channels Requires an Alpha1-Syntrophin-Dependent Complex in Skeletal Mouse Myotubes. J. Biol. Chem. 2009, 284, 36248–36261. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Antigny, F.; Sabourin, J.; Saüc, S.; Bernheim, L.; Koenig, S.; Frieden, M. TRPC1 and TRPC4 Channels Functionally Interact with STIM1L to Promote Myogenesis and Maintain Fast Repetitive Ca2+ Release in Human Myotubes. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 806–813. [Google Scholar] [CrossRef]
  67. Brownlow, S.L.; Harper, A.G.S.; Harper, M.T.; Sage, S.O. A Role for HTRPC1 and Lipid Raft Domains in Store-Mediated Calcium Entry in Human Platelets. Cell Calcium 2004, 35, 107–113. [Google Scholar] [CrossRef]
  68. Liu, X.; Cheng, K.T.; Bandyopadhyay, B.C.; Pani, B.; Dietrich, A.; Paria, B.C.; Swaim, W.D.; Beech, D.; Yildrim, E.; Singh, B.B.; et al. Attenuation of Store-Operated Ca2+ Current Impairs Salivary Gland Fluid Secretion in TRPC1(−/−) Mice. Proc. Natl. Acad. Sci. USA 2007, 104, 17542–17547. [Google Scholar] [CrossRef] [Green Version]
  69. Hong, J.H.; Li, Q.; Kim, M.S.; Shin, D.M.; Feske, S.; Birnbaumer, L.; Cheng, K.T.; Ambudkar, I.S.; Muallem, S. Polarized but Differential Localization and Recruitment of STIM1/Orai1 and STIM1/TRPC Channels in Secretory Cells. Traffic 2011, 12, 232–245. [Google Scholar] [CrossRef] [Green Version]
  70. Ma, X.; Cheng, K.-T.; Wong, C.-O.; O’Neil, R.G.; Birnbaumer, L.; Ambudkar, I.S.; Yao, X. Heteromeric TRPV4-C1 Channels Contribute to Store-Operated Ca2+ Entry in Vascular Endothelial Cells. Cell Calcium 2011, 50, 502–509. [Google Scholar] [CrossRef] [Green Version]
  71. Vazquez, G.; Wedel, B.J.; Trebak, M.; Bird, G.S.J.; Putney, J.W. Expression Level of the Canonical Transient Receptor Potential 3 (TRPC3) Channel Determines Its Mechanism of Activation. J. Biol. Chem. 2003, 278, 21649–21654. [Google Scholar] [CrossRef] [Green Version]
  72. Wu, X.; Zagranichnaya, T.K.; Gurda, G.T.; Eves, E.M.; Villereal, M.L. A TRPC1/TRPC3-Mediated Increase in Store-Operated Calcium Entry Is Required for Differentiation of H19-7 Hippocampal Neuronal Cells. J. Biol. Chem. 2004, 279, 43392–43402. [Google Scholar] [CrossRef] [Green Version]
  73. Antigny, F.; Jousset, H.; König, S.; Frieden, M. Thapsigargin Activates Ca2+ Entry Both by Store-Dependent, STIM1/Orai1-Mediated, and Store-Independent, TRPC3/PLC/PKC-Mediated Pathways in Human Endothelial Cells. Cell Calcium 2011, 49, 115–127. [Google Scholar] [CrossRef] [Green Version]
  74. Yuan, J.P.; Zeng, W.; Huang, G.N.; Worley, P.F.; Muallem, S. STIM1 Heteromultimerizes TRPC Channels to Determine Their Function as Store-Operated Channels. Nat. Cell Biol. 2007, 9, 636–645. [Google Scholar] [CrossRef] [PubMed]
  75. Kim, M.S.; Hong, J.H.; Li, Q.; Shin, D.M.; Abramowitz, J.; Birnbaumer, L.; Muallem, S. Deletion of TRPC3 in Mice Reduces Store-Operated Ca2+ Influx and the Severity of Acute Pancreatitis. Gastroenterology 2009, 137, 1509–1517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Kim, M.S.; Lee, K.P.; Yang, D.; Shin, D.M.; Abramowitz, J.; Kiyonaka, S.; Birnbaumer, L.; Mori, Y.; Muallem, S. Genetic and Pharmacological Inhibition of the Ca2+ Influx Channel TRPC3 Protects Secretory Epithelia from Ca2+-Dependent Toxicity. Gastroenterology 2011, 140, 2107–2115. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Kinoshita, M.; Akaike, A.; Satoh, M.; Kaneko, S. Positive Regulation of Capacitative Ca2+ Entry by Intracellular Ca2+ in Xenopus Oocytes Expressing Rat TRP4. Cell Calcium 2000, 28, 151–159. [Google Scholar] [CrossRef] [PubMed]
  78. Philipp, S.; Trost, C.; Warnat, J.; Rautmann, J.; Himmerkus, N.; Schroth, G.; Kretz, O.; Nastainczyk, W.; Cavalie, A.; Hoth, M.; et al. TRP4 (CCE1) Protein Is Part of Native Calcium Release-Activated Ca2+-like Channels in Adrenal Cells. J. Biol. Chem. 2000, 275, 23965–23972. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Yang, H.; Mergler, S.; Sun, X.; Wang, Z.; Lu, L.; Bonanno, J.A.; Pleyer, U.; Reinach, P.S. TRPC4 Knockdown Suppresses Epidermal Growth Factor-Induced Store-Operated Channel Activation and Growth in Human Corneal Epithelial Cells. J. Biol. Chem. 2005, 280, 32230–32237. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Fatherazi, S.; Presland, R.B.; Belton, C.M.; Goodwin, P.; Al-Qutub, M.; Trbic, Z.; Macdonald, G.; Schubert, M.M.; Izutsu, K.T. Evidence That TRPC4 Supports the Calcium Selective I(CRAC)-like Current in Human Gingival Keratinocytes. Pflug. Arch. 2007, 453, 879–889. [Google Scholar] [CrossRef]
  81. Sabourin, J.; Bartoli, F.; Antigny, F.; Gomez, A.M.; Benitah, J.-P. Transient Receptor Potential Canonical (TRPC)/Orai1-Dependent Store-Operated Ca2+ Channels: New targets of aldosterone in cardiomyocytes. J. Biol. Chem. 2016, 291, 13394–13409. [Google Scholar] [CrossRef] [Green Version]
  82. Sundivakkam, P.C.; Freichel, M.; Singh, V.; Yuan, J.P.; Vogel, S.M.; Flockerzi, V.; Malik, A.B.; Tiruppathi, C. The Ca2+ Sensor Stromal Interaction Molecule 1 (STIM1) Is Necessary and Sufficient for the Store-Operated Ca2+ Entry Function of Transient Receptor Potential Canonical (TRPC) 1 and 4 Channels in Endothelial Cells. Mol. Pharmacol. 2012, 81, 510–526. [Google Scholar] [CrossRef] [Green Version]
  83. Freichel, M.; Suh, S.H.; Pfeifer, A.; Schweig, U.; Trost, C.; Weißgerber, P.; Biel, M.; Philipp, S.; Freise, D.; Droogmans, G.; et al. Lack of an Endothelial Store-Operated Ca2+ Current Impairs Agonist-Dependent Vasorelaxation in TRP4−/− Mice. Nat. Cell Biol. 2001, 3, 121–127. [Google Scholar] [CrossRef]
  84. Tiruppathi, C.; Freichel, M.; Vogel, S.M.; Paria, B.C.; Mehta, D.; Flockerzi, V.; Malik, A.B. Impairment of Store-Operated Ca2+ Entry in TRPC4−/− Mice Interferes With Increase in Lung Microvascular Permeability. Circ. Res. 2002, 91, 70–76. [Google Scholar] [CrossRef] [Green Version]
  85. Xu, S.-Z.; Boulay, G.; Flemming, R.; Beech, D.J. E3-Targeted Anti-TRPC5 Antibody Inhibits Store-Operated Calcium Entry in Freshly Isolated Pial Arterioles. Am. J. Physiol.-Heart Circ. Physiol. 2006, 291, H2653–H2659. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Domínguez-Rodríguez, A.; Mayoral-Gonzalez, I.; Avila-Medina, J.; de Rojas-de Pedro, E.S.; Calderón-Sánchez, E.; Díaz, I.; Hmadcha, A.; Castellano, A.; Rosado, J.A.; Benitah, J.-P.; et al. Urocortin-2 Prevents Dysregulation of Ca2+ Homeostasis and Improves Early Cardiac Remodeling After Ischemia and Reperfusion. Front. Physiol. 2018, 9, 813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Bréchard, S.; Melchior, C.; Plançon, S.; Schenten, V.; Tschirhart, E.J. Store-Operated Ca2+ Channels Formed by TRPC1, TRPC6 and Orai1 and Non-Store-Operated Channels Formed by TRPC3 Are Involved in the Regulation of NADPH Oxidase in HL-60 Granulocytes. Cell Calcium 2008, 44, 492–506. [Google Scholar] [CrossRef] [PubMed]
  88. Jardín, I.; Redondo, P.C.; Salido, G.M.; Rosado, J.A. Phosphatidylinositol 4,5-Bisphosphate Enhances Store-Operated Calcium Entry through HTRPC6 Channel in Human Platelets. Biochim. Biophys. Acta 2008, 1783, 84–97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Jardin, I.; Gómez, L.J.; Salido, G.M.; Rosado, J.A. Dynamic Interaction of HTRPC6 with the Orai1-STIM1 Complex or HTRPC3 Mediates Its Role in Capacitative or Non-Capacitative Ca2+ Entry Pathways. Biochem. J. 2009, 420, 267–276. [Google Scholar] [CrossRef] [Green Version]
  90. Berna-Erro, A.; Galan, C.; Dionisio, N.; Gomez, L.J.; Salido, G.M.; Rosado, J.A. Capacitative and Non-Capacitative Signaling Complexes in Human Platelets. Biochim. Et Biophys. Acta (BBA)—Mol. Cell Res. 2012, 1823, 1242–1251. [Google Scholar] [CrossRef] [Green Version]
  91. El Boustany, C.; Bidaux, G.; Enfissi, A.; Delcourt, P.; Prevarskaya, N.; Capiod, T. Capacitative Calcium Entry and Transient Receptor Potential Canonical 6 Expression Control Human Hepatoma Cell Proliferation. Hepatology 2008, 47, 2068–2077. [Google Scholar] [CrossRef]
  92. DeHaven, W.I.; Jones, B.F.; Petranka, J.G.; Smyth, J.T.; Tomita, T.; Bird, G.S.; Putney, J.W. TRPC Channels Function Independently of STIM1 and Orai1. J. Physiol. 2009, 587, 2275–2298. [Google Scholar] [CrossRef]
  93. Zhang, X.; Trebak, M. Transient Receptor Potential Canonical 7 (TRPC7): A Diacylglycerol-Activated Non-Selective Cation Channel. Handb. Exp. Pharmacol. 2014, 222, 189–204. [Google Scholar] [CrossRef] [Green Version]
  94. Lemonnier, L.; Trebak, M.; Putney, J.W. Complex Regulation of the TRPC3,6,7 Channel Subfamily by Diacylglycerol and Phosphatidylinositol 4,5-Bisphosphate. Cell Calcium 2008, 43, 506–514. [Google Scholar] [CrossRef] [Green Version]
  95. Maroto, R.; Raso, A.; Wood, T.G.; Kurosky, A.; Martinac, B.; Hamill, O.P. TRPC1 Forms the Stretch-Activated Cation Channel in Vertebrate Cells. Nat. Cell. Biol. 2005, 7, 179–185. [Google Scholar] [CrossRef]
  96. Wes, P.D.; Chevesich, J.; Jeromin, A.; Rosenberg, C.; Stetten, G.; Montell, C. TRPC1, a Human Homolog of a Drosophila Store-Operated Channel. Proc. Natl. Acad. Sci. USA 1995, 92, 9652–9656. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Lucas, P.; Ukhanov, K.; Leinders-Zufall, T.; Zufall, F. A Diacylglycerol-Gated Cation Channel in Vomeronasal Neuron Dendrites Is Impaired in TRPC2 Mutant Mice: Mechanism of Pheromone Transduction. Neuron 2003, 40, 551–561. [Google Scholar] [CrossRef] [Green Version]
  98. Zitt, C.; Obukhov, A.G.; Strübing, C.; Zobel, A.; Kalkbrenner, F.; Lückhoff, A.; Schultz, G. Expression of TRPC3 in Chinese Hamster Ovary Cells Results in Calcium-Activated Cation Currents Not Related to Store Depletion. J. Cell Biol. 1997, 138, 1333–1341. [Google Scholar] [CrossRef] [PubMed]
  99. Hofmann, T.; Obukhov, A.G.; Schaefer, M.; Harteneck, C.; Gudermann, T.; Schultz, G. Direct Activation of Human TRPC6 and TRPC3 Channels by Diacylglycerol. Nature 1999, 397, 259–263. [Google Scholar] [CrossRef] [PubMed]
  100. Kamouchi, M.; Philipp, S.; Flockerzi, V.; Wissenbach, U.; Mamin, A.; Raeymaekers, L.; Eggermont, J.; Droogmans, G.; Nilius, B. Properties of Heterologously Expressed HTRP3 Channels in Bovine Pulmonary Artery Endothelial Cells. J. Physiol. 1999, 518, 345–358. [Google Scholar] [CrossRef]
  101. Schaefer, M.; Plant, T.D.; Obukhov, A.G.; Hofmann, T.; Gudermann, T.; Schultz, G. Receptor-Mediated Regulation of the Nonselective Cation Channels TRPC4 and TRPC5. J. Biol. Chem. 2000, 275, 17517–17526. [Google Scholar] [CrossRef] [Green Version]
  102. Kew, J.N.C.; Davies, C.H. Ion Channels: From Structure to Function; Oxford University Press: Oxford, UK, 2010; ISBN 978-0-19-929675-0. [Google Scholar]
  103. Feng, S. TRPC Channel Structure and Properties. Adv. Exp. Med. Biol. 2017, 976, 9–23. [Google Scholar] [CrossRef]
  104. Jung, S.; Mühle, A.; Schaefer, M.; Strotmann, R.; Schultz, G.; Plant, T.D. Lanthanides Potentiate TRPC5 Currents by an Action at Extracellular Sites Close to the Pore Mouth. J. Biol. Chem. 2003, 278, 3562–3571. [Google Scholar] [CrossRef] [Green Version]
  105. Spassova, M.A.; Hewavitharana, T.; Xu, W.; Soboloff, J.; Gill, D.L. A Common Mechanism Underlies Stretch Activation and Receptor Activation of TRPC6 Channels. Proc. Natl. Acad. Sci. USA 2006, 103, 16586–16591. [Google Scholar] [CrossRef] [Green Version]
  106. Okada, T.; Inoue, R.; Yamazaki, K.; Maeda, A.; Kurosaki, T.; Yamakuni, T.; Tanaka, I.; Shimizu, S.; Ikenaka, K.; Imoto, K.; et al. Molecular and Functional Characterization of a Novel Mouse Transient Receptor Potential Protein Homologue TRP7. J. Biol. Chem. 1999, 274, 27359–27370. [Google Scholar] [CrossRef] [Green Version]
  107. Berridge, M.J.; Irvine, R.F. Inositol Trisphosphate, a Novel Second Messenger in Cellular Signal Transduction. Nature 1984, 312, 315–321. [Google Scholar] [CrossRef]
  108. Streb, H.; Irvine, R.F.; Berridge, M.J.; Schulz, I. Release of Ca2+ from a Nonmitochondrial Intracellular Store in Pancreatic Acinar Cells by Inositol-1,4,5-Trisphosphate. Nature 1983, 306, 67–69. [Google Scholar] [CrossRef] [PubMed]
  109. Korzeniowski, M.K.; Manjarrés, I.M.; Varnai, P.; Balla, T. Activation of STIM1-ORAI1 involves an intramolecular switching mechanism. Sci. Signal. 2010, 3, ra82. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Korzeniowski, M.K.; Wisniewski, E.; Baird, B.; Holowka, D.A.; Balla, T. Molecular Anatomy of the Early Events in STIM1 Activation—Oligomerization or Conformational Change? J. Cell Sci. 2017, 130, 2821–2832. [Google Scholar] [CrossRef] [Green Version]
  111. Luik, R.M.; Wang, B.; Prakriya, M.; Wu, M.M.; Lewis, R.S. Oligomerization of STIM1 Couples ER Calcium Depletion to CRAC Channel Activation. Nature 2008, 454, 538–542. [Google Scholar] [CrossRef] [PubMed]
  112. Hoover, P.J.; Lewis, R.S. Stoichiometric Requirements for Trapping and Gating of Ca2+ Release-Activated Ca2+ (CRAC) Channels by Stromal Interaction Molecule 1 (STIM1). Proc. Natl. Acad. Sci. USA 2011, 108, 13299–13304. [Google Scholar] [CrossRef] [Green Version]
  113. 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]
  114. Lee, K.P.; Yuan, J.P.; So, I.; Worley, P.F.; Muallem, S. STIM1-Dependent and STIM1-Independent Function of Transient Receptor Potential Canonical (TRPC) Channels Tunes Their Store-Operated Mode. J. Biol. Chem. 2010, 285, 38666–38673. [Google Scholar] [CrossRef] [Green Version]
  115. Cheng, K.T.; Liu, X.; Ong, H.L.; Swaim, W.; Ambudkar, I.S. Local Ca2+ Entry Via Orai1 Regulates Plasma Membrane Recruitment of TRPC1 and Controls Cytosolic Ca2+ Signals Required for Specific Cell Functions. PLoS Biol. 2011, 9, e1001025. [Google Scholar] [CrossRef] [Green Version]
  116. de Souza, L.B.; Ong, H.L.; Liu, X.; Ambudkar, I.S. Fast Endocytic Recycling Determines TRPC1-STIM1 Clustering in ER-PM Junctions and Plasma Membrane Function of the Channel. Biochim. Biophys. Acta 2015, 1853, 2709–2721. [Google Scholar] [CrossRef] [Green Version]
  117. Mignen, O.; Thompson, J.L.; Shuttleworth, T.J. Ca2+ Selectivity and Fatty Acid Specificity of the Noncapacitative, Arachidonate-Regulated Ca2+ (ARC) Channels. J. Biol. Chem. 2003, 278, 10174–10181. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. 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 LeukotrieneC4: Role in Neointimal Hyperplasia. Circ. Res. 2013, 112, 1013–1025. [Google Scholar] [CrossRef] [Green Version]
  119. Mignen, O.; Thompson, J.L.; Shuttleworth, T.J. Both Orai1 and Orai3 Are Essential Components of the Arachidonate-Regulated Ca2+-Selective (ARC) Channels. J. Physiol. 2008, 586, 185–195. [Google Scholar] [CrossRef] [PubMed]
  120. Zhang, X.; Zhang, W.; González-Cobos, J.C.; Jardin, I.; Romanin, C.; Matrougui, K.; Trebak, M. Complex Role of STIM1 in the Activation of Store-Independent Orai1/3 Channels. J. Gen. Physiol. 2014, 143, 345–359. [Google Scholar] [CrossRef] [Green Version]
  121. Mignen, O.; Thompson, J.L.; Shuttleworth, T.J. STIM1 Regulates Ca2+ Entry via Arachidonate-Regulated Ca2+-Selective (ARC) Channels without Store Depletion or Translocation to the Plasma Membrane. J. Physiol. 2007, 579, 703–715. [Google Scholar] [CrossRef]
  122. Thompson, J.L.; Shuttleworth, T.J. Molecular Basis of Activation of the Arachidonate-Regulated Ca2+ (ARC) Channel, a Store-Independent Orai Channel, by Plasma Membrane STIM1. J. Physiol. 2013, 591, 3507–3523. [Google Scholar] [CrossRef] [PubMed]
  123. Humbert, M.; Guignabert, C.; Bonnet, S.; Dorfmüller, P.; Klinger, J.R.; Nicolls, M.R.; Olschewski, A.J.; Pullamsetti, S.S.; Schermuly, R.T.; Stenmark, K.R.; et al. Pathology and Pathobiology of Pulmonary Hypertension: State of the Art and Research Perspectives. Eur. Respir. J. 2019, 53, 1801887. [Google Scholar] [CrossRef] [Green Version]
  124. Simonneau, G.; Montani, D.; Celermajer, D.S.; Denton, C.P.; Gatzoulis, M.A.; Krowka, M.; Williams, P.G.; Souza, R. Haemodynamic Definitions and Updated Clinical Classification of Pulmonary Hypertension. Eur. Respir. J. 2019, 53, 1801913. [Google Scholar] [CrossRef]
  125. Galiè, N.; Humbert, M.; Vachiery, J.-L.; Gibbs, S.; Lang, I.; Torbicki, A.; Simonneau, G.; Peacock, A.; Noordegraaf, A.V.; Beghetti, M.; et al. 2015 ESC/ERS Guidelines for the Diagnosis and Treatment of Pulmonary Hypertension: The Joint Task Force for the Diagnosis and Treatment of Pulmonary Hypertension of the European Society of Cardiology (ESC) and the European Respiratory Society (ERS)Endorsed by: Association for European Paediatric and Congenital Cardiology (AEPC), International Society for Heart and Lung Transplantation (ISHLT). Eur. Respir. J. 2015, 46, 903–975. [Google Scholar] [CrossRef]
  126. Lau, E.M.T.; Giannoulatou, E.; Celermajer, D.S.; Humbert, M. Epidemiology and Treatment of Pulmonary Arterial Hypertension. Nat. Rev. Cardiol. 2017, 14, 603–614. [Google Scholar] [CrossRef]
  127. Farber, H.W.; Miller, D.P.; Poms, A.D.; Badesch, D.B.; Frost, A.E.; Muros-Le Rouzic, E.; Romero, A.J.; Benton, W.W.; Elliott, C.G.; McGoon, M.D.; et al. Five-Year Outcomes of Patients Enrolled in the REVEAL Registry. Chest 2015, 148, 1043–1054. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  128. Sitbon, O.; Gomberg-Maitland, M.; Granton, J.; Lewis, M.I.; Mathai, S.C.; Rainisio, M.; Stockbridge, N.L.; Wilkins, M.R.; Zamanian, R.T.; Rubin, L.J. Clinical Trial Design and New Therapies for Pulmonary Arterial Hypertension. Eur. Respir. J. 2019, 53, 1801908. [Google Scholar] [CrossRef] [Green Version]
  129. Antigny, F.; Girardin, N.; Frieden, M. Transient Receptor Potential Canonical Channels Are Required for in Vitro Endothelial Tube Formation. J. Biol. Chem. 2012, 287, 5917–5927. [Google Scholar] [CrossRef] [Green Version]
  130. Antigny, F.; Mercier, O.; Humbert, M.; Sabourin, J. Excitation-Contraction Coupling and Relaxation Alteration in Right Ventricular Remodelling Caused by Pulmonary Arterial Hypertension. Arch. Cardiovasc. Dis. 2020, 113, 70–84. [Google Scholar] [CrossRef] [PubMed]
  131. Sabourin, J.; Boet, A.; Rucker-Martin, C.; Lambert, M.; Gomez, A.-M.; Benitah, J.-P.; Perros, F.; Humbert, M.; Antigny, F. Ca2+ Handling Remodeling and STIM1L/Orai1/TRPC1/TRPC4 Upregulation in Monocrotaline-Induced Right Ventricular Hypertrophy. J. Mol. Cell. Cardiol. 2018, 118, 208–224. [Google Scholar] [CrossRef]
  132. Sankhe, S.; Manousakidi, S.; Antigny, F.; Arthur Ataam, J.; Bentebbal, S.; Ruchon, Y.; Lecerf, F.; Sabourin, J.; Price, L.; Fadel, E.; et al. T-Type Ca2+ Channels Elicit pro-Proliferative and Anti-Apoptotic Responses through Impaired PP2A/Akt1 Signaling in PASMCs from Patients with Pulmonary Arterial Hypertension. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2017, 1864, 1631–1641. [Google Scholar] [CrossRef]
  133. Ng, L.C.; Ramduny, D.; Airey, J.A.; Singer, C.A.; Keller, P.S.; Shen, X.-M.; Tian, H.; Valencik, M.; Hume, J.R. Orai1 Interacts with STIM1 and Mediates Capacitative Ca2+ Entry in Mouse Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol.-Cell Physiol. 2010, 299, C1079–C1090. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Wang, J. Orai1, 2, 3 and STIM1 Promote Store-Operated Calcium Entry in Pulmonary Arterial Smooth Muscle Cells. Cell Death Discov. 2017, 3, 17074. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Fernandez, R.A.; Wan, J.; Song, S.; Smith, K.A.; Gu, Y.; Tauseef, M.; Tang, H.; Makino, A.; Mehta, D.; Yuan, J.X.-J. 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]
  136. Song, M.Y.; Makino, A.; Yuan, J.X.-J. STIM2 Contributes to Enhanced Store-Operated Ca2+ Entry in Pulmonary Artery Smooth Muscle Cells from Patients with Idiopathic Pulmonary Arterial Hypertension. Pulm. Circ. 2011, 1, 84–94. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  137. Ogawa, A.; Firth, A.L.; Smith, K.A.; Maliakal, M.V.; Yuan, J.X.-J. PDGF Enhances Store-Operated Ca2+ Entry by Upregulating STIM1/Orai1 via Activation of Akt/MTOR in Human Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol.-Cell Physiol. 2011, 302, C405–C411. [Google Scholar] [CrossRef] [Green Version]
  138. Chen, T.-X.; Xu, X.-Y.; Zhao, Z.; Zhao, F.-Y.; Gao, Y.-M.; Yan, X.-H.; Wan, Y. Hydrogen Peroxide Is a Critical Regulator of the Hypoxia-Induced Alterations of Store-Operated Ca2+ Entry into Rat Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2017, 312, L477–L487. [Google Scholar] [CrossRef]
  139. Smith, K.A.; Voiriot, G.; Tang, H.; Fraidenburg, D.R.; Song, S.; Yamamura, H.; Yamamura, A.; Guo, Q.; Wan, J.; Pohl, N.M.; et al. Notch Activation of Ca2+ Signaling in the Development of Hypoxic Pulmonary Vasoconstriction and Pulmonary Hypertension. Am. J. Respir. Cell. Mol. Biol. 2015, 53, 355–367. [Google Scholar] [CrossRef] [Green Version]
  140. Hou, X.; Chen, J.; Luo, Y.; Liu, F.; Xu, G.; Gao, Y. Silencing of STIM1 Attenuates Hypoxia-Induced PASMCs Proliferation via Inhibition of the SOC/Ca2+/NFAT Pathway. Respir. Res. 2013, 14, 2. [Google Scholar] [CrossRef] [Green Version]
  141. Zhang, S.; Remillard, C.V.; Fantozzi, I.; Yuan, J.X.-J. ATP-Induced Mitogenesis Is Mediated by Cyclic AMP Response Element-Binding Protein-Enhanced TRPC4 Expression and Activity in Human Pulmonary Artery Smooth Muscle Cells. Am. J. Physiol.-Cell Physiol. 2004, 287, C1192–C1201. [Google Scholar] [CrossRef]
  142. Kunichika, N.; Yu, Y.; Remillard, C.V.; Platoshyn, O.; Zhang, S.; Yuan, J.X.-J. Overexpression of TRPC1 Enhances Pulmonary Vasoconstriction Induced by Capacitative Ca2+ Entry. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2004, 287, L962–L969. [Google Scholar] [CrossRef] [Green Version]
  143. Golovina, V.A.; Platoshyn, O.; Bailey, C.L.; Wang, J.; Limsuwan, A.; Sweeney, M.; Rubin, L.J.; Yuan, J.X.-J. Upregulated TRP and Enhanced Capacitative Ca2+ Entry in Human Pulmonary Artery Myocytes during Proliferation. Am. J. Physiol.-Heart Circ. Physiol. 2001, 280, H746–H755. [Google Scholar] [CrossRef] [Green Version]
  144. Sweeney, M.; Yu, Y.; Platoshyn, O.; Zhang, S.; McDaniel, S.S.; Yuan, J.X.-J. Inhibition of Endogenous TRP1 Decreases Capacitative Ca2+ Entry and Attenuates Pulmonary Artery Smooth Muscle Cell Proliferation. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2002, 283, L144–L155. [Google Scholar] [CrossRef]
  145. Lin, M.-J.; Leung, G.P.H.; Zhang, W.-M.; Yang, X.-R.; Yip, K.-P.; Tse, C.-M.; Sham, J.S.K. Chronic Hypoxia-Induced Upregulation of Store-Operated and Receptor-Operated Ca2+ Channels in Pulmonary Arterial Smooth Muscle Cells: A Novel Mechanism of Hypoxic Pulmonary Hypertension. Circ. Res. 2004, 95, 496–505. [Google Scholar] [CrossRef] [Green Version]
  146. Yu, Y.; Fantozzi, I.; Remillard, C.V.; Landsberg, J.W.; Kunichika, N.; Platoshyn, O.; Tigno, D.D.; Thistlethwaite, P.A.; Rubin, L.J.; Yuan, J.X.-J. Enhanced Expression of Transient Receptor Potential Channels in Idiopathic Pulmonary Arterial Hypertension. Proc. Natl. Acad. Sci. USA 2004, 101, 13861–13866. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Ng, L.C.; Gurney, A.M. Store-Operated Channels Mediate Ca2+ Influx and Contraction in Rat Pulmonary Artery. Circ. Res. 2001, 89, 923–929. [Google Scholar] [CrossRef] [Green Version]
  148. Yu, Y.; Sweeney, M.; Zhang, S.; Platoshyn, O.; Landsberg, J.; Rothman, A.; Yuan, J.X.-J. PDGF Stimulates Pulmonary Vascular Smooth Muscle Cell Proliferation by Upregulating TRPC6 Expression. Am. J. Physiol.-Cell Physiol. 2003, 284, C316–C330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  149. McDaniel, S.S.; Platoshyn, O.; Wang, J.; Yu, Y.; Sweeney, M.; Krick, S.; Rubin, L.J.; Yuan, J.X.-J. Capacitative Ca2+ Entry in Agonist-Induced Pulmonary Vasoconstriction. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2001, 280, L870–L880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  150. Wang, J.; Shimoda, L.A.; Sylvester, J.T. Capacitative Calcium Entry and TRPC Channel Proteins Are Expressed in Rat Distal Pulmonary Arterial Smooth Muscle. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2004, 286, L848–L858. [Google Scholar] [CrossRef]
  151. Ng, L.C.; O’Neill, K.G.; French, D.; Airey, J.A.; Singer, C.A.; Tian, H.; Shen, X.-M.; Hume, J.R. TRPC1 and Orai1 Interact with STIM1 and Mediate Capacitative Ca2+ Entry Caused by Acute Hypoxia in Mouse Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol.-Cell Physiol. 2012, 303, C1156–C1172. [Google Scholar] [CrossRef]
  152. Xia, Y.; Yang, X.-R.; Fu, Z.; Paudel, O.; Abramowitz, J.; Birnbaumer, L.; Sham, J.S.K. TRPC1 and TRPC6 Contribute to Hypoxic Pulmonary Hypertension through Differential Regulation of Pulmonary Vascular Functions RR. Hypertension 2014, 63, 173–180. [Google Scholar] [CrossRef] [Green Version]
  153. Malczyk, M.; Veith, C.; Fuchs, B.; Hofmann, K.; Storch, U.; Schermuly, R.T.; Witzenrath, M.; Ahlbrecht, K.; Fecher-Trost, C.; Flockerzi, V.; et al. Classical Transient Receptor Potential Channel 1 in Hypoxia-Induced Pulmonary Hypertension. Am. J. Respir. Crit. Care Med. 2013, 188, 1451–1459. [Google Scholar] [CrossRef]
  154. Lu, W.; Ran, P.; Zhang, D.; Lai, N.; Zhong, N.; Wang, J. Bone Morphogenetic Protein 4 Enhances Canonical Transient Receptor Potential Expression, Store-Operated Ca2+ Entry, and Basal [Ca2+]i in Rat Distal Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol.-Cell Physiol. 2010, 299, C1370–C1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  155. Wu, J.; Yu, Z.; Su, D. BMP4 Protects Rat Pulmonary Arterial Smooth Muscle Cells from Apoptosis by PI3K/AKT/Smad1/5/8 Signaling. Int. J. Mol. Sci. 2014, 15, 13738–13754. [Google Scholar] [CrossRef] [Green Version]
  156. Weissmann, N.; Dietrich, A.; Fuchs, B.; Kalwa, H.; Ay, M.; Dumitrascu, R.; Olschewski, A.; Storch, U.; Schnitzler, M.M.Y.; Ghofrani, A.; et al. Classical Transient Receptor Potential Channel 6 (TRPC6) Is Essential for Hypoxic Pulmonary Vasoconstriction and Alveolar Gas Exchange. Proc. Natl. Acad. Sci. USA 2006, 103, 19093–19098. [Google Scholar] [CrossRef] [Green Version]
  157. Wang, C.; Li, J.-F.; Zhao, L.; Liu, J.; Wan, J.; Wang, Y.X.; Wang, J.; Wang, C. Inhibition of SOC/Ca2+/NFAT Pathway Is Involved in the Anti-Proliferative Effect of Sildenafil on Pulmonary Artery Smooth Muscle Cells. Respir. Res. 2009, 10, 123. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  158. Ingueneau, C.; Huynh-Do, U.; Marcheix, B.; Athias, A.; Gambert, P.; Nègre-Salvayre, A.; Salvayre, R.; Vindis, C. TRPC1 Is Regulated by Caveolin-1 and Is Involved in Oxidized LDL-induced Apoptosis of Vascular Smooth Muscle Cells. J. Cell. Mol. Med. 2009, 13, 1620–1631. [Google Scholar] [CrossRef]
  159. Ng, L.C.; Airey, J.A.; Hume, J.R. The Contribution of TRPC1 and STIM1 to Capacitative Ca2+ Entry in Pulmonary Artery. Adv. Exp. Med. Biol. 2010, 661, 123–135. [Google Scholar] [CrossRef]
  160. Kumar, B.; Dreja, K.; Shah, S.S.; Cheong, A.; Xu, S.-Z.; Sukumar, P.; Naylor, J.; Forte, A.; Cipollaro, M.; McHugh, D.; et al. Upregulated TRPC1 Channel in Vascular Injury In Vivo and Its Role in Human Neointimal Hyperplasia. Circ. Res. 2006, 98, 557–563. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Takahashi, Y.; Watanabe, H.; Murakami, M.; Ono, K.; Munehisa, Y.; Koyama, T.; Nobori, K.; Iijima, T.; Ito, H. Functional Role of Stromal Interaction Molecule 1 (STIM1) in Vascular Smooth Muscle Cells. Biochem. Biophys. Res. Commun. 2007, 361, 934–940. [Google Scholar] [CrossRef]
  162. Li, J.; McKeown, L.; Ojelabi, O.; Stacey, M.; Foster, R.; O’Regan, D.; Porter, K.E.; Beech, D.J. Nanomolar Potency and Selectivity of a Ca2+ Release-Activated Ca2+ Channel Inhibitor against Store-Operated Ca2+ Entry and Migration of Vascular Smooth Muscle Cells. Br. J. Pharmacol. 2011, 164, 382–393. [Google Scholar] [CrossRef] [Green Version]
  163. Potier, M.; Gonzalez, J.C.; Motiani, R.K.; Abdullaev, I.F.; Bisaillon, J.M.; Singer, H.A.; Trebak, M. Evidence for STIM1- and Orai1-Dependent Store-Operated Calcium Influx through ICRAC in Vascular Smooth Muscle Cells: Role in Proliferation and Migration. FASEB J. 2009, 23, 2425–2437. [Google Scholar] [CrossRef] [Green Version]
  164. Guo, R.; Yang, L.; Li, M.; Pan, X.; Liu, B.; Deng, Y. Stim1- and Orai1-Mediated Store-Operated Calcium Entry Is Critical for Angiotensin II-Induced Vascular Smooth Muscle Cell Proliferation. Cardiovasc. Res. 2012, 93, 360–370. [Google Scholar] [CrossRef]
  165. Rodríguez-Moyano, M.; Díaz, I.; Dionisio, N.; Zhang, X.; Ávila-Medina, J.; Calderón-Sánchez, E.; Trebak, M.; Rosado, J.A.; Ordóñez, A.; Smani, T. Urotensin-II Promotes Vascular Smooth Muscle Cell Proliferation through Store-Operated Calcium Entry and EGFR Transactivation. Cardiovasc. Res. 2013, 100, 297–306. [Google Scholar] [CrossRef]
  166. Bisaillon, J.M.; Motiani, R.K.; Gonzalez-Cobos, J.C.; Potier, M.; Halligan, K.E.; Alzawahra, W.F.; Barroso, M.; Singer, H.A.; Jourd’heuil, D.; Trebak, M. Essential Role for STIM1/Orai1-Mediated Calcium Influx in PDGF-Induced Smooth Muscle Migration. Am. J. Physiol.-Cell Physiol. 2010, 298, C993–C1005. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Azimi, I.; Stevenson, R.J.; Zhang, X.; Meizoso-Huesca, A.; Xin, P.; Johnson, M.; Flanagan, J.U.; Chalmers, S.B.; Yoast, R.E.; Kapure, J.S.; et al. A New Selective Pharmacological Enhancer of the Orai1 Ca2+ Channel Reveals Roles for Orai1 in Smooth and Skeletal Muscle Functions. ACS Pharmacol. Transl. Sci. 2020, 3, 135–147. [Google Scholar] [CrossRef] [Green Version]
  168. Yang, H.; Chen, X.-Y.; Kuang, S.-J.; Zhou, M.-Y.; Zhang, L.; Zeng, Z.; Liu, L.; Wu, F.-L.; Zhang, M.-Z.; Mai, L.-P.; et al. Abnormal Ca2+ Handling Contributes to the Impairment of Aortic Smooth Muscle Contractility in Zucker Diabetic Fatty Rats. J. Mol. Cell. Cardiol. 2020, 141, 82–92. [Google Scholar] [CrossRef]
  169. Trebak, M. STIM/Orai Signalling Complexes in Vascular Smooth Muscle. J. Physiol. 2012, 590, 4201–4208. [Google Scholar] [CrossRef] [PubMed]
  170. Avila-Medina, J.; Mayoral-Gonzalez, I.; Dominguez-Rodriguez, A.; Gallardo-Castillo, I.; Ribas, J.; Ordoñez, A.; Rosado, J.A.; Smani, T. The Complex Role of Store Operated Calcium Entry Pathways and Related Proteins in the Function of Cardiac, Skeletal and Vascular Smooth Muscle Cells. Front. Physiol. 2018, 9, 257. [Google Scholar] [CrossRef]
  171. Kwan, H.-Y.; Shen, B.; Ma, X.; Kwok, Y.-C.; Huang, Y.; Man, Y.-B.; Yu, S.; Yao, X. TRPC1 Associates With BKCa Channel to Form a Signal Complex in Vascular Smooth Muscle Cells. Circ. Res. 2009, 104, 670–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Zhang, S.; Patel, H.H.; Murray, F.; Remillard, C.V.; Schach, C.; Thistlethwaite, P.A.; Insel, P.A.; Yuan, J.X.-J. Pulmonary Artery Smooth Muscle Cells from Normal Subjects and IPAH Patients Show Divergent CAMP-Mediated Effects on TRPC Expression and Capacitative Ca2+ Entry. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2007, 292, L1202–L1210. [Google Scholar] [CrossRef] [Green Version]
  173. He, X.; Song, S.; Ayon, R.J.; Balisterieri, A.; Black, S.M.; Makino, A.; Wier, W.G.; Zang, W.-J.; Yuan, J.X.-J. Hypoxia Selectively Upregulates Cation Channels and Increases Cytosolic [Ca2+] in Pulmonary, but Not Coronary, Arterial Smooth Muscle Cells. Am. J. Physiol.-Cell Physiol. 2018, 314, C504–C517. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Bonnet, S.; Michelakis, E.D.; Porter, C.J.; Andrade-Navarro, M.A.; Thébaud, B.; Bonnet, S.; Haromy, A.; Harry, G.; Moudgil, R.; McMurtry, M.S.; et al. An Abnormal Mitochondrial–Hypoxia Inducible Factor-1α–Kv Channel Pathway Disrupts Oxygen Sensing and Triggers Pulmonary Arterial Hypertension in Fawn Hooded Rats. Circulation 2006, 113, 2630–2641. [Google Scholar] [CrossRef] [Green Version]
  175. Dai, Z.; Li, M.; Wharton, J.; Zhu, M.M.; Zhao, Y.-Y. Prolyl-4 Hydroxylase 2 (PHD2) Deficiency in Endothelial Cells and Hematopoietic Cells Induces Obliterative Vascular Remodeling and Severe Pulmonary Arterial Hypertension in Mice and Humans Through Hypoxia-Inducible Factor-2α. Circulation 2016, 133, 2447–2458. [Google Scholar] [CrossRef] [Green Version]
  176. Lei, W.; He, Y.; Shui, X.; Li, G.; Yan, G.; Zhang, Y.; Huang, S.; Chen, C.; Ding, Y. Expression and Analyses of the HIF-1 Pathway in the Lungs of Humans with Pulmonary Arterial Hypertension. Mol. Med. Rep. 2016, 14, 4383–4390. [Google Scholar] [CrossRef]
  177. Chen, J.; Sysol, J.R.; Singla, S.; Zhao, S.; Yamamura, A.; Valdez-Jasso, D.; Abbasi, T.; Shioura, K.M.; Sahni, S.; Reddy, V.; et al. Nicotinamide Phosphoribosyltransferase Promotes Pulmonary Vascular Remodeling and Is a Therapeutic Target in Pulmonary Arterial Hypertension. Circulation 2017, 135, 1532–1546. [Google Scholar] [CrossRef]
  178. Lu, W.; Wang, J.; Shimoda, L.A.; Sylvester, J.T. Differences in STIM1 and TRPC Expression in Proximal and Distal Pulmonary Arterial Smooth Muscle Are Associated with Differences in Ca2+ Responses to Hypoxia. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2008, 295, L104–L113. [Google Scholar] [CrossRef] [Green Version]
  179. Giachini, F.R.C.; Chiao, C.-W.; Carneiro, F.S.; Lima, V.V.; Carneiro, Z.N.; Dorrance, A.M.; Tostes, R.C.; Webb, R.C. CHBPR: Increased Activation of STIM-1/Orai-1 in Aorta from Hypertensive Rats: A Novel Insight into Vascular Dysfunction. Hypertension 2009, 53, 409–416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  180. 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, 13–22. [Google Scholar] [CrossRef] [PubMed]
  181. Castillo-Galan, S.; Arenas, G.; Krause, B.; Iturriaga, R. Late Breaking Abstract—Contribution of STIM-Activated TRPC-ORAI Channels to the Intermittent Hypoxia-Induced Pulmonary Hypertension. Eur. Respir. J. 2020, 56. [Google Scholar] [CrossRef]
  182. Perros, F.; Montani, D.; Dorfmüller, P.; Durand-Gasselin, I.; Tcherakian, C.; Le Pavec, J.; Mazmanian, M.; Fadel, E.; Mussot, S.; Mercier, O.; et al. Platelet-Derived Growth Factor Expression and Function in Idiopathic Pulmonary Arterial Hypertension. Am. J. Respir. Crit. Care Med. 2008, 178, 81–88. [Google Scholar] [CrossRef]
  183. Schermuly, R.T.; Dony, E.; Ghofrani, H.A.; Pullamsetti, S.; Savai, R.; Roth, M.; Sydykov, A.; Lai, Y.J.; Weissmann, N.; Seeger, W.; et al. Reversal of Experimental Pulmonary Hypertension by PDGF Inhibition. J. Clin. Investig. 2005, 115, 2811–2821. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  184. Balasubramaniam, V.; Le Cras, T.D.; Ivy, D.D.; Grover, T.R.; Kinsella, J.P.; Abman, S.H. Role of Platelet-Derived Growth Factor in Vascular Remodeling during Pulmonary Hypertension in the Ovine Fetus. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2003, 284, L826–L833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Sun, C.-K.; Zhen, Y.-Y.; Lu, H.-I.; Sung, P.-H.; Chang, L.-T.; Tsai, T.-H.; Sheu, J.-J.; Chen, Y.-L.; Chua, S.; Chang, H.-W.; et al. Reducing TRPC1 Expression through Liposome-Mediated SiRNA Delivery Markedly Attenuates Hypoxia-Induced Pulmonary Arterial Hypertension in a Murine Model. Stem Cells Int. 2014, 2014, e316214. [Google Scholar] [CrossRef]
  186. Liu, X.-R.; Zhang, M.-F.; Yang, N.; Liu, Q.; Wang, R.-X.; Cao, Y.-N.; Yang, X.-R.; Sham, J.S.K.; Lin, M.-J. Enhanced Store-Operated Ca2+ Entry and TRPC Channel Expression in Pulmonary Arteries of Monocrotaline-Induced Pulmonary Hypertensive Rats. Am. J. Physiol.-Cell Physiol. 2012, 302, C77–C87. [Google Scholar] [CrossRef] [Green Version]
  187. Alzoubi, A.; Almalouf, P.; Toba, M.; O’Neill, K.; Qian, X.; Francis, M.; Taylor, M.S.; Alexeyev, M.; McMurtry, I.F.; Oka, M.; et al. TRPC4 Inactivation Confers a Survival Benefit in Severe Pulmonary Arterial Hypertension. Am. J. Pathol. 2013, 183, 1779–1788. [Google Scholar] [CrossRef] [Green Version]
  188. Yu, Y.; Keller, S.H.; Remillard, C.V.; Safrina, O.; Nicholson, A.; Zhang, S.L.; Jiang, W.; Vangala, N.; Landsberg, J.W.; Wang, J.-Y.; et al. A Functional Single-Nucleotide Polymorphism in the TRPC6 Gene Promoter Associated With Idiopathic Pulmonary Arterial Hypertension. Circulation 2009, 119, 2313–2322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Pousada, G.; Baloira, A.; Valverde, D. Molecular and Clinical Analysis of TRPC6 and AGTR1 Genes in Patients with Pulmonary Arterial Hypertension. Orphanet J. Rare Dis. 2015, 10, 1. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  190. Linde, C.I.; Karashima, E.; Raina, H.; Zulian, A.; Wier, W.G.; Hamlyn, J.M.; Ferrari, P.; Blaustein, M.P.; Golovina, V.A. Increased Arterial Smooth Muscle Ca2+ Signaling, Vasoconstriction, and Myogenic Reactivity in Milan Hypertensive Rats. Am. J. Physiol. Heart Circ. Physiol. 2012, 302, H611–H620. [Google Scholar] [CrossRef]
  191. Jain, P.P.; Lai, N.; Xiong, M.; Chen, J.; Babicheva, A.; Zhao, T.; Parmisano, S.; Zhao, M.; Paquin, C.; Matti, M.; et al. TRPC6, a Therapeutic Target for Pulmonary Hypertension. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2021. [Google Scholar] [CrossRef]
  192. Deng, Z.; Morse, J.H.; Slager, S.L.; Cuervo, N.; Moore, K.J.; Venetos, G.; Kalachikov, S.; Cayanis, E.; Fischer, S.G.; Barst, R.J.; et al. Familial Primary Pulmonary Hypertension (Gene PPH1) Is Caused by Mutations in the Bone Morphogenetic Protein Receptor-II Gene. Am. J. Hum. Genet. 2000, 67, 737–744. [Google Scholar] [CrossRef] [Green Version]
  193. Evans, J.D.W.; Girerd, B.; Montani, D.; Wang, X.-J.; Galiè, N.; Austin, E.D.; Elliott, G.; Asano, K.; Grünig, E.; Yan, Y.; et al. BMPR2 Mutations and Survival in Pulmonary Arterial Hypertension: An Individual Participant Data Meta-Analysis. Lancet Respir. Med. 2016, 4, 129–137. [Google Scholar] [CrossRef] [Green Version]
  194. Atkinson, C.; Stewart, S.; Upton, P.D.; Machado, R.; Thomson, J.R.; Trembath, R.C.; Morrell, N.W. Primary Pulmonary Hypertension Is Associated with Reduced Pulmonary Vascular Expression of Type II Bone Morphogenetic Protein Receptor. Circulation 2002, 105, 1672–1678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Yang, X.; Long, L.; Southwood, M.; Rudarakanchana, N.; Upton, P.D.; Jeffery, T.K.; Atkinson, C.; Chen, H.; Trembath, R.C.; Morrell, N.W. Dysfunctional Smad Signaling Contributes to Abnormal Smooth Muscle Cell Proliferation in Familial Pulmonary Arterial Hypertension. Circ. Res. 2005, 96, 1053–1063. [Google Scholar] [CrossRef] [Green Version]
  196. Yang, J.; Li, X.; Li, Y.; Southwood, M.; Ye, L.; Long, L.; Al-Lamki, R.S.; Morrell, N.W. Id Proteins Are Critical Downstream Effectors of BMP Signaling in Human Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2013, 305, L312–L321. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Orriols, M.; Gomez-Puerto, M.C.; ten Dijke, P. BMP Type II Receptor as a Therapeutic Target in Pulmonary Arterial Hypertension. Cell. Mol. Life Sci. 2017, 74, 2979–2995. [Google Scholar] [CrossRef] [Green Version]
  198. Anderson, L.; Lowery, J.W.; Frank, D.B.; Novitskaya, T.; Jones, M.; Mortlock, D.P.; Chandler, R.L.; de Caestecker, M.P. Bmp2 and Bmp4 Exert Opposing Effects in Hypoxic Pulmonary Hypertension. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2010, 298, R833–R842. [Google Scholar] [CrossRef] [Green Version]
  199. Zhang, Y.; Lu, W.; Yang, K.; Xu, L.; Lai, N.; Tian, L.; Jiang, Q.; Duan, X.; Chen, M.; Wang, J. Bone Morphogenetic Protein 2 Decreases TRPC Expression, Store-Operated Ca2+ Entry, and Basal [Ca2+]i in Rat Distal Pulmonary Arterial Smooth Muscle Cells. Am. J. Physiol. Cell Physiol. 2013, 304, C833–C843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  200. Kondratska, K.; Kondratskyi, A.; Yassine, M.; Lemonnier, L.; Lepage, G.; Morabito, A.; Skryma, R.; Prevarskaya, N. Orai1 and STIM1 Mediate SOCE and Contribute to Apoptotic Resistance of Pancreatic Adenocarcinoma. Biochim. Biophys. Acta (BBA)—Mol. Cell Res. 2014, 1843, 2263–2269. [Google Scholar] [CrossRef] [Green Version]
  201. 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. Arter. Thromb. Vasc. Biol. 2016, 36, 618–628. [Google Scholar] [CrossRef] [Green Version]
  202. Vacher, P.; Vacher, A.-M.; Pineau, R.; Latour, S.; Soubeyran, I.; Pangault, C.; Tarte, K.; Soubeyran, P.; Ducret, T.; Bresson-Bepoldin, L. Localized Store-Operated Calcium Influx Represses CD95-Dependent Apoptotic Effects of Rituximab in Non-Hodgkin B Lymphomas. J. Immunol. 2015, 195, 2207–2215. [Google Scholar] [CrossRef] [Green Version]
  203. Weigand, L.; Foxson, J.; Wang, J.; Shimoda, L.A.; Sylvester, J.T. Inhibition of Hypoxic Pulmonary Vasoconstriction by Antagonists of Store-Operated Ca2+ and Nonselective Cation Channels. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2005, 289, L5–L13. [Google Scholar] [CrossRef] [Green Version]
  204. Ward, J.P.T.; Robertson, T.P.; Aaronson, P.I. Capacitative Calcium Entry: A Central Role in Hypoxic Pulmonary Vasoconstriction? Am. J. Physiol.-Lung Cell. Mol. Physiol. 2005, 289, L2–L4. [Google Scholar] [CrossRef] [Green Version]
  205. Shibata, A.; Uchida, K.; Kodo, K.; Miyauchi, T.; Mikoshiba, K.; Takahashi, T.; Yamagishi, H. Type 2 Inositol 1,4,5-Trisphosphate Receptor Inhibits the Progression of Pulmonary Arterial Hypertension via Calcium Signaling and Apoptosis. Heart Vessels 2019, 34, 724–734. [Google Scholar] [CrossRef]
  206. Kiselyov, K.; Mignery, G.A.; Zhu, M.X.; Muallem, S. The N-Terminal Domain of the IP3 Receptor Gates Store-Operated HTrp3 Channels. Mol. Cell 1999, 4, 423–429. [Google Scholar] [CrossRef]
  207. 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]
  208. Yang, Z.; Song, T.; Truong, L.; Reyes-García, J.; Wang, L.; Zheng, Y.-M.; Wang, Y.-X. Important Role of Sarcoplasmic Reticulum Ca2+ Release via Ryanodine Receptor-2 Channel in Hypoxia-Induced Rieske Iron-Sulfur Protein-Mediated Mitochondrial Reactive Oxygen Species Generation in Pulmonary Artery Smooth Muscle Cells. Antioxid. Redox Signal. 2020, 32, 447–462. [Google Scholar] [CrossRef]
  209. Kaßmann, M.; Szijártó, I.A.; García-Prieto, C.F.; Fan, G.; Schleifenbaum, J.; Anistan, Y.-M.; Tabeling, C.; Shi, Y.; le Noble, F.; Witzenrath, M.; 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. [Google Scholar] [CrossRef] [Green Version]
  210. Gilbert, G.; Ducret, T.; Marthan, R.; Savineau, J.-P.; Quignard, J.-F. Stretch-Induced Ca2+ Signalling in Vascular Smooth Muscle Cells Depends on Ca2+ Store Segregation. Cardiovasc. Res. 2014, 103, 313–323. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  211. Dahan, D.; Ducret, T.; Quignard, J.-F.; Marthan, R.; Savineau, J.-P.; Estève, E. Implication of the Ryanodine Receptor in TRPV4-Induced Calcium Response in Pulmonary Arterial Smooth Muscle Cells from Normoxic and Chronically Hypoxic Rats. Am. J. Physiol.-Lung Cell. Mol. Physiol. 2012, 303, L824–L833. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Sitbon, O.; Humbert, M.; Jaïs, X.; Ioos, V.; Hamid, A.M.; Provencher, S.; Garcia, G.; Parent, F.; Hervé, P.; Simonneau, G. Long-Term Response to Calcium Channel Blockers in Idiopathic Pulmonary Arterial Hypertension. Circulation 2005, 111, 3105–3111. [Google Scholar] [CrossRef] [Green Version]
  213. The Effect of High Doses of Calcium-Channel Blockers on Survival in Primary Pulmonary Hypertension|NEJM. Available online: https://www.nejm.org/doi/full/10.1056/NEJM199207093270203 (accessed on 6 October 2021).
  214. He, L.-P.; Hewavitharana, T.; Soboloff, J.; Spassova, M.A.; Gill, D.L. A Functional Link between Store-Operated and TRPC Channels Revealed by the 3,5-Bis(Trifluoromethyl)Pyrazole Derivative, BTP2. J. Biol. Chem. 2005, 280, 10997–11006. [Google Scholar] [CrossRef] [Green Version]
  215. Takezawa, R.; Cheng, H.; Beck, A.; Ishikawa, J.; Launay, P.; Kubota, H.; Kinet, J.-P.; Fleig, A.; Yamada, T.; Penner, R. A Pyrazole Derivative Potently Inhibits Lymphocyte Ca2 Influx and Cytokine Production by Facilitating Transient Receptor Potential Melastatin 4 Channel Activity. Mol. Pharmacol. 2006, 69, 1413–1420. [Google Scholar] [CrossRef] [PubMed]
  216. Ishikawa, J.; Ohga, K.; Yoshino, T.; Takezawa, R.; Ichikawa, A.; Kubota, H.; Yamada, T. A Pyrazole Derivative, YM-58483, Potently Inhibits Store-Operated Sustained Ca2+ Influx and IL-2 Production in T Lymphocytes. J. Immunol. 2003, 170, 4441–4449. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Zitt, C.; Strauss, B.; Schwarz, E.C.; Spaeth, N.; Rast, G.; Hatzelmann, A.; Hoth, M. Potent Inhibition of Ca2+ Release-Activated Ca2+ Channels and T-Lymphocyte Activation by the Pyrazole Derivative BTP2. J. Biol. Chem. 2004, 279, 12427–12437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. Schleifer, H.; Doleschal, B.; Lichtenegger, M.; Oppenrieder, R.; Derler, I.; Frischauf, I.; Glasnov, T.; Kappe, C.; Romanin, C.; Groschner, K. Novel Pyrazole Compounds for Pharmacological Discrimination between Receptor-Operated and Store-Operated Ca2+ Entry Pathways. Br. J. Pharmacol. 2012, 167, 1712–1722. [Google Scholar] [CrossRef] [Green Version]
  219. Zhang, X.; Xin, P.; Yoast, R.E.; Emrich, S.M.; Johnson, M.T.; Pathak, T.; Benson, J.C.; Azimi, I.; Gill, D.L.; Monteith, G.R.; et al. Distinct Pharmacological Profiles of ORAI1, ORAI2, and ORAI3 Channels. Cell Calcium 2020, 91, 102281. [Google Scholar] [CrossRef]
  220. Derler, I.; Schindl, R.; Fritsch, R.; Heftberger, P.; Riedl, M.C.; Begg, M.; House, D.; Romanin, C. The Action of Selective CRAC Channel Blockers Is Affected by the Orai Pore Geometry. Cell Calcium 2013, 53, 139–151. [Google Scholar] [CrossRef]
  221. Shawer, H.; Norman, K.; Cheng, C.W.; Foster, R.; Beech, D.J.; Bailey, M.A. ORAI1 Ca2+ Channel as a Therapeutic Target in Pathological Vascular Remodelling. Front. Cell Dev. Biol. 2021, 9, 686. [Google Scholar] [CrossRef]
  222. Waldherr, L.; Tiffner, A.; Mishra, D.; Sallinger, M.; Schober, R.; Frischauf, I.; Schmidt, T.; Handl, V.; Sagmeister, P.; Köckinger, M.; et al. Blockage of Store-Operated Ca2+ Influx by Synta66 Is Mediated by Direct Inhibition of the Ca2+ Selective Orai1 Pore. Cancers 2020, 12, 2876. [Google Scholar] [CrossRef]
  223. Ng, S.W.; di Capite, J.; Singaravelu, K.; Parekh, A.B. Sustained Activation of the Tyrosine Kinase Syk by Antigen in Mast Cells Requires Local Ca2+ Influx through Ca2+ Release-Activated Ca2+ Channels. J. Biol. Chem. 2008, 283, 31348–31355. [Google Scholar] [CrossRef] [Green Version]
  224. Sabatino, A.D.; Rovedatti, L.; Kaur, R.; Spencer, J.P.; Brown, J.T.; Morisset, V.D.; Biancheri, P.; Leakey, N.A.B.; Wilde, J.I.; Scott, L.; et al. Targeting Gut T Cell Ca2+ Release-Activated Ca2+ Channels Inhibits T Cell Cytokine Production and T-Box Transcription Factor T-Bet in Inflammatory Bowel Disease. J. Immunol. 2009, 183, 3454–3462. [Google Scholar] [CrossRef] [Green Version]
  225. Bartoli, F.; Bailey, M.A.; Rode, B.; Mateo, P.; Antigny, F.; Bedouet, K.; Gerbaud, P.; Gosain, R.; Plante, J.; Norman, K.; et al. Orai1 Channel Inhibition Preserves Left Ventricular Systolic Function and Normal Ca 2+ Handling After Pressure Overload. Circulation 2020, 141, 199–216. [Google Scholar] [CrossRef]
  226. Trebak, M.; Earley, S. Signal Transduction and Smooth Muscle; CRC Press: Boca Raton, FL, USA, 2018; ISBN 978-1-4987-7423-9. [Google Scholar]
  227. Sadaghiani, A.M.; Lee, S.M.; Odegaard, J.I.; Leveson-Gower, D.B.; McPherson, O.M.; Novick, P.; Kim, M.R.; Koehler, A.N.; Negrin, R.; Dolmetsch, R.E.; et al. Identification of Orai1 Channel Inhibitors by Using Minimal Functional Domains to Screen Small Molecule Microarrays. Chem. Biol. 2014, 21, 1278–1292. [Google Scholar] [CrossRef]
  228. Kim, K.-D.; Srikanth, S.; Tan, Y.-V.; Yee, K.; Jew, M.; Damoiseaux, R.; Jung, M.; Shimizu, S.; An, D.; Ribalet, B.; et al. Calcium Signaling via Orai1 Is Essential for Induction of the Nuclear Orphan Receptor Pathway To Drive Th17 Differentiation. J. Immunol. 2013, 192, 110–122. [Google Scholar] [CrossRef] [Green Version]
  229. Smyth, J.T.; DeHaven, W.I.; Bird, G.S.; Putney, J.W., Jr. Ca2+-Store-Dependent and -Independent Reversal of Stim1 Localization and Function. J. Cell Sci. 2008, 121, 762–772. [Google Scholar] [CrossRef] [Green Version]
  230. Kiyonaka, S.; Kato, K.; Nishida, M.; Mio, K.; Numaga, T.; Sawaguchi, Y.; Yoshida, T.; Wakamori, M.; Mori, E.; Numata, T.; et al. Selective and Direct Inhibition of TRPC3 Channels Underlies Biological Activities of a Pyrazole Compound. Proc. Natl. Acad. Sci. USA 2009, 106, 5400–5405. [Google Scholar] [CrossRef] [Green Version]
  231. Miller, M.; Shi, J.; Zhu, Y.; Kustov, M.; Tian, J.; Stevens, A.; Wu, M.; Xu, J.; Long, S.; Yang, P.; et al. Identification of ML204, a Novel Potent Antagonist That Selectively Modulates Native TRPC4/C5 Ion Channels. J. Biol. Chem. 2011, 286, 33436–33446. [Google Scholar] [CrossRef] [Green Version]
  232. Bauer, C.C.; Minard, A.; Pickles, I.B.; Simmons, K.J.; Chuntharpursat-Bon, E.; Burnham, M.P.; Kapur, N.; Beech, D.J.; Muench, S.P.; Wright, M.H.; et al. Xanthine-Based Photoaffinity Probes Allow Assessment of Ligand Engagement by TRPC5 Channels. RSC Chem. Biol. 2020, 1, 436–448. [Google Scholar] [CrossRef]
  233. Just, S.; Chenard, B.L.; Ceci, A.; Strassmaier, T.; Chong, J.A.; Blair, N.T.; Gallaschun, R.J.; del Camino, D.; Cantin, S.; D’Amours, M.; et al. Treatment with HC-070, a Potent Inhibitor of TRPC4 and TRPC5, Leads to Anxiolytic and Antidepressant Effects in Mice. PLoS ONE 2018, 13, e0191225. [Google Scholar] [CrossRef]
  234. Rubaiy, H.N.; Ludlow, M.J.; Henrot, M.; Gaunt, H.J.; Miteva, K.; Cheung, S.Y.; Tanahashi, Y.; Hamzah, N.; Musialowski, K.E.; Blythe, N.M.; et al. Picomolar, Selective, and Subtype-Specific Small-Molecule Inhibition of TRPC1/4/5 Channels. J. Biol. Chem. 2017, 292, 8158–8173. [Google Scholar] [CrossRef] [Green Version]
  235. Minard, A.; Bauer, C.C.; Wright, D.J.; Rubaiy, H.N.; Muraki, K.; Beech, D.J.; Bon, R.S. Remarkable Progress with Small-Molecule Modulation of TRPC1/4/5 Channels: Implications for Understanding the Channels in Health and Disease. Cells 2018, 7, 52. [Google Scholar] [CrossRef] [Green Version]
  236. Yu, M.; Ledeboer, M.W.; Daniels, M.; Malojcic, G.; Tibbitts, T.T.; Coeffet-Le Gal, M.; Pan-Zhou, X.-R.; Westerling-Bui, A.; Beconi, M.; Reilly, J.F.; et al. Discovery of a Potent and Selective TRPC5 Inhibitor, Efficacious in a Focal Segmental Glomerulosclerosis Model. ACS Med. Chem. Lett. 2019, 10, 1579–1585. [Google Scholar] [CrossRef]
  237. Vinayagam, D.; Quentin, D.; Yu-Strzelczyk, J.; Sitsel, O.; Merino, F.; Stabrin, M.; Hofnagel, O.; Yu, M.; Ledeboer, M.W.; Nagel, G.; et al. Structural Basis of TRPC4 Regulation by Calmodulin and Pharmacological Agents. eLife 2020, 9, e60603. [Google Scholar] [CrossRef]
  238. Maier, T.; Follmann, M.; Hessler, G.; Kleemann, H.-W.; Hachtel, S.; Fuchs, B.; Weissmann, N.; Linz, W.; Schmidt, T.; Löhn, M.; et al. Discovery and Pharmacological Characterization of a Novel Potent Inhibitor of Diacylglycerol-Sensitive TRPC Cation Channels. Br. J. Pharmacol. 2015, 172, 3650–3660. [Google Scholar] [CrossRef]
  239. Merritt, J.E.; Armstrong, W.P.; Benham, C.D.; Hallam, T.J.; Jacob, R.; Jaxa-Chamiec, A.; Leigh, B.K.; McCarthy, S.A.; Moores, K.E.; Rink, T.J. SK&F 96365, a Novel Inhibitor of Receptor-Mediated Calcium Entry. Biochem. J. 1990, 271, 515–522. [Google Scholar] [CrossRef] [Green Version]
  240. Washburn, D.G.; Holt, D.A.; Dodson, J.; McAtee, J.J.; Terrell, L.R.; Barton, L.; Manns, S.; Waszkiewicz, A.; Pritchard, C.; Gillie, D.J.; et al. The Discovery of Potent Blockers of the Canonical Transient Receptor Channels, TRPC3 and TRPC6, Based on an Anilino-Thiazole Pharmacophore. Bioorg. Med. Chem. Lett. 2013, 23, 4979–4984. [Google Scholar] [CrossRef]
  241. Seo, K.; Rainer, P.P.; Shalkey Hahn, V.; Lee, D.; Jo, S.-H.; Andersen, A.; Liu, T.; Xu, X.; Willette, R.N.; Lepore, J.J.; et al. Combined TRPC3 and TRPC6 Blockade by Selective Small-Molecule or Genetic Deletion Inhibits Pathological Cardiac Hypertrophy. Proc. Natl. Acad. Sci. USA 2014, 111, 1551–1556. [Google Scholar] [CrossRef] [Green Version]
  242. Lin, B.L.; Matera, D.; Doerner, J.F.; Zheng, N.; del Camino, D.; Mishra, S.; Bian, H.; Zeveleva, S.; Zhen, X.; Blair, N.T.; et al. In Vivo Selective Inhibition of TRPC6 by Antagonist BI 749327 Ameliorates Fibrosis and Dysfunction in Cardiac and Renal Disease. Proc. Natl. Acad. Sci. USA 2019, 116, 10156–10161. [Google Scholar] [CrossRef] [Green Version]
  243. Häfner, S.; Burg, F.; Kannler, M.; Urban, N.; Mayer, P.; Dietrich, A.; Trauner, D.; Broichhagen, J.; Schaefer, M. A (+)-Larixol Congener with High Affinity and Subtype Selectivity toward TRPC6. ChemMedChem 2018, 13, 1028–1035. [Google Scholar] [CrossRef]
  244. Bai, Y.; Yu, X.; Chen, H.; Horne, D.; White, R.; Wu, X.; Lee, P.; Gu, Y.; Ghimire-Rijal, S.; Lin, D.C.-H.; et al. Structural Basis for Pharmacological Modulation of the TRPC6 Channel. eLife 2020, 9, e53311. [Google Scholar] [CrossRef]
  245. Wen, L.; Voronina, S.; Javed, M.A.; Awais, M.; Szatmary, P.; Latawiec, D.; Chvanov, M.; Collier, D.; Huang, W.; Barrett, J.; et al. Inhibitors of ORAI1 Prevent Cytosolic Calcium-Associated Injury of Human Pancreatic Acinar Cells and Acute Pancreatitis in 3 Mouse Models. Gastroenterology 2015, 149, 481–492. [Google Scholar] [CrossRef] [Green Version]
  246. Bruen, C.; Miller, J.; Wilburn, J.; Mackey, C.; Bollen, T.L.; Stauderman, K.; Hebbar, S. Auxora for the Treatment of Patients With Acute Pancreatitis and Accompanying Systemic Inflammatory Response Syndrome. Pancreas 2021, 50, 537–543. [Google Scholar] [CrossRef]
  247. Miller, J.; Bruen, C.; Schnaus, M.; Zhang, J.; Ali, S.; Lind, A.; Stoecker, Z.; Stauderman, K.; Hebbar, S. Auxora versus Standard of Care for the Treatment of Severe or Critical COVID-19 Pneumonia: Results from a Randomized Controlled Trial. Crit. Care 2020, 24, 502. [Google Scholar] [CrossRef]
  248. Barde, P.J.; Viswanadha, S.; Veeraraghavan, S.; Vakkalanka, S.V.; Nair, A. A First-in-Human Study to Evaluate the Safety, Tolerability and Pharmacokinetics of RP3128, an Oral Calcium Release-Activated Calcium (CRAC) Channel Modulator in Healthy Volunteers. J. Clin. Pharm. Ther. 2021, 46, 677–687. [Google Scholar] [CrossRef]
  249. PRCL Research Inc. A Phase 2a Study to Evaluate Safety, Tolerability, and Efficacy of PRCL-02 in Patients with Moderate to Severe Chronic Plaque Psoriasis; Clinicaltrials.gov: 2020. Available online: https://clinicaltrials.gov/ct2/show/NCT03614078 (accessed on 26 November 2021).
  250. Walsh, L.; Reilly, J.F.; Cornwall, C.; Gaich, G.A.; Gipson, D.S.; Heerspink, H.J.L.; Johnson, L.; Trachtman, H.; Tuttle, K.R.; Farag, Y.M.K.; et al. Safety and Efficacy of GFB-887, a TRPC5 Channel Inhibitor, in Patients With Focal Segmental Glomerulosclerosis, Treatment-Resistant Minimal Change Disease, or Diabetic Nephropathy: TRACTION-2 Trial Design. Kidney Int. Rep. 2021, 6, 2575–2584. [Google Scholar] [CrossRef]
  251. Li, J.; Zhang, X.; Song, X.; Liu, R.; Zhang, J.; Li, Z. The Structure of TRPC Ion Channels. Cell Calcium 2019, 80, 25–28. [Google Scholar] [CrossRef]
  252. Zhao, Y.; McVeigh, B.M.; Moiseenkova-Bell, V.Y. Structural Pharmacology of TRP Channels. J. Mol. Biol. 2021, 433, 166914. [Google Scholar] [CrossRef]
Figure 1. Topology and key domain organization of STIM1. A STIM1 monomer consists of an N-terminal luminal part, a transmembrane domain and a C-terminal cytosolic part. CC, coiled-coil; CAD, CRAC activation domain or SOAR, STIM–Orai activation region; EF hand, helix-loop structural domain; PBD, polybasic domain; SAM, sterile alpha motif; TM, transmembrane domain. Numbers correspond to the location of amino acid.
Figure 1. Topology and key domain organization of STIM1. A STIM1 monomer consists of an N-terminal luminal part, a transmembrane domain and a C-terminal cytosolic part. CC, coiled-coil; CAD, CRAC activation domain or SOAR, STIM–Orai activation region; EF hand, helix-loop structural domain; PBD, polybasic domain; SAM, sterile alpha motif; TM, transmembrane domain. Numbers correspond to the location of amino acid.
Biomolecules 11 01781 g001
Figure 2. Orai1 topology and key domain for interaction with STIM1. (A) The Orai1 monomer consists of an ETON domain in the N-terminal part, four transmembrane domains (TM1-TM4) and a C-terminal domain. ETON, extended transmembrane Orai1 N-terminal. (B) Schematic assembly of Orai1 subunits as a hexamer.
Figure 2. Orai1 topology and key domain for interaction with STIM1. (A) The Orai1 monomer consists of an ETON domain in the N-terminal part, four transmembrane domains (TM1-TM4) and a C-terminal domain. ETON, extended transmembrane Orai1 N-terminal. (B) Schematic assembly of Orai1 subunits as a hexamer.
Biomolecules 11 01781 g002
Figure 3. General topology and key domains of TRPC channels. (A) The TRPC subunit contains four to five ankyrin-like repeats on the N-terminal side, six transmembrane domains (TM1-TM6), a TRP domain and a CIRB domain on the C-terminal side. Ank, ankyrin-like repeats; CIRB, calmodulin–IP3R binding site. (B) Schematic assembly of TRPC subunits as a homotetramer or heterotetramer.
Figure 3. General topology and key domains of TRPC channels. (A) The TRPC subunit contains four to five ankyrin-like repeats on the N-terminal side, six transmembrane domains (TM1-TM6), a TRP domain and a CIRB domain on the C-terminal side. Ank, ankyrin-like repeats; CIRB, calmodulin–IP3R binding site. (B) Schematic assembly of TRPC subunits as a homotetramer or heterotetramer.
Biomolecules 11 01781 g003
Figure 4. Comparison between ICRAC and ISOC in HEK cells transfected with STIM1 + Orai1. Cells were transfected with Orai1 and STIM1 plasmids 24 h before electrophysiological recordings. (A) Representative ICRAC produced by STIM1+Orai1, inward rectifier current with almost no inward potential. In more than 60% of cells displaying ICRAC. (B) Representative ISOC produced by mixing STIM1+Orai1 and endogenous TRPC channels (linear current with inward potential close to 0 mV). In almost 40% of cells displaying ISOC. Patch-clamp recordings were performed in whole-cell mode using electrodes of resistance between 2.5–3 MΩ. The pipette solution contained (in mM) 130 caesium methansulfonate, 8 MgCl2, 10 HEPES, 10 BAPTA, and 0.001 thapsigargin, pH 7.2 (CsOH). Extracellular solution contained (in mM): 135 NaCl, 10 CsCl, 4.4 MgCl2, 2.8 KCl, 10 Hepes, 0.5 EDTA, 0.5 EGTA, 10 glucose, and 11 CaCl2, pH 7.4 (NaOH). After measuring the maximal current amplitude, 10 μM Gd3+ was added at the end of the recording to block the current and estimate the leak. Currents were recorded during a 1 s ramp of potentials ranging from −130 mV to +85 mV applied with low-pass filtering at 1 kHz.
Figure 4. Comparison between ICRAC and ISOC in HEK cells transfected with STIM1 + Orai1. Cells were transfected with Orai1 and STIM1 plasmids 24 h before electrophysiological recordings. (A) Representative ICRAC produced by STIM1+Orai1, inward rectifier current with almost no inward potential. In more than 60% of cells displaying ICRAC. (B) Representative ISOC produced by mixing STIM1+Orai1 and endogenous TRPC channels (linear current with inward potential close to 0 mV). In almost 40% of cells displaying ISOC. Patch-clamp recordings were performed in whole-cell mode using electrodes of resistance between 2.5–3 MΩ. The pipette solution contained (in mM) 130 caesium methansulfonate, 8 MgCl2, 10 HEPES, 10 BAPTA, and 0.001 thapsigargin, pH 7.2 (CsOH). Extracellular solution contained (in mM): 135 NaCl, 10 CsCl, 4.4 MgCl2, 2.8 KCl, 10 Hepes, 0.5 EDTA, 0.5 EGTA, 10 glucose, and 11 CaCl2, pH 7.4 (NaOH). After measuring the maximal current amplitude, 10 μM Gd3+ was added at the end of the recording to block the current and estimate the leak. Currents were recorded during a 1 s ramp of potentials ranging from −130 mV to +85 mV applied with low-pass filtering at 1 kHz.
Biomolecules 11 01781 g004
Figure 5. Activation mechanism of store-operated Ca2+ channels. Depletion of intraluminal Ca2+ concentration triggers Ca2+ dissociation of the EF hand of STIM1, allowing elongation and oligomerization of STIM1 at the ER–PM junction. This allows the interaction of STIM1 with Orai1 and/or TRPC channels, allowing Ca2+ entry into the cytoplasm. ADP, adenosine diphosphate; ATP, adenosine triphosphate; ER, endoplasmic reticulum; GPCR, G-protein coupled receptor; IP3R, inositol 1,4,5-trisphosphate receptor; PM, plasma membrane; SERCA, sarcoplasmic/endoplasmic reticulum Ca2+−ATPase; TRPC, transient receptor potential canonical channels.
Figure 5. Activation mechanism of store-operated Ca2+ channels. Depletion of intraluminal Ca2+ concentration triggers Ca2+ dissociation of the EF hand of STIM1, allowing elongation and oligomerization of STIM1 at the ER–PM junction. This allows the interaction of STIM1 with Orai1 and/or TRPC channels, allowing Ca2+ entry into the cytoplasm. ADP, adenosine diphosphate; ATP, adenosine triphosphate; ER, endoplasmic reticulum; GPCR, G-protein coupled receptor; IP3R, inositol 1,4,5-trisphosphate receptor; PM, plasma membrane; SERCA, sarcoplasmic/endoplasmic reticulum Ca2+−ATPase; TRPC, transient receptor potential canonical channels.
Biomolecules 11 01781 g005
Table 1. TRPC channels properties.
Table 1. TRPC channels properties.
ChannelSelectivity
PCa/PNa
Conductance (pS)Activation MechanismReferences
TRPC1116Store depletion, GPCR–PLC pathway, membrane stretching,[60,95,96]
TRPC22.742DAG[97]
TRPC31.666DAG, store depletion, PKC phosphorylation, membrane stretching[73,94,98,99,100]
TRPC41.1–7.742Store depletion, GPCR–PLC pathway[101,102,103]
TRPC51.8–9.563Store depletion, GPCR–PLC pathway[101,103,104]
TRPC6535DAG, membrane stretching[77,97,105]
TRPC71.925–50DAG, store depletion[50,94,106]
DAG, diacylglycerol; GPCR, G-protein coupled receptor; PCa/PNa, permeability ratio between Ca2+ and Na+; PKC, protein kinase C; PLC, phospholipase C.
Table 2. TRPC channels expression in PASMC.
Table 2. TRPC channels expression in PASMC.
TRPC1TRPC3TRPC4TRPC5TRPC6References
ExpressionmRNAProteinmRNAProteinmRNAProteinmRNAProteinmRNAProtein
Human--[141,142,143,144,145,146]
Rat-[147,148,149,150]
Mouse---[133,151,152,153,154,155,156]
-, no data available.
Table 3. Changes in SOC expression (mRNA and protein) and function in PASMC from PAH and from experimental models of PH.
Table 3. Changes in SOC expression (mRNA and protein) and function in PASMC from PAH and from experimental models of PH.
mRNAProteinSOCEhPASMCRodents ModelsConsequencesReferences
STIM1--Hypoxia Increased proliferation of rat PASMC
Increased NFATc3 nuclear translocation
[140,173]
- Rat CH Distal PA
STIM2-PAH Increased proliferation of iPAH hPASMC[136,173]
--Hypoxia
Orai1--Hypoxia [134,173]
Rat CH PA
Rat CH PASMC
Orai2--Hypoxia [134,173]
Rat CH PA
Rat CH PASMC
TRPC1 Rat CH/MCT PA
Rat CH PASMC
Reduced CH-induced PH phenotype and PASMC proliferation in trpc1−/− mice
Pharmacological inhibition normalized vascular tone in PAs of CH-induced PH rats
Increased murine PASMC proliferation
[145,152,153,185,186]
-- Mouse PASMC exposed to hypoxia
TRPC3PAH [173]
TRPC4 Rat SU/Hx PA
Rat SU/Hx PASMC
trpc4 gene deletion reduces PH[186,187]
TRPC6-PAH hPASMC proliferation
trpc6 gene deletion reduces CH-induced PH in mice
Pharmacological inhibition normalized vascular tone in PAs of CH-induced PH rats
[145,146,152,173]
--Hypoxia
Rat CH PA and PASMC
SOCE--PAH Increased STIM1–Orai1–TRPC1 interaction in hypoxic mouse PASMC[151,157,172]
-- Mouse PASMC exposed to hypoxia
The upwards pointing arrow shows an increase; -, no available information.
Table 4. SOCs inhibitors used as research tool.
Table 4. SOCs inhibitors used as research tool.
CompoundMode of ActionIC50Side EffectsReferences
Orai1 Inhibitor
YM-58483 (BTP2 or Pyr2)-10–590 nMInhibits TRPC3 and -C6 (IC50: 0.3 µM)
Activates TRPM4 channels (EC50: 8 nM)
Inhibit Orai2 and Orai3 at 10 µM
[214,215,216,217,218,219]
GSK7975A and GSK-5503APotentially allosteric effect on the selectivity filter of Orai4 µMOrai2 and Orai3 at 10 µM, L-type Ca2+ (IC50: 8 µM), and TRPV6 channels[219,220]
Synta-66Binds TM1 and TM3 helices and the extracellular loop segments26 nM-3 µMPotentiate Orai2 at 10 µM[219,221,222,223,224]
JPIII-244 nM-[225]
Gd3+ or La3+Binds extracellular loop of Orai1200 nMInhibit Orai2/3[219,226]
AnCoA4Binds the C-terminus of Orai1880 nM-[227]
5J-4-10 µM-[228]
STIM1 Inhibitor
ML-9Inhibit STIM1 puncta formation10 µMInhibit Myosin light chain kinase[229]
TRPC3 Inhibitor
Pyr3Direct binding0.7 μM-[230]
Pyr10-0.72 µM-[218]
TRPC4 Inhibitor
ML-204-1–3 μMInhibit TRPC5 and weakly TRPC6[231]
HC-070Direct binding9.3 nMInhibit TRPC4 (IC50:46 nM) and
TRPC3 (IC50: 1 µM)
[232,233]
HC-608 (Pico145)-0.35 nMInhibit TRPC5 (IC50: 1.3 nM), TRPC1-4 complex (IC50: 0.03 nM) and TRPC1-5 complex (IC50: 0.2 nM)[234]
TRPC5 Inhibitor
AC1903-13.6 μMInhibit TRPC4 (IC50 > 100 µM)[235]
GFB-8438Direct binding0.18 µMInhibit TRPC4 (IC50: 0.29 µM)[236,237]
TRPC6 Inhibitor
SAR7334-9.5 nMInhibit TRPC3 (IC50: 282 nM) and
TRPC7 (IC50: 226 nM)
[238]
SKF-96365-10 µM-[239]
GSK2833503-3 nMInhibit TRPC3 (IC50: 21 nM)[240,241]
BI 749327-19 nM-[242]
SH045-7.9 nMInhibit TRPC3 (IC50: 282 nM) and
TRPC7 (IC50: 226 nM)
[243]
AM-1473Direct binding0.22 nM-[244]
Inhibitors are classified according to their smallest IC50. EC50, half maximal effective concentration; IC50, half maximal inhibitory concentration; -, no data available.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Masson, B.; Montani, D.; Humbert, M.; Capuano, V.; Antigny, F. Role of Store-Operated Ca2+ Entry in the Pulmonary Vascular Remodeling Occurring in Pulmonary Arterial Hypertension. Biomolecules 2021, 11, 1781. https://doi.org/10.3390/biom11121781

AMA Style

Masson B, Montani D, Humbert M, Capuano V, Antigny F. Role of Store-Operated Ca2+ Entry in the Pulmonary Vascular Remodeling Occurring in Pulmonary Arterial Hypertension. Biomolecules. 2021; 11(12):1781. https://doi.org/10.3390/biom11121781

Chicago/Turabian Style

Masson, Bastien, David Montani, Marc Humbert, Véronique Capuano, and Fabrice Antigny. 2021. "Role of Store-Operated Ca2+ Entry in the Pulmonary Vascular Remodeling Occurring in Pulmonary Arterial Hypertension" Biomolecules 11, no. 12: 1781. https://doi.org/10.3390/biom11121781

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