Open access peer-reviewed chapter

Regulation of Calcium Signaling by STIM1 and ORAI1

Written By

Francisco Javier Martin-Romero, Carlos Pascual-Caro, Aida Lopez- Guerrero, Noelia Espinosa-Bermejo and Eulalia Pozo-Guisado

Submitted: 01 February 2018 Reviewed: 10 May 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.78587

From the Edited Volume

Calcium and Signal Transduction

Edited by John N. Buchholz and Erik J. Behringer

Chapter metrics overview

1,633 Chapter Downloads

View Full Metrics

Abstract

STIM1 and ORAI1 proteins are regulators of intracellular Ca2+ mobilization. This Ca2+ mobilization is essential to shape Ca2+ signaling in eukaryotic cells. STIM1 is a transmembrane protein located at the endoplasmic reticulum, where it acts as an intraluminal Ca2+ sensor. The transient drop of intraluminal Ca2+ concentration triggers STIM1 activation, which relocates to plasma membrane-endoplasmic reticulum junctions to bind and activate ORAI1, a plasma membrane Ca2+ channel. Thus, the Ca2+ influx pathway mediated by STIM1/ORAI1 is termed store-operated Ca2+ entry (SOCE). STIM and ORAI proteins are also involved in non-SOCE Ca2+ influx pathways, as we discuss here. In this chapter, we review the current knowledge regarding the role of SOCE, STIM1, and ORAI1 in cell signaling, with special focus on the modulation of the activity of kinases, phosphatases, and transcription factors that are strongly influenced by the extracellular Ca2+ influx mediated by these regulators.

Keywords

  • calcium
  • signaling
  • SOCE
  • STIM
  • ORAI

1. Introduction

Cell signaling is the network of reactions and interaction of molecules that allow cells to react to a wide range of stimuli. In this response, many pathways are involved, so cells are able to adapt to changing conditions. One of the mechanisms to respond to external stimuli is mediated by receptors, that is, proteins located at the plasma membrane that communicate the extracellular and the intracellular medium. A significant strategy that cells acquired early in their evolution was the modification of the composition of the intracellular milieu, so the ionic composition is different across the plasma membrane. This strategy is expensive in terms of the consumption of energy, since the ionic composition of the intracellular medium is modified by pumping out some ions from the cytosol. However, this is cost-efficient because it provided the possibility to proliferate and to gain cellular specialization. In this regard, free calcium (Ca2+) concentration in the cytosol of cells is much lower than that observed in the external medium, so there are mechanisms to remove the excess of free Ca2+ from the cytosol, such as extruding Ca2+ to the extracellular medium or to intracellular Ca2+ stores. This pumping is carried out by plasma membrane Ca2+ pumps and by endoplasmic reticulum Ca2+ pumps, respectively. Also, buffering of Ca2+ with Ca2+-binding proteins is another strategy to keep cytosolic free Ca2+ concentration ([Ca2+]i) within the low nanomolar range (~100 nM). The reason why the [Ca2+]i is tightly controlled is because this level is a second messenger in cell signaling, that is, transient variations of [Ca2+]i communicate a signal to be transmitted. For instance, during fertilization of mammalian oocytes, a series of short-term cytosolic increases of [Ca2+]i occurs in the oocyte for ~20 h after the fusion with sperm. These transient and short spikes are required to release the arrest of the cell cycle and to stimulate the transition from the fertilized oocyte to 1-cell embryo (zygote). The level of [Ca2+]i is also involved in many other cellular events, like the control of gene expression, vesicular trafficking, neurotransmitter release, cytoskeletal dynamics, and so on.

Cytosolic Ca2+ spikes and Ca2+ waves are generated by the opening of Ca2+-specific ion channels located at the plasma membrane and subcellular organelles. When they become activated, plasma membrane Ca2+ channels let the influx of extracellular Ca2+ so the [Ca2+]i rapidly increases, triggering the activation of Ca2+-sensitive effectors. As the main intracellular Ca2+ store, the endoplasmic reticulum (ER) also contains Ca2+ channels that become activated upon certain stimuli to let the transient release of Ca2+ to the cytosol. Then, elevated [Ca2+]i activates Ca2+ pumps to reduce the level of free Ca2+ in the cytosol, making possible the temporal increase of [Ca2+]i which is essential for its role as a messenger. The speed of the Ca2+ rise, as well as the Ca2+ removal, together with the time that this elevation lasts, define the temporal Ca2+ signaling, or Ca2+ signature, a critical point in the activation of subsequent events. Similarly, the specific distribution of Ca2+ channels and pumps define the spatial Ca2+ signature. The spatiotemporal control of the Ca2+ signaling is relevant for determining the regulation of different signaling pathways that finally lead to diverse actions. In summary, it is not only important to know how Ca2+ levels are altered upon specific stimuli, but also their specific duration, shape, and subcellular localization.

In this chapter, we summarize the current knowledge regarding the role of the STIM and ORAI proteins family. Because of their role as ER intraluminal Ca2+ sensors, STIM proteins have been recently involved in the modulation of several Ca2+-dependent signaling pathways. ORAI proteins are Ca2+ channels located at the plasma membrane that regulate the influx of Ca2+, in some cases under the control of STIM proteins. Thus, cooperation of both proteins is critical for Ca2+ influx, Ca2+ signaling, and cell physiology.

Advertisement

2. General overview of STIM and ORAI proteins

In humans, there are two different genes coding for STIM proteins: STIM1 and STIM2. STIM1 gene shows three known transcriptional variants that generate the proteins STIM1 (canonical), STIM1L (the longest isoform), and STIM1S (the shortest isoform). For STIM2 gene, also three transcriptional variants have been described coding for proteins STIM2, STIM2.1 (or STIM2 beta), and STIM2.2 (or STIM2 alpha) (see Table 1).

Gene Transcript(s) Protein Protein official name
ENSG00000167323 NM_001277961.1 NP_001264890.1 STIM1 isoform 1, or STIM1L
NM_003156.3 NP_003147.2 STIM1 isoform 2 (canonical)
NM_001277962.1 NP_001264891.1 STIM1 isoform 3, or STIMS
ENSG00000109689 NM_001169117.1 NP_001162588.1 STIM2 isoform 3
NM_001169118.1 NP_001162589.1 STIM2.1, STIM2β
NM_020860.3 NP_065911.3 STIM2.2, STIM2α
ENSG00000276045 NM_032790.3 NP_116179.2 ORAI1
ENSG00000160991 NM_001126340.2 NP_001119812.1 ORAI2 isoform a
NM_001271818.1 NP_001258747.1 ORAI2 isoform a
NM_001271819.1 NP_001258748.1 ORAI2 isoform b
NM_032831.3 NP_116220.1 ORAI2 isoform a
ENSG00000175938 NM_152288.2 NP_689501.1 ORAI3

Table 1.

Accession number for genes and reference sequences (RefSeq) of transcriptional variants and proteins.

Also in humans, three different genes code for ORAI proteins: ORAI1, ORAI2, and ORAI3. ORAI1 gene yields a single product (ORAI1 protein, also known as calcium release-activated calcium channel protein 1), whereas ORAI2 gene produces two variants (isoforms 1 and 2), and ORAI3 gene generates a single transcriptional variant and a single protein isoform (Table 1).

STIM1 protein is a positive regulator of store-operated Ca2+ entry (SOCE) [1, 2], a Ca2+ influx pathway regulated by the filling status of intracellular Ca2+ stores, mainly the ER. Although there is a significant pool of STIM1 at the plasma membrane, most STIM1 is ER-resident. When located at the ER, STIM1 shows a single transmembrane domain (TM) with the N-terminus toward the intraluminal space of this organelle. The Ca2+-sensitive EF-hand domain, together with a sterile-α-motif (SAM), constitute an intraluminal Ca2+ sensor, with an apparent dissociation constant for Ca2+ of 250 μM [3]. When the intraluminal Ca2+ concentration drops below this Kd, the dissociation of Ca2+ from the EF-hand domain is transmitted to the SAM domain, and to the cytosolic domain of the protein leading to its activation [4]. The cytosolic domain shows a well-studied calcium release-activated calcium (CRAC) activation domain (CAD), with a series of short coiled-coil (CC) domains that bind to ORAI1 plasma membrane channels to activate Ca2+ influx [5]. STIM1 protein also shows a Ser/Pro rich domain, close to a short sequence of four amino acids that binds to the microtubule plus-end binding protein EB1 [6], and finally a terminal Lys-rich domain which is critical for the activation of non-ORAI1 Ca2+ channels, such as TRPCs [7].

STIM2 and STIM1 share >60% sequence identity, and STIM2 also senses intraluminal Ca2+ concentration although with different sensitivity, since the dissociation constant for Ca2+(~500 μM) is twofold higher than that of STIM1 [8], suggesting that STIM2 becomes activated with smaller changes in intraluminal Ca2+ levels, whereas STIM1 activates Ca2+ entry upon more severe conditions [9].

ORAI1 is a plasma membrane protein with four transmembrane domains with the N- and C-termini oriented to the cytosol. The Ca2+ channel is formed by a hexamer of ORAI1 monomers, with the Ca2+ pore in the center of the hexamer [10, 11]. Both the N- and C-terminal domains are involved in the binding to STIM1 [12]. The paralogues ORAI2 and ORAI3 share 63% and 58% sequence identity with ORAI1, being the extracellular loop 3 that connects TM domain 3 and 4, significantly larger in ORAI3.

Advertisement

3. STIM1-ORAI1-mediated Ca2+ influx

The mechanism of activation has been well documented for the complex STIM1-ORAI1. At resting state, STIM1 is distributed on the surface of the ER, where its cytosolic domain is folded in a tight state due to an intramolecular clamp between domains CC1a1 and CC3 [13]. The inactive and resting STIM1 is a dimer [14], and the activation of Ca2+-release from intracellular stores due to the activation of the phosphoinositide pathway, leads to the transient depletion of Ca2+ levels within the ER. The consequent Ca2+ dissociation from the intraluminal domain of STIM1 triggers the activation of the protein in a more extended state that lets STIM1 to form oligomers [15]. In contrast to what it has been observed for inactive STIM1, which shows a high mobility on the ER surface while it is bound to EB1 and microtubules, STIM1 oligomers are quite immobile when they reach ER-PM junctions. This oligomerization has been extensively documented when STIM1 is targeted with fluorescent tags (Figure 1).

Figure 1.

HEK293 cells stably expressing STIM1-GFP were incubated in Hank’s balanced salt solution (HBSS) (left) or in Ca2+-free HBSS with 1 μM thapsigargin (right) to trigger Ca2+ store depletion. After 10-min incubation cells were fixed and visualized under wide-field fluorescence microscopy. In control cells (left panel), STIM1-GFP showed a localization that matched with endoplasmic reticulum. Thapsigargin induced aggregation of STIM1-GFP revealed by the clustering of GFP fluorescence (right panel), a result that demonstrated that the recombinant protein STIM1-GFP was sensitive to store depletion (reprinted from reference [25]).

The binding of STIM1 to the microtubule plus-end tracking protein EB1, ensures the targeting of STIM1 to ER-PM junctions [16]. However, this binding to EB1 is not required for the activation of ORAI1. STIM1 dissociates from EB1 by a mechanism regulated by the phosphorylation of a set of serine residues (Ser575, Ser608, and Ser621) adjacent to the EB1-binding site [17]. This STIM1 phosphorylation is mediated by the kinases ERK1/2, which become activated in the absence of Ca2+-influx by the activation of tyrosine kinase receptors at the plasma membrane [18, 19, 20]. Thus, STIM1 bound to EB1 travels to ER-PM junctions [6], but it dissociates from EB1 to bind to ORAI1 [17]. The physical interaction between STIM1 and ORAI1 is fully required for gating the Ca2+ channel. This interaction is mediated by the aforementioned CRAC activation domain (CAD) of STIM1 and both the cytosolic C-terminus and the N-terminus of ORAI1 [21], although the binding to the N-terminus is slightly weaker [22]. The stoichiometry of the complex STIM1-ORAI1 is also in the center of debate, but the current accepted proposal supports a 1:1 to 2:1 ratio in order to activate the channel [23, 24].

The high selectivity of ORAI1 for Ca2+ is due to the pore design, with a selectivity filter mediated by the acidic glutamate residue E106 at the first transmembrane domain [26]. Once Ca2+ influx is activated, a negative feedback controls the excessive Ca2+ entry, and Ca2+/calmodulin mediates this mechanism of inactivation. A short domain (residues 470–491) with seven acidic amino acids, close to the CAD binding domain, is directly involved in the Ca2+-dependent inactivation [27]. Also in this report, Mullins et al. identified a membrane-proximal N-terminal domain of ORAI1 (residues 68–91) that binds calmodulin (CaM) in a Ca2+-dependent manner [27], supporting a model in which Ca2+/CaM binds to the N-terminus of ORAI1 to trigger channel inactivation.

The large Rab GTPase CRACR2A mediates another mechanism that controls and prevents excessive Ca2+ entry. At low intracellular Ca2+ levels, CRACR2A enhances the binding of STIM1 to ORAI1, but at higher [Ca2+]i, that is, after ORAI1 activation, CRACR2A dissociates from ORAI1, inhibiting SOCE [28]. ORAI1 residues involved in the binding to CRACR2A are the same as those that bind Ca2+/CaM, so Ca2+/CaM and Ca2+-free CRACR2A are competitors for ORAI1.

STIMATE, a protein encoded by TMEM110 gene, is an ER-resident protein and a modulator for the activity of the STIM1-ORAI1 complex [29]. When Ca2+ dissociates from STIM1, the conformational change to a more extended structure of STIM1 facilitates the binding between STIMATE and STIM1-CC1 domain, avoiding the inhibition of CAD domain. This is the reason why STIMATE promotes the full extended conformation and the formation of STIM1 clustering at ER-PM junctions. In addition, the protein SARAF has been described as a negative regulator of SOCE [30]. As an ER membrane-resident protein, SARAF associates with STIM1 to promote Ca2+-dependent inactivation of SOCE. In this regard, a conserved STIM1 (448–530) C-terminal inhibitory domain (CTID) has been reported to regulate Ca2+-dependent inhibition [31]. CTID shows the capability to promote access of SARAF to the STIM1-ORAI1 activation region (SOAR or CAD), thus promoting inactivation of SOCE.

Additional regulators of the complex STIM1-ORAI1 have been reported, including septins [32] and RASSF4 [33]. Septin filaments and phosphatidylinositol-4,5-bisphosphate (PIP2) polarize in ER-PM junctions before store-depletion and facilitate STIM1 targeting to these junctions, where STIM1 recruits ORAI1. On the other hand, RASSF4 (RAS association domain family 4) also regulates SOCE by affecting the translocation of STIM1 to ER-PM junctions. Finally, a recent report has shown that ORAI1, as well as STIM1 phosphorylated at ERK1/2-target sites, are recruited at the leading edge of migrating cells, where ORAI1 binds cortactin, a regulator of plasma membrane ruffling [34]. This membrane ruffling is the reorganization of the cortical cytoskeleton required for the formation of filopodia and lamellipodia, and STIM1-KO (knockout) and ORAI1-KO cells, engineered by CRISPR/Cas9 genome editing, showed defective membrane ruffling and largely diminished cell migration [34], demonstrating that Ca2+ influx through STIM1-activated ORAI1 is essential for these events.

Advertisement

4. The role of STIM1 and ORAI1 on cell signaling

Given the importance of Ca2+ in many signaling pathways, the impact of STIM1, ORAI, and SOCE on cell signaling is also remarkable. Increasing evidence prove the significant role of this Ca2+ entry pathway in cell physiology and tissue homeostasis, and we focus here on the role of STIM and ORAI proteins on modulators of signaling pathways, such as kinases, phosphatases, and transcription factors. We also describe recent findings that unravel how STIM1 and ORAI1 are modulated by posttranslational modifications.

4.1. MAPK pathway

The close relationship between SOCE and mitogen-activated protein kinases (MAPKs) was revealed by Machaca and Haun [35], when they investigated the inactivation of SOCE in Xenopus oocyte maturation. SOCE is an active pathway in almost all eukaryotic cells, but during M-phase of cell cycle it becomes inactivated [36, 37, 38]. Machaca and Haun demonstrated that SOCE inactivation at germinal vesicle breakdown of Xenopus oocytes coincided with an increase in levels of MAPK and maturation-promoting factor (MPF), but they also demonstrated that MPF triggered SOCE inactivation by inhibiting the coupling between store depletion and SOCE activation, and not by blocking Ca2+ influx through SOCE channels [35].

In cells at interphase, some evidence supports a role for SOCE on ERK1/2 activation. In this regard, it has been proposed that SOCE activates extracellular signal-regulated kinases 1/2 (ERK1/2) in parotid acinar cells [39] and melanoma cells [40]. This proposal fits well with SOCE as an upstream regulator of ERK1/2. However, this proposal does not seem to be applicable to all cell lines, since ERK1/2 can be activated in the absence of extracellular Ca2+ and therefore in the absence of Ca2+ influx in HEK293 cells [20], the Ishikawa adenocarcinoma cell line [18], osteosarcoma U2OS cells [34], and prostate PC3 cells [19]. In addition, STIM1 knockdown did not modify phosphorylation of MEK1/2-ERK1/2 in gastric cancer cells [41], and STIM1 and ORAI1 knockdown did not inhibit the activation of ERK1/2 in response to EGF [42]. Moreover, ERK1/2 is fully activatable in STIM1-KO PC3 cells [19], with no active SOCE, demonstrating that SOCE is dispensable for ERK1/2 activation.

On the contrary, increasing evidence demonstrates that SOCE is a target for ERK1/2 activity, and that ERK1/2 is an upstream regulator of STIM1 and SOCE (reviewed in [43]). Pozo-Guisado et al. reported that STIM1 is phosphorylated by ERK1/2 at residues Ser575, Ser608, and Ser621 [25]. This phosphorylation is required for the full activation of STIM1 and for triggering the dissociation of STIM1 from microtubules [17]. Accordingly, phospho-STIM1 is enriched at the leading edge of migrating cells, that is, in the vicinity of receptor tyrosine kinases [34], where phospho-STIM1 acts in cooperation with ORAI1 to regulate the Ca2+ influx that rules cell migration. Consequently, phospho-STIM1 is an effector of ERK1/2 and an essential mediator for the activation of Ca2+ influx upon stimulation of cells with IGF-1 [20], or EGF [18, 19, 34].

Other MAPKs, such as p38 MAPK, have been shown to regulate SOCE, although by an indirect mechanism. Transforming growth factor beta (TGFβ) regulates megakaryocyte maturation and platelet formation by upregulating the expression of the serum-glucocorticoid inducible kinase SGK1 [44, 45], which is p38 MAPK-dependent [46]. SGK1 stimulates nuclear translocation of transcription factor NF-κB, which upregulates ORAI1 expression, increasing SOCE. This increase was demonstrated to be sensitive to p38 MAPK inhibition, SGK1 inhibition, and NF-κB inhibition, demonstrating the role of p38 MAPK in the upregulation of SOCE [47]. Another proposal was reported by Sundivakkam et al. [48], who reported that p38 MAPK directly phosphorylates STIM1. In this report, it was shown that pharmacological inhibition of p38 MAPK increased SOCE and that p38β knockdown prevented STIM1 phosphorylation and potentiated SOCE. However, this report did not identify the phosphorylated Ser/Thr residue(s), since the findings were based on the use of a phospho-Ser antibody [48], but not a site-specific phospho-specific antibody.

4.2. cAMP and PKA

The crosstalk between SOCE and cAMP-activated pathways has been investigated thoroughly since the molecular description of STIM1 and ORAI1. For instance, the Ca2+/CaM-stimulated adenylyl cyclase 8 (AC8) was found to be activated by SOCE and co-localized with STIM1 and ORAI1 in lipid rafts [49]. Interestingly, other authors found that lowering the concentration of free Ca2+ within the ER led to recruitment of adenylyl cyclases, enhancing the production of cAMP with the subsequent PKA activation, being this action independent of the [Ca2+]i [50]. Because activation of STIM1 and translocation to ER-PM junctions were required for coupling ER-Ca2+ depletion and adenylyl cyclase activity, without altering [Ca2+]i, those authors proposed the occurrence of a pathway termed store-operated cAMP signaling (SOcAMPS), a pathway that was later confirmed for other AC isoforms, such as AC3 [51]. More recently, it was confirmed that other Ca2+ channels, including TRPC1, were also involved in the activation of ACs [52]. It has been reported that STIM1 interacts with the plasma membrane adenylyl cyclase 6 to regulate melanogenesis [53], and this interaction is mediated by the Ser/Pro-rich C-terminal region of STIM1. These reports, together with the finding that ORAI1- and SOCE-deficient fibroblasts showed impaired cAMP production and cAMP-dependent signaling [54], strongly support the direct relationship between Ca2+ depletion at ER stores, STIM1 activation, and enhanced production of cAMP.

As for other pathways, the regulation between STIM1 and PKA seems to be reciprocal. In addition to the CRAC channel ORAI1, the plasma membrane-resident STIM1 activates store-independent arachidonic acid regulated Ca2+ (ARC) channels, and this activation depends on the phosphorylation of STIM1 at Thr389 by PKA, which requires the scaffold protein AKAP79 [55]. This phosphorylation triggers a structural change in the SOAR region of STIM1 (also known as CAD) being essential for the selective activation of ARC channels [56].

4.3. Other kinases and pathways

It is known that Ca2+-influx is upregulated by phosphoinositide 3-kinase (PI3K) signaling in platelets [57, 58]. Because PI3K signaling involves activation of SGK1, and this kinase has been shown to be a stimulator of ORAI1 expression [59], it is accepted that PI3K modulates SOCE by upregulation of the CRAC channel. In B16B6 melanoma cells, constitutive activation of Src and PKB/Akt was revealed to be due to the activation of SOCE in lipid rafts, which promoted Ca2+-dependence of the Src activity to trigger tumor signaling events [60], as it reported for lung metastasis of melanoma cells in a xenograft mouse model [61]. However, this is not shared by other cancer cells, as in prostate PC3 cells, with low levels of active ERK1/2 due to constitutive activation of PKB/Akt, Src is fully activatable in a Ca2+-independent manner by epidermal growth factor (EGF) [19].

Protein kinase C (PKC) phosphorylates ORAI1 at residues Ser27 and Ser30 [62]. More precisely, Kawasaki et al. demonstrated that the knockdown of the isoform PKCβ led to an increase of Ca2+ influx, and that recombinant PKC phosphorylated ORAI1 in vitro and in vivo at these two amino acids, an effect that inhibited the Ca2+ transport through ORAI1 [62]. No other phospho-residues have been characterized in detail in ORAI1. In airway smooth muscle cells, rottlerin, a PKCδ-selective inhibitor, reduced phorbol esters-triggered SOCE, without affecting total levels of STIM1 and ORAI1 [63]. However, the mechanism of this inhibition remains to be elucidated. PKC also inhibited SOCE in hepatocytes treated with amiodarone, an experimental design to mimic the accumulation of lipids during steatosis [64]. Because selective inhibition of PKC reversed SOCE to normal values, it was concluded that lipid accumulation triggers PKC-dependent SOCE impairment. Also, accumulation of palmitate is cytotoxic in kidney cells, and high levels of palmitate triggered Ca2+ depletion in the ER, in addition to mitochondrial stress. This depletion is antagonized by the inhibition of fatty acid transporters, inhibition of phospholipase C (PLC), and inhibition of PKC [65]. Once again, the mechanism that links PKC and the regulation of STIM1/ORAI1 remains elusive.

In 2011, Mungai et al. reported that hypoxia augmented cellular reactive oxygen species (ROS), without a significant alteration of energy charge values [66]. Hypoxia triggered an increase in [Ca2+]i, relocalization of STIM1 to ER-PM junctions, and phosphorylation of AMP-activated protein kinase (AMPK), in the absence of its upstream regulator LKB1 (liver kinase B1). These events were due to the ROS-dependent activation of ORAI1, which led to an increase in [Ca2+]i and activation of CaMKK2 (or CaMKKβ), an upstream activator of AMPK [66, 67]. Finally, a recent report from Yang et al. described how STIM1-ORAI1 mediated autophagy in endothelial progenitor cells exposed to oxidized low-density lipoprotein to mimic hypercholesterolemia. This treatment caused stimulation of Ca2+ influx mediated by STIM1-ORAI1, activation of CAMKK2 and decrease of mTOR activity with the subsequent activation of autophagy [68]. Similarly, in hepatocarcinoma cells, mitochondrial fission increased cytosolic Ca2+ levels that activated the NF-κB pathway, upregulating STIM1 expression and the subsequent SOCE [69]. The relative increase of [Ca2+]i also activated NFAT-dependent upregulation of Drp1, promoting a positive loop to rise levels of mitochondrial fission.

4.4. NF-κB, NFAT, CREB, and other transcription factors

As we mentioned above, SOCE is essential for platelet activation, and it is known the key role of ORAI1 in response to thrombin [70, 71]. ORAI1 transcripts were found significantly reduced in platelets and megakaryocytes from SGK1-KO mice, and transfection of megakaryocyte with constitutively active SGK1 increased phosphorylation of the IκB kinase (IKKα/β), which phosphorylates the inhibitor protein IκBα, promoting nuclear translocation of NF-κB subunit p65 [59]. In addition, Eylenstein et al. defined, by chromatin immunoprecipitation (ChIP) experiments, the promoter regions accounting for NF-κB-sensitive genomic regulation of STIM1 and ORAI1 [72], supporting further the conclusion that upregulation of ORAI1 and STIM1 by SGK1-dependent NF-κB signaling leads to the upregulation of SOCE, which in turn upregulates expression of other transcription factors like fibroblast growth factor 23 (FGF23) [73]. Other transcription factors are known to regulate STIM1 expression, such as Wilms tumor suppressor 1 (WT1) and early growth response 1 (EGR1), which were found by analyzing the STIM1 promoter with the TESS search system (University of Pennsylvania) [74]. Finally, NEUROD2, a neurogenic transcription factor, has been described as a negatively regulator of STIM1 expression, an activity that limits the level of STIM1 in cortical neurons [75].

Regarding NF-κB, reciprocal regulation seems to link this transcription factor and SOCE. In addition to the activation of SOCE by NF-κB described above, Liu et al. reported that T-cell activation triggered by the binding of antigen to T-cell receptor stimulated SOCE and that this Ca2+ entry activated a PKCα-mediated phosphorylation of p65 NF-κB at Ser536, an event that controls nuclear localization and transcriptional activity of NF-κB [76].

STIM1 and ORAI1 are also well-known activators of the protein phosphatase calcineurin, which activates the nuclear factor of activated T cells (NFAT) [77]. Once NFAT becomes dephosphorylated by calcineurin the transcription factor is internalized into the nucleus. Indeed, the activation and nuclear translocation of NFAT was the reporter used by Feske et al. when they searched for regulators of SOCE using a Drosophila RNA interference screening, a report that led to the description of ORAI1 as the channel that mediates the Ca2+ release-activated Ca2+ current, or CRAC [78]. Because NFAT modulates the expression of a wide range of genes, it is involved in many pathways, and also in the regulation of the expression of other regulators of transcription, including IRF4, BATF, and Bcl-6 [79]. NFAT is not the only transcription factor activated by the axis STIM1-ORAI1-Ca2+/CaM-calcineurin because Ca2+ influx through ORAI1 stimulates the transcription factor EB (TFEB), promoting the activation of chemokines genes [80]. SOCE also activates the Ca2+/cAMP response element binding protein (CREB), a transcription factor that regulates expression of many genes, at least in cultured smooth muscle cells and intact arteries [81]. In this regard, it was observed that mitochondrial Ca2+ uptake was reduced in lymphocytes lacking STIM1 or ORAI1, an effect that was due to reduced mitochondrial Ca2+ uniporter (MCU) expression [82]. ChIP and promoter analyses revealed that CREB directly binds the MCU promoter, revealing that SOCE regulates the Ca2+ uptake capability of mitochondria by regulating Ca2+-dependent activation of CREB [82].

Ca2+ influx regulates myoblasts differentiation, and shortly after the molecular description of STIM and ORAI proteins, it was reported that silencing STIM1, Orai1, or Orai3 reduced SOCE and myoblast differentiation [83]. This positive effect on myoblasts correlated with the expression of MEF2 and myogenin, two transcription factors involved in skeletal muscle development, although it is still unclear the molecular pathway that links STIM1/ORAI1 and the activation of the transcription factors. In cerebellar granule neurons cultured in low concentration of extracellular potassium, mimicking resting conditions, SOCE promoted the degradation of transcription factor Sp4, a regulator of neuronal morphogenesis and function [84].

Another important molecular interactor of STIM1 is the hypoxia-inducible factor-1 alpha (HIF-1α), which is upregulated during hepatocarcinoma growth [85]. Li et al. found that HIF-1α directly controls STIM1 transcription, but also that STIM1-mediated SOCE is required for HIF-1α accumulation in hepatocarcinoma cells via activation of Ca2+/CaM-dependent protein kinase II, revealing a mutual dependence of STIM1 and HIF-1α in the regulation of Ca2+ transport and tumor growth [85].

High levels of ORAI1 and STIM1 are found in many types of cancer cells. In gastric cancer tumor progression, this higher expression is associated with a negative impact on survival rates of patients, an effect that was partially due to targeting expression of metastasis-associated in colon cancer-1 (MACC1) [86], an essential regulator of the transcription for the gene coding for the hepatocyte growth factor receptor, MET. Similarly, a recent report described that STIM1 promotes cell migration and the epithelial-to-mesenchymal transition (EMT) by activating TGF-β, Snail and Wnt/β-catenin pathways in prostate cancer cells [87].

Finally, an excellent report from Stephan Feske laboratory [54], described how SOCE is crucial for mitochondrial fatty acid oxidation, and that Ca2+ entry through ORAI1 was essential to activate adenylyl cyclase, cyclic AMP production, the transcriptional regulator peroxisome proliferator-activated receptor gamma coactivator 1 alpha (PGC-1α) and peroxisome proliferator-activated receptor alpha (PPARα), which is mediated by the activation of CREB.

Advertisement

5. ORAI3 and calcium signaling

Shortly after the molecular description of ORAI1 as the CRAC channel, it was revealed the essential role of other ORAI proteins in cell signaling. The involvement of ORAI3, together with ORAI1, in arachidonic acid-regulated Ca2+ (ARC) channels was early proposed [88, 89]. In contrast to SOC channels, the activation of ARC channels depends on the pool of STIM1 resident in the plasma membrane [90]. More interestingly, ORAI1 and ORAI3 show a differential sensitivity to reactive oxygen species, due to the extracellularly located Cys195 residue which is found in ORAI1, but not in ORAI3. The differential redox sensitivity underlies the differential responses between naïve and T helper lymphocytes, an event that lets T(H) cells proliferate and secrete cytokines in oxidative environments [91].

ORAI3, but not ORAI1, was also involved in the activation of PLCδ in response to arachidonic acid, an activation that controls oscillation frequency of Ca2+ spikes triggered by carbachol [92]. ORAI3 channels are overexpressed in estrogen receptor-positive breast cancer cells [93], and it was later demonstrated, using the MCF-7 cancer cell line, that silencing ORAI3 slows down cell cycle and triggers arrest at G1 phase [94]. EGF triggers Ca2+ entry through ORAI3, and the channel is transcriptionally upregulated by the estrogen receptor alpha (ERα) [95]. It is now accepted that cancer cells show a remodeling of ORAI proteins, with an enhanced participation of ORAI3 compared to noncancerous cells, suggesting that heteromerization of ORAI3 and ORAI1 is a common feature in malignant transformation [96].

Advertisement

6. Future directions

During the past decade, a significant progress was made regarding the molecular description of the proteins involved in store-operated Ca2+ entry. Although some details remain unclear, a topology of STIM1-ORAI1 contact sites, selectivity filters in ORAI1, and the role of posttranslational modifications have been reported for both proteins. The involvement of STIM and ORAI proteins in different pathways is now much clearer, and they are now considered master regulators of Ca2+-dependent signaling pathways. However, in many cases pathways were studied in cancer cell lines in vitro, so physiological models are required to evaluate the importance of STIM1 and ORAI1 in the pathophysiology of cells in vivo. Nevertheless, primary cell cultures and established cell lines constitute a widely accepted experimental approach for basic studies in cell signaling and understanding the role of STIM1/ORAI1 in cell biology and cell signaling. With these tools, we have reached the conclusion that STIM1 and ORAI1 are involved in the control of Ca2+ refilling within the ER. More important, STIM1 and ORAI1 directly modulate Ca2+ signaling in a wide variety of pathways, with a significant role in gene expression, cell migration, and tumor cell metastasis (Figure 2). Because the expression of STIM1/ORAI1 is deregulated in cancer cells, it is required to evaluate the relative importance of STIM1/ORAI1 as pharmacological targets for the treatment of disease, not only with the use of in vitro cell cultures, but also in animal models for the study of human disease.

Figure 2.

Activation of SOCE by STIM1/ORAI1 and pathways involved in SOCE-dependent signaling. Panel A: diverse stimuli that triggers the activation of the phospholipase C pathway, such as activation of EGF receptor (EGFR), stimulate the production of inositol 1,4,5-trisphosphate (IP3) which binds and activates IP3 receptor (IP3R) at the endoplasmic reticulum (ER). This activation leads to the release of Ca2+ from the ER, with the subsequent transient depletion of intraluminal [Ca2+] and the activation of STIM1. Ca2+-unbound STIM1 aggregates in oligomers and translocates to plasma membrane (PM)-ER junctions where it binds and activates ORAI1. Extracellular Ca2+ entry through ORAI1 activates multiple Ca2+-dependent targets, as shown in panel B, but also provides a Ca2+ source to replenish intraluminal Ca2+ levels. This replenishment is accomplished by the ER-Ca2+-ATPase which pumps Ca2+ into the ER lumen. Panel B: schematic illustration of the most important pathways regulated by STIM1/ORAI1. AC, adenylyl cyclase; GPCR, G protein-coupled receptor; PM-STIM1, plasma membrane-resident STIM1; p-STIM1, phosphorylated STIM1.

Advertisement

Acknowledgments

The authors wish to acknowledge financial support from the Spanish Ministerio de Economía y Competitividad (Grant BFU2014-52401-P) and Junta de Extremadura (Grant IB16088). This financial support was co-funded by European Regional Development Funds.

Advertisement

Conflict of interest

The authors declare no conflict of interests.

References

  1. 1. Roos J, DiGregorio PJ, Yeromin AV, Ohlsen K, Lioudyno M, Zhang S, Safrina O, Kozak JA, Wagner SL, Cahalan MD, Veliçelebi G, Stauderman KA. STIM1, an essential and conserved component of store-operated Ca2+ channel function. The Journal of Cell Biology. 2005;169(3):435-445
  2. 2. Zhang SL, Yu Y, Roos J, Kozak JA, Deerinck TJ, Ellisman MH, Stauderman KA, Cahalan MD. STIM1 is a Ca2+ sensor that activates CRAC channels and migrates from the Ca2+ store to the plasma membrane. Nature. 2005;437(7060):902-905
  3. 3. Stathopulos PB, Li GY, Plevin MJ, Ames JB, Ikura M. Stored Ca2+ depletion-induced oligomerization of stromal interaction molecule 1 (STIM1) via the EF-SAM region: An initiation mechanism for capacitive Ca2+ entry. Journal of Biological Chemistry. 2006;281(47):35855-35862
  4. 4. Zheng L, Stathopulos PB, Schindl R, Li GY, Romanin C, Ikura M. Auto-inhibitory role of the EF-SAM domain of STIM proteins in store-operated calcium entry. Proceedings of the National Academy of Sciences of the United States of America. 2011;108(4):1337-1342
  5. 5. Huang GN, Zeng W, Kim JY, Yuan JP, Han L, Muallem S, Worley PF. STIM1 carboxyl-terminus activates native SOC, I(crac) and TRPC1 channels. Nature Cell Biology. 2006;8(9):1003-1010
  6. 6. Grigoriev I, Gouveia SM, van der Vaart B, Demmers J, Smyth JT, Honnappa S, Splinter D, Steinmetz MO, Putney JW Jr, Hoogenraad CC, Akhmanova A. STIM1 is a MT-plus-end-tracking protein involved in remodeling of the ER. Current Biology. 2008;18(3):177-182
  7. 7. Zeng W, Yuan JP, Kim MS, Choi YJ, Huang GN, Worley PF, Muallem S. STIM1 gates TRPC channels, but not Orai1, by electrostatic interaction. Molecular Cell. 2008;32(3):439-448
  8. 8. Zheng L, Stathopulos PB, Li GY, Ikura M. Biophysical characterization of the EF-hand and SAM domain containing Ca2+ sensory region of STIM1 and STIM2. Biochemical and Biophysical Research Communications. 2008;369(1):240-246
  9. 9. Brandman O, Liou J, Park WS, Meyer T. STIM2 is a feedback regulator that stabilizes basal cytosolic and endoplasmic reticulum Ca2+ levels. Cell. 2007;131(7):1327-1339
  10. 10. Hou X, Pedi L, Diver MM, Long SB. Crystal structure of the calcium release-activated calcium channel Orai. Science. 2012;338(6112):1308-1313
  11. 11. Yen M, Lokteva LA, Lewis RS. Functional analysis of Orai1 concatemers supports a hexameric stoichiometry for the CRAC channel. Biophysical Journal. 2016;111(9):1897-1907
  12. 12. McNally BA, Somasundaram A, Jairaman A, Yamashita M, Prakriya M. The C- and N-terminal STIM1 binding sites on Orai1 are required for both trapping and gating CRAC channels. The Journal of Physiology. 2013;591(11):2833-2850
  13. 13. Muik M, Fahrner M, Schindl R, Stathopulos P, Frischauf I, Derler I, Plenk P, Lackner B, Groschner K, Ikura M, Romanin C. STIM1 couples to ORAI1 via an intramolecular transition into an extended conformation. The EMBO Journal. 2011;30(9):1678-1689
  14. 14. Muik M, Fahrner M, Derler I, Schindl R, Bergsmann J, Frischauf I, Groschner K, Romanin C. A cytosolic homomerization and a modulatory domain within STIM1 C terminus determine coupling to ORAI1 channels. The Journal of Biological Chemistry. 2009;284(13):8421-8426
  15. 15. Liou J, Kim ML, Heo WD, Jones JT, Myers JW, Ferrell JE, Meyer T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Current Biology. 2005;15(13):1235-1241
  16. 16. Chang CL, Chen YJ, Quintanilla CG, Hsieh TS, Liou J. EB1 binding restricts STIM1 translocation to ER-PM junctions and regulates store-operated Ca2+ entry. Journal of Cell Biology. 2018;217(6):2047-2058
  17. 17. Pozo-Guisado E, Casas-Rua V, Tomas-Martin P, Lopez-Guerrero AM, Alvarez-Barrientos A, Martin-Romero FJ. Phosphorylation of STIM1 at ERK1/2 target sites regulates interaction with the microtubule plus-end binding protein EB1. Journal of Cell Science. 2013;126(Pt 14):3170-3180
  18. 18. Casas-Rua V, Tomas-Martin P, Lopez-Guerrero AM, Alvarez IS, Pozo-Guisado E, Martin-Romero FJ. STIM1 phosphorylation triggered by epidermal growth factor mediates cell migration. Biochimica et Biophysica Acta. 2015;1853(1):233-243
  19. 19. Lopez-Guerrero AM, Pascual-Caro C, Martin-Romero FJ, Pozo-Guisado E. Store-operated calcium entry is dispensable for the activation of ERK1/2 pathway in prostate cancer cells. Cellular Signalling. 2017;40:44-52
  20. 20. Tomas-Martin P, Lopez-Guerrero AM, Casas-Rua V, Pozo-Guisado E, Martin-Romero FJ. Phospho-STIM1 is a downstream effector that mediates the signaling triggered by IGF-1 in HEK293 cells. Cellular Signalling. 2015;27(3):545-554
  21. 21. Park CY, Hoover PJ, Mullins FM, Bachhawat P, Covington ED, Raunser S, Walz T, Garcia KC, Dolmetsch RE, Lewis RS. STIM1 clusters and activates CRAC channels via direct binding of a cytosolic domain to Orai1. Cell. 2009;136(5):876-890
  22. 22. Derler I, Plenk P, Fahrner M, Muik M, Jardin I, Schindl R, Gruber HJ, Groschner K, Romanin C. The extended transmembrane Orai1 N-terminal (ETON) region combines binding interface and gate for Orai1 activation by STIM1. The Journal of Biological Chemistry. 2013;288(40):29025-29034
  23. 23. Hoover PJ, Lewis RS. Stoichiometric requirements for trapping and gating of Ca2+ release-activated Ca2+ (CRAC) channels by stromal interaction molecule 1 (STIM1). Proceedings of the National Academy of Sciences of the United States of America. 2011;108(32):13299-13304
  24. 24. Stathopulos PB, Schindl R, Fahrner M, Zheng L, Gasmi-Seabrook GM, Muik M, Romanin C, Ikura M. STIM1/Orai1 coiled-coil interplay in the regulation of store-operated calcium entry. Nature Communications. 2013;4(2013):2963
  25. 25. Pozo-Guisado E, Campbell DG, Deak M, Alvarez-Barrientos A, Morrice NA, Alvarez IS, Alessi DR, Martin-Romero FJ. Phosphorylation of STIM1 at ERK1/2 target sites modulates store-operated calcium entry. Journal of Cell Science. 2010;123(Pt 18):3084-3093
  26. 26. Zhou Y, Meraner P, Kwon HT, Machnes D, Masatsugu OH, Zimmer J, Huang Y, Stura A, Rao A, Hogan PG. STIM1 gates the store-operated calcium channel ORAI1 in vitro. Nature Structural & Molecular Biology. 2010;17(1):112-116
  27. 27. Mullins FM, Park CY, Dolmetsch RE, Lewis RS. STIM1 and calmodulin interact with Orai1 to induce Ca2+-dependent inactivation of CRAC channels. Proceedings of the National Academy of Sciences of the United States of America. 2009;106(36):15495-15500
  28. 28. Srikanth S, Jung HJ, Kim KD, Souda P, Whitelegge J, Gwack Y. A novel EF-hand protein, CRACR2A, is a cytosolic Ca2+ sensor that stabilizes CRAC channels in T cells. Nature Cell Biology. 2010;12(5):436-446
  29. 29. Jing J, He L, Sun A, Quintana A, Ding Y, Ma G, Tan P, Liang X, Zheng X, Chen L, Shi X, Zhang SL, Zhong L, Huang Y, Dong MQ, Walker CL, Hogan PG, Wang Y, Zhou Y. Proteomic mapping of ER-PM junctions identifies STIMATE as a regulator of Ca2+ influx. Nature Cell Biology. 2015;17(10):1339-1347
  30. 30. Palty R, Raveh A, Kaminsky I, Meller R, Reuveny E. SARAF inactivates the store operated calcium entry machinery to prevent excess calcium refilling. Cell. 2012;149(2):425-438
  31. 31. Jha A, Ahuja M, Maléth J, Moreno CM, Yuan JP, Kim MS, Muallem S. The STIM1 CTID domain determines access of SARAF to SOAR to regulate Orai1 channel function. The Journal of Cell Biology. 2013;202(1):71-79
  32. 32. Sharma S, Quintana A, Findlay GM, Mettlen M, Baust B, Jain M, Nilsson R, Rao A, Hogan PG. An siRNA screen for NFAT activation identifies septins as coordinators of store-operated Ca2+ entry. Nature. 2013;499(7457):238-242
  33. 33. Chen YJ, Chang CL, Lee WR, Liou J. RASSF4 controls SOCE and ER-PM junctions through regulation of PI(4,5)P. The Journal of Cell Biology. 2017;216(7):2011-2025
  34. 34. Lopez-Guerrero AM, Tomas-Martin P, Pascual-Caro C, Macartney T, Rojas-Fernandez A, Ball G, Alessi DR, Pozo-Guisado E, Martin-Romero FJ. Regulation of membrane ruffling by polarized STIM1 and ORAI1 in cortactin-rich domains. Scientific Reports. 2017;7(1):383
  35. 35. Machaca K, Haun S. Induction of maturation-promoting factor during Xenopus oocyte maturation uncouples Ca2+ store depletion from store-operated Ca2+ entry. The Journal of Cell Biology. 2002;156(1):75-85
  36. 36. Volpi M, Berlin RD. Intracellular elevations of free calcium induced by activation of histamine H1 receptors in interphase and mitotic HeLa cells: Hormone signal transduction is altered during mitosis. The Journal of Cell Biology. 1988;107(6 Pt 2):2533-2539
  37. 37. Preston SF, Sha'afi RI, Berlin RD. Regulation of Ca2+ influx during mitosis: Ca2+ influx and depletion of intracellular Ca2+ stores are coupled in interphase but not mitosis. Cell Regulation. 1991;2(11):915-925
  38. 38. Machaca K, Haun S. Store-operated calcium entry inactivates at the germinal vesicle breakdown stage of Xenopus meiosis. The Journal of Biological Chemistry. 2000;275(49):38710-38715
  39. 39. Soltoff SP, Lannon WA. Activation of ERK1/2 by store-operated calcium entry in rat parotid acinar cells. PLoS One. 2013;8(8):e72881
  40. 40. Umemura M, Baljinnyam E, Feske S, De Lorenzo MS, Xie LH, Feng X, Oda K, Makino A, Fujita T, Yokoyama U, Iwatsubo M, Chen S, Goydos JS, Ishikawa Y, Iwatsubo K. Store-operated Ca2+ entry (SOCE) regulates melanoma proliferation and cell migration. PLoS One. 2014;9(2):e89292
  41. 41. Xu JM, Zhou Y, Gao L, Zhou SX, Liu WH, Li XA. Stromal interaction molecule 1 plays an important role in gastric cancer progression. Oncology Reports. 2016;35(6):3496-3504
  42. 42. Yang IH, Tsai YT, Chiu SJ, Liu LT, Lee HH, Hou MF, Hsu WL, Chen BK, Chang WC. Involvement of STIM1 and Orai1 in EGF-mediated cell growth in retinal pigment epithelial cells. Journal of Biomedical Science. 2013;20:41
  43. 43. Pozo-Guisado E, Martin-Romero FJ. The regulation of STIM1 by phosphorylation. Communicative & Integrative Biology. 2013;6(6):e26283
  44. 44. Waldegger S, Klingel K, Barth P, Sauter M, Rfer ML, Kandolf R, Lang F. h-sgk serine-threonine protein kinase gene as transcriptional target of transforming growth factor beta in human intestine. Gastroenterology. 1999;116(5):1081-1088
  45. 45. Al-Maghout T, Pelzl L, Sahu I, Sukkar B, Hosseinzadeh Z, Gutti R, Laufer S, Voelkl J, Pieske B, Gawaz M, Lang F. P38 kinase. SGK1 and NF-κB dependent up-regulation of Na+/Ca2+ exchanger expression and activity following TGFß1 treatment of megakaryocytes. Cellular Physiology and Biochemistry. 2017;42(6):2169-2181
  46. 46. Waerntges S, Klingel K, Weigert C, Fillon S, Buck M, Schleicher E, Rodemann HP, Knabbe C, Kandolf R, Lang F. Excessive transcription of the human serum and glucocorticoid dependent kinase hSGK1 in lung fibrosis. Cellular Physiology and Biochemistry. 2002;12(2-3):135-142
  47. 47. Yan J, Schmid E, Almilaji A, Shumilina E, Borst O, Laufer S, Gawaz M, Lang F. Effect of TGFβ on calcium signaling in megakaryocytes. Biochemical and Biophysical Research Communications. 2015;461(1):8-13
  48. 48. Sundivakkam PC, Natarajan V, Malik AB, Tiruppathi C. Store-operated Ca2+ entry (SOCE) induced by protease-activated receptor-1 mediates STIM1 protein phosphorylation to inhibit SOCE in endothelial cells through AMP-activated protein kinase and p38β mitogen-activated protein kinase. The Journal of Biological Chemistry. 2013;288(23):17030-17041
  49. 49. Martin AC, Willoughby D, Ciruela A, Ayling LJ, Pagano M, Wachten S, Tengholm A, Cooper DM. Capacitative Ca2+ entry via Orai1 and stromal interacting molecule 1 (STIM1) regulates adenylyl cyclase type 8. Molecular Pharmacology. 2009;75(4):830-842
  50. 50. Lefkimmiatis K, Srikanthan M, Maiellaro I, Moyer MP, Curci S, Hofer AM. Store-operated cyclic AMP signalling mediated by STIM1. Nature Cell Biology. 2009;11(4):433-442
  51. 51. Maiellaro I, Lefkimmiatis K, Moyer MP, Curci S, Hofer AM. Termination and activation of store-operated cyclic AMP production. Journal of Cellular and Molecular Medicine. 2012;16(11):2715-2725
  52. 52. Willoughby D, Ong HL, De Souza LB, Wachten S, Ambudkar IS, Cooper DM. TRPC1 contributes to the Ca2+-dependent regulation of adenylate cyclases. The Biochemical Journal. 2014;464(1):73-84
  53. 53. Motiani RK, Tanwar J, Raja DA, Vashisht A, Khanna S, Sharma S, Srivastava S, Sivasubbu S, Natarajan VT, Gokhale RS. STIM1 activation of adenylyl cyclase 6 connects Ca2+ and cAMP signaling during melanogenesis. The EMBO Journal. 2018;37(5):e97597. DOI: 10.15252/embj.201797597
  54. 54. Maus M, Cuk M, Patel B, Lian J, Ouimet M, Kaufmann U, Yang J, Horvath R, Hornig-Do HT, Chrzanowska-Lightowlers ZM, Moore KJ, Cuervo AM, Feske S. Store-operated Ca2+ entry controls induction of lipolysis and the transcriptional reprogramming to lipid metabolism. Cell Metabolism. 2017;25(3):698-712
  55. 55. Thompson JL, Shuttleworth TJ. Anchoring protein AKAP79-mediated PKA phosphorylation of STIM1 determines selective activation of the ARC channel, a store-independent Orai channel. The Journal of Physiology. 2015;593(3):559-572
  56. 56. Thompson JL, Zhao Y, Stathopulos PB, Grossfield A, Shuttleworth TJ. Phosphorylation-mediated structural changes within the SOAR domain of stromal interaction molecule 1 enable specific activation of distinct Orai channels. The Journal of Biological Chemistry. 2018;293(9):3145-3155
  57. 57. Chen J, De S, Damron DS, Chen WS, Hay N, Byzova TV. Impaired platelet responses to thrombin and collagen in AKT-1-deficient mice. Blood. 2004;104(6):1703-1710
  58. 58. Lian L, Wang Y, Draznin J, Eslin D, Bennett JS, Poncz M, Wu D, Abrams CS. The relative role of PLCbeta and PI3Kgamma in platelet activation. Blood. 2005;106(1):110-117
  59. 59. Borst O, Schmidt EM, Münzer P, Schönberger T, Towhid ST, Elvers M, Leibrock C, Schmid E, Eylenstein A, Kuhl D, May AE, Gawaz M, Lang F. The serum- and glucocorticoid-inducible kinase 1 (SGK1) influences platelet calcium signaling and function by regulation of Orai1 expression in megakaryocytes. Blood. 2012;119(1):251-261
  60. 60. Fedida-Metula S, Feldman B, Koshelev V, Levin-Gromiko U, Voronov E, Fishman D. Lipid rafts couple store-operated Ca2+ entry to constitutive activation of PKB/Akt in a Ca2+/calmodulin-, Src- and PP2A-mediated pathway and promote melanoma tumor growth. Carcinogenesis. 2012;33(4):740-750
  61. 61. Sun J, Lu F, He H, Shen J, Messina J, Mathew R, Wang D, Sarnaik AA, Chang WC, Kim M, Cheng H, Yang S. STIM1- and Orai1-mediated Ca2+ oscillation orchestrates invadopodium formation and melanoma invasion. The Journal of Cell Biology. 2014;207(4):535-548
  62. 62. Kawasaki T, Ueyama T, Lange I, Feske S, Saito N. Protein kinase C-induced phosphorylation of Orai1 regulates the intracellular Ca2+ level via the store-operated Ca2+ channel. The Journal of Biological Chemistry. 2010;285(33):25720-25730
  63. 63. Gao Y, Zou J, Geng S, Zheng J, Yang J. Role of protein kinase C in the activation of store-operated Ca2+ entry in airway smooth muscle cells. Journal of Huazhong University of Science and Technology. Medical Sciences. 2012;32(3):303-310
  64. 64. Wilson CH, Ali ES, Scrimgeour N, Martin AM, Hua J, Tallis GA, Rychkov GY, Barritt GJ. Steatosis inhibits liver cell store-operated Ca2+ entry and reduces ER Ca2+ through a protein kinase C-dependent mechanism. The Biochemical Journal. 2015;466(2):379-390
  65. 65. Xu S, Nam SM, Kim JH, Das R, Choi SK, Nguyen TT, Quan X, Choi SJ, Chung CH, Lee EY, Lee IK, Wiederkehr A, Wollheim CB, Cha SK, Park KS. Palmitate induces ER calcium depletion and apoptosis in mouse podocytes subsequent to mitochondrial oxidative stress. Cell Death & Disease. 2015;6:e1976
  66. 66. Mungai PT, Waypa GB, Jairaman A, Prakriya M, Dokic D, Ball MK, Schumacker PT. Hypoxia triggers AMPK activation through reactive oxygen species-mediated activation of calcium release-activated calcium channels. Molecular and Cellular Biology. 2011;31(17):3531-3545
  67. 67. Gusarova GA, Trejo HE, Dada LA, Briva A, Welch LC, Hamanaka RB, Mutlu GM, Chandel NS, Prakriya M, Sznajder JI. Hypoxia leads to Na,K-ATPase downregulation via Ca2+ release-activated Ca2+ channels and AMPK activation. Molecular and Cellular Biology. 2011;31(17):3546-3556
  68. 68. Yang J, Yu J, Li D, Yu S, Ke J, Wang L, Wang Y, Qiu Y, Gao X, Zhang J, Huang L. Store-operated calcium entry-activated autophagy protects EPC proliferation via the CAMKK2-MTOR pathway in ox-LDL exposure. Autophagy. 2017;13(1):82-98
  69. 69. Huang Q, Cao H, Zhan L, Sun X, Wang G, Li J, Guo X, Ren T, Wang Z, Lyu Y, Liu B, An J, Xing J. Mitochondrial fission forms a positive feedback loop with cytosolic calcium signaling pathway to promote autophagy in hepatocellular carcinoma cells. Cancer Letters. 2017;403:108-118
  70. 70. Galán C, Zbidi H, Bartegi A, Salido GM, Rosado JA. STIM1, Orai1 and hTRPC1 are important for thrombin- and ADP-induced aggregation in human platelets. Archives of Biochemistry and Biophysics. 2009;490(2):137-144
  71. 71. Gilio K, van Kruchten R, Braun A, Berna-Erro A, Feijge MA, Stegner D, van der Meijden PE, Kuijpers MJ, Varga-Szabo D, Heemskerk JW, Nieswandt B. Roles of platelet STIM1 and Orai1 in glycoprotein VI- and thrombin-dependent procoagulant activity and thrombus formation. The Journal of Biological Chemistry. 2010;285(31):23629-23638
  72. 72. Eylenstein A, Schmidt S, Gu S, Yang W, Schmid E, Schmidt EM, Alesutan I, Szteyn K, Regel I, Shumilina E, Lang F. Transcription factor NF-κB regulates expression of pore-forming Ca2+ channel unit, Orai1, and its activator, STIM1, to control Ca2+ entry and affect cellular functions. The Journal of Biological Chemistry. 2012;287(4):2719-2730
  73. 73. Zhang B, Yan J, Umbach AT, Fakhri H, Fajol A, Schmidt S, Salker MS, Chen H, Alexander D, Spichtig D, Daryadel A, Wagner CA, Föller M, Lang F. NFκB-sensitive Orai1 expression in the regulation of FGF23 release. Journal of Molecular Medicine (Berlin, Germany). 2016;94(5):557-566
  74. 74. Ritchie MF, Yue C, Zhou Y, Houghton PJ, Soboloff J. Wilms tumor suppressor 1 (WT1) and early growth response 1 (EGR1) are regulators of STIM1 expression. The Journal of Biological Chemistry. 2010;285(14):10591-10596
  75. 75. Guner G, Guzelsoy G, Isleyen FS, Sahin GS, Akkaya C, Bayam E, Kotan EI, Kabakcioglu A, Ince-Dunn G. NEUROD2 regulates Stim1 expression and store-operated calcium entry in cortical neurons. eNeuro. 2017;4(1):ENEURO.0255-16.2017
  76. 76. Liu X, Berry CT, Ruthel G, Madara JJ, MacGillivray K, Gray CM, Madge LA, McCorkell KA, Beiting DP, Hershberg U, May MJ, Freedman BD. T cell receptor-induced nuclear factor κB (NF-κB) signaling and transcriptional activation are regulated by STIM1- and Orai1-mediated calcium entry. The Journal of Biological Chemistry. 2016;291(16):8440-8452
  77. 77. Gwack Y, Feske S, Srikanth S, Hogan PG, Rao A. Signalling to transcription: Store-operated Ca2+ entry and NFAT activation in lymphocytes. Cell Calcium. 2007;42(2):145-156
  78. 78. Feske S, Gwack Y, Prakriya M, Srikanth S, Puppel SH, Tanasa B, Hogan PG, Lewis RS, Daly M, Rao A. A mutation in Orai1 causes immune deficiency by abrogating CRAC channel function. Nature. 2006;441(7090):179-185
  79. 79. Vaeth M, Eckstein M, Shaw PJ, Kozhaya L, Yang J, Berberich-Siebelt F, Clancy R, Unutmaz D, Feske S. Store-operated Ca2+ entry in follicular T cells controls humoral immune responses and autoimmunity. Immunity. 2016;44(6):1350-1364
  80. 80. Zhu ZD, Yu T, Liu HJ, Jin J, He J. SOCE induced calcium overload regulates autophagy in acute pancreatitis via calcineurin activation. Cell Death and Disease. 2018;9(2):50
  81. 81. Pulver RA, Rose-Curtis P, Roe MW, Wellman GC, Lounsbury KM. Store-operated Ca2+ entry activates the CREB transcription factor in vascular smooth muscle. Circulation Research. 2004;94(10):1351-1358
  82. 82. Shanmughapriya S, Rajan S, Hoffman NE, Zhang X, Guo S, Kolesar JE, Hines KJ, Ragheb J, Jog NR, Caricchio R, Baba Y, Zhou Y, Kaufman BA, Cheung JY, Kurosaki T, Gill DL, Madesh M. Ca2+ signals regulate mitochondrial metabolism by stimulating CREB-mediated expression of the mitochondrial Ca2+ uniporter gene MCU. Science Signaling. 2015;8(366):ra23
  83. 83. Darbellay B, Arnaudeau S, Konig S, Jousset H, Bader C, Demaurex N, Bernheim L. STIM1- and Orai1-dependent store-operated calcium entry regulates human myoblast differentiation. The Journal of Biological Chemistry. 2009;284(8):5370-5380
  84. 84. Lalonde J, Saia G, Gill G. Store-operated calcium entry promotes the degradation of the transcription factor Sp4 in resting neurons. Science Signaling. 2014;7(328):ra51
  85. 85. Li Y, Guo B, Xie Q, Ye D, Zhang D, Zhu Y, Chen H, Zhu B. STIM1 mediates hypoxia-driven hepatocarcinogenesis via interaction with HIF-1. Cell Reports. 2015;12(3):388-395
  86. 86. Xia J, Wang H, Huang H, Sun L, Dong S, Huang N, Shi M, Bin J, Liao Y, Liao W. Elevated Orai1 and STIM1 expressions upregulate MACC1 expression to promote tumor cell proliferation, metabolism, migration, and invasion in human gastric cancer. Cancer Letters. 2016;381(1):31-40
  87. 87. Xu Y, Zhang S, Niu H, Ye Y, Hu F, Chen S, Li X, Luo X, Jiang S, Liu Y, Chen Y, Li J, Xiang R, Li N. STIM1 accelerates cell senescence in a remodeled microenvironment but enhances the epithelial-to-mesenchymal transition in prostate cancer. Scientific Reports. 2015;5:11754
  88. 88. Mignen O, Thompson JL, Shuttleworth TJ. Both Orai1 and Orai3 are essential components of the arachidonate-regulated Ca2+-selective (ARC) channels. The Journal of Physiology. 2008;586(1):185-195
  89. 89. Mignen O, Thompson JL, Shuttleworth TJ. The molecular architecture of the arachidonate-regulated Ca2+-selective ARC channel is a pentameric assembly of Orai1 and Orai3 subunits. 2009;The Journal of Physiology 587(Pt 17):4181-4197
  90. 90. Thompson JL, Shuttleworth TJ. A plasma membrane-targeted cytosolic domain of STIM1 selectively activates ARC channels, an arachidonate-regulated store-independent Orai channel. Channels (Austin, Tex.). 2012;6(5):370-378
  91. 91. Bogeski I, Kummerow C, Al-Ansary D, Schwarz EC, Koehler R, Kozai D, Takahashi N, Peinelt C, Griesemer D, Bozem M, Mori Y, Hoth M, Niemeyer BA. Differential redox regulation of ORAI ion channels: A mechanism to tune cellular calcium signaling. Science Signaling. 2010;3(115):ra24
  92. 92. Thompson JL, Shuttleworth TJ. Orai channel-dependent activation of phospholipase C-δ: A novel mechanism for the effects of calcium entry on calcium oscillations. The Journal of Physiology. 2011;589(Pt 21):5057-5069
  93. 93. Motiani RK, Abdullaev IF, 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. The Journal of Biological Chemistry. 2010;285(25):19173-19183
  94. 94. Faouzi M, Hague F, Potier M, Ahidouch A, Sevestre H, Ouadid-Ahidouch H. Down-regulation of Orai3 arrests cell-cycle progression and induces apoptosis in breast cancer cells but not in normal breast epithelial cells. Journal of Cellular Physiology. 2011;226(2):542-551
  95. 95. Motiani RK, Zhang X, Harmon KE, Keller RS, Matrougui K, Bennett JA, Trebak M. Orai3 is an estrogen receptor α-regulated Ca2+ channel that promotes tumorigenesis. The FASEB Journal. 2013;27(1):63-75
  96. 96. Dubois C, Vanden Abeele F, Lehen'kyi V, Gkika D, Guarmit B, Lepage G, Slomianny C, Borowiec AS, Bidaux G, Benahmed M, Shuba Y, Prevarskaya N. Remodeling of channel-forming ORAI proteins determines an oncogenic switch in prostate cancer. Cancer Cell. 2014;26(1):19-32

Written By

Francisco Javier Martin-Romero, Carlos Pascual-Caro, Aida Lopez- Guerrero, Noelia Espinosa-Bermejo and Eulalia Pozo-Guisado

Submitted: 01 February 2018 Reviewed: 10 May 2018 Published: 05 November 2018