Next Article in Journal
High-Capacity Adenoviral Vectors: Expanding the Scope of Gene Therapy
Next Article in Special Issue
Nicotinamide Mononucleotide Administration Prevents Experimental Diabetes-Induced Cognitive Impairment and Loss of Hippocampal Neurons
Previous Article in Journal
Wisteria floribunda Agglutinin-Positive Mac-2 Binding Protein but not α-fetoprotein as a Long-Term Hepatocellular Carcinoma Predictor
Previous Article in Special Issue
Regulation of Vascular Function and Inflammation via Cross Talk of Reactive Oxygen and Nitrogen Species from Mitochondria or NADPH Oxidase—Implications for Diabetes Progression
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Structural Mechanisms of Store-Operated and Mitochondrial Calcium Regulation: Initiation Points for Drug Discovery

1
Department of Physiology and Pharmacology, Schulich School of Medicine and Dentistry, The University of Western Ontario, London, ON N6A5C1, Canada
2
Department of Medicine, Schulich School of Medicine and Dentistry, The University of Western Ontario, London, ON N6A5C1, Canada
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2020, 21(10), 3642; https://doi.org/10.3390/ijms21103642
Submission received: 13 April 2020 / Revised: 11 May 2020 / Accepted: 17 May 2020 / Published: 21 May 2020
(This article belongs to the Special Issue Mitochondria-Targeted Approaches in Health and Disease)

Abstract

:
Calcium (Ca2+) is a universal signaling ion that is essential for the life and death processes of all eukaryotes. In humans, numerous cell stimulation pathways lead to the mobilization of sarco/endoplasmic reticulum (S/ER) stored Ca2+, resulting in the propagation of Ca2+ signals through the activation of processes, such as store-operated Ca2+ entry (SOCE). SOCE provides a sustained Ca2+ entry into the cytosol; moreover, the uptake of SOCE-mediated Ca2+ by mitochondria can shape cytosolic Ca2+ signals, function as a feedback signal for the SOCE molecular machinery, and drive numerous mitochondrial processes, including adenosine triphosphate (ATP) production and distinct cell death pathways. In recent years, tremendous progress has been made in identifying the proteins mediating these signaling pathways and elucidating molecular structures, invaluable for understanding the underlying mechanisms of function. Nevertheless, there remains a disconnect between using this accumulating protein structural knowledge and the design of new research tools and therapies. In this review, we provide an overview of the Ca2+ signaling pathways that are involved in mediating S/ER stored Ca2+ release, SOCE, and mitochondrial Ca2+ uptake, as well as pinpoint multiple levels of crosstalk between these pathways. Further, we highlight the significant protein structures elucidated in recent years controlling these Ca2+ signaling pathways. Finally, we describe a simple strategy that aimed at applying the protein structural data to initiating drug design.

Graphical Abstract

1. Cytosolic and Stored Calcium

Regulating cytosolic calcium ion (Ca2+) levels is fundamental for cellular life. Under resting conditions, intracellular Ca2+ concentrations are maintained at ~100 nM by effectively segregating Ca2+ into the extracellular fluid and within intracellular stores, such as the mitochondria and sarco/endoplasmic reticulum (S/ER). Under basal conditions, Ca2+ levels within the cytosol and mitochondria are similar; conversely, in times of cellular stress or injury, mitochondria can uptake Ca2+ [1,2,3,4] in order to prevent cytosolic Ca2+ overload [5,6,7]. Persistently high levels of mitochondrial Ca2+ would lead to a cycle of higher respiratory chain activity and the generation of reactive oxygen species (ROS), resulting in oxidative stress, apoptosis, and/or necrosis [8,9]. Ca2+ is a cofactor that is required for mitochondrial dehydrogenase function, which are enzymes integral to the electron transport chain [10,11]. The primary mode of mitochondrial Ca2+ uptake is via the low affinity, high capacity mitochondrial Ca2+ uniporter (MCU) complex [1,12,13]. However, recent evidence suggests that the bi-directional leucine-zipper EF-hand containing transmembrane protein-1 (LETM1) Ca2+/proton (H+) antiporter is involved in fine-tuning these levels [14,15]. Thus, the MCU complex and LETM1 are emerging as two critical regulators of mitochondrial Ca2+ uptake, dictating both mitochondrial and cellular function.
The S/ER is the largest intracellular Ca2+ store, maintaining free Ca2+ in the ~400–700 µM range (depending on cell-type) [16,17,18,19,20,21], which can be released into the cytosol in order to mediate and/or trigger a multitude of downstream cellular processes [22]. In excitable and non-excitable cells, ER-membrane bound ryanodine receptors (RyRs) and inositol 1,4,5-trisphosphate receptors (IP3Rs), respectively, serve as the predominant channels that release Ca2+ into the cytosol from the S/ER stores [23,24,25]. A subsequent entry of extracellular Ca2+ is critical for the activation of Ca2+-dependent signaling cascades that require more chronic elevations in cytosolic Ca2+ because S/ER Ca2+ stores are limited in capacity. Store-operated Ca2+ entry (SOCE) is the process whereby S/ER Ca2+ store depletion leads to the activation of plasma membrane (PM) Ca2+ channels for a sustained Ca2+ entry into the cytosol from the extracellular space [26,27,28,29]. The principal molecular machinery mediating SOCE include the Ca2+-sensing S/ER-membrane inserted stromal interaction molecules (STIMs) [30,31] and the pore-forming Orai1 channels on the PM [32,33,34,35,36]. The elevated cytosolic Ca2+ levels that are driven by SOCE not only signal downstream physiological and pathophysiological cellular responses, but also replenish the depleted intracellular Ca2+ stores.

1.1. Mobilization of S/ER Calcium Stores

Across human cell types, several mechanisms of extracellular signaling exist, which ultimately culminate in the stimulation or sensitization of IP3Rs and/or RyRs, leading to the depletion of the S/ER Ca2+ stores [23,24,37]. Such extracellular signaling pathways can involve G-protein-coupled receptors (GPCRs), receptor tyrosine kinases (RTKs), T- and B-cell receptors (TCRs, BCRs), or voltage-gated ion channels (VGICs), presenting a variety of mechanisms for S/ER Ca2+ store release (Table 1), which activate Ca2+-dependent transcription factors, structural proteins, and enzymes that are critical for cellular function.
GPCRs constitute a major family of PM receptors that, when activated by the appropriate extracellular ligand, can serve to mediate the release of Ca2+ from the S/ER into the cytosol via IP3Rs [37,42]. Upon extracellular ligand binding-induced activation, GPCRs activate intracellular phospholipases, such as Cβ or Cγ, which cleave phosphatidylinositol 4,5-bisphosphate into diacylglycerol and inositol 1,4,5-trisphosphate (IP3) [42]. IP3 is a small diffusible second messenger that binds and activates IP3Rs, causing S/ER Ca2+ store release into the cytosol [65]. This IP3R-driven Ca2+ signaling mechanism is critical in platelets where thrombin binding to protease activated receptor (PAR)-1 or PAR4 mediates thrombosis [41], in vascular endothelial cells where bradykinin binding to the bradykinin receptors (BR)-1 or BR2 causes vasodilation [66], and in airway smooth muscle cells where histamine binding to the histamine H1 receptor (H1R) leads to bronchoconstriction [67], to name a few examples.
Nevertheless, it is not only GPCRs that can induce S/ER Ca2+ release through IP3R activation. RTKs, such as insulin growth factor-1 receptor in skeletal myoblasts activate phospholipase C and IP3 production upon binding insulin-like growth factor-1, leading to ER Ca2+ release necessary for differentiation in myogenesis [68,69]. In similar mechanisms, when the membrane immunoglobulin (Ig) component of a BCR, such as IgM-BCR, interacts with an antigen or when T-cells bind self-antigen presented by thymic epithelial cells via TCRs, phosphorylation cascades result in phospholipase C and IP3R activation, enabling humoral immunity signaling and apoptosis/elimination of self-reactive T-cells, respectively [51,70]. Likewise, this TCR signaling pathway is vital to mediating immune responses after interacting with foreign antigens [16,17]. Immunologically, antigen-antibody complexes can also bind to different types of Fc receptors on mast cells, macrophages, dendritic cells or natural killer cells, inducing phospholipase C-mediated IP3 generation and S/ER Ca2+ release [16,17,55,56].
RyRs are much larger structural homologues of IP3Rs and similarly function to release S/ER Ca2+ [24]. For example, in cardiomyocytes, VGICs on the PM (i.e., L-type Ca2+ channels) open in response to membrane depolarization, resulting in the entry of Ca2+ into the cytosol from the extracellular space [60,61]. This cytosol entered Ca2+ binds to S/ER RyRs, opening these channels, in contrast to IP3Rs, which require IP3. Remarkably, the RyR-mediated release of S/ER Ca2+ can propagate to adjacent channels and promote additional RyR activation in a process that is known as Ca2+-induced Ca2+ release [60,61].

1.2. Store-Operated Calcium Entry

SOCE is a ubiquitous Ca2+ signaling pathway that has evolved to use S/ER Ca2+ depletion as an activation signal, facilitating further and more sustained extracellular Ca2+ influx in response to the release of intracellular S/ER Ca2+ [26,27,28,29]. Stromal interaction molecule-1 (STIM1) is a single pass, type 1 transmembrane protein of 685 amino acids and the S/ER luminal Ca2+ sensor and SOCE activator [30,31] (Figure 1A). The luminal amino (N)-terminal region of STIM1 contains an ER signal peptide, EF-hand domain, and a sterile-alpha motif (SAM) domain. Included in the cytosolic segment of STIM1 is a long stretch of coiled-coil domains, a Pro-Ser-rich region, and a polybasic tail [71]. The EF-hand, together with the SAM domain, sense S/ER Ca2+ changes (Figure 1B–E). At resting S/ER Ca2+ concentrations, EF-SAM is bound to Ca2+ and in a quiescent monomeric state [72,73]. Upon the dissociation of Ca2+ from EF-SAM, destabilization and dimerization of this luminal domain occurs, causing conformational changes that propagate through to the cytosolic region of the molecule [74,75,76]. These Ca2+-depletion-induced structural changes lead to STIM1 oligomerization and trapping at S/ER-PM junctions [77,78,79,80]. At these junctions, a region of the STIM1 coiled-coil domains directly interacts with and gates Orai1 protein-composed Ca2+ channels on the PM, permitting Ca2+ influx into the cytosol from the extracellular space [81,82,83,84] (Figure 2).
A process known as Ca2+-dependent inactivation (CDI) deactivates Orai1 channels to terminate SOCE [85,86]. Multiple sites on STIM1 and Orai1 have been suggested to provide a negative feedback signal to SOCE, both dependent and independent of calmodulin function [87,88,89,90]. STIMs, like many proteins that are involved in mediating signaling cascades, are often subject to post-translational modifications, which affect both structure and function [27,91,92,93,94,95,96,97,98,99,100,101,102,103]. S-Glutathionylation, for example, is the covalent attachment of a glutathione (GSH) moiety to free cysteine (Cys) thiols. GSH modification of STIM1 at Cys56 has been shown to evoke constitutive Orai1 channel activation and Ca2+ entry independent of luminal Ca2+ concentrations [104]. It is worth speculating that the thiols are vital regulatory sites evolved for receiving distinct inputs depending on specific environmental cues given the high conservation of the luminal STIM Cys residues among vertebrates (Figure 1A).
Figure 1. Structural elucidation of individual human STIM1 domains. (A) Domain architecture of human STIM1 [National Center for Biotechnology Information (NCBI) accession AFZ76986]. The highly conserved domains and regions with well-defined functional roles are indicated with residue ranges shown below the respective segment. The residue ranges for EF-SAM and the coiled-coils are from the solution structure and Coils analysis (using a 0.1 probability cutoff and 14, 21, and 28 residue combination) [105], respectively. All other residue range annotations are from Uniprot accession Q13586. The location of the luminal Cys residues that undergo post-translation modifications to modify Ca2+-sensing function are indicated (red spheres). The topological orientation relative to the S/ER is indicated above the diagram. (B) Solution NMR structure of Ca2+ loaded STIM1 EF-SAM. (C) Cluster of four CC1 monomers showing several different interfaces revealed by STIM1 CC1 crystallization. (D) Solution structure of the CC1-CC2 dimer in complex with two Orai1 C-terminal peptides (orange spheres). (E) Crystal structure of the dimerized STIM1-Orai activation region (SOAR). In BE, colours are shaded, as depicted in (A), and overlapping coiled-coil region colours are preserved between structures. The pdb coordinate files for STIM1 EF-SAM, CC1, CC1-CC2:Orai1, and SOAR are from 2K60 [75], 4O9B [106], 2MAK [107], and 3TEQ [108], respectively. SP, signal peptide; cEF, canonical EF-hand; nEF, non-canonical EF-hand; SAM, sterile alpha motif; TMD, transmembrane domain; CC1 -2 -3, coiled-coil -1 -2 -3; ID, inhibitory domain; PS, pro/ser-rich region; K, Lys-rich (polybasic) region; N, amino terminus; C, carboxyl terminus.
Figure 1. Structural elucidation of individual human STIM1 domains. (A) Domain architecture of human STIM1 [National Center for Biotechnology Information (NCBI) accession AFZ76986]. The highly conserved domains and regions with well-defined functional roles are indicated with residue ranges shown below the respective segment. The residue ranges for EF-SAM and the coiled-coils are from the solution structure and Coils analysis (using a 0.1 probability cutoff and 14, 21, and 28 residue combination) [105], respectively. All other residue range annotations are from Uniprot accession Q13586. The location of the luminal Cys residues that undergo post-translation modifications to modify Ca2+-sensing function are indicated (red spheres). The topological orientation relative to the S/ER is indicated above the diagram. (B) Solution NMR structure of Ca2+ loaded STIM1 EF-SAM. (C) Cluster of four CC1 monomers showing several different interfaces revealed by STIM1 CC1 crystallization. (D) Solution structure of the CC1-CC2 dimer in complex with two Orai1 C-terminal peptides (orange spheres). (E) Crystal structure of the dimerized STIM1-Orai activation region (SOAR). In BE, colours are shaded, as depicted in (A), and overlapping coiled-coil region colours are preserved between structures. The pdb coordinate files for STIM1 EF-SAM, CC1, CC1-CC2:Orai1, and SOAR are from 2K60 [75], 4O9B [106], 2MAK [107], and 3TEQ [108], respectively. SP, signal peptide; cEF, canonical EF-hand; nEF, non-canonical EF-hand; SAM, sterile alpha motif; TMD, transmembrane domain; CC1 -2 -3, coiled-coil -1 -2 -3; ID, inhibitory domain; PS, pro/ser-rich region; K, Lys-rich (polybasic) region; N, amino terminus; C, carboxyl terminus.
Ijms 21 03642 g001
Figure 2. Molecular mechanism of store-operated Ca2+ entry (SOCE) signaling. (A) With replete S/ER luminal Ca2+ stores, STIM1 (green cartoon) adopts a quiescent conformation. Quiescent STIM1 is dimerized via interactions between cytosolic domains [108,109,110] and remains in a compact conformation due to Ca2+ (yellow spheres) binding by EF-SAM, which keeps these luminal domains in monomeric states [72,73,75]. (B) Upon cell stimulation, via G protein-mediated PLC activation for example, IP3 is generated and diffuses to the IP3Rs (brown cartoon) on the S/ER membrane. IP3 binding to IP3Rs, opens these receptor Ca2+ channels, depleting the S/ER Ca2+ stores [28,111,112]. (C) Ca2+ unbinding from EF-SAM causes luminal domain dimerization [73,75,113], which rearranges the transmembrane regions [114] and promotes coiled-coil-1 interactions, culminating in a conformational extension of the STIM1 cytosolic region [74,76,115]. This structural change also promotes further oligomerization of STIMs via the cytosolic domains [110,115]. (D) Extended/oligomerized STIM1 gets trapped at S/ER-PM junctions [18,77,78,80] via interactions between the polybasic C-terminal regions and PM [78,95,116,117] as well as interactions between the STIM1 SOAR/CAD and Orai1 (orange cartoon) [82,83,107,118]. The direct interactions between STIM1 and Orai1 gate Orai1-composed hexameric Ca2+ release-activated Ca2+ (CRAC) channels [119,120], allowing Ca2+ to move from the extracellular space into the cytosol in SOCE. GPCR, G-protein coupled receptor; Gαq, G-alpha subunit; Gβγ, G-beta-gamma subunits; PLCβ, phospholipase Cβ; IP3, inositol trisphosphate; IP3R, IP3 receptor; STIM1, stromal interaction molecule-1; N, amino terminus; C, carboxyl terminus. This figure and the organelles depicted in the graphical abstract were created with BioRender.com.
Figure 2. Molecular mechanism of store-operated Ca2+ entry (SOCE) signaling. (A) With replete S/ER luminal Ca2+ stores, STIM1 (green cartoon) adopts a quiescent conformation. Quiescent STIM1 is dimerized via interactions between cytosolic domains [108,109,110] and remains in a compact conformation due to Ca2+ (yellow spheres) binding by EF-SAM, which keeps these luminal domains in monomeric states [72,73,75]. (B) Upon cell stimulation, via G protein-mediated PLC activation for example, IP3 is generated and diffuses to the IP3Rs (brown cartoon) on the S/ER membrane. IP3 binding to IP3Rs, opens these receptor Ca2+ channels, depleting the S/ER Ca2+ stores [28,111,112]. (C) Ca2+ unbinding from EF-SAM causes luminal domain dimerization [73,75,113], which rearranges the transmembrane regions [114] and promotes coiled-coil-1 interactions, culminating in a conformational extension of the STIM1 cytosolic region [74,76,115]. This structural change also promotes further oligomerization of STIMs via the cytosolic domains [110,115]. (D) Extended/oligomerized STIM1 gets trapped at S/ER-PM junctions [18,77,78,80] via interactions between the polybasic C-terminal regions and PM [78,95,116,117] as well as interactions between the STIM1 SOAR/CAD and Orai1 (orange cartoon) [82,83,107,118]. The direct interactions between STIM1 and Orai1 gate Orai1-composed hexameric Ca2+ release-activated Ca2+ (CRAC) channels [119,120], allowing Ca2+ to move from the extracellular space into the cytosol in SOCE. GPCR, G-protein coupled receptor; Gαq, G-alpha subunit; Gβγ, G-beta-gamma subunits; PLCβ, phospholipase Cβ; IP3, inositol trisphosphate; IP3R, IP3 receptor; STIM1, stromal interaction molecule-1; N, amino terminus; C, carboxyl terminus. This figure and the organelles depicted in the graphical abstract were created with BioRender.com.
Ijms 21 03642 g002

1.3. STIM1, Orai1 and Disease

Several heritable mutations in STIM1 and Orai1 cause disease. The loss of function mutations in STIM1 occur as frameshift (i.e., E128RfsX9 [121]) and intronic (i.e., NM_003156.3:c.1538-1G>A [122]) variations leading to protein abrogation. Missense mutations exist that do not eliminate STIM1 protein (i.e., R429C [123] and P165Q [124]), but similarly lead to a loss of STIM1 function [125]. Orai1 loss of function mutations include missense (i.e., R91W [32] and A103E/L194P [126]) and frameshift (i.e., A88SfsX25 [126] and H165PfsX1 [127]), which can abrogate protein expression or disrupt the function of fully expressed protein. STIM1 and Orai1 loss of function mutations both lead to CRAC channelopathy syndrome. CRAC channelopathies are characterized by combined immunodeficiency with chronic infections in the central nervous system, respiratory tract, and gastrointestinal tracts, often leading to death. These channelopathies also present autoimmune hemolytic anemia and thrombocytopenia, muscular hypotonia and atrophy, anhidrosis, and defects in dental enamel development (reviewed in [128,129]). CRAC channelopathy due to STIM1 or Orai1 mutations is autosomal recessive.
Gain of function mutations occur in STIM1 (i.e., R304W [130], I115F [129], D84G, H109N, H109R, H72Q [131], N180T, L96V, F108I, F108L [132], and G81D [133]) and Orai1 (i.e., P245L [134], G98S and L138F [135]). All these variations are missense mutations that do not abrogate protein expression, but constitutively activate SOCE. The diseases that are caused by these mutations are autosomal dominant and include York platelet syndrome, Stormorken syndrome and tubular aggregate myopathy. Symptoms of tubular aggregate myopathy include muscle weakness, myalgia, and cramps; Stormorken and York platelet syndromes present with muscle weakness, miosis, thrombocytopenia, hyposplenism, ichthyosis, dyslexia, and short stature (reviewed in [128,129,136]).

1.4. Mitochondrial Calcium Uptake and SOCE Crosstalk

The first link between mitochondrial Ca2+ uptake and SOCE was identified well before the molecules that are involved in mediating these processes. It was demonstrated that the pharmacological disruption of the inner mitochondrial membrane (IMM) potential or depletion of sodium (Na+) could inhibit SOCE [137]. Indeed, the physiological relevance of this mitochondria-SOCE link was shown in studies comparing nuclear factor of activated T-cells (NFAT) translocation after T-cell activation in the presence and absence of mitochondrial Ca2+ uptake diminishers [138]. Mitochondrial Ca2+ uptake near store-operated Ca2+ channels is believed to prevent Ca2+-dependent inactivation by limiting the formation of high Ca2+ level domains. Interestingly, one study showed that the distance between mitochondria and PM SOCE channels decreases upon SOCE activation, allowing for a more sustained Ca2+ entry due to reduced CDI [139]. In contrast, follow-up work suggested that mitochondrial Ca2+ uptake, but not mitochondrial translocation/motility, was required for STIM1 and Orai1-mediated SOCE [140]. Furthermore, respiring mitochondria were independently shown to be required for SOCE activation and were found to regulate slow CDI of SOCE by buffering cytosolic Ca2+ levels [141].
Specific links between SOCE and MCU have also been established with MCU knockdown studies showing reduced mitochondrial Ca2+ uptake following SOCE [142]. MCU is not the only recently identified mitochondrial protein that has been directly linked to SOCE regulation. The Na+/Ca2+/lithium (Li+) exchanger (NCLX) on the IMM has been shown to be vital for clearance of mitochondrial Ca2+ (reviewed in [143]). SOCE is accompanied by a rise in cytosolic Na+, which is used by NCLX in order to drive Ca2+ efflux from the matrix. The removal of extracellular Na+ during SOCE activation inhibits NCLX and SOCE [144]. The activity of NCLX is tightly linked to Orai1 channel inactivation. In the absence of NCLX activity, the loss of Ca2+ efflux from the matrix enhances ROS-dependent oxidation of a Cys residue on Orai1, which inhibits SOCE [144]. It is becoming clear that the crosstalk between SOCE and mitochondria involve several different molecules that can both positively and negatively regulate Orai1 channel function [145].

2. Mitochondrial Calcium Uniporter (MCU)

Mitochondria are primarily recognized for their role in adenosine triphosphate (ATP) production, a process that requires an electrochemical gradient across the IMM [146]. Mitochondrial Ca2+ uptake plays a central role in regulating this energy production and it is also instrumental in regulating apoptosis and shaping cytosolic Ca2+ transients [147]. While Ca2+ can readily move through the voltage dependent anion channel (VDAC) of the outer mitochondrial membrane, Ca2+ uptake into the mitochondrial matrix is more precisely controlled [148]. MCU plays a major role in mediating Ca2+ uptake into the matrix, functioning as a multimeric protein complex that forms a Ca2+ selective pore through the IMM [12,13]. Under resting conditions, the MCU complex minimally permits the movement of Ca2+ into the matrix, despite the highly negative IMM potential (~−180 mV); following a rise in cytosolic Ca2+ levels, the MCU channel open probability increases, allowing for Ca2+ to move into the matrix, being driven by the IMM potential [146,149,150]. The MCU complex is composed of the channel-forming protein, MCU, which tetramerizes to form a functional pore [151,152,153,154,155], and several additional regulatory protein components. The best studied protein regulators include a family of EF-hand containing proteins, mitochondrial Ca2+ uptake 1, 2, and 3 (MICU1, MICU2 and MICU3) [149,156,157]; a small protein that is involved in bridging MICU proteins to the MCU pore, essential MCU regulator (EMRE) [158]; an MCU paralog, MCU dominant negative beta subunit (MCUb) [159]; and a complex stabilizing protein, MCU regulator 1 (MCUR1) [160].
The MCU pore forming subunit is an ~40 kDa protein (including the mitochondrial targeting sequence; NCBI accession NP_612366.1) composed of two transmembrane domains spanning the IMM, with both the N- and carboxyl (C)-termini localized in the mitochondrial matrix [12,13,161] (Figure 3A). An Asp-Ile-Met-Glu (DIME) motif confers Ca2+ selectivity and controls permeability through the channel due to binding of Ca2+ to the acidic residues [12,13]. Amino acid substitutions at D261 and E264 lead to a loss of Ca2+ uptake, which is likely due to the removal of these negative charges [12,13]. MCU subunits homo-oligomerize to form a functioning Ca2+-permeable pore [13]. Initially, a solution nuclear magnetic resonance (NMR)-driven model of Caenorhabditis elegans MCU was suggested to exist as a pentamer, with the DIME motifs forming an unstructured loop at the opening of the channel [162]. Several resolved structures since the C. elegans NMR model have established metazoan MCU as a tetramer with the DIME motifs lining the pore as part of the helical transmembrane regions [151,152,154] (Figure 3B).
Interestingly, the C-terminal domain of MCU alone can form a channel, because it contains the TM1 and TM2 helices that create the IMM channel pore [151,154,162,163]; however, the N-terminal domain (NTD) of human MCU, which resides in the matrix, has been demonstrated to be an important hub for inputs that regulate the channel function (Figure 3C). For example, the self-association of MCU-NTD promotes channel assembly and activation [164], oxidative modification of C97 within MCU-NTD alters the channel architecture and function [165], and the S92A mutation within MCU-NTD dominant negatively disrupts mitochondrial Ca2+ uptake [163]. It is noteworthy that MCU-NTD from lower order organisms assembles as a dimer of dimers directly under the pore [151,152,153,155] and a dimer of crescent-arranged tetramers across two human MCU channels, as revealed by cryo-electron microscopy (cryo-EM) [154].

2.1. Essential Mitochondrial Calcium Uniporter Regulator (EMRE)

EMRE is a metazoan-specific ~11 kDa (including the mitochondrial targeting sequence; (NCBI accession NP_201575.3) single-pass membrane protein that spans the IMM [158]. EMRE interacts with MCU at the IMM and with the MICU1/MICU2 heterodimer in the intermembrane space (IMS) [158,166]. Thus, EMRE acts to bridge the MICU1/MICU2 Ca2+-sensing properties to MCU channel activity. The presence of an Asp-rich C-terminus is thought to be important in MCU complex regulation [158,166]; however, the specific topology of EMRE has been debated in the past [166,167]. The most recent human MCU complex structure has shown that the EMRE C-terminus is positioned in the IMS [154] (Figure 3B). EMRE is ubiquitously expressed in all tissues, and the knockdown of EMRE leads to the loss of MCU-mediated Ca2+ uptake in mammals; subsequent overexpression of MCU is unable to recover the channel function [158,166]. The knockdown of MCU concomitantly reduces EMRE expression levels, indicating a mechanism of co-stabilization [158]. It is important to note that EMRE also plays an integral role in gating MCU by stabilizing the position of an MCU juxta-membrane loop away from the pore exit [154] (Figure 3B).

2.2. Mitochondrial Calcium Uptake (MICU) Proteins

MICU1 was the first component of the MCU complex discovered [156]. MICU1 is an ~55 kDa (including the mitochondrial targeting sequence; NCBI accession NP_001350442.1) membrane associated protein, containing two conserved EF-hand Ca2+-binding domains [156]. MICU1 is localized to the IMS and indirectly modulates MCU activity via interactions with EMRE [156,158,168]. The initial findings suggested MICU1 functions independently to alter MCU-facilitated Ca2+ uptake; however, two additional MICU isoforms, MICU2 and MICU3, have since been discovered to play a role in this Ca2+-dependent regulation [157,169]. MICU2 and MICU3 contain the EF-hand Ca2+-binding domains that are similarly found in MICU1, despite only sharing 25% sequence identity [157]. MICU1 and MICU2 are ubiquitously expressed in all tissues, while MICU3 appears to be more enriched in the nervous system and skeletal muscle [157,170,171]. MICU1 and MICU2 play synergistic roles in MCU regulation and, together in a complex, function as gatekeepers of the channel [170]. It has been proposed that, at low cytosolic Ca2+ levels, MICU1 and MICU2 form a loose dimer, which inactivates MCU [172,173,174,175,176]. Following a rise in cytosolic and IMS Ca2+ concentrations, Ca2+ binds to the MICU EF-hands and triggers a conformational rearrangement that promotes tighter MICU1/MICU2 interactions and relieves MCU inhibition [172,173,174,175,176]. MICU3 appears to have a specific role in regulating neuronal function and synaptic activity. Specifically, MICU3 is believed to dimerize with MICU1 via disulfide bonds, functioning as a potent enhancer of mitochondrial Ca2+ uptake that is controlled by MCU in skeletal muscle and the central nervous system [171].
Several high-resolution crystal structures of MICU1, MICU2, and MICU3 have been elucidated alone or co-complexed and in the presence and absence of Ca2+ [173,175,176,177,178]. However, a super-complex cryo-EM structure of human MCU-EMRE-MICU1-MICU2 recently submitted to bioRxiv has suggested that a single MICU1-MICU2 heterodimer regulates the pore exit gate on each channel via Ca2+-sensitive MICU1-EMRE contacts [179]. Further, each heterodimer is observed to bridge separate channels via MICU2-MICU2 interactions [179].

2.3. Mitochondrial Calcium Uniporter Dominant Negative Beta (MCUb) Subunit

MCUb, an MCU paralog, is an ~39 kDa protein (including the mitochondrial targeting sequence; NCBI accession NP_060388.2) that shares ~50% sequence identity with MCU [159] (Figure 3A). Much like MCU, MCUb has two transmembrane domains, two coiled-coil domains, and matrix-oriented N- and C-termini. Although MCUb contains a similar DIME motif as MCU, MCUb has two key amino acid substitutions (i.e., R251W and E256V) in the pore lining region, which might underlie the inability of MCUb to function as a Ca2+-permeable pore [159]. Interestingly, MCU pore subunits have been shown to hetero-oligomerize with MCUb [159]. Mechanistically, MCUb might displace an MCU subunit from the complex to dominant negatively inhibit the channel function [180]. Once incorporated into the MCU pore, MCUb leads to a drastic decrease in mitochondrial Ca2+ uptake by diminishing the open probability of the channel and Ca2+ permeability [159,180]. Nevertheless, the precise atomic basis by which MCUb inhibits MCU function remains elusive.

2.4. Mitochondrial Calcium Uniporter Regulator-1 (MCUR1)

MCUR1 has been identified as a positive regulator of MCU, functioning as a scaffold factor that binds to both MCU and EMRE [149,181]. MCUR1 is an ~40 kDa protein (including the mitochondrial targeting sequence; NCBI accession NCBI: NP_001026883.1), which has two transmembrane domains, two conserved coiled-coil domains, and N- and C-termini residing in the IMS. The interaction between MCU and MCUR1 likely occurs in the mitochondrial matrix, stabilizing the MCU complex [181]. MCUR1 knockdown decreases mitochondrial Ca2+ uptake, reduces the IMM potential, and disrupts oxidative phosphorylation [181,182]. Collectively, these data suggest that MCUR1 not only regulates MCU-dependent Ca2+ uptake, but also plays a key role in controlling cellular bioenergetics.

2.5. MCU and Disease

Mitochondrial Ca2+ signaling has been implicated in a variety of pathophysiological disorders. Aberrant mitochondrial Ca2+ uptake not only affects processes reliant on maintaining tight cytosolic Ca2+ concentrations, but also impacts oxidative metabolism and can trigger cell death pathways [146,147,183]. Under cell stress conditions, mitochondrial Ca2+ overload and/or oxidative stress results in an opening of the mitochondrial permeability transition pore (mPTP) [183,184]. High conductance mPTP allows the free movement of ions and small molecules less than ~1.5 kDa, dissipating the IMM potential and aberrating ATP production [185,186]. Ultimately, this type of mPTP opening leads to the swelling and rupture of mitochondria and release of larger pro-apoptotic factors [183,184]. Thus, dysregulated mitochondrial Ca2+ uptake has dire implications for cell survival and has been associated with various diseases, including cancer, neurodegenerative pathologies, diabetes, and learning and muscular disorders, among others [187,188,189,190,191].
Patients have been identified with heritable MICU1 mutations (i.e., NM_006077.3:c.1078-1G>C and NM_006077.3:c.741+1G>A in splice donor and acceptor sites, respectively), which result in intronic insertions causing frameshifts, nonsense mediated mRNA decay, and loss of MICU1 protein [187,192,193]. Patient cohorts harboring MICU1 exon 1 deletions (2,776 nucleotides) [194] and MICU1 nonsense mutations (i.e., NM_006077.3:c.553C>T:p.Q185* [195]) have been identified, also abolishing the MICU1 protein. Similarly, a heritable MICU2 nonsense mutation (i.e., NM_152726:c.42G>A:p.W14*) has been discovered, which eliminates full-length MICU2 protein [196]. All of these heritable mutations lead to loss of function MICU disorders, characterized by muscle weakness, fatigue, lethargy, developmental delay, and learning disabilities [193,194,195,196]. Patient fibroblasts with MICU1 protein abrogation have conflictingly demonstrated both increased [193] and decreased [194] rates of mitochondrial Ca2+ uptake. Further, MICU2 protein abrogation also suppressed mitochondrial Ca2+ uptake rates [196]. Nevertheless, all of the patient studies have shown enhanced resting mitochondrial Ca2+ levels due to either MICU1 or MICU2 protein downregulation [193,194,196], which is likely caused by the absence of MCU gatekeeping.
Enhanced mitochondrial Ca2+ uptake can suppress cytosolic Ca2+ signals in fibroblasts from MICU1-deficient patients [193], which is consistent with past studies showing mitochondria can suppress cytosolic Ca2+ signals [197,198,199,200]. Further, this enhanced mitochondrial Ca2+ uptake may be related to work showing deletion of MICU1 in mouse hepatocytes causes sensitization to Ca2+-overload-induced mPTP opening [201]. The identification of heritable mutations in MCU complex components that lead to disease underscore the importance of not only the MCU channel, but also the diverse regulatory controls of MCU function.

3. Leucine Zipper EF-Hand Containing Transmembrane Protein-1 (LETM1)

LETM1 is an essential IMM protein linked to mitochondrial ion transport, regulation of cell cycle, mitochondrial oxidative stress and bioenergetic function [202]. Interestingly, LETM1 has been shown to play a role in mitochondrial Ca2+ and K+ ion homeostasis, regulating key facets of mitochondrial physiology, such as osmotic balance and ATP production [14,203,204,205,206,207,208]. While the molecular mechanisms by which LETM1 functions remain incompletely understood, it is clear that LETM1 is pivotal in mitochondrial function and cellular health. The deletion of the LETM1 homologue in yeast, MDM38, results in mitochondrial swelling, loss of cristae, and disruption of cellular respiration [205]. Homozygous LETM1 deletion is embryonically lethal within ~6 days in mice [204]. Clinically, LETM1 haploinsufficiency in humans is thought to be responsible for seizures in patients with Wolf–Hirschhorn syndrome (WHS) [209,210] (see below). Consistent with these findings, two independent studies identified LETM1 as one of ~2000 essential genes in human cell cultures [211,212].
In the context of mitochondrial Ca2+ ion transport, while MCU facilitates low affinity/high throughput Ca2+ uptake triggered at ~µM cytosolic Ca2+ levels (see above), LETM1 has been shown to transport Ca2+ at ~nM cytosolic Ca2+ levels, suggesting that LETM1 provides cells with a high affinity/low capacity pathway necessary for precise mitochondrial Ca2+ homeostasis [208,213,214,215]. Indeed, liposome reconstitution assays and cell culture work have shown that LETM1 directly facilitates selective bidirectional Ca2+/H+ exchange, independent of MCU [203,206,208]. Congruently, LETM1 knockdown decreases mitochondrial Ca2+ uptake, decreases ATP production, and increases ROS production [203,216].
Human LETM1 is an ~83 kDa transmembrane protein (including the mitochondrial targeting sequence; NCBI accession NP_036450.1) that localizes in the IMM [206,207] (Figure 4A). LETM1 is thought to topologically orient with the N-terminus facing the IMS. This IMS region includes a putative coiled-coil domain and a notable PTEN-induced kinase 1 (PINK1) mediated threonine phosphorylation site [217]. LETM1 contains a single sequence-identifiable transmembrane domain and a large matrix-oriented C-terminal region, which includes a putative ribosome binding domain, coiled-coil domains, and a Ca2+-binding EF-hand motif [15]. This topology for LETM1 is consistent with restricted proteinase K digestion studies [204,206,207]. However, a recent study using probes targeting tyrosine residues has proposed a new topology for LETM1, where two transmembrane domains exist and both the N- and C-termini are located in the matrix [218] (Figure 4A).
LETM1 self-oligomerizes and directly facilitates selective Ca2+/H+ exchange within liposomes without any accessory proteins [204,206,208]. A low-resolution EM study suggests human LETM1 forms a hexamer, and pH may modulate the conformation [206]. Interestingly, the crystal structure of the C-terminal domain of the yeast homologue MDM38 has been solved and exhibits a 14-3-3 like protein interaction domain, which might indicate potential sites for inter- or intra-protomer binding [219,220] (Figure 4B,C). It is tempting to speculate that these interfaces could be targeted by drugs and small molecules aimed at modulating LETM1 assembly and function.

LETM1 and Disease

Located on the chromosome band 4p16.3, LETM1 is included in one of two critical regions lost in almost all patients exhibiting the full phenotype of WHS, which causes delayed growth and development, microcephaly, distinctive facial abnormalities, and seizures [204,209]. Deficiency in mitochondrial Ca2+ transport, decreased ATP levels, and particularly increased ROS may be key determinants of the neurodegeneration linked to seizures in this disease [210]. Given that loss of LETM1 suppresses ATP production, enhances ROS, and inhibits cell proliferation [221], it is no surprise the overexpression of LETM1 has been found in many cancers, including ovarian, rectal, stomach, esophagus, breast, colon, and lung non-small cell carcinoma [222,223,224].

4. Generating New Therapeutics and Diagnostics from Protein Structures

Over the past ~15 years, a number of protein structures involved in regulating cytosolic and stored Ca2+ have been elucidated. This surge of structural information has not only been fueled by high throughput proteomic and genomic approaches that have solved decades old mysteries regarding the molecular identities of proteins that are involved in mediating ubiquitous cellular processes, such as SOCE [30,31] and mitochondrial Ca2+ uptake [12,13], but also by improving technologies aimed at revealing protein structures, such as cryo-EM, solution NMR, and X-ray crystallography. For example, several high-resolution structures of the molecular machinery involved in mediating SOCE have been elucidated, including the STIM Ca2+-sensing domains by solution NMR [75,113], the cytosolic coiled-coil domains by solution NMR [107], and X-ray crystallography [106,108] (Figure 1), as well as the Orai1 Ca2+ channel itself by cryo-EM and X-ray crystallography [120,225,226]. Cryo-EM has driven medium/high-resolution structures of MCU from lower order and metazoan organisms, including humans [151,152,153,154,155] (Figure 3). Importantly, X-ray crystallography has also contributed to understanding the underlying mechanisms of MCU regulation with elucidations of the important gatekeeping proteins [173,175,176,177,178], environment-specific conformations [163,164,227], and scaffolds [154,228]. The first elucidations of IP3Rs by crystallography dates back almost two decades with structures of the N-terminal domain [229,230]. The analogous N-terminal region in RyR has also been studied by both crystallography and NMR [231,232]. In fact, crystal structures of the full N-terminal domains of both IP3Rs and RyRs revealed a remarkable structural and functional conservation between these receptor cousins at the apical positions of the receptors [233,234]. These tetrameric channels have also been excellent candidates for cryo-EM, due to their inherently large size [235,236,237]. Other Ca2+ toolkit components, such as Na+/Ca2+ exchangers (NCX) [238], S/ER Ca2+ ATPase (SERCA) pumps [239,240], and VGICs, have also been structurally resolved [241,242]. This list of structures only scratches the surface of the available Ca2+ toolkit structural data, yet reinforces a major question at the nexus of the structural biology, cell biology, and clinical research fields. How can we use this wealth of protein structural information to design new diagnostic tools and drugs for research and treatment of disease?
Indeed, while obtaining structural information has been invaluable for informing the broad scientific community on the molecular and structural basis for the function of hundreds or thousands of Ca2+ signaling proteins, there is a disconnect between applying this protein structural knowledge and rationally developing research tools or drugs. Perhaps this disconnect exists due to the relatively static insights provided by many of the resolved structures. Most (if not all) protein function involves dynamic changes in conformation and stability. Rationalizing effective drugs without a detailed understanding of the conformational ensembles present in the inactive, transition, and active states of proteins is practically challenging.

4.1. Store-Operated and Mitochondrial Ca2+ Entry Proteins as Drug Targets

The SOCE and mitochondrial Ca2+ uptake protein machinery are appealing drug targets for several reasons. First, mutations causing disease have been identified in the proteins mediating these processes (see above). Second, complexes of multiple different proteins regulate SOCE and mitochondrial Ca2+ uptake; hence, the modulation of function, in some cases, could be achieved by targeting native protein interaction partner components. Third, drugs targeting ion channels are demonstrated to be effective at treating a wide range of human diseases and acute conditions [243,244,245]. Fourth, clinically approved immunosuppressive drugs already exist that block signaling processes that are directly mediated by SOCE (e.g., cyclosporin [246,247]). Finally, the acute administration of MCU inhibitors is known to protect cells from hypoxia/ischemia-induced injury [248,249,250,251]. The specific structures available for the SOCE and mitochondrial Ca2+ uptake molecular machinery provide a unique opportunity for the development of new diagnostics and/or therapeutics that can biasedly target different aspects of these pathways.
Small molecule screens targeting MCU function have already been performed. An initial study employed Saccharomyces cerevisiae as an MCU-null organism to screen for small molecule modulators of MCU function. Human MCU, human EMRE, and a mitochondrial-targeted Ca2+ indicator (i.e., aequorin) were expressed in these yeasts to establish a reporter system for the screening of ~600 clinically approved drugs [252]. In a second approach, HEK293A cells expressing mitochondrial aequorin were used to screen a 120,000 compound library, successfully identifying an MCU inhibitor [253]. In the most recently published screen, HeLa cells expressing mitochondrially-targeted aequorin were used to identify two MCU inhibitors from a library of 44,000 compounds [254]. All of these in cellulo screens used genetically encoded aequorin, reliably targeted to the matrix, for luminescence readout of changes in Ca2+.
Since the identification of STIM1 and Orai1, published drug screens targeting SOCE have largely involved the over-expression of STIM1 and Orai1 in mammalian cells or the use of cells with endogenous levels of STIM1 and Orai1, in combination with indicator dyes or proteins monitoring cytosolic Ca2+ or NFAT translocation to the nucleus [255,256,257,258]. One in vitro study used recombinant functional domains to screen for binding to immobilized small molecules by immunofluorescence [259]. In silico tools have also been used, but only to identify drug candidates through small molecule homology to known SOCE inhibitors [260,261].
Thus, while drug screens targeting the protein machinery that mediates both store-operated and mitochondrial Ca2+ uptake have been performed, these published approaches have not used the available protein structure information as initiation points to drug discovery.

4.2. Initiating Protein Structure-based Drug Discovery

A starting point for drug development could be initiated by a simple high-throughput small molecule in silico screen of millions of compounds. This computational screen would use available protein structures and knowledge of the vital interfaces modulating activity. The structures of the small molecules/drugs would also need to be known. Such small molecule databases and computational docking programs exist and are freely available (e.g., see ZINC15 [262] and AUTODOCK [263]). The major bottleneck for this initial step is computing the resources that are required to dock millions of compounds in a relatively short time. Nevertheless, once top leads are identified, subsequent wet laboratory experiments are required in order to validate the interaction and physiological effect(s). Downstream structure-activity-relationship studies could be subsequently used to better understand the small molecule chemical groups that are required for binding and eliciting a cellular effect and improving the compound [264].
Of course, there are limitations to these large-scale computational screens. For example, the protein structures are generally treated as rigid bodies, unless using higher level molecular dynamics simulations. Water and/or lipid molecules are typically not modeled in these fast docking schemes, even though these molecules play integral roles in biological interactions. Most critically, default scoring schemes may not be optimal for a system under study, since not all of the interactions equally rely on specific stabilization forces (e.g., H-bonding, hydrophobic, charge, shape complementarity).
High resolution structural information can also be used to develop tools for high-throughput drug discovery. For example, fluorescence-resonance energy transfer (FRET)-based biosensors can be rationalized based on proximity information that is revealed by isolated and co-complexed structures. With respect to Ca2+ signaling, these types of FRET-based sensors have been engineered and used in screens for the identification of SERCA and RyR modulators. A green fluorescent protein (GFP) and red fluorescent protein (RFP) FRET pair was fused into the cytosolic headpiece of SERCA, which undergoes large scale structural changes during pump activity. This intramolecular FRET sensor was successfully used to identify the small molecule modulators of SERCA function [265,266]. Since the sensor is genetically encoded, the cell lines stably expressing this sensor could be used in the screening. An analogous approach was used by the same group to identify modulators of RyR function. However, the RyR screen relied on FRET between FK binding protein 12.6 (FKBP12) and calmodulin, which co-complex with RyR and show different FRET signals that are based on RyR conformation [267,268]. In these RyR experiments, Alexa Fluors were chemically linked to FKBP12 and calmodulin in vitro, and vesicles containing RyRs were prepared from muscle tissues. Two aspects of STIM/Orai and MCU signaling make this approach to identifying drugs particularly appealing: (i) these proteins function in large complexes that are amenable to intermolecular FRET monitoring and (ii) STIMs are known to undergo large conformational changes, making them amenable to intramolecular FRET monitoring.

4.3. Concluding Remarks

Not discussed herein are the countless additional experimental (involving molecules, cells, animals, and humans), logistical, and administrative hurdles that are required to move any small molecule lead into clinical use. Nevertheless, we provide a simple strategy for applying high-resolution structural knowledge of proteins as a starting point to drug development. Excitingly, several key protein structures are available that inform on the mechanisms regulating the SOCE-mitochondrial Ca2+ signaling axis, providing several possible initiating points to protein structure-based drug design. The elucidation of Ca2+ signaling protein structures is vital for understanding the fundamental mechanisms underlying cellular processes that are integral to life and death. The structural information can also provide an entry point for the development of tools and therapeutics targeting these pathways.

Author Contributions

Writing—Original Draft Preparation, M.N., Q.-T.L., C.S., J.A.H., M.J.N. and P.B.S.; Writing-Review and Editing, M.N., Q.-T.L., C.S., J.A.H., M.J.N. and P.B.S.; Visualization, M.N., Q.-T.L., C.S., J.A.H., M.J.N. and P.B.S.; Supervision, P.B.S.; Project Administration, P.B.S.; and Funding Acquisition, P.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by a Natural Sciences and Engineering Research Council of Canada Discovery Grant to P.B.S. (05239).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kirichok, Y.; Krapivinsky, G.; Clapham, D.E. The mitochondrial calcium uniporter is a highly selective ion channel. Nature 2004, 427, 360–364. [Google Scholar] [CrossRef] [PubMed]
  2. Pinton, P.; Giorgi, C.; Siviero, R.; Zecchini, E.; Rizzuto, R. Calcium and apoptosis: ER-mitochondria Ca2+ transfer in the control of apoptosis. Oncogene 2008, 27, 6407–6418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Rimessi, A.; Giorgi, C.; Pinton, P.; Rizzuto, R. The versatility of mitochondrial calcium signals: From stimulation of cell metabolism to induction of cell death. Biochim. Biophys. Acta 2008, 1777, 808–816. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Deluca, H.F.; Engstrom, G.W. Calcium uptake by rat kidney mitochondria. Proc. Natl. Acad. Sci. USA 1961, 47, 1744–1750. [Google Scholar] [CrossRef] [Green Version]
  5. Amberger, A.; Weiss, H.; Haller, T.; Kock, G.; Hermann, M.; Widschwendter, M.; Margreiter, R. A subpopulation of mitochondria prevents cytosolic calcium overload in endothelial cells after cold ischemia/reperfusion. Transplantation 2001, 71, 1821–1827. [Google Scholar] [CrossRef]
  6. Ly, L.D.; Ly, D.D.; Nguyen, N.T.; Kim, J.H.; Yoo, H.; Chung, J.; Lee, M.S.; Cha, S.K.; Park, K.S. Mitochondrial Ca(2+) Uptake Relieves Palmitate-Induced Cytosolic Ca(2+) Overload in MIN6 Cells. Mol. Cells 2020, 43, 66–75. [Google Scholar]
  7. Yi, M.; Weaver, D.; Hajnoczky, G. Control of mitochondrial motility and distribution by the calcium signal: A homeostatic circuit. J. Cell Biol. 2004, 167, 661–672. [Google Scholar] [CrossRef]
  8. Peng, T.I.; Jou, M.J. Oxidative stress caused by mitochondrial calcium overload. Ann. N. Y. Acad. Sci. 2010, 1201, 183–188. [Google Scholar] [CrossRef]
  9. Patron, M.; Raffaello, A.; Granatiero, V.; Tosatto, A.; Merli, G.; De Stefani, D.; Wright, L.; Pallafacchina, G.; Terrin, A.; Mammucari, C.; et al. The mitochondrial calcium uniporter (MCU): Molecular identity and physiological roles. J. Biol. Chem. 2013, 288, 10750–10758. [Google Scholar] [CrossRef] [Green Version]
  10. Cardenas, C.; Miller, R.A.; Smith, I.; Bui, T.; Molgo, J.; Muller, M.; Vais, H.; Cheung, K.H.; Yang, J.; Parker, I.; et al. Essential regulation of cell bioenergetics by constitutive InsP3 receptor Ca2+ transfer to mitochondria. Cell 2010, 142, 270–283. [Google Scholar] [CrossRef] [Green Version]
  11. Jouaville, L.S.; Pinton, P.; Bastianutto, C.; Rutter, G.A.; Rizzuto, R. Regulation of mitochondrial ATP synthesis by calcium: Evidence for a long-term metabolic priming. Proc. Natl. Acad. Sci. USA 1999, 96, 13807–13812. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Baughman, J.M.; Perocchi, F.; Girgis, H.S.; Plovanich, M.; Belcher-Timme, C.A.; Sancak, Y.; Bao, X.R.; Strittmatter, L.; Goldberger, O.; Bogorad, R.L.; et al. Integrative genomics identifies MCU as an essential component of the mitochondrial calcium uniporter. Nature 2011, 476, 341–345. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. De Stefani, D.; Raffaello, A.; Teardo, E.; Szabo, I.; Rizzuto, R. A forty-kilodalton protein of the inner membrane is the mitochondrial calcium uniporter. Nature 2011, 476, 336–340. [Google Scholar] [CrossRef] [PubMed]
  14. Jiang, D.; Zhao, L.; Clapham, D.E. Genome-wide RNAi screen identifies Letm1 as a mitochondrial Ca2+/H+ antiporter. Science 2009, 326, 144–147. [Google Scholar] [CrossRef] [Green Version]
  15. Lin, Q.T.; Stathopulos, P.B. Molecular Mechanisms of Leucine Zipper EF-Hand Containing Transmembrane Protein-1 Function in Health and Disease. Int. J. Mol. Sci. 2019, 20, 286. [Google Scholar] [CrossRef] [Green Version]
  16. Feske, S. Calcium signalling in lymphocyte activation and disease. Nat. Rev. Immunol. 2007, 7, 690–702. [Google Scholar] [CrossRef]
  17. Feske, S.; Skolnik, E.Y.; Prakriya, M. Ion channels and transporters in lymphocyte function and immunity. Nat. Rev. Immunol. 2012, 12, 532–547. [Google Scholar] [CrossRef] [Green Version]
  18. 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]
  19. Montero, M.; Barrero, M.J.; Torrecilla, F.; Lobaton, C.D.; Moreno, A.; Alvarez, J. Stimulation by thimerosal of histamine-induced Ca(2+) release in intact HeLa cells seen with aequorin targeted to the endoplasmic reticulum. Cell Calcium 2001, 30, 181–190. [Google Scholar] [CrossRef]
  20. Suzuki, J.; Kanemaru, K.; Ishii, K.; Ohkura, M.; Okubo, Y.; Iino, M. Imaging intraorganellar Ca2+ at subcellular resolution using CEPIA. Nat. Commun. 2014, 5, 4153. [Google Scholar] [CrossRef]
  21. Yu, R.; Hinkle, P.M. Rapid turnover of calcium in the endoplasmic reticulum during signaling. Studies with cameleon calcium indicators. J. Biol. Chem. 2000, 275, 23648–23653. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Berridge, M.J.; Lipp, P.; Bootman, M.D. The versatility and universality of calcium signalling. Nat. Rev. Mol. Cell Biol. 2000, 1, 11–21. [Google Scholar] [CrossRef] [PubMed]
  23. Stathopulos, P.B.; Seo, M.D.; Enomoto, M.; Amador, F.J.; Ishiyama, N.; Ikura, M. Themes and variations in ER/SR calcium release channels: Structure and function. Physiology (Bethesda) 2012, 27, 331–342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Amador, F.J.; Stathopulos, P.B.; Enomoto, M.; Ikura, M. Ryanodine receptor calcium release channels: Lessons from structure-function studies. FEBS J. 2013, 280, 5456–5470. [Google Scholar] [CrossRef]
  25. Fedorenko, O.A.; Popugaeva, E.; Enomoto, M.; Stathopulos, P.B.; Ikura, M.; Bezprozvanny, I. Intracellular calcium channels: Inositol-1,4,5-trisphosphate receptors. Eur. J. Pharmacol. 2014, 739, 39–48. [Google Scholar] [CrossRef] [Green Version]
  26. Stathopulos, P.B.; Ikura, M. Store operated calcium entry: From concept to structural mechanisms. Cell Calcium 2017, 63, 3–7. [Google Scholar] [CrossRef]
  27. Novello, M.J.; Zhu, J.; Feng, Q.; Ikura, M.; Stathopulos, P.B. Structural elements of stromal interaction molecule function. Cell Calcium 2018, 73, 88–94. [Google Scholar] [CrossRef]
  28. Putney, J.W., Jr. A model for receptor-regulated calcium entry. Cell Calcium 1986, 7, 1–12. [Google Scholar] [CrossRef]
  29. Smyth, J.T.; Dehaven, W.I.; Jones, B.F.; Mercer, J.C.; Trebak, M.; Vazquez, G.; Putney, J.W., Jr. Emerging perspectives in store-operated Ca2+ entry: Roles of Orai, Stim and TRP. Biochim. Biophys. Acta 2006, 1763, 1147–1160. [Google Scholar] [CrossRef] [Green Version]
  30. Liou, J.; Kim, M.L.; Heo, W.D.; Jones, J.T.; Myers, J.W.; Ferrell, J.E., Jr.; Meyer, T. STIM is a Ca2+ sensor essential for Ca2+-store-depletion-triggered Ca2+ influx. Curr. Biol. 2005, 15, 1235–1241. [Google Scholar] [CrossRef] [Green Version]
  31. 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] [PubMed] [Green Version]
  32. 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] [PubMed]
  33. 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]
  34. Vig, M.; Beck, A.; Billingsley, J.M.; Lis, A.; Parvez, S.; Peinelt, C.; Koomoa, D.L.; Soboloff, J.; Gill, D.L.; Fleig, A.; et al. CRACM1 multimers form the ion-selective pore of the CRAC channel. Curr. Biol. 2006, 16, 2073–2079. [Google Scholar] [CrossRef] [Green Version]
  35. 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]
  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. Hohendanner, F.; McCulloch, A.D.; Blatter, L.A.; Michailova, A.P. Calcium and IP3 dynamics in cardiac myocytes: Experimental and computational perspectives and approaches. Front. Pharmacol. 2014, 5, 35. [Google Scholar] [CrossRef] [Green Version]
  38. Hilger, D.; Masureel, M.; Kobilka, B.K. Structure and dynamics of GPCR signaling complexes. Nat. Struct. Mol. Biol. 2018, 25, 4–12. [Google Scholar] [CrossRef]
  39. Kadamur, G.; Ross, E.M. Mammalian phospholipase C. Annu. Rev. Physiol. 2013, 75, 127–154. [Google Scholar] [CrossRef] [Green Version]
  40. Smrcka, A.V.; Fisher, I. G-protein betagamma subunits as multi-functional scaffolds and transducers in G-protein-coupled receptor signaling. Cell Mol. Life Sci. 2019, 76, 4447–4459. [Google Scholar] [CrossRef]
  41. Nieman, M.T. Protease-activated receptors in hemostasis. Blood 2016, 128, 169–177. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Scarlata, S. The role of phospholipase Cbeta on the plasma membrane and in the cytosol: How modular domains enable novel functions. Adv. Biol. Regul. 2019, 73, 100636. [Google Scholar] [CrossRef] [PubMed]
  43. Konieczny, V.; Tovey, S.C.; Mataragka, S.; Prole, D.L.; Taylor, C.W. Cyclic AMP Recruits a Discrete Intracellular Ca(2+) Store by Unmasking Hypersensitive IP3 Receptors. Cell. Rep. 2017, 18, 711–722. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Meena, A.; Tovey, S.C.; Taylor, C.W. Sustained signalling by PTH modulates IP3 accumulation and IP3 receptors through cyclic AMP junctions. J. Cell Sci. 2015, 128, 408–420. [Google Scholar] [CrossRef] [Green Version]
  45. Tovey, S.C.; Taylor, C.W. Cyclic AMP directs inositol (1,4,5)-trisphosphate-evoked Ca2+ signalling to different intracellular Ca2+ stores. J. Cell Sci. 2013, 126, 2305–2313. [Google Scholar] [CrossRef] [Green Version]
  46. Du, Z.; Lovly, C.M. Mechanisms of receptor tyrosine kinase activation in cancer. Mol. Cancer 2018, 17, 58. [Google Scholar] [CrossRef]
  47. Lemmon, M.A.; Schlessinger, J. Cell signaling by receptor tyrosine kinases. Cell 2010, 141, 1117–1134. [Google Scholar] [CrossRef] [Green Version]
  48. Trenker, R.; Jura, N. Receptor tyrosine kinase activation: From the ligand perspective. Curr. Opin. Cell Biol. 2020, 63, 174–185. [Google Scholar] [CrossRef]
  49. Feng, Y.; Wang, Y.; Zhang, S.; Haneef, K.; Liu, W. Structural and immunogenomic insights into B-cell receptor activation. J. Genet. Genom. 2020, 47, 27–35. [Google Scholar] [CrossRef]
  50. Treanor, B. B-cell receptor: From resting state to activate. Immunology 2012, 136, 21–27. [Google Scholar] [CrossRef]
  51. Kim, Y.J.; Sekiya, F.; Poulin, B.; Bae, Y.S.; Rhee, S.G. Mechanism of B-cell receptor-induced phosphorylation and activation of phospholipase C-gamma2. Mol. Cell Biol. 2004, 24, 9986–9999. [Google Scholar] [CrossRef] [Green Version]
  52. Mahtani, T.; Treanor, B. Beyond the CRAC: Diversification of ion signaling in B cells. Immunol. Rev. 2019, 291, 104–122. [Google Scholar] [CrossRef] [Green Version]
  53. Schamel, W.W.; Alarcon, B.; Minguet, S. The TCR is an allosterically regulated macromolecular machinery changing its conformation while working. Immunol. Rev. 2019, 291, 8–25. [Google Scholar] [CrossRef]
  54. Xu, X.; Li, H.; Xu, C. Structural understanding of T cell receptor triggering. Cell Mol. Immunol. 2020, 17, 193–202. [Google Scholar] [CrossRef] [PubMed]
  55. Olivera, A.; Beaven, M.A.; Metcalfe, D.D. Mast cells signal their importance in health and disease. J. Allergy Clin. Immunol. 2018, 142, 381–393. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Bournazos, S.; Ravetch, J.V. Fcgamma receptor pathways during active and passive immunization. Immunol. Rev. 2015, 268, 88–103. [Google Scholar] [CrossRef] [PubMed]
  57. Rougier, J.S.; Abriel, H. Cardiac voltage-gated calcium channel macromolecular complexes. Biochim. Biophys. Acta 2016, 1863, 1806–1812. [Google Scholar] [CrossRef] [PubMed]
  58. Wu, J.; Yan, N.; Yan, Z. Structure-Function Relationship of the Voltage-Gated Calcium Channel Cav1.1 Complex. Adv. Exp. Med. Biol. 2017, 981, 23–39. [Google Scholar]
  59. Pallien, T.; Klussmann, E. New aspects in cardiac L-type Ca2+ channel regulation. Biochem. Soc. Trans. 2020, 48, 39–49. [Google Scholar] [CrossRef]
  60. Eisner, D.A.; Caldwell, J.L.; Kistamas, K.; Trafford, A.W. Calcium and Excitation-Contraction Coupling in the Heart. Circ. Res. 2017, 121, 181–195. [Google Scholar] [CrossRef]
  61. Thul, R. Translating intracellular calcium signaling into models. Cold Spring Harb. Protoc. 2014, 2014, 463–471. [Google Scholar] [CrossRef] [Green Version]
  62. Guse, A.H.; da Silva, C.P.; Berg, I.; Skapenko, A.L.; Weber, K.; Heyer, P.; Hohenegger, M.; Ashamu, G.A.; Schulze-Koops, H.; Potter, B.V.; et al. Regulation of calcium signalling in T lymphocytes by the second messenger cyclic ADP-ribose. Nature 1999, 398, 70–73. [Google Scholar] [CrossRef] [PubMed]
  63. Kiselyov, K.; Shin, D.M.; Shcheynikov, N.; Kurosaki, T.; Muallem, S. Regulation of Ca2+-release-activated Ca2+ current (Icrac) by ryanodine receptors in inositol 1,4,5-trisphosphate-receptor-deficient DT40 cells. Biochem. J. 2001, 360, 17–22. [Google Scholar] [CrossRef] [PubMed]
  64. Schwarzmann, N.; Kunerth, S.; Weber, K.; Mayr, G.W.; Guse, A.H. Knock-down of the type 3 ryanodine receptor impairs sustained Ca2+ signaling via the T cell receptor/CD3 complex. J. Biol. Chem. 2002, 277, 50636–50642. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Smrcka, A.V. Regulation of phosphatidylinositol-specific phospholipase C at the nuclear envelope in cardiac myocytes. J. Cardiovasc. Pharmacol. 2015, 65, 203–210. [Google Scholar] [CrossRef] [Green Version]
  66. Siltari, A.; Korpela, R.; Vapaatalo, H. Bradykinin -induced vasodilatation: Role of age, ACE1-inhibitory peptide, mas- and bradykinin receptors. Peptides 2016, 85, 46–55. [Google Scholar] [CrossRef] [Green Version]
  67. Thangam, E.B.; Jemima, E.A.; Singh, H.; Baig, M.S.; Khan, M.; Mathias, C.B.; Church, M.K.; Saluja, R. The Role of Histamine and Histamine Receptors in Mast Cell-Mediated Allergy and Inflammation: The Hunt for New Therapeutic Targets. Front. Immunol. 2018, 9, 1873. [Google Scholar] [CrossRef] [Green Version]
  68. Galvin, C.D.; Hardiman, O.; Nolan, C.M. IGF-1 receptor mediates differentiation of primary cultures of mouse skeletal myoblasts. Mol. Cell. Endocrinol. 2003, 200, 19–29. [Google Scholar] [CrossRef]
  69. Molhoek, K.R.; Shada, A.L.; Smolkin, M.; Chowbina, S.; Papin, J.; Brautigan, D.L.; Slingluff, C.L., Jr. Comprehensive analysis of receptor tyrosine kinase activation in human melanomas reveals autocrine signaling through IGF-1R. Melanoma Res. 2011, 21, 274–284. [Google Scholar] [CrossRef] [Green Version]
  70. Goodnow, C.C.; Sprent, J.; Fazekas de St Groth, B.; Vinuesa, C.G. Cellular and genetic mechanisms of self tolerance and autoimmunity. Nature 2005, 435, 590–597. [Google Scholar] [CrossRef]
  71. Cai, X. Molecular evolution and functional divergence of the Ca(2+) sensor protein in store-operated Ca(2+) entry: Stromal interaction molecule. PLoS ONE 2007, 2, e609. [Google Scholar] [CrossRef] [PubMed]
  72. Enomoto, M.; Nishikawa, T.; Back, S.I.; Ishiyama, N.; Zheng, L.; Stathopulos, P.B.; Ikura, M. Coordination of a Single Calcium Ion in the EF-hand Maintains the Off State of the Stromal Interaction Molecule Luminal Domain. J. Mol. Biol. 2020, 432, 367–383. [Google Scholar] [CrossRef] [PubMed]
  73. 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: An initiation mechanism for capacitive Ca2+ entry. J. Biol. Chem. 2006, 281, 35855–35862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. 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] [PubMed]
  75. Stathopulos, P.B.; Zheng, L.; Li, G.Y.; Plevin, M.J.; Ikura, M. Structural and mechanistic insights into STIM1-mediated initiation of store-operated calcium entry. Cell 2008, 135, 110–122. [Google Scholar] [CrossRef] [Green Version]
  76. Zhou, Y.; Srinivasan, P.; Razavi, S.; Seymour, S.; Meraner, P.; Gudlur, A.; Stathopulos, P.B.; Ikura, M.; Rao, A.; Hogan, P.G. Initial activation of STIM1, the regulator of store-operated calcium entry. Nat. Struct. Mol. Biol. 2013, 20, 973–981. [Google Scholar] [CrossRef] [Green Version]
  77. Baba, Y.; Hayashi, K.; Fujii, Y.; Mizushima, A.; Watarai, H.; Wakamori, M.; Numaga, T.; Mori, Y.; Iino, M.; Hikida, M.; et al. Coupling of STIM1 to store-operated Ca2+ entry through its constitutive and inducible movement in the endoplasmic reticulum. Proc. Natl. Acad. Sci. USA 2006, 103, 16704–16709. [Google Scholar] [CrossRef] [Green Version]
  78. 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]
  79. Luik, R.M.; Wu, M.M.; Buchanan, J.; Lewis, R.S. The elementary unit of store-operated Ca2+ entry: Local activation of CRAC channels by STIM1 at ER-plasma membrane junctions. J. Cell Biol. 2006, 174, 815–825. [Google Scholar] [CrossRef] [PubMed]
  80. Wu, M.M.; Buchanan, J.; Luik, R.M.; Lewis, R.S. Ca2+ store depletion causes STIM1 to accumulate in ER regions closely associated with the plasma membrane. J. Cell Biol. 2006, 174, 803–813. [Google Scholar] [CrossRef] [PubMed]
  81. Muik, M.; Frischauf, I.; Derler, I.; Fahrner, M.; Bergsmann, J.; Eder, P.; Schindl, R.; Hesch, C.; Polzinger, B.; Fritsch, R.; et al. Dynamic coupling of the putative coiled-coil domain of ORAI1 with STIM1 mediates ORAI1 channel activation. J. Biol. Chem. 2008, 283, 8014–8022. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. 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]
  83. 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] [PubMed]
  84. Kawasaki, T.; Lange, I.; Feske, S. A minimal regulatory domain in the C terminus of STIM1 binds to and activates ORAI1 CRAC channels. Biochem. Biophys. Res. Commun. 2009, 385, 49–54. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  85. Hoth, M.; Penner, R. Calcium release-activated calcium current in rat mast cells. J. Physiol. 1993, 465, 359–386. [Google Scholar] [CrossRef] [PubMed]
  86. Zweifach, A.; Lewis, R.S. Rapid inactivation of depletion-activated calcium current (ICRAC) due to local calcium feedback. J. Gen. Physiol. 1995, 105, 209–226. [Google Scholar] [CrossRef] [Green Version]
  87. Derler, I.; Fahrner, M.; Muik, M.; Lackner, B.; Schindl, R.; Groschner, K.; Romanin, C. A Ca2(+)release-activated Ca2(+) (CRAC) modulatory domain (CMD) within STIM1 mediates fast Ca2(+)-dependent inactivation of ORAI1 channels. J. Biol. Chem. 2009, 284, 24933–24938. [Google Scholar] [CrossRef] [Green Version]
  88. Mullins, F.M.; Lewis, R.S. The inactivation domain of STIM1 is functionally coupled with the Orai1 pore to enable Ca2+-dependent inactivation. J. Gen. Physiol. 2016, 147, 153–164. [Google Scholar] [CrossRef] [Green Version]
  89. Mullins, F.M.; Yen, M.; Lewis, R.S. Orai1 pore residues control CRAC channel inactivation independently of calmodulin. J. Gen. Physiol. 2016, 147, 137–152. [Google Scholar] [CrossRef] [Green Version]
  90. Mullins, F.M.; Park, C.Y.; Dolmetsch, R.E.; Lewis, R.S. STIM1 and calmodulin interact with Orai1 to induce Ca2+-dependent inactivation of CRAC channels. Proc. Natl. Acad. Sci. USA 2009, 106, 15495–15500. [Google Scholar] [CrossRef] [Green Version]
  91. Choi, Y.J.; Zhao, Y.; Bhattacharya, M.; Stathopulos, P.B. Structural perturbations induced by Asn131 and Asn171 glycosylation converge within the EFSAM core and enhance stromal interaction molecule-1 mediated store operated calcium entry. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 1054–1063. [Google Scholar] [CrossRef] [PubMed]
  92. Gui, L.; Zhu, J.; Lu, X.; Sims, S.M.; Lu, W.Y.; Stathopulos, P.B.; Feng, Q. S-Nitrosylation of STIM1 by Neuronal Nitric Oxide Synthase Inhibits Store-Operated Ca(2+) Entry. J. Mol. Biol. 2018, 430, 1773–1785. [Google Scholar] [CrossRef] [PubMed]
  93. Zhu, J.; Lu, X.; Feng, Q.; Stathopulos, P.B. A charge-sensing region in the stromal interaction molecule 1 luminal domain confers stabilization-mediated inhibition of SOCE in response to S-nitrosylation. J. Biol. Chem. 2018, 293, 8900–8911. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhu-Mauldin, X.; Marsh, S.A.; Zou, L.; Marchase, R.B.; Chatham, J.C. Modification of STIM1 by O-linked N-acetylglucosamine (O-GlcNAc) attenuates store-operated calcium entry in neonatal cardiomyocytes. J. Biol. Chem. 2012, 287, 39094–39106. [Google Scholar] [CrossRef] [Green Version]
  95. Korzeniowski, M.K.; Popovic, M.A.; Szentpetery, Z.; Varnai, P.; Stojilkovic, S.S.; Balla, T. Dependence of STIM1/Orai1-mediated calcium entry on plasma membrane phosphoinositides. J. Biol. Chem. 2009, 284, 21027–21035. [Google Scholar] [CrossRef] [Green Version]
  96. Lopez, E.; Jardin, I.; Berna-Erro, A.; Bermejo, N.; Salido, G.M.; Sage, S.O.; Rosado, J.A.; Redondo, P.C. STIM1 tyrosine-phosphorylation is required for STIM1-Orai1 association in human platelets. Cell Signal 2012, 24, 1315–1322. [Google Scholar] [CrossRef]
  97. Manji, S.S.; Parker, N.J.; Williams, R.T.; van Stekelenburg, L.; Pearson, R.B.; Dziadek, M.; Smith, P.J. STIM1: A novel phosphoprotein located at the cell surface. Biochim. Biophys. Acta 2000, 1481, 147–155. [Google Scholar] [CrossRef]
  98. Pozo-Guisado, E.; Campbell, D.G.; Deak, M.; Alvarez-Barrientos, A.; Morrice, N.A.; Alvarez, I.S.; Alessi, D.R.; Martin-Romero, F.J. Phosphorylation of STIM1 at ERK1/2 target sites modulates store-operated calcium entry. J. Cell Sci. 2010, 123, 3084–3093. [Google Scholar] [CrossRef] [Green Version]
  99. Smyth, J.T.; Beg, A.M.; Wu, S.; Putney, J.W., Jr.; Rusan, N.M. Phosphoregulation of STIM1 leads to exclusion of the endoplasmic reticulum from the mitotic spindle. Curr. Biol. 2012, 22, 1487–1493. [Google Scholar] [CrossRef] [Green Version]
  100. Smyth, J.T.; Petranka, J.G.; Boyles, R.R.; DeHaven, W.I.; Fukushima, M.; Johnson, K.L.; Williams, J.G.; Putney, J.W., Jr. Phosphorylation of STIM1 underlies suppression of store-operated calcium entry during mitosis. Nat. Cell Biol. 2009, 11, 1465–1472. [Google Scholar] [CrossRef]
  101. Thompson, J.L.; Lai-Zhao, Y.; Stathopulos, P.B.; Grossfield, A.; Shuttleworth, T.J. Phosphorylation-mediated structural changes within the SOAR domain of stromal interaction molecule 1 enable specific activation of distinct Orai channels. J. Biol. Chem. 2018, 293, 3145–3155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. 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]
  103. Yazbeck, P.; Tauseef, M.; Kruse, K.; Amin, M.R.; Sheikh, R.; Feske, S.; Komarova, Y.; Mehta, D. STIM1 Phosphorylation at Y361 Recruits Orai1 to STIM1 Puncta and Induces Ca(2+) Entry. Sci. Rep. 2017, 7, 42758. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Hawkins, B.J.; Irrinki, K.M.; Mallilankaraman, K.; Lien, Y.C.; Wang, Y.; Bhanumathy, C.D.; Subbiah, R.; Ritchie, M.F.; Soboloff, J.; Baba, Y.; et al. S-glutathionylation activates STIM1 and alters mitochondrial homeostasis. J. Cell Biol. 2010, 190, 391–405. [Google Scholar] [CrossRef] [Green Version]
  105. Lupas, A.; Van Dyke, M.; Stock, J. Predicting coiled coils from protein sequences. Science 1991, 252, 1162–1164. [Google Scholar] [CrossRef]
  106. Cui, B.; Yang, X.; Li, S.; Lin, Z.; Wang, Z.; Dong, C.; Shen, Y. The inhibitory helix controls the intramolecular conformational switching of the C-terminus of STIM1. PLoS ONE 2013, 8, e74735. [Google Scholar] [CrossRef] [Green Version]
  107. 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]
  108. Yang, X.; Jin, H.; Cai, X.; Li, S.; Shen, Y. Structural and mechanistic insights into the activation of Stromal interaction molecule 1 (STIM1). Proc. Natl. Acad. Sci. USA 2012, 109, 5657–5662. [Google Scholar] [CrossRef] [Green Version]
  109. 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. J. Biol. Chem. 2009, 284, 8421–8426. [Google Scholar] [CrossRef] [Green Version]
  110. Covington, E.D.; Wu, M.M.; Lewis, R.S. Essential role for the CRAC activation domain in store-dependent oligomerization of STIM1. Mol. Biol. Cell 2010, 21, 1897–1907. [Google Scholar] [CrossRef] [Green Version]
  111. Berridge, M.J.; Irvine, R.F. Inositol trisphosphate, a novel second messenger in cellular signal transduction. Nature 1984, 312, 315–321. [Google Scholar] [CrossRef] [PubMed]
  112. 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]
  113. Zheng, L.; Stathopulos, P.B.; Schindl, R.; Li, G.Y.; Romanin, C.; Ikura, M. Auto-inhibitory role of the EF-SAM domain of STIM proteins in store-operated calcium entry. Proc. Natl. Acad. Sci. USA 2011, 108, 1337–1342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  114. 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 Ca(2+) signalling prompted by STIM1 conformational switch. Nat. Commun. 2015, 6, 7826. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. 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] [PubMed] [Green Version]
  116. Calloway, N.; Owens, T.; Corwith, K.; Rodgers, W.; Holowka, D.; Baird, B. Stimulated association of STIM1 and Orai1 is regulated by the balance of PtdIns(4,5)P(2) between distinct membrane pools. J. Cell Sci. 2011, 124, 2602–2610. [Google Scholar] [CrossRef] [Green Version]
  117. Walsh, C.M.; Chvanov, M.; Haynes, L.P.; Petersen, O.H.; Tepikin, A.V.; Burgoyne, R.D. Role of phosphoinositides in STIM1 dynamics and store-operated calcium entry. Biochem. J. 2009, 425, 159–168. [Google Scholar] [CrossRef] [Green Version]
  118. 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]
  119. McNally, B.A.; 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. J. Physiol. 2013, 591, 2833–2850. [Google Scholar] [CrossRef]
  120. 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] [Green Version]
  121. Picard, C.; McCarl, C.A.; Papolos, A.; Khalil, S.; Luthy, K.; Hivroz, C.; LeDeist, F.; Rieux-Laucat, F.; Rechavi, G.; Rao, A.; et al. STIM1 mutation associated with a syndrome of immunodeficiency and autoimmunity. N. Engl. J. Med. 2009, 360, 1971–1980. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  122. Byun, M.; Abhyankar, A.; Lelarge, V.; Plancoulaine, S.; Palanduz, A.; Telhan, L.; Boisson, B.; Picard, C.; Dewell, S.; Zhao, C.; et al. Whole-exome sequencing-based discovery of STIM1 deficiency in a child with fatal classic Kaposi sarcoma. J. Exp. Med. 2010, 207, 2307–2312. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  123. Fuchs, S.; Rensing-Ehl, A.; Speckmann, C.; Bengsch, B.; Schmitt-Graeff, A.; Bondzio, I.; Maul-Pavicic, A.; Bass, T.; Vraetz, T.; Strahm, B.; et al. Antiviral and regulatory T cell immunity in a patient with stromal interaction molecule 1 deficiency. J. Immunol. 2012, 188, 1523–1533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Schaballie, H.; Rodriguez, R.; Martin, E.; Moens, L.; Frans, G.; Lenoir, C.; Dutre, J.; Canioni, D.; Bossuyt, X.; Fischer, A.; et al. A novel hypomorphic mutation in STIM1 results in a late-onset immunodeficiency. J. Allergy Clin. Immunol. 2015, 136, 816–819 e4. [Google Scholar] [CrossRef] [Green Version]
  125. Maus, M.; Jairaman, A.; Stathopulos, P.B.; Muik, M.; Fahrner, M.; Weidinger, C.; Benson, M.; Fuchs, S.; Ehl, S.; Romanin, C.; et al. Missense mutation in immunodeficient patients shows the multifunctional roles of coiled-coil domain 3 (CC3) in STIM1 activation. Proc. Natl. Acad. Sci. USA 2015, 112, 6206–6211. [Google Scholar] [CrossRef] [Green Version]
  126. McCarl, C.A.; Picard, C.; Khalil, S.; Kawasaki, T.; Rother, J.; Papolos, A.; Kutok, J.; Hivroz, C.; Ledeist, F.; Plogmann, K.; et al. ORAI1 deficiency and lack of store-operated Ca2+ entry cause immunodeficiency, myopathy, and ectodermal dysplasia. J. Allergy Clin. Immunol. 2009, 124, 1311–1318 e7. [Google Scholar] [CrossRef] [Green Version]
  127. Chou, J.; Badran, Y.R.; Yee, C.S.K.; Bainter, W.; Ohsumi, T.K.; Al-Hammadi, S.; Pai, S.Y.; Feske, S.; Geha, R.S. A novel mutation in ORAI1 presenting with combined immunodeficiency and residual T-cell function. J. Allergy Clin. Immunol. 2015, 136, 479–482 e1. [Google Scholar] [CrossRef] [Green Version]
  128. Feske, S. CRAC channels and disease-From human CRAC channelopathies and animal models to novel drugs. Cell Calcium 2019, 80, 112–116. [Google Scholar] [CrossRef]
  129. Lacruz, R.S.; Feske, S. Diseases caused by mutations in ORAI1 and STIM1. Ann. N. Y. Acad. Sci. 2015, 1356, 45–79. [Google Scholar] [CrossRef] [Green Version]
  130. Misceo, D.; Holmgren, A.; Louch, W.E.; Holme, P.A.; Mizobuchi, M.; Morales, R.J.; De Paula, A.M.; Stray-Pedersen, A.; Lyle, R.; Dalhus, B.; et al. A dominant STIM1 mutation causes Stormorken syndrome. Hum. Mutat. 2014, 35, 556–564. [Google Scholar] [CrossRef]
  131. Bohm, J.; Chevessier, F.; Maues De Paula, A.; Koch, C.; Attarian, S.; Feger, C.; Hantai, D.; Laforet, P.; Ghorab, K.; Vallat, J.M.; et al. Constitutive activation of the calcium sensor STIM1 causes tubular-aggregate myopathy. Am. J. Hum. Genet. 2013, 92, 271–278. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Bohm, J.; Chevessier, F.; Koch, C.; Peche, G.A.; Mora, M.; Morandi, L.; Pasanisi, B.; Moroni, I.; Tasca, G.; Fattori, F.; et al. Clinical, histological and genetic characterisation of patients with tubular aggregate myopathy caused by mutations in STIM1. J. Med. Genet. 2014, 51, 824–833. [Google Scholar] [CrossRef] [PubMed]
  133. Walter, M.C.; Rossius, M.; Zitzelsberger, M.; Vorgerd, M.; Muller-Felber, W.; Ertl-Wagner, B.; Zhang, Y.; Brinkmeier, H.; Senderek, J.; Schoser, B. 50 years to diagnosis: Autosomal dominant tubular aggregate myopathy caused by a novel STIM1 mutation. Neuromuscul. Disord. 2015, 25, 577–584. [Google Scholar] [CrossRef] [PubMed]
  134. Nesin, V.; Wiley, G.; Kousi, M.; Ong, E.C.; Lehmann, T.; Nicholl, D.J.; Suri, M.; Shahrizaila, N.; Katsanis, N.; Gaffney, P.M.; et al. Activating mutations in STIM1 and ORAI1 cause overlapping syndromes of tubular myopathy and congenital miosis. Proc. Natl. Acad. Sci. USA 2014, 111, 4197–4202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. Endo, Y.; Noguchi, S.; Hara, Y.; Hayashi, Y.K.; Motomura, K.; Miyatake, S.; Murakami, N.; Tanaka, S.; Yamashita, S.; Kizu, R.; et al. Dominant mutations in ORAI1 cause tubular aggregate myopathy with hypocalcemia via constitutive activation of store-operated Ca(2)(+) channels. Hum. Mol. Genet. 2015, 24, 637–648. [Google Scholar] [CrossRef] [Green Version]
  136. Bohm, J.; Laporte, J. Gain-of-function mutations in STIM1 and ORAI1 causing tubular aggregate myopathy and Stormorken syndrome. Cell Calcium 2018, 76, 1–9. [Google Scholar] [CrossRef]
  137. Hoth, M.; Fanger, C.M.; Lewis, R.S. Mitochondrial regulation of store-operated calcium signaling in T lymphocytes. J. Cell Biol. 1997, 137, 633–648. [Google Scholar] [CrossRef] [Green Version]
  138. Hoth, M.; Button, D.C.; Lewis, R.S. Mitochondrial control of calcium-channel gating: A mechanism for sustained signaling and transcriptional activation in T lymphocytes. Proc. Natl. Acad. Sci. USA 2000, 97, 10607–10612. [Google Scholar] [CrossRef] [Green Version]
  139. Quintana, A.; Schwarz, E.C.; Schwindling, C.; Lipp, P.; Kaestner, L.; Hoth, M. Sustained activity of calcium release-activated calcium channels requires translocation of mitochondria to the plasma membrane. J. Biol. Chem. 2006, 281, 40302–40309. [Google Scholar] [CrossRef] [Green Version]
  140. Naghdi, S.; Waldeck-Weiermair, M.; Fertschai, I.; Poteser, M.; Graier, W.F.; Malli, R. Mitochondrial Ca2+ uptake and not mitochondrial motility is required for STIM1-Orai1-dependent store-operated Ca2+ entry. J. Cell Sci. 2010, 123, 2553–2564. [Google Scholar] [CrossRef] [Green Version]
  141. Gilabert, J.A.; Parekh, A.B. Respiring mitochondria determine the pattern of activation and inactivation of the store-operated Ca(2+) current I(CRAC). EMBO J. 2000, 19, 6401–6407. [Google Scholar] [CrossRef] [PubMed]
  142. Samanta, K.; Douglas, S.; Parekh, A.B. Mitochondrial calcium uniporter MCU supports cytoplasmic Ca2+ oscillations, store-operated Ca2+ entry and Ca2+-dependent gene expression in response to receptor stimulation. PLoS ONE 2014, 9, e101188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  143. Kostic, M.; Sekler, I. Functional properties and mode of regulation of the mitochondrial Na(+)/Ca(2+) exchanger, NCLX. Semin Cell Dev. Biol. 2019, 94, 59–65. [Google Scholar] [CrossRef] [PubMed]
  144. Ben-Kasus Nissim, T.; Zhang, X.; Elazar, A.; Roy, S.; Stolwijk, J.A.; Zhou, Y.; Motiani, R.K.; Gueguinou, M.; Hempel, N.; Hershfinkel, M.; et al. Mitochondria control store-operated Ca(2+) entry through Na(+) and redox signals. EMBO J. 2017, 36, 797–815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Villalobos, C.; Gutierrez, L.G.; Hernandez-Morales, M.; Del Bosque, D.; Nunez, L. Mitochondrial control of store-operated Ca(2+) channels in cancer: Pharmacological implications. Pharmacol. Res. 2018, 135, 136–143. [Google Scholar] [CrossRef] [PubMed]
  146. Marchi, S.; Pinton, P. The mitochondrial calcium uniporter complex: Molecular components, structure and physiopathological implications. J. Physiol. 2014, 592, 829–839. [Google Scholar] [CrossRef]
  147. Giacomello, M.; Drago, I.; Pizzo, P.; Pozzan, T. Mitochondrial Ca2+ as a key regulator of cell life and death. Cell Death Differ. 2007, 14, 1267–1274. [Google Scholar] [CrossRef]
  148. Shoshan-Barmatz, V.; De, S. Mitochondrial VDAC, the Na(+)/Ca(2+) Exchanger, and the Ca(2+) Uniporter in Ca(2+) Dynamics and Signaling. Adv. Exp. Med. Biol. 2017, 981, 323–347. [Google Scholar]
  149. Mallilankaraman, K.; Doonan, P.; Cardenas, C.; Chandramoorthy, H.C.; Muller, M.; Miller, R.; Hoffman, N.E.; Gandhirajan, R.K.; Molgo, J.; Birnbaum, M.J.; et al. MICU1 is an essential gatekeeper for MCU-mediated mitochondrial Ca(2+) uptake that regulates cell survival. Cell 2012, 151, 630–644. [Google Scholar] [CrossRef] [Green Version]
  150. Rizzuto, R.; Pozzan, T. Microdomains of intracellular Ca2+: Molecular determinants and functional consequences. Physiol. Rev. 2006, 86, 369–408. [Google Scholar] [CrossRef]
  151. Baradaran, R.; Wang, C.; Siliciano, A.F.; Long, S.B. Cryo-EM structures of fungal and metazoan mitochondrial calcium uniporters. Nature 2018, 559, 580–584. [Google Scholar] [CrossRef]
  152. Fan, C.; Fan, M.; Orlando, B.J.; Fastman, N.M.; Zhang, J.; Xu, Y.; Chambers, M.G.; Xu, X.; Perry, K.; Liao, M.; et al. X-ray and cryo-EM structures of the mitochondrial calcium uniporter. Nature 2018, 559, 575–579. [Google Scholar] [CrossRef]
  153. Nguyen, N.X.; Armache, J.P.; Lee, C.; Yang, Y.; Zeng, W.; Mootha, V.K.; Cheng, Y.; Bai, X.C.; Jiang, Y. Cryo-EM structure of a fungal mitochondrial calcium uniporter. Nature 2018, 559, 570–574. [Google Scholar] [CrossRef]
  154. Wang, Y.; Nguyen, N.X.; She, J.; Zeng, W.; Yang, Y.; Bai, X.C.; Jiang, Y. Structural Mechanism of EMRE-Dependent Gating of the Human Mitochondrial Calcium Uniporter. Cell 2019, 177, 1252–1261 e13. [Google Scholar] [CrossRef] [PubMed]
  155. Yoo, J.; Wu, M.; Yin, Y.; Herzik, M.A., Jr.; Lander, G.C.; Lee, S.Y. Cryo-EM structure of a mitochondrial calcium uniporter. Science 2018, 361, 506–511. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Perocchi, F.; Gohil, V.M.; Girgis, H.S.; Bao, X.R.; McCombs, J.E.; Palmer, A.E.; Mootha, V.K. MICU1 encodes a mitochondrial EF hand protein required for Ca(2+) uptake. Nature 2010, 467, 291–296. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Plovanich, M.; Bogorad, R.L.; Sancak, Y.; Kamer, K.J.; Strittmatter, L.; Li, A.A.; Girgis, H.S.; Kuchimanchi, S.; De Groot, J.; Speciner, L.; et al. MICU2, a paralog of MICU1, resides within the mitochondrial uniporter complex to regulate calcium handling. PLoS ONE 2013, 8, e55785. [Google Scholar] [CrossRef] [Green Version]
  158. Sancak, Y.; Markhard, A.L.; Kitami, T.; Kovacs-Bogdan, E.; Kamer, K.J.; Udeshi, N.D.; Carr, S.A.; Chaudhuri, D.; Clapham, D.E.; Li, A.A.; et al. EMRE is an essential component of the mitochondrial calcium uniporter complex. Science 2013, 342, 1379–1382. [Google Scholar] [CrossRef] [Green Version]
  159. Raffaello, A.; De Stefani, D.; Sabbadin, D.; Teardo, E.; Merli, G.; Picard, A.; Checchetto, V.; Moro, S.; Szabo, I.; Rizzuto, R. The mitochondrial calcium uniporter is a multimer that can include a dominant-negative pore-forming subunit. EMBO J. 2013, 32, 2362–2376. [Google Scholar] [CrossRef] [Green Version]
  160. Mallilankaraman, K.; Cardenas, C.; Doonan, P.J.; Chandramoorthy, H.C.; Irrinki, K.M.; Golenar, T.; Csordas, G.; Madireddi, P.; Yang, J.; Muller, M.; et al. MCUR1 is an essential component of mitochondrial Ca2+ uptake that regulates cellular metabolism. Nat. Cell Biol. 2012, 14, 1336–1343. [Google Scholar] [CrossRef] [Green Version]
  161. Martell, J.D.; Deerinck, T.J.; Sancak, Y.; Poulos, T.L.; Mootha, V.K.; Sosinsky, G.E.; Ellisman, M.H.; Ting, A.Y. Engineered ascorbate peroxidase as a genetically encoded reporter for electron microscopy. Nat. Biotechnol. 2012, 30, 1143–1148. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Oxenoid, K.; Dong, Y.; Cao, C.; Cui, T.; Sancak, Y.; Markhard, A.L.; Grabarek, Z.; Kong, L.; Liu, Z.; Ouyang, B.; et al. Architecture of the mitochondrial calcium uniporter. Nature 2016, 533, 269–273. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Lee, Y.; Min, C.K.; Kim, T.G.; Song, H.K.; Lim, Y.; Kim, D.; Shin, K.; Kang, M.; Kang, J.Y.; Youn, H.S.; et al. Structure and function of the N-terminal domain of the human mitochondrial calcium uniporter. EMBO Rep. 2015, 16, 1318–1333. [Google Scholar] [CrossRef] [PubMed]
  164. Lee, S.K.; Shanmughapriya, S.; Mok, M.C.Y.; Dong, Z.; Tomar, D.; Carvalho, E.; Rajan, S.; Junop, M.S.; Madesh, M.; Stathopulos, P.B. Structural Insights into Mitochondrial Calcium Uniporter Regulation by Divalent Cations. Cell Chem. Biol. 2016, 23, 1157–1169. [Google Scholar] [CrossRef] [Green Version]
  165. Dong, Z.; Shanmughapriya, S.; Tomar, D.; Siddiqui, N.; Lynch, S.; Nemani, N.; Breves, S.L.; Zhang, X.; Tripathi, A.; Palaniappan, P.; et al. Mitochondrial Ca(2+) Uniporter Is a Mitochondrial Luminal Redox Sensor that Augments MCU Channel Activity. Mol. Cell 2017, 65, 1014–1028 e7. [Google Scholar] [CrossRef] [Green Version]
  166. Vais, H.; Mallilankaraman, K.; Mak, D.D.; Hoff, H.; Payne, R.; Tanis, J.E.; Foskett, J.K. EMRE Is a Matrix Ca(2+) Sensor that Governs Gatekeeping of the Mitochondrial Ca(2+) Uniporter. Cell Rep. 2016, 14, 403–410. [Google Scholar] [CrossRef] [Green Version]
  167. Tsai, M.F.; Phillips, C.B.; Ranaghan, M.; Tsai, C.W.; Wu, Y.; Willliams, C.; Miller, C. Dual functions of a small regulatory subunit in the mitochondrial calcium uniporter complex. Elife 2016, 5, e15545. [Google Scholar] [CrossRef]
  168. Hung, V.; Zou, P.; Rhee, H.W.; Udeshi, N.D.; Cracan, V.; Svinkina, T.; Carr, S.A.; Mootha, V.K.; Ting, A.Y. Proteomic mapping of the human mitochondrial intermembrane space in live cells via ratiometric APEX tagging. Mol. Cell 2014, 55, 332–341. [Google Scholar] [CrossRef] [Green Version]
  169. Csordas, G.; Golenar, T.; Seifert, E.L.; Kamer, K.J.; Sancak, Y.; Perocchi, F.; Moffat, C.; Weaver, D.; de la Fuente Perez, S.; Bogorad, R.; et al. MICU1 controls both the threshold and cooperative activation of the mitochondrial Ca(2)(+) uniporter. Cell Metab. 2013, 17, 976–987. [Google Scholar] [CrossRef] [Green Version]
  170. Patron, M.; Checchetto, V.; Raffaello, A.; Teardo, E.; Vecellio Reane, D.; Mantoan, M.; Granatiero, V.; Szabo, I.; De Stefani, D.; Rizzuto, R. MICU1 and MICU2 finely tune the mitochondrial Ca2+ uniporter by exerting opposite effects on MCU activity. Mol. Cell 2014, 53, 726–737. [Google Scholar] [CrossRef] [Green Version]
  171. Patron, M.; Granatiero, V.; Espino, J.; Rizzuto, R.; De Stefani, D. MICU3 is a tissue-specific enhancer of mitochondrial calcium uptake. Cell Death Differ. 2019, 26, 179–195. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  172. Kamer, K.J.; Grabarek, Z.; Mootha, V.K. High-affinity cooperative Ca(2+) binding by MICU1-MICU2 serves as an on-off switch for the uniporter. EMBO Rep. 2017, 18, 1397–1411. [Google Scholar] [CrossRef] [PubMed]
  173. Kamer, K.J.; Jiang, W.; Kaushik, V.K.; Mootha, V.K.; Grabarek, Z. Crystal structure of MICU2 and comparison with MICU1 reveal insights into the uniporter gating mechanism. Proc. Natl. Acad. Sci. USA 2019, 116, 3546–3555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Kamer, K.J.; Mootha, V.K. MICU1 and MICU2 play nonredundant roles in the regulation of the mitochondrial calcium uniporter. EMBO Rep. 2014, 15, 299–307. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Xing, Y.; Wang, M.; Wang, J.; Nie, Z.; Wu, G.; Yang, X.; Shen, Y. Dimerization of MICU Proteins Controls Ca(2+) Influx through the Mitochondrial Ca(2+) Uniporter. Cell Rep. 2019, 26, 1203–1212 e4. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  176. Park, J.; Lee, Y.; Park, T.; Kang, J.Y.; Mun, S.A.; Jin, M.; Yang, J.; Eom, S.H. Structure of the MICU1-MICU2 heterodimer provides insights into the gatekeeping threshold shift. IUCrJ 2020, 7, 355–365. [Google Scholar] [CrossRef] [Green Version]
  177. Wang, L.; Yang, X.; Li, S.; Wang, Z.; Liu, Y.; Feng, J.; Zhu, Y.; Shen, Y. Structural and mechanistic insights into MICU1 regulation of mitochondrial calcium uptake. EMBO J. 2014, 33, 594–604. [Google Scholar] [CrossRef]
  178. Wu, W.; Shen, Q.; Lei, Z.; Qiu, Z.; Li, D.; Pei, H.; Zheng, J.; Jia, Z. The crystal structure of MICU2 provides insight into Ca(2+) binding and MICU1-MICU2 heterodimer formation. EMBO Rep. 2019, 20, e47488. [Google Scholar] [CrossRef]
  179. Zhuo, W.; Zhou, H.; Guo, R.; Yi, J.; Sui, Y.; Zhang, L.; Zeng, W.; Wang, P.; Yang, M. Structure of intact human MCU supercomplex with the auxiliary MICU subunits. bioRxiv 2020. [Google Scholar] [CrossRef] [Green Version]
  180. Lambert, J.P.; Luongo, T.S.; Tomar, D.; Jadiya, P.; Gao, E.; Zhang, X.; Lucchese, A.M.; Kolmetzky, D.W.; Shah, N.S.; Elrod, J.W. MCUB Regulates the Molecular Composition of the Mitochondrial Calcium Uniporter Channel to Limit Mitochondrial Calcium Overload During Stress. Circulation 2019, 140, 1720–1733. [Google Scholar] [CrossRef]
  181. Tomar, D.; Dong, Z.; Shanmughapriya, S.; Koch, D.A.; Thomas, T.; Hoffman, N.E.; Timbalia, S.A.; Goldman, S.J.; Breves, S.L.; Corbally, D.P.; et al. MCUR1 Is a Scaffold Factor for the MCU Complex Function and Promotes Mitochondrial Bioenergetics. Cell Rep. 2016, 15, 1673–1685. [Google Scholar] [CrossRef] [PubMed]
  182. Vais, H.; Tanis, J.E.; Muller, M.; Payne, R.; Mallilankaraman, K.; Foskett, J.K. MCUR1, CCDC90A, Is a Regulator of the Mitochondrial Calcium Uniporter. Cell Metab. 2015, 22, 533–535. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Rizzuto, R.; De Stefani, D.; Raffaello, A.; Mammucari, C. Mitochondria as sensors and regulators of calcium signalling. Nat. Rev. Mol. Cell Biol. 2012, 13, 566–578. [Google Scholar] [CrossRef] [PubMed]
  184. Sileikyte, J.; Forte, M. The Mitochondrial Permeability Transition in Mitochondrial Disorders. Oxid Med. Cell Longev. 2019, 2019, 3403075. [Google Scholar] [CrossRef] [Green Version]
  185. Kwong, J.Q.; Molkentin, J.D. Physiological and pathological roles of the mitochondrial permeability transition pore in the heart. Cell Metab. 2015, 21, 206–214. [Google Scholar] [CrossRef] [Green Version]
  186. Zoratti, M.; Szabo, I. The mitochondrial permeability transition. Biochim. Biophys. Acta 1995, 1241, 139–176. [Google Scholar] [CrossRef]
  187. Bhosale, G.; Sharpe, J.A.; Koh, A.; Kouli, A.; Szabadkai, G.; Duchen, M.R. Pathological consequences of MICU1 mutations on mitochondrial calcium signalling and bioenergetics. Biochim. Biophys. Acta Mol. Cell Res. 2017, 1864, 1009–1017. [Google Scholar] [CrossRef]
  188. Halestrap, A.P.; Richardson, A.P. The mitochondrial permeability transition: A current perspective on its identity and role in ischaemia/reperfusion injury. J. Mol. Cell Cardiol. 2015, 78, 129–141. [Google Scholar] [CrossRef]
  189. Liao, Y.; Dong, Y.; Cheng, J. The Function of the Mitochondrial Calcium Uniporter in Neurodegenerative Disorders. Int. J. Mol. Sci. 2017, 18, 248. [Google Scholar] [CrossRef]
  190. Tarasov, A.I.; Semplici, F.; Ravier, M.A.; Bellomo, E.A.; Pullen, T.J.; Gilon, P.; Sekler, I.; Rizzuto, R.; Rutter, G.A. The mitochondrial Ca2+ uniporter MCU is essential for glucose-induced ATP increases in pancreatic beta-cells. PLoS ONE 2012, 7, e39722. [Google Scholar] [CrossRef] [Green Version]
  191. Vultur, A.; Gibhardt, C.S.; Stanisz, H.; Bogeski, I. The role of the mitochondrial calcium uniporter (MCU) complex in cancer. Pflugers Arch. 2018, 470, 1149–1163. [Google Scholar] [CrossRef] [PubMed]
  192. Debattisti, V.; Horn, A.; Singh, R.; Seifert, E.L.; Hogarth, M.W.; Mazala, D.A.; Huang, K.T.; Horvath, R.; Jaiswal, J.K.; Hajnoczky, G. Dysregulation of Mitochondrial Ca(2+) Uptake and Sarcolemma Repair Underlie Muscle Weakness and Wasting in Patients and Mice Lacking MICU1. Cell Rep. 2019, 29, 1274–1286. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Logan, C.V.; Szabadkai, G.; Sharpe, J.A.; Parry, D.A.; Torelli, S.; Childs, A.M.; Kriek, M.; Phadke, R.; Johnson, C.A.; Roberts, N.Y.; et al. Loss-of-function mutations in MICU1 cause a brain and muscle disorder linked to primary alterations in mitochondrial calcium signaling. Nat. Genet. 2014, 46, 188–193. [Google Scholar] [CrossRef] [PubMed]
  194. Lewis-Smith, D.; Kamer, K.J.; Griffin, H.; Childs, A.M.; Pysden, K.; Titov, D.; Duff, J.; Pyle, A.; Taylor, R.W.; Yu-Wai-Man, P.; et al. Homozygous deletion in MICU1 presenting with fatigue and lethargy in childhood. Neurol. Genet. 2016, 2, e59. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  195. Musa, S.; Eyaid, W.; Kamer, K.; Ali, R.; Al-Mureikhi, M.; Shahbeck, N.; Al Mesaifri, F.; Makhseed, N.; Mohamed, Z.; AlShehhi, W.A.; et al. A Middle Eastern Founder Mutation Expands the Genotypic and Phenotypic Spectrum of Mitochondrial MICU1 Deficiency: A Report of 13 Patients. JIMD Rep. 2019, 43, 79–83. [Google Scholar]
  196. Shamseldin, H.E.; Alasmari, A.; Salih, M.A.; Samman, M.M.; Mian, S.A.; Alshidi, T.; Ibrahim, N.; Hashem, M.; Faqeih, E.; Al-Mohanna, F.; et al. A null mutation in MICU2 causes abnormal mitochondrial calcium homeostasis and a severe neurodevelopmental disorder. Brain 2017, 140, 2806–2813. [Google Scholar] [CrossRef] [Green Version]
  197. Gordienko, D.V.; Greenwood, I.A.; Bolton, T.B. Direct visualization of sarcoplasmic reticulum regions discharging Ca(2+)sparks in vascular myocytes. Cell Calcium 2001, 29, 13–28. [Google Scholar] [CrossRef]
  198. Hajnoczky, G.; Hager, R.; Thomas, A.P. Mitochondria suppress local feedback activation of inositol 1,4, 5-trisphosphate receptors by Ca2+. J. Biol. Chem. 1999, 274, 14157–14162. [Google Scholar] [CrossRef] [Green Version]
  199. Marchant, J.S.; Ramos, V.; Parker, I. Structural and functional relationships between Ca2+ puffs and mitochondria in Xenopus oocytes. Am. J. Physiol. Cell Physiol. 2002, 282, C1374–C1386. [Google Scholar] [CrossRef]
  200. Pacher, P.; Thomas, A.P.; Hajnoczky, G. Ca2+ marks: Miniature calcium signals in single mitochondria driven by ryanodine receptors. Proc. Natl. Acad. Sci. USA 2002, 99, 2380–2385. [Google Scholar] [CrossRef] [Green Version]
  201. Antony, A.N.; Paillard, M.; Moffat, C.; Juskeviciute, E.; Correnti, J.; Bolon, B.; Rubin, E.; Csordas, G.; Seifert, E.L.; Hoek, J.B.; et al. MICU1 regulation of mitochondrial Ca(2+) uptake dictates survival and tissue regeneration. Nat. Commun. 2016, 7, 10955. [Google Scholar] [CrossRef] [PubMed]
  202. Austin, S.; Nowikovsky, K. LETM1: Essential for Mitochondrial Biology and Cation Homeostasis? Trends Biochem. Sci. 2019, 44, 648–658. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  203. Doonan, P.J.; Chandramoorthy, H.C.; Hoffman, N.E.; Zhang, X.; Cardenas, C.; Shanmughapriya, S.; Rajan, S.; Vallem, S.; Chen, X.; Foskett, J.K.; et al. LETM1-dependent mitochondrial Ca2+ flux modulates cellular bioenergetics and proliferation. FASEB J. 2014, 28, 4936–4949. [Google Scholar] [CrossRef] [Green Version]
  204. Jiang, D.; Zhao, L.; Clish, C.B.; Clapham, D.E. Letm1, the mitochondrial Ca2+/H+ antiporter, is essential for normal glucose metabolism and alters brain function in Wolf-Hirschhorn syndrome. Proc. Natl. Acad. Sci. USA 2013, 110, E2249–E2254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Nowikovsky, K.; Reipert, S.; Devenish, R.J.; Schweyen, R.J. Mdm38 protein depletion causes loss of mitochondrial K+/H+ exchange activity, osmotic swelling and mitophagy. Cell Death Differ. 2007, 14, 1647–1656. [Google Scholar] [CrossRef] [PubMed]
  206. Shao, J.; Fu, Z.; Ji, Y.; Guan, X.; Guo, S.; Ding, Z.; Yang, X.; Cong, Y.; Shen, Y. Leucine zipper-EF-hand containing transmembrane protein 1 (LETM1) forms a Ca(2+)/H(+) antiporter. Sci. Rep. 2016, 6, 34174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Tamai, S.; Iida, H.; Yokota, S.; Sayano, T.; Kiguchiya, S.; Ishihara, N.; Hayashi, J.; Mihara, K.; Oka, T. Characterization of the mitochondrial protein LETM1, which maintains the mitochondrial tubular shapes and interacts with the AAA-ATPase BCS1L. J. Cell Sci. 2008, 121, 2588–2600. [Google Scholar] [CrossRef] [Green Version]
  208. Tsai, M.F.; Jiang, D.; Zhao, L.; Clapham, D.; Miller, C. Functional reconstitution of the mitochondrial Ca2+/H+ antiporter Letm1. J. Gen. Physiol. 2014, 143, 67–73. [Google Scholar] [CrossRef] [Green Version]
  209. Endele, S.; Fuhry, M.; Pak, S.J.; Zabel, B.U.; Winterpacht, A. LETM1, a novel gene encoding a putative EF-hand Ca(2+)-binding protein, flanks the Wolf-Hirschhorn syndrome (WHS) critical region and is deleted in most WHS patients. Genomics 1999, 60, 218–225. [Google Scholar] [CrossRef]
  210. Rutherford, E.L.; Lowery, L.A. Exploring the developmental mechanisms underlying Wolf-Hirschhorn Syndrome: Evidence for defects in neural crest cell migration. Dev. Biol. 2016, 420, 1–10. [Google Scholar] [CrossRef] [Green Version]
  211. Blomen, V.A.; Majek, P.; Jae, L.T.; Bigenzahn, J.W.; Nieuwenhuis, J.; Staring, J.; Sacco, R.; van Diemen, F.R.; Olk, N.; Stukalov, A.; et al. Gene essentiality and synthetic lethality in haploid human cells. Science 2015, 350, 1092–1096. [Google Scholar] [CrossRef] [PubMed]
  212. Wang, T.; Birsoy, K.; Hughes, N.W.; Krupczak, K.M.; Post, Y.; Wei, J.J.; Lander, E.S.; Sabatini, D.M. Identification and characterization of essential genes in the human genome. Science 2015, 350, 1096–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  213. Collins, T.J.; Lipp, P.; Berridge, M.J.; Bootman, M.D. Mitochondrial Ca(2+) uptake depends on the spatial and temporal profile of cytosolic Ca(2+) signals. J. Biol. Chem. 2001, 276, 26411–26420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  214. Santo-Domingo, J.; Demaurex, N. Calcium uptake mechanisms of mitochondria. Biochim. Biophys. Acta 2010, 1797, 907–912. [Google Scholar] [CrossRef]
  215. Santo-Domingo, J.; Demaurex, N. Perspectives on: SGP symposium on mitochondrial physiology and medicine: The renaissance of mitochondrial pH. J. Gen. Physiol. 2012, 139, 415–423. [Google Scholar] [CrossRef] [Green Version]
  216. Aral, C.; Demirkesen, S.; Bircan, R.; Yasar Sirin, D. Melatonin reverses the oxidative stress and mitochondrial dysfunction caused by LETM1 silencing. Cell Biol. Int. 2020, 44, 795–807. [Google Scholar] [CrossRef]
  217. Huang, E.; Qu, D.; Huang, T.; Rizzi, N.; Boonying, W.; Krolak, D.; Ciana, P.; Woulfe, J.; Klein, C.; Slack, R.S.; et al. PINK1-mediated phosphorylation of LETM1 regulates mitochondrial calcium transport and protects neurons against mitochondrial stress. Nat. Commun. 2017, 8, 1399. [Google Scholar] [CrossRef]
  218. Yoo, C.M.; Rhee, H.W. APEX, a Master Key To Resolve Membrane Topology in Live Cells. Biochemistry 2020, 59, 250–259. [Google Scholar] [CrossRef]
  219. Lupo, D.; Vollmer, C.; Deckers, M.; Mick, D.U.; Tews, I.; Sinning, I.; Rehling, P. Mdm38 is a 14-3-3-like receptor and associates with the protein synthesis machinery at the inner mitochondrial membrane. Traffic 2011, 12, 1457–1466. [Google Scholar] [CrossRef]
  220. Nakamura, S.; Matsui, A.; Akabane, S.; Tamura, Y.; Hatano, A.; Miyano, Y.; Omote, H.; Kajikawa, M.; Maenaka, K.; Moriyama, Y.; et al. The mitochondrial inner membrane protein LETM1 modulates cristae organization through its LETM domain. Commun. Biol. 2020, 3, 99. [Google Scholar] [CrossRef]
  221. Huang, B.; Zhang, J.; Zhang, X.; Huang, C.; Hu, G.; Li, S.; Xie, T.; Liu, M.; Xu, Y. Suppression of LETM1 by siRNA inhibits cell proliferation and invasion of bladder cancer cells. Oncol. Rep. 2017, 38, 2935–2940. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Piao, L.; Li, Y.; Kim, S.J.; Byun, H.S.; Huang, S.M.; Hwang, S.K.; Yang, K.J.; Park, K.A.; Won, M.; Hong, J.; et al. Association of LETM1 and MRPL36 contributes to the regulation of mitochondrial ATP production and necrotic cell death. Cancer Res. 2009, 69, 3397–3404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  223. Piao, L.; Li, Y.; Kim, S.J.; Sohn, K.C.; Yang, K.J.; Park, K.A.; Byun, H.S.; Won, M.; Hong, J.; Hur, G.M.; et al. Regulation of OPA1-mediated mitochondrial fusion by leucine zipper/EF-hand-containing transmembrane protein-1 plays a role in apoptosis. Cell Signal. 2009, 21, 767–777. [Google Scholar] [CrossRef] [PubMed]
  224. Yang, Z.; Ni, W.; Cui, C.; Qi, W.; Piao, L.; Xuan, Y. Identification of LETM1 as a marker of cancer stem-like cells and predictor of poor prognosis in esophageal squamous cell carcinoma. Hum. Pathol. 2018, 81, 148–156. [Google Scholar] [CrossRef]
  225. Hou, X.; Burstein, S.R.; Long, S.B. Structures reveal opening of the store-operated calcium channel Orai. Elife 2018, 7, e36758. [Google Scholar] [CrossRef] [PubMed]
  226. Liu, X.; Wu, G.; Yu, Y.; Chen, X.; Ji, R.; Lu, J.; Li, X.; Zhang, X.; Yang, X.; Shen, Y. Molecular understanding of calcium permeation through the open Orai channel. PLoS Biol. 2019, 17, e3000096. [Google Scholar] [CrossRef]
  227. Yuan, Y.; Cao, C.; Wen, M.; Li, M.; Dong, Y.; Wu, L.; Wu, J.; Cui, T.; Li, D.; Chou, J.J.; et al. Structural Characterization of the N-Terminal Domain of the Dictyostelium discoideum Mitochondrial Calcium Uniporter. ACS Omega. 2020, 5, 6452–6460. [Google Scholar] [CrossRef] [Green Version]
  228. Adlakha, J.; Karamichali, I.; Sangwallek, J.; Deiss, S.; Bar, K.; Coles, M.; Hartmann, M.D.; Lupas, A.N.; Hernandez Alvarez, B. Characterization of MCU-Binding Proteins MCUR1 and CCDC90B-Representatives of a Protein Family Conserved in Prokaryotes and Eukaryotic Organelles. Structure 2019, 27, 464–475 e6. [Google Scholar] [CrossRef] [Green Version]
  229. Bosanac, I.; Alattia, J.R.; Mal, T.K.; Chan, J.; Talarico, S.; Tong, F.K.; Tong, K.I.; Yoshikawa, F.; Furuichi, T.; Iwai, M.; et al. Structure of the inositol 1,4,5-trisphosphate receptor binding core in complex with its ligand. Nature 2002, 420, 696–700. [Google Scholar] [CrossRef]
  230. Bosanac, I.; Yamazaki, H.; Matsu-Ura, T.; Michikawa, T.; Mikoshiba, K.; Ikura, M. Crystal structure of the ligand binding suppressor domain of type 1 inositol 1,4,5-trisphosphate receptor. Mol. Cell 2005, 17, 193–203. [Google Scholar] [CrossRef]
  231. Amador, F.J.; Kimlicka, L.; Stathopulos, P.B.; Gasmi-Seabrook, G.M.; Maclennan, D.H.; Van Petegem, F.; Ikura, M. Type 2 ryanodine receptor domain A contains a unique and dynamic alpha-helix that transitions to a beta-strand in a mutant linked with a heritable cardiomyopathy. J. Mol. Biol. 2013, 425, 4034–4046. [Google Scholar] [CrossRef] [PubMed]
  232. Amador, F.J.; Liu, S.; Ishiyama, N.; Plevin, M.J.; Wilson, A.; MacLennan, D.H.; Ikura, M. Crystal structure of type I ryanodine receptor amino-terminal beta-trefoil domain reveals a disease-associated mutation “hot spot” loop. Proc. Natl. Acad. Sci. USA 2009, 106, 11040–11044. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  233. Seo, M.D.; Velamakanni, S.; Ishiyama, N.; Stathopulos, P.B.; Rossi, A.M.; Khan, S.A.; Dale, P.; Li, C.; Ames, J.B.; Ikura, M.; et al. Structural and functional conservation of key domains in InsP3 and ryanodine receptors. Nature 2012, 483, 108–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. Tung, C.C.; Lobo, P.A.; Kimlicka, L.; Van Petegem, F. The amino-terminal disease hotspot of ryanodine receptors forms a cytoplasmic vestibule. Nature 2010, 468, 585–588. [Google Scholar] [CrossRef] [PubMed]
  235. des Georges, A.; Clarke, O.B.; Zalk, R.; Yuan, Q.; Condon, K.J.; Grassucci, R.A.; Hendrickson, W.A.; Marks, A.R.; Frank, J. Structural Basis for Gating and Activation of RyR1. Cell 2016, 167, 145–157 e17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  236. Peng, W.; Shen, H.; Wu, J.; Guo, W.; Pan, X.; Wang, R.; Chen, S.R.; Yan, N. Structural basis for the gating mechanism of the type 2 ryanodine receptor RyR2. Science 2016, 354, aah5324. [Google Scholar] [CrossRef]
  237. Fan, G.; Baker, M.L.; Wang, Z.; Baker, M.R.; Sinyagovskiy, P.A.; Chiu, W.; Ludtke, S.J.; Serysheva, I.I. Gating machinery of InsP3R channels revealed by electron cryomicroscopy. Nature 2015, 527, 336–341. [Google Scholar] [CrossRef] [Green Version]
  238. Liao, J.; Li, H.; Zeng, W.; Sauer, D.B.; Belmares, R.; Jiang, Y. Structural insight into the ion-exchange mechanism of the sodium/calcium exchanger. Science 2012, 335, 686–690. [Google Scholar] [CrossRef]
  239. Toyoshima, C.; Nomura, H.; Tsuda, T. Lumenal gating mechanism revealed in calcium pump crystal structures with phosphate analogues. Nature 2004, 432, 361–368. [Google Scholar] [CrossRef]
  240. Tsunekawa, N.; Ogawa, H.; Tsueda, J.; Akiba, T.; Toyoshima, C. Mechanism of the E2 to E1 transition in Ca(2+) pump revealed by crystal structures of gating residue mutants. Proc. Natl. Acad. Sci. USA 2018, 115, 12722–12727. [Google Scholar] [CrossRef] [Green Version]
  241. Zhao, Y.; Huang, G.; Wu, J.; Wu, Q.; Gao, S.; Yan, Z.; Lei, J.; Yan, N. Molecular Basis for Ligand Modulation of a Mammalian Voltage-Gated Ca(2+) Channel. Cell 2019, 177, 1495–1506 e12. [Google Scholar] [CrossRef] [PubMed]
  242. Tang, L.; Gamal El-Din, T.M.; Swanson, T.M.; Pryde, D.C.; Scheuer, T.; Zheng, N.; Catterall, W.A. Structural basis for inhibition of a voltage-gated Ca(2+) channel by Ca(2+) antagonist drugs. Nature 2016, 537, 117–121. [Google Scholar] [CrossRef] [PubMed]
  243. Imbrici, P.; Nicolotti, O.; Leonetti, F.; Conte, D.; Liantonio, A. Ion Channels in Drug Discovery and Safety Pharmacology. Methods Mol. Biol. 2018, 1800, 313–326. [Google Scholar] [PubMed]
  244. Santos, R.; Ursu, O.; Gaulton, A.; Bento, A.P.; Donadi, R.S.; Bologa, C.G.; Karlsson, A.; Al-Lazikani, B.; Hersey, A.; Oprea, T.I.; et al. A comprehensive map of molecular drug targets. Nat. Rev. Drug Discov. 2017, 16, 19–34. [Google Scholar] [CrossRef] [PubMed]
  245. Wulff, H.; Christophersen, P.; Colussi, P.; Chandy, K.G.; Yarov-Yarovoy, V. Antibodies and venom peptides: New modalities for ion channels. Nat. Rev. Drug Discov. 2019, 18, 339–357. [Google Scholar] [CrossRef]
  246. Faulds, D.; Goa, K.L.; Benfield, P. Cyclosporin. A review of its pharmacodynamic and pharmacokinetic properties, and therapeutic use in immunoregulatory disorders. Drugs 1993, 45, 953–1040. [Google Scholar] [CrossRef]
  247. Flores, C.; Fouquet, G.; Moura, I.C.; Maciel, T.T.; Hermine, O. Lessons to Learn From Low-Dose Cyclosporin-A: A New Approach for Unexpected Clinical Applications. Front. Immunol. 2019, 10, 588. [Google Scholar] [CrossRef] [Green Version]
  248. Garcia-Rivas Gde, J.; Carvajal, K.; Correa, F.; Zazueta, C. Ru360, a specific mitochondrial calcium uptake inhibitor, improves cardiac post-ischaemic functional recovery in rats in vivo. Br. J. Pharmacol. 2006, 149, 829–837. [Google Scholar] [CrossRef] [Green Version]
  249. Zhang, S.Z.; Gao, Q.; Cao, C.M.; Bruce, I.C.; Xia, Q. Involvement of the mitochondrial calcium uniporter in cardioprotection by ischemic preconditioning. Life Sci. 2006, 78, 738–745. [Google Scholar] [CrossRef]
  250. de Jesus Garcia-Rivas, G.; Guerrero-Hernandez, A.; Guerrero-Serna, G.; Rodriguez-Zavala, J.S.; Zazueta, C. Inhibition of the mitochondrial calcium uniporter by the oxo-bridged dinuclear ruthenium amine complex (Ru360) prevents from irreversible injury in postischemic rat heart. FEBS J. 2005, 272, 3477–3488. [Google Scholar] [CrossRef]
  251. Woods, J.J.; Nemani, N.; Shanmughapriya, S.; Kumar, A.; Zhang, M.; Nathan, S.R.; Thomas, M.; Carvalho, E.; Ramachandran, K.; Srikantan, S.; et al. A Selective and Cell-Permeable Mitochondrial Calcium Uniporter (MCU) Inhibitor Preserves Mitochondrial Bioenergetics after Hypoxia/Reoxygenation Injury. ACS Cent. Sci. 2019, 5, 153–166. [Google Scholar] [CrossRef] [PubMed]
  252. Arduino, D.M.; Wettmarshausen, J.; Vais, H.; Navas-Navarro, P.; Cheng, Y.; Leimpek, A.; Ma, Z.; Delrio-Lorenzo, A.; Giordano, A.; Garcia-Perez, C.; et al. Systematic Identification of MCU Modulators by Orthogonal Interspecies Chemical Screening. Mol. Cell 2017, 67, 711–723. [Google Scholar] [CrossRef] [PubMed]
  253. Kon, N.; Murakoshi, M.; Isobe, A.; Kagechika, K.; Miyoshi, N.; Nagayama, T. DS16570511 is a small-molecule inhibitor of the mitochondrial calcium uniporter. Cell Death Discov. 2017, 3, 17045. [Google Scholar] [CrossRef] [PubMed]
  254. Di Marco, G.; Vallese, F.; Jourde, B.; Bergsdorf, C.; Sturlese, M.; De Mario, A.; Techer-Etienne, V.; Haasen, D.; Oberhauser, B.; Schleeger, S.; et al. A High-Throughput Screening Identifies MICU1 Targeting Compounds. Cell Rep. 2020, 30, 2321–2331 e6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Goto, J.; Suzuki, A.Z.; Ozaki, S.; Matsumoto, N.; Nakamura, T.; Ebisui, E.; Fleig, A.; Penner, R.; Mikoshiba, K. Two novel 2-aminoethyl diphenylborinate (2-APB) analogues differentially activate and inhibit store-operated Ca(2+) entry via STIM proteins. Cell Calcium 2010, 47, 1–10. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Kim, K.D.; Srikanth, S.; Tan, Y.V.; Yee, M.K.; Jew, M.; Damoiseaux, R.; Jung, M.E.; Shimizu, S.; An, D.S.; Ribalet, B.; et al. Calcium signaling via Orai1 is essential for induction of the nuclear orphan receptor pathway to drive Th17 differentiation. J. Immunol. 2014, 192, 110–122. [Google Scholar] [CrossRef] [Green Version]
  257. Stauderman, K.A. CRAC channels as targets for drug discovery and development. Cell Calcium 2018, 74, 147–159. [Google Scholar] [CrossRef]
  258. Zhang, H.Z.; Xu, X.L.; Chen, H.Y.; Ali, S.; Wang, D.; Yu, J.W.; Xu, T.; Nan, F.J. Discovery and structural optimization of 1-phenyl-3-(1-phenylethyl)urea derivatives as novel inhibitors of CRAC channel. Acta Pharmacol. Sin. 2015, 36, 1137–1144. [Google Scholar] [CrossRef] [Green Version]
  259. 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]
  260. Azimi, I.; Flanagan, J.U.; Stevenson, R.J.; Inserra, M.; Vetter, I.; Monteith, G.R.; Denny, W.A. Evaluation of known and novel inhibitors of Orai1-mediated store operated Ca(2+) entry in MDA-MB-231 breast cancer cells using a Fluorescence Imaging Plate Reader assay. Bioorg. Med. Chem. 2017, 25, 440–449. [Google Scholar] [CrossRef] [Green Version]
  261. Rahman, S.; Rahman, T. Unveiling some FDA-approved drugs as inhibitors of the store-operated Ca(2+) entry pathway. Sci. Rep. 2017, 7, 12881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  262. Sterling, T.; Irwin, J.J. ZINC 15--Ligand Discovery for Everyone. J. Chem. Inf. Model. 2015, 55, 2324–2337. [Google Scholar] [CrossRef] [PubMed]
  263. Trott, O.; Olson, A.J. AutoDock Vina: Improving the speed and accuracy of docking with a new scoring function, efficient optimization, and multithreading. J. Comput. Chem. 2010, 31, 455–461. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Andricopulo, A.D.; Montanari, C.A. Structure-activity relationships for the design of small-molecule inhibitors. Mini. Rev. Med. Chem. 2005, 5, 585–593. [Google Scholar] [CrossRef]
  265. Gruber, S.J.; Cornea, R.L.; Li, J.; Peterson, K.C.; Schaaf, T.M.; Gillispie, G.D.; Dahl, R.; Zsebo, K.M.; Robia, S.L.; Thomas, D.D. Discovery of enzyme modulators via high-throughput time-resolved FRET in living cells. J. Biomol. Screen. 2014, 19, 215–222. [Google Scholar] [CrossRef] [Green Version]
  266. Schaaf, T.M.; Peterson, K.C.; Grant, B.D.; Bawaskar, P.; Yuen, S.; Li, J.; Muretta, J.M.; Gillispie, G.D.; Thomas, D.D. High-Throughput Spectral and Lifetime-Based FRET Screening in Living Cells to Identify Small-Molecule Effectors of SERCA. SLAS Discov. 2017, 22, 262–273. [Google Scholar]
  267. Rebbeck, R.T.; Essawy, M.M.; Nitu, F.R.; Grant, B.D.; Gillispie, G.D.; Thomas, D.D.; Bers, D.M.; Cornea, R.L. High-Throughput Screens to Discover Small-Molecule Modulators of Ryanodine Receptor Calcium Release Channels. SLAS Discov. 2017, 22, 176–186. [Google Scholar] [CrossRef] [Green Version]
  268. Rebbeck, R.T.; Singh, D.P.; Janicek, K.A.; Bers, D.M.; Thomas, D.D.; Launikonis, B.S.; Cornea, R.L. RyR1-targeted drug discovery pipeline integrating FRET-based high-throughput screening and human myofiber dynamic Ca(2+) assays. Sci. Rep. 2020, 10, 1791. [Google Scholar] [CrossRef]
Figure 3. Structural elucidation of the human MCU pore. (A) Domain architecture of human mitochondrial Ca2+ uniporter (MCU) (Uniprot accession Q8NE86) and MCUb (Uniprot accession Q9NWR8). The conserved coiled-coil and transmembrane regions are shaded salmon and orange, respectively. The residue ranges based on Uniprot annotation are indicated above the respective domain. The topological orientation relative to the inner mitochondrial membrane (IMM) is indicated below the diagrams and the amino (N)-terminal and carboxyl (C)-terminal domains that have been the focus of separate structural studies [162,163,164] are indicated above the diagrams. (B) Human MCU tetramer in complex with four essential MCU regulator (EMRE) peptides. The EMRE peptides (black spheres) are oriented with the N-termini in the matrix and the C-termini in the IMS. The EMRE N-termini are situated adjacent to the JML (green spheres), stabilizing the loop conformation. The MCU N- and C-termini are oriented in the matrix. The MCU C-terminal domain (salmon cartoon representation) assembles as a symmetric tetramer, while the N-terminal domain (cyan cartoon representation) exhibits a more linear/crescent tetramer assembly. The Asp-Ile-Met-Glu (DIME) motif (red sticks), important for Ca2+ selectivity and permeation, is indicated near the IMS opening of the channel. (C) Human MCU-N-terminal domain (MCU-NTD) structure showing the location of various sensory input sites. The glutathionylation C97 and phosphorylation S92 post-translation modification sites (blue sticks) are indicated. The negatively charged Asp residues (red sticks) in close proximity to the Mg2+ (orange sphere) binding site are shown. In B and C, colours of MCU regions are consistent with panel A. JML, juxtamembrane loop; IMM, inner mitochondrial membrane; IMS, intermembrane space; NTD, N-terminal domain. The structure figures were made using the 6O58 [154] and 5KUJ [164] pdb coordinate files for the MCU-EMRE complex and MCU-NTD, respectively. MTS, mitochondrial targeting sequence; CC1, -2, coiled-coil-1, -2; TM1, -2, transmembrane-1, -2; N, amino terminus; C, carboxyl terminus.
Figure 3. Structural elucidation of the human MCU pore. (A) Domain architecture of human mitochondrial Ca2+ uniporter (MCU) (Uniprot accession Q8NE86) and MCUb (Uniprot accession Q9NWR8). The conserved coiled-coil and transmembrane regions are shaded salmon and orange, respectively. The residue ranges based on Uniprot annotation are indicated above the respective domain. The topological orientation relative to the inner mitochondrial membrane (IMM) is indicated below the diagrams and the amino (N)-terminal and carboxyl (C)-terminal domains that have been the focus of separate structural studies [162,163,164] are indicated above the diagrams. (B) Human MCU tetramer in complex with four essential MCU regulator (EMRE) peptides. The EMRE peptides (black spheres) are oriented with the N-termini in the matrix and the C-termini in the IMS. The EMRE N-termini are situated adjacent to the JML (green spheres), stabilizing the loop conformation. The MCU N- and C-termini are oriented in the matrix. The MCU C-terminal domain (salmon cartoon representation) assembles as a symmetric tetramer, while the N-terminal domain (cyan cartoon representation) exhibits a more linear/crescent tetramer assembly. The Asp-Ile-Met-Glu (DIME) motif (red sticks), important for Ca2+ selectivity and permeation, is indicated near the IMS opening of the channel. (C) Human MCU-N-terminal domain (MCU-NTD) structure showing the location of various sensory input sites. The glutathionylation C97 and phosphorylation S92 post-translation modification sites (blue sticks) are indicated. The negatively charged Asp residues (red sticks) in close proximity to the Mg2+ (orange sphere) binding site are shown. In B and C, colours of MCU regions are consistent with panel A. JML, juxtamembrane loop; IMM, inner mitochondrial membrane; IMS, intermembrane space; NTD, N-terminal domain. The structure figures were made using the 6O58 [154] and 5KUJ [164] pdb coordinate files for the MCU-EMRE complex and MCU-NTD, respectively. MTS, mitochondrial targeting sequence; CC1, -2, coiled-coil-1, -2; TM1, -2, transmembrane-1, -2; N, amino terminus; C, carboxyl terminus.
Ijms 21 03642 g003
Figure 4. Structural insights into potential LETM1 interaction interfaces. (A) Domain architecture of human LETM1 (Uniprot accession O95202) and the yeast LETM1 homologue (MDM38; Uniprot accession Q08179). The predicted/putative location of four coiled-coil domains (red), two transmembranes helices (green), EF-hand Ca2+-binding motif (yellow), mitochondrial targeting sequence (blue), ribosome binding domain (magenta) and 14-3-3 like domain (cyan) are indicated with residue ranges shown above the respective segment based on the Uniprot annotation. (B) Yeast MDM38 ribosome binding domain crystal structure exhibiting a 14-3-3 like fold. The amino (N)-terminal and carboxyl (C) terminal helices are coloured blue and red, respectively. The N-terminal helix has been previously implicated in mediating protein-protein interactions of the structurally homologous human 14-3-3 epsilon protein [219]. The conserved residue positions necessary for functional human LETM1 assembly and growth complementation of yeast deficient in MDM38 [220] are coloured magenta (i.e., D295, R318, G319, M320 in yeast MDM38; shown as sticks). (C) Surface representation of the yeast MDM38 ribosome binding domain showing a putative substrate/protein interaction cavity (orange), which may interact with phosphorylated proteins akin to human 14-3-3 structural homologues [219]. CC1 -2, -3, 4, coiled-coil-1, -2, -3; TM1, -2, transmembrane-1, -2; EF, EF-hand; RBD, ribosome binding domain; MTS, mitochondrial targeting sequence. The structure figures were made using the 3SKQ [219] pdb coordinate file for yeast MDM38 ribosome binding domain.
Figure 4. Structural insights into potential LETM1 interaction interfaces. (A) Domain architecture of human LETM1 (Uniprot accession O95202) and the yeast LETM1 homologue (MDM38; Uniprot accession Q08179). The predicted/putative location of four coiled-coil domains (red), two transmembranes helices (green), EF-hand Ca2+-binding motif (yellow), mitochondrial targeting sequence (blue), ribosome binding domain (magenta) and 14-3-3 like domain (cyan) are indicated with residue ranges shown above the respective segment based on the Uniprot annotation. (B) Yeast MDM38 ribosome binding domain crystal structure exhibiting a 14-3-3 like fold. The amino (N)-terminal and carboxyl (C) terminal helices are coloured blue and red, respectively. The N-terminal helix has been previously implicated in mediating protein-protein interactions of the structurally homologous human 14-3-3 epsilon protein [219]. The conserved residue positions necessary for functional human LETM1 assembly and growth complementation of yeast deficient in MDM38 [220] are coloured magenta (i.e., D295, R318, G319, M320 in yeast MDM38; shown as sticks). (C) Surface representation of the yeast MDM38 ribosome binding domain showing a putative substrate/protein interaction cavity (orange), which may interact with phosphorylated proteins akin to human 14-3-3 structural homologues [219]. CC1 -2, -3, 4, coiled-coil-1, -2, -3; TM1, -2, transmembrane-1, -2; EF, EF-hand; RBD, ribosome binding domain; MTS, mitochondrial targeting sequence. The structure figures were made using the 3SKQ [219] pdb coordinate file for yeast MDM38 ribosome binding domain.
Ijms 21 03642 g004
Table 1. Various signaling mechanisms leading to sarco/endoplasmic reticulum (S/ER) Ca2+ store release.
Table 1. Various signaling mechanisms leading to sarco/endoplasmic reticulum (S/ER) Ca2+ store release.
Surface Receptor or ChannelExtracellular SignalImmediate Effect of Receptor ActivationDownstream S/ER Ca2+ Release ChannelReferences
GPCR
(e.g., PAR1)
Protein and small molecule ligands
(e.g., thrombin)
PLCβ activationIP3R[38,39,40,41,42]
a Polymodal GPCR
(e.g., M3R and PTHR1)
Protein and small molecule ligands
(e.g., carbachol/PTH)
PLCβ and adenylyl cyclase activationIP3R[43,44,45]
RTK
(e.g., IGF-1R)
Protein ligands
(e.g., IGF-1)
PLCγ activationIP3R[46,47,48]
b BCRIgM-binding antigensPLCγ activationIP3R[49,50,51,52]
b TCRAntigen presenting cell (MHC)PLCγ activationIP3R[16,17,53,54]
FcAntigen-antibody complexPLCγ activationIP3R[16,17,55,56]
VGIC
(e.g., L-type Ca2+ channel)
Membrane depolarizationCa2+ influx from the extracellular spaceRyR[57,58,59,60,61]
a Simultaneous distinct GPCR stimuli can lead to the production of IP3 (via PLC) and cAMP (via adenylyl cyclase). cAMP sensitizes a subset of high affinity IP3Rs in a discrete S/ER Ca2+ store to IP3, which would otherwise remain inactive in the absence of cAMP. b TCR and BCR activation can also lead to the generation cyclic adenosine diphosphate ribose (cADPr), resulting in S/ER Ca2+ release via RyR activation [16,17,62,63,64]. PTH, parathyroid hormone; PTHR1, PTH receptor type-1; cAMP, 3′,5′-cyclic adenosine monophosphate; M3R, muscarinic type-3 acetylcholine receptor.

Share and Cite

MDPI and ACS Style

Noble, M.; Lin, Q.-T.; Sirko, C.; Houpt, J.A.; Novello, M.J.; Stathopulos, P.B. Structural Mechanisms of Store-Operated and Mitochondrial Calcium Regulation: Initiation Points for Drug Discovery. Int. J. Mol. Sci. 2020, 21, 3642. https://doi.org/10.3390/ijms21103642

AMA Style

Noble M, Lin Q-T, Sirko C, Houpt JA, Novello MJ, Stathopulos PB. Structural Mechanisms of Store-Operated and Mitochondrial Calcium Regulation: Initiation Points for Drug Discovery. International Journal of Molecular Sciences. 2020; 21(10):3642. https://doi.org/10.3390/ijms21103642

Chicago/Turabian Style

Noble, Megan, Qi-Tong Lin, Christian Sirko, Jacob A. Houpt, Matthew J. Novello, and Peter B. Stathopulos. 2020. "Structural Mechanisms of Store-Operated and Mitochondrial Calcium Regulation: Initiation Points for Drug Discovery" International Journal of Molecular Sciences 21, no. 10: 3642. https://doi.org/10.3390/ijms21103642

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