Abstract
Calcium ions (Ca2+) are prominent cell signaling effectors that regulate a wide variety of cellular processes. Among the different players in Ca2+ homeostasis, primary active Ca2+ transporters are responsible for keeping low basal Ca2+ levels in the cytosol while establishing steep Ca2+ gradients across intracellular membranes or the plasma membrane. This review summarizes our current knowledge on the three types of primary active Ca2+-ATPases: the sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA) pumps, the secretory pathway Ca2+- ATPase (SPCA) isoforms, and the plasma membrane Ca2+-ATPase (PMCA) Ca2+-transporters. We first discuss the Ca2+ transport mechanism of SERCA1a, which serves as a reference to describe the Ca2+ transport of other Ca2+ pumps. We further highlight the common and unique features of each isoform and review their structure–function relationship, expression pattern, regulatory mechanisms, and specific physiological roles. Finally, we discuss the increasing genetic and in vivo evidence that links the dysfunction of specific Ca2+-ATPase isoforms to a broad range of human pathologies, and highlight emerging therapeutic strategies that target Ca2+ pumps.
Ca2+ signaling is crucial for many physiological processes and is dysregulated in a multitude of pathological conditions. Ca2+ influx from outside the cell or Ca2+ release from intracellular reservoirs increases cytosolic Ca2+ levels in the nano- to micromolar range, leading to a Ca2+ signal that can vary in amplitude, frequency, and subcellular localization. Afterward, resting cytosolic Ca2+ levels must be restored by primary and secondary active transport systems, which are referred to as Ca2+ pumps and exchangers, respectively. In this review, we will focus on the primary active Ca2+-transporters or Ca2+-ATPases, which are responsible for keeping low basal Ca2+ levels in the cytosol while establishing vitally important Ca2+ gradients across intracellular membranes or the plasma membrane. All Ca2+-ATPases belong to the family of P-type ATPases: the sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA), the Golgi/secretory pathway Ca2+-ATPase (SPCA), and the plasma membrane Ca2+-ATPase (PMCA) (Fig. 1A). SERCA and SPCA share 43% sequence similarity and belong to the P2A subfamily, whereas the more distal PMCA shares 33% sequence similarity with SERCA and belongs to the P2B subfamily (Vangheluwe et al. 2009).
The transport process of a P-type Ca2+- ATPase follows the Post–Albers cycle, that is, alternating between a Ca2+-bound E1 state and a Ca2+-free E2 state (Fig. 1B; Albers 1967; Post et al. 1972). During transport, Ca2+-ATPases undergo reversible autophosphorylation on a critically conserved Asp residue in one of the cytosolic domains, which controls the opening and closure of the Ca2+-binding sites in the membrane region. Since 2000, many structures of SERCA1a in various conformations were solved (Toyoshima et al. 2000, 2013; Olesen et al. 2004, 2007; Toyoshima and Mizutani 2004; Jensen et al. 2006; Clausen et al. 2016), and more recently, also, PMCA1 (Gong et al. 2018) and SERCA2a and SERCA2b structures were reported (Inoue et al. 2019; Sitsel et al. 2019). These structures revealed the Ca2+-transporter architecture, which involves a transmembrane (TM) domain of 10 TM helices and three cytosolic domains (Fig. 2B). The TM region contains the Ca2+-binding sites and ion entrance/exit pathways. Although SERCA pumps contain two Ca2+-binding sites (I and II, formed by helices M4, M5, M6, and M8 in SERCA isoforms), SPCA and PMCA only contain one ion-binding site, closely resembling the Ca2+-binding site II of SERCA (Fig. 2C; Toyoshima 2009; Vangheluwe et al. 2009). The cytosolic nucleotide-binding (N-) domain contains a highly conserved Lys residue for ATP coordination in the KGA motif (Møller et al. 2010). The phosphorylation (P-) domain carries the Asp acceptor residue for autophosphorylation found in the DKTGT P-type ATPase signature motif. The TGE motif in the actuator (A-) domain regulates the access of water for the dephosphorylation reaction (Fig. 2; Olesen et al. 2004; Møller et al. 2010).
THE Ca2+ TRANSPORT MECHANISM EXEMPLIFIED BY SERCA1a
The crystal structures of the skeletal muscle isoform SERCA1a in the major conformational states have been solved. SERCA1a, therefore, became the archetypical Ca2+ pump for which the Ca2+ transport mechanism is described in great molecular detail, and which is summarized below (Fig. 2A; Toyoshima 2009; Møller et al. 2010; Primeau et al. 2018). In the high Ca2+ affinity E1 state, the cytosolic gate of SERCA1a is open, allowing 2–3 H+ to be displaced by two Ca2+ ions from the cytosol. The two Ca2+ ions bind sequentially and cooperatively at the Ca2+-binding sites I and II, leading to the stepwise repositioning of the Ca2+-binding residues (Fig. 2C). The induced fit following the binding of Ca2+ in the TM region is transmitted to the cytoplasmic domain via movement of M1–M4 (Sorensen et al. 2004; Gorski et al. 2017), which triggers ATP binding to a pocket in the N-domain, close to F487, K492, and K515 (Toyoshima et al. 2000). The adenosine of ATP binds at the N-domain and, with the help of the cofactor Mg2+, the γ phosphate of ATP is bridged to D351 at the P-domain. The subsequent SN2 nucleophilic reaction generates a high-energy phospho-intermediate of the pump (E1∼P) (Sorensen et al. 2004; Toyoshima and Mizutani 2004). At the same time, M1 is lifted toward the cytosolic side of the membrane and forms a kink that closes the cytosolic entry gate, leading to an occluded state. ATP forms a bridge between the N- and P-domains that generates tension, which is relieved by ATP hydrolysis causing the N-domain to move. This, in turn, creates tension between M3 and the A-domain, allowing the A-domain to rotate nearly 90°, which positions the 181TGE loop of the A-domain at the phosphorylation site in the P-domain. This loop prevents ADP or bulk H2O from reacting with the aspartylphosphate. The major rotation of the A-domain is also transmitted to the TM region, which distorts the high-affinity Ca2+-binding sites and creates a luminal gate through which Ca2+ can exit. Hence, the low Ca2+ affinity E2-P state is formed, which displays open Ca2+-binding sites facing the lumen (Olesen et al. 2007). The empty ion-binding sites are stabilized by two to three protons triggering the dephosphorylation reaction in the cytosolic domains and the closure of the luminal pathway in the TM domain. This is caused by a further rotation of the A-domain, which positions the TGE loop so that E183 fixes a water molecule and catalyzes an attack on the aspartyl phosphate. Consequently, phosphate and Mg2+ are released from the P-domain, which repositions the membrane helices and renders the occluded E2 state (Toyoshima and Nomura 2002; Toyoshima et al. 2004). Finally, the A-domain rotates away from the P-domain, which repositions the TM helices and recreates the high-affinity Ca2+-binding sites, thereby returning the pump to the E1 state (Ma et al. 2003). Although SERCA pumps countertransport protons when importing Ca2+ to the ER, it does not lead to a more basic ER luminal store because of the permeability of the ER membrane to small molecules (Le Gall et al. 2004; Bultynck et al. 2014).
An additional, Mg2+-bound structure was solved, representing a transition state between the closed Ca2+-free E2 and the open Ca2+-bound E1 state (Toyoshima et al. 2013; Winther et al. 2013). The high-affinity Ca2+-binding sites are only half formed and occupied by one or two Mg2+ ions, but the protein is prevented from undergoing autophosphorylation (Toyoshima et al. 2013; Winther et al. 2013). Mg2+ binding may temporarily delay the Ca2+ transport in muscle (Winther et al. 2013), but because Ca2+ entry is not blocked by Mg2+, Mg2+ ions may actually help to form the high-affinity Ca2+-binding sites (Toyoshima et al. 2013). The conformational transitions of SERCA1a are further facilitated by the protein–phospholipid interplay in the membrane. Interactions between phospholipids and specific R/K residues promote conformational transitions, whereas phospholipid interactions with W residues determine the protein tilt in the membrane. Together, lipid interactions lower the energy cost of the major movements of TM helices during the transport cycle (Norimatsu et al. 2017).
Besides crystallography, biochemical and biophysical approaches, such as fast kinetics and the intramolecular FRET method, provided insights on the dynamics of the Post–Albers cycle, revealing valuable information such as the rate-limiting steps (Dyla et al. 2017) and transition of conformations (Raguimova et al. 2018).
CONSERVATION AND MODULATION OF THE Ca2+ TRANSPORT MECHANISM
All reported P-type ATPase structures, including the Ca2+-ATPases SERCA1a (Toyoshima et al. 2000), SERCA2a (Sitsel et al. 2019), SERCA2b (Inoue et al. 2019), and PMCA (Gong et al. 2018), display an identical domain organization (Fig. 3A–C) and contain the key signature motifs for ATP hydrolysis and coupled Ca2+ transport (Vangheluwe et al. 2009; Palmgren and Nissen 2011). Although this indicates that the Ca2+ transport mechanism is highly conserved among Ca2+-ATPases (Møller et al. 2010; Bublitz et al. 2011), all Ca2+ pumps also present distinct properties that depend on isoform-specific sequences. Indeed, SERCA1a and the cardiac muscle isoform SERCA2a display different kinetic properties (Dode et al. 2003), but present strikingly similar structures (Inoue et al. 2019; Sitsel et al. 2019). The discrete properties of the different Ca2+-ATPases likely arise from isoform-specific residues that alter the intramolecular network of salt bridges and hydrogen bonds, which may change the molecular dynamics of the pump and affect the rate of conformational transitions (Sitsel et al. 2019). Other Ca2+-ATPase isoforms, like SERCA2b, SPCA1-2, and PMCA1-4, contain extra protein stretches, mainly at the amino and/or carboxyl terminus, which provide additional regulatory control (Chen et al. 2016). Finally, isoform-specific residues participate in regulation by including sites for protein interactions or posttranslational modifications ([PTMs]; Sitsel et al. 2019). In conclusion, each Ca2+-pump isoform presents a unique dynamic behavior and regulatory control. Together with a cell-type-specific expression profile and/or limited subcellular distribution, this ensures that each Ca2+-ATPase fulfils a distinct physiological role, and when dysregulated, may lead to specific pathological conditions, which will be further reviewed.
SERCA
The SERCA pump was identified and purified from skeletal muscle (Hasselbach 1964; MacLennan 1970), in which it plays an important role in muscle relaxation. Mammals contain three genes (ATP2A1–3) that express SERCA isoforms (SERCA1–3), which comprise around 1000 amino acids. Interestingly, no SERCA pumps are found in yeast, whereas invertebrates express only one SERCA gene, which corresponds to the mammalian SERCA2 isoform. In humans, alternative splicing of the messenger RNA (mRNA) transcripts of the three genes introduces additional variations in the carboxyl terminus rendering, in total, more than 10 SERCA protein variants. These variants are differently regulated and display distinct enzymatic properties and expression profiles, thereby fulfilling tissue-specific functions (Table 1; Periasamy and Kalyanasundaram 2007). SERCA proteins are selectively inhibited by the plant extract thapsigargin or its derivatives, by the synthetic compound 2,5-di(tert-butyl)-hydroquinone (BHQ), and the mycotoxin cyclopiazonic acid (CPA). Structural complexes of SERCA1a with these inhibitors show that thapsigargin binds to a pocket formed by M3, M5, and M7, whereas BHQ and CPA bind to overlapping pockets occupying the Ca2+ access channel delimited by TM segments M1–M4 (Toyoshima and Nomura 2002; Obara et al. 2005; Moncoq et al. 2007).
Table 1.
Gene | Isoform | Tissue distribution | Human disease and associated alterations on genetic/protein level | OMIM |
---|---|---|---|---|
ATP2A1 | SERCA1a | Fast twitch skeletal muscle (adult) | Brody disease (autosomal recessive inheritance): splice site mutations, premature stop codons, missense mutations | 108730 |
SERCA1b | Fast twitch skeletal muscle (fetal) | Myotonic dystrophy type 1: SERCA1b alternative mRNA splicing and dysregulated expression | ||
ATP2A2 | SERCA2a | Highly expressed in cardiac and slow twitch skeletal muscle, smooth muscle, neuronal cells | Darier–White disease (autosomal dominant): acrokeratosis verruciformis | 108740 |
SERCA2b | Ubiquitous | Heart failure: impaired SERCA2a protein expression and ATPase activity, mutations in PLB and reduced DWORF expression | ||
SERCA2c | Cardiac muscle (slow and fast twitch) skeletal muscle, myeloid and nonmyeloid cells, primary blood monocytes | Cancer: dysregulated protein expression, targeted SERCA inhibition as cancer therapy | ||
ATP2A3 | SERCA3a–f | 3a, d, f: cardiac muscle; nonisoform specific: smooth muscle and other nonmuscle cells (endothelial, epithelial cells, lung, and pancreas) | Gastric carcinomas, colon and lung cancer: dysregulated protein expression Diabetes: dysregulated protein expression |
601929 |
ATP2C1 | SPCA1a–d | Ubiquitous | Hailey–Hailey disease (autosomal inheritance): nonsense, splice-site, and nonconservative missense mutations; frameshift insertion and deletions Breast cancer: up-regulated protein expression |
604384 |
ATP2C2 | SPCA2 | Gastrointestinal tract, trachea, thyroid gland, salivary gland, mammary gland, prostate, brain (hippocampal neurons), keratinocytes (mRNA level) | Breast cancer: up-regulated protein expression | 613082 |
ATP2B1 | PMCA1a–e | Ubiquitous (1b: fetal; 1a: adult) 1a, 1c, 1e: brain; 1c in skeletal muscle |
Cardiovascular disease risk, preeclampsia, salt sensitivity: associated SNPs | 108741 |
ATP2B2 | PMCA2a–f | Brain (fetal and adult), cerebellar Purkinje cells, hair cells in the inner ear, mammary gland | Autosomal-recessive deafness: heterozygous point mutation (one case report) Autism: associated SNPs |
108733 |
ATP2B3 | PMCA3a–c | Widely expressed in the embryo Brain: cerebellum |
Early-onset spinocerebellar ataxia-1 (X-linked): point mutation | 300014 |
Aldosterone-producing adenomas: somatic point mutation | ||||
ATP2B4 | PMCA4a–g | Ubiquitous | Familial spastic paraplegia: point mutation Malaria resistance: associated SNPs |
108732 |
Overview of the specific characteristics of expression profile and disease links of Ca2+-ATPases. SERCA, SPCA, and PMCA are closely related based on phylogeny and their function as active Ca2+-transporters. However, differences in physiological function between these proteins arise through the expression of multiple isoforms, their structural differences, and specific tissue distribution. Additionally, many of these genes/proteins are linked to several human diseases and therefore may serve as interesting therapeutic targets. OMIM, Online Mendelian Inheritance in Man; mRNA, messenger RNA; PLB, phospholamban; DWORF, DWARF open reading frame; GWAS, genome-wide association studies; SNPs, single-nucleotide polymorphisms.
SERCA Isoforms
The two major skeletal muscle variants, SERCA1a and SERCA1b, are encoded by the ATP2A1 gene. The splice variant SERCA1b is expressed during embryonic myogenesis and in the neonatal stage in myotubes and myoblasts, as well as in regenerating adult muscles (Brandl et al. 1987; Zádor et al. 2007). Conversely, SERCA1a is the adult variant, which replaces SERCA1b during development (Brandl et al. 1987; Periasamy and Kalyanasundaram 2007). Its expression is highest in the fast twitch skeletal-muscle fibers, in which it serves as the relaxing factor by removing all Ca2+ from the myofilaments at the end of contraction. Functionally, SERCA1a displays a twofold higher maximal activity as compared with SERCA1b at high luminal Ca2+ concentrations (Zhao et al. 2015), which is a consequence of the SERCA1b-specific carboxy-terminal tail (994DPEDERRK in SERCA1b, instead of 994G in SERCA1a). Interestingly, a truncated variant of SERCA1 was identified and named S1T (Chami et al. 2001). It lacks a large part of the N- and P-domain, as well as M5–M10, hence, is inactive in Ca2+ pumping (Chami et al. 2001). The presence of S1T in ER-mitochondria contact sites increases Ca2+ leakage through ER, which results in increased Ca2+ transfer into mitochondria inducing apoptosis (Chami et al. 2008).
The housekeeping ATP2A2 gene generates three variants (SERCA2a–c) that were confirmed at the protein level and differ at their carboxyl termini (Vandecaetsbeek et al. 2009). SERCA2a shares 84% sequence identity with SERCA1a, and is found in the heart, slow-twitch skeletal muscle, and smooth muscle cells (Lytton et al. 1989; Zarain-Herzberg et al. 1990). In humans, SERCA2a removes between 70% and 90% of the elevated cytosolic Ca2+ after cardiomyocyte contraction (Bers 2002), which is stored inside the sarcoplasmic reticulum (SR) for the next contraction. Therefore, SERCA2a is a major determinant of cardiomyocyte relaxation, but it also determines the SR Ca2+ content that controls the Ca2+ release for contraction (Periasamy et al. 2008).
SERCA2b is ubiquitously expressed in the ER and is the main SERCA isoform in the brain (Miller et al. 1991; Baba-Aissa et al. 1998; Mata and Sepúlveda 2005). SERCA2b keeps cytosolic Ca2+ levels in the submicromolar range while filling the ER with 0.5–1 mm Ca2+ (Suzuki et al. 2016), which is important for ER homeostasis and protein maturation such as N-linked glycosylation (Helenius and Aebi 2001) and disulfide bridge formation (Michalak et al. 2002). In addition, SERCA2b mRNA levels increase 3- to 4-fold when ER stress is induced (Caspersen et al. 2000). The four amino acids that comprise the short SERCA2a-specific carboxyl terminus (994AILE) are replaced by 49 amino acids in SERCA2b, which is referred to as the 2b-tail. The 2b-tail consists of an additional TM helix (M11) and a luminal extension (Lytton et al. 1992; Verboomen et al. 1992; Vandecaetsbeek et al. 2009). The luminal extension is predicted to interact near luminal loop L7/8 and M11 near M10 (Vandecaetsbeek et al. 2009). Both regions independently alter the kinetic properties of the pump. The luminal extension increases the intrinsic Ca2+ affinity and slows the E1∼P to E2-P transition, whereas M11 lowers the maximal turnover rate, mainly by reducing E2-P to E2 and E2 to E1 conversion rates (Verboomen et al. 1992; Dode et al. 2003; Vandecaetsbeek et al. 2009; Clausen et al. 2012; Gorski et al. 2012). Recently, a first structure of SERCA2b was solved that depicts the position of M11 on M10, that is, isolated from the other TM segments (Inoue et al. 2019), at a similar position as neuroplastin (NPTN) binds to PMCA1 (Gong et al. 2018). The high B-factor values of M11 in SERCA2b and the missing electron densities in the cytosolic and luminal extensions may suggest that the 2b-tail interaction is flexible and may be prone to regulation (Inoue et al. 2019).
SERCA2c (with the carboxy-terminal tail 994VLSSEL) mRNA was detected in epithelial, mesenchymal, hematopoietic cell lines, and primary human monocytes (Gélébart et al. 2003), and the protein form is confirmed in the human heart (Dally et al. 2010). It displays a lower Ca2+ affinity than SERCA2a and SERCA2b, and a similar Vmax compared with SERCA2b (Dally et al. 2006).
In humans, six different SERCA3 variants (SERCA3a–f) have been identified. SERCA3 is expressed at high levels in the intestine and lymphatic tissue, platelets (Wuytack et al. 1994; Bobe et al. 2004), Purkinje neurons (Baba-Aissa et al. 1996), hematopoietic cell lineages, epithelial cells, fibroblasts, and endothelial cells (Anger et al. 1994). SERCA3 has also been found in low levels in smooth muscle cells (Wu et al. 2001). Together with SERCA2c, SERCA3a, -3d, and -3f were described in cardiomyocytes (Dally et al. 2009, 2010). In contrast to the more uniform distribution of SERCA3a, SERCA3d and SERCA3f displayed restricted localizations around the nucleus and in the subplasmalemmal area, respectively (Dally et al. 2009, 2010). Although the SERCA3 sequence is about 75% similar to other SERCA isoforms, all three isoforms displayed a lower apparent Ca2+ affinity than SERCA1/2 isoforms, turning them only active at high cytosolic Ca2+ concentrations. This is because of a decreased E2 to E1 transition rate and an increased rate of Ca2+ dissociation in E1 as compared with SERCA1a (Wuytack et al. 1995; Dode et al. 2002; Chandrasekera et al. 2009). SERCA3 also displays an increased rate of E2-P dephosphorylation and a higher pH optimum at 7.5–9.0 (Dode et al. 2002; Periasamy and Kalyanasundaram 2007). A potential role of SERCA3 in cell differentiation associated with remodeling of ER Ca2+ homeostasis was suggested by several studies on endothelial, myeloid, and colon epithelial cells (Launay et al. 1999; Mountian et al. 1999; Gélébart et al. 2002). However, the specific function of the individual SERCA3 variants in many other tissues remains incompletely understood.
Regulation of SERCA Isoforms
As a major regulator of cytosolic and ER/SR luminal Ca2+ concentrations, SERCA isoforms are extensively regulated to fine-tune their activity according to the physiological requirements. This tight regulation takes place both at the expression and Ca2+ transport level (see detailed reviews elsewhere; Vangheluwe et al. 2005a; Vandecaetsbeek et al. 2011; Stammers et al. 2015). Here, we will focus on an emerging general concept of SERCA regulation by a family of small TM proteins that regulate SERCA Ca2+ affinity in an isoform- and/or cell-type-specific manner (Primeau et al. 2018).
Historically, phospholamban (PLB) was identified as the first and main small TM protein regulator of SERCA2a that is part of the β-adrenergic control system in the heart (Wegener et al. 1989). PLB is a 52-amino-acid, single-span integral membrane protein that is highly expressed in cardiac muscle together with SERCA2a, and at a lower level in smooth and slow-twitch skeletal muscles. The monomeric PLB forms a reversible one-to-one complex with SERCA by binding to a groove involving M2, M4, M6, and M9. This direct interaction reduces the apparent Ca2+ affinity of the pump (Periasamy and Kalyanasundaram 2007). The SERCA2a inhibition is reversed at high cytosolic Ca2+ concentrations (>1 µm) or during β-adrenergic stimulation via phosphorylation of PLB by protein kinase A and/or Ca2+/calmodulin (CaM) kinase II. Subsequently, PLB congregates into pentamers, serving mainly as a reserve pool of the protein (Kimura et al. 1997). The higher SERCA2a activity improves cardiac relaxation and, via its impact on the SR Ca2+ load, also the contraction (for reviews, see Simmerman and Jones 1998; MacLennan and Kranias 2003; Kranias and Hajjar 2012; Akin et al. 2013).
The small integral membrane protein sarcolipin ([SLN], 31 residues) appeared as a functional and structural homolog of PLB (reviewed in MacLennan and Kranias 2003) that interacts with SERCA at the same binding site as PLB (Toyoshima et al. 2013; Winther et al. 2013), lowering both the apparent Ca2+ affinity and Vmax of the pump (Gorski et al. 2013; Sahoo et al. 2013). SLN contains a shorter amino terminus that holds a regulatory phosphorylation site, and a longer carboxy-terminal extension that is functionally important by docking SLN to the luminal side of the SERCA protein (Gorski et al. 2012). SLN appears more stably associated with the pump than PLB because high Ca2+ concentrations do not dissociate SLN (Shaikh et al. 2016). SLN is coexpressed with SERCA1a and SERCA2a in skeletal muscle or with SERCA2a in atrial tissue, but is absent in the ventricles of the heart (Vangheluwe et al. 2005b; Babu et al. 2007a). SLN modifies the atrial contractility and β-adrenergic response (Babu et al. 2007b), which differ from ventricles (Vangheluwe et al. 2005b). Moreover, SLN knockout mice are susceptible to atrial arrhythmias and remodeling dependent on aging (Xie et al. 2012), further showing the physiological relevance of SLN in the atria. In HEK-293 cells coexpressing SLN, PLN, and SERCA, SLN and PLB potentially form a ternary, superinhibitory complex with SERCA (Asahi et al. 2002), whereas SLN directly binds to PLN, inhibiting the formation of PLN pentamers, which may contribute to the superinhibitory effect on SERCA (Asahi et al. 2004). However, physiological evidence showing that SERCA would be superinhibited in atria is not yet available. In skeletal muscle with a high SLN/SERCA1a coexpression, SLN may uncouple SERCA1a's ATPase activity from Ca2+ transport (Sahoo et al. 2013), resulting in energy dissipation and heat production without transporting Ca2+. This uncoupling behavior appears physiologically important for muscle-based nonshivering thermogenesis in mammals. Interestingly, nonshivering thermogenic mechanisms through SLN-mediated uncoupling of SERCA1a may be considered as a target for obesity treatment (Bal et al. 2012; Periasamy et al. 2017).
Solved complexes of SERCA1a with SLN (Toyoshima et al. 2013; Winther et al. 2013) or PLB4, a gain-of-function mutant of PLB (Akin et al. 2013), provided structural insights into the modulation of Ca2+ transport by small TM regulators. Both regulators bind in the same TM region of the pump, but the observed structural differences may indicate that PLB and SLN stabilize a distinct conformation of SERCA, suggesting that the two regulators present a different mode of action. Indeed, the SERCA1a-SLN complex adopts an E1-like conformation with Mg2+ at the ion-binding sites (Toyoshima et al. 2013; Winther et al. 2013), whereas the SERCA1a-PLB4 structure represents an E2-like structure in which Ca2+ and Mg2+ binding is precluded (Akin et al. 2013).
More recently, several other related small TM proteins were discovered in annotated long noncoding RNAs such as myoregulin (MLN), DWARF open reading frame (DWORF), endoregulin (ELN), and another-regulin (ALN). These proteins present a similar primary and secondary structure as PLB and SLN, contain a conserved LFxxF sequence in the TM region, presumably bind to the same pocket in the SERCA TM region, and regulate the apparent Ca2+ affinity of the Ca2+ pump. ELN and ALN are mainly expressed in nonmuscle tissue (Anderson et al. 2016), in which they diminish the apparent Ca2+ affinity of SERCA2b and SERCA3a. MLN appears as a skeletal muscle–specific regulator of SERCA1a and SERCA2a (Anderson et al. 2015). DWORF is expressed in soleus, the ventricles of the heart, and the diaphragm, in which it may regulate SERCA1a, SERCA2a/b, and SERCA3a/b isoforms (Nelson et al. 2016). Interestingly, DWORF exerts no direct effect on the apparent Ca2+ affinity of the pump, but instead may stimulate SERCA activity by displacing other regulators like PLB (Makarewich et al. 2018). This points to a complex, tissue-specific interplay between various regulatory proteins, but how this dynamically adapts the SERCA activity to the local physiological demand and how this is modified in disease conditions remains to be further elucidated.
Many other regulators of SERCA have been documented, which directly interact with the Ca2+ pump and regulate its activity (e.g., Atrap, sarcalumenin, S100A1, histidine-rich Ca2+-binding protein, ERdj5; reviewed in Vangheluwe et al. 2005a; Vandecaetsbeek et al. 2011). Other interactions control organelle contact sites at the ER. The ER-localized metazoan-specific autophagy protein VMP1 prevents the PLB/SLN inhibition of SERCA, which activates SERCA and reduces contact formation between the ER and isolation membranes (autophagosome precursors), mitochondria, lipid precursors, and endosomes (Zhao et al. 2017). The importance of SERCA in ER-mitochondria contact sites has been recently reviewed in Krols et al. (2016), Chemaly et al. (2018), and Gutiérrez and Simmen (2018). The role of SERCA pumps in ER-mitochondria communication and cell death is also emerging as SERCA is regulated by mitochondrial antiapoptotic proteins HS1-associated protein HAX-1 (Vafiadaki et al. 2009; Bidwell et al. 2018) and B-cell lymphoma 2 (Bcl-2) (Dremina et al. 2004), proapoptotic protein p53 (Giorgi et al. 2015), and palmitoylated calnexin during short-term ER stress (Lynes et al. 2013).
SERCA in Disease
SERCA isoforms and their regulators preserve the required Ca2+ balance in cells, whereas mutations and dysregulation of these proteins are implicated in a variety of pathological conditions.
Homozygous or compound heterozygous loss-of-function mutations in the skeletal muscle isoform SERCA1 are associated with autosomal recessive Brody myopathy (Odermatt et al. 1996; Guglielmi et al. 2013), an exercise-induced impairment of fast-twitch skeletal muscle relaxation (Odermatt et al. 1996). Brody myopathy was also described in cattle (Charlier et al. 2008; Drögemüller et al. 2008) and zebrafish (Hirata et al. 2004), whereas loss of SERCA1 in mice is lethal as a result of the respiratory failure and hypercontracture injury of the diaphragm (Pan et al. 2003). ATP2A1 is also one of the genes that is aberrantly spliced in myotonic dystrophy type 1 muscle, leading to expression of the SERCA1b splice variant instead of SERCA1a in normal muscle (Zhao et al. 2015). Also, improving SERCA1a activity is considered for therapy to restore abnormalities of Ca2+ homeostasis in Duchenne muscular dystrophy patients (Morine et al. 2010; Goonasekera et al. 2011).
Homozygous SERCA2 knockout mice are not viable, whereas heterozygous SERCA2 mice appeared healthy, but showed reduced cardiac muscle contractility (Periasamy et al. 1999). Also, heterozygous SERCA2 mice develop squamous cell tumors more frequently at older age (Liu et al. 2001). When put on a high-fat diet, SERCA2 heterozygous mice develop glucose intolerance, diminished insulin secretion, and elevated β-cell ER stress and death, suggesting that restoring SERCA2 activity may represent a viable strategy to improve glucose homeostasis (Kang et al. 2016; Tong et al. 2016). In humans, the absence of one functional allele of SERCA2 triggers an inherited, dominant skin disorder called Darier disease, which is characterized by distinctive nail abnormalities, warty papules, and plaques mainly on the chest, neck, back, ears, forehead, and groin (Sakuntabhai et al. 1999; Dhitavat et al. 2003; Engin et al. 2015). Most disease mutations display a loss-of-function phenotype causing haploinsufficiency, although a gain-of-function because of a leaky Ca2+ pump was also proposed for some Darier mutants (Kaneko et al. 2014).
A reduced expression and activity of SERCA2a strongly contributes to the poor cardiac contractility in patients with end-stage heart failure (for review, see Lipskaia et al. 2010). However, Darier disease patients show no predisposition to develop cardiomyopathy (Mayosi et al. 2006). Although this observation may raise the question on the causality of SERCA2a dysfunction in heart failure, cardiac-specific SERCA2a knockout in adult mice induces heart failure within weeks (Andersson et al. 2009). Moreover, heterozygous SERCA2 mice are predisposed to some, but not all forms of heart failure (Prasad et al. 2015). In humans, several disease-causing PLB mutations were identified in heritable forms of dilated cardiomyopathy that lead to a chronic inhibition of SERCA2a (MacLennan and Kranias 2003). DWORF expression was also reduced in ischemic failing human hearts, which may lead to a stronger PLB inhibition (Nelson et al. 2016). Consequently, restoring SERCA2a activity is considered as a key therapeutic strategy for end-stage heart failure. Indeed, increasing DWORF expression (Makarewich et al. 2018) or lowering PLB expression (Minamisawa et al. 1999) enhances contractility and prevents heart failure in mouse models of dilated cardiomyopathy. Moreover, adeno-associated viral gene transfer of SERCA2a has beneficial effects on the contractility and remodeling in small and large animal models of heart failure (for review, see Lipskaia et al. 2010; Park and Oh 2013; Gorski et al. 2015). However, in patients, the outcome of the clinical trials was disappointing, presumably caused by an inadequate gene delivery in the diseased human heart (Greenberg et al. 2016). Alternative strategies are currently explored to achieve SERCA2a activation, for example, via small activator compounds or better viral gene delivery methods (Samuel et al. 2018).
A strong reduction in SERCA2 is also observed in limb-girdle muscular dystrophy type 2A, suggesting that restoring SERCA2 activity may also be of therapeutic interest (Toral-Ojeda et al. 2016). In contrast, SERCA inhibition by analogs of the highly potent and selective SERCA inhibitor thapsigargin is considered for prostate cancer treatment. This approach relies on an inactive thapsigargin prodrug that is cleaved by a prostate-specific antigen in the malignant tissue environment, leading to local SERCA2b inhibition and apoptosis of the cancer cells (Denmeade et al. 2012; Mahalingam et al. 2016).
Altered SERCA3 expression levels are reported in diabetes and cancer (Varadi et al. 1999; Xu et al. 2012). Also, SERCA3f mRNA is up-regulated in idiopathic forms of dilated cardiomyopathy, which may be a marker of ER stress induction (Dally et al. 2009). SERCA3 further plays a role in progesterone-triggered Ca2+ signaling in the MCF-7 breast cancer cell line, modulating cell proliferation and death (Azeez et al. 2018). However, SERCA3 knockout mice do not show structural malformations and grow normally into adulthood without a clear disease phenotype. Instead, they show an extended bleeding time, defective platelet adhesion, and thrombus growth as a consequence of reduced ADP secretion (Elaib et al. 2016).
SPCA
SPCA is the most recently identified active Ca2+-transporter and was first described in Saccharomyces cerevisiae, named plasma membrane-related Ca2+-ATPase (PMR1) (Rudolph et al. 1989). Later, the mammalian SPCA isoforms were cloned and characterized (Gunteski-Hamblin et al. 1992; Wootton et al. 2004; Vanoevelen et al. 2005; Xiang et al. 2005). In humans, two genes (ATP2C1 and ATP2C2) code for SPCA proteins that share 63% sequence identity. SPCA1 represents evolutionarily the older and most widespread isoform, whereas SPCA2 emerged later in vertebrate evolution at the rise of tetrapods (Table 1; Vangheluwe et al. 2009; Pestov et al. 2012).
SPCA proteins contain about 950 residues and most likely present a similar domain organization and Ca2+ transport mechanism as SERCA1a (Fig. 3B). However, SPCA isoforms are also structurally and mechanistically different from SERCA1a. In addition to Ca2+, SPCA proteins also transport Mn2+ ions, which depend on structural elements in the amino terminus and TM domain that regulate ion selectivity (Wei et al. 1999; Mandal et al. 2003; Vangheluwe et al. 2009). Also, SPCA isoforms contain a single Ca2+-binding site that corresponds to site II of SERCA proteins (Fig. 2C). SPCA pumps transport Ca2+ at lower maximal turnover rates, but with much higher apparent affinities than SERCA1a (Van Baelen et al. 2001) because of a slower conversion of E1∼P(Ca2+) to E2-P (Dode et al. 2005). Different from SERCA, SPCA isoforms present an enhanced rate of E2-P hydrolysis, which is pH independent, suggesting that SPCA1 may not countertransport protons (Dode et al. 2005, 2006). SPCA proteins are generally more compact and lack several regulatory elements that are present in SERCA (Fig. 3A,B). In addition, SPCA1 and SPCA2 contain longer amino and carboxyl termini, of which the regulatory roles are gradually emerging (Wei et al. 1999; Feng et al. 2010; Smaardijk et al. 2017; Chen et al. 2019). SPCA proteins are inhibited by high (micromolar) concentrations of the commonly used SERCA inhibitors thapsigargin, BHQ, and CPA. Specific and potent SPCA inhibitors are not yet reported, but the distinct structure–activity relationship of thapsigargin inhibition in SERCA versus SPCA1 indicates that SPCA-specific inhibitors may be developed based on the thapsigargin scaffold (Chen et al. 2017).
The higher apparent Ca2+ affinity renders the SPCA activity less sensitive to fluctuations in the cytosolic Ca2+ concentration than SERCA (Dode et al. 2005). SPCA ensures a constant filling of the Golgi with Ca2+ and also Mn2+. Both ions are cofactors of many Golgi enzymes required for adequate protein processing by posttranslational modifications and trafficking. In particular, there is a Ca2+ gradient across the secretory pathway from the ER (0.5–1 mm), cis-Golgi (250 µm) to the trans-Golgi network (TGN) (100 µm) (Pizzo et al. 2011; Suzuki et al. 2016), which is generated by the combined activities of SERCA (in ER and early Golgi compartments) and SPCA (all Golgi compartments) (Fig. 1A). The luminal Ca2+/Mn2+ levels need to fall in a physiological window because either a diminished (Okunade et al. 2007) or excessive SPCA1 (Smaardijk et al. 2018) activity induces signs of Golgi stress such as Golgi swelling and fragmentation.
SPCA Isoforms
SPCA1 is ubiquitously expressed and displays a predominant distribution in the trans-Golgi (reviewed in Vangheluwe et al. 2009). In humans, alternative splicing of the ATP2C1 gene generates four experimentally confirmed protein variants that differ at the carboxyl terminus (SPCA1a–d), among which the SPCA1c is an inactive form and SPCA1d is the longest (Table 1; Missiaen et al. 2004; Micaroni et al. 2016). However, their tissue-specific expression pattern and subcellular localization remain incompletely understood. SPCA1 is the main SPCA isoform in brain, in which it plays a crucial role in establishing neural polarity during development (Sepúlveda et al. 2009). Neurons are highly sensitive to Ca2+ dyshomeostasis in the Golgi (Sepúlveda et al. 2009) and to Mn2+ toxicity (Olanow 2004; Sepúlveda et al. 2012b).
SPCA2 expression is more restricted to the brain, testis, gastrointestinal and respiratory tracts, and to actively secreting cells like prostate, thyroid, salivary, and mammary glands (Table 1), suggesting more specialized functions than SPCA1 (Vanoevelen et al. 2005; Xiang et al. 2005). Compared to SPCA1, SPCA2 displays a broader subcellular localization in the secretory pathway such as the Golgi, ER, and secretory vesicles (Vanoevelen et al. 2005; Xiang et al. 2005; Feng et al. 2010; Pestov et al. 2012).
The SPCA1 and SPCA2 proteins have distinct functional properties. The SPCA2 maximal activity is 2.5-fold higher than SPCA1d, but presents a similar apparent Ca2+ affinity. This relates to an increased rate of E1∼P to E2-P conversion, a reduced rate of the E2 to E1 transition, and an enhanced rate of the E2-P dephosphorylation (Dode et al. 2006). However, purified SPCA1a displays a lower apparent Ca2+ affinity than purified SPCA2, but a twofold higher ATPase activity in the presence of Ca2+. In contrast, the maximal activity and apparent affinity for Mn2+ are comparable for both isoforms. The distinct Ca2+-dependent properties of SPCA1a relate to the presence of an amino-terminal Ca2+-binding EF-hand-like motif that is absent in SPCA2. Ca2+ binding to the amino-terminal EF-hand-like motif promotes the activity of SPCA1a by facilitating the autophosphorylation step. This may be important in cells with a high Ca2+ load, such as mammary gland cells during lactation, or in cells with a low ATP content, such as keratinocytes (Chen et al. 2019). Also, the S. cerevisiae ortholog PMR1 contains a Ca2+-binding EF-hand-like motif in the amino terminus that is important for substrate affinity and ion selectivity (Wei et al. 1999). Based on a partial proteolytic digestion analysis, a model was proposed that the amino terminus of PMR1 interacts with the carboxy-terminal half of the protein to control its functional properties (Wei et al. 1999), but the precise molecular mechanism remains unclear.
Regulation of SPCA
The activity and expression of multiple Ca2+ transporters in the mammary gland is tightly coordinated during pregnancy, lactation after parturition, and the process of involution (reviewed in Cross et al. 2014). These remarkable changes are required to support the release of large amounts of Ca2+ in the milk (8–60 mm, depending on the species) (Neville 2005). SPCA1 and SPCA2 expression is highly up-regulated in mammary glands during lactation (McAndrew et al. 2011; Cross et al. 2013). Via its amino- and carboxy-terminal extensions, SPCA2 directly interacts with and activates the plasma membrane Ca2+ channel Orai1, which leads to an increased Ca2+ influx. This happens independent of a change in the intracellular Ca2+ store content or the relocalization of STIM1, which typically activates Orai1 dependent on Ca2+ store depletion. The SPCA2-Orai1 coupling is therefore dubbed “store-independent Ca2+ entry” (SICE) (Feng et al. 2010). Once activated, SPCA2 transfers the incoming Ca2+ into the Golgi/secretory pathway (Smaardijk et al. 2017). Later, SPCA1 was also found to induce SICE through Orai1, which controls the luminal Ca2+ content of the Golgi/secretory pathway (Smaardijk et al. 2018). The incoming Ca2+ via Orai1 occurs at the basolateral side of the mammary gland epithelial cells (Cross et al. 2013). At the apical side, Ca2+ is delivered into the milk by direct Ca2+ transport by PMCA2 over the plasma membrane (60%) and by secretion of Ca2+ that was sequestered in the secretory pathway by SERCA and SPCA (40%) (Reinhardt et al. 2004; Faddy et al. 2008; Cross et al. 2013).
The TGN sorts proteins destined for endo-lysosomes, secretory storage granules, and the plasma membrane via several parallel sorting systems that package cargo into specific vesicles. Of interest, SPCA is activated by the actin filament severing protein cofilin-1, which determines protein sorting and secretion (Kienzle et al. 2014). The transient increase of the local luminal Ca2+ concentration in the TGN induces the polymerization of luminal 45-kDa, Ca2+-binding protein (Cab45) (Crevenna et al. 2016). In its polymerized state, Cab45 selectively interacts with specific cargo molecules for secretion (von Blume et al. 2012; Crevenna et al. 2016). More recently, it was shown that SPCA1 also influences cell contractility by disrupting actin dynamics and the localization of cofilin-1. This process is important during embryonic development to organize neuronal tube closure, which goes wrong in the ATP2C1 knockout embryo (Brown and Garcia-Garcia 2018).
SPCA Isoforms in Disease
In humans, heterozygous mutations in ATP2A2 result in Darier disease, whereas heterozygous ATP2C1 mutations cause Hailey–Hailey disease (HHD), a related chronic skin disease with similar symptoms, showing blisters and itchy erosions mainly at the sites of sweating and friction such as the groin and the axillar regions (Hu et al. 2000). Although most studied SPCA1 mutations show a loss of transport function (Fairclough et al. 2003), some mutants are unable to couple with Orai1 to elicit a SICE response (Smaardijk et al. 2018). Irrespective of the mechanism, disease-causing mutations appear less potent to fill the Golgi/secretory pathway with Ca2+ (Smaardijk et al. 2018), which may cause haploinsufficiency.
Remarkably, both Atp2a2+/− (Prasad et al. 2005) and Atp2c1+/− (Okunade et al. 2007) heterozygous mice display an increased incidence of squamous tumors, which markedly differs from the skin disease phenotype in humans. Loss of Atp2c1 is embryonically lethal as a result of the failure of neural tube closure (Okunade et al. 2007; Brown and Garcia-Garcia 2018).
SPCA1 levels are increased in basal-type breast cancer, whereas SPCA2 is increased mainly in luminal types of breast cancer (Grice et al. 2010; Dang et al. 2017). Knockdown of SPCA1 in the basal-type breast cancer cell line MDA-MB-231 leads to impaired processing of the insulin-like growth factor 1 receptor (IGF1R) (Grice et al. 2010), suggesting that SPCA1 inhibition may be of interest for breast cancer therapy (Christopoulos et al. 2018). Furthermore, the up-regulation of SPCA2 leads to a constitutive activation of Orai1 and, consequently, a pathological elevated cytosolic Ca2+ concentration that increases proliferation and oncogenic activity (Feng et al. 2010). Preventing the pathological Ca2+ influx from the Orai1-SPCA2 complex, or inhibition of SPCA1 activity, may be considered for breast cancer therapy (Feng et al. 2010). Furthermore, breast calcification is a radiographic feature that is linked to poorer survival in breast cancer patients. In vitro microcalcifications in human breast cancer cells depend on SPCA Ca2+ transport activity (Dang et al. 2017).
SPCA1 is required as a host factor for virus maturation and spreading possibly by maintaining high luminal Ca2+ and Mn2+ levels for (viral) protein maturation. Thus, SPCA1 may be an interesting target for viral diseases, in particular, involving flaviviruses and togaviruses (Hoffmann et al. 2017). Furthermore, disrupting PMR1 in various model organisms counters the cytosolic Ca2+ toxicity that is elicited by the ectopic expression of α-synuclein, indicating that inhibition of PMR1/SPCA may be of interest to lower α-synuclein toxicity, the major constituent of the brain plaques present in the brain of Parkinson's disease patients (Büttner et al. 2013).
Finally, SPCA1 may be implicated in manganism, a Parkinson-like disease that affects miners, welders, and steel and battery workers because of Mn2+ intoxication. Toxic Mn2+ accumulation in brain areas correlates with a higher presence of SPCA1 (Sepúlveda et al. 2012a). In addition, Mn2+ overload affects survival of neurons and glia by inhibiting SPCA activity and inducing Golgi fragmentation (Sepúlveda et al. 2012b), a phenomenon also described in other neurodegenerative diseases (Gonatas et al. 2006). Conversely, PMR1 in yeast is important for Mn2+ detoxification by uptake of Mn2+ in the secretory pathway for subsequent secretion. Likewise, SPCA1 may play a role in Mn2+ detoxification in the liver (Leitch et al. 2011).
PMCA
PMCA was identified (Schatzmann 1966) and purified (Niggli et al. 1979) in erythrocytes, in which it is the only type of Ca2+ pump. Over the years, the simple view that PMCAs reduce cytosolic Ca2+ to avoid cellular overload shifted to a more complex picture in which PMCA isoforms fine-tune local Ca2+ events that originate close to the plasma membrane. Because it is impossible to include the vast literature on PMCAs in this review, the reader is also referred to Brini and Carafoli (2011), Strehler (2015), and Stafford et al. (2017).
In mammals, four PMCA isoforms and more than 20 spliced variants were described with diverse kinetic properties (Strehler and Zacharias 2001), each comprising about 1200 amino acids (Table 1). PMCA displays a Ca2+/ATP molar stoichiometry of 1/1 as does SPCA while it behaves like a Ca2+/H+ countertransporter akin to SERCA (Hao et al. 1994; Salvador et al. 1998). Like other Ca2+-ATPases, PMCA isoforms contain three cytoplasmic domains and 10 TM helices (Fig. 3C), but they fundamentally differ in regulatory regions. PMCAs are marked by the presence of an autoinhibitor domain at the carboxyl terminus, which constitutes the CaM-binding domain. Recently, a cryo-EM structure of PMCA1 in the E1-Mg2+ intermediate state has been resolved to about 4 Å in complex with the protein NPTN/basigin (BASI) (Gong et al. 2018), which recently identified auxiliary subunits that form a complex with PMCA to regulate Ca2+ clearance in different types of cells (Korthals et al. 2017; Schmidt et al. 2017). NPTN/BASI binds primarily with M10 of PMCA1 (Fig. 3C; Go and Soboloff 2018; Gong et al. 2018), apparently at the same site as the 2b-tail binds in the SERCA2b protein (Inoue et al. 2019). Heterotetramers of two PMCA1s and two regulatory subunits have been proposed to improve the efficiency of Ca2+ transport by facilitating the transition from E2 to E1 (Gong et al. 2018).
PMCA Isoforms
The diverse set of PMCA protein variants is generated via alternative processing of the primary transcripts of the four PMCA genes (named ATP2B1–4) in two main sites (A and C) (Table 1). Splicing site A (rendering splice variants w, x, z) is located in the first cytosolic domain, between the interaction sites for the CaM-binding domain and a binding site of acidic phospholipids. Splicing at this site not only changes the length of the domain, but also alters the autoinhibition and sensitivity toward acidic phospholipid regulation (Brini et al. 2010). Splicing site C (variants a, b, c, d, e) is situated in the middle of the carboxy-terminal CaM-binding domain, and affects the regulatory and functional properties of PMCA, such as the affinity of the pumps for Ca2+-CaM and the kinetics of their activation (reviewed in Strehler and Zacharias 2001; Krebs 2015; Strehler 2015). In general, “a” variants display a shortened C-tail and present a lower CaM affinity, but increased basal activity and moderate CaM stimulation than “b” splice variants. However, there are also differences between “b” variants; for example, PMCA4b shows high basal activity and is highly stimulated by CaM, but PMCA2b is marked by high basal activity, but a modest stimulation by CaM (Elwess et al. 1997).
All tissues express at least one PMCA isoform, but their abundance and distribution appear to be isoform specific. However, not all variants have been confirmed at the protein level because of the lack of specific antibodies. PMCA1 is ubiquitous, although variant expression levels and distribution change especially during development. It is the earliest PMCA isoform to be expressed in the embryo and seems to exert an essential housekeeping or developmental function (Okunade et al. 2004). There is a switch from PMCA1b to PMCA1a during brain development (Brandt and Neve 1992) that reflects a differential expression of variants according to the developmental stage. PMCA2 displays a more restricted distribution, but is abundant in cerebellar Purkinje neurons and hair cells in the inner ear (Furuta et al. 1998). PMCA2 is also highly expressed in the apical membrane of the epithelia of lactating mammary glands, in which it is important for Ca2+ extrusion into the milk (Reinhardt et al. 2004). PMCA3 is the least characterized isoform and is broadly expressed in the developing embryo, but, like PMCA2, shows a more restricted distribution in adults, in which it is mainly expressed in the brain, particularly in cerebellar synaptic terminals and choroid plexus (Eakin et al. 1995; Marcos et al. 2009). Finally, PMCA4 is present in most cells and tissues as PMCA1, and therefore fulfills a housekeeping role.
Regulation of PMCA
In resting conditions at about 100 nm cytosolic Ca2+, PMCAs are trapped in an autoinhibited state because of blocking of the ATP-binding site by the carboxy-terminal CaM-binding domain that interacts with two cytosolic regions of PMCA. The first region is located within the A-domain, preceding an acidic phospholipids-binding site. The second region is located within the N-domain after the autophosphorylation site. At increased cytosolic Ca2+ concentrations, Ca2+-bound CaM interacts with the CaM-binding domain of PMCA, which relieves the autoinhibition. This leads to a significant increase in the Ca2+ affinity and turnover rate of PMCA (for review, see Brini and Carafoli 2009; Lopreiato et al. 2014). Ca2+-loaded CaM binds to two sites along the carboxy-terminal regulatory domain, which facilitates a two-step PMCA activation mechanism that allows a tight control of intracellular Ca2+ levels over a broad range of physiological conditions in eukaryotic cells (Tidow et al. 2012).
The C-tail of PMCA is subjected to phosphorylation by protein kinases A and C (Zylinska et al. 1998) and tyrosine kinases such as Src (Ghosh et al. 2011, 2016), regulates self-association (Kosk-Kosicka and Bzdega 1988), and undergoes cleavage by proteases such as caspases or calpain (Pászty et al. 2002; Guerini et al. 2003). Furthermore, “b” splice variants contain a PDZ domain for protein–protein interactions. Different PDZ proteins from the membrane-associated guanylate kinase (MAGUK) family of scaffold proteins are involved in the recruitment and retention of different PMCA splice variants in specific Ca2+ microdomains (Kim et al. 1998; DeMarco and Strehler 2001; Schuh et al. 2003; Strehler 2015). These interactions also allow the regulation of specific PMCAs by affecting the affinity for Ca2+ or CaM. Interestingly, besides suppressing SERCA expression and stability (Dremina et al. 2004), Bcl-2 may also influence cell fate by suppressing the Ca2+ extrusion via PMCA (Ferdek et al. 2012).
Protein–lipid interactions also play a key role in PMCA function. There are two binding sites for acidic phospholipids that modulate PMCA activity (Niggli et al. 1981), one that precedes M3, and another one that is positioned in the carboxy-terminal region. The lipid composition of cellular membranes or subdomains like lipid rafts also affect PMCA activity and localization (Sepúlveda et al. 2006; Jiang et al. 2007; Marques-da-Silva and Gutiérrez-Merino 2014). Conversely, changes in the lipid environment, for example, in the aging brain or in the development of age-dependent neurodegenerative disorders, drastically affect activity and distribution of PMCAs (Michaelis et al. 1996; Farooqui et al. 1997; Jiang et al. 2012).
PMCA in Disease
The relatively low number of diseases that are linked to PMCA mutations contrasts with the high diversity of PMCA variants and physiological roles, which may point to functional redundancy between isoforms. Still, all PMCA isoforms have been implicated in specific human disorders.
Single-nucleotide polymorphisms in the ATP2B1 gene are associated with higher risk for hypertension and cardiovascular disease (Cho et al. 2009; Hong et al. 2010; Kobayashi et al. 2012; Wang et al. 2017). A reduced PMCA1 expression may lead to a raised blood pressure associated with elevated cytosolic Ca2+ and vascular remodeling (Kobayashi et al. 2012; Shin et al. 2013). PMCA1 defects are also associated with impaired bone mineralization and dietary Ca2+ absorption in the small intestine (Kim et al. 2012; Liu et al. 2013; Ryan et al. 2015; Ehara et al. 2018). In addition, the key role of PMCA1 in the embryo becomes clear from the homozygous Atp2b1 knockout, which caused embryo lethality, whereas heterozygous mutants presented no overt disease phenotype (Okunade et al. 2004).
PMCA2 deficiency mainly affects tissues in which PMCA2 is highly expressed, highlighting specific physiological functions of PMCA2. PMCA2 null mice display balance control and hearing problems (Kozel et al. 1998; Kurnellas et al. 2007). Moreover, spontaneous mutations in the ATP2B2 gene are associated with hearing loss, deafness, or ataxia in humans and mice (Street et al. 1998; Takahashi and Kitamura 1999; Ueno et al. 2002; Ficarella et al. 2007; Spiden et al. 2008; Vicario et al. 2018), which is the result of impaired sensory transduction in hair cells of the inner ear. During lactation, PMCA2 contributes to Ca2+ uptake in the milk, whereas the reduction in PMCA2 expression during physiological mammary gland involution induces apoptosis. Conversely, PMCA2 overexpression in breast cancer cells is coupled to proliferation and resistance to apoptosis (VanHouten et al. 2010; Peters et al. 2016). Genetic studies further linked PMCA2 mutations with autism in line with the abundant expression of PMCA2 in the brain (Hu et al. 2009; Carayol et al. 2011).
PMCA3 mutations were associated with congenital cerebellar ataxia in humans, which may relate to the high PMCA3 expression in cerebellum (Zanni et al. 2012; Calì et al. 2016; Vicario et al. 2017). However, other PMCA3 mutations were identified in adrenal aldosterone-producing adenomas. These PMCA3 mutants present a lower Ca2+-ATPase activity and may trigger Ca2+ influx because of an increased Ca2+ leak (Beuschlein et al. 2013; Dutta et al. 2014; Tauber et al. 2016). Remarkably, the PMCA3 knockout mice did not present any of these phenotypes, but instead showed an interesting long-sleeper phenotype (Tatsuki et al. 2016).
PMCA4 is the major isoform in erythrocytes (Strehler et al. 1990) and several single-nucleotide polymorphisms in PMCA4 were reported to confer resistance to malaria infection (Bedu-Addo et al. 2013; Lessard et al. 2017), which makes PMCA4 an interesting target for antimalaria therapy. Loss of Atp2b4 in mice is not lethal, but induces sterility in males by impaired sperm motility (Okunade et al. 2004; Schuh et al. 2004). In addition, the Atp2b4 knockout mice present vascular smooth muscle dysfunction in support of a specific role of PMCA4 in cardiovascular physiology (Oceandy et al. 2007; Mohamed et al. 2011; Prasad et al. 2014). Isoform PMCA4 is also abundant in nervous tissue (Krizaj et al. 2002; Burette et al. 2003; Marcos et al. 2009), in which Ca2+ dyshomeostasis contributes to aging and neurodegenerative diseases (Berridge 2011; Rivero-Rios et al. 2014). In fact, PMCA4 is inhibited by amyloid-β peptide and tau, two hallmarks in Alzheimer's disease (Berrocal et al. 2009, 2015), is affected by glutamate excitotoxicity (Pottorf et al. 2006), and its mutation may cause a familial spastic paraplegia (Akin et al. 2013; Li et al. 2014; Ho et al. 2015).
CONCLUDING REMARKS
Over the last few decades, a plethora of Ca2+ pump isoforms and splice variants have been discovered, which are mapped differently across cell types and subcellular compartments, in which they fine tune Ca2+ transport according to the local physiological needs. With detailed insights into the Ca2+ transport mechanism of primary active transporters at hand, the field now focuses on understanding the isoform-specific regulation of Ca2+ pumps at the molecular level. Isoform-specific elements alter the molecular dynamics and kinetic behavior, allow posttranslational control, and provide regulation by intramolecular domains or protein interaction. In addition, genetic screenings continue to provide new correlations between primary Ca2+ transport dysfunction and human diseases, which helps to establish Ca2+ pumps as therapeutic targets. Without any doubt, insights into the molecular aspects of Ca2+ pump regulation will provide new therapeutic opportunities to restore aberrant Ca2+ transport in disease.
ACKNOWLEDGMENTS
This work was supported by Flanders Research Foundation FWO Grants G044212N and G0B1115N, and the Inter-University Attraction Poles Program P7/13 assigned to P.V. and PP2016-PJI05 from University of Granada to M.R.S. V.B. is supported by project funding from the Center for Drug Design and Discovery (Belgium) and A.S. is supported by the doctoral scholarship provided by Agentschap Innoveren and Ondernemen (VLAIO).
Footnotes
Editors: Geert Bultynck, Martin D. Bootman, Michael J. Berridge, and Grace E. Stutzmann
Additional Perspectives on Calcium Signaling available at www.cshperspectives.org
REFERENCES
- Akin BL, Hurley TD, Chen Z, Jones LR. 2013. The structural basis for phospholamban inhibition of the calcium pump in sarcoplasmic reticulum. J Biol Chem 288: 30181–30191. 10.1074/jbc.M113.501585 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Albers RW. 1967. Biochemical aspects of active transport. Annu Rev Biochem 36: 727–756. 10.1146/annurev.bi.36.070167.003455 [DOI] [PubMed] [Google Scholar]
- Anderson DM, Anderson KM, Chang CL, Makarewich CA, Nelson BR, McAnally JR, Kasaragod P, Shelton JM, Liou J, Bassel-Duby R, et al. 2015. A micropeptide encoded by a putative long noncoding RNA regulates muscle performance. Cell 160: 595–606. 10.1016/j.cell.2015.01.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Anderson DM, Makarewich CA, Anderson KM, Shelton JM, Bezprozvannaya S, Bassel-Duby R, Olson EN. 2016. Widespread control of calcium signaling by a family of SERCA-inhibiting micropeptides. Sci Signal 9: ra119 10.1126/scisignal.aaj1460 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Andersson KB, Birkeland JA, Finsen AV, Louch WE, Sjaastad I, Wang Y, Chen J, Molkentin JD, Chien KR, Sejersted OM, et al. 2009. Moderate heart dysfunction in mice with inducible cardiomyocyte-specific excision of the Serca2 gene. J Mol Cell Cardiol 47: 180–187. 10.1016/j.yjmcc.2009.03.013 [DOI] [PubMed] [Google Scholar]
- Anger M, Samuel JL, Marotte F, Wuytack F, Rappaport L, Lompré AM. 1994. In situ mRNA distribution of sarco(endo)plasmic reticulum Ca2+-ATPase isoforms during ontogeny in the rat. J Mol Cell Cardiol 26: 539–550. 10.1006/jmcc.1994.1064 [DOI] [PubMed] [Google Scholar]
- Asahi M, Kurzydlowski K, Tada M, MacLennan DH. 2002. Sarcolipin inhibits polymerization of phospholamban to induce superinhibition of sarco(endo)plasmic reticulum Ca2+-ATPases (SERCAs). J Biol Chem 277: 26725–26728. 10.1074/jbc.C200269200 [DOI] [PubMed] [Google Scholar]
- Asahi M, Otsu K, Nakayama H, Hikoso S, Takeda T, Gramolini AO, Trivieri MG, Oudit GY, Morita T, Kusakari Y, et al. 2004. Cardiac-specific overexpression of sarcolipin inhibits sarco(endo)plasmic reticulum Ca2+ ATPase (SERCA2a) activity and impairs cardiac function in mice. Proc Natl Acad Sci 101: 9199–9204. 10.1073/pnas.0402596101 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Azeez JM, Vini R, Remadevi V, Surendran A, Jaleel A, Santhosh Kumar TR, Sreeja S. 2018. VDAC1 and SERCA3 mediate progesterone-triggered Ca2+ signaling in breast cancer cells. J Proteome Res 17: 698–709. 10.1021/acs.jproteome.7b00754 [DOI] [PubMed] [Google Scholar]
- Baba-Aissa F, Raeymaekers L, Wuytack F, De Greef C, Missiaen L, Casteels R. 1996. Distribution of the organellar Ca2+ transport ATPase SERCA2 isoforms in the cat brain. Brain Res 743: 141–153. 10.1016/S0006-8993(96)01037-2 [DOI] [PubMed] [Google Scholar]
- Baba-Aissa F, Raeymaekers L, Wuytack F, Dode L, Casteels R. 1998. Distribution and isoform diversity of the organellar Ca2+ pumps in the brain. Mol Chem Neuropathol 33: 199–208. 10.1007/BF02815182 [DOI] [PubMed] [Google Scholar]
- Babu GJ, Bhupathy P, Carnes CA, Billman GE, Periasamy M. 2007a. Differential expression of sarcolipin protein during muscle development and cardiac pathophysiology. J Mol Cell Cardiol 43: 215–222. 10.1016/j.yjmcc.2007.05.009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Babu GJ, Bhupathy P, Timofeyev V, Petrashevskaya NN, Reiser PJ, Chiamvimonvat N, Periasamy M. 2007b. Ablation of sarcolipin enhances sarcoplasmic reticulum calcium transport and atrial contractility. Proc Natl Acad Sci 104: 17867–17872. 10.1073/pnas.0707722104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bal NC, Maurya SK, Sopariwala DH, Sahoo SK, Gupta SC, Shaikh SA, Pant M, Rowland LA, Bombardier E, Goonasekera SA, et al. 2012. Sarcolipin is a newly identified regulator of muscle-based thermogenesis in mammals. Nat Med 18: 1575–1579. 10.1038/nm.2897 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bedu-Addo G, Meese S, Mockenhaupt FP. 2013. An ATP2B4 polymorphism protects against malaria in pregnancy. J Infect Dis 207: 1600–1603. 10.1093/infdis/jit070 [DOI] [PubMed] [Google Scholar]
- Berridge MJ. 2011. Calcium signalling and Alzheimer's disease. Neurochem Res 36: 1149–1156. 10.1007/s11064-010-0371-4 [DOI] [PubMed] [Google Scholar]
- Berrocal M, Marcos D, Sepúlveda MR, Pérez M, Ávila J, Mata AM. 2009. Altered Ca2+ dependence of synaptosomal plasma membrane Ca2+-ATPase in human brain affected by Alzheimer's disease. FASEB J 23: 1826–1834. 10.1096/fj.08-121459 [DOI] [PubMed] [Google Scholar]
- Berrocal M, Corbacho I, Vázquez-Hernández M, Ávila J, Sepúlveda MR, Mata AM. 2015. Inhibition of PMCA activity by tau as a function of aging and Alzheimer's neuropathology. Biochim Biophys Acta 1852: 1465–1476. 10.1016/j.bbadis.2015.04.007 [DOI] [PubMed] [Google Scholar]
- Bers DM. 2002. Cardiac excitation-contraction coupling. Nature 415: 198–205. 10.1038/415198a [DOI] [PubMed] [Google Scholar]
- Beuschlein F, Boulkroun S, Osswald A, Wieland T, Nielsen HN, Lichtenauer UD, Penton D, Schack VR, Amar L, Fischer E, et al. 2013. Somatic mutations in ATP1A1 and ATP2B3 lead to aldosterone-producing adenomas and secondary hypertension. Nat Genet 45: 440–444. 10.1038/ng.2550 [DOI] [PubMed] [Google Scholar]
- Bidwell PA, Haghighi K, Kranias EG. 2018. The antiapoptotic protein HAX-1 mediates half of phospholamban's inhibitory activity on calcium cycling and contractility in the heart. J Biol Chem 293: 359–367. 10.1074/jbc.RA117.000128 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bobe R, Bredoux R, Corvazier E, Andersen JP, Clausen JD, Dode L, Kovács T, Enouf J. 2004. Identification, expression, function, and localization of a novel (sixth) isoform of the human sarco/endoplasmic reticulum Ca2+ ATPase 3 gene. J Biol Chem 279: 24297–24306. 10.1074/jbc.M314286200 [DOI] [PubMed] [Google Scholar]
- Brandl CJ, deLeon S, Martin DR, MacLennan DH. 1987. Adult forms of the Ca2+ ATPase of sarcoplasmic reticulum. Expression in developing skeletal muscle. J Biol Chem 262: 3768–3774. [PubMed] [Google Scholar]
- Brandt P, Neve RL. 1992. Expression of plasma membrane calcium-pumping ATPase mRNAs in developing rat brain and adult brain subregions: evidence for stage-specific expression. J Neurochem 59: 1566–1569. 10.1111/j.1471-4159.1992.tb08476.x [DOI] [PubMed] [Google Scholar]
- Brini M, Carafoli E. 2009. Calcium pumps in health and disease. Physiol Rev 89: 1341–1378. 10.1152/physrev.00032.2008 [DOI] [PubMed] [Google Scholar]
- Brini M, Carafoli E. 2011. The plasma membrane Ca2+ ATPase and the plasma membrane sodium calcium exchanger cooperate in the regulation of cell calcium. Cold Spring Harb Perspect Biol 3: a004168 10.1101/cshperspect.a004168 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brini M, Di Leva F, Ortega CK, Domi T, Ottolini D, Leonardi E, Tosatto SC, Carafoli E. 2010. Deletions and mutations in the acidic lipid-binding region of the plasma membrane Ca2+ pump: A study on different splicing variants of isoform 2. J Biol Chem 285: 30779–30791. 10.1074/jbc.M110.140475 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brown JM, García-García MJ. 2018. Secretory pathway calcium ATPase 1 (SPCA1) controls mouse neural tube closure by regulating cytoskeletal dynamics. Development 145: dev170019 10.1242/dev.170019 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bublitz M, Morth JP, Nissen P. 2011. P-type ATPases at a glance. J Cell Sci 124: 2515–2519. 10.1242/jcs.088716 [DOI] [PubMed] [Google Scholar]
- Bublitz M, Musgaard M, Poulsen H, Thøgersen L, Olesen C, Schiøtt B, Morth JP, Møller JV, Nissen P. 2013. Ion pathways in the sarcoplasmic reticulum Ca2+-ATPase. J Biol Chem 288: 10759–10765. 10.1074/jbc.R112.436550 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bultynck G, Kiviluoto S, Methner A. 2014. Bax inhibitor-1 is likely a pH-sensitive calcium leak channel, not a H+/Ca2+ exchanger. Sci Signal 7: pe22 10.1126/scisignal.2005764 [DOI] [PubMed] [Google Scholar]
- Burette A, Rockwood JM, Strehler EE, Weinberg RJ. 2003. Isoform-specific distribution of the plasma membrane Ca2+ ATPase in the rat brain. J Comp Neurol 467: 464–476. 10.1002/cne.10933 [DOI] [PubMed] [Google Scholar]
- Büttner S, Faes L, Reichelt WN, Broeskamp F, Habernig L, Benke S, Kourtis N, Ruli D, Carmona-Gutierrez D, Eisenberg T, et al. 2013. The Ca2+/Mn2+ ion-pump PMR1 links elevation of cytosolic Ca2+ levels to α-synuclein toxicity in Parkinson's disease models. Cell Death Differ 20: 465–477. 10.1038/cdd.2012.142 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Calì T, Frizzarin M, Luoni L, Zonta F, Pantano S, Cruz C, Bonza MC, Bertipaglia I, Ruzzene M, De Michelis MI, et al. 2016. The ataxia related G1107D mutation of the plasma membrane Ca2+ ATPase isoform 3 affects its interplay with calmodulin and the autoinhibition process. Biochim Biophys Acta 1863: 165–173. 10.1016/j.bbadis.2016.09.007 [DOI] [PubMed] [Google Scholar]
- Carayol J, Sacco R, Tores F, Rousseau F, Lewin P, Hager J, Persico AM. 2011. Converging evidence for an association of ATP2B2 allelic variants with autism in male subjects. Biol Psychiatry 70: 880–887. 10.1016/j.biopsych.2011.05.020 [DOI] [PubMed] [Google Scholar]
- Caspersen C, Pedersen PS, Treiman M. 2000. The sarco/endoplasmic reticulum calcium-ATPase 2b is an endoplasmic reticulum stress-inducible protein. J Biol Chem 275: 22363–22372. 10.1074/jbc.M001569200 [DOI] [PubMed] [Google Scholar]
- Chami M, Gozuacik D, Lagorce D, Brini M, Falson P, Peaucellier G, Pinton P, Lecoeur H, Gougeon ML, le Maire M, et al. 2001. SERCA1 truncated proteins unable to pump calcium reduce the endoplasmic reticulum calcium concentration and induce apoptosis. J Cell Biol 153: 1301–1314. 10.1083/jcb.153.6.1301 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chami M, Oulès B, Szabadkai G, Tacine R, Rizzuto R, Paterlini-Bréchot P. 2008. Role of SERCA1 truncated isoform in the proapoptotic calcium transfer from ER to mitochondria during ER stress. Mol Cell 32: 641–651. 10.1016/j.molcel.2008.11.014 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chandrasekera PC, Kargacin ME, Deans JP, Lytton J. 2009. Determination of apparent calcium affinity for endogenously expressed human sarco(endo)plasmic reticulum calcium-ATPase isoform SERCA3. Am J Physiol Cell Physiol 296: C1105–C1114. 10.1152/ajpcell.00650.2008 [DOI] [PubMed] [Google Scholar]
- Charlier C, Coppieters W, Rollin F, Desmecht D, Agerholm JS, Cambisano N, Carta E, Dardano S, Dive M, Fasquelle C, et al. 2008. Highly effective SNP-based association mapping and management of recessive defects in livestock. Nat Genet 40: 449–454. 10.1038/ng.96 [DOI] [PubMed] [Google Scholar]
- Chemaly ER, Troncone L, Lebeche D. 2018. SERCA control of cell death and survival. Cell Calcium 69: 46–61. 10.1016/j.ceca.2017.07.001 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen JL, Smaardijk S, Vandecaetsbeek I, Vangheluwe P. 2016. Regulation of Ca2+ transport ATPases by amino- and carboxy-terminal extensions: Mechanisms and (patho)physiological implications. Adv Biochem Health D 14: 243–279. 10.1007/978-3-319-24780-9_14 [DOI] [Google Scholar]
- Chen J, De Raeymaecker J, Hovgaard JB, Smaardijk S, Vandecaetsbeek I, Wuytack F, Møller JV, Eggermont J, De Maeyer M, Christensen SB, et al. 2017. Structure/activity relationship of thapsigargin inhibition on the purified Golgi/secretory pathway Ca2+/Mn2+-transport ATPase (SPCA1a). J Biol Chem 292: 6938–6951. 10.1074/jbc.M117.778431 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen J, Smaardijk S, Mattelaer CA, Pamula F, Vandecaetsbeek I, Vanoevelen J, Wuytack F, Lescrinier E, Eggermont J, Vangheluwe P. 2019. An N-terminal Ca2+ binding motif regulates the secretory pathway Ca2+/Mn2+-transport ATPase SPCA1. J Biol Chem 294: 7878–7891. 10.1074/jbc.RA118.006250 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho YS, Go MJ, Kim YJ, Heo JY, Oh JH, Ban HJ, Yoon D, Lee MH, Kim DJ, Park M, et al. 2009. A large-scale genome-wide association study of Asian populations uncovers genetic factors influencing eight quantitative traits. Nat Genet 41: 527–534. 10.1038/ng.357 [DOI] [PubMed] [Google Scholar]
- Christopoulos PF, Corthay A, Koutsilieris M. 2018. Aiming for the insulin-like growth factor-1 system in breast cancer therapeutics. Cancer Treat Rev 63: 79–95. 10.1016/j.ctrv.2017.11.010 [DOI] [PubMed] [Google Scholar]
- Clausen JD, Vandecaetsbeek I, Wuytack F, Vangheluwe P, Andersen JP. 2012. Distinct roles of the C-terminal 11th transmembrane helix and luminal extension in the partial reactions determining the high Ca2+ affinity of sarco(endo)plasmic reticulum Ca2+-ATPase isoform 2b (SERCA2b). J Biol Chem 287: 39460–39469. 10.1074/jbc.M112.397331 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Clausen JD, Bublitz M, Arnou B, Olesen C, Andersen JP, Møller JV, Nissen P. 2016. Crystal structure of the vanadate-inhibited Ca2+-ATPase. Structure 24: 617–623. 10.1016/j.str.2016.02.018 [DOI] [PubMed] [Google Scholar]
- Crevenna AH, Blank B, Maiser A, Emin D, Prescher J, Beck G, Kienzle C, Bartnik K, Habermann B, Pakdel M, et al. 2016. Secretory cargo sorting by Ca2+-dependent Cab45 oligomerization at the trans-Golgi network. J Cell Biol 213: 305 10.1083/jcb.201601089 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cross BM, Hack A, Reinhardt TA, Rao R. 2013. SPCA2 regulates Orai1 trafficking and store independent Ca2+ entry in a model of lactation. PLoS ONE 8: e67348 10.1371/journal.pone.0067348 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cross BM, Breitwieser GE, Reinhardt TA, Rao R. 2014. Cellular calcium dynamics in lactation and breast cancer: From physiology to pathology. Am J Physiol Cell Physiol 306: C515–C526. 10.1152/ajpcell.00330.2013 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dally S, Bredoux R, Corvazier E, Andersen JP, Clausen JD, Dode L, Fanchaouy M, Gelebart P, Monceau V, Del Monte F, et al. 2006. Ca2+-ATPases in non-failing and failing heart: Evidence for a novel cardiac sarco/endoplasmic reticulum Ca2+-ATPase 2 isoform (SERCA2c). Biochem J 395: 249–258. 10.1042/BJ20051427 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dally S, Monceau V, Corvazier E, Bredoux R, Raies A, Bobe R, del Monte F, Enouf J. 2009. Compartmentalized expression of three novel sarco/endoplasmic reticulum Ca2+ ATPase 3 isoforms including the switch to ER stress, SERCA3f, in non-failing and failing human heart. Cell Calcium 45: 144–154. 10.1016/j.ceca.2008.08.002 [DOI] [PubMed] [Google Scholar]
- Dally S, Corvazier E, Bredoux R, Bobe R, Enouf J. 2010. Multiple and diverse coexpression, location, and regulation of additional SERCA2 and SERCA3 isoforms in nonfailing and failing human heart. J Mol Cell Cardiol 48: 633–644. 10.1016/j.yjmcc.2009.11.012 [DOI] [PubMed] [Google Scholar]
- Dang D, Prasad H, Rao R. 2017. Secretory pathway Ca2+-ATPases promote in vitro microcalcifications in breast cancer cells. Mol Carcinog 56: 2474–2485. 10.1002/mc.22695 [DOI] [PMC free article] [PubMed] [Google Scholar]
- DeMarco SJ, Strehler EE. 2001. Plasma membrane Ca2+- atpase isoforms 2b and 4b interact promiscuously and selectively with members of the membrane-associated guanylate kinase family of PDZ (PSD95/Dlg/ZO-1) domain-containing proteins. J Biol Chem 276: 21594–21600. 10.1074/jbc.M101448200 [DOI] [PubMed] [Google Scholar]
- Denmeade SR, Mhaka AM, Rosen DM, Brennen WN, Dalrymple S, Dach I, Olesen C, Gurel B, Demarzo AM, Wilding G, et al. 2012. Engineering a prostate-specific membrane antigen-activated tumor endothelial cell prodrug for cancer therapy. Sci Transl Med 4: 140ra186 10.1126/scitranslmed.3003886 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dhitavat J, Dode L, Leslie N, Sakuntabhai A, Lorette G, Hovnanian A. 2003. Mutations in the sarcoplasmic/endoplasmic reticulum Ca2+ ATPase isoform cause Darier's disease. J Invest Dermatol 121: 486–489. 10.1046/j.1523-1747.2003.12410.x [DOI] [PubMed] [Google Scholar]
- Dode L, Vilsen B, Van Baelen K, Wuytack F, Clausen JD, Andersen JP. 2002. Dissection of the functional differences between sarco(endo)plasmic reticulum Ca2+- ATPase (SERCA) 1 and 3 isoforms by steady-state and transient kinetic analyses. J Biol Chem 277: 45579–45591. 10.1074/jbc.M207778200 [DOI] [PubMed] [Google Scholar]
- Dode L, Andersen JP, Leslie N, Dhitavat J, Vilsen B, Hovnanian A. 2003. Dissection of the functional differences between sarco(endo)plasmic reticulum Ca2+-ATPase (SERCA) 1 and 2 isoforms and characterization of Darier disease (SERCA2) mutants by steady-state and transient kinetic analyses. J Biol Chem 278: 47877–47889. 10.1074/jbc.M306784200 [DOI] [PubMed] [Google Scholar]
- Dode L, Andersen JP, Raeymaekers L, Missiaen L, Vilsen B, Wuytack F. 2005. Functional comparison between secretory pathway Ca2+/Mn2+-ATPase (SPCA) 1 and sarcoplasmic reticulum Ca2+-ATPase (SERCA) 1 isoforms by steady-state and transient kinetic analyses. J Biol Chem 280: 39124–39134. 10.1074/jbc.M506181200 [DOI] [PubMed] [Google Scholar]
- Dode L, Andersen JP, Vanoevelen J, Raeymaekers L, Missiaen L, Vilsen B, Wuytack F. 2006. Dissection of the functional differences between human secretory pathway Ca2+/Mn2+-ATPase (SPCA) 1 and 2 isoenzymes by steady-state and transient kinetic analyses. J Biol Chem 281: 3182–3189. 10.1074/jbc.M511547200 [DOI] [PubMed] [Google Scholar]
- Dremina ES, Sharov VS, Kumar K, Zaidi A, Michaelis EK, Schöneich C. 2004. Anti-apoptotic protein Bcl-2 interacts with and destabilizes the sarcoplasmic/endoplasmic reticulum Ca2+-ATPase (SERCA). Biochem J 383: 361–370. 10.1042/BJ20040187 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Drögemüller C, Drögemüller M, Leeb T, Mascarello F, Testoni S, Rossi M, Gentile A, Damiani E, Sacchetto R. 2008. Identification of a missense mutation in the bovine ATP2A1 gene in congenital pseudomyotonia of Chianina cattle: An animal model of human Brody disease. Genomics 92: 474–477. 10.1016/j.ygeno.2008.07.014 [DOI] [PubMed] [Google Scholar]
- Dutta RK, Welander J, Brauckhoff M, Walz M, Alesina P, Arnesen T, Söderkvist P, Gimm O. 2014. Complementary somatic mutations of KCNJ5, ATP1A1, and ATP2B3 in sporadic aldosterone producing adrenal adenomas. Endocr Relat Cancer 21: L1–L4. 10.1530/ERC-13-0466 [DOI] [PubMed] [Google Scholar]
- Dyla M, Terry DS, Kjaergaard M, Sørensen TL, Lauwring Andersen J, Andersen JP, Rohde Knudsen C, Altman RB, Nissen P, Blanchard SC. 2017. Dynamics of P-type ATPase transport revealed by single-molecule FRET. Nature 551: 346–351. 10.1038/nature24296 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eakin TJ, Antonelli MC, Malchiodi EL, Baskin DG, Stahl WL. 1995. Localization of the plasma membrane Ca2+-ATPase isoform PMCA3 in rat cerebellum, choroid plexus and hippocampus. Brain Res Mol Brain Res 29: 71–80. 10.1016/0169-328X(94)00231-3 [DOI] [PubMed] [Google Scholar]
- Ehara Y, Hirawa N, Sumida K, Fujiwara A, Kagimoto M, Ooki-Okuyama Y, Fujita M, Katsumata M, Kobayashi Y, Saka S, et al. 2018. Reduced secretion of parathyroid hormone and hypocalcemia in systemic heterozygous ATP2B1-null hypertensive mice. Hypertens Res 41: 699–707. 10.1038/s41440-018-0067-8 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elaib Z, Adam F, Berrou E, Bordet JC, Prevost N, Bobe R, Bryckaert M, Rosa JP. 2016. Full activation of mouse platelets requires ADP secretion regulated by SERCA3 ATPase-dependent calcium stores. Blood 128: 1129–1138. 10.1182/blood-2015-10-678383 [DOI] [PubMed] [Google Scholar]
- Elwess NL, Filoteo AG, Enyedi A, Penniston JT. 1997. Plasma membrane Ca2+ pump isoforms 2a and 2b are unusually responsive to calmodulin and Ca2+. J Biol Chem 272: 17981–17986. 10.1074/jbc.272.29.17981 [DOI] [PubMed] [Google Scholar]
- Engin B, Kutlubay Z, Erkan E, Tüzün Y. 2015. Darier disease: A fold (intertriginous) dermatosis. Clin Dermatol 33: 448–451. 10.1016/j.clindermatol.2015.04.009 [DOI] [PubMed] [Google Scholar]
- Faddy HM, Smart CE, Xu R, Lee GY, Kenny PA, Feng M, Rao R, Brown MA, Bissell MJ, Roberts-Thomson SJ, et al. 2008. Localization of plasma membrane and secretory calcium pumps in the mammary gland. Biochem Biophys Res Commun 369: 977–981. 10.1016/j.bbrc.2008.03.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fairclough RJ, Dode L, Vanoevelen J, Andersen JP, Missiaen L, Raeymaekers L, Wuytack F, Hovnanian A. 2003. Effect of Hailey–Hailey disease mutations on the function of a new variant of human secretory pathway Ca2+/Mn2+-ATPase (hSPCA1). J Biol Chem 278: 24721–24730. 10.1074/jbc.M300509200 [DOI] [PubMed] [Google Scholar]
- Farooqui AA, Rapoport SI, Horrocks LA. 1997. Membrane phospholipid alterations in Alzheimer's disease: Deficiency of ethanolamine plasmalogens. Neurochem Res 22: 523–527. 10.1023/A:1027380331807 [DOI] [PubMed] [Google Scholar]
- Feng M, Grice DM, Faddy HM, Nguyen N, Leitch S, Wang Y, Muend S, Kenny PA, Sukumar S, Roberts-Thomson SJ, et al. 2010. Store-independent activation of Orai1 by SPCA2 in mammary tumors. Cell 143: 84–98. 10.1016/j.cell.2010.08.040 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ferdek PE, Gerasimenko JV, Peng S, Tepikin AV, Petersen OH, Gerasimenko OV. 2012. A novel role for Bcl-2 in regulation of cellular calcium extrusion. Curr Biol 22: 1241–1246. 10.1016/j.cub.2012.05.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ficarella R, Di Leva F, Bortolozzi M, Ortolano S, Donaudy F, Petrillo M, Melchionda S, Lelli A, Domi T, Fedrizzi L, et al. 2007. A functional study of plasma-membrane calcium-pump isoform 2 mutants causing digenic deafness. Proc Natl Acad Sci 104: 1516–1521. 10.1073/pnas.0609775104 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Furuta H, Luo L, Hepler K, Ryan AF. 1998. Evidence for differential regulation of calcium by outer versus inner hair cells: Plasma membrane Ca-ATPase gene expression. Hear Res 123: 10–26. 10.1016/S0378-5955(98)00091-4 [DOI] [PubMed] [Google Scholar]
- Gélébart P, Kovács T, Brouland JP, van Gorp R, Grossmann J, Rivard N, Panis Y, Martin V, Bredoux R, Enouf J, et al. 2002. Expression of endomembrane calcium pumps in colon and gastric cancer cells. Induction of SERCA3 expression during differentiation. J Biol Chem 277: 26310–26320. 10.1074/jbc.M201747200 [DOI] [PubMed] [Google Scholar]
- Gélébart P, Martin V, Enouf J, Papp B. 2003. Identification of a new SERCA2 splice variant regulated during monocytic differentiation. Biochem Biophys Res Commun 303: 676–684. 10.1016/S0006-291X(03)00405-4 [DOI] [PubMed] [Google Scholar]
- Ghosh B, Li Y, Thayer SA. 2011. Inhibition of the plasma membrane Ca2+ pump by CD44 receptor activation of tyrosine kinases increases the action potential afterhyperpolarization in sensory neurons. J Neurosci 31: 2361–2370. 10.1523/jneurosci.5764-10.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ghosh B, Green MV, Krogh KA, Thayer SA. 2016. Interleukin-1β activates an Src family kinase to stimulate the plasma membrane Ca2+ pump in hippocampal neurons. J Neurophysiol 115: 1875–1885. 10.1152/jn.00541.2015 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Giorgi C, Bonora M, Sorrentino G, Missiroli S, Poletti F, Suski JM, Ramirez FG, Rizzuto R, Di Virgilio F, Zito E, et al. 2015. p53 at the endoplasmic reticulum regulates apoptosis in a Ca2+-dependent manner. Proc Natl Acad Sci 112: 1779–1784. 10.1073/pnas.1410723112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Go CK, Soboloff J. 2018. Hold the door: hPMCA1/neuroplastin interactions regulate Ca2+-binding site accessibility. Cell Calcium 76: 135–136. 10.1016/j.ceca.2018.10.003 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gonatas NK, Stieber A, Gonatas JO. 2006. Fragmentation of the Golgi apparatus in neurodegenerative diseases and cell death. J Neurol Sci 246: 21–30. 10.1016/j.jns.2006.01.019 [DOI] [PubMed] [Google Scholar]
- Gong D, Chi X, Ren K, Huang G, Zhou G, Yan N, Lei J, Zhou Q. 2018. Structure of the human plasma membrane Ca2+-ATPase 1 in complex with its obligatory subunit neuroplastin. Nat Commun 9: 3623 10.1038/s41467-018-06075-7 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goonasekera SA, Lam CK, Millay DP, Sargent MA, Hajjar RJ, Kranias EG, Molkentin JD. 2011. Mitigation of muscular dystrophy in mice by SERCA overexpression in skeletal muscle. J Clin Invest 121: 1044–1052. 10.1172/JCI43844 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorski PA, Trieber CA, Larivière E, Schuermans M, Wuytack F, Young HS, Vangheluwe P. 2012. Transmembrane helix 11 is a genuine regulator of the endoplasmic reticulum Ca2+ pump and acts as a functional parallel of β-subunit on α-Na+,K+-ATPase. J Biol Chem 287: 19876–19885. 10.1074/jbc.M111.335620 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorski PA, Glaves JP, Vangheluwe P, Young HS. 2013. Sarco(endo)plasmic reticulum calcium ATPase (SERCA) inhibition by sarcolipin is encoded in its luminal tail. J Biol Chem 288: 8456–8467. 10.1074/jbc.M112.446161 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorski PA, Ceholski DK, Hajjar RJ. 2015. Altered myocardial calcium cycling and energetics in heart failure—A rational approach for disease treatment. Cell Metab 21: 183–194. 10.1016/j.cmet.2015.01.005 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gorski PA, Ceholski DK, Young HS. 2017. Structure–function relationship of the SERCA pump and its regulation by phospholamban and sarcolipin. Adv Exp Med Biol 981: 77–119. 10.1007/978-3-319-55858-5_5 [DOI] [PubMed] [Google Scholar]
- Greenberg B, Butler J, Felker GM, Ponikowski P, Voors AA, Desai AS, Barnard D, Bouchard A, Jaski B, Lyon AR, et al. 2016. Calcium upregulation by percutaneous administration of gene therapy in patients with cardiac disease (CUPID 2): A randomised, multinational, double-blind, placebo-controlled, phase 2b trial. Lancet 387: 1178–1186. 10.1016/S0140-6736(16)00082-9 [DOI] [PubMed] [Google Scholar]
- Grice DM, Vetter I, Faddy HM, Kenny PA, Roberts-Thomson SJ, Monteith GR. 2010. Golgi calcium pump secretory pathway calcium ATPase 1 (SPCA1) is a key regulator of insulin-like growth factor receptor (IGF1R) processing in the basal-like breast cancer cell line MDA-MB-231. J Biol Chem 285: 37458–37466. 10.1074/jbc.M110.163329 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guerini D, Pan B, Carafoli E. 2003. Expression, purification, and characterization of isoform 1 of the plasma membrane Ca2+ pump: Focus on calpain sensitivity. J Biol Chem 278: 38141–38148. 10.1074/jbc.M302400200 [DOI] [PubMed] [Google Scholar]
- Guglielmi V, Vattemi G, Gualandi F, Voermans NC, Marini M, Scotton C, Pegoraro E, Oosterhof A, Kósa M, Zádor E, et al. 2013. SERCA1 protein expression in muscle of patients with Brody disease and Brody syndrome and in cultured human muscle fibers. Mol Genet Metab 110: 162–169. 10.1016/j.ymgme.2013.07.015 [DOI] [PubMed] [Google Scholar]
- Gunteski-Hamblin AM, Clarke DM, Shull GE. 1992. Molecular cloning and tissue distribution of alternatively spliced mRNAs encoding possible mammalian homologues of the yeast secretory pathway calcium pump. Biochemistry (Mosc) 31: 7600–7608. 10.1021/bi00148a023 [DOI] [PubMed] [Google Scholar]
- Gutiérrez T, Simmen T. 2018. Endoplasmic reticulum chaperones tweak the mitochondrial calcium rheostat to control metabolism and cell death. Cell Calcium 70: 64–75. 10.1016/j.ceca.2017.05.015 [DOI] [PubMed] [Google Scholar]
- Hao L, Rigaud JL, Inesi G. 1994. Ca2+/H+ countertransport and electrogenicity in proteoliposomes containing erythrocyte plasma membrane Ca-ATPase and exogenous lipids. J Biol Chem 269: 14268–14275. [PubMed] [Google Scholar]
- Hasselbach W. 1964. Atp-driven active transport of calcium in the membranes of the sarcoplasmic reticulum. Proc R Soc Lond B Biol Sci 160: 501–504. 10.1098/rspb.1964.0064 [DOI] [PubMed] [Google Scholar]
- Helenius A, Aebi M. 2001. Intracellular functions of N-linked glycans. Science 291: 2364–2369. 10.1126/science.291.5512.2364 [DOI] [PubMed] [Google Scholar]
- Hirata H, Saint-Amant L, Waterbury J, Cui W, Zhou W, Li Q, Goldman D, Granato M, Kuwada JY. 2004. Accordion, a zebrafish behavioral mutant, has a muscle relaxation defect due to a mutation in the ATPase Ca2+ pump SERCA1. Development 131: 5457–5468. 10.1242/dev.01410 [DOI] [PubMed] [Google Scholar]
- Ho PW, Pang SY, Li M, Tse ZH, Kung MH, Sham PC, Ho SL. 2015. PMCA4 (ATP2B4) mutation in familial spastic paraplegia causes delay in intracellular calcium extrusion. Brain Behav 5: e00321 10.1002/brb3.321 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hoffmann HH, Schneider WM, Blomen VA, Scull MA, Hovnanian A, Brummelkamp TR, Rice CM. 2017. Diverse viruses require the calcium transporter SPCA1 for maturation and spread. Cell Host Microbe 22: 460–470.e465. 10.1016/j.chom.2017.09.002 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hong KW, Go MJ, Jin HS, Lim JE, Lee JY, Han BG, Hwang SY, Lee SH, Park HK, Cho YS, et al. 2010. Genetic variations in ATP2B1, CSK, ARSG and CSMD1 loci are related to blood pressure and/or hypertension in two Korean cohorts. J Hum Hypertens 24: 367–372. 10.1038/jhh.2009.86 [DOI] [PubMed] [Google Scholar]
- Hu Z, Bonifas JM, Beech J, Bench G, Shigihara T, Ogawa H, Ikeda S, Mauro T, Epstein EH Jr. 2000. Mutations in ATP2C1, encoding a calcium pump, cause Hailey–Hailey disease. Nat Genet 24: 61–65. 10.1038/71701 [DOI] [PubMed] [Google Scholar]
- Hu VW, Nguyen A, Kim KS, Steinberg ME, Sarachana T, Scully MA, Soldin SJ, Luu T, Lee NH. 2009. Gene expression profiling of lymphoblasts from autistic and nonaffected sib pairs: Altered pathways in neuronal development and steroid biosynthesis. PLoS ONE 4: e5775 10.1371/journal.pone.0005775 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Inoue M, Sakuta N, Watanabe S, Zhang Y, Yoshikaie K, Tanaka Y, Ushioda R, Kato Y, Takagi J, Tsukazaki T, et al. 2019. Structural basis of sarco/endoplasmic reticulum Ca2+-ATPase 2b regulation via transmembrane helix interplay. Cell Rep 27: 1221–1230 e1223. 10.1016/j.celrep.2019.03.106 [DOI] [PubMed] [Google Scholar]
- Jensen AM, Sørensen TL, Olesen C, Møller JV, Nissen P. 2006. Modulatory and catalytic modes of ATP binding by the calcium pump. EMBO J 25: 2305–2314. 10.1038/sj.emboj.7601135 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang L, Fernandes D, Mehta N, Bean JL, Michaelis ML, Zaidi A. 2007. Partitioning of the plasma membrane Ca2+- ATPase into lipid rafts in primary neurons: Effects of cholesterol depletion. J Neurochem 102: 378–388. 10.1111/j.1471-4159.2007.04480.x [DOI] [PubMed] [Google Scholar]
- Jiang L, Bechtel MD, Galeva NA, Williams TD, Michaelis EK, Michaelis ML. 2012. Decreases in plasma membrane Ca2+-ATPase in brain synaptic membrane rafts from aged rats. J Neurochem 123: 689–699. 10.1111/j.1471-4159.2012.07918.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaneko M, Desai BS, Cook B. 2014. Ionic leakage underlies a gain-of-function effect of dominant disease mutations affecting diverse P-type ATPases. Nat Genet 46: 144–151. 10.1038/ng.2850 [DOI] [PubMed] [Google Scholar]
- Kang S, Dahl R, Hsieh W, Shin A, Zsebo KM, Buettner C, Hajjar RJ, Lebeche D. 2016. Small molecular allosteric activator of the sarco/endoplasmic reticulum Ca2+- ATPase (SERCA) attenuates diabetes and metabolic disorders. J Biol Chem 291: 5185–5198. 10.1074/jbc.M115.705012 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kienzle C, Basnet N, Crevenna AH, Beck G, Habermann B, Mizuno N, von Blume J. 2014. Cofilin recruits F-actin to SPCA1 and promotes Ca2+-mediated secretory cargo sorting. J Cell Biol 206: 635–654. 10.1083/jcb.201311052 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim E, DeMarco SJ, Marfatia SM, Chishti AH, Sheng M, Strehler EE. 1998. Plasma membrane Ca2+ ATPase isoform 4b binds to membrane-associated guanylate kinase (MAGUK) proteins via their PDZ (PSD-95/Dlg/ZO-1) domains. J Biol Chem 273: 1591–1595. 10.1074/jbc.273.3.1591 [DOI] [PubMed] [Google Scholar]
- Kim HJ, Prasad V, Hyung SW, Lee ZH, Lee SW, Bhargava A, Pearce D, Lee Y, Kim HH. 2012. Plasma membrane calcium ATPase regulates bone mass by fine-tuning osteoclast differentiation and survival. J Cell Biol 199: 1145–1158. 10.1083/jcb.201204067 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kimura Y, Kurzydlowski K, Tada M, MacLennan DH. 1997. Phospholamban inhibitory function is activated by depolymerization. J Biol Chem 272: 15061–15064. 10.1074/jbc.272.24.15061 [DOI] [PubMed] [Google Scholar]
- Kobayashi Y, Hirawa N, Tabara Y, Muraoka H, Fujita M, Miyazaki N, Fujiwara A, Ichikawa Y, Yamamoto Y, Ichihara N, et al. 2012. Mice lacking hypertension candidate gene ATP2B1 in vascular smooth muscle cells show significant blood pressure elevation. Hypertension 59: 854–860. 10.1161/hypertensionaha.110.165068 [DOI] [PubMed] [Google Scholar]
- Korthals M, Langnaese K, Smalla KH, Kähne T, Herrera-Molina R, Handschuh J, Lehmann AC, Mamula D, Naumann M, Seidenbecher C, et al. 2017. A complex of neuroplastin and plasma membrane Ca2+ ATPase controls T cell activation. Sci Rep 7: 8358 10.1038/s41598-017-08519-4 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kosk-Kosicka D, Bzdega T. 1988. Activation of the erythrocyte Ca2+-ATPase by either self-association or interaction with calmodulin. J Biol Chem 263: 18184–18189. 10.1021/bi00467a025 [DOI] [PubMed] [Google Scholar]
- Kozel PJ, Friedman RA, Erway LC, Yamoah EN, Liu LH, Riddle T, Duffy JJ, Doetschman T, Miller ML, Cardell EL, et al. 1998. Balance and hearing deficits in mice with a null mutation in the gene encoding plasma membrane Ca2+-ATPase isoform 2. J Biol Chem 273: 18693–18696. 10.1074/jbc.273.30.18693 [DOI] [PubMed] [Google Scholar]
- Kranias EG, Hajjar RJ. 2012. Modulation of cardiac contractility by the phospholamban/SERCA2a regulatome. Circ Res 110: 1646–1660. 10.1161/circresaha.111.259754 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krebs J. 2015. The plethora of PMCA isoforms: Alternative splicing and differential expression. Biochim Biophys Acta 1853: 2018–2024. 10.1016/j.bbamcr.2014.12.020 [DOI] [PubMed] [Google Scholar]
- Krizaj D, Demarco SJ, Johnson J, Strehler EE, Copenhagen DR. 2002. Cell-specific expression of plasma membrane calcium ATPase isoforms in retinal neurons. J Comp Neurol 451: 1–21. 10.1002/cne.10281 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krols M, Bultynck G, Janssens S. 2016. ER-mitochondria contact sites: A new regulator of cellular calcium flux comes into play. J Cell Biol 214: 367–370. 10.1083/jcb.201607124 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kurnellas MP, Lee AK, Li H, Deng L, Ehrlich DJ, Elkabes S. 2007. Molecular alterations in the cerebellum of the plasma membrane calcium ATPase 2 (PMCA2)-null mouse indicate abnormalities in Purkinje neurons. Mol Cell Neurosci 34: 178–188. 10.1016/j.mcn.2006.10.010 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Launay S, Gianni M, Kovacs T, Bredoux R, Bruel A, Gelebart P, Zassadowski F, Chomienne C, Enouf J, Papp B. 1999. Lineage-specific modulation of calcium pump expression during myeloid differentiation. Blood 93: 4395–4405. [PubMed] [Google Scholar]
- Le Gall S, Neuhof A, Rapoport T. 2004. The endoplasmic reticulum membrane is permeable to small molecules. Mol Biol Cell 15: 447–455. 10.1091/mbc.e03-05-0325 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Leitch S, Feng M, Muend S, Braiterman LT, Hubbard AL, Rao R. 2011. Vesicular distribution of secretory pathway Ca2+-ATPase isoform 1 and a role in manganese detoxification in liver-derived polarized cells. Biometals 24: 159–170. 10.1007/s10534-010-9384-3 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lessard S, Gatof ES, Beaudoin M, Schupp PG, Sher F, Ali A, Prehar S, Kurita R, Nakamura Y, Baena E, et al. 2017. An erythroid-specific ATP2B4 enhancer mediates red blood cell hydration and malaria susceptibility. J Clin Invest 127: 3065–3074. 10.1172/JCI94378 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li M, Ho PW, Pang SY, Tse ZH, Kung MH, Sham PC, Ho SL. 2014. PMCA4 (ATP2B4) mutation in familial spastic paraplegia. PLoS ONE 9: e104790 10.1371/journal.pone.0104790 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lipskaia L, Chemaly ER, Hadri L, Lompre AM, Hajjar RJ. 2010. Sarcoplasmic reticulum Ca2+ ATPase as a therapeutic target for heart failure. Expert Opin Biol Ther 10: 29–41. 10.1517/14712590903321462 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu LH, Boivin GP, Prasad V, Periasamy M, Shull GE. 2001. Squamous cell tumors in mice heterozygous for a null allele of Atp2a2, encoding the sarco(endo)plasmic reticulum Ca2+-ATPase isoform 2 Ca2+ pump. J Biol Chem 276: 26737–26740. 10.1074/jbc.C100275200 [DOI] [PubMed] [Google Scholar]
- Liu C, Weng H, Chen L, Yang S, Wang H, Debnath G, Guo X, Wu L, Mohandas N, An X. 2013. Impaired intestinal calcium absorption in protein 4.1R-deficient mice due to altered expression of plasma membrane calcium ATPase 1b (PMCA1b). J Biol Chem 288: 11407–11415. 10.1074/jbc.M112.436659 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lopreiato R, Giacomello M, Carafoli E. 2014. The plasma membrane calcium pump: New ways to look at an old enzyme. J Biol Chem 289: 10261–10268. 10.1074/jbc.O114.555565 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lynes EM, Raturi A, Shenkman M, Ortiz Sandoval C, Yap MC, Wu J, Janowicz A, Myhill N, Benson MD, Campbell RE, et al. 2013. Palmitoylation is the switch that assigns calnexin to quality control or ER Ca2+ signaling. J Cell Sci 126: 3893–3903. 10.1242/jcs.125856 [DOI] [PubMed] [Google Scholar]
- Lytton J, Zarain-Herzberg A, Periasamy M, MacLennan DH. 1989. Molecular cloning of the mammalian smooth muscle sarco(endo)plasmic reticulum Ca2+-ATPase. J Biol Chem 264: 7059–7065. [PubMed] [Google Scholar]
- Lytton J, Westlin M, Burk SE, Shull GE, MacLennan DH. 1992. Functional comparisons between isoforms of the sarcoplasmic or endoplasmic reticulum family of calcium pumps. J Biol Chem 267: 14483–14489. [PubMed] [Google Scholar]
- Ma H, Inesi G, Toyoshima C. 2003. Substrate-induced conformational fit and headpiece closure in the Ca2+ ATPase (SERCA). J Biol Chem 278: 28938–28943. 10.1074/jbc.M304120200 [DOI] [PubMed] [Google Scholar]
- MacLennan DH. 1970. Purification and properties of an adenosine triphosphatase from sarcoplasmic reticulum. J Biol Chem 245: 4508–4518. [PubMed] [Google Scholar]
- MacLennan DH, Kranias EG. 2003. Phospholamban: A crucial regulator of cardiac contractility. Nat Rev Mol Cell Biol 4: 566–577. 10.1038/nrm1151 [DOI] [PubMed] [Google Scholar]
- Mahalingam D, Wilding G, Denmeade S, Sarantopoulas J, Cosgrove D, Cetnar J, Azad N, Bruce J, Kurman M, Allgood VE, et al. 2016. Mipsagargin, a novel thapsigargin-based PSMA-activated prodrug: results of a first-in-man phase I clinical trial in patients with refractory, advanced or metastatic solid tumours. Br J Cancer 114: 986–994. 10.1038/bjc.2016.72 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Makarewich CA, Munir AZ, Schiattarella GG, Bezprozvannaya S, Raguimova ON, Cho EE, Vidal AH, Robia SL, Bassel-Duby R, Olson EN. 2018. The DWORF micropeptide enhances contractility and prevents heart failure in a mouse model of dilated cardiomyopathy. eLife 7: e38319 10.7554/eLife.38319 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mandal D, Rulli SJ, Rao R. 2003. Packing interactions between transmembrane helices alter ion selectivity of the yeast Golgi Ca2+/Mn2+-ATPase PMR1. J Biol Chem 278: 35292–35298. 10.1074/jbc.M306166200 [DOI] [PubMed] [Google Scholar]
- Marcos D, Sepulveda MR, Berrocal M, Mata AM. 2009. Ontogeny of ATP hydrolysis and isoform expression of the plasma membrane Ca2+-ATPase in mouse brain. BMC Neurosci 10: 112 10.1186/1471-2202-10-112 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marques-da-Silva D, Gutierrez-Merino C. 2014. Caveolin-rich lipid rafts of the plasma membrane of mature cerebellar granule neurons are microcompartments for calcium/reactive oxygen and nitrogen species cross-talk signaling. Cell Calcium 56: 108–123. 10.1016/j.ceca.2014.06.002 [DOI] [PubMed] [Google Scholar]
- Mata AM, Sepúlveda MR. 2005. Calcium pumps in the central nervous system. Brain Res Brain Res Rev 49: 398–405. 10.1016/j.brainresrev.2004.11.004 [DOI] [PubMed] [Google Scholar]
- Mayosi BM, Kardos A, Davies CH, Gumedze F, Hovnanian A, Burge S, Watkins H. 2006. Heterozygous disruption of SERCA2a is not associated with impairment of cardiac performance in humans: Implications for SERCA2a as a therapeutic target in heart failure. Heart 92: 105–109. 10.1136/hrt.2004.051037 [DOI] [PMC free article] [PubMed] [Google Scholar]
- McAndrew D, Grice DM, Peters AA, Davis FM, Stewart T, Rice M, Smart CE, Brown MA, Kenny PA, Roberts-Thomson SJ, et al. 2011. ORAI1-mediated calcium influx in lactation and in breast cancer. Mol Cancer Therapeut 10: 448–460. 10.1158/1535-7163.MCT-10-0923 [DOI] [PubMed] [Google Scholar]
- Micaroni M, Giacchetti G, Plebani R, Xiao GG, Federici L. 2016. ATP2C1 gene mutations in Hailey–Hailey disease and possible roles of SPCA1 isoforms in membrane trafficking. Cell Death Dis 7: e2259 10.1038/cddis.2016.147 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Michaelis ML, Bigelow DJ, Schöneich C, Williams TD, Ramonda L, Yin D, Hühmer AF, Yao Y, Gao J, Squier TC. 1996. Decreased plasma membrane calcium transport activity in aging brain. Life Sci 59: 405–412. 10.1016/0024-3205(96)00319-0 [DOI] [PubMed] [Google Scholar]
- Michalak M, Robert Parker JM, Opas M. 2002. Ca2+ signaling and calcium binding chaperones of the endoplasmic reticulum. Cell Calcium 32: 269–278. 10.1016/S0143416002001884 [DOI] [PubMed] [Google Scholar]
- Miller KK, Verma A, Snyder SH, Ross CA. 1991. Localization of an endoplasmic reticulum calcium ATPase mRNA in rat brain by in situ hybridization. Neuroscience 43: 1–9. 10.1016/0306-4522(91)90410-P [DOI] [PubMed] [Google Scholar]
- Minamisawa S, Hoshijima M, Chu GX, Ward CA, Frank K, Gu YS, Martone ME, Wang YB, Ross J, Kranias EG, et al. 1999. Chronic phospholamban-sarcoplasmic reticulum calcium ATPase interaction is the critical calcium cycling defect in dilated cardiomyopathy. Cell 99: 313–322. 10.1016/S0092-8674(00)81662-1 [DOI] [PubMed] [Google Scholar]
- Missiaen L, Raeymaekers L, Dode L, Vanoevelen J, Van Baelen K, Parys JB, Callewaert G, De Smedt H, Segaert S, Wuytack F. 2004. SPCA1 pumps and Hailey–Hailey disease. Biochem Biophys Res Commun 322: 1204–1213. 10.1016/j.bbrc.2004.07.128 [DOI] [PubMed] [Google Scholar]
- Mohamed TM, Oceandy D, Zi M, Prehar S, Alatwi N, Wang Y, Shaheen MA, Abou-Leisa R, Schelcher C, Hegab Z, et al. 2011. Plasma membrane calcium pump (PMCA4)-neuronal nitric-oxide synthase complex regulates cardiac contractility through modulation of a compartmentalized cyclic nucleotide microdomain. J Biol Chem 286: 41520–41529. 10.1074/jbc.M111.290411 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Møller JV, Olesen C, Winther AM, Nissen P. 2010. The sarcoplasmic Ca2+-ATPase: Design of a perfect chemi-osmotic pump. Q Rev Biophys 43: 501–566. 10.1017/S003358351000017X [DOI] [PubMed] [Google Scholar]
- Moncoq K, Trieber CA, Young HS. 2007. The molecular basis for cyclopiazonic acid inhibition of the sarcoplasmic reticulum calcium pump. J Biol Chem 282: 9748–9757. 10.1074/jbc.M611653200 [DOI] [PubMed] [Google Scholar]
- Morine KJ, Sleeper MM, Barton ER, Sweeney HL. 2010. Overexpression of SERCA1a in the mdx diaphragm reduces susceptibility to contraction-induced damage. Hum Gene Ther 21: 1735–1739. 10.1089/hum.2010.077 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mountian I, Manolopoulos VG, De Smedt H, Parys JB, Missiaen L, Wuytack F. 1999. Expression patterns of sarco/endoplasmic reticulum Ca2+-ATPase and inositol 1,4,5-trisphosphate receptor isoforms in vascular endothelial cells. Cell Calcium 25: 371–380. 10.1054/ceca.1999.0034 [DOI] [PubMed] [Google Scholar]
- Nelson BR, Makarewich CA, Anderson DM, Winders BR, Troupes CD, Wu F, Reese AL, McAnally JR, Chen X, Kavalali ET, et al. 2016. A peptide encoded by a transcript annotated as long noncoding RNA enhances SERCA activity in muscle. Science 351: 271–275. 10.1126/science.aad4076 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Neville MC. 2005. Calcium secretion into milk. J Mammary Gland Biol Neoplasia 10: 119–128. 10.1007/s10911-005-5395-z [DOI] [PubMed] [Google Scholar]
- Niggli V, Penniston JT, Carafoli E. 1979. Purification of the (Ca2+-Mg2+)-ATPase from human erythrocyte membranes using a calmodulin affinity column. J Biol Chem 254: 9955–9958. 10.1016/0003-9861(79)90402-8 [DOI] [PubMed] [Google Scholar]
- Niggli V, Adunyah ES, Carafoli E. 1981. Acidic phospholipids, unsaturated fatty acids, and limited proteolysis mimic the effect of calmodulin on the purified erythrocyte Ca2+-ATPase. J Biol Chem 256: 8588–8592. 10.1016/0014-5793(82)80274-3 [DOI] [PubMed] [Google Scholar]
- Norimatsu Y, Hasegawa K, Shimizu N, Toyoshima C. 2017. Protein-phospholipid interplay revealed with crystals of a calcium pump. Nature 545: 193–198. 10.1038/nature22357 [DOI] [PubMed] [Google Scholar]
- Obara K, Miyashita N, Xu C, Toyoshima I, Sugita Y, Inesi G, Toyoshima C. 2005. Structural role of countertransport revealed in Ca2+ pump crystal structure in the absence of Ca2+. Proc Natl Acad Sci 102: 14489–14496. 10.1073/pnas.0506222102 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Oceandy D, Cartwright EJ, Emerson M, Prehar S, Baudoin FM, Zi M, Alatwi N, Venetucci L, Schuh K, Williams JC, et al. 2007. Neuronal nitric oxide synthase signaling in the heart is regulated by the sarcolemmal calcium pump 4b. Circulation 115: 483–492. 10.1161/circulationaha.106.643791 [DOI] [PubMed] [Google Scholar]
- Odermatt A, Taschner PE, Khanna VK, Busch HF, Karpati G, Jablecki CK, Breuning MH, MacLennan DH. 1996. Mutations in the gene-encoding SERCA1, the fast-twitch skeletal muscle sarcoplasmic reticulum Ca2+ ATPase, are associated with Brody disease. Nat Genet 14: 191–194. 10.1038/ng1096-191 [DOI] [PubMed] [Google Scholar]
- Okunade GW, Miller ML, Pyne GJ, Sutliff RL, O'Connor KT, Neumann JC, Andringa A, Miller DA, Prasad V, Doetschman T, et al. 2004. Targeted ablation of plasma membrane Ca2+-ATPase (PMCA) 1 and 4 indicates a major housekeeping function for PMCA1 and a critical role in hyperactivated sperm motility and male fertility for PMCA4. J Biol Chem 279: 33742–33750. 10.1074/jbc.M404628200 [DOI] [PubMed] [Google Scholar]
- Okunade GW, Miller ML, Azhar M, Andringa A, Sanford LP, Doetschman T, Prasad V, Shull GE. 2007. Loss of the Atp2c1 secretory pathway Ca2+-ATPase (SPCA1) in mice causes Golgi stress, apoptosis, and midgestational death in homozygous embryos and squamous cell tumors in adult heterozygotes. J Biol Chem 282: 26517–26527. 10.1074/jbc.M703029200 [DOI] [PubMed] [Google Scholar]
- Olanow CW. 2004. Manganese-induced parkinsonism and Parkinson's disease. Ann NY Acad Sci 1012: 209–223. 10.1196/annals.1306.018 [DOI] [PubMed] [Google Scholar]
- Olesen C, Sorensen TL, Nielsen RC, Moller JV, Nissen P. 2004. Dephosphorylation of the calcium pump coupled to counterion occlusion. Science 306: 2251–2255. 10.1126/science.1106289 [DOI] [PubMed] [Google Scholar]
- Olesen C, Picard M, Winther AM, Gyrup C, Morth JP, Oxvig C, Møller JV, Nissen P. 2007. The structural basis of calcium transport by the calcium pump. Nature 450: 1036–1042. 10.1038/nature06418 [DOI] [PubMed] [Google Scholar]
- Palmgren MG, Nissen P. 2011. P-type ATPases. Annu Rev Biophys 40: 243–266. 10.1146/annurev.biophys.093008.131331 [DOI] [PubMed] [Google Scholar]
- Pan Y, Zvaritch E, Tupling AR, Rice WJ, de Leon S, Rudnicki M, McKerlie C, Banwell BL, MacLennan DH. 2003. Targeted disruption of the ATP2A1 gene encoding the sarco(endo)plasmic reticulum Ca2+ ATPase isoform 1 (SERCA1) impairs diaphragm function and is lethal in neonatal mice. J Biol Chem 278: 13367–13375. 10.1074/jbc.M213228200 [DOI] [PubMed] [Google Scholar]
- Park WJ, Oh JG. 2013. SERCA2a: A prime target for modulation of cardiac contractility during heart failure. BMB Rep 46: 237–243. 10.5483/BMBRep.2013.46.5.077 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pászty K, Verma AK, Padányi R, Filoteo AG, Penniston JT, Enyedi A. 2002. Plasma membrane Ca2+ ATPase isoform 4b is cleaved and activated by caspase-3 during the early phase of apoptosis. J Biol Chem 277: 6822–6829. 10.1074/jbc.M109548200 [DOI] [PubMed] [Google Scholar]
- Periasamy M, Kalyanasundaram A. 2007. SERCA pump isoforms: Their role in calcium transport and disease. Muscle Nerve 35: 430–442. 10.1002/mus.20745 [DOI] [PubMed] [Google Scholar]
- Periasamy M, Reed TD, Liu LH, Ji Y, Loukianov E, Paul RJ, Nieman ML, Riddle T, Duffy JJ, Doetschman T, et al. 1999. Impaired cardiac performance in heterozygous mice with a null mutation in the sarco(endo)plasmic reticulum Ca2+-ATPase isoform 2 (SERCA2) gene. J Biol Chem 274: 2556–2562. 10.1074/jbc.274.4.2556 [DOI] [PubMed] [Google Scholar]
- Periasamy M, Bhupathy P, Babu GJ. 2008. Regulation of sarcoplasmic reticulum Ca2+ ATPase pump expression and its relevance to cardiac muscle physiology and pathology. Cardiovasc Res 77: 265–273. 10.1093/cvr/cvm056 [DOI] [PubMed] [Google Scholar]
- Periasamy M, Herrera JL, Reis FCG. 2017. Skeletal muscle thermogenesis and its role in whole body energy metabolism. Diabetes Metab J 41: 327–336. 10.4093/dmj.2017.41.5.327 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pestov NB, Dmitriev RI, Kostina MB, Korneenko TV, Shakhparonov MI, Modyanov NN. 2012. Structural evolution and tissue-specific expression of tetrapod-specific second isoform of secretory pathway Ca2+-ATPase. Biochem Biophys Res Commun 417: 1298–1303. 10.1016/j.bbrc.2011.12.135 [DOI] [PubMed] [Google Scholar]
- Peters AA, Milevskiy MJ, Lee WC, Curry MC, Smart CE, Saunus JM, Reid L, da Silva L, Marcial DL, Dray E, et al. 2016. The calcium pump plasma membrane Ca2+- ATPase 2 (PMCA2) regulates breast cancer cell proliferation and sensitivity to doxorubicin. Sci Rep 6: 25505 10.1038/srep25505 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pizzo P, Lissandron V, Capitanio P, Pozzan T. 2011. Ca2+ signalling in the Golgi apparatus. Cell Calcium 50: 184–192. 10.1016/j.ceca.2011.01.006 [DOI] [PubMed] [Google Scholar]
- Post RL, Hegyvary C, Kume S. 1972. Activation by adenosine triphosphate in the phosphorylation kinetics of sodium and potassium ion transport adenosine triphosphatase. J Biol Chem 247: 6530–6540. [PubMed] [Google Scholar]
- Pottorf WJ II, Johanns TM, Derrington SM, Strehler EE, Enyedi A, Thayer SA. 2006. Glutamate-induced protease-mediated loss of plasma membrane Ca2+ pump activity in rat hippocampal neurons. J Neurochem 98: 1646–1656. 10.1111/j.1471-4159.2006.04063.x [DOI] [PMC free article] [PubMed] [Google Scholar]
- Prasad V, Boivin GP, Miller ML, Liu LH, Erwin CR, Warner BW, Shull GE. 2005. Haploinsufficiency of Atp2a2, encoding the sarco(endo)plasmic reticulum Ca2+-ATPase isoform 2 Ca2+ pump, predisposes mice to squamous cell tumors via a novel mode of cancer susceptibility. Cancer Res 65: 8655–8661. 10.1158/0008-5472.CAN-05-0026 [DOI] [PubMed] [Google Scholar]
- Prasad V, Lorenz JN, Lasko VM, Nieman ML, Jiang M, Gao X, Rubinstein J, Wieczorek DF, Shull GE. 2014. Ablation of plasma membrane Ca2+-ATPase isoform 4 prevents development of hypertrophy in a model of hypertrophic cardiomyopathy. J Mol Cell Cardiol 77: 53–63. 10.1016/j.yjmcc.2014.09.025 [DOI] [PubMed] [Google Scholar]
- Prasad V, Lorenz JN, Lasko VM, Nieman ML, Huang W, Wang Y, Wieczorek DW, Shull GE. 2015. SERCA2 haploinsufficiency in a mouse model of Darier disease causes a selective predisposition to heart failure. BioMed Res Int 2015: 251598 10.1155/2015/251598 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Primeau JO, Armanious GP, Fisher ME, Young HS. 2018. The sarcoendoplasmic reticulum calcium ATPase. Subcell Biochem 87: 229–258. 10.1007/978-981-10-7757-9_8 [DOI] [PubMed] [Google Scholar]
- Raguimova ON, Smolin N, Bovo E, Bhayani S, Autry JM, Zima AV, Robia SL. 2018. Redistribution of SERCA calcium pump conformers during intracellular calcium signaling. J Biol Chem 293: 10843–10856. 10.1074/jbc.RA118.002472 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reinhardt TA, Lippolis JD, Shull GE, Horst RL. 2004. Null mutation in the gene encoding plasma membrane Ca2+-ATPase isoform 2 impairs calcium transport into milk. J Biol Chem 279: 42369–42373. 10.1074/jbc.M407788200 [DOI] [PubMed] [Google Scholar]
- Rivero-Rios P, Gomez-Suaga P, Fdez E, Hilfiker S. 2014. Upstream deregulation of calcium signaling in Parkinson's disease. Front Mol Neurosci 7: 53 10.3389/fnmol.2014.00053 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rudolph HK, Antebi A, Fink GR, Buckley CM, Dorman TE, LeVitre J, Davidow LS, Mao JI, Moir DT. 1989. The yeast secretory pathway is perturbed by mutations in PMR1, a member of a Ca2+ ATPase family. Cell 58: 133–145. 10.1016/0092-8674(89)90410-8 [DOI] [PubMed] [Google Scholar]
- Ryan ZC, Craig TA, Filoteo AG, Westendorf JJ, Cartwright EJ, Neyses L, Strehler EE, Kumar R. 2015. Deletion of the intestinal plasma membrane calcium pump, isoform 1, Atp2b1, in mice is associated with decreased bone mineral density and impaired responsiveness to 1, 25-dihydroxyvitamin D3. Biochem Biophys Res Commun 467: 152–156. 10.1016/j.bbrc.2015.09.087 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sahoo SK, Shaikh SA, Sopariwala DH, Bal NC, Periasamy M. 2013. Sarcolipin protein interaction with sarco(endo)plasmic reticulum Ca2+ ATPase (SERCA) is distinct from phospholamban protein, and only sarcolipin can promote uncoupling of the SERCA pump. J Biol Chem 288: 6881–6889. 10.1074/jbc.M112.436915 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sakuntabhai A, Ruiz-Perez V, Carter S, Jacobsen N, Burge S, Monk S, Smith M, Munro CS, O'Donovan M, Craddock N, et al. 1999. Mutations in ATP2A2, encoding a Ca2+ pump, cause Darier disease. Nat Genet 21: 271–277. 10.1038/6784 [DOI] [PubMed] [Google Scholar]
- Salvador JM, Inesi G, Rigaud JL, Mata AM. 1998. Ca2+ transport by reconstituted synaptosomal ATPase is associated with H+ countertransport and net charge displacement. J Biol Chem 273: 18230–18234. 10.1074/jbc.273.29.18230 [DOI] [PubMed] [Google Scholar]
- Samuel TJ, Rosenberry RP, Lee S, Pan Z. 2018. Correcting calcium dysregulation in chronic heart failure using SERCA2a gene therapy. Int J Mol Sci 19: E1086 10.3390/ijms19041086 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schatzmann HJ. 1966. ATP-dependent Ca++-extrusion from human red cells. Experientia 22: 364–365. 10.1007/BF01901136 [DOI] [PubMed] [Google Scholar]
- Schmidt N, Kollewe A, Constantin CE, Henrich S, Ritzau-Jost A, Bildl W, Saalbach A, Hallermann S, Kulik A, Fakler B, et al. 2017. Neuroplastin and basigin are essential auxiliary subunits of plasma membrane Ca2+-ATPases and key regulators of Ca2+ clearance. Neuron 96: 827–838.e9. 10.1016/j.neuron.2017.09.038 [DOI] [PubMed] [Google Scholar]
- Schuh K, Uldrijan S, Gambaryan S, Roethlein N, Neyses L. 2003. Interaction of the plasma membrane Ca2+ pump 4b/CI with the Ca2+/calmodulin-dependent membrane-associated kinase CASK. J Biol Chem 278: 9778–9783. 10.1074/jbc.M212507200 [DOI] [PubMed] [Google Scholar]
- Schuh K, Cartwright EJ, Jankevics E, Bundschu K, Liebermann J, Williams JC, Armesilla AL, Emerson M, Oceandy D, Knobeloch KP, et al. 2004. Plasma membrane Ca2+ ATPase 4 is required for sperm motility and male fertility. J Biol Chem 279: 28220–28226. 10.1074/jbc.M312599200 [DOI] [PubMed] [Google Scholar]
- Sepúlveda MR, Berrocal-Carrillo M, Gasset M, Mata AM. 2006. The plasma membrane Ca2+-ATPase isoform 4 is localized in lipid rafts of cerebellum synaptic plasma membranes. J Biol Chem 281: 447–453. 10.1074/jbc.M506950200 [DOI] [PubMed] [Google Scholar]
- Sepúlveda MR, Vanoevelen J, Raeymaekers L, Mata AM, Wuytack F. 2009. Silencing the SPCA1 (secretory pathway Ca2+-ATPase isoform 1) impairs Ca2+ homeostasis in the Golgi and disturbs neural polarity. J Neurosci 29: 12174–12182. 10.1523/jneurosci.2014-09.2009 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sepúlveda MR, Dresselaers T, Vangheluwe P, Everaerts W, Himmelreich U, Mata AM, Wuytack F. 2012a. Evaluation of manganese uptake and toxicity in mouse brain during continuous MnCl2 administration using osmotic pumps. Contrast Media Mol I 7: 426–434. 10.1002/cmmi.1469 [DOI] [PubMed] [Google Scholar]
- Sepúlveda MR, Wuytack F, Mata AM. 2012b. High levels of Mn2+ inhibit secretory pathway Ca2+/Mn2+-ATPase (SPCA) activity and cause Golgi fragmentation in neurons and glia. J Neurochem 123: 824–836. 10.1111/j.1471-4159.2012.07888.x [DOI] [PubMed] [Google Scholar]
- Shaikh SA, Sahoo SK, Periasamy M. 2016. Phospholamban and sarcolipin: Are they functionally redundant or distinct regulators of the sarco(endo)plasmic reticulum calcium ATPase? J Mol Cell Cardiol 91: 81–91. 10.1016/j.yjmcc.2015.12.030 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shin YB, Lim JE, Ji SM, Lee HJ, Park SY, Hong KW, Lim M, McCarthy MI, Lee YH, Oh B. 2013. Silencing of Atp2b1 increases blood pressure through vasoconstriction. J Hypertens 31: 1575–1583. 10.1097/HJH.0b013e32836189e9 [DOI] [PubMed] [Google Scholar]
- Simmerman HK, Jones LR. 1998. Phospholamban: Protein structure, mechanism of action, and role in cardiac function. Physiol Rev 78: 921–947. 10.1152/physrev.1998.78.4.921 [DOI] [PubMed] [Google Scholar]
- Sitsel A, De Raeymaecker J, Drachmann ND, Derua R, Smaardijk S, Andersen JL, Vandecaetsbeek I, Chen J, De Maeyer M, Waelkens E, et al. 2019. Structures of the heart specific SERCA2a Ca2+-ATPase. EMBO J 38: e100020 10.15252/embj.2018100020 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smaardijk S, Chen J, Wuytack F, Vangheluwe P. 2017. SPCA2 couples Ca2+ influx via Orai1 to Ca2+ uptake into the Golgi/secretory pathway. Tissue Cell 49: 141–149. 10.1016/j.tice.2016.09.004 [DOI] [PubMed] [Google Scholar]
- Smaardijk S, Chen J, Kerselaers S, Voets T, Eggermont J, Vangheluwe P. 2018. Store-independent coupling between the secretory pathway Ca2+ transport ATPase SPCA1 and Orai1 in Golgi stress and Hailey–Hailey disease. Biochim Biophys Acta 1865: 855–862. 10.1016/j.bbamcr.2018.03.007 [DOI] [PubMed] [Google Scholar]
- Sorensen TL, Moller JV, Nissen P. 2004. Phosphoryl transfer and calcium ion occlusion in the calcium pump. Science 304: 1672–1675. 10.1126/science.1099366 [DOI] [PubMed] [Google Scholar]
- Spiden SL, Bortolozzi M, Di Leva F, de Angelis MH, Fuchs H, Lim D, Ortolano S, Ingham NJ, Brini M, Carafoli E, et al. 2008. The novel mouse mutation Oblivion inactivates the PMCA2 pump and causes progressive hearing loss. PLoS Genet 4: e1000238 10.1371/journal.pgen.1000238 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stafford N, Wilson C, Oceandy D, Neyses L, Cartwright EJ. 2017. The plasma membrane calcium ATPases and their role as major new players in human disease. Physiol Rev 97: 1089–1125. 10.1152/physrev.00028.2016 [DOI] [PubMed] [Google Scholar]
- Stammers AN, Susser SE, Hamm NC, Hlynsky MW, Kimber DE, Kehler DS, Duhamel TA. 2015. The regulation of sarco(endo)plasmic reticulum calcium-ATPases (SERCA). Can J Physiol Pharmacol 93: 843–854. 10.1139/cjpp-2014-0463 [DOI] [PubMed] [Google Scholar]
- Street VA, McKee-Johnson JW, Fonseca RC, Tempel BL, Noben-Trauth K. 1998. Mutations in a plasma membrane Ca2+-ATPase gene cause deafness in deafwaddler mice. Nat Genet 19: 390–394. 10.1038/1284 [DOI] [PubMed] [Google Scholar]
- Strehler EE. 2015. Plasma membrane calcium ATPases: From generic Ca2+ sump pumps to versatile systems for fine-tuning cellular Ca2+. Biochem Biophys Res Commun 460: 26–33. 10.1016/j.bbrc.2015.01.121 [DOI] [PubMed] [Google Scholar]
- Strehler EE, Zacharias DA. 2001. Role of alternative splicing in generating isoform diversity among plasma membrane calcium pumps. Physiol Rev 81: 21–50. 10.1152/physrev.2001.81.1.21 [DOI] [PubMed] [Google Scholar]
- Strehler EE, James P, Fischer R, Heim R, Vorherr T, Filoteo AG, Penniston JT, Carafoli E. 1990. Peptide sequence analysis and molecular cloning reveal two calcium pump isoforms in the human erythrocyte membrane. J Biol Chem 265: 2835–2842. [PubMed] [Google Scholar]
- Suzuki J, Kanemaru K, Iino M. 2016. Genetically encoded fluorescent indicators for organellar calcium imaging. Biophys J 111: 1119–1131. 10.1016/j.bpj.2016.04.054 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi K, Kitamura K. 1999. A point mutation in a plasma membrane Ca2+-ATPase gene causes deafness in Wriggle Mouse Sagami. Biochem Biophys Res Commun 261: 773–778. 10.1006/bbrc.1999.1102 [DOI] [PubMed] [Google Scholar]
- Tatsuki F, Sunagawa GA, Shi S, Susaki EA, Yukinaga H, Perrin D, Sumiyama K, Ukai-Tadenuma M, Fujishima H, Ohno R, et al. 2016. Involvement of Ca2+-dependent hyperpolarization in sleep duration in mammals. Neuron 90: 70–85. 10.1016/j.neuron.2016.02.032 [DOI] [PubMed] [Google Scholar]
- Tauber P, Aichinger B, Christ C, Stindl J, Rhayem Y, Beuschlein F, Warth R, Bandulik S. 2016. Cellular pathophysiology of an adrenal adenoma-associated mutant of the plasma membrane Ca2+-ATPase ATP2B3. Endocrinology 157: 2489–2499. 10.1210/en.2015-2029 [DOI] [PubMed] [Google Scholar]
- Tidow H, Poulsen LR, Andreeva A, Knudsen M, Hein KL, Wiuf C, Palmgren MG, Nissen P. 2012. A bimodular mechanism of calcium control in eukaryotes. Nature 491: 468–472. 10.1038/nature11539 [DOI] [PubMed] [Google Scholar]
- Tong X, Kono T, Anderson-Baucum EK, Yamamoto W, Gilon P, Lebeche D, Day RN, Shull GE, Evans-Molina C. 2016. SERCA2 deficiency impairs pancreatic β-cell function in response to diet-induced obesity. Diabetes 65: 3039–3052. 10.2337/db16-0084 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toral-Ojeda I, Aldanondo G, Lasa-Elgarresta J, Lasa-Fernández H, Fernández-Torrón R, de Munain A L, Vallejo-Illarramendi A. 2016. Calpain 3 deficiency affects SERCA expression and function in the skeletal muscle. Exp Rev Mol Med 18: e7 10.1017/erm.2016.9 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toyoshima C. 2009. How Ca2+-ATPase pumps ions across the sarcoplasmic reticulum membrane. Biochim Biophys Acta 1793: 941–946. 10.1016/j.bbamcr.2008.10.008 [DOI] [PubMed] [Google Scholar]
- Toyoshima C, Mizutani T. 2004. Crystal structure of the calcium pump with a bound ATP analogue. Nature 430: 529–535. 10.1038/nature02680 [DOI] [PubMed] [Google Scholar]
- Toyoshima C, Nomura H. 2002. Structural changes in the calcium pump accompanying the dissociation of calcium. Nature 418: 605–611. 10.1038/nature00944 [DOI] [PubMed] [Google Scholar]
- Toyoshima C, Nakasako M, Nomura H, Ogawa H. 2000. Crystal structure of the calcium pump of sarcoplasmic reticulum at 2.6 Å resolution. Nature 405: 647–655. 10.1038/35015017 [DOI] [PubMed] [Google Scholar]
- Toyoshima C, Nomura H, Tsuda T. 2004. Lumenal gating mechanism revealed in calcium pump crystal structures with phosphate analogues. Nature 432: 361–368. 10.1038/nature02981 [DOI] [PubMed] [Google Scholar]
- Toyoshima C, Iwasawa S, Ogawa H, Hirata A, Tsueda J, Inesi G. 2013. Crystal structures of the calcium pump and sarcolipin in the Mg2+-bound E1 state. Nature 495: 260–264. 10.1038/nature11899 [DOI] [PubMed] [Google Scholar]
- Ueno T, Kameyama K, Hirata M, Ogawa M, Hatsuse H, Takagaki Y, Ohmura M, Osawa N, Kudo Y. 2002. A mouse with a point mutation in plasma membrane Ca2+-ATPase isoform 2 gene showed the reduced Ca2+ influx in cerebellar neurons. Neurosci Res 42: 287–297. 10.1016/S0168-0102(02)00008-1 [DOI] [PubMed] [Google Scholar]
- Vafiadaki E, Arvanitis DA, Pagakis SN, Papalouka V, Sanoudou D, Kontrogianni-Konstantopoulos A, Kranias EG. 2009. The anti-apoptotic protein HAX-1 interacts with SERCA2 and regulates its protein levels to promote cell survival. Mol Biol Cell 20: 306–318. 10.1091/mbc.e08-06-0587 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Baelen K, Vanoevelen J, Missiaen L, Raeymaekers L, Wuytack F. 2001. The Golgi PMR1 P-type ATPase of Caenorhabditis elegans. Identification of the gene and demonstration of calcium and manganese transport. J Biol Chem 276: 10683–10691. 10.1074/jbc.M010553200 [DOI] [PubMed] [Google Scholar]
- Vandecaetsbeek I, Trekels M, De Maeyer M, Ceulemans H, Lescrinier E, Raeymaekers L, Wuytack F, Vangheluwe P. 2009. Structural basis for the high Ca2+ affinity of the ubiquitous SERCA2b Ca2+ pump. Proc Natl Acad Sci 106: 18533–18538. 10.1073/pnas.0906797106 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vandecaetsbeek I, Vangheluwe P, Raeymaekers L, Wuytack F, Vanoevelen J. 2011. The Ca2+ pumps of the endoplasmic reticulum and Golgi apparatus. Cold Spring Harb Perspect Biol 3: a004184 10.1101/cshperspect.a004184 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vangheluwe P, Raeymaekers L, Dode L, Wuytack F. 2005a. Modulating sarco(endo)plasmic reticulum Ca2+ ATPase 2 (SERCA2) activity: Cell biological implications. Cell Calcium 38: 291–302. 10.1016/j.ceca.2005.06.033 [DOI] [PubMed] [Google Scholar]
- Vangheluwe P, Schuermans M, Zádor E, Waelkens E, Raeymaekers L, Wuytack F. 2005b. Sarcolipin and phospholamban mRNA and protein expression in cardiac and skeletal muscle of different species. Biochem J 389: 151–159. 10.1042/BJ20050068 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vangheluwe P, Sepulveda MR, Missiaen L, Raeymaekers L, Wuytack F, Vanoevelen J. 2009. Intracellular Ca2+- and Mn2+-transport ATPases. Chem Rev 109: 4733–4759. 10.1021/cr900013m [DOI] [PubMed] [Google Scholar]
- VanHouten J, Sullivan C, Bazinet C, Ryoo T, Camp R, Rimm DL, Chung G, Wysolmerski J. 2010. PMCA2 regulates apoptosis during mammary gland involution and predicts outcome in breast cancer. Proc Natl Acad Sci 107: 11405–11410. 10.1073/pnas.0911186107 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vanoevelen J, Dode L, Van Baelen K, Fairclough RJ, Missiaen L, Raeymaekers L FW. 2005. The secretory pathway Ca2+/Mn2+-ATPase 2 is a Golgi-localized pump with high affinity for Ca2+ ions. J Biol Chem 280: 22800–22808. 10.1074/jbc.M501026200 [DOI] [PubMed] [Google Scholar]
- Varadi A, Lebel L, Hashim Y, Mehta Z, Ashcroft SJH, Turner R. 1999. Sequence variants of the sarco(endo)plasmic reticulum Ca2+-transport ATPase 3 gene (SERCA3) in Caucasian type II diabetic patients (UH Prospective Diabetes Study 48). Diabetologia 42: 1240–1243. 10.1007/s001250051298 [DOI] [PubMed] [Google Scholar]
- Verboomen H, Wuytack F, De Smedt H, Himpens B, Casteels R. 1992. Functional difference between SERCA2a and SERCA2b Ca2+ pumps and their modulation by phospholamban. Biochem J 286: 591–595. 10.1042/bj2860591 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vicario M, Calì T, Cieri D, Vallese F, Bortolotto R, Lopreiato R, Zonta F, Nardella M, Micalizzi A, Lefeber DJ, et al. 2017. A novel PMCA3 mutation in an ataxic patient with hypomorphic phosphomannomutase 2 (PMM2) heterozygote mutations: Biochemical characterization of the pump defect. Biochim Biophys Acta 1863: 3303–3312. 10.1016/j.bbadis.2017.08.006 [DOI] [PubMed] [Google Scholar]
- Vicario M, Zanni G, Vallese F, Santorelli F, Grinzato A, Cieri D, Berto P, Frizzarin M, Lopreiato R, Zonta F, et al. 2018. A V1143F mutation in the neuronal-enriched isoform 2 of the PMCA pump is linked with ataxia. Neurobiol Dis 115: 157–166. 10.1016/j.nbd.2018.04.009 [DOI] [PubMed] [Google Scholar]
- von Blume J, Alleaume AM, Kienzle C, Carreras-Sureda A, Valverde M, Malhotra V. 2012. Cab45 is required for Ca2+-dependent secretory cargo sorting at the trans-Golgi network. J Cell Biol 199: 1057–1066. 10.1083/jcb.201207180 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y, Wilson C, Cartwright EJ, Lei M. 2017. Plasma membrane Ca2+ -ATPase 1 is required for maintaining atrial Ca2+ homeostasis and electrophysiological stability in the mouse. J Physiol 595: 7383–7398. 10.1113/JP274110 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wegener AD, Simmerman HK, Lindemann JP, Jones LR. 1989. Phospholamban phosphorylation in intact ventricles. Phosphorylation of serine 16 and threonine 17 in response to β-adrenergic stimulation. J Biol Chem 264: 11468–11474. [PubMed] [Google Scholar]
- Wei Y, Marchi V, Wang R, Rao R. 1999. An N-terminal EF hand-like motif modulates ion transport by Pmr1, the yeast Golgi Ca2+/Mn2+-ATPase. Biochemistry 38: 14534–14541. 10.1021/bi9911233 [DOI] [PubMed] [Google Scholar]
- Winther AM, Bublitz M, Karlsen JL, Moller JV, Hansen JB, Nissen P, Buch-Pedersen MJ. 2013. The sarcolipin-bound calcium pump stabilizes calcium sites exposed to the cytoplasm. Nature 495: 265–269. 10.1038/nature11900 [DOI] [PubMed] [Google Scholar]
- Wootton LL, Argent CC, Wheatley M, Michelangeli F. 2004. The expression, activity and localisation of the secretory pathway Ca2+-ATPase (SPCA1) in different mammalian tissues. Biochim Biophys Acta 1664: 189–197. 10.1016/j.bbamem.2004.05.009 [DOI] [PubMed] [Google Scholar]
- Wu KD, Bungard D, Lytton J. 2001. Regulation of SERCA Ca2+ pump expression by cytoplasmic Ca2+ in vascular smooth muscle cells. Am J Physiol Cell Physiol 280: C843–C851. 10.1152/ajpcell.2001.280.4.C843 [DOI] [PubMed] [Google Scholar]
- Wuytack F, Papp B, Verboomen H, Raeymaekers L, Dode L, Bobe R, Enouf J, Bokkala S, Authi KS, Casteels R. 1994. A sarco/endoplasmic reticulum Ca2+-ATPase 3-type Ca2+ pump is expressed in platelets, in lymphoid cells, and in mast cells. J Biol Chem 269: 1410–1416. [PubMed] [Google Scholar]
- Wuytack F, Dode L, Baba-Aissa F, Raeymaekers L. 1995. The SERCA3-type of organellar Ca2+ pumps. Biosci Rep 15: 299–306. 10.1007/BF01788362 [DOI] [PubMed] [Google Scholar]
- Xiang M, Mohamalawari D, Rao R. 2005. A novel isoform of the secretory pathway Ca2+, Mn2+-ATPase, hSPCA2, has unusual properties and is expressed in the brain. J Biol Chem 280: 11608–11614. 10.1074/jbc.M413116200 [DOI] [PubMed] [Google Scholar]
- Xie LH, Shanmugam M, Park JY, Zhao Z, Wen H, Tian B, Periasamy M, Babu GJ. 2012. Ablation of sarcolipin results in atrial remodeling. Am J Physiol Cell Physiol 302: C1762–C1771. 10.1152/ajpcell.00425.2011 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu XY, Gou WF, Yang X, Wang GL, Takahashi H, Yu M, Mao XY, Takano Y, Zheng HC. 2012. Aberrant SERCA3 expression is closely linked to pathogenesis, invasion, metastasis, and prognosis of gastric carcinomas. Tumour Biol 33: 1845–1854. 10.1007/s13277-012-0444-x [DOI] [PubMed] [Google Scholar]
- Zádor E, Vangheluwe P, Wuytack F. 2007. The expression of the neonatal sarcoplasmic reticulum Ca2+ pump (SERCA1b) hints to a role in muscle growth and development. Cell Calcium 41: 379–388. 10.1016/j.ceca.2006.08.001 [DOI] [PubMed] [Google Scholar]
- Zanni G, Cali T, Kalscheuer VM, Ottolini D, Barresi S, Lebrun N, Montecchi-Palazzi L, Hu H, Chelly J, Bertini E, et al. 2012. Mutation of plasma membrane Ca2+ ATPase isoform 3 in a family with X-linked congenital cerebellar ataxia impairs Ca2+ homeostasis. Proc Natl Acad Sci 109: 14514–14519. 10.1073/pnas.1207488109 [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zarain-Herzberg A, MacLennan DH, Periasamy M. 1990. Characterization of rabbit cardiac sarco(endo)plasmic reticulum Ca2+-ATPase gene. J Biol Chem 265: 4670–4677. 10.1100/tsw.2002.228 [DOI] [PubMed] [Google Scholar]
- Zhao Y, Ogawa H, Yonekura S, Mitsuhashi H, Mitsuhashi S, Nishino I, Toyoshima C, Ishiura S. 2015. Functional analysis of SERCA1b, a highly expressed SERCA1 variant in myotonic dystrophy type 1 muscle. Biochim Biophys Acta 1852: 2042–2047. 10.1016/j.bbadis.2015.07.006 [DOI] [PubMed] [Google Scholar]
- Zhao YG, Chen Y, Miao G, Zhao H, Qu W, Li D, Wang Z, Liu N, Li L, Chen S, et al. 2017. The ER-localized transmembrane protein EPG-3/VMP1 regulates SERCA activity to control ER-isolation membrane contacts for autophagosome formation. Mol Cell 67: 974–989.e976. [DOI] [PubMed] [Google Scholar]
- Zylinska L, Guerini D, Gromadzinska E, Lachowicz L. 1998. Protein kinases A and C phosphorylate purified Ca2+- ATPase from rat cortex, cerebellum and hippocampus. Biochim Biophys Acta 1448: 99–108. 10.1016/S0167-4889(98)00128-1 [DOI] [PubMed] [Google Scholar]