Next Article in Journal
Urinary Exosomes Identify Inflammatory Pathways in Vancomycin Associated Acute Kidney Injury
Next Article in Special Issue
Calmodulin and Its Binding Proteins in Parkinson’s Disease
Previous Article in Journal
TRP Channels as Cellular Targets of Particulate Matter
Previous Article in Special Issue
Structural Aspects and Prediction of Calmodulin-Binding Proteins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Crosstalk among Calcium ATPases: PMCA, SERCA and SPCA in Mental Diseases

1
Department of Molecular Neurochemistry, Medical University of Lodz, 92215 Lodz, Poland
2
Department of Pharmaceutical Toxicology, China Medical University, Shenyang 110122, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2021, 22(6), 2785; https://doi.org/10.3390/ijms22062785
Submission received: 24 February 2021 / Revised: 7 March 2021 / Accepted: 8 March 2021 / Published: 10 March 2021
(This article belongs to the Special Issue Calmodulin Function in Health and Disease 2.0)

Abstract

:
Calcium in mammalian neurons is essential for developmental processes, neurotransmitter release, apoptosis, and signal transduction. Incorrectly processed Ca2+ signal is well-known to trigger a cascade of events leading to altered response to variety of stimuli and persistent accumulation of pathological changes at the molecular level. To counterbalance potentially detrimental consequences of Ca2+, neurons are equipped with sophisticated mechanisms that function to keep its concentration in a tightly regulated range. Calcium pumps belonging to the P-type family of ATPases: plasma membrane Ca2+-ATPase (PMCA), sarco/endoplasmic Ca2+-ATPase (SERCA) and secretory pathway Ca2+-ATPase (SPCA) are considered efficient line of defense against abnormal Ca2+ rises. However, their role is not limited only to Ca2+ transport, as they present tissue-specific functionality and unique sensitive to the regulation by the main calcium signal decoding protein—calmodulin (CaM). Based on the available literature, in this review we analyze the contribution of these three types of Ca2+-ATPases to neuropathology, with a special emphasis on mental diseases.

1. Introduction

Ca2+-ATPases are key components of Ca2+ extrusion machinery and thus are pivotal for preservation of neuronal function. Among three main calcium pumps, the plasma membrane Ca2+-ATPase (PMCA) and sarco/endoplasmic Ca2+-ATPase (SERCA) are known for decades while the secretory pathway Ca2+-ATPase has been discovered in 2000s by two independent laboratories that described novel mutations leading to Hailey-Hailey disease [1,2,3]. All pumps have high affinity for Ca2+ and function to restore cytosolic Ca2+ concentration [Ca2+]c to the resting, nanomolar level following neuronal stimulations. They belong to the superfamily of mammalian P-type ATPases and are characterized by formation of a phosphorylated enzyme intermediate during catalytic cycle [1]. However, they have a low (~15%) degree of sequency identity [1], and differ in several other key features including tissue distribution, regulatory mechanisms, and contribution to neuronal Ca2+ homeostasis. Each pump is encoded by multiple genes giving rise to a number of isoforms and further splice variants, which often possess distinguishable kinetic parameters and are dedicated to unique and highly regulated neural processes [4]. Naturally, the pumps share essential basic properties such as membrane topology, catalytic mechanism and probably the general features of 3D structure [5,6], although the structure of SPCA pump has not been solved yet (Figure 1). The rapid expansion of the knowledge on pumps peculiar role, which run parallel to the advances in neuronal Ca2+ signaling, led to the identification of several diseases associated either directly or indirectly with Ca2+ pumps malfunction. Most of these defects have genetic background and the number of studies have been aimed to characterize their severity, effect on neuronal Ca2+ homeostasis and signaling as well as neuronal survival. Besides known neuropathologies, defects in Ca2+ pumps and alterations in the mechanisms regulating their activity may also produce subtle, tissue-specific disturbances that are not clinically manifested, yet they may affect neuronal machinery controlling and processing Ca2+ signal. In this review, we focus on the contribution of PMCA, SERCA and SPCA to mental diseases and give a special emphasis to altered regulation by calmodulin (CaM) that often accompanies pump defects.

2. Calmodulin—Ubiquitous Ca2+ Sensor in Neurons

Most commonly, detection and transduction of Ca2+ signals in neurons are orchestrated by ubiquitous messenger called calmodulin. CaM is known as a relatively small (149aa; 16.7 kDa) and highly conserved calcium-binding sensor synthesized in all eukaryotic cells. It is particularity involved in synaptic signaling processes, neurotransmitter release and neuroplasticity by modulation (called “calmodulation”) of a large array of binding partners such as enzymes (e.g., adenylate cyclase, calcineurin, cyclic nucleotide phosphodiesterase, nitric oxide synthase, and certain kinases), transcription factors (e.g., CREB, NeuroD2, NFAT and MEF2) as well as various ion channels and transporters [7,8,9,10].
In human, CaM is encoded by three independent genes CALM1, CALM2, CALM3 located on chromosomes 14q32.11; 2p21; and 19q13.32, respectively, which are collectively transcribed into at least eight mRNAs using different alternative polyadenylation signals (reviewed in [7,11]). Next, the resulting protein is susceptible to undergo various post-translational modifications, mainly phosphorylation on tyrosine (Thr26, Thr29, Thr44, Thr79, Tyr99, Thr117, and Tyr138) and serine (Ser81, and Ser101) sites [12]; acetylation of the N-terminal alanine [13]; trimethylation of the Lys115 [14]; and proteolytic cleavage at the C-terminal domain [15], all collectively regulating CaM biological activity. The crystal structure of mature CaM contains two independently folded lobes (N-lobe and C-lobe) connected by a flexible central α-helical linker, that differ by calcium affinity and kinetics of calcium dissociation. Each of these globular clusters can bind up to two free Ca2+ ions via a pair of helix-loop-helix motives (EF-hands) in a cooperative manner (Kd = 5 · 10−7 M to 5 · 10−6 M) [16,17,18]. Because of subtle structural differences between these lobes resulting from evolutionary processes [19], EF hands in the C-lobe exhibit a three- to five times higher affinity for Ca2+. However, they possess slower rate of ion binding than the regions of EF hands located in the N-lobe, establishing the broad range of CaM sensitivity to the changes in calcium concentrations in the intracellular space [20]. CaM is susceptible to dramatic structural rearrangements via partially exposed hydrophobic patch on the C-terminal domain which may interact with CaM-binding proteins (CaMBPs) in a Ca2+ -free (apo-CaM) state or in partially calcium-saturated forms (two Ca2+ ions bound to the C-terminus) [16]. Up to date, over three hundred different calmodulin targets with specific binding sites and unique affinities for CaM, many of which located in the central nervous system (CNS) neurons [21], have been validated and extensively characterized [22]. The analysis of over 80 CaM complexes compiled in the Protein Data Bank (PDB) has revealed that CaM binding sites not always contain defined consensus sequence but rather share some common biochemical and biophysical properties such as high helix-forming propensities, positively charged binding region and the presence of hydrophobic anchor residues [8,22]. Thus, the classification of several CaM-binding motifs is determined by the spacing between these anchor residues as was extensively discussed by Mruk and colleagues [23]. As observed from sequence analysis of several CaMBPs, their IQ motif ([FILV]Qxxx[RK]Gxxx[RK]xx[FILVWY]) with highly conserved amino acid residues at positions 1, 2, 5, 6, 11, and 14 or IQ-like ([FILV]Qxxx[RK]Gxxxxxxxx) motif may also bind CaM in the presence or absence of Ca2+ [23,24].
Considering the diversity of CaM interactions and its abundance in the brain (up to 100 μM range) [25], it seems rational to suspect that disruption of these multifunctional interactions regulating Ca2+-dependent intracellular signal transduction cascades may be implicated in the development of numerous neuropsychiatric disorders. Moreover, there is increasing evidence suggesting that pathophysiology of these states is intimately related to the disturbed neuronal calcium homeostasis also mediated by ATP-driven pumps located in the plasma membrane, in the membranes of the endoplasmic reticulum (ER), or Golgi compartments.

3. Plasma Membrane Ca2+-ATPase (PMCA)—The Only Calcium Pump Directly Regulated by Calmodulin

PMCA is one of the most important and sensitive players in maintaining of low resting Ca2+ concentration, and ensuring a fast recovery of [Ca2+]c to the basal level following neuronal excitation [26]. The enzyme was first described by Schatzmann in 1960s as ATP-powered mechanism that removes calcium from red blood cells [27], whereas further studies revealed the presence of PMCA in other cells, including neurons [28,29,30] Structurally, PMCA comprises of ten transmembrane segments with N- and C- terminal tails both located on the cytosolic site [31]. Most of the regulatory regions including acidic phospholipids, protein kinase C (PKC), protein kinase A (PKA) and the crucial natural activator—CaM, are located at the C- terminus. The important regulatory role of CaM in stimulating of PMCA is associated with increasing the affinity of the pump for calcium and the maximum rate of calcium extrusion. In the activation process, CaM removes the auto-inhibitory C-terminal domain from the active site and releases the enzyme from auto-inhibition [32]. It is also worth mentioning that PMCA is so far the only known calcium pump directly activated by CaM [26].
In mammals, four isoforms of PMCA (PMCA1-PMCA4), structurally similar to each other, have been found [4] but their expression depends on cell type (Table 1). The PMCA1 and PMCA4 are widely expressed in virtually all animal tissues and both play a house-keeping role.
Expression of PMCA2 and PMCA3 is highly restricted to excitable cells and their high concentration has been detected in the CNS [4]. PMCA2 is especially abundant in cerebellar Purkinje cells and granule cells, but it also localizes to the cerebral cortex and hippocampus [33]. PMCA3, in turn, is present predominantly in cerebellar granule cells and in the choroid plexus [34] what suggests its role in generation and release of cerebrospinal fluid. Additionally, PMCA isoforms are characterized by distinct calmodulin sensitivity (Table 1) and specific kinetic properties. PMCA2 and PMCA3 are referred to as “fast” isoforms due to their high basal activity and high affinity for CaM, whereas PMCA1 and PMCA4 are much slower despite their strong stimulation by CaM [1]. It has been suggested that the cell response to a physiological stimulus depends on significant differences in the kinetic parameters of the individual isoforms. In the brain, distribution of PMCA isoforms clearly alters during development, what may indicate their specific role in embryogenesis and further in postnatal period [35].
In addition to control critical neuronal functions such as synaptic transition and neurotransmitter release, neuronal Ca2+ also participates in the regulation of survival and differentiation, processes common to other cell types [36]. Early in vitro study on differentiated pheochromocytoma-derived cells, a model frequently used to mimic the physiology of sympathetic neurons, has shown that PMCA1 knockdown impaired neuritogenesis and axonal elongation [37]. Similar effect was seen when PMCA2 or PMCA3 expression was decreased with isoform-specific antisense oligonucleotides [38,39] suggesting a role of PMCA in neuronal differentiation. Moreover, cells deficient in neuron-specific PMCA isoforms were unable to maintain Ca2+ and pH homeostasis which translated into altered activity of main signaling- and energy-generating pathways [40,41,42]. There is also a compelling evidence that PMCA4 is of paramount importance for neuronal survival in the conditions of Ca2+ overload as pheochromocytoma viability was preserved or impaired when this isoform was overexpressed or downregulated, respectively [43,44]. The sections below further explore the association and the specific role of the PMCA in neurodegenerative disorders.

3.1. PMCA in Neuropathology

3.1.1. PMCA in Aging

Contribution of PMCA to age-related neuropathologies was first suggested by Michaelis and coworkers [45,46,47]. These authors showed for the first time that PMCA activity and abundance in the synaptosomal membranes is reduced with age, similar to the PMCA activation by aged CaM. Zaidi and coworkers further demonstrated that the decline in PMCA activity was progressive with increasing biological age and was associated with lowered maximal velocity (Vmax) with no apparent changes in the affinity for Ca2+ [48]. The age-dependent alterations in PMCA are likely to be a consequence of oxidative stress as PMCA was identified to be a target for reactive oxygen/nitrogen species as does for CaM [49]. For instance, exposure of purified pump protein to H2O2 inhibited both basal and CaM-stimulated activity. However, neither CaM binding to the oxidized protein nor the concentration-dependent CaM effect on PMCA were affected, suggesting that C-terminal CaM binding domain is not primarily targeted by the oxidant. Pretreatment of PMCA with CaM almost completely preserved PMCA activity in the presence of H2O2 indicating that conformational state upon CaM binding may be more resistant to oxidation [50]. Several oxidative agents have been demonstrated to abolish PMCA sensitivity to CaM in a concentration-dependent manner [51], induce proteolytic degradation in synaptic membranes [52] or promote internalization and subsequent lack of detectable PMCA expression in hippocampal neurons [53].
PMCA activity is also affected by lipid-surrounding environment. It was demonstrated that PMCA and CaM are partitioned to the cholesterol-rich lipid rafts and PMCA activity in these membrane microdomains is higher than in non-raft regions [54]. Moreover, raft-localized PMCA is more sensitive to age-dependent loss of the activity [55]. Depletion of cholesterol drastically inhibited the activity of raft-associated PMCA but did not produce any effect on non-raft PMCA [54]. Therefore, increasing lipid order may be beneficial for protection of PMCA activity in the aged membranes but cannot overcome age-dependent loss of PMCA.

3.1.2. PMCA in Alzheimer’s Disease

Besides brain aging, altered PMCA expression and activity was detected postmortem in the brain cortex of patients affected by Alzheimer’s disease (AD) [56,57]. One of the histological hallmarks of this disease are the presence of a senile plaques of the amyloid β-peptide (Aβ) and accumulation of an abnormal tau protein [58]. Biochemical studies have revealed that Aβ decreased the activity of purified PMCA and the strongest inhibitory effect was seen for PMCA4 [57]. Moreover, cholesterol was shown to abolish the inhibitory effect of Aβ and the level of inhibition was lower in the lipid rafts of synaptosomal membranes than in non-rafts [57]. The Aβ inhibitory effect can be blocked by CaM, and the activity of PMCA lacking C-terminal CaM-binding domain was unaffected by Aβ. This antagonistic action of CaM is due to physical association with Aβ [59] or competing for PMCA binding [57]. Hence, CaM can protect PMCA activity by masking Aβ-PMCA interacting sites making them unavailable for Aβ. The accumulated data present a clear link between Ca2+, CaM and amyloid plaque formation indicating how the dysregulation in neuronal Ca2+ homeostasis and Aβ formation affect each other, and how CaM function in the center of this crosstalk. Interestingly, CaM content in the frontal, temporal, parietal cortex and subjacent white matter in AD was reduced by nearly 66% compared to the normal control brains [60]. The mechanism of progressive decline in PMCA and CaM in AD is unknown. However, the recent study on differentiated pheochromocytoma suggests that PMCA downregulation may be a trigger initiating calcineurin/NFAT-dependent repression of CalmII and CalmIII genes [40]. The altered expression of CaM is thus expected to deepen the physiological decline of PMCA function with increasing age and offer insufficient protection not only against Aβ, but also proteolytic and oxidative pump deactivation.
Growing body of evidence suggests that soluble forms of Aβ and tau cooperate with each other to drive healthy neurons into diseased state and the toxicity of Aβ requires tau [58]. The study of Berrocal and colleagues demonstrated that tau, which hyperphosphorylated form is predominant in neurofibrillary tangles, can directly interact with PMCA and inhibit its activity [56]. In this study, PMCA function was solely affected by this interaction as neither SERCA not SPCA were targeted. In differentiated pheochromocytoma, tau was concentrated in growth cones and interacted with PMCA through its N-terminal projection domain [61]. Overexpression of this amino-terminal fragment, but not full-length protein, suppressed nerve growth factor (NGF)-induced axonal outgrowth [61] establishing tau as a mediator of microtubule-plasma membrane interactions during neuritic development. Tau is also required for Fyn kinase-mediated NMDR receptor activation in the postsynaptic densities [62], which strengthen the interaction between NMDA receptor and postsynaptic density protein 95 (PSD-95) [63]. Considering the recruitment of PMCA via PSD-95 to a close proximity to NMDA receptor-mediated Ca2+ entry, a direct physical interaction between PMCA and tau in vivo would complement the contribution of PMCA-tau interaction to Ca2+ homeostasis dysregulation in the pathogenesis of AD.
Several early studies demonstrated that association between tau and CaM in vitro is Ca2+-dependent and it prevents tau interaction with microtubules [64,65]. Binding to CaM also prevents tau phosphorylation by PKC [66]. Although several CaM-dependent kinases and phosphatases are involved in tau posttranslational modification, for instance Ca2+/CaM-dependent protein kinase II (CaMKII), cyclin-dependent kinase 5 or protein phosphatase 2B (PP2B or calcineurin) [67], no recent studies have further expanded the functional significant of direct interaction with CaM and tau in the AD.

3.1.3. PMCA in Parkinson’s Disease

Altered PMCA function may significantly contribute to neuronal Ca2+ dyshomeostasis and increase the duration and frequency of intracellular Ca2+ spikes which may in turn influence the formation of pathological proteins such as the alpha synuclein in Parkinson’s disease (PD) [68]. This hypothesis is supported by the studies of Brendel and coworkers who demonstrated increased Ca2+ level and reduction in PMCA2 expression in primary midbrain neurons and neuroblastoma SH-SY5Y treated with Parkinsonian mimetic 1-methyl-4-phenylpyridinium (MPP) [69]. Interestingly, other Ca2+ efflux systems such as SERCA pump or Na+/Ca2+ exchanger remained largely unaffected. The same authors showed that PMCA2 knockdown with siRNA decreased the survival of mesencephalic neurons, but overexpression significantly increased the resistance of midbrain neurons to MPP toxicity [69]. These data indicate that PMCA2, which possesses the highest affinity for CaM, is particularly vulnerable to the inhibition by MPP. The mechanistic explanation of this phenomenon may lie in the oxidative stress and partial oxidative inactivation of PMCA, as membrane protein-selective antioxidants fully prevented MPP toxicity [70].
PMCA inactivation or even PMCA2 knockdown are known to irreversibly deprive neurons of substantial part of their Ca2+ clearing potency leading to dysregulation of Ca2+ homeostasis [39,68,71]. It is known that Ca2+ is a key controller of synuclein formation and Ca2+-dependent binding of CaM to α-synuclein accelerates formation of protein fibrils in vitro [72]. Recent study demonstrated that calcineurin also binds α-synuclein and this interaction is mediated by Ca2+/CaM signaling [73]. These and other authors demonstrated that increased calcineurin activity was associated with α-synuclein toxicity [73,74]. The activity of calcineurin is known to be regulated by the interaction with PMCA2 [75] and disruption of this interaction elevates intracellular activity of this phosphatase [76,77]. Although no direct interaction between α-synuclein and PMCA has been established, as does for SERCA pump [78], PMCA by regulating the activity of Ca2+-dependent signaling may potentially contribute to PD pathogenesis.

3.1.4. PMCA in Schizophrenia and Bipolar Disorder

Ca2+has been placed in the center of dopaminergic hypothesis of schizophrenia, mainly because of its essential role in dopamine receptors D1 and D2-mediated synaptic plasticity and signal transduction [79]. Therefore, it is not surprising that many of calcium signaling proteins, including PMCA and CaM, have been found to be differentially regulated in schizophrenia. In early study, Kluge and Kuhne demonstrated that kinetic properties of CaM-stimulated PMCA were altered in erythrocytes of patients with affective psychoses and hyper- or para-kinetic schizophrenics [80]. Unexpectedly, schizophrenia proteomic studies revealed PMCA4 to be up-regulated in anterior temporal lobe in affected patients [81] what may be seen as compensatory change to counterbalance elevated [Ca2+]c. On the other hand, proteins such as CaM (CALM1 and CALM2), CaMKII (CAMK2B, CAMK2D, CAMK2G) and CaM-like proteins were found to be downregulated in the brain or secretion fluids [81,82]. These findings limit, but not exclude, the special role of PMCA4 in neuronal signaling in schizophrenia, however the disadvantage of the proteomic studies is a fact they were performed not on isolated cells but on whole tissue. Therefore, it is likely that non-neuronal cells could contribute to revealed changes but because of high sample heterogeneity, these data should be interpreted with caution.
In the subgroup of patients involving mostly those with unipolar maniac and bipolar psychoses, the activation of human erythrocyte PMCA by CaM measured in the presence of sub-optimal concentration of lithium was stronger in all individuals with maniac-depressive episodes [80,83,84]. Interestingly, the activity measured in the conditions of optimal concentration of monovalent ions (Na+ and K+) was higher in lithium-treated groups than in control and untreated patients suggesting that the CaM-activated PMCA may be differentially regulated in maniac-depressive patients.
The potential involvement of PMCA in schizophrenia is also supported by data derived from pharmacological in vivo models. Recent study on animals challenged with 30 mg/kg ketamine, a drug that is known to mimic a wide spectrum of psychotomimetic and cognitive aberrations observed in schizophrenia in humans [85], demonstrated differential regulation of PMCA expression in functionally distinct brain regions [86]. Moreover, the basal and CaM-stimulated PMCA activities were reduced in the synaptosomal membranes mainly due to a direct interaction of the drug within large catalytic loop and competing with CaM for binding in the C-terminal domain of the pump [87,88]. Similarly, emerging studies also suggest the interaction of ketamine with CaM-dependent enzymes, in particular CaMKII [89,90], but no specific functional studies largely limit the conclusions on CaM and PMCA-mediated Ca2+ regulation in schizophrenia.

3.1.5. PMCA in Cerebellar Disorders

Immunohistochemical studies showed that neuron-specific PMCA2 and PMCA3 are more abundant in the cerebellum than PMCA1 and PMCA4. Moreover, they are concentrated in synaptic terminals of Purkinje cells while PMCA1 and PMCA4 localize mostly to granular layer [91]. Purkinje cells integrate the excitatory input to the cerebellar cortex being an essential component for the regulation and coordination of motor movements [92]. It is therefore not surprising that one of the most visible cerebellar dysfunctions during ataxia are uncoordinated movements and inability to maintain body balance. The link between ataxia and PMCA3 became evident when Zanni and coworkers identified a point mutations in ATP2B3 gene (located on the human chromosome X) in a family affected by X-linked congenital cerebellar ataxia [93]. The mutation located in exon 20 of ATP2B3/PMCA3 gene replaces a conserved Gly by Asp in C-terminal CaM-binding domain. The mutated pump demonstrated decreased ability to extrude Ca2+ what can be a direct consequence of altered interaction of its C-terminus with CaM. This is supported by mathematical modeling showing reduced ability of CaM to interact with the mutated binding domain what tends to depress the basal PMCA activity and decrease autoinhibitory interaction of CaM-binding domain with the main body of the pump. The possible defective interplay between mutated pump and PMCA-interacting signaling molecules (see Table 2) organized within Ca2+ nanodomains could contribute to the phenotype of cerebellar disease. It is even more plausible as more than a dozen X-linked gene defects have been demonstrated to contribute to cerebellar phenotype [94]. Among them, mutation in Ca2+/CaM-dependent serine protein kinase (CASK) may be significant [95,96] as CASK interacts with PMCA through PDZ domain located in the C-terminal end of the pump.
Cali and colleagues [97] identified another mutation in ATP2B3 gene in a patient with cerebellar ataxia and global developmental delay. The mutation (R482H) significantly reduced the Ca2+ clearing potency of the pump and resulted in the inability to handle intracellular Ca2+ transients evoked by cell stimulation. Interestingly, the patient also carried two additional mutations in the LAMA1 gene encoding laminin 1α. It has been well-characterized that mutations or deletions in this gene are associated with cerebellar dysplasia phenotype [98]. Therefore, on the basis of the family pedigree of the patient, it is reasonable to suspect that mutations in PMCA3 along with those in LAMA1 could synergistically contribute to the ataxic symptoms.
Recently, a novel mutation in PMCA3 (G733R) has been identified in a patient carrying a defect in phosphomannomutase 2 (PMM2), indicating a possible link between these mutations in generating ataxia phenotype [99]. PMM2 is an enzyme catalyzing the isomerization of mannose-6-phosphate to mannose1-phosphate and two missense mutations (R123Q and G214S) in PMM2 gene are known to be associated with type I congenital disorder of glycosylation [100,101]. The G733R substitution in the pump disturbed the ability to handle Ca2+ rises upon Ca2+ release from the ER or influx through the plasma membrane without affecting pump expression or subcellular targeting. The basal activity of autoinhibited pump and constitutive active variant of PMCA3 lacking C-terminal domain was also compromised. The coexistence of both PMCA3 and PMM2 mutations in the patient affected by non-progressive ataxia and muscular hypotonia is of special significance as PMM2 was found to be Ca2+-regulated enzyme [102]. Therefore, mutated PMCA3 may give rise to increased Ca2+ concentration in microdomains where PMM2 is located, inhibiting it. This, along with the mutations in PMM2, would deepen the decline in PMM2 enzymatic activity.
There is also convincing evidence coming from PMCA2 knockout mice that this isoform, apart from its function in auditory system discussed elsewhere [103,104,105], plays an essential role in cerebellar function. PMCA2−/− mice exhibited severe ataxia that became apparent by 12 days of age and had great difficulty in maintaining body balance [106]. Histological examination showed increased density of Purkinje neurons and reduced density of granular layer in cerebellum. Similarly, the deafwaddler (dfw) and wriggle mouse sagami (wri) strains displayed similar phenotypes such as tremor and vestibular/motor imbalance. Both dfw and wri genes were reported to be associated with a mutation in PMCA2 gene [107,108,109]. Interestingly, despite prolonged accumulation of Ca2+ in the cytosol of Purkinje neurons, heterozygous PMCA2+/− mice exhibited outwardly normal behavior but presented clear deficits in hindlimb coordination when challenged with exercise [110].
The first mutation in PMCA2 (V1143F) associated with congenital cerebellar ataxia has been recently identified by Vicario and coworkers [111]. In contrast to the hearing loss phenotype [105], the ataxia phenotype was generated without corresponding mutations in cadherin 23 and the hearing ability was fully retained. As V1143F substitution was located within CaM binding domain, it affected the interplay between mutated pump and CaM and the effect was particularly visible for full-length PMCA2 variant. Like other mutations, one of the consequences was prolonged duration of Ca2+ transients and compromised ability to maintain Ca2+ homeostasis in neurons. The list of PMCA mutations associated with cerebellar defects are summarized in Table 2.
Table 2. List of PMCA mutations and associated phenotype. Modified based on [112].
Table 2. List of PMCA mutations and associated phenotype. Modified based on [112].
SpeciesMutationPhenotypeReference
PMCA2MouseG283S (Dfw)Vestibular/motor imbalance[107]
MouseI655N (Elfin)Ataxia[113]
MouseS877F (Obv)Ataxia[114]
MouseE629K (Tmy)Ataxia[115]
MouseE412K (Wri)Abnormal movements[116]
HumanV1143FAtaxia[111]
PMCA3HumanG1107DAtaxia[93]
HumanR482HAtaxia[97]
RatR35CAtaxia[117]
HumanG773RAtaxia[99]
There is substantial body of evidence to suggest that CaM-regulated PMCA isoforms play an important role in neuronal survival and synaptic transmission, thus contributing to several pathological states of the CNS. However, at the present stage of development, the exact molecular mechanisms by which the defective PMCA function leads to generation of disease phenotype is still under investigation. Plausibly, the answer lies in the regulation of Ca2+ within discrete plasma membrane microdomains hosting PMCA-associated signaling molecules with CaM, due to its universality as a Ca2+ decoding molecule, being central to the regulation of neuronal signaling.

3.2. PMCA-Interacting Proteins in Mental Diseases

Besides the well-known regulation of PMCA by CaM, the pump is also affected in less well-studied ways by multiple interacting partners that have been associated with many of the neurodegenerative diseases (Table 3) [118]. These interactions are thought to target PMCA to highly-specialized membrane microdomains, regulate pump activity or recruit it to multiprotein complexes—“signalosomes” responsible for orchestration of local Ca2+ signaling. The dynamic and fidelity of these interactions determined by structural differences further augments functional specialization of particular PMCA isoforms. For instance, it has been demonstrated that PMCA2b and PMCA4b interact with postsynaptic density protein-95 (PSD-95, not indicated in the table but widely discussed elsewhere [119,120,121,122]), which tether the pump to microdomains enriched in NMDA receptor [123]. Alterations in NMDA receptor are, in turn, frequently reported in schizophrenia, mood disorders, Huntington’s disease (HD), AD and substance-induced psychosis [124]. The functional coupling between NMDA receptor and PMCA would allow rapid response to local Ca2+ rises due to bring the pump to a local Ca2+ entry sites. The formation of tertiary PMCA/PSD95/NMDA receptor complexes could modulate the amplitude of Ca2+ increases and affect the neurotransmission. Such modulation has been recently showed for glutamate in a rat model of ketamine-induced psychosis [87]. As PMCA plays a pivotal role in the regulation of local Ca2+ fluxes, it cannot be ruled out that interaction with the binding partners may affect their clustering into large signaling complexes and influence the neurosecretory process.

4. The Sarco/Endoplasmic Reticulum Ca2+-ATPase (SERCA)

The SERCA pump is the product of a multigene family. It is a 110-kDa single polypeptide located in the sarco/endoplasmic reticulum (SR/ER) which primary role is to transport Ca2+ back to the internal stores. Mammals express three isoforms of the pump, called SERCA1 through SERCA3, and post-translational modifications increase the total number of identified subtypes to 10 [140]. The expression profile of individual variants is not only tissue-dependent, but also undergoes changes during development. The distribution of SERCA1 is limited to fast- and slow-twitch skeletal muscles, and the role of this pump is to accumulate calcium in SR of skeletal muscle. The alternative splicing of SERCA1 gene generates two variants which expression pattern is developmentally regulated. SERCA1b is predominantly expressed in neonatal stages, then is entirely replaced by SERCA1a in adults [141]. SERCA2 exists in two variants—SERCA2a and SERCA 2b [141]. SERCA2a is mostly found in cardiac and skeletal muscle with slow contractile characteristics, but also exhibits minor expression in the brain, where it is almost exclusively restricted to the Purkinje neurons of the cerebellum [142]. SERCA2a is considered to be involved in contraction and relaxation of cardiac muscle. SERCA2b is the most abundant variant expressed widely in all tissue types, including neurons. In the brain, SERCA2b expression has been identified in both, cerebrum, and cerebellum [143]. Moreover, SERCA2b is the only SERCA variant present in astrocytes [144]. Further studies confirmed that both variants of SERCA2 pump are present in substantia nigra—the structure involved in dopamine release [145]. In humans, there are six possible variants of SERCA3 [141]. However, due to the complex alternative splicing, their functional and structure aspects remain poorly understood. SERCA3a-3f are located in most tissues, especially in secretory cells. Because of predominant presence of SERCA3 in pancreatic β-cells, this isoform is recognized as being involved in metabolic homeostasis [146].
Despite high structural homology, all isoforms of SERCA possess different affinity for calcium and unique kinetic properties. For instance, SERCA2a has a two-fold lower affinity for [Ca2+] and shows two-fold higher velocity of Ca2+ transport compared to ubiquitous SERCA2b [147]. This translates into primary role of SERCA2a in cardiac function, whereas SERCA2b is considered more as a house-keeping form. In turn, among all SERCA isoforms the lowest affinity for calcium and the slowest turnover rate for Ca2+ uptake can be ascribed to SERCA3 [148]. This feature is a consequence of the distribution of this isoform in non-excitable cells. Accumulating evidence suggests that SERCA2 is ubiquitously present in different brain areas and therefore may be the isoform of paramount importance for neuronal function [143,149,150,151].

4.1. SERCA Pumps in Neuropathology

Disturbances in ER Ca2+ homeostasis can lead, among others, to ER stress and accumulation of unfolded or misfolded proteins, which are detected for most neurodegenerative diseases, including AD, PD, HD, and ALS [152,153,154]. It has been reported that truncated isoform of human SERCA1 (S1T) triggered and amplified ER stress response, leading to apoptosis [155]. Furthermore, the increase of human S1T protein expression has been demonstrated in sporadic AD-derived post-mortem brains and in a cellular AD model, confirming that S1T can induce neuroinflammatory response in vitro and in vivo [156].
An interesting observation was identification of several ATP2A2 mutations in autosomal dominant skin disorder—Darier’s disease (DD), a disorder frequently associated with several mental diseases (bipolar disorder, schizophrenia, affective psychosis, epilepsy) [157,158,159,160,161]. Numerous SERCA2 mutations in DD have been detected, including missense and nonsense types, which produced the insoluble truncated, misfolded and/or aggregated proteins, finally decreasing the amount of fully active SERCA. All these results indicate that ATP2A2 mutations may have pleiotropic effects on the brain as well as skin.

4.2. Calmodulin-Controlled Regulation of SERCA Pumps

Although the direct regulation by CaM has been shown only for PMCA, a growing body of evidence indicates that CaM can significantly, but indirectly, participate in modulation of SERCA activity. The well-recognized mechanism exists in cardiac and skeletal muscles and is based on regulation of SERCA activity by endogenous molecule—phospholamban (PLN) [162,163]. This small, 52-amino-acid transmembrane protein, which is expressed almost exclusively in muscle cells, in non-phosphorylated form binds SERCA2a, the muscles-specific variant, and lowers its affinity for Ca2+, thus attenuating transport rate by ~50% [164]. It was shown that PLN can inhibit SERCA2a and SERCA2b isoforms to the same extent [165]. Phosphorylation at Ser16 by PKA or by Ca2+-CaMKII at Thr17 causes PLN dissociation from SERCA, thereby removes its inhibitory effect, and increases Ca2+ uptake to the SR [166]. The decrease in the phosphorylation level of PLN by protein phosphatases PP1 and/or PP2B restores interaction between SERCA2a and PLN, inhibiting pump activity [167]. Although up to now the presence of PLN protein has not been confirmed in neurons, its expression was demonstrated in astrocytes what may have profound implications for the function of CNS.
Astrocytes, the important elements of glia, are an integral part of the CNS that couple the activity of neurons and blood-brain barrier (BBB). The proper function of BBB requires cooperation between endothelium, astrocytes, neurons, and extracellular matrix. An important role in a progression of neurological diseases is a disruption of the BBB integrity [168,169]. Degeneration of BBB has been confirmed in different disease—AD, PD, ALS, multiple sclerosis (MS), vascular dementia, stroke, hypoxia, or ischemia [170]. It leads to altered permeability with subsequent infiltration of serum components what can trigger abnormal intrinsic signaling pathways, including those involving calcium and/or CaM. Under physiological conditions astrocytes secrete neurotrophic factors, growth factors and cytokines that regulate neurogenesis, synaptogenesis, neuromodulation, and neuronal survival, but abnormal function of astrocytes has been observed in many neurodegenerative diseases [171,172,173]. For example, accumulation of α-synuclein detected in PD astrocytes clearly indicates their critical role in the course of disease [174]. It has been recently demonstrated that in α-synuclein-treated brain astrocytes, a leucine-rich repeat kinase 2 mutant G2019S (LRRK2-GS) can act through SERCA inactivation triggering ER stress [175]. LRRK2 is a multifunctional protein kinase, localized in the cytoplasm and associated with cellular membrane structures, that contains many domains capable of protein–protein interactions [176]. The GS mutation in the kinase domain of LRRK2 is one of the most common genetic basis of PD [177,178,179]. SERCA directly interacts with LRRK2-GS and, after translocation to the ER, forms persistently inactive SERCA–PLN complex. Disruption of SERCA function causes Ca2+ depletion from ER, finally leading to the cell death.

5. Secretory Pathway Ca2+-ATPase (SPCA)—The Golgi-Resident Ca2+/Mn2+ Pump

SPCA pump, as a member of the P2A subfamily, shares some common structural and mechanistic properties of SERCA [180], yet in addition to high Ca2+ affinity, SPCA also represents strong preference for Mn2+ ions. This Ca2+/Mn2+ transporter possesses unique structural elements in the N-terminus and transmembrane (TM) region determining orientation and selectivity of the ion transport during phosphoryl-transfer reactions. Particularly, SPCA pump is crucial for maintaining the sufficient supply of Mn2+ for glycosyltransferases and sulfotransferases in the Golgi lumen [181]. On the other hand, in the cytosol, SPCA pump prevents excessive accumulation of Ca2+ and Mn2+, while the overload of these ions may trigger neurotoxicity resulting in several neurological disorders [182].
In human, two SPCA isoforms, SPCA1 and SPCA2 are encoded by ATP2C1 and ATP2C2 genes, respectively. These isoforms share nearly 60% of sequence identity and they exhibit distinct expression pattern and tissue distributions in the human body [183]. Although SPCA1 is ubiquitously presented in the Golgi membranes of almost all mammalian tissues [184], it displays predominant localization in the brain where it helps to maintain appropriate physiological features of neurons including reception, conduction, and transmission of signals [185,186]. For SPCA2, its expression is more restricted to respiratory, gastrointestinal, genitourinary systems, as well as salivary and mammary glands [187,188]. The abundance and the specific role of SPCA2 in the brain is still a contentious issue, however its subcellular localization was confined mainly to the Golgi apparatus (GA) and the secretory vesicles [181,187]. Through alternative splicing at the 3′-end of human ATP2C1 pre-mRNA, four splice variants are generated, leading to additional SPCA1a-d isoforms with different length and sequence at the COOH-terminal domain. However, functional differences among these mature variants still remain largely unexplored [189].
SPCA pump is presumed to be composed of ten membrane-spanning helices (M1-M10) and cytosolic headpiece containing conservative motifs, which are critical for transport functions in a SERCA-like manner, including the Thr–Gly–Glu loop in A site, the phosphate-accepting Asp residue, the ATP- and FITC-binding region, and the Asp–Pro–Pro–Arg loop between the N and P sites. Unlike SERCA, SPCA pump lacks long cytoplasmic tail and elongated luminal loops linking some of the transmembrane domains, and importantly, only one calcium-binding region in SPCA protein overlaps with site II in the transmembrane domains of SERCA [183]. The research on the PMR1 yeast enzyme has shown that interface between Gln-783 and Val-335 in M6 and M4 domains can be critical for Mn2+ transport by SPCA pump [180]. Human SPCA1a/2 isoforms exhibit comparable Mn2+ affinity while SPCA1a displays significantly lower apparent Ca2+ affinity than SPCA2. Remarkably, SPCA1a displays a two-fold higher maximal ATPase activity in the presence of Ca2+ as compared to Mn2+ conditions, whereas SPCA2 shows similar maximal turnover rates for both ions. These differences in biochemical properties of SPCA isoforms seem to be determined, at least in part, by EF hand-like motif that is present in SPCA1a N-terminus, but absent in SPCA2. It is believed that such motif influences lower affinity and higher Ca2+ capacity of this pump promoting stronger Ca2+-dependent autophosphorylation. This feature may have physiological implications in cells with a high calcium load and/or fulfilling secretory functions [190].

5.1. SPCA Pumps in Neuropathology

The GA is associated with post-translational processing of proteins destined for secretion as well as their incorporation into the ER, lysosomes or plasma membranes. In the GA, calcium can reach ~300 μM, due to the action of SERCA and SPCA localized in this organelle [191,192]. Stored Ca2+ can be released mainly by IP3 receptors following extracellular stimuli. It is now widely accepted that GA, together with ER and mitochondria, plays an important role in the regulation of cytosolic Ca2+. It has been shown that the maintenance of the Ca2+ content provided by SPCA1 appears essential for the correct structure of the entire GA and for important functions of the secretory pathway.
Early study on SPCA1 indicated its vital role during brain development and closure of neural tube [185,193]. Additionally, downregulation of SPCA1 disrupted the proper processes of neuronal growth and differentiation, leading to altered GA morphology (like its fragmentation), as well as slowed down protein transport in the Golgi compartments [194]. In neural tissue SPCA1 exhibited an important role in the control of cytoskeletal dynamics in mice neuroepithelial cells and perturbation of calcium homeostasis impaired apical constriction during neural tube closure [195]. Since the GA is an important platform for a number of signaling cascades, inactivation of SPCA1 can also induce the disturbances in mitochondrial structure and metabolism, increasing their sensitivity to stress conditions [196]. Thereby, the abnormal function of the GA in several neuropathologies can be initiated by one or more aforementioned mechanisms. Moreover, GA in neurons is suggested to participate in modification of calcium signaling in some diseases, including AD, HD, or ALS [197].
A unique SPCA1 expression pattern has been described during ischemic/reperfusion brain injury (IRI), which was affected by pre-ischemic challenge. Whereas IRI induced the depression of SPCA-mediated calcium transport in hippocampus, ischemic preconditioning (IPC) had a partial protective effect on SPCA activity [198]. Similar relationship between the expression of SPCA1 and calcium concentration in neuronal cytoplasm and GA was observed during cerebral ischemia and reperfusion [199]. In naive ischemia, SPCA1 declined at early reperfusion, but increased in late reperfusion [200]. In rat brain cortex and hippocampus down-regulation of SPCA2 due to IRI has also been shown [201]. Based on these findings, it can be concluded that both SPCA isoforms may play a vital role in the GA stress during brain IRI, showing dual type of action. In preconditioned rat brain, SPCA may reduce calcium overload and the level of oxidative stress, but on the other hand, SPCA level was downregulated after prolonged ischemia.
Besides Ca2+ transport, SPCA also controls manganese homeostasis, which is important for brain development. Moreover, it represents the only known system for cellular Mn2+ detoxification. In rat brain, SPCA1 was upregulated following Mn2+ exposure, but impaired regulation of Mn2+ transport may trigger neuronal pathology. Although SPCA2 shows more restricted distribution, its relatively high level in brain could be more important taking into account the correlations between manganese neurotoxicity and PD [202,203]. High extracellular Mn2+ concentration exerted the toxic, inhibitory effect on SPCA activity in cultured mice neurons and glia [204]. In addition, since Mn2+ and Ca2+ can occupy the same ion-transport site, Mn2+ toxicity may also affect Ca2+ homeostasis. At elevated levels, Mn2+ can also compete with magnesium-binding sites in many functional proteins, enhance oxidative stress or lead to accumulation of intracellular toxic metabolites [205]. Finally, disturbances in ions homeostasis can induce ER and mitochondrial dysfunction, leading to neuronal and/or glial apoptosis. Although glial cells seem to be more resistant to Mn2+ poisoning than neurons, prolonged dyshomeostasis may result in fragmentation of the GA. The role of Mn2+ in the etiology of neurodegenerative diseases has been widely documented. Excess Mn2+ was shown to induce a neurological disorder with symptoms similar to PD [206]. Moreover, Mn2+ contributed to AD associated with impaired cognitive function and cognitive decline [207].

5.2. Calmodulin-Controlled Regulation of Calcium in Golgi Apparatus

Changes in the GA morphology observed in a number of neuronal diseases such as AD, ALS, and stroke include GA fragmentation and intracellular Ca2+ overload, which can lead to additional injury. These processes are also associated with specific brain areas, as well as with BBB dysfunction. Mechanisms underlying the neuropathological changes are observed in different cell types and frequently result in improper communication between brain structures, including cooperation between neurons and glial cells [208]. Astrocytes, which represent the prevalent glial cell types in mammalian brain, are responsible for maintenance of neurotransmitter and ion balance [209]. The active, bidirectional regulation of synapses by astrocytes has been shown to influence neuronal and synaptic functions accompanied by the altered Ca2+ circulation. Furthermore, astrocytes are an integral part of the BBB [210]. Interestingly, many transport systems responsible for balancing cytosolic Ca2+ concentration are under feedback control of Ca2+/CaM complex [211,212,213]. Up to now, the direct regulation of SPCA by CaM has not be detected, but there are several indirect CaM-induced processes that can affect the GA function. One of them is linked with production of nitric oxide (NO). The initial release of calcium forms a Ca2+/CaM complex, which binds to nitric oxide synthase (NOS), and actives the enzyme producing NO from L-arginine. The most important target for NO is guanylate cyclase (GC). High level of NO can contribute to excitotoxicity following a stroke and neurodegenerative diseases. In addition, NOS generates superoxide, which is involved in both cell injury and signaling [214].
It has been demonstrated that Ca2+ released after the GA fragmentation caused by pathological conditions triggers overactivation of eNOS in a Ca2+/CaM-dependent way. Moreover, increased Mn2+ levels in astrocytes elevated the expression of iNOS and activated soluble guanylate cyclase (sGC), which is suggested to be a causative factor for PD [215]. sGC expression and activity appear to be higher in the striatum than in any other brain regions [216]. The second interesting regulation by Ca2+/CaM comprises the activation of CaMKII, the enzyme controlling multiple signaling pathways in the brain. It should be noted that the effects of CaMKII may be reversed by antagonistic action of protein phosphatases, particularly PP2B, which is also a Ca2+/CaM-dependent enzyme. Phosphorylation of nNOS at Ser847 by CaMKII was shown to decrease NO generation, but increase superoxide generation [217]. Subsequently, reactive oxygen species and reactive nitrogen species may damage intracellular membranes of ER, GA and mitochondria what results in the leakage of Ca 2+ into the cytoplasm. Nitrosative/oxidative stress may also induce BBB rupture and impair metabolic network of astrocytes [218].
Ca2+ in the ER and the GA is mainly released through the inositol-1,4,5-trisphosphate receptor (IP3R) and ryanodine receptor (RYR) channels, which are also regulated by Ca2+/CaM complex [219,220]. The role of IP3Rs has been shown in ataxia and neurodegenerative diseases such as AD and HD [221]. Recent studies also suggest that alterations in the expression and function of RyRs are related to neurodegenerative diseases such as AD [222].
The functionality of GA appears to be more complex, especially in its ability to counterbalance excessive Ca2+ and/or Mn2+ concentration. Participation of several ion transporting systems and cooperation of intracellular organelles is required for adaptation to stress conditions that allows protection against neurotoxicity. The importance of cooperation between GA, ER, and mitochondria in maintenance of calcium homeostasis has been suggested in many studies. Under pathological conditions improper control of Ca2+ circulation, its excessive accumulation and disturbed mitochondrial energy production preclude the formation of physiological Ca2+ signaling networks leading to neurotoxicity and cell death.

6. Concluding Remarks

There is now substantial evidence that defective Ca2+ signaling is frequently observed in majority of mental and neurodegenerative diseases. Because one of the most distinguishable features of Ca2+ is ambivalence, the dysfunctions of neuronal Ca2+-ATPases have recently acquired growing awareness and prominence. So far, no massive and uncontrolled neuronal death due to pumps dysfunctions have been identified. Instead, they are usually linked to less severe phenotypic changes that possibly originate from mild alterations in Ca2+ concentration in subtle cellular compartments or domains. It should be underlined that at different steps of subsequent Ca2+-induced processes, the Ca2+/CaM complex plays a decisive control function. Up to now, the studies revealed only certain human diseases, mostly with genetic defects in Ca2+ pumps. These phenotypes have gained considerable attention, however an intensified research on animal models is soon expected to describe the global effects of nongenetic pump alterations as well. Undoubtedly, knowing the consequences of Ca2+ pumps dysfunctions will help to understand the importance of Ca2+ signaling in development of neuropathologies. In view of that, future work should reveal the relationship between PMCA, SERCA and SPCA and their integration into dynamic processes of neuronal death, adaptation, and repair. The special attention should be given to CaM-regulatory mechanisms as primarily responsible for neuronal signaling in physiological and pathological states. This may significantly contribute to the identification of therapeutic strategies centered on neuronal calcium pumps.

Author Contributions

T.B., M.S., J.M., M.L., B.F., F.G. and L.Z. individually contributed to manuscript writing, tables and figures preparation, manuscript check and preparation of its final version. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by National Science Centre grant no. 2019/33/B/NZ4/00587 and by Medical University of Lodz grant no. 503/6-086-02/503-61-001.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Brini, M.; Carafoli, E. Calcium Pumps in Health and Disease. Physiol. Rev. 2009, 89, 1341–1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Hu, Z.; Bonifas, J.M.; Beech, J.; Bench, G.; Shigihara, T.; Ogawa, H.; Ikeda, S.; Mauro, T.M.; Epstein, E.H., Jr. Mutations in ATP2C1, encoding a calcium pump, cause Hailey-Hailey disease. Nat. Genet. 2000, 24, 61–65. [Google Scholar] [CrossRef] [PubMed]
  3. Sudbrak, R.; Brown, J.M.; Dobson-Stone, C.; Carter, S.A.; Ramser, J.; White, J.; Healy, E.; Dissanayake, M.; Larrègue, M.; Perrussel, M.; et al. Hailey-Hailey disease is caused by mutations in ATP2C1 encoding a novel Ca2+ pump. Hum. Mol. Genet. 2000, 9, 1131–1140. [Google Scholar] [CrossRef] [PubMed]
  4. Strehler, E.E.; Zacharias, D.A. Role of Alternative Splicing in Generating Isoform Diversity Among Plasma Membrane Calcium Pumps. Physiol. Rev. 2001, 81, 21–50. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Gong, D.; Chi, X.; Ren, K.; Huang, G.; Zhou, G.; Yan, N.; Lei, J.; Zhou, Q. Structure of the human plasma membrane Ca. Nat. Commun. 2018, 9, 3623. [Google Scholar] [CrossRef]
  6. Toyoshima, C.; Iwasawa, S.; Ogawa, H.; Hirata, A.; Tsueda, J.; Inesi, G. Crystal structures of the calcium pump and sarcolipin in the Mg2+-bound E1 state. Nat. Cell Biol. 2013, 495, 260–264. [Google Scholar] [CrossRef]
  7. Friedberg, F.; Rhoads, A.R.; Friedberg, A.R. Evolutionary Aspects of Calmodulin. IUBMB Life 2001, 51, 215–221. [Google Scholar] [CrossRef]
  8. Hoeflich, K.P.; Ikura, M. Calmodulin in action: Diversity in target recognition and activation mechanisms. Cell 2002, 108, 739–742. [Google Scholar] [CrossRef] [Green Version]
  9. Vetter, S.W.; Leclerc, E. Novel aspects of calmodulin target recognition and activation. JBIC J. Biol. Inorg. Chem. 2003, 270, 404–414. [Google Scholar] [CrossRef]
  10. Jiang, X.; Lautermilch, N.J.; Watari, H.; Westenbroek, R.E.; Scheuer, T.; Catterall, W.A. Modulation of CaV2.1 channels by Ca2+/calmodulin-dependent protein kinase II bound to the C-terminal domain. Proc. Natl. Acad. Sci. USA 2007, 105, 341–346. [Google Scholar] [CrossRef] [Green Version]
  11. Kortvely, E.; Gulya, K. Calmodulin, and various ways to regulate its activity. Life Sci. 2004, 74, 1065–1070. [Google Scholar] [CrossRef] [PubMed]
  12. Villalobo, A. The multifunctional role of phospho-calmodulin in pathophysiological processes. Biochem. J. 2018, 475, 4011–4023. [Google Scholar] [CrossRef] [Green Version]
  13. Cobb, J.A.; Roberts, D.M. Structural Requirements for N-Trimethylation of Lysine 115 of Calmodulin. J. Biol. Chem. 2000, 275, 18969–18975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Hofmann, N.R. Calmodulin methylation: Another layer of regulation in calcium signaling. Plant Cell 2013, 25, 4284. [Google Scholar] [CrossRef] [Green Version]
  15. Thulin, E.; Andersson, A.; Drakenberg, T.; Forsén, S.; Vogel, H.J. Metal ion and drug binding to proteolytic fragments of calmodulin: Proteolytic cadmium-113 and proton nuclear magnetic resonance studies. Biochem. J. 1984, 23, 1862–1870. [Google Scholar] [CrossRef] [PubMed]
  16. Chin, D.; Means, A.R. Calmodulin: A prototypical calcium sensor. Trends Cell Biol. 2000, 10, 322–328. [Google Scholar] [CrossRef]
  17. Grabarek, Z. Structural Basis for Diversity of the EF-hand Calcium-binding Proteins. J. Mol. Biol. 2006, 359, 509–525. [Google Scholar] [CrossRef]
  18. Gifford, J.L.; Walsh, M.P.; Vogel, H.J. Structures and metal-ion-binding properties of the Ca2+-binding helix–loop–helix EF-hand motifs. Biochem. J. 2007, 405, 199–221. [Google Scholar] [CrossRef]
  19. Kawasaki, H.; Soma, N.; Kretsinger, R.H. Molecular Dynamics Study of the Changes in Conformation of Calmodulin with Calcium Binding and/or Target Recognition. Sci. Rep. 2019, 9, 10688. [Google Scholar] [CrossRef] [PubMed]
  20. Jensen, H.H.; Brohus, M.; Nyegaard, M.; Overgaard, M.T. Human Calmodulin Mutations. Front. Mol. Neurosci. 2018, 11, 396. [Google Scholar] [CrossRef] [Green Version]
  21. Burgoyne, R.D. Neuronal calcium sensor proteins: Generating diversity in neuronal Ca2+ signalling. Nat. Rev. Neurosci. 2007, 8, 182–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Tidow, H.; Nissen, P. Structural diversity of calmodulin binding to its target sites. FEBS J. 2013, 280, 5551–5565. [Google Scholar] [CrossRef] [PubMed]
  23. Mruk, K.; Farley, B.M.; Ritacco, A.W.; Kobertz, W.R. Calmodulation meta-analysis: Predicting calmodulin binding via canonical motif clustering. J. Gen. Physiol. 2014, 144, 105–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Bähler, M.; Rhoads, A. Calmodulin signaling via the IQ motif. FEBS Lett. 2001, 513, 107–113. [Google Scholar] [CrossRef]
  25. Biber, A.; Schmid, G.; Hempel, K. Calmodulin content in specific brain areas. Exp. Brain Res. 1984, 56, 323–326. [Google Scholar] [CrossRef]
  26. Calì, T.; Brini, M.; Carafoli, E. Regulation of Cell Calcium and Role of Plasma Membrane Calcium ATPases. Int. Rev. Cell Mol. Biol. 2017, 332, 259–296. [Google Scholar] [CrossRef]
  27. Schatzmann, H.J. ATP-dependent Ca+-Extrusion from human red cells. Cell. Mol. Life Sci. 1966, 22, 364–365. [Google Scholar] [CrossRef]
  28. Boczek, T.; Radzik, T.; Ferenc, B.; Zylinska, L. The Puzzling Role of Neuron-Specific PMCA Isoforms in the Aging Process. Int. J. Mol. Sci. 2019, 20, 6338. [Google Scholar] [CrossRef] [Green Version]
  29. Guerini, D.; García-Martin, E.; Gerber, A.; Volbracht, C.; Leist, M.; Merino, C.G.; Carafoli, E. The Expression of Plasma Membrane Ca2+ Pump Isoforms in Cerebellar Granule Neurons Is Modulated by Ca2+. J. Biol. Chem. 1999, 274, 1667–1676. [Google Scholar] [CrossRef] [Green Version]
  30. Padányi, R.; Pászty, K.; Hegedűs, L.; Varga, K.; Papp, B.; Penniston, J.T.; Enyedi, Á. Multifaceted plasma membrane Ca2+ pumps: From structure to intracellular Ca2+ handling and cancer. Biochim. Biophys. Acta BBA Bioenerg. 2016, 1863, 1351–1363. [Google Scholar] [CrossRef] [Green Version]
  31. Monteith, G.R.; Wanigasekara, Y.; Roufogalis, B.D. The plasma membrane calcium pump, its role and regulation: New complexities and possibilities. J. Pharmacol. Toxicol. Methods 1998, 40, 183–190. [Google Scholar] [CrossRef]
  32. Falchetto, R.; Vorherr, T.; Brunner, J.; Carafoli, E. The plasma membrane Ca2+ pump contains a site that interacts with its calmodulin-binding domain. J. Biol. Chem. 1991, 266, 2930–2936. [Google Scholar] [CrossRef]
  33. Burette, A.; Rockwood, J.M.; Strehler, E.E.; Weinberg, R.J. Isoform-specific distribution of the plasma membrane Ca2+ ATPase in the rat brain. J. Comp. Neurol. 2003, 467, 464–476. [Google Scholar] [CrossRef] [PubMed]
  34. Eakin, T.J.; Antonelli, M.C.; Malchiodi, E.L.; Baskin, D.G.; Stahl, W.L. Localization of the plasma membrane Ca2+-ATPase isoform PMCA3 in rat cerebellum, choroid plexus and hippocampus. Mol. Brain Res. 1995, 29, 71–80. [Google Scholar] [CrossRef]
  35. Zacharias, D.; Kappen, C. Developmental expression of the four plasma membrane calcium ATPase (Pmca) genes in the mouse. Biochim. Biophys. Acta BBA Gen. Subj. 1999, 1428, 397–405. [Google Scholar] [CrossRef]
  36. Brini, M.; Calì, T.; Ottolini, D.; Carafoli, E. Neuronal calcium signaling: Function and dysfunction. Cell. Mol. Life Sci. 2014, 71, 2787–2814. [Google Scholar] [CrossRef]
  37. Brandt, P.C.; Sisken, J.E.; Neve, R.L.; Vanaman, T.C. Blockade of plasma membrane calcium pumping ATPase isoform I impairs nerve growth factor-induced neurite extension in pheochromocytoma cells. Proc. Natl. Acad. Sci. USA 1996, 93, 13843–13848. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Szemraj, J.; Kawecka, I.; Bartkowiak, J.; Zylińska, L. The effect of antisense oligonucleotide treatment of plasma membrane Ca(+2)-ATPase in PC12 cells. Cell. Mol. Biol. Lett. 2004, 9, 451–464. [Google Scholar]
  39. Boczek, T.; Lisek, M.; Kowalski, A.; Pikula, S.; Niewiarowska, J.; Wiktorska, M.; Zylinska, L. Downregulation of PMCA2 or PMCA3 reorganizes Ca2+ handling systems in differentiating PC12 cells. Cell Calcium 2012, 52, 433–444. [Google Scholar] [CrossRef]
  40. Boczek, T.; Lisek, M.; Ferenc, B.; Zylinska, L. Cross talk among PMCA, calcineurin and NFAT transcription factors in control of calmodulin gene expression in differentiating PC12 cells. Biochim. Biophys. Acta BBA Bioenerg. 2017, 1860, 502–515. [Google Scholar] [CrossRef]
  41. Boczek, T.; Lisek, M.; Ferenc, B.; Kowalski, A.; Wiktorska, M.; Zylinska, L. Silencing of Plasma Membrane Ca2+-ATPase Isoforms 2 and 3 Impairs Energy Metabolism in Differentiating PC12 Cells. BioMed Res. Int. 2014, 2014, 735106. [Google Scholar] [CrossRef] [PubMed]
  42. Boczek, T.; Lisek, M.; Ferenc, B.; Kowalski, A.; Stepinski, D.; Wiktorska, M.; Zylinska, L. Plasma Membrane Ca2+-ATPase Isoforms Composition Regulates Cellular pH Homeostasis in Differentiating PC12 Cells in a Manner Dependent on Cytosolic Ca2+ Elevations. PLoS ONE 2014, 9, e102352. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Garcia, M.L.; Usachev, Y.M.; Thayer, S.A.; Strehler, E.E.; Windebank, A.J. Plasma membrane calcium ATPase plays a role in reducing Ca2+-mediated cytotoxicity in PC12 cells. J. Neurosci. Res. 2001, 64, 661–669. [Google Scholar] [CrossRef] [PubMed]
  44. Usachev, Y.M.; Demarco, S.J.; Campbell, C.; Strehler, E.E.; Thayer, S.A. Bradykinin and ATP Accelerate Ca2+ Efflux from Rat Sensory Neurons via Protein Kinase C and the Plasma Membrane Ca2+ Pump Isoform 4. Neuron 2002, 33, 113–122. [Google Scholar] [CrossRef] [Green Version]
  45. Michaelis, M.; Johe, K.; Kitos, T. Age-dependent alterations in synaptic membrane systems for Ca2+ regulation. Mech. Ageing Dev. 1984, 25, 215–225. [Google Scholar] [CrossRef]
  46. Michaelis, M.; Bigelow, D.; Schöneich, C.; Williams, T.; Ramonda, L.; Yin, D.; Hühmer, A.; Yao, Y.; Gao, J.; Squier, T. Decreased plasma membrane calcium transport activity in aging brain. Life Sci. 1996, 59, 405–412. [Google Scholar] [CrossRef]
  47. Michaelis, M.L. Ca2+Handling Systems and Neuronal Aging. Ann. N. Y. Acad. Sci. 1989, 568, 89–94. [Google Scholar] [CrossRef]
  48. Zaidi, A.; Gao, J.; Squier, T.C.; Michaelis, M.L. Age-related decrease in brain synaptic membrane Ca2+-ATPase in F344/BNF1 rats. Neurobiol. Aging 1998, 19, 487–495. [Google Scholar] [CrossRef]
  49. McCarthy, M.R.; Thompson, A.R.; Nitu, F.; Moen, R.J.; Olenek, M.J.; Klein, J.C.; Thomas, D.D. Impact of methionine oxidation on calmodulin structural dynamics. Biochem. Biophys. Res. Commun. 2015, 456, 567–572. [Google Scholar] [CrossRef] [Green Version]
  50. Zaidi, A.; Barŕon, L.; Sharov, V.S.; Schöneich, C.; Michaelis, E.K.; Michaelis, M.L. Oxidative Inactivation of Purified Plasma Membrane Ca2+-ATPase by Hydrogen Peroxide and Protection by Calmodulin. Biochemistry 2003, 42, 12001–12010. [Google Scholar] [CrossRef]
  51. Zaidi, A.; Fernandes, D.; Bean, J.L.; Michaelis, M.L. Effects of paraquat-induced oxidative stress on the neuronal plasma membrane Ca2+-ATPase. Free. Radic. Biol. Med. 2009, 47, 1507–1514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Zaidi, A.; Michaelis, M.L. Effects of reactive oxygen species on brain synaptic plasma membrane Ca(2+)-ATPase. Free. Radic. Biol. Med. 1999, 27, 810–821. [Google Scholar] [CrossRef]
  53. Kip, S.N.; Strehler, E.E. Rapid Downregulation of NCX and PMCA in Hippocampal Neurons Following H2O2 Oxidative Stress. Ann. N. Y. Acad. Sci. 2007, 1099, 436–439. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Jiang, L.; Fernandes, D.; Mehta, N.; Bean, J.L.; Michaelis, M.L.; Zaidi, A. Partitioning of the plasma membrane Ca2+-ATPase into lipid rafts in primary neurons: Effects of cholesterol depletion. J. Neurochem. 2007, 102, 378–388. [Google Scholar] [CrossRef] [PubMed]
  55. Jiang, L.; Bechtel, M.D.; Galeva, N.A.; Williams, T.D.; Michaelis, E.K.; Michaelis, M.L. Decreases in plasma membrane Ca2+-ATPase in brain synaptic membrane rafts from aged rats. J. Neurochem. 2012, 123, 689–699. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Berrocal, M.; Corbacho, I.; Vázquez-Hernández, M.; Ávila, J.; Sepúlveda, M.R.; Mata, A.M. Inhibition of PMCA activity by tau as a function of aging and Alzheimer’s neuropathology. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2015, 1852, 1465–1476. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Berrocal, M.; Marcos, D.; Sepúlveda, M.R.; Pérez, M.; Ávila, J.; Mata, A.M. Altered Ca2+ dependence of synaptosomal plasma membrane Ca2+ -ATPase in human brain affected by Alzheimer’s disease. FASEB J. 2009, 23, 1826–1834. [Google Scholar] [CrossRef]
  58. Bloom, G.S. Amyloid-β and tau: The trigger and bullet in Alzheimer disease pathogenesis. JAMA Neurol. 2014, 71, 505–508. [Google Scholar] [CrossRef] [Green Version]
  59. Corbacho, I.; Berrocal, M.; Török, K.; Mata, A.M.; Gutierrez-Merino, C. High affinity binding of amyloid β-peptide to calmodulin: Structural and functional implications. Biochem. Biophys. Res. Commun. 2017, 486, 992–997. [Google Scholar] [CrossRef] [Green Version]
  60. McLachlan, D.R.C.; Wong, L.; Bergeron, C.; Baimbridge, K.G. Calmodulin and calbindin D28K in Alzheimer disease. Alzheimer Dis. Assoc. Disord. 1987, 1, 171–179. [Google Scholar] [CrossRef]
  61. Brandt, R.; Léger, J.; Lee, G. Interaction of tau with the neural plasma membrane mediated by tau’s amino-terminal projection domain. J. Cell Biol. 1995, 131, 1327–1340. [Google Scholar] [CrossRef] [PubMed]
  62. Ittner, L.M.; Ke, Y.D.; Delerue, F.; Bi, M.; Gladbach, A.; Van Eersel, J.; Wölfing, H.; Chieng, B.C.; Christie, M.J.; Napier, I.A.; et al. Dendritic Function of Tau Mediates Amyloid-β Toxicity in Alzheimer’s Disease Mouse Models. Cell 2010, 142, 387–397. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Rong, Y.; Lu, X.; Bernard, A.; Khrestchatisky, M.; Baudry, M. Tyrosine phosphorylation of ionotropic glutamate receptors by Fyn or Src differentially modulates their susceptibility to calpain and enhances their binding to spectrin and PSD-95. J. Neurochem. 2008, 79, 382–390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Padilla, R.; Maccioni, R.; Avila, J. Calmodulin binds to a tubulin binding site of the microtubule-associated protein tau. Mol. Cell. Biochem. 1990, 97, 35–41. [Google Scholar] [CrossRef]
  65. Lee, Y.C.; Wolff, J. Calmodulin binds to both microtubule-associated protein 2 and tau proteins. J. Biol. Chem. 1984, 259, 1226–1230. [Google Scholar] [CrossRef]
  66. Baudier, J.; Mochly-Rosen, D.; Newton, A.; Lee, S.H.; Koshland, D.E.; Cole, R.D. Comparison of S100b protein with calmodulin: Interactions with melittin and microtubule-associated. tau. proteins and inhibition of phosphorylation of. tau. proteins by protein kinase C. Biochem. J. 1987, 26, 2886–2893. [Google Scholar] [CrossRef]
  67. O’Day, D.H.; Eshak, K.; Myre, M.A. Calmodulin Binding Proteins and Alzheimer’s Disease. J. Alzheimers Dis. 2015, 46, 553–569. [Google Scholar] [CrossRef] [Green Version]
  68. Zaidi, A. Plasma membrane Ca2+-ATPases: Targets of oxidative stress in brain aging and neurodegeneration. World J. Biol. Chem. 2010, 1, 271–280. [Google Scholar] [CrossRef]
  69. Brendel, A.; Renziehausen, J.; Behl, C.; Hajieva, P. Downregulation of PMCA2 increases the vulnerability of midbrain neurons to mitochondrial complex I inhibition. NeuroToxicology 2014, 40, 43–51. [Google Scholar] [CrossRef]
  70. Mocko, J.B.; Kern, A.; Moosmann, B.; Behl, C.; Hajieva, P. Phenothiazines interfere with dopaminergic neurodegeneration in Caenorhabditis elegans models of Parkinson’s disease. Neurobiol. Dis. 2010, 40, 120–129. [Google Scholar] [CrossRef]
  71. Fernandes, D.; Zaidi, A.; Bean, J.; Hui, D.; Michaelis, M.L. RNAi- induced silencing of the plasma membrane Ca2+- ATPase 2 in neuronal cells: Effects on Ca2+ homeostasis and cell viability. J. Neurochem. 2007, 102, 454–465. [Google Scholar] [CrossRef] [PubMed]
  72. Martinez, J.; Moeller, I.; Erdjument-Bromage, H.; Tempst, P.; Lauring, B. Parkinson’s Disease-associated α-Synuclein Is a Calmodulin Substrate. J. Biol. Chem. 2003, 278, 17379–17387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Shi, X.; Sun, Y.; Wang, P.; Gu, L.; Wang, L.; Yang, H.; Wei, Q.; Li, Z.; Luo, J. The interaction between calcineurin and α-synuclein is regulated by calcium and calmodulin. Biochem. Biophys. Res. Commun. 2018, 496, 1109–1114. [Google Scholar] [CrossRef] [PubMed]
  74. Caraveo, G.; Soste, M.; Cappelleti, V.; Fanning, S.; Van Rossum, D.B.; Whitesell, L.; Huang, Y.; Chung, C.Y.; Baru, V.; Zaichick, S.; et al. FKBP12 contributes to α-synuclein toxicity by regulating the calcineurin-dependent phosphoproteome. Proc. Natl. Acad. Sci. USA 2017, 114, E11313–E11322. [Google Scholar] [CrossRef] [Green Version]
  75. Holton, M.; Yang, D.; Wang, W.; Mohamed, T.M.; Neyses, L.; Armesilla, A.L. The interaction between endogenous calcineurin and the plasma membrane calcium-dependent ATPase is isoform specific in breast cancer cells. FEBS Lett. 2007, 581, 4115–4119. [Google Scholar] [CrossRef] [PubMed]
  76. Kosiorek, M.; Podszywalow-Bartnicka, P.; Zylinska, L.; Zablocki, K.; Pikula, S. Interaction of plasma membrane Ca2+-ATPase isoform 4 with calcineurin A: Implications for catecholamine secretion by PC12 cells. Biochem. Biophys. Res. Commun. 2011, 411, 235–240. [Google Scholar] [CrossRef]
  77. Boczek, T.; Ferenc, B.; Lisek, M.; Zylinska, L. Regulation of GAP43/calmodulin complex formation via calcineurin-dependent mechanism in differentiated PC12 cells with altered PMCA isoforms composition. Mol. Cell. Biochem. 2015, 407, 251–262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Betzer, C.; Lassen, L.B.; Olsen, A.; Kofoed, R.H.; Reimer, L.; Gregersen, E.; Zheng, J.; Calì, T.; Gai, W.; Chen, T.; et al. Alpha-synuclein aggregates activate calcium pump SERCA leading to calcium dysregulation. EMBO Rep. 2018, 19, 5. [Google Scholar] [CrossRef]
  79. Lidow, M.S. Calcium signaling dysfunction in schizophrenia: A unifying approach. Brain Res. Rev. 2003, 43, 70–84. [Google Scholar] [CrossRef]
  80. Kluge, H.; Kühne, G. Preliminary findings on calmodulin-stimulated Ca2+-ATPase of erythrocyte ghosts in psychotic patients. Eur. Arch. Psychiatry Clin. Neurosci. 1985, 235, 57–59. [Google Scholar] [CrossRef]
  81. Martins-De-Souza, D.; Gattaz, W.F.; Schmitt, A.; Rewerts, C.; Marangoni, S.; Novello, J.C.; Maccarrone, G.; Turck, C.W.; Dias-Neto, E. Alterations in oligodendrocyte proteins, calcium homeostasis and new potential markers in schizophrenia anterior temporal lobe are revealed by shotgun proteome analysis. J. Neural. Transm. 2008, 116, 275–289. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Föcking, M.; Lopez, L.M.; English, J.A.; Dicker, P.; Wolff, A.; Brindley, E.; Wynne, K.; Cagney, G.; Cotter, D.R. Proteomic and genomic evidence implicates the postsynaptic density in schizophrenia. Mol. Psychiatry 2014, 20, 424–432. [Google Scholar] [CrossRef] [PubMed]
  83. Meltzer, H.L.; Kassir, S.; Goodnick, P.J.; Fieve, R.R.; Chrisomalis, L.; Feliciano, M.; Szypula, D. Calmodulin-activated calcium ATPase in bipolar illness. Neuropsychobiology 1988, 20, 169–173. [Google Scholar] [CrossRef]
  84. Meltzer, H.L.; Kassir, S. Abnormal calmodulin-activated CaATPase in manic-depressive subjects. J. Psychiatr. Res. 1982, 17, 29–35. [Google Scholar] [CrossRef]
  85. Krystal, J.H.; Karper, L.P.; Seibyl, J.P.; Freeman, G.K.; Delaney, R.; Bremner, J.D.; Heninger, G.R.; Bowers, M.B.; Charney, D.S. Subanesthetic effects of the noncompetitive NMDA antagonist, ketamine, in humans. Psychotomimetic, perceptual, cognitive, and neuroendocrine responses. Arch. Gen. Psychiatry 1994, 51, 199–214. [Google Scholar] [CrossRef]
  86. Lisek, M.; Boczek, T.; Ferenc, B.; Zylinska, L. Regional brain dysregulation of Ca2+-handling systems in ketamine-induced rat model of experimental psychosis. Cell Tissue Res. 2015, 363, 609–620. [Google Scholar] [CrossRef] [Green Version]
  87. Lisek, M.; Ferenc, B.; Studzian, M.; Pulaski, L.; Guo, F.; Zylinska, L.; Boczek, T. Glutamate Deregulation in Ketamine-Induced Psychosis—A Potential Role of PSD95, NMDA Receptor and PMCA Interaction. Front. Cell. Neurosci. 2017, 11, 181. [Google Scholar] [CrossRef] [Green Version]
  88. Boczek, T.; Lisek, M.; Ferenc, B.; Zylinska, L. Plasma membrane Ca2+-ATPase is a novel target for ketamine action. Biochem. Biophys. Res. Commun. 2015, 465, 312–317. [Google Scholar] [CrossRef]
  89. Adaikkan, C.; Taha, E.; Barrera, I.; David, O.; Rosenblum, K. Calcium/Calmodulin-Dependent Protein Kinase II and Eukaryotic Elongation Factor 2 Kinase Pathways Mediate the Antidepressant Action of Ketamine. Biol. Psychiatry 2018, 84, 65–75. [Google Scholar] [CrossRef]
  90. Xiao, Y.; Luo, H.; Yang, W.Z.; Zeng, Y.; Shen, Y.; Ni, X.; Shi, Z.; Zhong, J.; Liang, Z.; Fu, X.; et al. A Brain Signaling Framework for Stress-Induced Depression and Ketamine Treatment Elucidated by Phosphoproteomics. Front. Cell. Neurosci. 2020, 14, 48. [Google Scholar] [CrossRef]
  91. Sepulveda, M.R.; Mata, A.M. Localization of intracellular and plasma membrane Ca2+-ATPases in the cerebellum. Cerebellum 2005, 4, 82–89. [Google Scholar] [CrossRef] [PubMed]
  92. Van Der Heijden, M.E.; Sillitoe, R.V. Interactions between Purkinje Cells and Granule Cells Coordinate the Development of Functional Cerebellar Circuits. Neuroscience 2020. [Google Scholar] [CrossRef] [PubMed]
  93. Zanni, G.; Calì, T.; Kalscheuer, V.M.; Ottolini, D.; Barresi, S.; Lebrun, N.; Montecchi-Palazzi, L.; Hu, H.; Chelly, J.; Bertini, E.; et al. Mutation of plasma membrane Ca2+ ATPase isoform 3 in a family with X-linked congenital cerebellar ataxia impairs Ca2+ homeostasis. Proc. Natl. Acad. Sci. USA 2012, 109, 14514–14519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zanni, G.; Bertini, E.S. X-linked disorders with cerebellar dysgenesis. Orphanet J. Rare Dis. 2011, 6, 24. [Google Scholar] [CrossRef] [Green Version]
  95. Hsueh, Y.-P. Calcium/calmodulin-dependent serine protein kinase and mental retardation. Ann. Neurol. 2009, 66, 438–443. [Google Scholar] [CrossRef]
  96. Najm, J.; Horn, D.; Wimplinger, I.; Golden, J.A.; Chizhikov, V.V.; Sudi, J.; Christian, S.L.; Ullmann, R.; Kuechler, A.; Haas, C.; et al. Mutations of CASK cause an X-linked brain malformation phenotype with microcephaly and hypoplasia of the brainstem and cerebellum. Nat. Genet. 2008, 40, 1065–1067. [Google Scholar] [CrossRef]
  97. Calì, T.; Lopreiato, R.; Shimony, J.; Vineyard, M.; Frizzarin, M.; Zanni, G.; Zanotti, G.; Brini, M.; Shinawi, M.; Carafoli, E. A Novel Mutation in Isoform 3 of the Plasma Membrane Ca2+ Pump Impairs Cellular Ca2+ Homeostasis in a Patient with Cerebellar Ataxia and Laminin Subunit 1α Mutations. J. Biol. Chem. 2015, 290, 16132–16141. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Aldinger, K.A.; Mosca, S.J.; Tétreault, M.; Dempsey, J.C.; Ishak, G.E.; Hartley, T.; Phelps, I.G.; Lamont, R.E.; O’Day, D.R.; Basel, D.; et al. Mutations in LAMA1 Cause Cerebellar Dysplasia and Cysts with and without Retinal Dystrophy. Am. J. Hum. Genet. 2014, 95, 227–234. [Google Scholar] [CrossRef] [Green Version]
  99. Vicario, M.; Calì, T.; Cieri, D.; Vallese, F.; Bortolotto, R.; Lopreiato, R.; Zonta, F.; Nardella, M.; Micalizzi, A.; Lefeber, D.J.; et al. A novel PMCA3 mutation in an ataxic patient with hypomorphic phosphomannomutase 2 (PMM2) heterozygote mutations: Biochemical characterization of the pump defect. Biochim. Biophys. Acta BBA Mol. Basis Dis. 2017, 1863, 3303–3312. [Google Scholar] [CrossRef]
  100. Westphal, V.; Peterson, S.; Patterson, M.; Tournay, A.; Blumenthal, A.; Treacy, E.P.; Freeze, H.H. Functional significance of PMM2 mutations in mildly affected patients with congenital disorders of glycosylation Ia. Genet. Med. 2001, 3, 393–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Briones, P.; Vilaseca, M.A.; Schollen, E.; Ferrer, I.; Maties, M.; Busquets, C.; Artuch, R.; Gort, L.; Marco, M.; Van Schaftingen, E.; et al. Biochemical and molecular studies in 26 Spanish patients with congenital disorder of glycosylation type Ia. J. Inherit. Metab. Dis. 2003, 25, 635–646. [Google Scholar] [CrossRef]
  102. Andreotti, G.; de Vaca, I.C.; Poziello, A.; Monti, M.C.; Guallar, V.; Cubellis, M.V. Conformational response to ligand binding in phosphomannomutase2: Insights into inborn glycosylation disorder. J. Biol. Chem. 2014, 289, 34900–34910. [Google Scholar] [CrossRef] [Green Version]
  103. Bortolozzi, M.; Mammano, F. PMCA2 pump mutations and hereditary deafness. Neurosci. Lett. 2018, 663, 18–24. [Google Scholar] [CrossRef] [PubMed]
  104. Giacomello, M.; De Mario, A.; Primerano, S.; Brini, M.; Carafoli, E. Hair cells, plasma membrane Ca2+ ATPase and deafness. Int. J. Biochem. Cell Biol. 2012, 44, 679–683. [Google Scholar] [CrossRef]
  105. Carafoli, E. The plasma membrane calcium pump in the hearing process: Physiology and pathology. Sci. China Life Sci. 2011, 54, 686–690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Kozel, P.J.; Friedman, R.A.; Erway, L.C.; Yamoah, E.N.; Liu, L.H.; Riddle, T.; Duffy, J.J.; Doetschman, T.; Miller, M.L.; Cardell, E.L.; et al. Balance and Hearing Deficits in Mice with a Null Mutation in the Gene Encoding Plasma Membrane Ca2+-ATPase Isoform 2. J. Biol. Chem. 1998, 273, 18693–18696. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  107. Street, V.A.; McKee-Johnson, J.W.; Fonseca, R.C.; Tempel, B.L.; Noben-Trauth, K. Mutations in a plasma membrane Ca2+-ATPase gene cause deafness in deafwaddler mice. Nat. Genet. 1998, 19, 390–394. [Google Scholar] [CrossRef] [PubMed]
  108. Inoue, Y.; Matsumura, Y.; Inoue, K.; Ichikawa, R.; Takayama, C. Abnormal synaptic architecture in the cerebellar cortex of a new dystonic mutant mouse, Wriggle Mouse Sagami. Neurosci. Res. 1993, 16, 39–48. [Google Scholar] [CrossRef]
  109. Ueno, T.; Kameyama, K.; Hirata, M.; Ogawa, M.; Hatsuse, H.; Takagaki, Y.; Ohmura, M.; Osawa, N.; Kudo, Y. A mouse with a point mutation in plasma membrane Ca2+-ATPase isoform 2 gene showed the reduced Ca2+ influx in cerebellar neurons. Neurosci. Res. 2002, 42, 287–297. [Google Scholar] [CrossRef]
  110. Empson, R.M.; Turner, P.R.; Nagaraja, R.Y.; Beesley, P.W.; Knopfel, T. Reduced expression of the Ca2+transporter protein PMCA2 slows Ca2+dynamics in mouse cerebellar Purkinje neurones and alters the precision of motor coordination. J. Physiol. 2010, 588, 907–922. [Google Scholar] [CrossRef] [PubMed]
  111. Vicario, M.; Zanni, G.; Vallese, F.; Santorelli, F.; Grinzato, A.; Cieri, D.; Berto, P.; Frizzarin, M.; Lopreiato, R.; Zonta, F.; et al. A V1143F mutation in the neuronal-enriched isoform 2 of the PMCA pump is linked with ataxia. Neurobiol. Dis. 2018, 115, 157–166. [Google Scholar] [CrossRef] [PubMed]
  112. Brini, M.; Carafoli, E.; Calì, T. The plasma membrane calcium pumps: Focus on the role in (neuro)pathology. Biochem. Biophys. Res. Commun. 2017, 483, 1116–1124. [Google Scholar] [CrossRef] [PubMed]
  113. Xu, L.; Wang, Z.; Xiong, X.; Gu, X.; Gao, X.; Gao, X. Identification of a novel point mutation of mouse Atp2b2 induced by N-ethyl-N-nitrosourea mutagenesis. Exp. Anim. 2011, 60, 71–78. [Google Scholar] [CrossRef] [Green Version]
  114. Spiden, S.L.; Bortolozzi, M.; Di Leva, F.; De Angelis, M.H.; Fuchs, H.; Lim, D.; Ortolano, S.; Ingham, N.J.; Brini, M.; Carafoli, E.; et al. The Novel Mouse Mutation Oblivion Inactivates the PMCA2 Pump and Causes Progressive Hearing Loss. PLoS Genet. 2008, 4, e1000238. [Google Scholar] [CrossRef] [Green Version]
  115. Bortolozzi, M.; Brini, M.; Parkinson, N.; Crispino, G.; Scimemi, P.; De Siati, R.D.; Di Leva, F.; Parker, A.; Ortolano, S.; Arslan, E.; et al. The Novel PMCA2 Pump Mutation Tommy Impairs Cytosolic Calcium Clearance in Hair Cells and Links to Deafness in Mice. J. Biol. Chem. 2010, 285, 37693–37703. [Google Scholar] [CrossRef] [Green Version]
  116. Takahashi, K.; Kitamura, K. A Point Mutation in a Plasma Membrane Ca2+-ATPase Gene Causes Deafness in Wriggle Mouse Sagami. Biochem. Biophys. Res. Commun. 1999, 261, 773–778. [Google Scholar] [CrossRef] [PubMed]
  117. Figueroa, K.P.; Paul, S.; Cali’, T.; Lopreiato, R.; Karan, S.; Frizzarin, M.; Ames, D.; Zanni, G.; Brini, M.; Dansithong, W.; et al. Spontaneous shaker rat mutant—A new model for X-linked tremor/ataxia. Dis. Model. Mech. 2016, 9, 553–562. [Google Scholar] [CrossRef] [Green Version]
  118. Lopreiato, R.; Giacomello, M.; Carafoli, E. The Plasma Membrane Calcium Pump: New Ways to Look at an Old Enzyme. J. Biol. Chem. 2014, 289, 10261–10268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Lisek, M.; Boczek, T.; Zylinska, L. Calcium as a Trojan horse in mental diseases—The role of PMCA and PMCA-interacting proteins in bipolar disorder and schizophrenia. Neurosci. Lett. 2018, 663, 48–54. [Google Scholar] [CrossRef]
  120. Fromer, M.; Pocklington, A.J.; Kavanagh, D.H.; Williams, H.J.; Dwyer, S.; Gormley, P.; Georgieva, L.; Rees, E.; Palta, P.; Ruderfer, D.M.; et al. De novo mutations in schizophrenia implicate synaptic networks. Nat. Cell Biol. 2014, 506, 179–184. [Google Scholar] [CrossRef] [Green Version]
  121. Coley, A.A.; Gao, W.-J. PSD95: A synaptic protein implicated in schizophrenia or autism? Prog. Neuro Psychopharmacol. Biol. Psychiatry 2018, 82, 187–194. [Google Scholar] [CrossRef] [PubMed]
  122. Funk, A.J.; Mielnik, C.A.; Koene, R.; Newburn, E.; Ramsey, A.J.; Lipska, B.K.; McCullumsmith, R.E. Postsynaptic Density-95 Isoform Abnormalities in Schizophrenia. Schizophr. Bull. 2017, 43, 891–899. [Google Scholar] [CrossRef] [PubMed]
  123. Garside, M.L.; Turner, P.R.; Austen, B.; Strehler, E.E.; Beesley, P.W.; Empson, R.M. Molecular interactions of the plasma membrane calcium ATPase 2 at pre- and post-synaptic sites in rat cerebellum. Neuroscience 2009, 162, 383–395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Yamamoto, H.; Hagino, Y.; Kasai, S.; Ikeda, K. Specific Roles of NMDA Receptor Subunits in Mental Disorders. Curr. Mol. Med. 2015, 15, 193–205. [Google Scholar] [CrossRef] [Green Version]
  125. Strehler, E.E. Plasma Membrane Calcium ATPases as Novel Candidates for Therapeutic Agent Development. J. Pharm. Pharm. Sci. 2013, 16, 190–206. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Zhou, Q.-G.; Zhu, X.-H.; Nemes, A.D.; Zhu, D.-Y. Neuronal nitric oxide synthase and affective disorders. IBRO Rep. 2018, 5, 116–132. [Google Scholar] [CrossRef] [PubMed]
  127. Knott, A.B.; Bossy-Wetzel, E. Nitric Oxide in Health and Disease of the Nervous System. Antioxid. Redox Signal. 2009, 11, 541–553. [Google Scholar] [CrossRef] [Green Version]
  128. Becker, M.; Mastropasqua, F.; Reising, J.P.; Maier, S.; Ho, M.-L.; Rabkina, I.; Li, D.; Neufeld, J.; Ballenberger, L.; Myers, L.; et al. Presynaptic dysfunction in CASK-related neurodevelopmental disorders. Transl. Psychiatry 2020, 10, 1–17. [Google Scholar] [CrossRef]
  129. Voit, T.; Kramer, H.; Thomas, C.; Wechsler, W.; Reichmann, H.; Lenard, H.G. Mypopathy in Williams-Beuren syndrome. Eur. J. Nucl. Med. Mol. Imaging 1991, 150, 521–526. [Google Scholar] [CrossRef]
  130. Tarpey, P.; Parnau, J.; Blow, M.; Woffendin, H.; Bignell, G.; Cox, C.; Cox, J.; Davies, H.; Edkins, S.; Holden, S.; et al. Mutations in the DLG3 Gene Cause Nonsyndromic X-Linked Mental Retardation. Am. J. Hum. Genet. 2004, 75, 318–324. [Google Scholar] [CrossRef] [Green Version]
  131. Kristiansen, L.V.; Meador-Woodruff, J.H. Abnormal striatal expression of transcripts encoding NMDA interacting PSD proteins in schizophrenia, bipolar disorder and major depression. Schizophr. Res. 2005, 78, 87–93. [Google Scholar] [CrossRef] [PubMed]
  132. Gardoni, F.; Marcello, E.; Di Luca, M. Postsynaptic density–membrane associated guanylate kinase proteins (PSD–MAGUKs) and their role in CNS disorders. Neuroscience 2009, 158, 324–333. [Google Scholar] [CrossRef] [PubMed]
  133. Clifton, N.E.; Trent, S.; Thomas, K.L.; Hall, J. Regulation and Function of Activity-Dependent Homer in Synaptic Plasticity. Mol. Neuropsychiatry 2019, 5, 147–161. [Google Scholar] [CrossRef] [PubMed]
  134. Miyakawa, T.; Leiter, L.M.; Gerber, D.J.; Gainetdinov, R.R.; Sotnikova, T.D.; Zeng, H.; Caron, M.G.; Tonegawa, S. Conditional calcineurin knockout mice exhibit multiple abnormal behaviors related to schizophrenia. Proc. Natl. Acad. Sci. USA 2003, 100, 8987–8992. [Google Scholar] [CrossRef] [Green Version]
  135. Kipanyula, M.J.; Kimaro, W.H.; Etet, P.F.S. The Emerging Roles of the Calcineurin-Nuclear Factor of Activated T-Lymphocytes Pathway in Nervous System Functions and Diseases. J. Aging Res. 2016, 2016, 5081021. [Google Scholar] [CrossRef] [Green Version]
  136. Mathieu, F.; Miot, S.; Etain, B.; El Khoury, M.-A.; Chevalier, F.; Bellivier, F.; Leboyer, M.; Giros, B.; Tzavara, E.T. Association between the PPP3CC gene, coding for the calcineurin gamma catalytic subunit, and bipolar disorder. Behav. Brain Funct. 2008, 4, 2. [Google Scholar] [CrossRef] [Green Version]
  137. Donohoe, G.; Walters, J.; Morris, D.W.; Quinn, E.M.; Judge, R.; Norton, N.; Giegling, I.; Hartmann, A.M.; Möller, H.-J.; Muglia, P.; et al. Influence of NOS1 on Verbal Intelligence and Working Memory in Both Patients with Schizophrenia and Healthy Control Subjects. Arch. Gen. Psychiatry 2009, 66, 1045–1054. [Google Scholar] [CrossRef] [Green Version]
  138. Iourov, I.Y.; Vorsanova, S.G.; Kurinnaia, O.S.; Yurov, Y.B. An Interstitial 20q11.21 Microdeletion Causing Mild Intellectual Disability and Facial Dysmorphisms. Case Rep. Genet. 2013, 2013, 353028. [Google Scholar] [CrossRef] [Green Version]
  139. Foote, M.; Zhou, Y. 14-3-3 proteins in neurological disorders. Int. J. Biochem. Mol. Boil. 2012, 3, 152–164. [Google Scholar]
  140. Periasamy, M.; Kalyanasundaram, A. SERCA pump isoforms: Their role in calcium transport and disease. Muscle Nerve 2007, 35, 430–442. [Google Scholar] [CrossRef]
  141. Brandl, C.J.; DeLeon, S.; Martin, D.R.; MacLennan, D.H. Adult forms of the Ca2+ATPase of sarcoplasmic reticulum. Expression in developing skeletal muscle. J. Biol. Chem. 1987, 262, 3768–3774. [Google Scholar] [CrossRef]
  142. Verkhratsky, A. Physiology and Pathophysiology of the Calcium Store in the Endoplasmic Reticulum of Neurons. Physiol. Rev. 2005, 85, 201–279. [Google Scholar] [CrossRef]
  143. Sepulveda, M.R.; Hidalgo-Sánchez, M.; Mata, A.M. Localization of endoplasmic reticulum and plasma membrane Ca2+-ATPases in subcellular fractions and sections of pig cerebellum. Eur. J. Neurosci. 2004, 19, 542–551. [Google Scholar] [CrossRef]
  144. Morita, M.; Kudo, Y. Growth factors upregulate astrocyte [Ca2+]i oscillation by increasing SERCA2b expression. Glia 2010, 58, 1988–1995. [Google Scholar] [CrossRef] [PubMed]
  145. Patel, J.C.; Witkovsky, P.; Avshalumov, M.V.; Rice, M.E. Mobilization of Calcium from Intracellular Stores Facilitates Somatodendritic Dopamine Release. J. Neurosci. 2009, 29, 6568–6579. [Google Scholar] [CrossRef] [PubMed]
  146. Arredouani, A.; Guiot, Y.; Jonas, J.-C.; Liu, L.H.; Nenquin, M.; Pertusa, J.A.; Rahier, J.; Rolland, J.-F.; Shull, G.E.; Stevens, M.; et al. SERCA3 Ablation Does Not Impair Insulin Secretion but Suggests Distinct Roles of Different Sarcoendoplasmic Reticulum Ca2+ Pumps for Ca2+ Homeostasis in Pancreatic -cells. Diabetes 2002, 51, 3245–3253. [Google Scholar] [CrossRef] [PubMed]
  147. Wuytack, F.; Raeymaekers, L.; De Smedt, H.; Eggermont, J.A.; Missiaen, L.; Bosch, L.V.D.; De Jaegere, S.; Verboomen, H.; Plessers, L.; Casteels, R. Ca2+-Transport ATPases and Their Regulation in Muscle and Brain. Ann. N. Y. Acad. Sci. 1992, 671, 82–91. [Google Scholar] [CrossRef] [PubMed]
  148. Wuytack, F.; Dode, L.; Baba-Aissa, F.; Raeymaekers, L. The SERCA3-type of organellar Ca2+pumps. Biosci. Rep. 1995, 15, 299–306. [Google Scholar] [CrossRef]
  149. Campbell, A.M.; Wuytack, F.; Fambrough, D.M. Differential distribution of the alternative forms of the sarcoplasmic/endoplasmic reticulum Ca2+-ATPase, SERCA2b and SERCA2a, in the avian brain. Brain Res. 1993, 605, 67–76. [Google Scholar] [CrossRef]
  150. Baba-Aissa, F.; Raeymaekers, L.; Wuytack, F.; De Greef, C.; Missiaen, L.; Casteels, R. Distribution of the organellar Ca2+ transport ATPase SERCA2 isoforms in the cat brain. Brain Res. 1996, 743, 141–153. [Google Scholar] [CrossRef]
  151. Salvador, J.M.; Berengena, M.; Sepúlveda, M.R.; Mata, A.M. Distribution of the intracellular Ca(2+)-ATPase isoform 2b in pig brain subcellular fractions and cross-reaction with a monoclonal antibody raised against the enzyme isoform. J. Biochem. 2001, 129, 621–626. [Google Scholar] [CrossRef] [PubMed]
  152. Pchitskaya, E.; Popugaeva, E.; Bezprozvanny, I. Calcium signaling and molecular mechanisms underlying neurodegenerative diseases. Cell Calcium 2018, 70, 87–94. [Google Scholar] [CrossRef] [PubMed]
  153. Britzolaki, A.; Saurine, J.; Flaherty, E.; Thelen, C.; Pitychoutis, P.M. The SERCA2: A Gatekeeper of Neuronal Calcium Homeostasis in the Brain. Cell. Mol. Neurobiol. 2018, 38, 981–994. [Google Scholar] [CrossRef] [PubMed]
  154. Britzolaki, A.; Saurine, J.; Klocke, B.; Pitychoutis, P.M. A Role for SERCA Pumps in the Neurobiology of Neuropsychiatric and Neurodegenerative Disorders. Adv. Exp. Med. Biol. 2020, 1131, 131–161. [Google Scholar] [CrossRef]
  155. Chami, M.; Oulès, B.; Szabadkai, G.; Tacine, R.; Rizzuto, R.; Paterlini-Bréchot, P. Role of SERCA1 Truncated Isoform in the Proapoptotic Calcium Transfer from ER to Mitochondria during ER Stress. Mol. Cell 2008, 32, 641–651. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Bussiere, R.; Oulès, B.; Mary, A.; Vaillant-Beuchot, L.; Martin, C.; El Manaa, W.; Vallée, D.; Duplan, E.; Paterlini-Bréchot, P.; Da Costa, C.A.; et al. Upregulation of the Sarco-Endoplasmic Reticulum Calcium ATPase 1 Truncated Isoform Plays a Pathogenic Role in Alzheimer’s Disease. Cells 2019, 8, 1539. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  157. Wang, Y.; Bruce, A.T.; Tu, C.; Ma, K.; Zeng, L.; Zheng, P.; Liu, Y.; Liu, Y. Protein aggregation of SERCA2 mutants associated with Darier disease elicits ER stress and apoptosis in keratinocytes. J. Cell Sci. 2011, 124, 3568–3580. [Google Scholar] [CrossRef] [Green Version]
  158. Wang, S.-L.; Yang, S.-F.; Chen, C.-C.; Tsai, P.-T.; Chai, C.-Y. Darier’s disease associated with bipolar affective disorder: A case report. Kaohsiung J. Med. Sci. 2002, 18, 622–626. [Google Scholar]
  159. Nakamura, T.; Kazuno, A.; Nakajima, K.; Kusumi, I.; Tsuboi, T.; Kato, T. Loss of function mutations in ATP2A2 and psychoses: A case report and literature survey. Psychiatry Clin. Neurosci. 2016, 70, 342–350. [Google Scholar] [CrossRef]
  160. Gordon-Smith, K.; Green, E.; Grozeva, D.; Tavadia, S.; Craddock, N.; Jones, L. Genotype-phenotype correlations in Darier disease: A focus on the neuropsychiatric phenotype. Am. J. Med. Genet. Part B Neuropsychiatr. Genet. 2018, 177, e32679. [Google Scholar] [CrossRef] [PubMed]
  161. Takeichi, T.; Sugiura, K.; Nakamura, Y.; Fujio, Y.; Konohana, I.; Akiyama, M. Darier’s Disease Complicated by Schizophrenia Caused by a Novel ATP2A2 Mutation. Acta Derm. Venereol. 2016, 96, 993–994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  162. Simmerman, H.K.B.; Jones, L.R. Phospholamban: Protein structure, mechanism of action, and role in cardiac function. Physiol. Rev. 1998, 78, 921–947. [Google Scholar] [CrossRef]
  163. Shaikh, S.A.; Sahoo, S.K.; Periasamy, M. Phospholamban and sarcolipin: Are they functionally redundant or distinct regulators of the Sarco(Endo)Plasmic Reticulum Calcium ATPase? J. Mol. Cell. Cardiol. 2016, 91, 81–91. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  164. Kranias, E.G.; Hajjar, R.J. Modulation of Cardiac Contractility by the Phopholamban/SERCA2a Regulatome. Circ. Res. 2012, 110, 1646–1660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Vandecaetsbeek, I.; Vangheluwe, P.; Raeymaekers, L.; Wuytack, F.; Vanoevelen, J. The Ca2+ Pumps of the Endoplasmic Reticulum and Golgi Apparatus. Cold Spring Harb. Perspect. Biol. 2011, 3, a004184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Stammers, A.N.; Susser, S.E.; Hamm, N.C.; Hlynsky, M.W.; Kimber, D.E.; Kehler, D.S.; Duhamel, T.A. The regulation of sarco(endo)plasmic reticulum calcium-ATPases (SERCA). Can. J. Physiol. Pharmacol. 2015, 93, 843–854. [Google Scholar] [CrossRef]
  167. Münch, G.; Bölck, B.; Karczewski, P.; Schwinger, R.H. Evidence for Calcineurin-mediated Regulation of SERCA 2a Activity in Human Myocardium. J. Mol. Cell. Cardiol. 2002, 34, 321–334. [Google Scholar] [CrossRef]
  168. Hawkins, B.T.; Davis, T.P. The Blood-Brain Barrier/Neurovascular Unit in Health and Disease. Pharmacol. Rev. 2005, 57, 173–185. [Google Scholar] [CrossRef]
  169. Michalicova, A.; Majerova, P.; Kovac, A. Tau Protein and Its Role in Blood–Brain Barrier Dysfunction. Front. Mol. Neurosci. 2020, 13, 570045. [Google Scholar] [CrossRef]
  170. Sweeney, M.D.; Sagare, A.P.; Zlokovic, B.V. Blood–brain barrier breakdown in Alzheimer disease and other neurodegenerative disorders. Nat. Rev. Neurol. 2018, 14, 133–150. [Google Scholar] [CrossRef]
  171. Vasile, F.; Dossi, E.; Rouach, N. Human astrocytes: Structure and functions in the healthy brain. Brain Struct. Funct. 2017, 222, 2017–2029. [Google Scholar] [CrossRef] [Green Version]
  172. Dossi, E.; Vasile, F.; Rouach, N. Human astrocytes in the diseased brain. Brain Res. Bull. 2018, 136, 139–156. [Google Scholar] [CrossRef]
  173. Sanz, P.; Garcia-Gimeno, M.A. Reactive Glia Inflammatory Signaling Pathways and Epilepsy. Int. J. Mol. Sci. 2020, 21, 4096. [Google Scholar] [CrossRef] [PubMed]
  174. Sorrentino, Z.A.; Giasson, B.I.; Chakrabarty, P. α-Synuclein and astrocytes: Tracing the pathways from homeostasis to neurodegeneration in Lewy body disease. Acta Neuropathol. 2019, 138, 1–21. [Google Scholar] [CrossRef]
  175. Lee, J.H.; Han, J.-H.; Kim, H.; Park, S.M.; Joe, E.-H.; Jou, I. Parkinson’s disease-associated LRRK2-G2019S mutant acts through regulation of SERCA activity to control ER stress in astrocytes. Acta Neuropathol. Commun. 2019, 7, 68. [Google Scholar] [CrossRef]
  176. Chen, J.; Chen, Y.; Pu, J. Leucine-Rich Repeat Kinase 2 in Parkinson’s Disease: Updated from Pathogenesis to Potential Therapeutic Target. Eur. Neurol. 2018, 79, 256–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Zimprich, A.; Biskup, S.; Leitner, P.; Lichtner, P.; Farrer, M.; Lincoln, S.; Kachergus, J.; Hulihan, M.; Uitti, R.J.; Calne, D.B.; et al. Mutations in LRRK2 Cause Autosomal-Dominant Parkinsonism with Pleomorphic Pathology. Neuron 2004, 44, 601–607. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Cookson, M.R. The role of leucine-rich repeat kinase 2 (LRRK2) in Parkinson’s disease. Nat. Rev. Neurosci. 2010, 11, 791–797. [Google Scholar] [CrossRef] [PubMed]
  179. Martin, I.; Kim, J.W.; Dawson, V.L.; Dawson, T.M. LRRK2 pathobiology in Parkinson’s disease. J. Neurochem. 2014, 131, 554–565. [Google Scholar] [CrossRef] [Green Version]
  180. Mandal, D.; Rulli, S.J.; Rao, R. Packing Interactions between Transmembrane Helices Alter Ion Selectivity of the Yeast Golgi Ca2+/Mn2+-ATPase PMR1. J. Biol. Chem. 2003, 278, 35292–35298. [Google Scholar] [CrossRef] [Green Version]
  181. Xiang, M.; Mohamalawari, D.; Rao, R. A Novel Isoform of the Secretory Pathway Ca2+,Mn2+-ATPase, hSPCA2, Has Unusual Properties and Is Expressed in the Brain. J. Biol. Chem. 2005, 280, 11608–11614. [Google Scholar] [CrossRef] [Green Version]
  182. Sidoryk-Wegrzynowicz, M.; Aschner, M. Manganese toxicity in the central nervous system: The glutamine/glutamate-γ-aminobutyric acid cycle. J. Intern. Med. 2013, 273, 466–477. [Google Scholar] [CrossRef] [Green Version]
  183. Van Baelen, K.; Dode, L.; Vanoevelen, J.; Callewaert, G.; De Smedt, H.; Missiaen, L.; Parys, J.B.; Raeymaekers, L.; Wuytack, F. The Ca2+/Mn2+ pumps in the Golgi apparatus. Biochim. Biophys. Acta BBA Bioenerg. 2004, 1742, 103–112. [Google Scholar] [CrossRef] [Green Version]
  184. Wootton, L.L.; Argent, C.C.; Wheatley, M.; Michelangeli, F. The expression, activity and localisation of the secretory pathway Ca2+-ATPase (SPCA1) in different mammalian tissues. Biochim. Biophys. Acta BBA Biomembr. 2004, 1664, 189–197. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Sepulveda, M.R.; Vanoevelen, J.; Raeymaekers, L.; Mata, A.M.; Wuytack, F. Silencing the SPCA1 (Secretory Pathway Ca2+-ATPase Isoform 1) Impairs Ca2+ Homeostasis in the Golgi and Disturbs Neural Polarity. J. Neurosci. 2009, 29, 12174–12182. [Google Scholar] [CrossRef] [PubMed]
  186. Murín, R.; Verleysdonk, S.; Raeymaekers, L.; Kaplan, P.; Lehotský, J. Distribution of Secretory Pathway Ca2+ ATPase (SPCA1) in Neuronal and Glial Cell Cultures. Cell. Mol. Neurobiol. 2006, 26, 1353–1363. [Google Scholar] [CrossRef]
  187. Vanoevelen, J.; Dode, L.; Van Baelen, K.; Fairclough, R.J.; Missiaen, L.; Raeymaekers, L.; Wuytack, F. The Secretory Pathway Ca2+/Mn2+-ATPase 2 Is a Golgi-localized Pump with High Affinity for Ca2+ Ions. J. Biol. Chem. 2005, 280, 22800–22808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  188. Sepulveda, M.R.; Berrocal, M.; Marcos, D.; Wuytack, F.; Mata, A.M. Functional and immunocytochemical evidence for the expression and localization of the secretory pathway Ca2+-ATPase isoform 1 (SPCA1) in cerebellum relative to other Ca2+pumps. J. Neurochem. 2007, 103, 1009–1018. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  189. Micaroni, M.; Giacchetti, G.; Plebani, R.; Xiao, G.G.; Federici, L. ATP2C1 gene mutations in Hailey–Hailey disease and possible roles of SPCA1 isoforms in membrane trafficking. Cell Death Dis. 2016, 7, e2259. [Google Scholar] [CrossRef] [Green Version]
  190. Chen, J.; Smaardijk, S.; Mattelaer, C.-A.; Pamula, F.; Vandecaetsbeek, I.; Vanoevelen, J.; Wuytack, F.; Lescrinier, E.; Eggermont, J.; Vangheluwe, P. An N-terminal Ca2+-binding motif regulates the secretory pathway Ca2+/Mn2+-transport ATPase SPCA1. J. Biol. Chem. 2019, 294, 7878–7891. [Google Scholar] [CrossRef]
  191. Laude, A.J.; Simpson, A.W.M. Compartmentalized signalling: Ca2+ compartments, microdomains and the many facets of Ca2+signalling. FEBS J. 2009, 276, 1800–1816. [Google Scholar] [CrossRef]
  192. Missiaen, L.; Dode, L.; Vanoevelen, J.; Raeymaekers, L.; Wuytack, F. Calcium in the Golgi apparatus. Cell Calcium 2007, 41, 405–416. [Google Scholar] [CrossRef]
  193. Brunskill, E.W.; Potter, A.S.; Distasio, A.; Dexheimer, P.; Plassard, A.; Aronow, B.J.; Potter, S.S. A gene expression atlas of early craniofacial development. Dev. Biol. 2014, 391, 133–146. [Google Scholar] [CrossRef] [Green Version]
  194. Ramos-Castañeda, J.; Park, Y.-N.; Liu, M.; Hauser, K.; Rudolph, H.; Shull, G.E.; Jonkman, M.F.; Mori, K.; Ikeda, S.; Ogawa, H.; et al. Deficiency of ATP2C1, a Golgi Ion Pump, Induces Secretory Pathway Defects in Endoplasmic Reticulum (ER)-associated Degradation and Sensitivity to ER Stress. J. Biol. Chem. 2005, 280, 9467–9473. [Google Scholar] [CrossRef] [Green Version]
  195. Brown, J.M.; García-García, M.J. Secretory pathway calcium ATPase 1 (SPCA1) controls mouse neural tube closure by regulating cytoskeletal dynamics. Development 2018, 145, dev170019. [Google Scholar] [CrossRef] [Green Version]
  196. Okunade, G.W.; Miller, M.L.; Azhar, M.; Andringa, A.; Sanford, L.P.; Doetschman, T.; Prasad, V.; Shull, G.E. 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. 2007, 282, 26517–26527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Fan, J.; Hu, Z.; Zeng, L.; Lu, W.; Tang, X.; Zhang, J.; Li, T. Golgi apparatus and neurodegenerative diseases. Int. J. Dev. Neurosci. 2008, 26, 523–534. [Google Scholar] [CrossRef] [PubMed]
  198. Lehotský, J.; Racay, P.; Pavlíková, M.; Tatarková, Z.; Urban, P.; Chomová, M.; Kovalská, M.; Kaplán, P. Cross-talk of intracellular calcium stores in the response to neuronal ischemia and ischemic tolerance. Gen. Physiol. Biophys. 2009, 28, 104–114. [Google Scholar]
  199. Hu, Z.-P.; Li, L.-H.; Tian, X.-R. The key target of neuroprotection after the onset of ischemic stroke: Secretory pathway Ca2+-ATPase 1. Neural Regen. Res. 2015, 10, 1271–1278. [Google Scholar] [CrossRef]
  200. Pavlíková, M.; Tatarková, Z.; Sivoňová, M.; Kaplan, P.; Križanová, O.; Lehotský, J. Alterations Induced by Ischemic Preconditioning on Secretory Pathways Ca2+-ATPase (SPCA) Gene Expression and Oxidative Damage After Global Cerebral Ischemia/Reperfusion in Rats. Cell. Mol. Neurobiol. 2009, 29, 909–916. [Google Scholar] [CrossRef]
  201. Lu, T.; Hu, Z.; Zeng, L.; Jiang, Z. Changes in secretory pathway Ca2+-ATPase 2 following focal cerebral ischemia/reperfusion injury. Neural Regen. Res. 2013, 8, 76–82. [Google Scholar]
  202. Olanow, C.W. Manganese-Induced Parkinsonism and Parkinson’s Disease. Ann. N. Y. Acad. Sci. 2004, 1012, 209–223. [Google Scholar] [CrossRef]
  203. Aschner, M.; Erikson, K.M.; Hernández, E.H.; Tjalkens, R. Manganese and its Role in Parkinson’s Disease: From Transport to Neuropathology. Neuro Mol. Med. 2009, 11, 252–266. [Google Scholar] [CrossRef] [PubMed]
  204. Sepúlveda, M.R.; Wuytack, F.; Mata, A.M. High levels of Mn2+inhibit secretory pathway Ca2+/Mn2+-ATPase (SPCA) activity and cause Golgi fragmentation in neurons and glia. J. Neurochem. 2012, 123, 824–836. [Google Scholar] [CrossRef] [PubMed]
  205. Racette, B.A.; Aschner, M.; Guilarte, T.R.; Dydak, U.; Criswell, S.R.; Zheng, W. Pathophysiology of manganese-associated neurotoxicity. NeuroToxicology 2012, 33, 881–886. [Google Scholar] [CrossRef] [Green Version]
  206. Gubert, P.; Boas, G.R.V.; Paes, M.M.; Santamaría, A.; Lee, E.; Tinkov, A.A.; Bowman, A.B.; Aschner, M. Manganese-induced neurodegenerative diseases and possible therapeutic approaches. Expert Rev. Neurother. 2020, 20, 1109–1121. [Google Scholar] [CrossRef]
  207. Tong, Y.; Yang, H.; Tian, X.; Wang, H.; Zhou, T.; Zhang, S.; Yu, J.; Zhang, T.; Fan, D.; Guo, X.; et al. High Manganese, A Risk for Alzheimer’s Disease: High Manganese Induces Amyloid-β Related Cognitive Impairment. J. Alzheimers Dis. 2014, 42, 865–878. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  208. Stevenson, R.; Samokhina, E.; Rossetti, I.; Morley, J.W.; Buskila, Y. Neuromodulation of Glial Function During Neurodegeneration. Front. Cell. Neurosci. 2020, 14, 278. [Google Scholar] [CrossRef]
  209. Verkhratsky, A.; Nedergaard, M. Physiology of Astroglia. Physiol. Rev. 2018, 98, 239–389. [Google Scholar] [CrossRef]
  210. Von Bartheld, C.S.; Bahney, J.; Herculano-Houzel, S. The search for true numbers of neurons and glial cells in the human brain: A review of 150 years of cell counting. J. Comp. Neurol. 2016, 524, 3865–3895. [Google Scholar] [CrossRef] [Green Version]
  211. Perea, G.; Sur, M.; Araque, A. Neuron-glia networks: Integral gear of brain function. Front. Cell. Neurosci. 2014, 8, 378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Durkee, C.A.; Araque, A. Diversity and Specificity of Astrocyte–neuron Communication. Neuroscience 2019, 396, 73–78. [Google Scholar] [CrossRef]
  213. Caudal, L.C.; Gobbo, D.; Scheller, A.; Kirchhoff, F. The Paradox of Astroglial Ca2+ Signals at the Interface of Excitation and Inhibition. Front. Cell. Neurosci. 2020, 14, 609947. [Google Scholar] [CrossRef]
  214. Porasuphatana, S.; Tsai, P.; Rosen, G.M. The generation of free radicals by nitric oxide synthase. Comp. Biochem. Physiol. Part C Toxicol. Pharmacol. 2003, 134, 281–289. [Google Scholar] [CrossRef]
  215. Moreno, J.A.; Sullivan, K.A.; Carbone, D.L.; Hanneman, W.H.; Tjalkens, R.B. Manganese potentiates nuclear factor-κB-dependent expression of nitric oxide synthase 2 in astrocytes by activating soluble guanylate cyclase and extracellular responsive kinase signaling pathways. J. Neurosci. Res. 2008, 86, 2028–2038. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Tseng, K.Y.; Caballero, A.; Dec, A.; Cass, D.K.; Simak, N.; Sunu, E.; Park, M.J.; Blume, S.R.; Sammut, S.; Park, D.J.; et al. Inhibition of Striatal Soluble Guanylyl Cyclase-cGMP Signaling Reverses Basal Ganglia Dysfunction and Akinesia in Experimental Parkinsonism. PLoS ONE 2011, 6, e27187. [Google Scholar] [CrossRef] [PubMed]
  217. Araki, S.; Osuka, K.; Takata, T.; Tsuchiya, Y.; Watanabe, Y. Coordination between Calcium/Calmodulin-Dependent Protein Kinase II and Neuronal Nitric Oxide Synthase in Neurons. Int. J. Mol. Sci. 2020, 21, 7997. [Google Scholar] [CrossRef]
  218. Deng, S.; Liu, H.; Qiu, K.; You, H.; Lei, Q.; Lu, W. Role of the Golgi Apparatus in the Blood-Brain Barrier: Golgi Protection May Be a Targeted Therapy for Neurological Diseases. Mol. Neurobiol. 2017, 55, 4788–4801. [Google Scholar] [CrossRef]
  219. Parys, J.B.; Vervliet, T. New Insights in the IP3 Receptor and Its Regulation. Calcium Signal. 2020, 1131, 243–270. [Google Scholar]
  220. Kesherwani, V.; Agrawal, S.K. Regulation of Inositol 1,4,5-triphosphate receptor, type 1 (IP3R1) in hypoxic/reperfusion injury of white matter. Neurol. Res. 2012, 34, 504–511. [Google Scholar] [CrossRef]
  221. Seo, M.-D.; Enomoto, M.; Ishiyama, N.; Stathopulos, P.B.; Ikura, M. Structural insights into endoplasmic reticulum stored calcium regulation by inositol 1,4,5-trisphosphate and ryanodine receptors. Biochim. Biophys. Acta BBA Bioenerg. 2015, 1853, 1980–1991. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  222. Del Prete, D.; Checler, F.; Chami, M. Ryanodine receptors: Physiological function and deregulation in Alzheimer disease. Mol. Neurodegener. 2014, 9, 21. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The model of PMCA ((A), PDB entry code 6A69) and SERCA ((B), PDB entry code 3W5C) structures, and SPCA structure prediction using SWISS-MODEL (C). The cartoon models were generated with PyMOL.
Figure 1. The model of PMCA ((A), PDB entry code 6A69) and SERCA ((B), PDB entry code 3W5C) structures, and SPCA structure prediction using SWISS-MODEL (C). The cartoon models were generated with PyMOL.
Ijms 22 02785 g001
Table 1. Properties of PMCA isoforms. Modified based on [1].
Table 1. Properties of PMCA isoforms. Modified based on [1].
PMCA1PMCA2PMCA3PMCA4
Tissue DistributionUbiquitousRestricted
(brain)
Restricted
(brain)
Ubiquitous
Developmental Expression/SwitchIsoform switch fetal/adultIsoform switch fetal/adultIsoform switch fetal/adultIsoform switch fetal/adult
Affinity CaM (Kd nM)40–502–483–40
Table 3. Protein interacting with PMCA. Modified based on [125]. AD—Alzheimer’s disease; BP—bipolar disorder; HD—Huntington’s disease; PD—Parkinson’s disease; SZ—schizophrenia, MDD—major depressive disorder, ALS—amytrophic lateral sclerosis, ASD—autism spectrum disorder.
Table 3. Protein interacting with PMCA. Modified based on [125]. AD—Alzheimer’s disease; BP—bipolar disorder; HD—Huntington’s disease; PD—Parkinson’s disease; SZ—schizophrenia, MDD—major depressive disorder, ALS—amytrophic lateral sclerosis, ASD—autism spectrum disorder.
A Protein Interacting with PMCAAssociated DiseasePMCA Domain Involved in the InteractionA Protein Domain Involved in the InteractionFunctional Importance of the Interaction
NOSAD, BP, MDD, ALS, anxiety, stroke, HD [126,127]PDZ-domain binding sequencePDZ domainDecline in NOS activity, down-regulation of NO production
CASKMicrocephaly with pontine and cerebellar hypoplasia, X-linked intellectual disability, ASD [128]PDZ-domain binding sequencePDZ domainDown-regulation the T- dependent transcriptional activity
CLP36Williams-Beuren syndrome [129]PDZ-domain binding sequencePDZ domainPump translocation during platelet activation
MAGUKAD, PD, stroke, X-linked mental retardation, BD, MDD, SZ [130,131,132]PDZ-domain binding sequencePDZ domainBiding enables the localization of PMCAs in specific membrane domains and local control of Ca2+ concentration
Ania3/HomerSZ, ASD, MDD, suicide attempt, cocaine dependence, opiate abuse [133]PDZ-domain binding sequencePDZ domainStabilization of PMCA in domains near the sites of calcium influx into the cell
CalcineurinAD, HD, SZ, PD, ALS, BD epilepsy [134,135,136]catalytic domainAmino acids 58–143Inhibition of the phosphatase activity of calcineurin, decrease in the activity of the transcription factor NFAT
Syntrophin α1SZ, mild intellectual disability [137,138]catalytic domainAmino acids 399–447The formation of a triple complex with PMCA and NOS-1 inhibits the production of NO
ε 14-3-3AD, BP, PD, SZ [139]the N-terminal regionAmino acids 2–92Inhibition of PMCA activity
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Boczek, T.; Sobolczyk, M.; Mackiewicz, J.; Lisek, M.; Ferenc, B.; Guo, F.; Zylinska, L. Crosstalk among Calcium ATPases: PMCA, SERCA and SPCA in Mental Diseases. Int. J. Mol. Sci. 2021, 22, 2785. https://doi.org/10.3390/ijms22062785

AMA Style

Boczek T, Sobolczyk M, Mackiewicz J, Lisek M, Ferenc B, Guo F, Zylinska L. Crosstalk among Calcium ATPases: PMCA, SERCA and SPCA in Mental Diseases. International Journal of Molecular Sciences. 2021; 22(6):2785. https://doi.org/10.3390/ijms22062785

Chicago/Turabian Style

Boczek, Tomasz, Marta Sobolczyk, Joanna Mackiewicz, Malwina Lisek, Bozena Ferenc, Feng Guo, and Ludmila Zylinska. 2021. "Crosstalk among Calcium ATPases: PMCA, SERCA and SPCA in Mental Diseases" International Journal of Molecular Sciences 22, no. 6: 2785. https://doi.org/10.3390/ijms22062785

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

Article Metrics

Back to TopTop