Next Article in Journal
Lipoprotein(a) As a Potential Predictive Factor for Earlier Aortic Valve Replacement in Patients with Bicuspid Aortic Valve
Next Article in Special Issue
The Role of Dopamine in Repurposing Drugs for Oncology
Previous Article in Journal
Age, Dose, and Locomotion: Decoding Vulnerability to Ketamine in C57BL/6J and BALB/c Mice
Previous Article in Special Issue
Dopamine Receptor Ligand Selectivity—An In Silico/In Vitro Insight
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Na+,K+-ATPase and Cardiotonic Steroids in Models of Dopaminergic System Pathologies

by
Alisa A. Markina
1,2,
Rogneda B. Kazanskaya
1,3,
Julia A. Timoshina
3,4,
Vladislav A. Zavialov
1,2,
Denis A. Abaimov
3,
Anna B. Volnova
1,
Tatiana N. Fedorova
3,
Raul R. Gainetdinov
2,5 and
Alexander V. Lopachev
2,3,*
1
Biological Department, Saint Petersburg State University, Universitetskaya Emb. 7/9, 199034 Saint Petersburg, Russia
2
Institute of Translational Biomedicine, Saint Petersburg State University, Universitetskaya Emb. 7/9, 199034 Saint Petersburg, Russia
3
Research Center of Neurology, Volokolamskoye Ahosse 80, 125367 Moscow, Russia
4
Biological Department, Lomonosov Moscow State University, Leninskiye Gory 1, 119991 Moscow, Russia
5
Saint Petersburg University Hospital, 199034 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Biomedicines 2023, 11(7), 1820; https://doi.org/10.3390/biomedicines11071820
Submission received: 28 April 2023 / Revised: 13 June 2023 / Accepted: 17 June 2023 / Published: 25 June 2023
(This article belongs to the Special Issue Dopamine Signaling Pathway in Health and Disease)

Abstract

:
In recent years, enough evidence has accumulated to assert that cardiotonic steroids, Na+,K+-ATPase ligands, play an integral role in the physiological and pathophysiological processes in the body. However, little is known about the function of these compounds in the central nervous system. Endogenous cardiotonic steroids are involved in the pathogenesis of affective disorders, including depression and bipolar disorder, which are linked to dopaminergic system dysfunction. Animal models have shown that the cardiotonic steroid ouabain induces mania-like behavior through dopamine-dependent intracellular signaling pathways. In addition, mutations in the alpha subunit of Na+,K+-ATPase lead to the development of neurological pathologies. Evidence from animal models confirms the neurological consequences of mutations in the Na+,K+-ATPase alpha subunit. This review is dedicated to discussing the role of cardiotonic steroids and Na+,K+-ATPase in dopaminergic system pathologies—both the evidence supporting their involvement and potential pathways along which they may exert their effects are evaluated. Since there is an association between affective disorders accompanied by functional alterations in the dopaminergic system and neurological disorders such as Parkinson’s disease, we extend our discussion to the role of Na+,K+-ATPase and cardiotonic steroids in neurodegenerative diseases as well.

Graphical Abstract

1. Introduction

It is known that both neurons and glial cells need to constantly restore their resting membrane potentials. Maintenance and restoration of the resting potential is facilitated by Na+,K+-ATPase (NKA), a cytoplasmic membrane protein complex that exports three Na+ ions out of the cell in exchange for two K+ ions. This pump action is facilitated by the α subunit, part of a membrane protein complex that also includes the β and γ subunits [1]. In neurons, aside from the ubiquitous α1 isoform, a neuron-specific isoform is present—the α3, while glial cells express the α2 isoform in addition to α1 [2]. Na+ export is necessary for neurons to restore the resting potential after the propagation of an action potential, and it facilitates Na+-conjugated transport processes [3]. Glial cells use the Na+ and K+ gradient to transport various compounds across the membrane, including excess neurotransmitters from the synaptic cleft and energy-intensive substrates transported into neurons [4].
A large body of evidence hints at the association of NKA dysfunction with the development of neurodegenerative and neuropsychiatric diseases. For example, mutations in the ATP1A3 gene cause rapid-onset dystonia parkinsonism (RDP) and alternating hemiplegia of childhood (AHC) [5]. Neurotoxic α-synuclein aggregates, which are a hallmark of Parkinson’s disease, bind to the neuronal α3-subunit of NKA, disrupting its function [6]. Oxidative stress (OS), which can be caused by toxic dopamine metabolites [7], as well as protein kinase C (PKC) activation [8] also cause dysfunction of neuronal NKA. Thus, there is reason to further study the role of NKA dysfunction in pathophysiological processes in the central nervous system (CNS).
In addition to its role in maintaining resting membrane potential, NKA is also involved in a number of intracellular signaling pathways and is a receptor for cardiotonic steroids (CTS), which can induce changes in intracellular signaling when binding to the enzyme. To date, thanks to the use of mass spectrometric analysis, enough data have been accumulated that allow us to consider CTS as endogenous hormone-like compounds in mammals, including humans. Endogenous ouabain was identified in human blood plasma [8,9], and its role in the development of various diseases, including arterial hypertension, was shown [8,9,10]. The presence of marinobufagenin in human blood was identified [11]. Additionally, endogenous CTS were isolated from the bovine adrenal glands [12]. From the bovine hypothalamus, a compound with an integer mass measured by HPLC-mass spectrometry equal to ouabain was isolated by affinity chromatography [13]. Thus, it is assumed that endogenous ouabain can be produced in the brain and adrenal glands of mammals. It has been shown that its amount can increase in response to an increase in tissue NaCl concentration. Increased content of endogenous ouabain in the brain is associated with epilepsy and motor neuron dysfunction [14]. However, there is currently no complete understanding of the physiological role of CTS in the CNS. There is also virtually no knowledge about the pathways of their biosynthesis in the brain and their regulation.
In addition to endogenous CTS, the CNS can also be affected by exogenous factors: the use of the CTS digoxin to treat patients with heart failure can lead to a wide range of neuropsychiatric side effects, such as fatigue, depression, psychosis, and delirium [14,15]. In various experimental models, it was shown that CTS can affect the efficiency of Na+ and K+-dependent processes by inhibiting NKA [16]. Thus, inhibition of the α3 subunit in neurons leads to the inability to quickly restore the Na+ gradient and enable action potential generation [17]. It is also known that ouabain causes increased release of GABA and decreased rate of GABA reuptake [18]. In addition, the NKA in the CNS has a number of functions specific to each isoform that are not directly related to pump activity, including the regulation of other membrane proteins and the activity of intracellular signaling cascades [8]. Via binding to NKA, CTS can influence the work of membrane and cytoplasmic proteins with which they interact [19,20,21]. Experimental data obtained in an amphetamine-induced model of mania in mice indicated the possible involvement of endogenous CTS in the development of bipolar disorder [22]. When entering the bloodstream, endogenous CTS affect the excretory and cardiovascular systems [14]. However, there is currently no complete picture of the involvement of CTS in physiological and pathophysiological processes in the CNS.
In this review, we summarize the data obtained in various models on the role of NKAs and CTS in CNS pathologies related to dopaminergic system dysfunction.

2. Neurological Disorders in Animals with NKA Mutations

The α3 subunit of NKA is encoded by the ATP1A3 gene. To date, four mouse models used to study the in vivo consequences of mutations in the ATP1A3 gene have been described. The creation of model animals—mice in which the α3 subunit gene promoter (Atp1a3) is used to control the expression of the fluorescent protein ZsGreen1 (a3NKA-ZsGreen1 mouse model) [23]—made it possible to determine the localization of the α3 subunit in brain tissue. It was shown that the signal intensity was highest in the neuronal bodies located in the stem structures, including the substantia nigra, some nuclei of the thalamus and cerebellum. No fluorescence was detected in astrocytes and brain white matter.
Mutations in the ATP1A3 gene have an autosomal dominant inheritance pattern. Homozygous mutants die shortly after birth. Therefore, viable and fertile heterozygotes are used to study all four in vivo models. These models display symptoms and endophenotypes similar to those seen in the manic and depressive phases of bipolar disorder, rapid-onset dystonia parkinsonism, epilepsy, alternating hemiplegia of childhood, and CAPOS syndrome to varying degrees (Table 1) [3].
Heterozygous Myshkin mutants (NKA13AMyk/+; Myk/+) (1.1 in Table 1) carry a missense mutation with an amino acid substitution at position 810 (I810 N). Such NKA α3 subunits are expressed normally but are not functionally active. Myshkin mutants were originally developed as a preclinical model of epilepsy because heterozygotes exhibited spontaneous seizures [24]. By crossing with seizure-resistant C57BL/6NCr mice, mutants that did not exhibit seizures were obtained [32]. So far, Myk/+ mutants have been shown to be valid models of mania [33]. In behavioral tests, Myk/+ mutants demonstrated hyperactivity, circadian rhythm and sleep disturbances [34], risk-taking tendencies, and increased sensitivity to D-amphetamine [25,35]—these symptoms are seen in patients in the manic stage of bipolar disorder. Additionally, administration of lithium and valproic acid, effective in mania therapy, has been shown to normalize behavior in heterozygous mice. However, it is not known at this time whether an endophenotype of depression is possible in this model in response to stressors. Myk/+ mice were also shown to exhibit a number of disturbances in circadian behavioral rhythms related to the processing of sensory visual information but without disturbances in the function of clock genes [36]. The authors suggested a link between the identified circadian rhythm abnormalities in this mouse model and the sleep disorders observed in parkinsonism. Some reviews on rush-induced dystonia-parkinsonism suggested the use of Myshkin heterozygotes as models of this disease [37]. The 4-week-old Myk/+ displays a different gait than the wild type, unstable with a shorter stride and accompanied by tremor. Tremor and gait problems are symptoms characteristic of parkinsonism. Changes in glucose metabolism and functional brain connectivity have also been shown in mice of this line. However, Myk/+ heterozygotes are not adequate models of RDP and parkinsonism; their endophenotype is more similar to that of alternating hemiplegia of childhood [26].
Heterozygous mutants of Mashlool (α+/D801N; Mashl+/−) (1.2 In Table 1) also carry a missense mutation with an amino acid substitution at position 810. A similar amino acid substitution at the same position is found in AHC patients [38]. Hyperactivity, reduced learning ability, memory problems, tremor, and shorter stride length have been shown for this line of mice compared to wild-type mice. Dystonia, hemiplegia, and hyperexcitability were found in Mashl+/−. In vivo electrophysiology data show that heterozygotes require fewer electrical stimulations for full excitation than wild-type animals; in addition, registration of electrical activity of the amygdala and hippocampus shows that the duration of full excitation of these structures after stimulation is significantly longer in heterozygotes than in wild-type mice. Mashl+/− mutants show spontaneous seizures and have an increased mortality [27]. Mashlool mutant data show that this lineage can serve as an AHC model with some reservations, but it is difficult to judge whether it can be an adequate model for studying bipolar disorder.
Heterozygous mutants with a point mutation in the fourth intron (NKA1A3tm1Ling, NKA1A3+/−, α+/KOI4) (1.3 in Table 1) show an approximately 60% reduction in α3-subunit expression in the hippocampus [28] because of aberrant splicing. At the same time, total NKA activity is reduced by 15% compared to the wild type. Behavioral features of intact (unstressed) heterozygotes are hyperactivity, decreased anxiety, and sensitivity to methamphetamine. No behavioral manifestations of neurological disorders were found in intact heterozygotes [29]. High-performance liquid chromatography showed no change in the levels of serotonin, dopamine, and their metabolites in the striatum in heterozygotes compared to wild-type animals. However, heterozygotes showed increased locomotor activity when presented with methamphetamine, which may be related to disturbances in the dopaminergic system [28]. α+/KOI4 mice exposed to chronic variable stress (CVS) exhibit behaviors similar to those observed in the depressive phase of bipolar disorder: anhedonia, despair-like behavior, weight changes, increased anxiety, and impaired memory and socialization. At the same time, NKA1A3 activity was reduced by 33% compared to the stressed wild type, consistent with the endophenotype of depression [26]. Thus, CVS-treated α+/KOI4 mutants can serve as a model for the depressive phase of bipolar disorder. In males with this mutation, however, no overt symptoms of parkinsonism or dystonia were found before or after stressors. However, for females, chronic stress was shown to induce coordination problems. In addition, rearing in stressed heterozygotes of both sexes was shown to have a negative correlation with levels of dopamine and its metabolites, which was not observed in wild-type mice [29].
Heterozygous Atp1a3tm2Kwk/+ mutants (1.4 in Table 1) have directional deletion of exons 2–6. Hyperactivity in both cell and open field tests was shown for them, but their anxiety level is not significantly different from that of wild-type animals. Heterozygotes have a higher level of coordination and motor balance compared to the wild type. Stressors do not cause dystonia-like symptoms, but microinjections of kainate into the cerebellar vermis induced a similar state. Electrophysiological studies on slides showed a connection of the mutation to the GABAergic system but not to the dopaminergic system [30]. Heterozygotes at 4 weeks of age show a shorter stride length compared to the wild type. Older heterozygotes (6–12 weeks old) do not show gait abnormality in the absence of stressors. However, when exposed to stressors, they begin to take shorter steps when moving, compared to controls. This is very similar to the manifestation of RDP, the symptoms of which in humans can be triggered by stress. It has been suggested that Atp1a3tm2Kwk/+ mutants may be a good model for RDP, although researchers have not reported dystonia or other symptoms of parkinsonism (postural instability, bradykinesia) [31].
For all four genetic models, increased impulsivity, a propensity for risk-taking behavior, and decreased habituation have been shown to varying degrees. All of these behavioral traits are symptoms of mania. The most striking symptoms of a mania-like state are noted in Myshkin mutants. However, there is currently insufficient information about dopamine levels in this line of mice. The depressive phase of bipolar disorder is best reproduced in CVS-exposed NKA1A3tm1Ling mutants. A correlation was found between the activity of stressed mice of this lineage and dopamine levels, but the relationship between dopamine levels and the mania-like state of unstressed heterozygotes carrying this mutation is not well understood.
Gait impairment is one of the symptoms of parkinsonism, including RDP. Gait abnormalities in mice were shown for three of the four models. The Atp1a3tm2Kwk/+ model is the closest to RDP, but it does not demonstrate the full range of classic parkinsonism symptoms. Thus, no genetic model associated with a mutation in the ATP1A3 gene can be called sufficiently reliable to study parkinsonism, at least for the time being. Nevertheless, the manifestation of both manic behavior and motor disorders simultaneously in the models may indicate that mutations in the α3-subunit of NKA can phenotypically manifest these two pathologies. Further research is needed to understand the mechanisms of the relationship between these pathologies.
Mutations that disrupt the α2-subunit of NKA, which is expressed in the brain in glial cells, can also lead to the development of various neurological and neuropsychiatric disorders. Variants in the ATP1A2 gene, which encodes the α2-subunit of NKA, are associated with familial hemiplegic migraine. For example, patients with the G301R mutation are affected by a complex syndrome characterized by migraine comorbidity with epilepsy, motor symptoms, and depression or obsessive–compulsive disorder [39,40]. This mutation was successfully replicated in mice, which displayed impaired glutamate uptake and altered inflammatory cytokine signaling [39,40].

3. Using Cardiotonic Steroids to Model Dopaminergic System Dysfunction

In addition to using animal lines with mutations in the NKA genes, studies of the effect of NKA dysfunction on the dopaminergic system have been conducted using intracerebroventricular (ICV) administration of ouabain to laboratory animals. The first indication of CTS involvement in affective disorder pathogenesis was seen in patients with heart failure, who developed mania-like symptoms in response to treatment with digoxin [15]. After this discovery, a series of attempts was made to model BD using ouabain, which, like digoxin, is a cardenolide. The first report of mania-like behavior after ICV injection of ouabain in rats was published in 1995 [41]. Since then, two approaches to modeling BD using ICV ouabain injection have emerged.
The first category of models includes administration of highly concentrated ouabain during a stereotaxic operation into the lateral ventricle of an anesthetized animal, with a subsequent behavioral evaluation 7–10 days post injection. This approach showed that a single ICV injection of 5 µL of 1 mM ouabain causes increased locomotion and grooming frequency in rats 11 days post injection, accompanied by decreased phosphorylation of PI3K, Akt, and GSK3β, and unchanged ERK1/2 phosphorylation. Seven days post ouabain injection, oxidative changes were observed in brain tissue [42]. Both the manic and depressive phases of BD were present in this model [43]. Chronic administration of valproate, lithium, or AR-A014418 (an inhibitor of GSK3β) prevented all of the above [44,45]. Haloperidol, a D2 receptor antagonist, also prevented ouabain-induced hyperlocomotion in rats in concentrations that decrease locomotor activity in intact animals [46]. It was also shown that the mania-like behavior observed in this model was accompanied by PKC activation [47]. Fourteen days post ouabain injection, animals displayed locomotor depression and impaired memory. Levels of pro-BDNF and BDNF in the frontal cortex were found to be decreased on the 7th day post injection, while its receptor (TRKB) and CREB decreased on the 7th and 14th day post injection [48]. On the 14th day post ouabain administration, the observed depressive symptoms were accompanied by increased levels of interleukin IL-1β, IL-6, IL-10, TNF-α, and CINC-1 in the frontal cortex and hippocampus [49], which may indicate the development of neuroinflammatory processes. A similar model was developed in mice, where anesthetized animals were given ICV injections of 0.625 pmol ouabain. After 8 days, the animals developed signs of mania-like behavior accompanied by c-fos activation. Administration of lithium chloride and haloperidol also neutralized the effects of ouabain in this model [50].
The second category includes models where ouabain is administered to unanesthetized animals using a surgically implanted cannula, and the effects are observed immediately post injection and/or several days later. Thus, an increase in motor activity within 30 min after ICV injection of ouabain (5 µL 0.5–1 mM) was described. At the same time in the striatum, there was an increase in the phosphorylation of ERK1/2 and tyrosine hydroxylase (TH). Administration of the MEK1/2 inhibitor (ERK1/2 MAP kinase kinase) U0126 leveled the effect of ouabain on the motor activity of the animals [51]. In another study of this model, animals injected with ouabain were shown to have increased phosphorylation of Akt, GSK3β, FOXO1, and eNOS amid increased motor activity 1–8 h after ICV injection [52]. At the same time, chronic administration of lithium chloride (for 7 days before ICV injection of 5 μL of 1 mM ouabain) was shown to prevent an increase in motor activity in rats [53]. The mechanism of the effect of ouabain in this model was attributed to mTOR activation mediated by Akt and ERK1/2 activation, with a subsequent effect on the expression of a number of proteins [54]. In a recently published paper, we described a model of ouabain-induced mania in mice. ICV injection of 0.5 μL of 50 μM ouabain into the lateral ventricles of the brain caused an increase in motor activity and stereotypic movements, as well as a decrease in anxiety in the animals within 1 h after the injection. At the same time, ouabain was shown to cause a decrease in the rate of dopamine reuptake. Inhibitor analysis with haloperidol showed that the effects of ouabain were mediated by the activation of D2 dopamine receptors and were associated with Akt activation, GSK3β deactivation, and ERK1/2 kinase activation, but not with neurodegenerative changes, which were not detected in animals 24 h after ouabain administration [55].
It would seem that the models described above are associated with the administration of CTS in doses that significantly exceed physiological ones. However, there is evidence that in other models of mania in laboratory animals endogenous CTS play a role in the development of the pathophysiological process. Thus, the administration of anti-ouabain antibodies, which reduced amphetamine-induced hyperactivity, protected against OS in the brain [56]. Moreover, administration of the ouabain antagonist rostafuroxin ameliorated behavioral and brain biochemical changes in the dextromethorphan-induced mania model [57].
Based on the data obtained in these models, we can conclude that NKA dysfunction induced by the administration of both exogenous CTS and endogenous CTS may be associated with dopaminergic system dysfunction, causing symptoms of neuropsychiatric diseases. However, these studies have not shown neurological abnormalities and degeneration of dopaminergic neurons. The only study on modeling Parkinson’s disease (PD) with CTS was conducted on Danio Rerio, where the CTS neriifolin [58] was used as a parkinsonism inducer.

4. Evidence for NKA Dysfunction in Experimental PD Models

At the same time, there is ample evidence of NKA impairment in classical models of parkinsonism. The most widespread method for modeling PD in laboratory rodents utilizes mitochondrial toxins, such as 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), rotenone, and 6-hydroxydopamine (6-OHDA). Mitochondrial dysfunction accompanied by OS are both involved in DAergic neuron degeneration. Because of the high energy demands of NKA—its activity can account for over 50% of neuronal ATP consumption [59]—it is extremely sensitive to mitochondrial dysfunction. Because of its large number of modifiable sites, it is also sensitive to reactive oxygen species (ROS) [60].
In mouse MPTP-induced parkinsonism, an approximately 40% decrease in total NKA activity in the striatum occurs, accompanied by a 60% decrease in dopamine levels [61]. A 60% decrease in NKA activity was also observed in conditions of 1-methyl-4-phenylpyridinium (MPP+)-induced OS in NGF-differentiated pheochromocytoma of the rat adrenal medulla (PC12) cells [62]. NKA activity also decreases in rotenone-induced parkinsonism models—total brain activity by approximately 25–40% [63,64] and the midbrain and striatum by 22% and 28%, respectively [65]. One of the first effects of exposure to rotenone is intracellular accumulation of sodium, which causes early hyperpolarization and a build-up of intracellular calcium following depolarization [66]. This coincides with changes in ion traffic that occur post NKA inhibition with ouabain [67], suggesting that these changes in rotenone-induced parkinsonism are a direct consequence of NKA dysfunction due to impaired ATP synthesis. Exposure to 6-OHDA, which inhibits all four mitochondrial electron transport chain complexes, causes a 28–43% decrease in NKA activity accompanied by a significant decrease in DA and its metabolites [68,69].

5. Mechanisms and Positive Feedback Loops: NKA and DAergic System Dysfunction

Factors affecting NKA function can be divided into two broad categories, the first including non-specific factors such as ATP concentration; Na+, K+, Mg2+ concentrations; OS; phosphorylation by various intracellular kinases; misfolded protein aggregates (α-synuclein, β-amyloid, superoxide dismutase); and various modifications (including glutathionylation), and the second including specific ligands—CTS. Non-specific factors can be divided into factors that increase NKA activity and those that decrease it. For example, phosphorylation by PKC, OS, low ATP, and interaction with misfolded protein aggregates cause a decrease in NKA activity. Factors such as glutathionylation and increased intracellular Na+ or extracellular K+ concentrations cause an increase in NKA activity. In turn, CTS exert different effects depending on the CTS and the concentration—concentrations below 10 nM can induce an increase in NKA activity, while concentrations exceeding 10 nM inhibit it [70,71]. Via binding to E2P conformation of NKA [72], different CTS can lead to the activation of various intracellular signaling pathways, which was discussed previously by other authors [73,74]. As such, in this review, we will focus specifically on the effects that altered NKA function may have on dopamine signaling and metabolism.
As one of the main functions of NKA is the maintenance of the electrochemical gradient, alterations in its function inevitably affect Ca2+ signaling. Since the pacemaking activity of dopaminergic neurons, specifically those in the substantia nigra, is dependent on intracellular Ca2+ oscillations and continuous Ca2+ influx [75,76], dysregulation of Ca2+ oscillations via NKA inhibition may synergize with exposure to other risk factors, causing mitochondrial damage via oxidative stress [75,77]. Indeed, it was shown previously that Ca2+ influx in dopaminergic neurons is a feed-forward mechanism that stimulates mitochondrial oxidative phosphorylation [78], thus increasing metabolic load. Considering that dopaminergic neurons experience high basal metabolic load compared to other neuron types, NKA dysfunction-induced Ca2+ homeostasis alterations could contribute to dopaminergic neuron degeneration.
Ca2+ and NKA signaling in neurons was extensively discussed in a recent review by Kinoshita et al. [67], and as such we will not go into detail on the subject. In brief, CTS are known to influence Ca2+ homeostasis in different ways depending on the CTS and concentration. In low, nanomolar concentrations, CTS can cause Ca2+ oscillations in neurons, mediated by the direct protein interaction of NKA with the inositol 1,4,5-trisphosphate receptor (IP3R). Low concentration ouabain-induced Ca2+ oscillations were shown to promote dendritic growth in an embryonic culture of primary cortical neurons [79] and improve long-term spatial reference memory in rats when administered into the hippocampus [80]. As such, at low concentrations ouabain is considered to have a neuroprotective effect on some neurons through its activation of CREB, the Wnt/β-catenin pathway, and NF-κB [81]. In subnanomolar concentrations, ouabain also protects against NMDA-induced cytotoxicity via direct protein-to-protein interactions between NKA and the Na+/Ca2+ exchanger [20]. In concentrations that inhibit NKA, CTS binding slows down or reverses the action of the Na+/Ca2+ exchanger, which co-localizes with NKA, thus increasing local cytoplasmic Ca2+ and leading to glutamate-mediated excitotoxicity [82] (Figure 1).
On the basis of the available data, it is possible to suggest several hypotheses of how changes in NKA functioning, due to both fluctuating CTS levels and other factors, can lead to dopaminergic neuron death. In the above-described models, CTS cause an increase in dopamine receptor activation. This may be a consequence of impaired dopamine reuptake, increased dopamine release, or increased dopamine synthesis, as has been demonstrated in various studies [51,55]. For different tissue types and cell cultures, it was shown that CTS in non-inhibitory concentrations can cause OS via activation of the Src-ERK1/2 signaling pathway [83,84] (Figure 2A). In turn, we propose a pathway that can lead to non-inhibitory CTS concentrations causing OS specifically in dopaminergic neurons (Figure 2B).
Inhibition of NKA activity by 40–50 µM ouabain in mouse striatum slices was shown to induce a decrease in the rate of DA reuptake by the dopamine active transporter (DAT) and an increase in its duration in the synaptic cleft. Reduced DAT activity normally causes activation of D2 dopamine autoreceptors on the presynaptic membrane, increasing the rate of dopamine transport from the cytoplasm to vesicles via the vesicular monoamine transporter-2 (VMAT2) [85,86,87]. Long-term dysfunction of DAT leads to increased duration of DA circulation in the synaptic cleft [55]. In DAT gene knockout mice, it was shown that DAT dysfunction leads to a decrease in presynaptic D2 autoreceptors [88]. Thus, long-term DAT dysfunction can lead to both an increase in DA synthesis [89] and a decrease in its uptake into vesicles by VMAT2. VMAT2 dysfunction is known to be associated with the development of PD owing to the accumulation of toxic products of DA metabolism [90]. Moreover, people with DAT dysfunction develop juvenile parkinsonism (with complete loss of function in the first months of life, with partial loss of function in adolescence), whereas partial loss of function leads to the development of bipolar disorder [91] (Figure 1).
Previously it was shown that ICV ouabain administration causes an increase in TH phosphorylation via ERK1/2 activation, indicating that DA synthesis increases as well [51]. ERK1/2 activation post ouabain injection was demonstrated several times, both in vivo in rodents [51,55,92] and in vitro on neuron cultures [93]. Although it is known that ouabain can activate PKA and PKC in rat cortex neuron cultures [94], to our knowledge there have been no studies showing that ouabain-induced TH activation is mediated by these kinases. In various cell cultures, however, it has been shown that PKA activates TH via Ser40 phosphorylation [95,96]. Furthermore, PKA activation leads to an increase in TH expression [97]. It was also shown that phorscoline-induced PKA activation causes an increase in DA release in rat striatum slices [98] and increased D2R expression in the striatum post ICV administration in rats [99]. PKA activation also causes an increase in DAT activity in rat striatum-derived synaptosomes [100]. On the other hand, PKA inhibition in PC12 cell culture causes an increase in VMAT2 amounts in “synaptic” vesicles [101].
It is likely that post a single injection of ouabain, TH activity eventually returns to normal. However, if endogenous CTS levels in the brain remain elevated chronically, similar to blood plasma levels of CTS in hypertension [14], it is possible that the observed neuronal TH hyperactivity is sustained chronically as well. It is known that TH hyperactivation in neurons leads to the accumulation of toxic dopamine oxidation products, OS, and eventually neuron death [102]. In addition, it was shown that hyperstimulation of dopamine receptors can lead to neuronal death [103]. Prolonged D2R activation is known to trigger a β-arrestin-dependent signaling pathway, leading to increased GSK3β activity [104,105]. Pathological GSK3β activity is known to be associated with DA neuronal degeneration and PD [106]. Activation of GSK3β also causes NURR1 degradation [107], which is vital to VMAT2 expression [108]. One of the mechanisms responsible for neuronal death during GSK3β hyperactivation is an increase in NR2B-containing NMDAR activity followed by Ca2+ overload [109]. Thus, we can assume that NKA dysfunction is associated with OS and other stressors (including products of DA metabolism [7]).
As mentioned above, ERK1/2 activation and increased TH phosphorylation in the striatum is characteristic of CTS-induced mania-like behavior models [51]. It is known that activation of ERK1/2 in primary culture neurons can be induced by various CTS and is associated with the neurotoxic effect of ouabain [91,110]. In the described models, activation of ERK1/2 also occurs upon administration of ouabain. ERK1/2 is known to play an ambiguous role in the pathogenesis of PD. ERK1/2 activation is necessary for the implementation of protective mechanisms in neurons when exposed to stress factors that lead to the initiation of neurodegeneration. PI3K/Akt and ERK1/2 signaling pathways are known to be involved in protecting dopaminergic neurons from MPTP/MPP+-induced neurotoxicity [111]. Previously, it was shown that ERK1/2 is involved in neuronal antioxidant defense and translocating to the nucleus via binding to the DJ-1 protein [112]. Increased amounts of p-ERK1/2 were found in the mitochondria of degenerating neurons from PD patients and patients with dementia with Levi’s corpuscles [113]. Other studies supported the idea that ERK1/2 inhibition causes activation of both apoptotic and necrotic pathways, leading to neuronal death [114]. On the other hand, activation of ERK1/2 and JNK is known to be associated with L-DOPA-induced neurotoxicity to dopaminergic neurons in a cellular model of PD [115]. In PD models, ERK1/2 activation mediates the occurrence of OS in pro-inflammatory factor-activated microglia. ERK1/2 is also involved in the development of L-DOPA-induced dyskinesia by affecting synaptic plasticity in the striatum [116,117]. Using the CG4 oligodendroglial cell line, it was shown that H2O2-induced cell death is prevented by the ERK1/2 pathway inhibitor PD98059 [118]. PD98059 can also prevent neuronal degeneration caused by nitric oxide released by glial cells through ERK1/2 activation [119]. The use of another inhibitor, U0126, also demonstrated that dopamine-induced striatal neuronal death is associated with ERK1/2 activation [120].
Dopamine binding to dopamine receptors can decrease NKA activity through PKC and PKA activation [70]. Dopamine binding to the D1 dopamine receptor in striatum neurons leads to a decrease in NKA activity. Binding of dopamine to the D2 dopamine receptor induces sodium channels to open, causing a spike in intracellular Na+ concentration and activating NKA [121]. Using co-immunoprecipitation and mass spectrometry, it was shown that D1 and D2 dopamine receptors form a protein complex with NKA. Transfection of the D1 or D2 dopamine receptor into HEK293T cells without dopamine addition resulted in a marked decrease in α1-containing NKA activity but had no effect on its amount [122]. Furthermore, as mentioned earlier, OS and PKC activation also cause a decrease in NKA activity, closing the positive feedback loop.
Thus, there are many ways in which chronic NKA dysfunction due to a chronic increase in endogenous CTS in the brain or due to other factors affecting NKA may lead to the degeneration of dopaminergic neurons.

6. Conclusions

Although there is currently no clear picture of the role of CTS and NKA abnormalities in the development of neurodegenerative diseases of the dopaminergic system, there is an understanding of their role in the development of affective disorders associated with functional dopaminergic pathologies. That being said, there is a significant amount of evidence suggesting that CTS and NKA abnormalities may be key players in the development of neurodegenerative disorders of the DA system such as PD. Further study of changes in both NKA functioning and the amount of endogenous CTS in neurodegenerative disorders of the DA system, and mechanisms of CTS influence on the dopaminergic system in various models at the physiological, neurochemical, and biochemical levels could open up potential new pharmacological targets and biomarkers for both PD and affective disorders.

Author Contributions

Conceptualization, A.V.L. and R.B.K.; writing—original draft preparation, A.A.M., R.B.K., J.A.T., V.A.Z. and D.A.A.; writing—review and editing, A.B.V., T.N.F., R.R.G. and A.V.L.; supervision, R.R.G. and A.V.L.; project administration, A.B.V. and T.N.F.; funding acquisition, R.R.G. and A.V.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 22-75-10131; ABV and RRG are supported by the project ID: 94030300 of the St. Petersburg State University, St. Petersburg, Russia.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data sharing not applicable.

Acknowledgments

The authors thank the Russian Science Foundation and St. Petersburg State University for their funding for this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fedosova, N.U.; Habeck, M.; Nissen, P. Structure and Function of Na,K-ATPase-The Sodium-Potassium Pump. Compr. Physiol. 2021, 12, 2659–2679. [Google Scholar] [CrossRef]
  2. Sundaram, S.M.; Safina, D.; Ehrkamp, A.; Faissner, A.; Heumann, R.; Dietzel, I.D. Differential expression patterns of sodium potassium ATPase alpha and beta subunit isoforms in mouse brain during postnatal development. Neurochem. Int. 2019, 128, 163–174. [Google Scholar] [CrossRef]
  3. Holm, T.H.; Lykke-Hartmann, K. Insights into the Pathology of the α3 Na+/K+-ATPase Ion Pump in Neurological Disorders; Lessons from Animal Models. Front. Physiol. 2016, 7, 209. [Google Scholar] [CrossRef] [Green Version]
  4. Boscia, F.; Begum, G.; Pignataro, G.; Sirabella, R.; Cuomo, O.; Casamassa, A.; Sun, D.; Annunziato, L. Glial Na(+) -dependent ion transporters in pathophysiological conditions. Glia 2016, 64, 1677–1697. [Google Scholar] [CrossRef] [Green Version]
  5. Heinzen, E.L.; Arzimanoglou, A.; Brashear, A.; Clapcote, S.J.; Gurrieri, F.; Goldstein, D.B.; Jóhannesson, S.H.; Mikati, M.A.; Neville, B.; Nicole, S.; et al. Distinct neurological disorders with ATP1A3 mutations. Lancet Neurol. 2014, 13, 503–514. [Google Scholar] [CrossRef] [Green Version]
  6. Shrivastava, A.N.; Redeker, V.; Fritz, N.; Pieri, L.; Almeida, L.G.; Spolidoro, M.; Liebmann, T.; Bousset, L.; Renner, M.; Léna, C.; et al. α-synuclein assemblies sequester neuronal α3-Na+/K+-ATPase and impair Na+ gradient. EMBO J. 2015, 34, 2408–2423. [Google Scholar] [CrossRef] [Green Version]
  7. Khan, F.H.; Sen, T.; Chakrabarti, S. Dopamine oxidation products inhibit Na+, K+-ATPase activity in crude synaptosomal-mitochondrial fraction from rat brain. Free Radic. Res. 2003, 37, 597–601. [Google Scholar] [CrossRef]
  8. de Lores Arnaiz, G.R.; Ordieres, M.G.L. Brain Na(+), K(+)-ATPase Activity In Aging and Disease. Int. J. Biomed. Sci. 2014, 10, 85–102. Available online: https://www.ncbi.nlm.nih.gov/pubmed/25018677 (accessed on 15 April 2023).
  9. Hamlyn, J.M.; Blaustein, M.P.; Bova, S.; DuCharme, D.W.; Harris, D.W.; Mandel, F.; Mathews, W.R.; Ludens, J.H. Identification and characterization of a ouabain-like compound from human plasma. Proc. Natl. Acad. Sci. USA 1991, 88, 6259–6263. [Google Scholar] [CrossRef] [Green Version]
  10. Hamlyn, J.M.; Ringel, R.; Schaeffer, J.; Levinson, P.D.; Hamilton, B.P.; Kowarski, A.A.; Blaustein, M.P. A circulating inhibitor of (Na+ + K+)ATPase associated with essential hypertension. Nature 1982, 300, 650–652. [Google Scholar] [CrossRef]
  11. Komiyama, Y.; Dong, X.H.; Nishimura, N.; Masaki, H.; Yoshika, M.; Masuda, M.; Takahashi, H. A novel endogenous digitalis, telocinobufagin, exhibits elevated plasma levels in patients with terminal renal failure. Clin. Biochem. 2005, 38, 36–45. [Google Scholar] [CrossRef]
  12. Schneider, R.; Wray, V.; Nimtz, M.; Lehmann, W.D.; Kirch, U.; Antolovic, R.; Schoner, W. Bovine adrenals contain, in addition to ouabain, a second inhibitor of the sodium pump. J. Biol. Chem. 1998, 273, 784–792. [Google Scholar] [CrossRef] [Green Version]
  13. Tymiak, A.A.; Norman, J.A.; Bolgar, M.; DiDonato, G.C.; Lee, H.; Parker, W.L.; Lo, L.C.; Berova, N.; Nakanishi, K.; Haber, E. Physicochemical characterization of a ouabain isomer isolated from bovine hypothalamus. Proc. Natl. Acad. Sci. USA 1993, 90, 8189–8193. [Google Scholar] [CrossRef] [Green Version]
  14. Pavlovic, D. Endogenous cardiotonic steroids and cardiovascular disease, where to next? Cell Calcium. 2020, 86, 102156. [Google Scholar] [CrossRef]
  15. Keller, S.; Frishman, W.H. Neuropsychiatric effects of cardiovascular drug therapy. Cardiol. Rev. 2003, 11, 73–93. [Google Scholar] [CrossRef]
  16. Lingrel, J.B. The physiological significance of the cardiotonic steroid/ouabain-binding site of the Na,K-ATPase. Annu. Rev. Physiol. 2010, 72, 395–412. [Google Scholar] [CrossRef] [Green Version]
  17. Azarias, G.; Kruusmägi, M.; Connor, S.; Akkuratov, E.E.; Liu, X.-L.; Lyons, D.; Brismar, H.; Broberger, C.; Aperia, A. A specific and essential role for Na,K-ATPase α3 in neurons co-expressing α1 and α3. J. Biol. Chem. 2013, 288, 2734–2743. [Google Scholar] [CrossRef] [Green Version]
  18. Santos, M.S.; Goncalves, P.P.; Carvalho, A.P. Effect of ouabain on the gamma-[3H] aminobutyric acid uptake and release in the absence of Ca (+)+ and K (+)-depolarization. J. Pharmacol. Exp. Ther. 1990, 253, 620–627. Available online: https://jpet.aspetjournals.org/content/253/2/620.short (accessed on 28 April 2023).
  19. Akkuratov, E.E.; Westin, L.; Vazquez-Juarez, E.; de Marothy, M.; Melnikova, A.K.; Blom, H.; Lindskog, M.; Brismar, H.; Aperia, A. Ouabain Modulates the Functional Interaction Between Na,K-ATPase and NMDA Receptor. Mol. Neurobiol. 2020, 57, 4018–4030. [Google Scholar] [CrossRef]
  20. Sibarov, D.A.; Bolshakov, A.E.; Abushik, P.A.; Krivoi, I.I.; Antonov, S.M. Na+,K+-ATPase functionally interacts with the plasma membrane Na+,Ca2+ exchanger to prevent Ca2+ overload and neuronal apoptosis in excitotoxic stress. J. Pharmacol. Exp. Ther. 2012, 343, 596–607. [Google Scholar] [CrossRef]
  21. Yuan, Z.; Cai, T.; Tian, J.; Ivanov, A.V.; Giovannucci, D.R.; Xie, Z. Na/K-ATPase tethers phospholipase C and IP3 receptor into a calcium-regulatory complex. Mol. Biol. Cell 2005, 16, 4034–4045. [Google Scholar] [CrossRef] [Green Version]
  22. Hodes, A.; Rosen, H.; Deutsch, J.; Lifschytz, T.; Einat, H.; Lichtstein, D. Endogenous cardiac steroids in animal models of mania. Bipolar. Disord. 2016, 18, 451–459. [Google Scholar] [CrossRef]
  23. Dobretsov, M.; Hayar, A.; Kockara, N.T.; Kozhemyakin, M.; Light, K.E.; Patyal, P.; Pierce, D.R.; Wight, P.A. A Transgenic Mouse Model to Selectively Identify α3 Na,K-ATPase Expressing Cells in the Nervous System. Neuroscience 2019, 398, 274–294. [Google Scholar] [CrossRef]
  24. Clapcote, S.J.; Duffy, S.; Xie, G.; Kirshenbaum, G.; Bechard, A.R.; Rodacker Schack, V.; Petersen, J.; Sinai, L.; Saab, B.J.; Lerch, J.P.; et al. Mutation I810N in the alpha3 isoform of Na+,K+-ATPase causes impairments in the sodium pump and hyperexcitability in the CNS. Proc. Natl. Acad. Sci. USA 2009, 106, 14085–14090. [Google Scholar] [CrossRef] [Green Version]
  25. Kirshenbaum, G.S.; Clapcote, S.J.; Duffy, S.; Burgess, C.R.; Petersen, J.; Jarowek, K.J.; Yücel, Y.H.; Cortez, M.A.; Snead, O.C.; Vilsen, B.; et al. Mania-like behavior induced by genetic dysfunction of the neuron-specific Na+,K+-ATPase α3 sodium pump. Proc. Natl. Acad. Sci. USA 2011, 108, 18144–18149. [Google Scholar] [CrossRef] [Green Version]
  26. Kirshenbaum, G.S.; Dawson, N.; Mullins, J.G.L.; Johnston, T.H.; Drinkhill, M.J.; Edwards, I.J.; Fox, S.H.; Pratt, J.A.; Brotchie, J.M.; Roder, J.C.; et al. Alternating hemiplegia of childhood-related neural and behavioural phenotypes in Na+,K+-ATPase α3 missense mutant mice. PLoS ONE 2013, 8, e60141. [Google Scholar] [CrossRef] [Green Version]
  27. Hunanyan, A.S.; Fainberg, N.A.; Linabarger, M.; Arehart, E.; Leonard, A.S.; Adil, S.M.; Helseth, A.R.; Swearingen, A.K.; Forbes, S.L.; Rodriguiz, R.M.; et al. Knock-in mouse model of alternating hemiplegia of childhood: Behavioral and electrophysiologic characterization. Epilepsia 2015, 56, 82–93. [Google Scholar] [CrossRef]
  28. Moseley, A.E.; Williams, M.T.; Schaefer, T.L.; Bohanan, C.S.; Neumann, J.C.; Behbehani, M.M.; Vorhees, C.V.; Lingrel, J.B. Deficiency in Na,K-ATPase α Isoform Genes Alters Spatial Learning, Motor Activity, and Anxiety in Mice. J. Neurosci. 2007, 27, 616–626. [Google Scholar] [CrossRef] [Green Version]
  29. DeAndrade, M.P.; Yokoi, F.; van Groen, T.; Lingrel, J.B.; Li, Y. Characterization of Atp1a3 mutant mice as a model of rapid-onset dystonia with parkinsonism. Behav. Brain Res. 2011, 216, 659–665. [Google Scholar] [CrossRef] [Green Version]
  30. Ikeda, K.; Satake, S.; Onaka, T.; Sugimoto, H.; Takeda, N.; Imoto, K.; Kawakami, K. Enhanced inhibitory neurotransmission in the cerebellar cortex of Atp1a3-deficient heterozygous mice. J. Physiol. 2013, 591, 3433–3449. [Google Scholar] [CrossRef]
  31. Sugimoto, H.; Ikeda, K.; Kawakami, K. Heterozygous mice deficient in Atp1a3 exhibit motor deficits by chronic restraint stress. Behav. Brain Res. 2014, 272, 100–110. [Google Scholar] [CrossRef]
  32. Kosobud, A.E.; Crabbe, J.C. Genetic correlations among inbred strain sensitivities to convulsions induced by 9 convulsant drugs. Brain Res. 1990, 526, 8–16. [Google Scholar] [CrossRef]
  33. Logan, R.W.; McClung, C.A. Animal models of bipolar mania: The past, present and future. Neuroscience 2016, 321, 163–188. [Google Scholar] [CrossRef] [Green Version]
  34. Harvey, A.G. Sleep and circadian rhythms in bipolar disorder: Seeking synchrony, harmony, and regulation. Am. J. Psychiatry 2008, 165, 820–829. [Google Scholar] [CrossRef] [Green Version]
  35. Anand, A.; Verhoeff, P.; Seneca, N.; Zoghbi, S.S.; Seibyl, J.P.; Charney, D.S.; Innis, R.B. Brain SPECT imaging of amphetamine-induced dopamine release in euthymic bipolar disorder patients. Am. J. Psychiatry 2000, 157, 1108–1114. [Google Scholar] [CrossRef]
  36. Timothy, J.W.S.; Klas, N.; Sanghani, H.R.; Al-Mansouri, T.; Hughes, A.T.L.; Kirshenbaum, G.S.; Brienza, V.; Belle, M.D.; Ralph, M.R.; Clapcote, S.J.; et al. Circadian Disruptions in the Myshkin Mouse Model of Mania Are Independent of Deficits in Suprachiasmatic Molecular Clock Function. Biol. Psychiatry 2018, 84, 827–837. [Google Scholar] [CrossRef] [Green Version]
  37. Calderon, D.P.; Khodakhah, K. Chapter 29—Modeling Dystonia-Parkinsonism. In Movement Disorders, 2nd ed.; LeDoux, M.S., Ed.; Academic Press: Boston, MA, USA, 2015; pp. 507–515. [Google Scholar] [CrossRef]
  38. Heinzen, E.L.; Swoboda, K.J.; Hitomi, Y.; Gurrieri, F.; Nicole, S.; de Vries, B.; Tiziano, F.D.; Fontaine, B.; Walley, N.M.; Heavin, S.; et al. De novo mutations in ATP1A3 cause alternating hemiplegia of childhood. Nat. Genet. 2012, 44, 1030–1034. [Google Scholar] [CrossRef] [Green Version]
  39. Bøttger, P.; Glerup, S.; Gesslein, B.; Illarionova, N.B.; Isaksen, T.J.; Heuck, A.; Clausen, B.H.; Füchtbauer, E.-M.; Gramsbergen, J.B.; Gunnarson, E.; et al. Glutamate-system defects behind psychiatric manifestations in a familial hemiplegic migraine type 2 disease-mutation mouse model. Sci. Rep. 2016, 6, 22047. [Google Scholar] [CrossRef] [Green Version]
  40. Ellman, D.G.; Isaksen, T.J.; Lund, M.C.; Dursun, S.; Wirenfeldt, M.; Jørgensen, L.H.; Lykke-Hartmann, K.; Lambertsen, K.L. The loss-of-function disease-mutation G301R in the Na+/K+-ATPase α2 isoform decreases lesion volume and improves functional outcome after acute spinal cord injury in mice. BMC Neurosci. 2017, 18, 66. [Google Scholar] [CrossRef] [Green Version]
  41. el-Mallakh, R.S.; Harrison, L.T.; Li, R.; Changaris, D.G.; Levy, R.S. An animal model for mania: Preliminary results. Prog. Neuropsychopharmacol. Biol. Psychiatry 1995, 19, 955–962. [Google Scholar] [CrossRef]
  42. Riegel, R.E.; Valvassori, S.S.; Elias, G.; Réus, G.Z.; Steckert, A.V.; de Souza, B.; Petronilho, F.; Gavioli, E.C.; Dal-Pizzol, F.; Quevedo, J. Animal model of mania induced by ouabain: Evidence of oxidative stress in submitochondrial particles of the rat brain. Neurochem. Int. 2009, 55, 491–495. [Google Scholar] [CrossRef]
  43. Valvassori, S.S.; Dal-Pont, G.C.; Resende, W.R.; Varela, R.B.; Lopes-Borges, J.; Cararo, J.H.; Quevedo, J. Validation of the animal model of bipolar disorder induced by Ouabain: Face, construct and predictive perspectives. Transl. Psychiatry 2019, 9, 158. [Google Scholar] [CrossRef] [Green Version]
  44. Valvassori, S.S.; Dal-Pont, G.C.; Resende, W.R.; Jornada, L.K.; Peterle, B.R.; Machado, A.G.; Farias, H.R.; de Souza, C.T.; Carvalho, A.F.; Quevedo, J. Lithium and valproate act on the GSK-3β signaling pathway to reverse manic-like behavior in an animal model of mania induced by ouabain. Neuropharmacology 2017, 117, 447–459. [Google Scholar] [CrossRef]
  45. Jornada, L.K.; Valvassori, S.S.; Steckert, A.V.; Moretti, M.; Mina, F.; Ferreira, C.L.; Arent, C.O.; Dal-Pizzol, F.; Quevedo, J. Lithium and valproate modulate antioxidant enzymes and prevent ouabain-induced oxidative damage in an animal model of mania. J. Psychiatr Res. 2011, 45, 162–168. [Google Scholar] [CrossRef]
  46. El-Mallakh, R.S.; Decker, S.; Morris, M.; Li, X.-P.; Huff, M.O.; El-Masri, M.A.; Levy, R.S. Efficacy of olanzapine and haloperidol in an animal model of mania. Prog. Neuropsychopharmacol. Biol. Psychiatry 2006, 30, 1261–1264. [Google Scholar] [CrossRef]
  47. Valvassori, S.S.; Dal-Pont, G.C.; Resende, W.R.; Varela, R.B.; Peterle, B.R.; Gava, F.F.; Mina, F.G.; Cararo, J.H.; Carvalho, A.F.; Quevedo, J. Lithium and Tamoxifen Modulate Behavior and Protein Kinase C Activity in the Animal Model of Mania Induced by Ouabain. Int. J. Neuropsychopharmacol. 2017, 20, 877–885. [Google Scholar] [CrossRef] [Green Version]
  48. Valvassori, S.S.; Dal-Pont, G.C.; Varela, R.B.; Resende, W.R.; Gava, F.F.; Mina, F.G.; Budni, J.; Quevedo, J. Ouabain induces memory impairment and alter the BDNF signaling pathway in an animal model of bipolar disorder: Cognitive and neurochemical alterations in BD model. J. Affect. Disord. 2021, 282, 1195–1202. [Google Scholar] [CrossRef]
  49. Valvassori, S.S.; Aguiar-Geraldo, J.M.; Possamai-Della, T.; da-Rosa, D.D.; Peper-Nascimento, J.; Cararo, J.H.; Quevedo, J. Depressive-like behavior accompanies neuroinflammation in an animal model of bipolar disorder symptoms induced by ouabain. Pharmacol. Biochem. Behav. 2022, 219, 173434. [Google Scholar] [CrossRef]
  50. Kurauchi, Y.; Yoshimaru, Y.; Kajiwara, Y.; Yamada, T.; Matsuda, K.; Hisatsune, A.; Seki, T.; Katsuki, H. Na+, K+-ATPase inhibition causes hyperactivity and impulsivity in mice via dopamine D2 receptor-mediated mechanism. Neurosci. Res. 2019, 146, 54–64. [Google Scholar] [CrossRef]
  51. Yu, H.S.; Kim, S.H.; Park, H.G.; Kim, Y.S.; Ahn, Y.M. Intracerebroventricular administration of ouabain, a Na/K-ATPase inhibitor, activates tyrosine hydroxylase through extracellular signal-regulated kinase in rat striatum. Neurochem. Int. 2011, 59, 779–786. [Google Scholar] [CrossRef]
  52. Yu, H.-S.; Kim, S.H.; Park, H.G.; Kim, Y.S.; Ahn, Y.M. Activation of Akt signaling in rat brain by intracerebroventricular injection of ouabain: A rat model for mania. Prog. Neuropsychopharmacol. Biol. Psychiatry 2010, 34, 888–894. [Google Scholar] [CrossRef]
  53. Li, R.; el-Mallakh, R.S.; Harrison, L.; Changaris, D.G.; Levy, R.S. Lithium prevents ouabain-induced behavioral changes. Toward an animal model for manic depression. Mol. Chem. Neuropathol. 1997, 31, 65–72. [Google Scholar] [CrossRef]
  54. Kim, S.H.; Yu, H.-S.; Park, H.G.; Ha, K.; Kim, Y.S.; Shin, S.Y.; Ahn, Y.M. Intracerebroventricular administration of ouabain, a Na/K-ATPase inhibitor, activates mTOR signal pathways and protein translation in the rat frontal cortex. Prog. Neuropsychopharmacol. Biol. Psychiatry 2013, 45, 73–82. [Google Scholar] [CrossRef]
  55. Lopachev, A.; Volnova, A.; Evdokimenko, A.; Abaimov, D.; Timoshina, Y.; Kazanskaya, R.; Lopacheva, O.; Deal, A.; Budygin, E.; Fedorova, T.; et al. Intracerebroventricular injection of ouabain causes mania-like behavior in mice through D2 receptor activation. Sci. Rep. 2019, 9, 15627. [Google Scholar] [CrossRef] [Green Version]
  56. Hodes, A.; Lifschytz, T.; Rosen, H.; Cohen Ben-Ami, H.; Lichtstein, D. Reduction in endogenous cardiac steroids protects the brain from oxidative stress in a mouse model of mania induced by amphetamine. Brain Res. Bull. 2018, 137, 356–362. [Google Scholar] [CrossRef]
  57. Shin, E.-J.; Nguyen, B.-T.; Jeong, J.H.; Hoai Nguyen, B.-C.; Tran, N.K.C.; Sharma, N.; Kim, D.-J.; Nah, S.-Y.; Lichtstein, D.; Nabeshima, T.; et al. Ouabain inhibitor rostafuroxin attenuates dextromethorphan-induced manic potential. Food Chem. Toxicol. 2021, 158, 112657. [Google Scholar] [CrossRef]
  58. Sun, Y.; Dong, Z.; Khodabakhsh, H.; Chatterjee, S.; Guo, S. Zebrafish chemical screening reveals the impairment of dopaminergic neuronal survival by cardiac glycosides. PLoS ONE 2012, 7, e35645. [Google Scholar] [CrossRef] [Green Version]
  59. Howarth, C.; Gleeson, P.; Attwell, D. Updated energy budgets for neural computation in the neocortex and cerebellum. J. Cereb. Blood Flow Metab. 2012, 32, 1222–1232. [Google Scholar] [CrossRef]
  60. Bogdanova, A.; Petrushanko, I.Y.; Hernansanz-Agustín, P.; Martínez-Ruiz, A. “Oxygen Sensing” by Na,K-ATPase: These Miraculous Thiols. Front. Physiol. 2016, 7, 314. [Google Scholar] [CrossRef] [Green Version]
  61. Lv, C.; Hong, T.; Yang, Z.; Zhang, Y.; Wang, L.; Dong, M.; Zhao, J.; Mu, J.; Meng, Y. Effect of Quercetin in the 1-Methyl-4-phenyl-1, 2, 3, 6-tetrahydropyridine-Induced Mouse Model of Parkinson’s Disease. Evid. Based. Complement Altern. Med. 2012, 2012, 928643. [Google Scholar] [CrossRef] [Green Version]
  62. Lin, K.-H.; Li, C.-Y.; Hsu, Y.-M.; Tsai, C.-H.; Tsai, F.-J.; Tang, C.-H.; Yang, J.-S.; Wang, Z.-H.; Yin, M.-C. Oridonin, A natural diterpenoid, protected NGF-differentiated PC12 cells against MPP+- and kainic acid-induced injury. Food Chem. Toxicol. 2019, 133, 110765. [Google Scholar] [CrossRef]
  63. Ilesanmi, O.B.; Akinmoladun, A.C.; Josiah, S.S.; Olaleye, M.T.; Akindahunsi, A.A. Modulation of key enzymes linked to Parkinsonism and neurologic disorders by Antiaris africana in rotenone-toxified rats. J. Basic Clin. Physiol. Pharmacol. 2019, 31. [Google Scholar] [CrossRef]
  64. Anusha, C.; Sumathi, T.; Joseph, L.D. Protective role of apigenin on rotenone induced rat model of Parkinson’s disease: Suppression of neuroinflammation and oxidative stress mediated apoptosis. Chem. Biol. Interact. 2017, 269, 67–79. [Google Scholar] [CrossRef]
  65. Khadrawy, Y.A.; Salem, A.M.; El-Shamy, K.A.; Ahmed, E.K.; Fadl, N.N.; Hosny, E.N. Neuroprotective and Therapeutic Effect of Caffeine on the Rat Model of Parkinson’s Disease Induced by Rotenone. J. Diet. Suppl. 2017, 14, 553–572. [Google Scholar] [CrossRef]
  66. Bonsi, P.; Calabresi, P.; De Persis, C.; Papa, M.; Centonze, D.; Bernardi, G.; Pisani, A. Early ionic and membrane potential changes caused by the pesticide rotenone in striatal cholinergic interneurons. Exp. Neurol. 2004, 185, 169–181. [Google Scholar] [CrossRef]
  67. Kinoshita, P.F.; Orellana, A.M.M.; Nakao, V.W.; de Souza Port’s, N.M.; Quintas, L.E.M.; Kawamoto, E.M.; Scavone, C. The Janus face of ouabain in Na+ /K+ -ATPase and calcium signalling in neurons. Br. J. Pharmacol. 2022, 179, 1512–1524. [Google Scholar] [CrossRef]
  68. Antunes, M.S.; Ladd, F.V.L.; Ladd, A.A.B.L.; Moreira, A.L.; Boeira, S.P.; Cattelan Souza, L. Hesperidin protects against behavioral alterations and loss of dopaminergic neurons in 6-OHDA-lesioned mice: The role of mitochondrial dysfunction and apoptosis. Metab. Brain Dis. 2021, 36, 153–167. [Google Scholar] [CrossRef]
  69. Del Fabbro, L.; Goes, A.R.; Jesse, C.R.; de Gomes, M.G.; Souza, L.C.; Ladd, F.V.L.; Ladd, A.A.L.; Arantes, R.V.N.; Simionato, A.R.; Oliveira, M.S.; et al. Chrysin protects against behavioral, cognitive and neurochemical alterations in a 6-hydroxydopamine model of Parkinson’s disease. Neurosci. Lett. 2019, 706, 158–163. [Google Scholar] [CrossRef]
  70. Therien, A.G.; Blostein, R. Mechanisms of sodium pump regulation. Am. J. Physiol. Cell Physiol. 2000, 279, C541–C566. [Google Scholar] [CrossRef] [Green Version]
  71. Holthouser, K.A.; Mandal, A.; Merchant, M.L.; Schelling, J.R.; Delamere, N.A.; Valdes, R.R., Jr.; Tyagi, S.C.; Lederer, E.D.; Khundmiri, S.J. Ouabain stimulates Na-K-ATPase through a sodium/hydrogen exchanger-1 (NHE-1)-dependent mechanism in human kidney proximal tubule cells. Am. J. Physiol. Renal. Physiol. 2010, 299, F77–F90. [Google Scholar] [CrossRef] [Green Version]
  72. Kanai, R.; Cornelius, F.; Ogawa, H.; Motoyama, K.; Vilsen, B.; Toyoshima, C. Binding of cardiotonic steroids to Na+,K+-ATPase in the E2P state. Proc. Natl. Acad. Sci. USA 2021, 118. [Google Scholar] [CrossRef]
  73. Orlov, S.N.; Tverskoi, A.M.; Sidorenko, S.V.; Smolyaninova, L.V.; Lopina, O.D.; Dulin, N.O.; Klimanova, E.A. Na,K-ATPase as a target for endogenous cardiotonic steroids: What’s the evidence? Genes Dis. 2021, 8, 259–271. [Google Scholar] [CrossRef]
  74. Xu, Y.; Marck, P.; Huang, M.; Xie, J.X.; Wang, T.; Shapiro, J.I.; Cai, L.; Feng, F.; Xie, Z. Biased Effect of Cardiotonic Steroids on Na/K-ATPase-Mediated Signal Transduction. Mol. Pharmacol. 2021, 99, 217–225. [Google Scholar] [CrossRef]
  75. Guzman, J.N.; Ilijic, E.; Yang, B.; Sanchez-Padilla, J.; Wokosin, D.; Galtieri, D.; Kondapalli, J.; Schumacker, P.T.; Surmeier, D.J. Systemic isradipine treatment diminishes calcium-dependent mitochondrial oxidant stress. J. Clin. Investig. 2018, 128, 2266–2280. [Google Scholar] [CrossRef] [Green Version]
  76. Guzman, J.N.; Sánchez-Padilla, J.; Chan, C.S.; Surmeier, D.J. Robust pacemaking in substantia nigra dopaminergic neurons. J. Neurosci. 2009, 29, 11011–11019. [Google Scholar] [CrossRef] [Green Version]
  77. Surmeier, D.J.; Guzman, J.N.; Sanchez-Padilla, J.; Schumacker, P.T. The role of calcium and mitochondrial oxidant stress in the loss of substantia nigra pars compacta dopaminergic neurons in Parkinson’s disease. Neuroscience 2011, 198, 221–231. [Google Scholar] [CrossRef] [Green Version]
  78. Zampese, E.; Wokosin, D.L.; Gonzalez-Rodriguez, P.; Guzman, J.N.; Tkatch, T.; Kondapalli, J.; Surmeier, W.C.; D’alessandro, K.B.; De Stefani, D.; Rizzuto, R.; et al. Ca2+ channels couple spiking to mitochondrial metabolism in substantia nigra dopaminergic neurons. Sci. Adv. 2022, 8, eabp8701. [Google Scholar] [CrossRef]
  79. Desfrere, L.; Karlsson, M.; Hiyoshi, H.; Malmersjö, S.; Nanou, E.; Estrada, M.; Miyakawa, A.; Lagercrantz, H.; El Manira, A.; Lal, M.; et al. Na,K-ATPase signal transduction triggers CREB activation and dendritic growth. Proc. Natl. Acad. Sci. USA 2009, 106, 2212–2217. [Google Scholar] [CrossRef] [Green Version]
  80. Orellana, A.M.; Leite, J.A.; Kinoshita, P.F.; Vasconcelos, A.R.; Andreotti, D.Z.; de Sá Lima, L.; Xavier, G.F.; Kawamoto, E.M.; Scavone, C. Ouabain increases neuronal branching in hippocampus and improves spatial memory. Neuropharmacology 2018, 140, 260–274. [Google Scholar] [CrossRef]
  81. de Sá Lima, L.; Kawamoto, E.M.; Munhoz, C.D.; Kinoshita, P.F.; Orellana, A.M.M.; Curi, R.; Rossoni, L.; Avellar, M.; Scavone, C. Ouabain activates NFκB through an NMDA signaling pathway in cultured cerebellar cells. Neuropharmacology 2013, 73, 327–336. [Google Scholar] [CrossRef]
  82. Veldhuis, W.B.; van der Stelt, M.; Delmas, F.; Gillet, B.; Veldink, G.A.; Vliegenthart, J.F.G.; Nicolay, K.; Bär, P.R. In vivo excitotoxicity induced by ouabain, a Na+/K+-ATPase inhibitor. J. Cereb. Blood Flow Metab. 2003, 23, 62–74. [Google Scholar] [CrossRef] [Green Version]
  83. Liu, J.; Lilly, M.N.; Shapiro, J.I. Targeting Na/K-ATPase Signaling: A New Approach to Control Oxidative Stress. Curr. Pharm. Des. 2018, 24, 359–364. [Google Scholar] [CrossRef]
  84. Yan, Y.; Wang, J.; Chaudhry, M.A.; Nie, Y.; Sun, S.; Carmon, J.; Shah, P.T.; Bai, F.; Pratt, R.; Brickman, C.; et al. Metabolic Syndrome and Salt-Sensitive Hypertension in Polygenic Obese TALLYHO/JngJ Mice: Role of Na/K-ATPase Signaling. Int. J. Mol. Sci. 2019, 20, 3495. [Google Scholar] [CrossRef] [Green Version]
  85. Brown, J.M.; Hanson, G.R.; Fleckenstein, A.E. Cocaine-induced increases in vesicular dopamine uptake: Role of dopamine receptors. J. Pharmacol. Exp. Ther. 2001, 298, 1150–1153. Available online: https://www.ncbi.nlm.nih.gov/pubmed/11504813 (accessed on 17 April 2023).
  86. Truong, J.G.; Rau, K.S.; Hanson, G.R.; Fleckenstein, A.E. Pramipexole increases vesicular dopamine uptake: Implications for treatment of Parkinson’s neurodegeneration. Eur. J. Pharmacol. 2003, 474, 223–226. [Google Scholar] [CrossRef]
  87. Truong, J.G.; Hanson, G.R.; Fleckenstein, A.E. Apomorphine increases vesicular monoamine transporter-2 function: Implications for neurodegeneration. Eur. J. Pharmacol. 2004, 492, 143–147. [Google Scholar] [CrossRef]
  88. Jones, S.R.; Gainetdinov, R.R.; Hu, X.T.; Cooper, D.C.; Wightman, R.M.; White, F.J.; Caron, M.G. Loss of autoreceptor functions in mice lacking the dopamine transporter. Nat. Neurosci. 1999, 2, 649–655. [Google Scholar] [CrossRef]
  89. Gainetdinov, R.R.; Jones, S.R.; Fumagalli, F.; Wightman, R.M.; Caron, M.G. Re-evaluation of the role of the dopamine transporter in dopamine system homeostasis. Brain Res. Brain Res. Rev. 1998, 26, 148–153. [Google Scholar] [CrossRef]
  90. Goldstein, D.S.; Sullivan, P.; Holmes, C.; Miller, G.W.; Alter, S.; Strong, R.; Mash, D.C.; Kopin, I.J.; Sharabi, Y. Determinants of buildup of the toxic dopamine metabolite DOPAL in Parkinson’s disease. J. Neurochem. 2013, 126, 591–603. [Google Scholar] [CrossRef] [Green Version]
  91. Ng, J.; Zhen, J.; Meyer, E.; Erreger, K.; Li, Y.; Kakar, N.; Ahmad, J.; Thiele, H.; Kubisch, C.; Rider, N.L.; et al. Dopamine transporter deficiency syndrome: Phenotypic spectrum from infancy to adulthood. Brain 2014, 137, 1107–1119. [Google Scholar] [CrossRef] [Green Version]
  92. Kim, S.H.; Yu, H.-S.; Park, H.G.; Jeon, W.J.; Song, J.Y.; Kang, U.G.; Ahn, Y.M.; Lee, Y.H.; Kim, Y.S. Dose-dependent effect of intracerebroventricular injection of ouabain on the phosphorylation of the MEK1/2-ERK1/2-p90RSK pathway in the rat brain related to locomotor activity. Prog Neuropsychopharmacol Biol. Psychiatry 2008, 32, 1637–1642. [Google Scholar] [CrossRef]
  93. Lopachev, A.V.; Lopacheva, O.M.; Osipova, E.A.; Vladychenskaya, E.A.; Smolyaninova, L.V.; Fedorova, T.N.; Koroleva, O.V.; Akkuratov, E.E. Ouabain-induced changes in MAP kinase phosphorylation in primary culture of rat cerebellar cells. Cell Biochem. Funct. 2016, 34, 367–377. [Google Scholar] [CrossRef]
  94. Ivanova, M.A.; Kokorina, A.D.; Timofeeva, P.D.; Karelina, T.V.; Abushik, P.A.; Stepanenko, J.D.; Sibarov, D.A.; Antonov, S.M. Calcium Export from Neurons and Multi-Kinase Signaling Cascades Contribute to Ouabain Neuroprotection in Hyperhomocysteinemia. Biomolecules 2020, 10, 1104. [Google Scholar] [CrossRef]
  95. Salvatore, M.F.; Waymire, J.C.; Haycock, J.W. Depolarization-stimulated catecholamine biosynthesis: Involvement of protein kinases and tyrosine hydroxylase phosphorylation sites in situ. J. Neurochem. 2001, 79, 349–360. [Google Scholar] [CrossRef]
  96. Daubner, S.C.; Le, T.; Wang, S. Tyrosine hydroxylase and regulation of dopamine synthesis. Arch. Biochem. Biophys. 2011, 508, 1–12. [Google Scholar] [CrossRef] [Green Version]
  97. Kim, K.S.; Tinti, C.; Song, B.; Cubells, J.F.; Joh, T.H. Cyclic AMP-dependent protein kinase regulates basal and cyclic AMP-stimulated but not phorbol ester-stimulated transcription of the tyrosine hydroxylase gene. J. Neurochem. 1994, 63, 834–842. [Google Scholar] [CrossRef]
  98. Yamada, S.; Yokoo, H.; Nishi, S. Effects of N-ethylmaleimide on dopamine release in the rat striatum after repeated treatment with methamphetamine. Eur. J. Pharmacol. 1994, 257, 243–248. [Google Scholar] [CrossRef]
  99. Wanderoy, M.H.; Westlind-Danielsson, A.; Ahlenius, S. Dopamine D2 receptor upregulation in rat neostriatum following in vivo infusion of forskolin. Neuroreport 1997, 8, 2971–2976. [Google Scholar] [CrossRef]
  100. Page, G.; Barc-Pain, S.; Pontcharraud, R.; Cante, A.; Piriou, A.; Barrier, L. The up-regulation of the striatal dopamine transporter’s activity by cAMP is PKA-, CaMK II- and phosphatase-dependent. Neurochem. Int. 2004, 45, 627–632. [Google Scholar] [CrossRef]
  101. Yao, J.; Erickson, J.D.; Hersh, L.B. Protein kinase A affects trafficking of the vesicular monoamine transporters in PC12 cells. Traffic 2004, 5, 1006–1016. [Google Scholar] [CrossRef]
  102. Vecchio, L.M.; Sullivan, P.; Dunn, A.R.; Bermejo, M.K.; Fu, R.; Masoud, S.T.; Gregersen, E.; Urs, N.M.; Nazari, R.; Jensen, P.H.; et al. Enhanced tyrosine hydroxylase activity induces oxidative stress, causes accumulation of autotoxic catecholamine metabolites, and augments amphetamine effects in vivo. J. Neurochem. 2021, 158, 960–979. [Google Scholar] [CrossRef]
  103. Nguyen, P.-T.; Dang, D.-K.; Tran, H.-Q.; Shin, E.-J.; Jeong, J.H.; Nah, S.-Y.; Cho, M.C.; Lee, Y.S.; Jang, C.-G.; Kim, H.-C. Methiopropamine, a methamphetamine analogue, produces neurotoxicity via dopamine receptors. Chem. Biol. Interact. 2019, 305, 134–147. [Google Scholar] [CrossRef]
  104. Beaulieu, J.-M.; Sotnikova, T.D.; Marion, S.; Lefkowitz, R.J.; Gainetdinov, R.R.; Caron, M.G. An Akt/beta-arrestin 2/PP2A signaling complex mediates dopaminergic neurotransmission and behavior. Cell 2005, 122, 261–273. [Google Scholar] [CrossRef] [Green Version]
  105. Beaulieu, J.-M.; Del’guidice, T.; Sotnikova, T.D.; Lemasson, M.; Gainetdinov, R.R. Beyond cAMP: The Regulation of Akt and GSK3 by Dopamine Receptors. Front. Mol. Neurosci. 2011, 4, 38. [Google Scholar] [CrossRef] [Green Version]
  106. Gianferrara, T.; Cescon, E.; Grieco, I.; Spalluto, G.; Federico, S. Glycogen Synthase Kinase 3β Involvement in Neuroinflammation and Neurodegenerative Diseases. Curr. Med. Chem. 2022, 29, 4631–4697. [Google Scholar] [CrossRef]
  107. García-Yagüe, Á.J.; Lastres-Becker, I.; Stefanis, L.; Vassilatis, D.K.; Cuadrado, A. α-Synuclein Induces the GSK-3-Mediated Phosphorylation and Degradation of NURR1 and Loss of Dopaminergic Hallmarks. Mol. Neurobiol. 2021, 58, 6697–6711. [Google Scholar] [CrossRef]
  108. Hermanson, E.; Joseph, B.; Castro, D.; Lindqvist, E.; Aarnisalo, P.; Wallén, A.; Benoit, G.; Hengerer, B.; Olson, L.; Perlmann, T. Nurr1 regulates dopamine synthesis and storage in MN9D dopamine cells. Exp. Cell Res. 2003, 288, 324–334. [Google Scholar] [CrossRef]
  109. Chen, P.; Gu, Z.; Liu, W.; Yan, Z. Glycogen synthase kinase 3 regulates N-methyl-D-aspartate receptor channel trafficking and function in cortical neurons. Mol. Pharmacol. 2007, 72, 40–51. [Google Scholar] [CrossRef] [Green Version]
  110. Lopachev, A.V.; Lopacheva, O.M.; Nikiforova, K.A.; Filimonov, I.S.; Fedorova, T.N.; Akkuratov, E.E. Comparative Action of Cardiotonic Steroids on Intracellular Processes in Rat Cortical Neurons. Biochemistry 2018, 83, 140–151. [Google Scholar] [CrossRef]
  111. Cao, Q.; Qin, L.; Huang, F.; Wang, X.; Yang, L.; Shi, H.; Wu, H.; Zhang, B.; Chen, Z.; Wu, X. Amentoflavone protects dopaminergic neurons in MPTP-induced Parkinson’s disease model mice through PI3K/Akt and ERK signaling pathways. Toxicol. Appl. Pharmacol. 2017, 319, 80–90. [Google Scholar] [CrossRef]
  112. Wang, Z.; Liu, J.; Chen, S.; Wang, Y.; Cao, L.; Zhang, Y.; Kang, W.; Li, H.; Gui, Y.; Chen, S.; et al. DJ-1 modulates the expression of Cu/Zn-superoxide dismutase-1 through the Erk1/2-Elk1 pathway in neuroprotection. Ann. Neurol. 2011, 70, 591–599. [Google Scholar] [CrossRef]
  113. Zhu, J.-H.; Guo, F.; Shelburne, J.; Watkins, S.; Chu, C.T. Localization of phosphorylated ERK/MAP kinases to mitochondria and autophagosomes in Lewy body diseases. Brain Pathol. 2003, 13, 473–481. [Google Scholar] [CrossRef] [Green Version]
  114. Monick, M.M.; Powers, L.S.; Barrett, C.W.; Hinde, S.; Ashare, A.; Groskreutz, D.J.; Nyunoya, T.; Coleman, M.; Spitz, D.R.; Hunninghake, G.W. Constitutive ERK MAPK activity regulates macrophage ATP production and mitochondrial integrity. J. Immunol. 2008, 180, 7485–7496. [Google Scholar] [CrossRef] [Green Version]
  115. Park, K.H.; Shin, K.S.; Zhao, T.T.; Park, H.J.; Lee, K.E.; Lee, M.K. L-DOPA modulates cell viability through the ERK-c-Jun system in PC12 and dopaminergic neuronal cells. Neuropharmacology 2016, 101, 87–97. [Google Scholar] [CrossRef]
  116. Santini, E.; Valjent, E.; Usiello, A.; Carta, M.; Borgkvist, A.; Girault, J.-A.; Herve, D.; Greengard, P.; Fisone, G. Critical involvement of cAMP/DARPP-32 and extracellular signal-regulated protein kinase signaling in L-DOPA-induced dyskinesia. J. Neurosci. 2007, 27, 6995–7005. [Google Scholar] [CrossRef] [Green Version]
  117. Valjent, E.; Pascoli, V.; Svenningsson, P.; Paul, S.; Enslen, H.; Corvol, J.-C.; Stipanovich, A.; Caboche, J.; Lombroso, P.J.; Nairn, A.C.; et al. Regulation of a protein phosphatase cascade allows convergent dopamine and glutamate signals to activate ERK in the striatum. Proc. Natl. Acad. Sci. USA 2005, 102, 491–496. [Google Scholar] [CrossRef] [Green Version]
  118. Bhat, N.R.; Zhang, P. Hydrogen peroxide activation of multiple mitogen-activated protein kinases in an oligodendrocyte cell line: Role of extracellular signal-regulated kinase in hydrogen peroxide-induced cell death. J. Neurochem. 1999, 72, 112–119. [Google Scholar] [CrossRef]
  119. Canals, S.; Casarejos, M.J.; de Bernardo, S.; Solano, R.M.; Mena, M.A. Selective and persistent activation of extracellular signal-regulated protein kinase by nitric oxide in glial cells induces neuronal degeneration in glutathione-depleted midbrain cultures. Mol. Cell Neurosci. 2003, 24, 1012–1026. [Google Scholar] [CrossRef]
  120. Chen, J.; Rusnak, M.; Lombroso, P.J.; Sidhu, A. Dopamine promotes striatal neuronal apoptotic death via ERK signaling cascades. Eur. J. Neurosci. 2009, 29, 287–306. [Google Scholar] [CrossRef] [Green Version]
  121. Aizman, O.; Brismar, H.; Uhlén, P.; Zettergren, E.; Levey, A.I.; Forssberg, H.; Greengard, P.; Aperia, A. Anatomical and physiological evidence for D1 and D2 dopamine receptor colocalization in neostriatal neurons. Nat. Neurosci. 2000, 3, 226–230. [Google Scholar] [CrossRef]
  122. Hazelwood, L.A.; Free, R.B.; Cabrera, D.M.; Skinbjerg, M.; Sibley, D.R. Reciprocal modulation of function between the D1 and D2 dopamine receptors and the Na+,K+-ATPase. J. Biol. Chem. 2008, 283, 36441–36453. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Potential consequences that NKA dysfunction, whether from high CTS or other pathological conditions, may have on dopaminergic signaling. Green arrows represent downstream activation, while red arrows represent inhibitory processes. Short black arrows to the right of a given element denote increased activation or increase in concentration (pointing up), or decreased activation or decrease in concentration (pointing down).
Figure 1. Potential consequences that NKA dysfunction, whether from high CTS or other pathological conditions, may have on dopaminergic signaling. Green arrows represent downstream activation, while red arrows represent inhibitory processes. Short black arrows to the right of a given element denote increased activation or increase in concentration (pointing up), or decreased activation or decrease in concentration (pointing down).
Biomedicines 11 01820 g001
Figure 2. OS caused by non-inhibitory concentrations of CTS, mediated via the Src-ERK1/2 pathway (A); possible consequences of chronic elevation of endogenous (non-inhibiting NKA) CTS concentrations in dopaminergic system neurons (B). Green arrows represent downstream activation, while red arrows represent inhibitory processes. Short black arrows to the right of a given element denote increased activation or increase in concentration (pointing up), or decreased activation or decrease in concentration (pointing down).
Figure 2. OS caused by non-inhibitory concentrations of CTS, mediated via the Src-ERK1/2 pathway (A); possible consequences of chronic elevation of endogenous (non-inhibiting NKA) CTS concentrations in dopaminergic system neurons (B). Green arrows represent downstream activation, while red arrows represent inhibitory processes. Short black arrows to the right of a given element denote increased activation or increase in concentration (pointing up), or decreased activation or decrease in concentration (pointing down).
Biomedicines 11 01820 g002
Table 1. ATP1A3 genetically modified models.
Table 1. ATP1A3 genetically modified models.
ModelSymptoms of Affective DisordersSymptoms of Neurological DisordersIn Vivo Electrophysiology DataChanges in Dopamine LevelsReferences
1.1
Myk/+
Mania:
Hyperactivity
Sleep disturbances
Dysregulated circadian rhythm
Tendency to engage in high-risk behavior
Increased sensitivity to amphetamine
Decreased anxiety
High impulsivity
Lower spatial memory
Tremor
Impaired gait
--[24,25,26]
1.2
Mashl+/−
Mania:
Hyperactivity
Increased excitability
Decreased anxiety
High impulsivity
Lower spatial memory
Tremor
Impaired gait
High excitability,
prolonged arousal after a threshold stimulus
-[27]
1.3
NKA1A3tm1Ling
Mania:
Hyperactivity
Increased sensitivity to amphetamine
Decreased anxiety
Impulsivity
Low habituation
Depression:
Anhedonia
Despair-like behavior
Increased anxiety
Impaired learning and memory
Decreased socialization
--Mania:
Not different from wild type
Depression:
Negative correlation with vertical activity
[25,28,29]
1.4
Atp1a3tm2Kwk/+
Mania:
Hyperactivity
Impulsivity
Lower spatial memory
Impaired gait
Symptoms similar to RDP
--[30]
[31]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Markina, A.A.; Kazanskaya, R.B.; Timoshina, J.A.; Zavialov, V.A.; Abaimov, D.A.; Volnova, A.B.; Fedorova, T.N.; Gainetdinov, R.R.; Lopachev, A.V. Na+,K+-ATPase and Cardiotonic Steroids in Models of Dopaminergic System Pathologies. Biomedicines 2023, 11, 1820. https://doi.org/10.3390/biomedicines11071820

AMA Style

Markina AA, Kazanskaya RB, Timoshina JA, Zavialov VA, Abaimov DA, Volnova AB, Fedorova TN, Gainetdinov RR, Lopachev AV. Na+,K+-ATPase and Cardiotonic Steroids in Models of Dopaminergic System Pathologies. Biomedicines. 2023; 11(7):1820. https://doi.org/10.3390/biomedicines11071820

Chicago/Turabian Style

Markina, Alisa A., Rogneda B. Kazanskaya, Julia A. Timoshina, Vladislav A. Zavialov, Denis A. Abaimov, Anna B. Volnova, Tatiana N. Fedorova, Raul R. Gainetdinov, and Alexander V. Lopachev. 2023. "Na+,K+-ATPase and Cardiotonic Steroids in Models of Dopaminergic System Pathologies" Biomedicines 11, no. 7: 1820. https://doi.org/10.3390/biomedicines11071820

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