Entry - *182350 - ATPase, Na+/K+ TRANSPORTING, ALPHA-3 POLYPEPTIDE; ATP1A3 - OMIM
* 182350

ATPase, Na+/K+ TRANSPORTING, ALPHA-3 POLYPEPTIDE; ATP1A3


Alternative titles; symbols

SODIUM-POTASSIUM-ATPase, ALPHA-3 POLYPEPTIDE
ATPase, Na+/K+, ALPHA III


HGNC Approved Gene Symbol: ATP1A3

Cytogenetic location: 19q13.2     Genomic coordinates (GRCh38): 19:41,966,582-41,994,230 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
19q13.2 Alternating hemiplegia of childhood 2 614820 AD 3
CAPOS syndrome 601338 AD 3
Developmental and epileptic encephalopathy 99 619606 AD 3
Dystonia-12 128235 AD 3

TEXT

Description

The ATP1A3 gene encodes the alpha-3 catalytic subunit of the Na+/K(+)-ATPase transmembrane ion pump. The ATP1A3 isoform is exclusively expressed in neurons of various brain regions, including the basal ganglia, hippocampus, and cerebellum (summary by Rosewich et al., 2012).

Na+/K(+)-ATPases are heterooligomers of a catalytic alpha subunit, such as ATP1A3, and a glycosylated beta subunit. Na+/K(+)-ATPases catalyze ATP-driven exchange of 3 intracellular Na+ ions for 2 extracellular K+ ions across the plasma membrane. This exchange involves 2 major conformational changes, ATP hydrolysis and transitory phosphorylation of the ATPase, and temporary occlusion of 3 Na+ ions, followed by 2 K+ ions, within the ATPase in each conformation (summary by Rodacker et al., 2006).


Cloning and Expression

Ovchinnikov et al. (1988) cloned ATP1A3, which they called alpha III, from a human brain cDNA library. The deduced 1,013-amino acid protein has a calculated molecular mass of 111.7 kD. Alpha III is predicted to have an N-terminal signal sequence, 7 transmembrane segments, and a cytoplasmic ATPase catalytic site.

By immunohistochemical analysis of rat and mouse retina, Wetzel et al. (1999) found that alpha-3 was expressed in photoreceptors, horizontal cells, bipolar cells, amacrine cells, and ganglion cells. In photoreceptors, alpha-3 was expressed in rod inner segments, as well as in cell somas in the outer nuclear layer and their presumptive terminals in the outer plexiform layer. Alpha-3 colocalized with beta-2 (ATP1B2; 182331) in photoreceptors and with beta-1 (ATP1B1; 182330) in horizontal cells. Various Na,K-ATPase isoforms exhibited marked changes in distribution during mouse postnatal development. Alpha-3 was detected in undifferentiated photoreceptor somas at birth, and was later targeted to inner segments and synaptic terminals.

Using in situ hybridization, Sugimoto et al. (2014) detected widespread Atp1a3 expression in mouse central nervous system, including expression in almost all brain regions and major neuronal cells.

Allocco et al. (2019) found expression of the Atp1a3 gene in neurons in all cortical layers of embryonic mouse brain. It was identified in differentiated neurons at the cortical plate and in neural stem cells at the ventricular zone lining the lateral ventricles; immunostaining was also observed in choroid plexus endothelial cells.


Gene Structure

Ovchinnikov et al. (1988) determined that the ATP1A3 gene spans about 25 kb and that its protein-coding region includes 23 exons.


Mapping

By Southern analysis of DNA from panels of rodent/human somatic cell hybrid lines, Yang-Feng et al. (1988) mapped the ATP1A3 gene to chromosome 19q12-q13.2. Harley et al. (1988) concluded that the order is qter--DM--APOC2--ATP1A3--cen.

Gross (2021) mapped the ATP1A3 gene to chromosome 19q13.2 based on an alignment of the ATP1A3 sequence (GenBank BC009282) with the genomic sequence (GRCh38).


Gene Function

Agrin (AGRN; 103320) mediates accumulation of acetylcholine receptors at the developing neuromuscular junction through its interaction with MUSK (601296), and it has also been implicated in brain development. Through biochemical studies, Hilgenberg et al. (2006) found that agrin bound Atp1a3 in mouse cortical neurons. Immunohistochemical analysis showed that Atp1a3 colocalized with agrin-binding sites at synapses. Agrin inhibited Atp1a3 activity, resulting in membrane depolarization and increased action potential frequency in mouse cortical neurons in culture and acute slice. An agrin fragment that acted as a competitive antagonist depressed action potential frequency, indicating that endogenous agrin regulates native Atp1a3 function. Hilgenberg et al. (2006) concluded that agrin regulates activity-dependent processes in neurons through its interaction with ATP1A3.


Molecular Genetics

Dystonia 12

In 7 unrelated families with rapid-onset dystonia parkinsonism, or dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified 6 different heterozygous mutations in the ATP1A3 gene (182350.0001-182350.0006). Functional expression studies and structural analysis suggested that the mutations impaired enzyme activity or stability.

Anselm et al. (2009) and Blanco-Arias et al. (2009) reported de novo heterozygous ATP1A3 mutations (182350.0007 and 182350.0008, respectively) in patients with DYT12.

In Drosophila, Kaneko et al. (2014) identified a dominant missense mutation (A617T) in the calcium ATPase Serca gene (see SERCA2 (ATP2A2); 108740) that conferred temperature-sensitive motor uncoordination in a gain-of-function manner. The homologous residue is conserved by different type II P-type ATPases, including ATP1A2 (182340). Introduction of an R751Q mutation in the Drosophila Serca gene also caused a temperature-sensitive uncoordination phenotype. The corresponding residue in human SERCA2, ATP1A2, and ATP1A3 is mutated in the human diseases Darier disease (124200), FHM2 (602481), and dystonia-12, respectively. Cellular expression of Drosophila A617T resulted in temperature-induced decreased levels of stored calcium compared to wildtype, whereas cellular expression of R751Q elicited depletion of stored calcium even without heating. These calcium changes were due to leakage through the mutant channel pores that overwhelmed the pumping capacity of the cell. Similar results occurred after transfection of these mutations, as well as other disease-causing mutations that affected different parts of the protein, into mouse cells. Kaneko et al. (2014) concluded that ionic leakage is a gain-of-function mechanism that underlies a variety of dominant type II P-type ATPase-related diseases.

Alternating Hemiplegia of Childhood 2

In 82 of 105 patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified 19 different heterozygous mutations in the ATP1A3 gene (see, e.g., 182350.0009-182350.0012). The first mutations were identified through exome sequencing of affected individuals. Thirteen of the 18 mutations observed in sporadic cases were confirmed to occur de novo. Since it was possible that some variants represented polymorphisms, Heinzen et al. (2012) estimated that mutations in the ATP1A3 gene may be responsible for up to 74% of patients with sporadic, typical AHC. Several mutations were recurrent, and some occurred within hypermutable sequences. All patients had infantile onset of hemiplegia attacks, usually associated with episodes of quadriparesis, abnormal eye movements, autonomic signs, seizures, dystonia, ataxia, chorea, and developmental delay. Transfection of several of the mutations in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased. In contrast, transfection of DYT12-associated mutations resulted in decreased protein levels as well as decreased activity. The report expanded the spectrum of phenotypes associated with mutations in the ATP1A3 gene.

Simultaneously and independent to the report of Heinzen et al. (2012), Rosewich et al. (2012) identified de novo heterozygous mutations in the ATP1A3 gene (see, e.g., 182350.0009; 182350.0010; 182350.0015-182350.0017) in 24 unrelated patients with AHC2. Mutations in the first 3 patients were found by whole-exome sequencing of 3 affected child-parent trios, and subsequent mutations were found by direct Sanger sequencing of the ATP1A3 gene in additional patients. There were 2 main recurrent mutations: D801N (182350.0009) and E815K (182350.0010), found in 9 (38%) and 7 (29%) patients, respectively, suggesting mutational hotspots. None of the mutations resulted in a truncated protein, although there was 1 splice site mutation (182350.0017). Functional studies of the variants and studies of patients cells were not performed. Rosewich et al. (2012) noted the phenotypic overlap between AHC2 and DYT12.

In 45 (95.7%) of 47 Chinese children with typical AHC2, Yang et al. (2014) identified 19 different heterozygous missense mutations in the ATP1A3 gene. Three mutation hotspots, D801N (182350.0009), E815K (182350.0010), and G947R (182350.0012 and 182350.0013), were detected in 14 (31.1%), 9 (20.0%), and 7 (15.6%) ATP1A3-positive patients, respectively. Except for 1 patient who had inherited a mutation from her affected mother, all patients for whom parental DNA was available were found to have de novo mutations. Heterozygous ATP1A3 mutations were also found in 4 additional Chinese patients with atypical AHC2 who had onset of the disorder after 18 months of age. The initial mutations were found by whole-exome sequencing of several patients, and the subsequent mutations were found by direct sequencing of the ATP1A3 gene in a larger cohort. Presence of the E815K mutation was associated with epilepsy. A review of published disease-associated ATP1A3 mutations suggested that mutations associated with AHC2 were predominantly located in the transmembrane domain, whereas those associated with DYT12 had no location bias. Molecular modeling of the variants identified 2 statistically significant molecular features, solvent accessibility and distance to metal ion, that distinguished disease-associated mutations from neutral variants. In vitro functional studies were not performed on any of the variants.

Cerebellar Ataxia, Areflexia, Pes Cavus, Optic Atrophy, And Sensorineural Hearing Loss

In 10 patients from 3 unrelated families with cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss (CAPOS; 601338), Demos et al. (2014) identified the same heterozygous missense mutation in the ATP1A3 gene (E818K; 182350.0014).

Developmental And Epileptic Encephalopathy 99

In 16 patients from 15 families with developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified 14 heterozygous mutations in the ATP1A2 gene (see, e.g., 182350.0019-182350.0021). The mutations occurred de novo in all except for a mother and son pair (patients 14 and 15). Mutations were mostly missense, with a few small in-frame deletions or insertions. All occurred at conserved residues, and none were present in the gnomAD database. In vitro functional expression studies showed that all of the mutations caused variable functional defects in the Na+/(K+)ATPase. Variants with more severe functional deficits were associated with a more severe phenotype. The findings were consistent with a loss-of-function effect. Vetro et al. (2021) estimated that about 12% of ATP1A3 mutations may be associated with DEE. Polymicrogyria was estimated to occur in about 5.5% of patients with ATP1A3 mutations.

Associations Pending Confirmation

For discussion of a possible association between autosomal recessive congenital hydrocephalus due to aqueductal stenosis (see, e.g., 236635) and variation in the ATP1A3 gene, see 182350.0022 and 182350.0023.


Genotype/Phenotype Correlations

Rosewich et al. (2014) identified 16 patients with AHC2 and 3 with DYT12 confirmed by genetic analysis. A review of the clinical and molecular findings of these patients and of 164 previously reported patients with ATP1A3 mutations indicated that although mutations were distributed over almost all protein domains, those affecting transmembrane and functional domains tended to be associated with AHC2 as the more severe phenotype. The majority of mutations associated with AHC2 were located in exons 17 and 18, whereas those associated with DYT12 were located in exons 8 and 14; however, there was overlap, particularly in exon 17. Clinical analysis suggested that the 2 disorders represent a continuous phenotypic spectrum, with intermediate phenotypes combining criteria of both conditions. Shared clinical characteristics of both disorders include asymmetric movement disorder, rostrocaudal gradient of involvement with prominent bulbar symptoms, and triggering of symptoms by different stressors.

Using a formulated questionnaire, Panagiotakaki et al. (2015) assessed clinical data from 155 patients with AHC, including 132 confirmed to have ATP1A3 mutations by genetic analysis. Among those with AHC2, the most frequent mutations were D801N (in 43%), E815K (in 16%), and G947R (182350.0012 and 182350.0013, which were considered together) (in 11%). E815K was associated with a severe phenotype, with greater intellectual and motor disability; D801N appeared to confer a milder phenotype; and G947R correlated with the most favorable prognosis. For those with epilepsy, the age at seizure onset was earlier for patients with the E815K or G947R mutations than for those with the D801N mutation. Several mutational clusters within the gene were identified.


Animal Model

Ashmore et al. (2009) identified 6 different EMS-induced missense mutations in the Atp1a2 and Atp1a3 genes in Drosophila. All mutations resulted in reduced respiration activity consistent with a loss of ATPase function and a hypomorphic effect. Different mutant strains exhibited some abnormalities, including progressive temperature-dependent paralysis, progressive stress-sensitive paralysis, and decreased locomotor activity in response to startle, suggesting a decrease in maximal locomotion capacity. Neuromuscular studies showed allele-specific pathology, including brain vacuoles and myopathology, and biochemical studies showed decreased metabolic rates. An unexpected finding was the some mutant strains had increased longevity, which was not related to caloric restriction. Low doses of ouabain showed a similar effect on longevity in control groups. Ashmore et al. (2009) suggested that these findings may be relevant for studying the pathogenesis of FHM2 and DYT12 (128235).

In a mouse mutagenesis screen, Clapcote et al. (2009) identified a mutant mouse strain, Myshkin (Myk), that showed autosomal dominant complex and partial and secondarily generalized seizures, a reduced threshold for seizures in hippocampal slices, and neuronal degeneration in the hippocampus. Heterozygous mice were also smaller than wildtype, and homozygosity for the mutation resulted in perinatal death. Positional cloning and functional analysis identified a heterozygous ile810-to-asn (I810N) substitution in the Atp1a3 gene as responsible for the phenotype. In vitro cellular functional expression studies showed that the I810N substitution disrupted enzymatic function. The mutant protein had 42% reduced activity in mouse brain, and the phenotype could be rescued by transfection of the wildtype gene, consistent with loss of function as a pathogenic mechanism. The findings indicated the importance of ion homeostasis in maintaining normal neuronal excitability.

Doganli et al. (2013) found that the 2 ATP1A3 orthologs in zebrafish, Atp1a3a and Atp1a3b, were expressed in distinct but overlapping sets of brain structures. Morpholino-mediated knockdown (KD) of either Atp1a3a or Atp1a3b caused dilation of brain ventricles. Dilation in Atp1a3a KD embryos was not rescued by coinjection of Atp1a3b, and Atp1a3b KD embryos were not rescued by coinjection of Atp1a3a. Atp1a3a KD also caused depolarization of the resting membrane potential of Rohon-Beard neurons, which are mechanosensory neurons localized in the dorsal spinal cord. Atp1a3a KD and Atp1a3b KD embryos showed abnormal but distinct spontaneous motility and responses to tactile stimuli. Proteomic analysis revealed that Atp1a3a KD and Atp1a3b KD altered expression of overlapping sets of genes.

Sugimoto et al. (2014) found that Atp1a3 +/- mice showed shortened stride length and decreased motor strength in hanging box test following restraint stress compared with wildtype mice. Male and female Atp1a3 +/- mice showed some differences in the effects of restraint stress.

Balestrini et al. (2020) studied electrocardiogram abnormalities in a mouse model with heterozygosity for an Atp1a3 D801N knock-in mutation (Mashl +/-). Compared to 15 wildtype mice, 3 Mashl +/- mice had an increased heart rate, prolonged QRS and PR interval, and a longer QTc interval. After intraamygdala injection of kainic acid to induce seizures, all mice had elevation of JT intervals. One of the Mashl +/- mice had a period with JT-segment depression, and 2 Mashl +/- mice died from atrioventricular block.


ALLELIC VARIANTS ( 23 Selected Examples):

.0001 DYSTONIA 12

ATP1A3, THR613MET
  
RCV000013772...

In a sporadic patient (Linazasoro et al., 2002) and affected members of a second family (Zaremba et al., 2004) with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 1838C-T transition in the ATP1A3 gene, resulting in a thr613-to-met (T613M) substitution in a highly conserved residue near the phosphorylation domain on the cytoplasmic face of the protein. The mutation was not identified in 500 northern European control chromosomes.

Brashear et al. (2007) identified the T613M mutation in a family with DYT12 reported by Pittock et al. (2000).

Rodacker et al. (2006) noted that T613 is universally conserved among Na+/K(+)-ATPases, H+/K(+)-ATPases, and Ca(2+)-ATPases. Using rat Atp1a1 (182310) for technical reasons, they presented evidence that the T613M substitution in ATP1A3 alters the affinity of ATP1A3 for Na+ and ATP and alters the conformation equilibrium in favor of the potassium-bound form.


.0002 DYSTONIA 12

ATP1A3, ILE274THR
  
RCV000013773

In a patient with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 821T-C transition in the ATP1A3 gene, resulting in an ile274-to-thr (I274T) substitution in a highly conserved residue in the transmembrane domain of the protein. The mutation was not identified in 500 northern European control chromosomes. The patient had disease onset at age 37 years.


.0003 DYSTONIA 12

ATP1A3, GLU277LYS
  
RCV000013774

In a patient with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 829G-A transition in exon 8 of the ATP1A3 gene, resulting in a glu277-to-lys (E277K) substitution in a highly conserved residue in the transmembrane domain of the protein. The mutation was not identified in 500 northern European control chromosomes. The patient had disease onset at age 20 years.

Tarsy et al. (2010) identified the E277K mutation in a 29-year-old woman of African Caribbean descent with DYT12. She had onset at age 26 years of weakness and flexion of the left hand and ankle, which progressed rapidly over the next few years to become frank dystonia of the left arm and bulbar symptoms, including dysphagia, laryngeal dysfunction with task-specific dysphonia, and oropharyngeal dysmotility. She also had mild parkinsonism, with hypomimia and wide-based gait. Treatment with oral trihexyphenidyl and botulinum injection into selected laryngeal muscles resulted in clinical improvement.


.0004 DYSTONIA 12

ATP1A3, ILE758SER
  
RCV000013775...

In 12 affected members of a family with dystonia-12 (DYT12; 128235) reported by Dobyns et al. (1993), de Carvalho Aguiar et al. (2004) identified a heterozygous 2273T-G transversion in the ATP1A3 gene, resulting in an ile758-to-ser (I758S) substitution in a highly conserved residue in the transmembrane domain of the protein. The mutation was not identified in 500 northern European control chromosomes.


.0005 DYSTONIA 12

ATP1A3, PHE780LEU
  
RCV000013776

In 2 affected members of a family with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 2338T-C transition in the ATP1A3 gene, resulting in a phe780-to-leu (F780L) substitution in a highly conserved residue in the transmembrane region of the protein close to the extracellular surface. The mutation was not identified in 500 northern European control chromosomes.

Rodacker et al. (2006) noted that F780 is fully conserved in all known Na+/K(+)-ATPases from different species. Using rat Atp1a1 for technical reasons, they presented evidence that the F780L substitution reduced the affinity of ATP1A3 for Na+ and reduced the V(max) for ATP-dependent phosphorylation. The mutation was not expected to affect either the affinity of ATP1A3 for K+ nor the K(+)-induced dephosphorylation event.


.0006 DYSTONIA 12

ATP1A3, ASP801TYR
  
RCV000013777

In 4 affected members of a family with dystonia-12 (DYT12; 128235) reported by Brashear et al. (1997), de Carvalho Aguiar et al. (2004) identified a heterozygous 2401G-T transversion in the ATP1A3 gene, resulting in an asp801-to-tyr (D801Y) substitution in a highly conserved residue in the transmembrane region of the protein. The mutation was not identified in 500 northern European control chromosomes.


.0007 DYSTONIA 12

ALTERNATING HEMIPLEGIA OF CHILDHOOD 2, INCLUDED
ATP1A3, ASP923ASN
  
RCV000013778...

In a boy with early-onset dystonia-12 (DYT12; 128235) at age 4 years, Anselm et al. (2009) identified a heterozygous de novo 2767G-A transition in exon 20 of the ATP1A3 gene, resulting in an asp923-to-asn (D923N) substitution. The mutation was not found in 338 Caucasian control chromosomes. The substitution was predicted to occur in a residue buried in the membrane close to the ion-binding residue gln920, suggesting that it may affect enzyme activity. He was born of an unaffected Caucasian father and Chinese mother. The onset of dystonia was abrupt, occurring after mild trauma to the forehead. He developed mutism, eye convergence, and inability to walk, which later evolved into severe dystonia, severe dysarthria, and drooling. The condition stabilized over several months, and he showed mild improvement over the next 8 years. About a year after onset, he developed unusual episodes of flaccidity lasting for hours, later replaced by shorter episodes of stiffness. Treatment with L-DOPA was not effective. At the time of the report, he had bulbar symptoms, striking oromotor dystonia with inability to speak or swallow well, and apraxia.

Yang et al. (2014) identified a de novo heterozygous D923N mutation in a Chinese boy with atypical alternating hemiplegia of childhood-2 (AHC2; 614820). The phenotype was considered atypical due to relatively late onset of symptoms at age 30 months. Otherwise, the boy had typical features of quadriplegia as well as abnormal eye movements, dystonia, and developmental delay.


.0008 DYSTONIA 12

ATP1A3, 3-BP DUP, 3191TAC
  
RCV000148335

In a 16-year-old female with dystonia-12 (DYT12; 128235), Blanco-Arias et al. (2009) reported a de novo heterozygous 3-bp duplication (3191dupTAC) in exon 23 of the ATP1A3 gene, resulting in duplication of tyr1013, the C-terminal amino acid of the protein before the stop codon. The mutation was not found in either parent, her brother, or in 218 control individuals. HeLa cells expressing the mutant protein showed decreased survival in response to ouabain challenge, but no defect was detected in protein expression or plasma membrane targeting. Functional analysis demonstrated a drastic 40- to 50-fold reduction in Na+ affinity in the mutant. Blanco-Arias et al. (2009) suggested a crucial role for the C terminus of the alpha-subunit in the function of the Na+/K(+)-ATPase and emphasized a key impact of Na+ affinity in the pathophysiology of DYT12.


.0009 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, ASP801ASN
   RCV000030749...

In 36 of 95 unrelated patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a heterozygous 2401G-A transition in the ATP1A3 gene, resulting in an asp801-to-asn (D801N) substitution in the sixth transmembrane domain. The mutation was demonstrated to occur de novo in cases where parental material was available. All patients had infantile onset of hemiplegia attacks, usually associated with episodes of quadriparesis, abnormal eye movements, autonomic signs, seizures, dystonia, ataxia, chorea, and developmental delay. Transfection of the mutation in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased. Evaluation of the crystal structure of the protein predicted that the D801N substitution would prevent the binding of potassium ions to the pump.

Rosewich et al. (2012) identified a de novo heterozygous D801N mutation in 9 (38%) of 24 AHC2 patients. D801N occurs in the functionally conserved C-terminal cation-transporting ATPase domain and the P-type ATPase domain that is also a transmembrane domain. Functional studies of the variant and studies of patient cells were not performed.

Yang et al. (2014) identified a de novo heterozygous D801N mutation in 14 unrelated Chinese children with AHC2. All had typical features of the disorder, including abnormal eye movements and developmental delay, but none had seizures.


.0010 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, GLU815LYS
  
RCV000030750...

In 19 patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a heterozygous 2443G-A transition in the ATP1A3 gene, resulting in a glu815-to-lys (E815K) substitution in the sixth transmembrane domain. The mutation was shown to occur de novo in all patients whose parents were available for study. Transfection of the mutation in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased.

Rosewich et al. (2012) identified a de novo heterozygous E815K mutation in 7 (29%) of 24 AHC2 patients. E815K occurs in the functionally conserved C-terminal cation-transporting ATPase domain and the P-type ATPase domain that is also a transmembrane domain. Functional studies of the variant and studies of patient cells were not performed.

Yang et al. (2014) identified a de novo heterozygous E815K mutation in 9 unrelated Chinese children with AHC2. Seven of the patients had epilepsy.


.0011 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, SER811PRO
  
RCV000030751...

In 4 unrelated patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a de novo heterozygous 2431T-C transition in the ATP1A3 gene, resulting in a ser811-to-pro (S811P) substitution in the sixth transmembrane domain. Transfection of the mutation in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased.


.0012 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, NT2839G-A, GLY947ARG
  
RCV000030752...

In 5 patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a heterozygous 2839G-A transition in the ATP1A3 gene, resulting in a gly947-to-arg (G947R) substitution in the ninth transmembrane domain. The mutation was shown to occur de novo in all patients whose parents were available for study.

Yang et al. (2014) identified a de novo heterozygous G947R mutation in 5 unrelated Chinese children with AHC2. This same amino acid substitution can also result from a c.2839G-C transversion (182350.0013).


.0013 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, NT2839G-C, GLY947ARG
  
RCV000128466...

In 3 unrelated Chinese children with alternating hemiplegia of childhood-2 (AHC2; 614820), Yang et al. (2014) identified a heterozygous c.2839G-C transversion in exon 21 of the ATP1A3 gene, resulting in a gly947-to-arg (G947R) substitution at a highly conserved residue. The mutation occurred de novo in 2 of the patients and was inherited from an affected mother in the third patient. This same amino acid substitution can also result from a c.2839G-A transition (182350.0012). One of the patients had so-called atypical AHC2, with onset at 30 months of age. The mutation was not found in the 1000 Genomes Project or Exome Sequencing Project databases, or in 100 normal controls.


.0014 CEREBELLAR ATAXIA, AREFLEXIA, PES CAVUS, OPTIC ATROPHY, AND SENSORINEURAL HEARING LOSS

ATP1A3, GLU818LYS
  
RCV000144250...

In 10 patients from 3 unrelated families with cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss (CAPOS; 601338), including the original family reported by Nicolaides et al. (1996), Demos et al. (2014) identified a heterozygous c.2452G-A transition in the ATP1A3 gene, resulting in a glu818-to-lys (E818K) substitution at a highly conserved residue. The mutation, which was found by whole-exome sequencing of 2 of the families and confirmed by Sanger sequencing, was filtered against the dbSNP (builds 129 and 130) and 1000 Genomes Project databases and was not found in 1,834 controls. The mutation occurred de novo in the oldest affected generation of 1 family, but haplotype analysis could not rule out the possibility of a remote relationship between the other 2 families. All families were of Caucasian European descent. Functional studies of the E818K variant were not performed, but Demos et al. (2014) postulated a gain-of-function effect.

In a German boy with CAPOS, Rosewich et al. (2014) identified a de novo heterozygous E818K mutation in the ATP1A3 gene. Functional studies were not performed.

Tranebjaerg et al. (2018) reported that residue 818 of ATP1A3 is located at the cytoplasmic side of transmembrane helix-6, where it forms a salt bridge with the backbone carbonyl of arg930, a residue that stabilizes the C terminus. Tranebjaerg et al. (2018) expressed ATP1A3 with the E818K mutation in Xenopus laevis oocytes. Electrophysiologic analysis showed that the mutation disrupted the C terminus, caused opening of the C-terminal structure of ATP1A3, and affected sodium binding to and release from the binding site in the molecule. Molecular dynamic simulations confirmed that E818K opened the C-terminal pathway, allowing rapid entry of water molecules toward the ion-binding site.


.0015 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, ASP923TYR
  
RCV000148329...

In a 19-year-old man with alternating hemiplegia of childhood-2 (AHC2; 614820), Rosewich et al. (2012) identified a de novo heterozygous c.2767G-T transversion (c.2767G-T, NM_152296.4) in exon 20 of the ATP1A3 gene, resulting in an asp923-to-tyr (D923Y) substitution at a highly conserved residue in the C-terminal cation ATPase domain. Functional studies of the mutation and studies of patient cells were not performed. The authors noted that a different substitution at this same residue (D923N; 182350.0007) has been identified in patients with dystonia-12 (DYT12; 128235).


.0016 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, ILE274ASN
  
RCV000148305

In a 17-year-old girl with alternating hemiplegia of childhood-2 (AHC2; 614820), Rosewich et al. (2012) identified a de novo heterozygous c.821T-A transversion (c.821T-A, NM_152296.4) in exon 8 of the ATP1A3 gene, resulting in an ile274-to-asn (I274N) substitution at a highly conserved residue in the E1-E2 ATPase domain. Functional studies of the mutation and studies of patient cells were not performed. The authors noted that a different substitution at this same residue (I274T; 182350.0002) has been identified in patients with dystonia-12 (DYT12; 128235).


.0017 ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, IVS18DS, G-A, +1
  
RCV000148326...

In a 24-year-old woman with alternating hemiplegia of childhood-2 (AHC2; 614820), Rosewich et al. (2012) identified a de novo heterozygous G-to-A transition in intron 18 of the ATP1A3 gene (c.2542+1G-A, NM_152296.4), predicted to result in exon skipping. Functional studies of the mutation and studies of patient cells were not performed.


.0018 DYSTONIA 12

ATP1A3, GLY316SER
  
RCV000210848...

In a 26-year-old man with dystonia-12 (DYT12; 128235), Sweadner et al. (2016) identified a de novo heterozygous c.946G-A transition (c.946G-A, NM_152296.3) in the ATP1A3 gene, resulting in a gly316-to-ser (G316S) substitution in the highly conserved fourth transmembrane domain and close to an ion binding pocket. The mutation was found by exome sequencing and confirmed by Sanger sequencing. In vitro functional studies showed that the mutation resulted in a growth defect when expressed in HEK293 cells, consistent with impaired Na/K-ATPase function. The patient had a somewhat unusual phenotype, presenting at age 19 with rapidly progressive ataxia and dysarthria and tremor, resulting in loss of independent ambulation, and minimal dystonia. Exome sequencing showed that the patient also carried a de novo heterozygous missense E482K variant in the UBQLN4 gene (605440), which may have played a role in the prominent cerebellar ataxia and cerebellar atrophy observed in this patient; functional studies of the UBQLN4 variant were not performed.


.0019 DEVELOPMENTAL AND EPILEPTIC ENCEPHALOPATHY 99

ATP1A3, LEU292ARG
  
RCV001777181

In a 2-month-old infant (patient 7) with lethal developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified a de novo heterozygous c.875T-G transversion (c.875T-G, NM_152296.4) in the ATP1A3 gene, resulting in a leu292-to-arg (L292R) substitution at a conserved residue. The mutation was not present in the gnomAD database. In vitro studies showed that the mutation was unable to support growth and survival of COS1 cells in culture and interfered with Na+ and K+ affinity, resulting in nearly absent Na+/(K+)ATPase activity, consistent with a loss-of-function effect. The patient presented at birth with migrating focal seizures that evolved to almost continuous seizure activity with status epilepticus, resulting in death at 2 months of age.


.0020 DEVELOPMENTAL AND EPILEPTIC ENCEPHALOPATHY 99

ATP1A3, GLY316VAL
  
RCV001777182

In a 6-year-old girl (patient 8) with developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified a de novo heterozygous c.947G-T transversion (c.947G-T, NM_152296.4) in the ATP1A3 gene, resulting in a gly316-to-val (G316V) substitution at a conserved residue. The mutation was not present in the gnomAD database. In vitro studies showed that the mutation was unable to support growth and survival of COS1 cells in culture and interfered with Na+ and K+ affinity, resulting in nearly absent Na+/(K+)ATPase activity, consistent with a loss-of-function effect. The patient had onset of migrating focal and severe generalized seizures at 4 years of age.


.0021 DEVELOPMENTAL AND EPILEPTIC ENCEPHALOPATHY 99

ATP1A3, SER361PRO
  
RCV001777183

In a 7-year-old girl (patient 9) with developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified a de novo heterozygous c.1081T-C transition (c.1081T-C, NM_152296.4) in the ATP1A3 gene, resulting in a ser361-to-pro (S361P) substitution at a conserved residue. The mutation was not present in the gnomAD database. In vitro studies showed that the mutation was unable to support growth and survival of COS1 cells in culture with decreased phosphorylation activity and nearly absent Na+/(K+)ATPase activity, consistent with a loss-of-function effect. The patient had onset of focal temporal seizures at 5 months of age.


.0022 VARIANT OF UNKNOWN SIGNIFICANCE

ATP1A3, ARG19CYS
  
RCV000803350...

This variant is classified as a variant of unknown significance because its contribution to congenital hydrocephalus due to aqueductal stenosis (see, e.g., 236635) has not been confirmed.

In a 23-year-old Caucasian woman with congenital hydrocephalus due to aqueductal stenosis, Allocco et al. (2019) identified compound heterozygous missense variants in the ATP1A3 gene: a c.55G-A transition (c.55G-A, NM_152296) in exon 2, resulting in an arg19-to-cys (R19C) substitution inherited from the unaffected mother, and a c.1387G-A transition in exon 11, resulting in an arg463-to-cys (R463C; 182350.0023) substitution inherited from the unaffected father. The variants were identified by whole-exome sequencing and confirmed by Sanger sequencing. The R19C variant was present in the heterozygous state at a low frequency in gnomAD (6.4 x 10(-5)). Both variants occurred at conserved residues and were predicted to have disruptive effects on protein stability, although functional studies of the variants and studies of patient cells were not performed. The patient had multiple brain malformations, including open schizencephaly, type 1 Chiari malformation, and dysgenesis of the corpus callosum. Clinical details were limited, but she was noted to have learning disabilities. The authors postulated that dysregulation of neural development may be the pathogenesis of the disorder in this patient.


.0023 VARIANT OF UNKNOWN SIGNIFICANCE

ATP1A3, ARG463CYS
  
RCV000441666...

This variant is classified as a variant of unknown significance because its contribution to congenital hydrocephalus due to aqueductal stenosis (see, e.g., 236635) has not been confirmed.

For discussion of the c.1387G-A transition (c.1387G-A, NM_152296) in exon 11 of the ATP1A3 gene, resulting in an arg463-to-cys (R463C) substitution, that was found in compound heterozygous state in a patient with congenital hydrocephalus due to aqueductal stenosis, by Allocco et al. (2019) see 182350.0022.


REFERENCES

  1. Allocco, A. A., Jin, S. C., Duy, P. Q., Furey, C. G., Zeng, X., Dong, W., Nelson-Williams, C., Karimy, J. K., DeSpenza, T., Hao, L. T., Reeves, B., Haider, S., Gunel, M., Lifton, R. P., Kahle, K. T. Recessive inheritance of congenital hydrocephalus with other structural brain abnormalities caused by compound heterozygous mutations in ATP1A3. Front. Cellular Neurosci. 13: 425, 2019.

  2. Anselm, I. A., Sweadner, K. J., Gollamudi, S., Ozelius, L. J., Darras, B. T. Rapid-onset dystonia-parkinsonism in a child with a novel ATP1A3 gene mutation. Neurology 73: 400-401, 2009. [PubMed: 19652145, related citations] [Full Text]

  3. Ashmore, L. J., Hrizo, S. L., Paul, S. M., Van Voorhies, W. A., Beitel, G. J., Palladino, M. J. Novel mutations affecting the Na, K ATPase alpha model complex neurological diseases and implicate the sodium pump in increased longevity. Hum. Genet. 126: 431-447, 2009. [PubMed: 19455355, images, related citations] [Full Text]

  4. Balestrini, S., Mikati, M. A., Alvarez-Garcia-Roves, R., Carboni, M. Hunanyan, A. S., Kherallah, B., McLean, M., Prange, L., De Grandis, E., Gagliardi, A., Pisciotta, L., Stagnaro, M., and 42 others. Cardiac phenotype in ATP1A3-related syndromes: a multicenter study. Neurology 95: e2866-e2879, 2020. [PubMed: 32913013, images, related citations] [Full Text]

  5. Blanco-Arias, P., Einholm, A. P., Mamsa, H., Concheiro, C., Gutierrez-de-Teran, H., Romero, J., Toustrup-Jensen, M. S., Carracedo, A., Jen, J. C., Vilsen, B., Sobrido, M.-J. A C-terminal mutation of ATP1A3 underscores the crucial role of sodium affinity in the pathophysiology of rapid-onset dystonia-parkinsonism. Hum. Molec. Genet. 18: 2370-2377, 2009. [PubMed: 19351654, related citations] [Full Text]

  6. Brashear, A., DeLeon, D., Bressman, S. B., Thyagarajan, D., Farlow, M. R., Dobyns, W. B. Rapid-onset dystonia-parkinsonism in a second family. Neurology 48: 1066-1069, 1997. [PubMed: 9109901, related citations] [Full Text]

  7. Brashear, A., Dobyns, W. B., de Carvalho Aguiar, P., Borg, M., Frijns, C. J. M., Gollamudi, S., Green, A., Guimaraes, J., Haake, B. C., Klein, C., Linazasoro, G., Munchau, A., Raymond, D., Riley, D., Saunders-Pullman, R., Tijssen, M. A. J., Webb, D., Zaremba, J., Bressman, S. B., Ozelius, L. J. The phenotypic spectrum of rapid-onset dystonia-parkinsonism (RDP) and mutations in the ATP1A3 gene. Brain 130: 828-835, 2007. [PubMed: 17282997, related citations] [Full Text]

  8. 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., Minassian, B. A., Ackerley, C. A., Sled, J. G., Cortez, M. A., Henderson, J. T., Vilsen, B., Roder, J. C. Mutation I810N in the alpha-3 isoform of Na+,K(+)-ATPase causes impairments in the sodium pump and hyperexcitability in the CNS. Proc. Nat. Acad. Sci. 106: 14085-14090, 2009. [PubMed: 19666602, images, related citations] [Full Text]

  9. de Carvalho Aguiar, P., Sweadner, K. J., Penniston, J. T., Zaremba, J., Liu, L., Caton, M., Linazasoro, G., Borg, M., Tijssen, M. A. J., Bressman, S. B., Dobyns, W. B., Brashear, A., Ozelius, L. J. Mutations in the Na+/K(+)-ATPase alpha-3 gene ATP1A3 are associated with rapid-onset dystonia parkinsonism. Neuron 43: 169-175, 2004. [PubMed: 15260953, related citations] [Full Text]

  10. Demos, M. K., van Karnebeek, C. D. M., Ross, C. J. D., Adam, S., Shen, Y., Zhan, S. H., Shyr, C., Horvath, G., Suri, M., Fryer, A., Jones, S. J. M., Friedman, J. M., FORGE Canada Consortium. A novel recurrent mutation in ATP1A3 causes CAPOS syndrome. Orphanet J. Rare Dis. 9: 15, 2014. Note: Electronic Article. [PubMed: 24468074, related citations] [Full Text]

  11. Dobyns, W. B., Ozelius, L. J., Kramer, P. L., Brashear, A., Farlow, M. R., Perry, T. R., Walsh, L. E., Kasarskis, E. J., Butler, I. J., Breakefield, X. O. Rapid-onset dystonia-parkinsonism. Neurology 43: 2596-2602, 1993. [PubMed: 8255463, related citations] [Full Text]

  12. Doganli, C., Beck, H. C., Ribera, A. B., Oxvig, C., Lykke-Hartmann, K. Alpha-3-Na+/K(+)-ATPase deficiency causes brain ventricle dilation and abrupt embryonic motility in zebrafish. J. Biol. Chem. 288: 8862-8874, 2013. [PubMed: 23400780, images, related citations] [Full Text]

  13. Gross, M. B. Personal Communication. Baltimore, Md. 2/1/2021.

  14. Harley, H. G., Brook, J. D., Jackson, C. L., Glaser, T., Walsh, K. V., Sarfarazi, M., Kent, R., Lager, M., Koch, M., Harper, P. S., Levenson, R., Housman, D. E., Shaw, D. J. Localization of a human Na+,K(+)-ATPase alpha subunit gene to chromosome 19q12-q13.2 and linkage to the myotonic dystrophy locus. Genomics 3: 380-384, 1988. [PubMed: 2907504, related citations] [Full Text]

  15. 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., Panagiotakaki, E, European Alternating Hemiplegia of Childhood (AHC) Genetics Consortium, and 33 others. De novo mutations in ATP1A3 cause alternating hemiplegia of childhood. Nature Genet. 44: 1030-1034, 2012. [PubMed: 22842232, images, related citations] [Full Text]

  16. Hilgenberg, L. G. W., Su, H., Gu, H., O'Dowd, D. K., Smith, M. A. Alpha-3-Na(+)/K(+)-ATPase is a neuronal receptor for agrin. Cell 125: 359-369, 2006. [PubMed: 16630822, related citations] [Full Text]

  17. Kaneko, M., Desai, B. S., Cook, B. Ionic leakage underlies a gain-of-function effect of dominant disease mutations affecting diverse P-type ATPases. Nature Genet. 46: 144-151, 2014. [PubMed: 24336169, related citations] [Full Text]

  18. Linazasoro, G., Indakoetxea, B., Ruiz, J., Van Blercom, N., Lasa, A. Possible sporadic rapid-onset dystonia-parkinsonism. Mov. Disord. 17: 608-609, 2002. [PubMed: 12112218, related citations] [Full Text]

  19. Nicolaides, P., Appleton, R. E., Fryer, A. Cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss (CAPOS): a new syndrome. J. Med. Genet. 33: 419-421, 1996. [PubMed: 8733056, related citations] [Full Text]

  20. Ovchinnikov, Y. A., Monastyrskaya, G. S., Broude, N. E., Ushkaryov, Y. A., Melkov, A. M., Smirnov, Y. V., Malyshev, I. V., Allikmets, R. L., Kostina, M. B., Dulubova, I. E., Kiyatkin, N. I., Grishin, A. V., Modyanov, N. N., Sverdlov, E. D. Family of human Na+,K(+)-ATPase genes: structure of the gene for the catalytic subunit (alpha-III-form) and its relationship with structural features of the protein. FEBS Lett. 233: 87-94, 1988. [PubMed: 2838329, related citations] [Full Text]

  21. Panagiotakaki, E., De Grandis, E., Stagnaro, M., Heinzen, E. L., Fons, C., Sisodiya, S., de Vries, B., Goubau, C., Weckhuysen, S., Kemlink, D., Scheffer, I., Lesca, G., and 24 others. Clinical profile of patients with ATP1A3 mutations in alternating hemiplegia of childhood--a study of 155 patients. Orphanet J. Rare Dis. 10: 123, 2015. Note: Electronic Article. [PubMed: 26410222, images, related citations] [Full Text]

  22. Pittock, S. J., Joyce, C., O'Keane, V., Hugle, B., Hardiman, O., Brett, F., Green, A. J., Barton, D. E., King, M. D., Webb, D. W. Rapid-onset dystonia-parkinsonism: a clinical and genetic analysis of a new kindred. Neurology 55: 991-995, 2000. [PubMed: 11061257, related citations] [Full Text]

  23. Rodacker, V., Toustrup-Jensen, M., Vilsen, B. Mutations Phe785Leu and Thr618Met in Na+,K(+)-ATPase, associated with familial rapid-onset dystonia parkinsonism, interfere with Na+ interaction by distinct mechanisms. J. Biol. Chem. 281: 18539-18548, 2006. [PubMed: 16632466, related citations] [Full Text]

  24. Rosewich, H., Ohlenbusch, A., Huppke, P., Schlotawa, L., Baethmann, M., Carrilho, I., Fiori, S., Lourenco, C. M., Sawyer, S., Steinfeld, R., Gartner, J., Brockmann, K. The expanding clinical and genetic spectrum of ATP1A3-related disorders. Neurology 82: 945-955, 2014. [PubMed: 24523486, related citations] [Full Text]

  25. Rosewich, H., Thiele, H., Ohlenbusch, A., Maschke, U., Altmuller, J., Frommolt, P., Zirn, B., Ebinger, F., Siemes, H., Nurnberg, P., Brockmann, K., Gartner, J. Heterozygous de-novo mutations in ATP1A3 in patients with alternating hemiplegia of childhood: a whole-exome sequencing gene-identification study. Lancet Neurol. 11: 764-773, 2012. [PubMed: 22850527, related citations] [Full Text]

  26. Rosewich, H., Weise, D., Ohlenbusch, A., Gartner, J., Brockmann, K. Phenotypic overlap of alternating hemiplegia of childhood and CAPOS syndrome. Neurology 83: 861-863, 2014. [PubMed: 25056583, related citations] [Full Text]

  27. Sugimoto, H., Ikeda, K., Kawakami, K. Heterozygous mice deficient in Atp1a3 exhibit motor deficits by chronic restraint stress. Behav. Brain Res. 272: 100-110, 2014. [PubMed: 24983657, related citations] [Full Text]

  28. Sweadner, K. J., Toro, C., Whitlow, C. T., Snively, B. M., Cook, J. F., Ozelius, L. J., Markello, T. C., Brashear, A. ATP1A3 mutation in adult rapid-onset ataxia. PLoS One 11: e0151429, 2016. Note: Electronic Article. [PubMed: 26990090, images, related citations] [Full Text]

  29. Tarsy, D., Sweadner, K. J., Song, P. C. Case 17-2010: a 29-year-old woman with flexion of the left hand and foot and difficulty speaking. New Eng. J. Med. 362: 2213-2219, 2010. [PubMed: 20558373, related citations] [Full Text]

  30. Tranebjaerg, L., Strenzke, N., Lindholm, S., Rendtorff, N . D., Poulsen, H., Khandelia, H., Kopec, W., Lyngbye, T. J. B., Hamel, C., Delettre, C., Bocquet, B., Bille, M., and 19 others. The CAPOS mutation in ATP1A3 alters Na/K-ATPase function and results in auditory neuropathy which has implications for management. Hum. Genet. 137: 111-127, 2018. Note: Erratum Hum. Genet. 137: 279-280, 2018. [PubMed: 29305691, related citations] [Full Text]

  31. Vetro, A., Nielsen, H. N., Holm, R., Hevner, R. F., Parrini, E., Powis, Z., Moller, R. S., Bellan, C., Simonati, A., Lesca, G., Helbig, K. L., Palmer, E. E., and 18 others. ATP1A2- and ATP1A3-associated early profound epileptic encephalopathy and polymicrogyria. Brain 144: 1435-1450, 2021. [PubMed: 33880529, related citations] [Full Text]

  32. Wetzel, R. K., Arystarkhova, E., Sweadner, K. J. Cellular and subcellular specification of Na,K-ATPase alpha and beta isoforms in the postnatal development of mouse retina. J. Neurosci. 19: 9878-9889, 1999. [PubMed: 10559397, images, related citations] [Full Text]

  33. Yang, X., Gao, H., Zhang, J., Xu, X., Liu, X., Wu, X., Wei, L., Zhang, Y. ATP1A3 mutations and genotype-phenotype correlation of alternating hemiplegia of childhood in Chinese patients. PLoS One 9: e97274, 2014. Note: Electronic Article. [PubMed: 24842602, images, related citations] [Full Text]

  34. Yang-Feng, T. L., Schneider, J. W., Lindgren, V., Shull, M. M., Benz, E. J., Jr., Lingrel, J. B., Francke, U. Chromosomal localization of human Na+,K(+)-ATPase alpha- and beta-subunit genes. Genomics 2: 128-138, 1988. [PubMed: 2842249, related citations] [Full Text]

  35. Zaremba, J., Mierzewska, H., Lysiak, Z., Kramer, P., Ozelius, L. J., Brashear, A. Rapid-onset dystonia-parkinsonism: a fourth family consistent with linkage to chromosome 19q13. Mov. Disord. 19: 1506-1510, 2004. [PubMed: 15390049, related citations] [Full Text]


Cassandra L. Kniffin - updated : 11/04/2021
Matthew B. Gross - updated : 02/02/2021
Hilary J. Vernon - updated : 02/01/2021
Bao Lige - updated : 01/09/2019
Cassandra L. Kniffin - updated : 4/13/2016
Patricia A. Hartz - updated : 4/12/2016
Cassandra L. Kniffin - updated : 2/25/2015
Cassandra L. Kniffin - updated : 9/29/2014
Patricia A. Hartz - updated : 9/22/2014
Cassandra L. Kniffin - updated : 6/30/2014
Cassandra L. Kniffin - updated : 3/4/2014
Cassandra L. Kniffin - updated : 9/13/2012
Cassandra L. Kniffin - updated : 6/10/2010
Cassandra L. Kniffin - updated : 5/24/2010
George E. Tiller - updated : 3/30/2010
Matthew B. Gross - updated : 3/8/2010
Cassandra L. Kniffin - updated : 12/17/2009
Cassandra L. Kniffin - updated : 3/10/2005
Creation Date:
Victor A. McKusick : 12/1/1987
alopez : 11/10/2021
ckniffin : 11/04/2021
mgross : 02/02/2021
carol : 02/01/2021
alopez : 04/09/2019
mgross : 01/09/2019
carol : 04/07/2017
carol : 04/14/2016
carol : 4/14/2016
ckniffin : 4/13/2016
mgross : 4/12/2016
mgross : 4/12/2016
carol : 4/7/2016
carol : 2/26/2015
mcolton : 2/25/2015
ckniffin : 2/25/2015
alopez : 9/29/2014
ckniffin : 9/29/2014
mgross : 9/26/2014
mcolton : 9/22/2014
alopez : 7/3/2014
mcolton : 7/1/2014
ckniffin : 6/30/2014
carol : 3/5/2014
ckniffin : 3/4/2014
mcolton : 2/21/2014
carol : 9/14/2012
carol : 9/14/2012
terry : 9/13/2012
ckniffin : 9/13/2012
wwang : 6/11/2010
ckniffin : 6/10/2010
wwang : 5/25/2010
ckniffin : 5/24/2010
wwang : 3/31/2010
terry : 3/30/2010
wwang : 3/11/2010
mgross : 3/8/2010
wwang : 1/8/2010
ckniffin : 12/17/2009
wwang : 3/16/2005
wwang : 3/10/2005
ckniffin : 3/10/2005
carol : 10/31/2000
mgross : 7/21/1999
terry : 6/18/1998
supermim : 3/16/1992
supermim : 3/20/1990
ddp : 10/27/1989
root : 8/14/1989
root : 1/9/1989
root : 12/20/1988

* 182350

ATPase, Na+/K+ TRANSPORTING, ALPHA-3 POLYPEPTIDE; ATP1A3


Alternative titles; symbols

SODIUM-POTASSIUM-ATPase, ALPHA-3 POLYPEPTIDE
ATPase, Na+/K+, ALPHA III


HGNC Approved Gene Symbol: ATP1A3

SNOMEDCT: 702323008, 720634003;  


Cytogenetic location: 19q13.2     Genomic coordinates (GRCh38): 19:41,966,582-41,994,230 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
19q13.2 Alternating hemiplegia of childhood 2 614820 Autosomal dominant 3
CAPOS syndrome 601338 Autosomal dominant 3
Developmental and epileptic encephalopathy 99 619606 Autosomal dominant 3
Dystonia-12 128235 Autosomal dominant 3

TEXT

Description

The ATP1A3 gene encodes the alpha-3 catalytic subunit of the Na+/K(+)-ATPase transmembrane ion pump. The ATP1A3 isoform is exclusively expressed in neurons of various brain regions, including the basal ganglia, hippocampus, and cerebellum (summary by Rosewich et al., 2012).

Na+/K(+)-ATPases are heterooligomers of a catalytic alpha subunit, such as ATP1A3, and a glycosylated beta subunit. Na+/K(+)-ATPases catalyze ATP-driven exchange of 3 intracellular Na+ ions for 2 extracellular K+ ions across the plasma membrane. This exchange involves 2 major conformational changes, ATP hydrolysis and transitory phosphorylation of the ATPase, and temporary occlusion of 3 Na+ ions, followed by 2 K+ ions, within the ATPase in each conformation (summary by Rodacker et al., 2006).


Cloning and Expression

Ovchinnikov et al. (1988) cloned ATP1A3, which they called alpha III, from a human brain cDNA library. The deduced 1,013-amino acid protein has a calculated molecular mass of 111.7 kD. Alpha III is predicted to have an N-terminal signal sequence, 7 transmembrane segments, and a cytoplasmic ATPase catalytic site.

By immunohistochemical analysis of rat and mouse retina, Wetzel et al. (1999) found that alpha-3 was expressed in photoreceptors, horizontal cells, bipolar cells, amacrine cells, and ganglion cells. In photoreceptors, alpha-3 was expressed in rod inner segments, as well as in cell somas in the outer nuclear layer and their presumptive terminals in the outer plexiform layer. Alpha-3 colocalized with beta-2 (ATP1B2; 182331) in photoreceptors and with beta-1 (ATP1B1; 182330) in horizontal cells. Various Na,K-ATPase isoforms exhibited marked changes in distribution during mouse postnatal development. Alpha-3 was detected in undifferentiated photoreceptor somas at birth, and was later targeted to inner segments and synaptic terminals.

Using in situ hybridization, Sugimoto et al. (2014) detected widespread Atp1a3 expression in mouse central nervous system, including expression in almost all brain regions and major neuronal cells.

Allocco et al. (2019) found expression of the Atp1a3 gene in neurons in all cortical layers of embryonic mouse brain. It was identified in differentiated neurons at the cortical plate and in neural stem cells at the ventricular zone lining the lateral ventricles; immunostaining was also observed in choroid plexus endothelial cells.


Gene Structure

Ovchinnikov et al. (1988) determined that the ATP1A3 gene spans about 25 kb and that its protein-coding region includes 23 exons.


Mapping

By Southern analysis of DNA from panels of rodent/human somatic cell hybrid lines, Yang-Feng et al. (1988) mapped the ATP1A3 gene to chromosome 19q12-q13.2. Harley et al. (1988) concluded that the order is qter--DM--APOC2--ATP1A3--cen.

Gross (2021) mapped the ATP1A3 gene to chromosome 19q13.2 based on an alignment of the ATP1A3 sequence (GenBank BC009282) with the genomic sequence (GRCh38).


Gene Function

Agrin (AGRN; 103320) mediates accumulation of acetylcholine receptors at the developing neuromuscular junction through its interaction with MUSK (601296), and it has also been implicated in brain development. Through biochemical studies, Hilgenberg et al. (2006) found that agrin bound Atp1a3 in mouse cortical neurons. Immunohistochemical analysis showed that Atp1a3 colocalized with agrin-binding sites at synapses. Agrin inhibited Atp1a3 activity, resulting in membrane depolarization and increased action potential frequency in mouse cortical neurons in culture and acute slice. An agrin fragment that acted as a competitive antagonist depressed action potential frequency, indicating that endogenous agrin regulates native Atp1a3 function. Hilgenberg et al. (2006) concluded that agrin regulates activity-dependent processes in neurons through its interaction with ATP1A3.


Molecular Genetics

Dystonia 12

In 7 unrelated families with rapid-onset dystonia parkinsonism, or dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified 6 different heterozygous mutations in the ATP1A3 gene (182350.0001-182350.0006). Functional expression studies and structural analysis suggested that the mutations impaired enzyme activity or stability.

Anselm et al. (2009) and Blanco-Arias et al. (2009) reported de novo heterozygous ATP1A3 mutations (182350.0007 and 182350.0008, respectively) in patients with DYT12.

In Drosophila, Kaneko et al. (2014) identified a dominant missense mutation (A617T) in the calcium ATPase Serca gene (see SERCA2 (ATP2A2); 108740) that conferred temperature-sensitive motor uncoordination in a gain-of-function manner. The homologous residue is conserved by different type II P-type ATPases, including ATP1A2 (182340). Introduction of an R751Q mutation in the Drosophila Serca gene also caused a temperature-sensitive uncoordination phenotype. The corresponding residue in human SERCA2, ATP1A2, and ATP1A3 is mutated in the human diseases Darier disease (124200), FHM2 (602481), and dystonia-12, respectively. Cellular expression of Drosophila A617T resulted in temperature-induced decreased levels of stored calcium compared to wildtype, whereas cellular expression of R751Q elicited depletion of stored calcium even without heating. These calcium changes were due to leakage through the mutant channel pores that overwhelmed the pumping capacity of the cell. Similar results occurred after transfection of these mutations, as well as other disease-causing mutations that affected different parts of the protein, into mouse cells. Kaneko et al. (2014) concluded that ionic leakage is a gain-of-function mechanism that underlies a variety of dominant type II P-type ATPase-related diseases.

Alternating Hemiplegia of Childhood 2

In 82 of 105 patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified 19 different heterozygous mutations in the ATP1A3 gene (see, e.g., 182350.0009-182350.0012). The first mutations were identified through exome sequencing of affected individuals. Thirteen of the 18 mutations observed in sporadic cases were confirmed to occur de novo. Since it was possible that some variants represented polymorphisms, Heinzen et al. (2012) estimated that mutations in the ATP1A3 gene may be responsible for up to 74% of patients with sporadic, typical AHC. Several mutations were recurrent, and some occurred within hypermutable sequences. All patients had infantile onset of hemiplegia attacks, usually associated with episodes of quadriparesis, abnormal eye movements, autonomic signs, seizures, dystonia, ataxia, chorea, and developmental delay. Transfection of several of the mutations in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased. In contrast, transfection of DYT12-associated mutations resulted in decreased protein levels as well as decreased activity. The report expanded the spectrum of phenotypes associated with mutations in the ATP1A3 gene.

Simultaneously and independent to the report of Heinzen et al. (2012), Rosewich et al. (2012) identified de novo heterozygous mutations in the ATP1A3 gene (see, e.g., 182350.0009; 182350.0010; 182350.0015-182350.0017) in 24 unrelated patients with AHC2. Mutations in the first 3 patients were found by whole-exome sequencing of 3 affected child-parent trios, and subsequent mutations were found by direct Sanger sequencing of the ATP1A3 gene in additional patients. There were 2 main recurrent mutations: D801N (182350.0009) and E815K (182350.0010), found in 9 (38%) and 7 (29%) patients, respectively, suggesting mutational hotspots. None of the mutations resulted in a truncated protein, although there was 1 splice site mutation (182350.0017). Functional studies of the variants and studies of patients cells were not performed. Rosewich et al. (2012) noted the phenotypic overlap between AHC2 and DYT12.

In 45 (95.7%) of 47 Chinese children with typical AHC2, Yang et al. (2014) identified 19 different heterozygous missense mutations in the ATP1A3 gene. Three mutation hotspots, D801N (182350.0009), E815K (182350.0010), and G947R (182350.0012 and 182350.0013), were detected in 14 (31.1%), 9 (20.0%), and 7 (15.6%) ATP1A3-positive patients, respectively. Except for 1 patient who had inherited a mutation from her affected mother, all patients for whom parental DNA was available were found to have de novo mutations. Heterozygous ATP1A3 mutations were also found in 4 additional Chinese patients with atypical AHC2 who had onset of the disorder after 18 months of age. The initial mutations were found by whole-exome sequencing of several patients, and the subsequent mutations were found by direct sequencing of the ATP1A3 gene in a larger cohort. Presence of the E815K mutation was associated with epilepsy. A review of published disease-associated ATP1A3 mutations suggested that mutations associated with AHC2 were predominantly located in the transmembrane domain, whereas those associated with DYT12 had no location bias. Molecular modeling of the variants identified 2 statistically significant molecular features, solvent accessibility and distance to metal ion, that distinguished disease-associated mutations from neutral variants. In vitro functional studies were not performed on any of the variants.

Cerebellar Ataxia, Areflexia, Pes Cavus, Optic Atrophy, And Sensorineural Hearing Loss

In 10 patients from 3 unrelated families with cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss (CAPOS; 601338), Demos et al. (2014) identified the same heterozygous missense mutation in the ATP1A3 gene (E818K; 182350.0014).

Developmental And Epileptic Encephalopathy 99

In 16 patients from 15 families with developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified 14 heterozygous mutations in the ATP1A2 gene (see, e.g., 182350.0019-182350.0021). The mutations occurred de novo in all except for a mother and son pair (patients 14 and 15). Mutations were mostly missense, with a few small in-frame deletions or insertions. All occurred at conserved residues, and none were present in the gnomAD database. In vitro functional expression studies showed that all of the mutations caused variable functional defects in the Na+/(K+)ATPase. Variants with more severe functional deficits were associated with a more severe phenotype. The findings were consistent with a loss-of-function effect. Vetro et al. (2021) estimated that about 12% of ATP1A3 mutations may be associated with DEE. Polymicrogyria was estimated to occur in about 5.5% of patients with ATP1A3 mutations.

Associations Pending Confirmation

For discussion of a possible association between autosomal recessive congenital hydrocephalus due to aqueductal stenosis (see, e.g., 236635) and variation in the ATP1A3 gene, see 182350.0022 and 182350.0023.


Genotype/Phenotype Correlations

Rosewich et al. (2014) identified 16 patients with AHC2 and 3 with DYT12 confirmed by genetic analysis. A review of the clinical and molecular findings of these patients and of 164 previously reported patients with ATP1A3 mutations indicated that although mutations were distributed over almost all protein domains, those affecting transmembrane and functional domains tended to be associated with AHC2 as the more severe phenotype. The majority of mutations associated with AHC2 were located in exons 17 and 18, whereas those associated with DYT12 were located in exons 8 and 14; however, there was overlap, particularly in exon 17. Clinical analysis suggested that the 2 disorders represent a continuous phenotypic spectrum, with intermediate phenotypes combining criteria of both conditions. Shared clinical characteristics of both disorders include asymmetric movement disorder, rostrocaudal gradient of involvement with prominent bulbar symptoms, and triggering of symptoms by different stressors.

Using a formulated questionnaire, Panagiotakaki et al. (2015) assessed clinical data from 155 patients with AHC, including 132 confirmed to have ATP1A3 mutations by genetic analysis. Among those with AHC2, the most frequent mutations were D801N (in 43%), E815K (in 16%), and G947R (182350.0012 and 182350.0013, which were considered together) (in 11%). E815K was associated with a severe phenotype, with greater intellectual and motor disability; D801N appeared to confer a milder phenotype; and G947R correlated with the most favorable prognosis. For those with epilepsy, the age at seizure onset was earlier for patients with the E815K or G947R mutations than for those with the D801N mutation. Several mutational clusters within the gene were identified.


Animal Model

Ashmore et al. (2009) identified 6 different EMS-induced missense mutations in the Atp1a2 and Atp1a3 genes in Drosophila. All mutations resulted in reduced respiration activity consistent with a loss of ATPase function and a hypomorphic effect. Different mutant strains exhibited some abnormalities, including progressive temperature-dependent paralysis, progressive stress-sensitive paralysis, and decreased locomotor activity in response to startle, suggesting a decrease in maximal locomotion capacity. Neuromuscular studies showed allele-specific pathology, including brain vacuoles and myopathology, and biochemical studies showed decreased metabolic rates. An unexpected finding was the some mutant strains had increased longevity, which was not related to caloric restriction. Low doses of ouabain showed a similar effect on longevity in control groups. Ashmore et al. (2009) suggested that these findings may be relevant for studying the pathogenesis of FHM2 and DYT12 (128235).

In a mouse mutagenesis screen, Clapcote et al. (2009) identified a mutant mouse strain, Myshkin (Myk), that showed autosomal dominant complex and partial and secondarily generalized seizures, a reduced threshold for seizures in hippocampal slices, and neuronal degeneration in the hippocampus. Heterozygous mice were also smaller than wildtype, and homozygosity for the mutation resulted in perinatal death. Positional cloning and functional analysis identified a heterozygous ile810-to-asn (I810N) substitution in the Atp1a3 gene as responsible for the phenotype. In vitro cellular functional expression studies showed that the I810N substitution disrupted enzymatic function. The mutant protein had 42% reduced activity in mouse brain, and the phenotype could be rescued by transfection of the wildtype gene, consistent with loss of function as a pathogenic mechanism. The findings indicated the importance of ion homeostasis in maintaining normal neuronal excitability.

Doganli et al. (2013) found that the 2 ATP1A3 orthologs in zebrafish, Atp1a3a and Atp1a3b, were expressed in distinct but overlapping sets of brain structures. Morpholino-mediated knockdown (KD) of either Atp1a3a or Atp1a3b caused dilation of brain ventricles. Dilation in Atp1a3a KD embryos was not rescued by coinjection of Atp1a3b, and Atp1a3b KD embryos were not rescued by coinjection of Atp1a3a. Atp1a3a KD also caused depolarization of the resting membrane potential of Rohon-Beard neurons, which are mechanosensory neurons localized in the dorsal spinal cord. Atp1a3a KD and Atp1a3b KD embryos showed abnormal but distinct spontaneous motility and responses to tactile stimuli. Proteomic analysis revealed that Atp1a3a KD and Atp1a3b KD altered expression of overlapping sets of genes.

Sugimoto et al. (2014) found that Atp1a3 +/- mice showed shortened stride length and decreased motor strength in hanging box test following restraint stress compared with wildtype mice. Male and female Atp1a3 +/- mice showed some differences in the effects of restraint stress.

Balestrini et al. (2020) studied electrocardiogram abnormalities in a mouse model with heterozygosity for an Atp1a3 D801N knock-in mutation (Mashl +/-). Compared to 15 wildtype mice, 3 Mashl +/- mice had an increased heart rate, prolonged QRS and PR interval, and a longer QTc interval. After intraamygdala injection of kainic acid to induce seizures, all mice had elevation of JT intervals. One of the Mashl +/- mice had a period with JT-segment depression, and 2 Mashl +/- mice died from atrioventricular block.


ALLELIC VARIANTS 23 Selected Examples):

.0001   DYSTONIA 12

ATP1A3, THR613MET
SNP: rs80356534, ClinVar: RCV000013772, RCV000726724, RCV001004717

In a sporadic patient (Linazasoro et al., 2002) and affected members of a second family (Zaremba et al., 2004) with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 1838C-T transition in the ATP1A3 gene, resulting in a thr613-to-met (T613M) substitution in a highly conserved residue near the phosphorylation domain on the cytoplasmic face of the protein. The mutation was not identified in 500 northern European control chromosomes.

Brashear et al. (2007) identified the T613M mutation in a family with DYT12 reported by Pittock et al. (2000).

Rodacker et al. (2006) noted that T613 is universally conserved among Na+/K(+)-ATPases, H+/K(+)-ATPases, and Ca(2+)-ATPases. Using rat Atp1a1 (182310) for technical reasons, they presented evidence that the T613M substitution in ATP1A3 alters the affinity of ATP1A3 for Na+ and ATP and alters the conformation equilibrium in favor of the potassium-bound form.


.0002   DYSTONIA 12

ATP1A3, ILE274THR
SNP: rs80356532, ClinVar: RCV000013773

In a patient with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 821T-C transition in the ATP1A3 gene, resulting in an ile274-to-thr (I274T) substitution in a highly conserved residue in the transmembrane domain of the protein. The mutation was not identified in 500 northern European control chromosomes. The patient had disease onset at age 37 years.


.0003   DYSTONIA 12

ATP1A3, GLU277LYS
SNP: rs80356533, ClinVar: RCV000013774

In a patient with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 829G-A transition in exon 8 of the ATP1A3 gene, resulting in a glu277-to-lys (E277K) substitution in a highly conserved residue in the transmembrane domain of the protein. The mutation was not identified in 500 northern European control chromosomes. The patient had disease onset at age 20 years.

Tarsy et al. (2010) identified the E277K mutation in a 29-year-old woman of African Caribbean descent with DYT12. She had onset at age 26 years of weakness and flexion of the left hand and ankle, which progressed rapidly over the next few years to become frank dystonia of the left arm and bulbar symptoms, including dysphagia, laryngeal dysfunction with task-specific dysphonia, and oropharyngeal dysmotility. She also had mild parkinsonism, with hypomimia and wide-based gait. Treatment with oral trihexyphenidyl and botulinum injection into selected laryngeal muscles resulted in clinical improvement.


.0004   DYSTONIA 12

ATP1A3, ILE758SER
SNP: rs80356535, ClinVar: RCV000013775, RCV001781260

In 12 affected members of a family with dystonia-12 (DYT12; 128235) reported by Dobyns et al. (1993), de Carvalho Aguiar et al. (2004) identified a heterozygous 2273T-G transversion in the ATP1A3 gene, resulting in an ile758-to-ser (I758S) substitution in a highly conserved residue in the transmembrane domain of the protein. The mutation was not identified in 500 northern European control chromosomes.


.0005   DYSTONIA 12

ATP1A3, PHE780LEU
SNP: rs80356536, ClinVar: RCV000013776

In 2 affected members of a family with dystonia-12 (DYT12; 128235), de Carvalho Aguiar et al. (2004) identified a heterozygous 2338T-C transition in the ATP1A3 gene, resulting in a phe780-to-leu (F780L) substitution in a highly conserved residue in the transmembrane region of the protein close to the extracellular surface. The mutation was not identified in 500 northern European control chromosomes.

Rodacker et al. (2006) noted that F780 is fully conserved in all known Na+/K(+)-ATPases from different species. Using rat Atp1a1 for technical reasons, they presented evidence that the F780L substitution reduced the affinity of ATP1A3 for Na+ and reduced the V(max) for ATP-dependent phosphorylation. The mutation was not expected to affect either the affinity of ATP1A3 for K+ nor the K(+)-induced dephosphorylation event.


.0006   DYSTONIA 12

ATP1A3, ASP801TYR
SNP: rs80356537, ClinVar: RCV000013777

In 4 affected members of a family with dystonia-12 (DYT12; 128235) reported by Brashear et al. (1997), de Carvalho Aguiar et al. (2004) identified a heterozygous 2401G-T transversion in the ATP1A3 gene, resulting in an asp801-to-tyr (D801Y) substitution in a highly conserved residue in the transmembrane region of the protein. The mutation was not identified in 500 northern European control chromosomes.


.0007   DYSTONIA 12

ALTERNATING HEMIPLEGIA OF CHILDHOOD 2, INCLUDED
ATP1A3, ASP923ASN
SNP: rs267606670, ClinVar: RCV000013778, RCV000128465, RCV000763432, RCV003233069, RCV003389231

In a boy with early-onset dystonia-12 (DYT12; 128235) at age 4 years, Anselm et al. (2009) identified a heterozygous de novo 2767G-A transition in exon 20 of the ATP1A3 gene, resulting in an asp923-to-asn (D923N) substitution. The mutation was not found in 338 Caucasian control chromosomes. The substitution was predicted to occur in a residue buried in the membrane close to the ion-binding residue gln920, suggesting that it may affect enzyme activity. He was born of an unaffected Caucasian father and Chinese mother. The onset of dystonia was abrupt, occurring after mild trauma to the forehead. He developed mutism, eye convergence, and inability to walk, which later evolved into severe dystonia, severe dysarthria, and drooling. The condition stabilized over several months, and he showed mild improvement over the next 8 years. About a year after onset, he developed unusual episodes of flaccidity lasting for hours, later replaced by shorter episodes of stiffness. Treatment with L-DOPA was not effective. At the time of the report, he had bulbar symptoms, striking oromotor dystonia with inability to speak or swallow well, and apraxia.

Yang et al. (2014) identified a de novo heterozygous D923N mutation in a Chinese boy with atypical alternating hemiplegia of childhood-2 (AHC2; 614820). The phenotype was considered atypical due to relatively late onset of symptoms at age 30 months. Otherwise, the boy had typical features of quadriplegia as well as abnormal eye movements, dystonia, and developmental delay.


.0008   DYSTONIA 12

ATP1A3, 3-BP DUP, 3191TAC
SNP: rs397515382, ClinVar: RCV000148335

In a 16-year-old female with dystonia-12 (DYT12; 128235), Blanco-Arias et al. (2009) reported a de novo heterozygous 3-bp duplication (3191dupTAC) in exon 23 of the ATP1A3 gene, resulting in duplication of tyr1013, the C-terminal amino acid of the protein before the stop codon. The mutation was not found in either parent, her brother, or in 218 control individuals. HeLa cells expressing the mutant protein showed decreased survival in response to ouabain challenge, but no defect was detected in protein expression or plasma membrane targeting. Functional analysis demonstrated a drastic 40- to 50-fold reduction in Na+ affinity in the mutant. Blanco-Arias et al. (2009) suggested a crucial role for the C terminus of the alpha-subunit in the function of the Na+/K(+)-ATPase and emphasized a key impact of Na+ affinity in the pathophysiology of DYT12.


.0009   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, ASP801ASN
ClinVar: RCV000030749, RCV000413511, RCV000515424, RCV000624579, RCV000644928, RCV001004008, RCV001265551, RCV002281545, RCV003934862

In 36 of 95 unrelated patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a heterozygous 2401G-A transition in the ATP1A3 gene, resulting in an asp801-to-asn (D801N) substitution in the sixth transmembrane domain. The mutation was demonstrated to occur de novo in cases where parental material was available. All patients had infantile onset of hemiplegia attacks, usually associated with episodes of quadriparesis, abnormal eye movements, autonomic signs, seizures, dystonia, ataxia, chorea, and developmental delay. Transfection of the mutation in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased. Evaluation of the crystal structure of the protein predicted that the D801N substitution would prevent the binding of potassium ions to the pump.

Rosewich et al. (2012) identified a de novo heterozygous D801N mutation in 9 (38%) of 24 AHC2 patients. D801N occurs in the functionally conserved C-terminal cation-transporting ATPase domain and the P-type ATPase domain that is also a transmembrane domain. Functional studies of the variant and studies of patient cells were not performed.

Yang et al. (2014) identified a de novo heterozygous D801N mutation in 14 unrelated Chinese children with AHC2. All had typical features of the disorder, including abnormal eye movements and developmental delay, but none had seizures.


.0010   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, GLU815LYS
SNP: rs387907281, ClinVar: RCV000030750, RCV000432504, RCV000469482, RCV000626997, RCV000763433, RCV001192636, RCV001267254, RCV001807744, RCV002243675

In 19 patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a heterozygous 2443G-A transition in the ATP1A3 gene, resulting in a glu815-to-lys (E815K) substitution in the sixth transmembrane domain. The mutation was shown to occur de novo in all patients whose parents were available for study. Transfection of the mutation in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased.

Rosewich et al. (2012) identified a de novo heterozygous E815K mutation in 7 (29%) of 24 AHC2 patients. E815K occurs in the functionally conserved C-terminal cation-transporting ATPase domain and the P-type ATPase domain that is also a transmembrane domain. Functional studies of the variant and studies of patient cells were not performed.

Yang et al. (2014) identified a de novo heterozygous E815K mutation in 9 unrelated Chinese children with AHC2. Seven of the patients had epilepsy.


.0011   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, SER811PRO
SNP: rs387907282, ClinVar: RCV000030751, RCV000541711

In 4 unrelated patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a de novo heterozygous 2431T-C transition in the ATP1A3 gene, resulting in a ser811-to-pro (S811P) substitution in the sixth transmembrane domain. Transfection of the mutation in HeLa cells showed protein levels similar to wildtype, but ATP1A3 activity was significantly decreased.


.0012   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, NT2839G-A, GLY947ARG
SNP: rs398122887, ClinVar: RCV000030752, RCV000415180, RCV000418823, RCV000476589, RCV000763431, RCV003982851

In 5 patients with alternating hemiplegia of childhood-2 (AHC2; 614820), Heinzen et al. (2012) identified a heterozygous 2839G-A transition in the ATP1A3 gene, resulting in a gly947-to-arg (G947R) substitution in the ninth transmembrane domain. The mutation was shown to occur de novo in all patients whose parents were available for study.

Yang et al. (2014) identified a de novo heterozygous G947R mutation in 5 unrelated Chinese children with AHC2. This same amino acid substitution can also result from a c.2839G-C transversion (182350.0013).


.0013   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, NT2839G-C, GLY947ARG
SNP: rs398122887, ClinVar: RCV000128466, RCV001849915

In 3 unrelated Chinese children with alternating hemiplegia of childhood-2 (AHC2; 614820), Yang et al. (2014) identified a heterozygous c.2839G-C transversion in exon 21 of the ATP1A3 gene, resulting in a gly947-to-arg (G947R) substitution at a highly conserved residue. The mutation occurred de novo in 2 of the patients and was inherited from an affected mother in the third patient. This same amino acid substitution can also result from a c.2839G-A transition (182350.0012). One of the patients had so-called atypical AHC2, with onset at 30 months of age. The mutation was not found in the 1000 Genomes Project or Exome Sequencing Project databases, or in 100 normal controls.


.0014   CEREBELLAR ATAXIA, AREFLEXIA, PES CAVUS, OPTIC ATROPHY, AND SENSORINEURAL HEARING LOSS

ATP1A3, GLU818LYS
SNP: rs587777771, ClinVar: RCV000144250, RCV000190725, RCV000195001, RCV000234480, RCV000314245

In 10 patients from 3 unrelated families with cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss (CAPOS; 601338), including the original family reported by Nicolaides et al. (1996), Demos et al. (2014) identified a heterozygous c.2452G-A transition in the ATP1A3 gene, resulting in a glu818-to-lys (E818K) substitution at a highly conserved residue. The mutation, which was found by whole-exome sequencing of 2 of the families and confirmed by Sanger sequencing, was filtered against the dbSNP (builds 129 and 130) and 1000 Genomes Project databases and was not found in 1,834 controls. The mutation occurred de novo in the oldest affected generation of 1 family, but haplotype analysis could not rule out the possibility of a remote relationship between the other 2 families. All families were of Caucasian European descent. Functional studies of the E818K variant were not performed, but Demos et al. (2014) postulated a gain-of-function effect.

In a German boy with CAPOS, Rosewich et al. (2014) identified a de novo heterozygous E818K mutation in the ATP1A3 gene. Functional studies were not performed.

Tranebjaerg et al. (2018) reported that residue 818 of ATP1A3 is located at the cytoplasmic side of transmembrane helix-6, where it forms a salt bridge with the backbone carbonyl of arg930, a residue that stabilizes the C terminus. Tranebjaerg et al. (2018) expressed ATP1A3 with the E818K mutation in Xenopus laevis oocytes. Electrophysiologic analysis showed that the mutation disrupted the C terminus, caused opening of the C-terminal structure of ATP1A3, and affected sodium binding to and release from the binding site in the molecule. Molecular dynamic simulations confirmed that E818K opened the C-terminal pathway, allowing rapid entry of water molecules toward the ion-binding site.


.0015   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, ASP923TYR
SNP: rs267606670, ClinVar: RCV000148329, RCV000489720, RCV000689821

In a 19-year-old man with alternating hemiplegia of childhood-2 (AHC2; 614820), Rosewich et al. (2012) identified a de novo heterozygous c.2767G-T transversion (c.2767G-T, NM_152296.4) in exon 20 of the ATP1A3 gene, resulting in an asp923-to-tyr (D923Y) substitution at a highly conserved residue in the C-terminal cation ATPase domain. Functional studies of the mutation and studies of patient cells were not performed. The authors noted that a different substitution at this same residue (D923N; 182350.0007) has been identified in patients with dystonia-12 (DYT12; 128235).


.0016   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, ILE274ASN
SNP: rs80356532, ClinVar: RCV000148305

In a 17-year-old girl with alternating hemiplegia of childhood-2 (AHC2; 614820), Rosewich et al. (2012) identified a de novo heterozygous c.821T-A transversion (c.821T-A, NM_152296.4) in exon 8 of the ATP1A3 gene, resulting in an ile274-to-asn (I274N) substitution at a highly conserved residue in the E1-E2 ATPase domain. Functional studies of the mutation and studies of patient cells were not performed. The authors noted that a different substitution at this same residue (I274T; 182350.0002) has been identified in patients with dystonia-12 (DYT12; 128235).


.0017   ALTERNATING HEMIPLEGIA OF CHILDHOOD 2

ATP1A3, IVS18DS, G-A, +1
SNP: rs606231441, ClinVar: RCV000148326, RCV001850014

In a 24-year-old woman with alternating hemiplegia of childhood-2 (AHC2; 614820), Rosewich et al. (2012) identified a de novo heterozygous G-to-A transition in intron 18 of the ATP1A3 gene (c.2542+1G-A, NM_152296.4), predicted to result in exon skipping. Functional studies of the mutation and studies of patient cells were not performed.


.0018   DYSTONIA 12

ATP1A3, GLY316SER
SNP: rs869320661, ClinVar: RCV000210848, RCV003335231

In a 26-year-old man with dystonia-12 (DYT12; 128235), Sweadner et al. (2016) identified a de novo heterozygous c.946G-A transition (c.946G-A, NM_152296.3) in the ATP1A3 gene, resulting in a gly316-to-ser (G316S) substitution in the highly conserved fourth transmembrane domain and close to an ion binding pocket. The mutation was found by exome sequencing and confirmed by Sanger sequencing. In vitro functional studies showed that the mutation resulted in a growth defect when expressed in HEK293 cells, consistent with impaired Na/K-ATPase function. The patient had a somewhat unusual phenotype, presenting at age 19 with rapidly progressive ataxia and dysarthria and tremor, resulting in loss of independent ambulation, and minimal dystonia. Exome sequencing showed that the patient also carried a de novo heterozygous missense E482K variant in the UBQLN4 gene (605440), which may have played a role in the prominent cerebellar ataxia and cerebellar atrophy observed in this patient; functional studies of the UBQLN4 variant were not performed.


.0019   DEVELOPMENTAL AND EPILEPTIC ENCEPHALOPATHY 99

ATP1A3, LEU292ARG
SNP: rs2145977887, ClinVar: RCV001777181

In a 2-month-old infant (patient 7) with lethal developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified a de novo heterozygous c.875T-G transversion (c.875T-G, NM_152296.4) in the ATP1A3 gene, resulting in a leu292-to-arg (L292R) substitution at a conserved residue. The mutation was not present in the gnomAD database. In vitro studies showed that the mutation was unable to support growth and survival of COS1 cells in culture and interfered with Na+ and K+ affinity, resulting in nearly absent Na+/(K+)ATPase activity, consistent with a loss-of-function effect. The patient presented at birth with migrating focal seizures that evolved to almost continuous seizure activity with status epilepticus, resulting in death at 2 months of age.


.0020   DEVELOPMENTAL AND EPILEPTIC ENCEPHALOPATHY 99

ATP1A3, GLY316VAL
SNP: rs2145977758, ClinVar: RCV001777182

In a 6-year-old girl (patient 8) with developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified a de novo heterozygous c.947G-T transversion (c.947G-T, NM_152296.4) in the ATP1A3 gene, resulting in a gly316-to-val (G316V) substitution at a conserved residue. The mutation was not present in the gnomAD database. In vitro studies showed that the mutation was unable to support growth and survival of COS1 cells in culture and interfered with Na+ and K+ affinity, resulting in nearly absent Na+/(K+)ATPase activity, consistent with a loss-of-function effect. The patient had onset of migrating focal and severe generalized seizures at 4 years of age.


.0021   DEVELOPMENTAL AND EPILEPTIC ENCEPHALOPATHY 99

ATP1A3, SER361PRO
SNP: rs2145972497, ClinVar: RCV001777183

In a 7-year-old girl (patient 9) with developmental and epileptic encephalopathy-99 (DEE99; 619606), Vetro et al. (2021) identified a de novo heterozygous c.1081T-C transition (c.1081T-C, NM_152296.4) in the ATP1A3 gene, resulting in a ser361-to-pro (S361P) substitution at a conserved residue. The mutation was not present in the gnomAD database. In vitro studies showed that the mutation was unable to support growth and survival of COS1 cells in culture with decreased phosphorylation activity and nearly absent Na+/(K+)ATPase activity, consistent with a loss-of-function effect. The patient had onset of focal temporal seizures at 5 months of age.


.0022   VARIANT OF UNKNOWN SIGNIFICANCE

ATP1A3, ARG19CYS
SNP: rs782229302, gnomAD: rs782229302, ClinVar: RCV000803350, RCV001128681, RCV001777177

This variant is classified as a variant of unknown significance because its contribution to congenital hydrocephalus due to aqueductal stenosis (see, e.g., 236635) has not been confirmed.

In a 23-year-old Caucasian woman with congenital hydrocephalus due to aqueductal stenosis, Allocco et al. (2019) identified compound heterozygous missense variants in the ATP1A3 gene: a c.55G-A transition (c.55G-A, NM_152296) in exon 2, resulting in an arg19-to-cys (R19C) substitution inherited from the unaffected mother, and a c.1387G-A transition in exon 11, resulting in an arg463-to-cys (R463C; 182350.0023) substitution inherited from the unaffected father. The variants were identified by whole-exome sequencing and confirmed by Sanger sequencing. The R19C variant was present in the heterozygous state at a low frequency in gnomAD (6.4 x 10(-5)). Both variants occurred at conserved residues and were predicted to have disruptive effects on protein stability, although functional studies of the variants and studies of patient cells were not performed. The patient had multiple brain malformations, including open schizencephaly, type 1 Chiari malformation, and dysgenesis of the corpus callosum. Clinical details were limited, but she was noted to have learning disabilities. The authors postulated that dysregulation of neural development may be the pathogenesis of the disorder in this patient.


.0023   VARIANT OF UNKNOWN SIGNIFICANCE

ATP1A3, ARG463CYS
SNP: rs150785666, gnomAD: rs150785666, ClinVar: RCV000441666, RCV000547051, RCV001131091, RCV001777161, RCV002524753, RCV003970106

This variant is classified as a variant of unknown significance because its contribution to congenital hydrocephalus due to aqueductal stenosis (see, e.g., 236635) has not been confirmed.

For discussion of the c.1387G-A transition (c.1387G-A, NM_152296) in exon 11 of the ATP1A3 gene, resulting in an arg463-to-cys (R463C) substitution, that was found in compound heterozygous state in a patient with congenital hydrocephalus due to aqueductal stenosis, by Allocco et al. (2019) see 182350.0022.


REFERENCES

  1. Allocco, A. A., Jin, S. C., Duy, P. Q., Furey, C. G., Zeng, X., Dong, W., Nelson-Williams, C., Karimy, J. K., DeSpenza, T., Hao, L. T., Reeves, B., Haider, S., Gunel, M., Lifton, R. P., Kahle, K. T. Recessive inheritance of congenital hydrocephalus with other structural brain abnormalities caused by compound heterozygous mutations in ATP1A3. Front. Cellular Neurosci. 13: 425, 2019.

  2. Anselm, I. A., Sweadner, K. J., Gollamudi, S., Ozelius, L. J., Darras, B. T. Rapid-onset dystonia-parkinsonism in a child with a novel ATP1A3 gene mutation. Neurology 73: 400-401, 2009. [PubMed: 19652145] [Full Text: https://doi.org/10.1212/WNL.0b013e3181b04acd]

  3. Ashmore, L. J., Hrizo, S. L., Paul, S. M., Van Voorhies, W. A., Beitel, G. J., Palladino, M. J. Novel mutations affecting the Na, K ATPase alpha model complex neurological diseases and implicate the sodium pump in increased longevity. Hum. Genet. 126: 431-447, 2009. [PubMed: 19455355] [Full Text: https://doi.org/10.1007/s00439-009-0673-2]

  4. Balestrini, S., Mikati, M. A., Alvarez-Garcia-Roves, R., Carboni, M. Hunanyan, A. S., Kherallah, B., McLean, M., Prange, L., De Grandis, E., Gagliardi, A., Pisciotta, L., Stagnaro, M., and 42 others. Cardiac phenotype in ATP1A3-related syndromes: a multicenter study. Neurology 95: e2866-e2879, 2020. [PubMed: 32913013] [Full Text: https://doi.org/10.1212/WNL.0000000000010794]

  5. Blanco-Arias, P., Einholm, A. P., Mamsa, H., Concheiro, C., Gutierrez-de-Teran, H., Romero, J., Toustrup-Jensen, M. S., Carracedo, A., Jen, J. C., Vilsen, B., Sobrido, M.-J. A C-terminal mutation of ATP1A3 underscores the crucial role of sodium affinity in the pathophysiology of rapid-onset dystonia-parkinsonism. Hum. Molec. Genet. 18: 2370-2377, 2009. [PubMed: 19351654] [Full Text: https://doi.org/10.1093/hmg/ddp170]

  6. Brashear, A., DeLeon, D., Bressman, S. B., Thyagarajan, D., Farlow, M. R., Dobyns, W. B. Rapid-onset dystonia-parkinsonism in a second family. Neurology 48: 1066-1069, 1997. [PubMed: 9109901] [Full Text: https://doi.org/10.1212/wnl.48.4.1066]

  7. Brashear, A., Dobyns, W. B., de Carvalho Aguiar, P., Borg, M., Frijns, C. J. M., Gollamudi, S., Green, A., Guimaraes, J., Haake, B. C., Klein, C., Linazasoro, G., Munchau, A., Raymond, D., Riley, D., Saunders-Pullman, R., Tijssen, M. A. J., Webb, D., Zaremba, J., Bressman, S. B., Ozelius, L. J. The phenotypic spectrum of rapid-onset dystonia-parkinsonism (RDP) and mutations in the ATP1A3 gene. Brain 130: 828-835, 2007. [PubMed: 17282997] [Full Text: https://doi.org/10.1093/brain/awl340]

  8. 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., Minassian, B. A., Ackerley, C. A., Sled, J. G., Cortez, M. A., Henderson, J. T., Vilsen, B., Roder, J. C. Mutation I810N in the alpha-3 isoform of Na+,K(+)-ATPase causes impairments in the sodium pump and hyperexcitability in the CNS. Proc. Nat. Acad. Sci. 106: 14085-14090, 2009. [PubMed: 19666602] [Full Text: https://doi.org/10.1073/pnas.0904817106]

  9. de Carvalho Aguiar, P., Sweadner, K. J., Penniston, J. T., Zaremba, J., Liu, L., Caton, M., Linazasoro, G., Borg, M., Tijssen, M. A. J., Bressman, S. B., Dobyns, W. B., Brashear, A., Ozelius, L. J. Mutations in the Na+/K(+)-ATPase alpha-3 gene ATP1A3 are associated with rapid-onset dystonia parkinsonism. Neuron 43: 169-175, 2004. [PubMed: 15260953] [Full Text: https://doi.org/10.1016/j.neuron.2004.06.028]

  10. Demos, M. K., van Karnebeek, C. D. M., Ross, C. J. D., Adam, S., Shen, Y., Zhan, S. H., Shyr, C., Horvath, G., Suri, M., Fryer, A., Jones, S. J. M., Friedman, J. M., FORGE Canada Consortium. A novel recurrent mutation in ATP1A3 causes CAPOS syndrome. Orphanet J. Rare Dis. 9: 15, 2014. Note: Electronic Article. [PubMed: 24468074] [Full Text: https://doi.org/10.1186/1750-1172-9-15]

  11. Dobyns, W. B., Ozelius, L. J., Kramer, P. L., Brashear, A., Farlow, M. R., Perry, T. R., Walsh, L. E., Kasarskis, E. J., Butler, I. J., Breakefield, X. O. Rapid-onset dystonia-parkinsonism. Neurology 43: 2596-2602, 1993. [PubMed: 8255463] [Full Text: https://doi.org/10.1212/wnl.43.12.2596]

  12. Doganli, C., Beck, H. C., Ribera, A. B., Oxvig, C., Lykke-Hartmann, K. Alpha-3-Na+/K(+)-ATPase deficiency causes brain ventricle dilation and abrupt embryonic motility in zebrafish. J. Biol. Chem. 288: 8862-8874, 2013. [PubMed: 23400780] [Full Text: https://doi.org/10.1074/jbc.M112.421529]

  13. Gross, M. B. Personal Communication. Baltimore, Md. 2/1/2021.

  14. Harley, H. G., Brook, J. D., Jackson, C. L., Glaser, T., Walsh, K. V., Sarfarazi, M., Kent, R., Lager, M., Koch, M., Harper, P. S., Levenson, R., Housman, D. E., Shaw, D. J. Localization of a human Na+,K(+)-ATPase alpha subunit gene to chromosome 19q12-q13.2 and linkage to the myotonic dystrophy locus. Genomics 3: 380-384, 1988. [PubMed: 2907504] [Full Text: https://doi.org/10.1016/0888-7543(88)90131-0]

  15. 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., Panagiotakaki, E, European Alternating Hemiplegia of Childhood (AHC) Genetics Consortium, and 33 others. De novo mutations in ATP1A3 cause alternating hemiplegia of childhood. Nature Genet. 44: 1030-1034, 2012. [PubMed: 22842232] [Full Text: https://doi.org/10.1038/ng.2358]

  16. Hilgenberg, L. G. W., Su, H., Gu, H., O'Dowd, D. K., Smith, M. A. Alpha-3-Na(+)/K(+)-ATPase is a neuronal receptor for agrin. Cell 125: 359-369, 2006. [PubMed: 16630822] [Full Text: https://doi.org/10.1016/j.cell.2006.01.052]

  17. Kaneko, M., Desai, B. S., Cook, B. Ionic leakage underlies a gain-of-function effect of dominant disease mutations affecting diverse P-type ATPases. Nature Genet. 46: 144-151, 2014. [PubMed: 24336169] [Full Text: https://doi.org/10.1038/ng.2850]

  18. Linazasoro, G., Indakoetxea, B., Ruiz, J., Van Blercom, N., Lasa, A. Possible sporadic rapid-onset dystonia-parkinsonism. Mov. Disord. 17: 608-609, 2002. [PubMed: 12112218] [Full Text: https://doi.org/10.1002/mds.10103]

  19. Nicolaides, P., Appleton, R. E., Fryer, A. Cerebellar ataxia, areflexia, pes cavus, optic atrophy, and sensorineural hearing loss (CAPOS): a new syndrome. J. Med. Genet. 33: 419-421, 1996. [PubMed: 8733056] [Full Text: https://doi.org/10.1136/jmg.33.5.419]

  20. Ovchinnikov, Y. A., Monastyrskaya, G. S., Broude, N. E., Ushkaryov, Y. A., Melkov, A. M., Smirnov, Y. V., Malyshev, I. V., Allikmets, R. L., Kostina, M. B., Dulubova, I. E., Kiyatkin, N. I., Grishin, A. V., Modyanov, N. N., Sverdlov, E. D. Family of human Na+,K(+)-ATPase genes: structure of the gene for the catalytic subunit (alpha-III-form) and its relationship with structural features of the protein. FEBS Lett. 233: 87-94, 1988. [PubMed: 2838329] [Full Text: https://doi.org/10.1016/0014-5793(88)81361-9]

  21. Panagiotakaki, E., De Grandis, E., Stagnaro, M., Heinzen, E. L., Fons, C., Sisodiya, S., de Vries, B., Goubau, C., Weckhuysen, S., Kemlink, D., Scheffer, I., Lesca, G., and 24 others. Clinical profile of patients with ATP1A3 mutations in alternating hemiplegia of childhood--a study of 155 patients. Orphanet J. Rare Dis. 10: 123, 2015. Note: Electronic Article. [PubMed: 26410222] [Full Text: https://doi.org/10.1186/s13023-015-0335-5]

  22. Pittock, S. J., Joyce, C., O'Keane, V., Hugle, B., Hardiman, O., Brett, F., Green, A. J., Barton, D. E., King, M. D., Webb, D. W. Rapid-onset dystonia-parkinsonism: a clinical and genetic analysis of a new kindred. Neurology 55: 991-995, 2000. [PubMed: 11061257] [Full Text: https://doi.org/10.1212/wnl.55.7.991]

  23. Rodacker, V., Toustrup-Jensen, M., Vilsen, B. Mutations Phe785Leu and Thr618Met in Na+,K(+)-ATPase, associated with familial rapid-onset dystonia parkinsonism, interfere with Na+ interaction by distinct mechanisms. J. Biol. Chem. 281: 18539-18548, 2006. [PubMed: 16632466] [Full Text: https://doi.org/10.1074/jbc.M601780200]

  24. Rosewich, H., Ohlenbusch, A., Huppke, P., Schlotawa, L., Baethmann, M., Carrilho, I., Fiori, S., Lourenco, C. M., Sawyer, S., Steinfeld, R., Gartner, J., Brockmann, K. The expanding clinical and genetic spectrum of ATP1A3-related disorders. Neurology 82: 945-955, 2014. [PubMed: 24523486] [Full Text: https://doi.org/10.1212/WNL.0000000000000212]

  25. Rosewich, H., Thiele, H., Ohlenbusch, A., Maschke, U., Altmuller, J., Frommolt, P., Zirn, B., Ebinger, F., Siemes, H., Nurnberg, P., Brockmann, K., Gartner, J. Heterozygous de-novo mutations in ATP1A3 in patients with alternating hemiplegia of childhood: a whole-exome sequencing gene-identification study. Lancet Neurol. 11: 764-773, 2012. [PubMed: 22850527] [Full Text: https://doi.org/10.1016/S1474-4422(12)70182-5]

  26. Rosewich, H., Weise, D., Ohlenbusch, A., Gartner, J., Brockmann, K. Phenotypic overlap of alternating hemiplegia of childhood and CAPOS syndrome. Neurology 83: 861-863, 2014. [PubMed: 25056583] [Full Text: https://doi.org/10.1212/WNL.0000000000000735]

  27. Sugimoto, H., Ikeda, K., Kawakami, K. Heterozygous mice deficient in Atp1a3 exhibit motor deficits by chronic restraint stress. Behav. Brain Res. 272: 100-110, 2014. [PubMed: 24983657] [Full Text: https://doi.org/10.1016/j.bbr.2014.06.048]

  28. Sweadner, K. J., Toro, C., Whitlow, C. T., Snively, B. M., Cook, J. F., Ozelius, L. J., Markello, T. C., Brashear, A. ATP1A3 mutation in adult rapid-onset ataxia. PLoS One 11: e0151429, 2016. Note: Electronic Article. [PubMed: 26990090] [Full Text: https://doi.org/10.1371/journal.pone.0151429]

  29. Tarsy, D., Sweadner, K. J., Song, P. C. Case 17-2010: a 29-year-old woman with flexion of the left hand and foot and difficulty speaking. New Eng. J. Med. 362: 2213-2219, 2010. [PubMed: 20558373] [Full Text: https://doi.org/10.1056/NEJMcpc1002112]

  30. Tranebjaerg, L., Strenzke, N., Lindholm, S., Rendtorff, N . D., Poulsen, H., Khandelia, H., Kopec, W., Lyngbye, T. J. B., Hamel, C., Delettre, C., Bocquet, B., Bille, M., and 19 others. The CAPOS mutation in ATP1A3 alters Na/K-ATPase function and results in auditory neuropathy which has implications for management. Hum. Genet. 137: 111-127, 2018. Note: Erratum Hum. Genet. 137: 279-280, 2018. [PubMed: 29305691] [Full Text: https://doi.org/10.1007/s00439-017-1862-z]

  31. Vetro, A., Nielsen, H. N., Holm, R., Hevner, R. F., Parrini, E., Powis, Z., Moller, R. S., Bellan, C., Simonati, A., Lesca, G., Helbig, K. L., Palmer, E. E., and 18 others. ATP1A2- and ATP1A3-associated early profound epileptic encephalopathy and polymicrogyria. Brain 144: 1435-1450, 2021. [PubMed: 33880529] [Full Text: https://doi.org/10.1093/brain/awab052]

  32. Wetzel, R. K., Arystarkhova, E., Sweadner, K. J. Cellular and subcellular specification of Na,K-ATPase alpha and beta isoforms in the postnatal development of mouse retina. J. Neurosci. 19: 9878-9889, 1999. [PubMed: 10559397] [Full Text: https://doi.org/10.1523/JNEUROSCI.19-22-09878.1999]

  33. Yang, X., Gao, H., Zhang, J., Xu, X., Liu, X., Wu, X., Wei, L., Zhang, Y. ATP1A3 mutations and genotype-phenotype correlation of alternating hemiplegia of childhood in Chinese patients. PLoS One 9: e97274, 2014. Note: Electronic Article. [PubMed: 24842602] [Full Text: https://doi.org/10.1371/journal.pone.0097274]

  34. Yang-Feng, T. L., Schneider, J. W., Lindgren, V., Shull, M. M., Benz, E. J., Jr., Lingrel, J. B., Francke, U. Chromosomal localization of human Na+,K(+)-ATPase alpha- and beta-subunit genes. Genomics 2: 128-138, 1988. [PubMed: 2842249] [Full Text: https://doi.org/10.1016/0888-7543(88)90094-8]

  35. Zaremba, J., Mierzewska, H., Lysiak, Z., Kramer, P., Ozelius, L. J., Brashear, A. Rapid-onset dystonia-parkinsonism: a fourth family consistent with linkage to chromosome 19q13. Mov. Disord. 19: 1506-1510, 2004. [PubMed: 15390049] [Full Text: https://doi.org/10.1002/mds.20258]


Contributors:
Cassandra L. Kniffin - updated : 11/04/2021
Matthew B. Gross - updated : 02/02/2021
Hilary J. Vernon - updated : 02/01/2021
Bao Lige - updated : 01/09/2019
Cassandra L. Kniffin - updated : 4/13/2016
Patricia A. Hartz - updated : 4/12/2016
Cassandra L. Kniffin - updated : 2/25/2015
Cassandra L. Kniffin - updated : 9/29/2014
Patricia A. Hartz - updated : 9/22/2014
Cassandra L. Kniffin - updated : 6/30/2014
Cassandra L. Kniffin - updated : 3/4/2014
Cassandra L. Kniffin - updated : 9/13/2012
Cassandra L. Kniffin - updated : 6/10/2010
Cassandra L. Kniffin - updated : 5/24/2010
George E. Tiller - updated : 3/30/2010
Matthew B. Gross - updated : 3/8/2010
Cassandra L. Kniffin - updated : 12/17/2009
Cassandra L. Kniffin - updated : 3/10/2005

Creation Date:
Victor A. McKusick : 12/1/1987

Edit History:
alopez : 11/10/2021
ckniffin : 11/04/2021
mgross : 02/02/2021
carol : 02/01/2021
alopez : 04/09/2019
mgross : 01/09/2019
carol : 04/07/2017
carol : 04/14/2016
carol : 4/14/2016
ckniffin : 4/13/2016
mgross : 4/12/2016
mgross : 4/12/2016
carol : 4/7/2016
carol : 2/26/2015
mcolton : 2/25/2015
ckniffin : 2/25/2015
alopez : 9/29/2014
ckniffin : 9/29/2014
mgross : 9/26/2014
mcolton : 9/22/2014
alopez : 7/3/2014
mcolton : 7/1/2014
ckniffin : 6/30/2014
carol : 3/5/2014
ckniffin : 3/4/2014
mcolton : 2/21/2014
carol : 9/14/2012
carol : 9/14/2012
terry : 9/13/2012
ckniffin : 9/13/2012
wwang : 6/11/2010
ckniffin : 6/10/2010
wwang : 5/25/2010
ckniffin : 5/24/2010
wwang : 3/31/2010
terry : 3/30/2010
wwang : 3/11/2010
mgross : 3/8/2010
wwang : 1/8/2010
ckniffin : 12/17/2009
wwang : 3/16/2005
wwang : 3/10/2005
ckniffin : 3/10/2005
carol : 10/31/2000
mgross : 7/21/1999
terry : 6/18/1998
supermim : 3/16/1992
supermim : 3/20/1990
ddp : 10/27/1989
root : 8/14/1989
root : 1/9/1989
root : 12/20/1988