1,044
Views
30
CrossRef citations to date
0
Altmetric
REVIEW ARTICLE

Voltage‐gated sodium channels: Action players with many faces

, &
Pages 472-482 | Published online: 08 Jul 2009

Abstract

Voltage‐gated sodium channels are responsible for the upstroke of the action potential and thereby play an important role in propagation of the electrical impulse in excitable tissues like muscle, nerve and the heart. Duplication of the sodium channels encoding genes during evolution generated the sodium channel gene family with the different isoforms differing in biophysical properties and tissue distribution. In this review article, mutations in these genes leading to various inherited disorders are discussed.

Abbreviations
BFNIS=

benign familial neonatal‐infantile seizures

BrS=

Brugada syndrome

CMD1E=

dilated cardiomyopathy‐1E

FEB3=

familial febrile convulsions‐3

FHM=

familial hemiplegic migraine

FRP=

familial rectal pain

GEFS+=

generalized epilepsy with febrile seizures plus

HOKPP=

hypokalemic periodic paralysis

HYPP=

hyperkalemic periodic paralysis

ICEGTC=

intractable childhood epilepsies and frequent generalized tonic‐clonic seizures

LQT3=

long QT syndrome type 3

LQTS=

long QT syndrome

PCCD=

progressive cardiac conduction defect

PMC=

paramyotonia congenital

SIDS=

sudden infant death syndrome

SMEB=

borderline SMEI

SMEI=

severe myoclonic epilepsy of infancy

SSS=

sick sinus syndrome

Introduction

Voltage‐gated sodium (Na+) channels are responsible for the upstroke of the action potential and thereby play an important role in propagation of the electrical excitable impulse in tissues including muscle, nerve and the heart. The channels are composed of pore‐forming α‐subunits of ∼260 kDa associated with one or two β‐subunits of 30–40 kDa that alter the properties of the channel. The α‐subunit gene family consists of nine genes (and one additional sodium channel‐like gene, see below) that are highly conserved across species. The channels are characterized by differential sensitivities to the sodium channel blocker tetrodotoxin (TTX) and inactivation kinetics: highly TTX‐sensitive α‐subunits (encoded by SCN1A, SCN2A, SCN3A, SCN4A, SCN8A, SCN9A) have faster inactivation kinetics compared to α‐subunits that are less sensitive to TTX (encoded by SCN5A, SCN10A, SCN11A).

Mutations in these genes cause various inherited disorders, which will be discussed in this review article.

The sodium channel genes

Duplication of genes encoding the α‐subunits during evolution generated the sodium channel gene family with the different genes differing in their biophysical properties and tissue distribution () Citation1. SCN9A, SCN10A and SCN11A are expressed in the peripheral nervous system. SCN1A, SCN2A, SCN3A and SCN8A are also expressed in the peripheral nervous system, but are more abundant in the central nervous system (including the brain). SCN4A and SCN5A are highly expressed in muscle: SCN4A is expressed in adult skeletal muscle and SCN5A in embryonic and denervated skeletal muscle and heart muscle. SCN6A/SCN7A (probably referring to the same gene) is expressed in a diversity of tissues, including peripheral nervous system, heart, uterus and skeletal muscle Citation2,3. Since SCN6A/SCN7A is the only sodium channel encoding gene that has not been expressed in an exogenous system and consensus amino acid sequences essential for voltage sensitivity and channel inactivation are not well conserved Citation4, it was debated whether this gene encodes a functional voltage‐gated Na+ channel. On the other hand, SCN6A/SCN7A has been implicated in regulation of salt‐intake behavior in a knockout mouse model Citation5,6.

Table I. The voltage‐gated sodium channel gene family: tissue specificity and clinical association of mutations in these genes.

Alternative RNA‐splicing has been described for several α‐subunits (): SCN1A (extended exon 11), SCN3A (three different variants of exon 12), SCN5A (exon 18, difference of one amino acid), SCN8A (extended exon 12), SCN9A (extended exon 11) and SCN11A (lacking exon 16) Citation7–13. Additionally, developmentally regulated splice variants of SCN2A (exon 5N and 5A), SCN3A (exon 5N and 5A), SCN8A (exon 5N and 5A) and SCN9A (exon 5N and 5A) that differ in a few amino acids and are either predominantly expressed neonatally (N) or in adults (A) have been described Citation10,Citation14–16. In the case of SCN8A, another splice variant (exon 18N and 18A) contains an in‐frame stop‐codon in the neonatally expressed variant which encodes a truncated two‐domain protein, the function of which is unknown Citation8. In addition, four alternatively spliced noncoding exons which generate alternative 5′UTRs for transcripts of this gene have been reported Citation17.

Table II. Different splice variants of voltage‐gated sodium channels.

Electrophysiological studies on splice variants of SCN2A did not show any functional differences Citation18, while splice variants of SCN3A, SCN5A and SCN8A resulted in channels that have altered kinetics Citation7,Citation12,Citation19. The functional significance of the other alternative splicing events is still unknown.

The four β‐subunit isoforms can be divided into two groups. β1 (SCN1B; localized in brain, skeletal and cardiac muscle Citation20) and β3 (SCN3B; localized primarily in neuronal tissue Citation21,22, but also detected in the heart Citation22–24) are most similar in amino acid sequence and are noncovalently associated with α‐subunits Citation22,Citation25. β2 (SCN2B; localized in the central nervous system and cardiac muscle) and β4 (SCN4B; localized in many tissues, including brain, heart, and skeletal muscle Citation26) subunits are also closely related in amino acid sequence to one another but as opposed to β1 and β3 are disulfide‐linked to the α‐subunits Citation21,Citation26. Alternative RNA‐splicing has also been described for β‐subunits: SCN1Ba and SCN1Bb differ beyond the immunoglobulin (Ig)‐loop region, resulting in a distinct C‐terminal cytoplasmic domain and a longer SCN1Bb transcript Citation27,28.

Key messages

  • Voltage‐gated sodium channels are expressed in excitable tissues namely heart, muscle and nerve.

  • Mutations in the genes encoding these channels lead to various clinical phenotypes.

  • Our understanding of the sodium channelopathies is complicated by several factors, among which is the complexity of the clinical phenotypes, the allelic heterogeneity involved, the diverse impact of the different mutations on channel function, and the contribution of other (yet‐unknown) genetic and environmental factors to the final form or severity of the disease.

Structure and function of sodium channel subunits

α‐Subunits

The pore‐forming α‐subunits are large transmembrane proteins that contain four structurally homologous domains (DI–DIV), each composed of six helical transmembrane segments (S1–S6) (see Catterall et al. for review Citation29; see ). The S5 and S6 segments and the P‐loop between them from each domain line the channel pore. The pore contains the selectivity filter also referred to as the DEKA ring (consisting of aspartic acid, glutamate, lysine, alanine; one of these amino acids per P‐loop), which attracts positive Na+ ions and excludes negatively charged ions Citation30. The lysine residue in the P‐loop of DIII is important for discrimination for Na+ over Ca2+Citation31,32.

Figure 1. Diagrammatic representation of voltage‐gated sodium channels showing locations of mutations causing channelopathies. A: SCN1A; generalized epilepsy with febrile seizures plus (blue), severe myoclonic epilepsy of infancy (red), intractable childhood epilepsy and frequent generalized tonic‐clonic seizures (orange), borderline severe myoclonic epilepsy of infancy (green), familial febrile convulsions‐3 (purple), familial autism (light blue), familial hemiplegic migraine (yellow) and mixed phenotypes (white). B: SCN2A; benign familial neonatal‐infantile seizures (red), severe myoclonic epilepsy of infancy (green), febrile and afebrile seizures (blue), and familial autism (orange). C: SCN4A; hyperkalemic periodic paralysis (red), hypokalemic periodic paralysis (green), paramyotonia congenita (orange), potassium‐aggravated myotonia (blue) and mixed phenotypes (white). D: SCN5A; Brugada syndrome (red), long QT syndrome type 3 (green), (progressive) cardiac conduction defects (orange), sick sinus syndrome (blue), atrial standstill (purple), sudden infant death syndrome (light blue), drug‐induced torsades de pointes (yellow), AV block (black), dilated cardiomyopathy (brown) and mixed phenotype (white). E: SCN8A; ataxia. F: SCN9A; primary erythermalgia.

Figure 1. Diagrammatic representation of voltage‐gated sodium channels showing locations of mutations causing channelopathies. A: SCN1A; generalized epilepsy with febrile seizures plus (blue), severe myoclonic epilepsy of infancy (red), intractable childhood epilepsy and frequent generalized tonic‐clonic seizures (orange), borderline severe myoclonic epilepsy of infancy (green), familial febrile convulsions‐3 (purple), familial autism (light blue), familial hemiplegic migraine (yellow) and mixed phenotypes (white). B: SCN2A; benign familial neonatal‐infantile seizures (red), severe myoclonic epilepsy of infancy (green), febrile and afebrile seizures (blue), and familial autism (orange). C: SCN4A; hyperkalemic periodic paralysis (red), hypokalemic periodic paralysis (green), paramyotonia congenita (orange), potassium‐aggravated myotonia (blue) and mixed phenotypes (white). D: SCN5A; Brugada syndrome (red), long QT syndrome type 3 (green), (progressive) cardiac conduction defects (orange), sick sinus syndrome (blue), atrial standstill (purple), sudden infant death syndrome (light blue), drug‐induced torsades de pointes (yellow), AV block (black), dilated cardiomyopathy (brown) and mixed phenotype (white). E: SCN8A; ataxia. F: SCN9A; primary erythermalgia.

Depending on the membrane potential, voltage‐gated Na+ channels can switch between three functional states: resting (closed), activated (open), and inactivated (closed). The highly conserved S4 region in each domain has a positive amino acid at every third position, and is considered the voltage sensor. The transition from the resting state to the activated state occurs when a change in transmembrane voltage moves S4 from within the pore towards the extracellular side of the cell, activating the channel which becomes permeable to ions Citation33. Inactivation is mediated mainly by the inactivation gate (DIII–DIV linker), that blocks the inside of the channel shortly after it has been activated, and the C‐terminal cytoplasmic domain Citation34–36. During an action potential the channel normally remains open for only a few milliseconds after depolarization before it is being inactivated. When the membrane potential becomes negative after the repolarization phase of the action potential, the channels return to their resting state and can be activated again during the next action potential.

β‐Subunits

β‐Subunits consist of one transmembrane segment, an intracellular domain and a glycosylated extracellular domain. The structure of the extracellular domain resembles the structure of the V‐like family of Ig‐fold proteins, containing domains similar to the variable regions of antibodies and including motifs as found in cell adhesion molecules Citation37.

The multifunctional β‐subunits control channel‐gating, regulate the level of expression of the α‐subunit at the plasma membrane Citation38,39 and are involved in cell adhesion through interaction with the cytoskeleton, extracellular matrix, and other cell adhesion molecules that regulate cell migration and aggregation Citation40.

Channelopathies: Cardiac sodium channelopathies

SCN5A (Figure)

Mutations in the cardiac voltage‐gated Na+ channel α‐subunit gene SCN5A result in multiple arrhythmia syndromes Citation41. Mutations leading to loss of Nav1.5 channel function can result in Brugada syndrome (BrS; MIM 601144) Citation42, (progressive) cardiac conduction defect (PCCD; MIM 113900) Citation43, sick sinus syndrome (SSS; MIM 608567) Citation44, sudden infant death syndrome (SIDS; MIM 272120) Citation45 and dilated cardiomyopathy associated with conduction defects and arrhythmias (CMD1E; MIM 601154) Citation46. In combination with modifier genes, a loss‐of‐function defect causes atrial standstill Citation47. Mutations leading to a gain‐of‐function of the channel cause long QT syndrome type 3 (LQT3; MIM 603830) Citation48. Some mutations in this gene lead to more than one disease phenotype, referred to as overlap syndromes of cardiac Na+ channelopathy, which are usually only recognized in large families Citation49,50.

Loss‐of‐function mutations: Brugada syndrome and conduction defects

The Brugada syndrome, with an estimated 5–50 cases per 10,000 individuals (with a higher incidence in Asia than in the United States and Europe Citation51), is an autosomal dominant disorder characterized by sudden cardiac death from ventricular tachyarrhythmias, in combination with a typical electrocardiogram (ECG) pattern of ST segment elevation in leads V1–V3. It is believed to cause 4%–12% of all sudden cardiac deaths and ∼20% of deaths in patients without structural abnormalities. To date, 80 mutations in SCN5A (of which 14% are nonsense or frameshift mutations, leading to truncation of the protein) have been described in BrS patients or in BrS patients with a mixed (overlap) phenotype (Inherited Arrhythmias Database: http://www.fsm.it/cardmoc/) Citation52. These loss‐of‐function mutations are associated with dysfunctional channels or with a reduction of membrane expression of the channel due to a trafficking defect. Loss of Na+ channel function reduces the upstroke of the action potential and may slow down action potential propagation. Thus, not surprisingly, patients with BrS often present with (progressive) conduction defects Citation53,54. Loss‐of‐function mutations in SCN5A can also cause isolated cardiac conduction disease, i.e. without ECG features of BrS. Recently, a haplotype in the promoter region of SCN5A that occurs frequently in Asians was found to be associated with slower cardiac conduction Citation55, suggesting that decreased expression of SCN5A transcripts may contribute to differences in BrS prevalence as a function of ethnicity.

Gain‐of‐function mutations: long QT syndrome

Multiple genes have been associated with LQTS, an inherited cardiac arrhythmia associated with syncope and sudden death from torsades de pointes polymorphic ventricular tachycardia, estimated to affect 1 per 5,000 individuals. One subtype of this syndrome is associated with mutations in SCN5A (LQT3). Gain‐of‐function mutations in Nav1.5 result in an increase in the late component of the Na+ current by slowing of inactivation or an increase in the reversibility of inactivation, resulting in a slow and constant entry of Na+ in the plateau phase of the action potential, leading to a prolonged QT interval on the surface electrocardiogram (ECG) Citation56. Because in SCN5A‐related LQTS QT‐prolongation is most pronounced at lower heart rates, bradycardia presents an important factor in developing lethal arrhythmias in LQTS families with mutations in SCN5ACitation57. Thus far, 62 LQTS‐causing missense mutations and small (in frame) insertions and deletions have been identified in SCN5A (Inherited Arrhythmias Database: http://www.fsm.it/cardmoc/).

SCN4B

Very recently, the first mutation in the SCN4B gene encoding Navβ.4 was presented Citation58. This missense mutation functionally disturbs Nav1.5 in a LQTS patient and therefore SCN4B very likely is a new LQTS susceptibility gene.

Neuronal channelopathies

Brain sodium channelopathies

SCN1A (Figure)

SCN1A, the neuronal voltage‐gated Na+ channel α‐subunit gene encoding Nav1.1, is part of the SCN1ASCN2ASCN3A gene cluster on chromosome 2q24. Missense mutations have been identified in patients with generalized epilepsy with febrile seizures plus (GEFS+; MIM 604233) Citation59, an autosomal dominant epilepsy characterized by febrile seizures in children, and afebrile seizures in adults Citation60,61. A large number of mutations of SCN1A have been identified in patients with severe myoclonic epilepsy of infancy (SMEI; MIM 607208) Citation62, a rare disorder characterized by various types of generalized and partial seizures, including myoclonic seizures. Many of the mutations characterized in children with SMEI are de novo (69 out of 75 cases) Citation63. Also, many of the observed mutations in SMEI patients are nonsense or frameshift mutations that cause protein truncation Citation64, with deletion of the C‐terminal cytoplasmic domain resulting in disease of similar severity to deletion of the N‐terminal cytoplasmic domain Citation63. No nonsense or frameshift mutations in SCN1A have been described so far in patients with GEFS+. One report suggests that in SMEI, pore region missense mutations are associated with a more severe phenotype Citation65.

SCN1A mutations have also been associated with intractable childhood epilepsies and frequent generalized tonic‐clonic seizures (ICEGTC) Citation66, familial febrile convulsions‐3 (FEB3; MIM: 604403) Citation67, and borderline SMEI (SMEB, when not all the SMEI criteria are fulfilled) Citation68.

In 2005, Dichgans et al. Citation69 described a mutation in SCN1A leading to familial hemiplegic migraine (FHM; MIM 141500), an autosomal dominant severe subtype of migraine with aura. This gain‐of‐function missense mutation in the inactivation gate of the channel was present in three families with the same disorder. This finding underlines the molecular links between migraine and epilepsy, two common paroxysmal disorders.

The association of SCN1A mutations with familial autism is also being investigated Citation70.

SCN2A (Figure)

Despite the fact that SCN1A and SCN2A are closely related genes, only a few epilepsy mutations have been identified in SCN2A encoding Nav1.2 Citation14,Citation71,72. Loss‐of‐function missense mutations in SCN2A were found in patients with benign familial neonatal‐infantile seizures (BFNIS; MIM 607745), a mild autosomal dominant syndrome in which afebrile seizures occur in clusters during the first year of life but which does not progress to adult epilepsy. Interestingly, one mutation (located in the conserved transmembrane segment S4 of DIII) was identified in affected members of three families, which occurred independently according to haplotype analysis Citation73.

Unlike SCN1A, SCN2A does not show evidence of haploinsufficiency. Only one (de novo) truncation mutation has been identified in SCN2A in a patient with intractable epilepsy and mental decline, a severe form of epilepsy resembling SMEI Citation71. In this case the truncated protein had a dominant‐negative effect possibly arising from direct or indirect cytoskeletal interactions of the mutant protein.

One mutation in SCN2A has been described that might play a role in autism, but needs to be further analyzed Citation70.

SCN3A

Studies in rats have indicated that seizure activity induces alterations in the developmental splicing of neonatal and adult SCN2A and SCN3A transcripts Citation74, whose genes are located side by side on the chromosome: seizures were found to re‐activate the neonatal splicing event, causing an increase in the presence of the neonatal SCN2A and SCN3A transcripts in localized regions of the adult rat brain. In contrast to SCN2A, SCN3A mRNA was found to be expressed at significantly higher levels in CA4 hilar cells in the epileptic hippocampus when compared with control, and therefore possibly contributes to the pathophysiology of epilepsy Citation75. However, no mutations in this gene have been reported.

SCN8A (Figure)

Mutations in the mouse ortholog of SCN8A cause ataxia and other movement disorders Citation76. So far, only one protein truncation mutation has been described in humans, causing cerebellar atrophy, ataxia and mental retardation Citation77. This loss‐of‐function frameshift mutation was located in the pore loop of domain IV, resulting in truncation of the C‐terminal cytoplasmic domain. Interestingly, Wasserman et al. described a single‐nucleotide polymorphism in SCN8A that may contribute to the risk for suicide attempts, possibly through alterations in neuronal conduction which hypothetically could lead to disturbed analysis of incoming information in periods of emotional and physical stress Citation78.

SCN1B

Mutations in the Navβ.1 encoding SCN1B gene are associated with GEFS+ Citation79. The loss‐of‐function mutation C121W has been described in two families with GEFS+ Citation80. In another family with febrile seizures plus and early‐onset absence epilepsy, a mutation in the splice acceptor site of SCN1B that predicts a deletion of five amino acids in the extracellular Ig‐loop region was identified that could lead to loss‐of‐function Citation81. Both mutations are expected to disrupt proper folding of the protein and therefore can inhibit interaction with α‐subunits or impair subcellular distribution, which will reduce the inactivation rate of Na+ channels and results in neuronal hyperexcitability Citation81,82.

To study the loss‐of‐function effects of SCN1B in vivo, knockout mice were generated, which appear ataxic and reveal spontaneous seizures, growth retardation and premature death Citation83. These phenotypes were the result of slowing of neuronal action potential conduction, reduced number of mature nodes of Ranvier, alterations in nodal architecture, loss of Na+ channel‐contactin interactions, and abnormalities in the expression of SCN1A and SCN3A. From this, it was clear that SCN1B regulates Na+ channel density and localization, is involved in axo‐glial communication at nodes of Ranvier, and is required for normal action potential conduction and control of excitability in vivo.

Peripheral nerve sodium channelopathies

SCN9A (Figure)

Mutations in the SCN9A gene encoding the Nav1.7 channel cause primary erythermalgia (MIM 133020), a rare autosomal dominant disorder characterized by sporadic intense burning pain with redness and heat in the extremities Citation84. These gain‐of‐function mutations modify thresholds of activation and are therefore likely to contribute to increased excitability of spinal sensory neurons that express the channels and may cause the abnormal pain sensations in patients suffering from this disorder Citation85,86. Missense mutations in SCN9A have also been associated with familial rectal pain (FRP; MIM 167400), a disorder characterized by brief episodes of excruciating pain of the submandibular, ocular, and rectal areas with flushing of the surrounding skin Citation87. Additionally, conditional inactivation of SCN9A in sensory neurons of the mouse resulted in increased threshold for mechanical, thermal, and inflammatory pain Citation88. The association of SCN9A mutations with pain syndromes shows that this Na+ channel could be a target for local anesthetics.

SCN10A

SCN10A encodes Nav1.8, a channel that is restricted to the peripheral sensory nervous system Citation89. The down‐regulation of SCN10A expression in rat can 1) prevent thermal hyperalgesia (hypersensitivity to noxious stimuli) and allodynia (pain response to non‐noxious stimuli) in a rat model of neuropathic pain Citation90, and can 2) suppress responses caused by pain in a rat model of visceral pain Citation91. No mutations in humans have been described thus far, but SCN10A could be a potential target for analgesic drugs.

SCN11A

The Nav1.9 channel encoded by SCN11A is expected to contribute to setting the resting membrane potential and modulating subthreshold electrogenesis in nociceptive neurons Citation92. Its expression is adjusted in response to axotomy Citation92 and inflammation Citation93. Priest et al. Citation94 observed no differences in passive membrane properties and action potential characteristics between acutely dissociated peripheral sensory neurons in the dorsal root ganglia between wildtype and NaV1.9 knockout mice. However, expression of SCN11A contributes to the persistent thermal hypersensitivity and spontaneous pain behavior after peripheral administration of inflammatory agents Citation94. Although no mutations in SCN11A in humans have been described to date, this gene can possibly act as a target for analgesic drugs.

Muscle sodium channelopathies

SCN4A (Figure)

Mutations in the muscle voltage‐gated Na+ channel Nav1.4 encoding gene SCN4A have been identified in a group of related muscular disorders, including hyperkalemic periodic paralysis (HYPP; MIM 170500) Citation95 and hypokalemic periodic paralysis (HOKPP; MIM 170400) Citation96, paramyotonia congenita (PMC; MIM 168300) Citation97, and a group of disorders classified as potassium‐aggravated myotonia (MIM 608390) Citation98. The gain‐of‐function mutations in SCN4A associated with HYPP, PMC and myotonia cause a disruption of fast inactivation Citation99–101, which results in channel re‐opening and intracellular Na+ accumulation. In that case, muscle cells depolarize and generate recurrent action potentials. This can lead to enduring hyperexcitability which causes myotonia Citation102, or it can lead to general opening of the channel which can cause paralysis Citation103. In contrast, HOKPP is associated with loss‐of‐function mutations leading to hypoexcitability of the fiber membrane resulting in muscle weakness Citation104–106. However, the mechanism leading to this decreased excitability is still poorly understood.

Discussion

In the last decades, much research has been done to identify new phenotype‐linked mutations in voltage‐gated Na+ channels. Despite the fact that the relationship between mutations, altered Na+ channel function and disease phenotypes has become clearer, many questions still remain. Our understanding of the sodium channelopathies is complicated by several factors, among which is the complexity of the clinical phenotypes (pleiotropy), the allelic heterogeneity involved, the diverse impact of the different mutations on channel function, and the contribution of other (yet‐unknown) genetic and environmental factors to the final form or severity of the disease Citation107,108. Gain‐of‐function and loss‐of‐function mutations in the same gene lead to different diseases. In contrast to the obvious association between truncation mutations and loss of channel function, no relationships have been described so far between the locations within the channel protein of amino‐acid‐changing mutations and these mechanisms, which appear to be random. In rare instances, a given mutation may even harbor biophysical defects associated with both gain and loss of channel function Citation63.

Genetic factors other than the causal mutation itself that play a role in modulation of disease severity are starting to be uncovered. For example, the combined effect of a mutation in SCN5A and polymorphisms in the atrial‐specific gap junction protein connexin40 gene has been reported to cause familial atrial standstill Citation47. In mice, genetic variation in a putative RNA‐splicing factor (SCNM1) has been shown to modulate movement disorder severity in SCN8A mutant mice Citation109. In another mouse study, an interaction between two mild mutations, one in SCN2A and the other in the potassium channel gene KCNQ2, has been described to result in severe epilepsy Citation110. Furthermore, two modifier loci affecting epilepsy severity caused by a SCN2A mutation have been reported Citation111. The genetic factors modulating disease severity may also reside in the gene affected itself Citation55 and even on the same allele. In SCN5A, for example, a common polymorphism that attenuates the biophysical defect of a mutation on the same allele has been described Citation112.

Not all voltage‐gated Na+ channels have been linked to human disease. Although no mutations have been described in SCN3A, SCN10A and SCN11A, in vivo animal studies indicate that other Na+ channelopathies probably exist. Furthermore, the different splice variants of the Na+ channel genes could play a role in the pathogenesis of the sodium channelopathies.

Since voltage‐gated Na+ channels are action players with many faces, our understanding of how mutations cause disease has lagged somewhat behind. In the next years, our increasing knowledge may lead to better targeted treatment of patients suffering from these disorders.

Acknowledgements

The authors wish to thank A.C. Linnenbank for preparing the figure and A.C.G. van Ginneken for critical reading of the manuscript. This work is supported by Netherlands Heart Foundation grants 2003B195 (CRB) and 2003T302 (AAMW), and the Interuniversity Cardiology Institute of the Netherlands (project 27, AAMW). Connie R. Bezzina is an Established Investigator of the Netherlands Heart Foundation (Grant 2005/T024).

References

  • Plummer N. W., Meisler M. H. Evolution and diversity of mammalian sodium channel genes. Genomics 1999; 57: 323–31
  • George A. L., Jr., Knittle T. J., Tamkun M. M. Molecular cloning of an atypical voltage‐gated sodium channel expressed in human heart and uterus: evidence for a distinct gene family. Proc Natl Acad Sci U S A 1992; 89: 4893–7
  • Watanabe E., Hiyama T. Y., Kodama R., Noda M. NaX sodium channel is expressed in non‐myelinating Schwann cells and alveolar type II cells in mice. Neurosci Lett 2002; 330: 109–13
  • Goldin A. L. Resurgence of sodium channel research. Annu Rev Physiol 2001; 63: 871–94
  • Hiyama T. Y., Watanabe E., Okado H., Noda M. The subfornical organ is the primary locus of sodium‐level sensing by Na(x) sodium channels for the control of salt‐intake behavior. J Neurosci 2004; 24: 9276–81
  • Watanabe E., Fujikawa A., Matsunaga H., Yasoshima Y., Sako N., Yamamoto T., et al. Nav2/NaG channel is involved in control of salt‐intake behavior in the CNS. J Neurosci 2000; 20: 7743–51
  • Makielski J. C., Ye B., Valdivia C. R., Pagel M. D., Pu J., Tester D. J., et al. A ubiquitous splice variant and a common polymorphism affect heterologous expression of recombinant human SCN5A heart sodium channels. Circ Res 2003; 93: 821–8
  • Plummer N. W., McBurney M. W., Meisler M. H. Alternative splicing of the sodium channel SCN8A predicts a truncated two‐domain protein in fetal brain and non‐neuronal cells. J Biol Chem 1997; 272: 24008–15
  • Plummer N. W., Galt J., Jones J. M., Burgess D. L., Sprunger L. K., Kohrman D. C., et al. Exon organization, coding sequence, physical mapping, and polymorphic intragenic markers for the human neuronal sodium channel gene SCN8A. Genomics 1998; 54: 287–96
  • Raymond C. K., Castle J., Garrett‐Engele P., Armour C. D., Kan Z., Tsinoremas N., et al. Expression of alternatively spliced sodium channel alpha‐subunit genes. Unique splicing patterns are observed in dorsal root ganglia. J Biol Chem 2004; 279: 46234–41
  • Schaller K. L., Krzemien D. M., McKenna N. M., Caldwell J. H. Alternatively spliced sodium channel transcripts in brain and muscle. J Neurosci 1992; 12: 1370–81
  • Thimmapaya R., Neelands T., Niforatos W., vis‐Taber R. A., Choi W., Putman C. B., et al. Distribution and functional characterization of human Nav1.3 splice variants. Eur J Neurosci 2005; 22: 1–9
  • Dietrich P. S., McGivern J. G., Delgado S. G., Koch B. D., Eglen R. M., Hunter J. C., et al. Functional analysis of a voltage‐gated sodium channel and its splice variant from rat dorsal root ganglia. J Neurochem 1998; 70: 2262–72
  • Heron S. E., Crossland K. M., Andermann E., Phillips H. A., Hall A. J., Bleasel A., et al. Sodium‐channel defects in benign familial neonatal‐infantile seizures. Lancet 2002; 360: 851–2
  • Kasai N., Fukushima K., Ueki Y., Prasad S., Nosakowski J., Sugata K., et al. Genomic structures of SCN2A and SCN3A. Gene 2001; 264: 113–22
  • Lu C. M., Brown G. B. Isolation of a human‐brain sodium‐channel gene encoding two isoforms of the subtype III alpha‐subunit. J Mol Neurosci 1998; 10: 67–70
  • Drews V. L., Lieberman A. P., Meisler M. H. Multiple transcripts of sodium channel SCN8A (Na(V)1.6) with alternative 5′‐ and 3′‐untranslated regions and initial characterization of the SCN8A promoter. Genomics 2005; 85: 245–57
  • Auld V. J., Goldin A. L., Krafte D. S., Catterall W. A., Lester H. A., Davidson N., et al. A neutral amino acid change in segment IIS4 dramatically alters the gating properties of the voltage‐dependent sodium channel. Proc Natl Acad Sci U S A 1990; 87: 323–7
  • Tan J., Liu Z., Nomura Y., Goldin A. L., Dong K. Alternative splicing of an insect sodium channel gene generates pharmacologically distinct sodium channels. J Neurosci 2002; 22: 5300–9
  • Makita N., Bennett P. B., Jr., George A. L., Jr. Voltage‐gated Na+ channel beta 1 subunit mRNA expressed in adult human skeletal muscle, heart, and brain is encoded by a single gene. J Biol Chem 1994; 269: 7571–8
  • Isom L. L., Ragsdale D. S., De Jongh K. S., Westenbroek R. E., Reber B. F., Scheuer T., et al. Structure and function of the beta 2 subunit of brain sodium channels, a transmembrane glycoprotein with a CAM motif. Cell 1995; 83: 433–42
  • Morgan K., Stevens E. B., Shah B., Cox P. J., Dixon A. K., Lee K., et al. beta 3: an additional auxiliary subunit of the voltage‐sensitive sodium channel that modulates channel gating with distinct kinetics. Proc Natl Acad Sci U S A 2000; 97: 2308–13
  • Dhar M. J., Chen C., Rivolta I., Abriel H., Malhotra R., Mattei L. N., et al. Characterization of sodium channel alpha‐ and beta‐subunits in rat and mouse cardiac myocytes. Circulation 2001; 103: 1303–10
  • Fahmi A. I., Patel M., Stevens E. B., Fowden A. L., John J. E., III., Lee K., et al. The sodium channel beta‐subunit SCN3b modulates the kinetics of SCN5a and is expressed heterogeneously in sheep heart. J Physiol 2001; 537: 693–700
  • Isom L. L., De Jongh K. S., Patton D. E., Reber B. F., Offord J., Charbonneau H., et al. Primary structure and functional expression of the beta 1 subunit of the rat brain sodium channel. Science 1992; 256: 839–42
  • Yu F. H., Westenbroek R. E., Silos‐Santiago I., McCormick K. A., Lawson D., Ge P., et al. Sodium channel beta4, a new disulfide‐linked auxiliary subunit with similarity to beta2. J Neurosci 2003; 23: 7577–85
  • Kazen‐Gillespie K. A., Ragsdale D. S., D'Andrea M. R., Mattei L. N., Rogers K. E., Isom L. L. Cloning, localization, and functional expression of sodium channel beta1A subunits. J Biol Chem 2000; 275: 1079–88
  • Qin N., D'Andrea M. R., Lubin M. L., Shafaee N., Codd E. E., Correa A. M. Molecular cloning and functional expression of the human sodium channel beta1B subunit, a novel splicing variant of the beta1 subunit. Eur J Biochem 2003; 270: 4762–70
  • Catterall W. A. From ionic currents to molecular mechanisms: the structure and function of voltage‐gated sodium channels. Neuron 2000; 26: 13–25
  • Backx P. H., Yue D. T., Lawrence J. H., Marban E., Tomaselli G. F. Molecular localization of an ion‐binding site within the pore of mammalian sodium channels. Science 1992; 257: 248–51
  • Heinemann S. H., Terlau H., Stuhmer W., Imoto K., Numa S. Calcium channel characteristics conferred on the sodium channel by single mutations. Nature 1992; 356: 441–3
  • Perez‐Garcia M. T., Chiamvimonvat N., Ranjan R., Balser J. R., Tomaselli G. F., Marban E. Mechanisms of sodium/calcium selectivity in sodium channels probed by cysteine mutagenesis and sulfhydryl modification. Biophys J 1997; 72: 989–96
  • Yang N., Horn R. Evidence for voltage‐dependent S4 movement in sodium channels. Neuron 1995; 15: 213–8
  • Eaholtz G., Zagotta W. N., Catterall W. A. Kinetic analysis of block of open sodium channels by a peptide containing the isoleucine, phenylalanine, and methionine (IFM) motif from the inactivation gate. J Gen Physiol 1998; 111: 75–82
  • West J. W., Patton D. E., Scheuer T., Wang Y., Goldin A. L., Catterall W. A. A cluster of hydrophobic amino acid residues required for fast Na(+)‐channel inactivation. Proc Natl Acad Sci U S A 1992; 89: 10910–4
  • Motoike H. K., Liu H., Glaaser I. W., Yang A. S., Tateyama M., Kass R. S. The Na+ channel inactivation gate is a molecular complex: a novel role of the COOH‐terminal domain. J Gen Physiol 2004; 123: 155–65
  • Isom L. L., Catterall W. A. Na+ channel subunits and Ig domains. Nature 1996; 383: 307–8
  • Goldin A. L. Accessory subunits and sodium channel inactivation. Curr Opin Neurobiol 1993; 3: 272–7
  • Isom L. L. I. Cellular and molecular biology of sodium channel beta‐subunits: therapeutic implications for pain?. Am J Physiol Gastrointest Liver Physiol 2000; 278: G349–G353
  • Isom L. L. Sodium channel beta subunits: anything but auxiliary. Neuroscientist 2001; 7: 42–54
  • Tan H. L., Bezzina C. R., Smits J. P., Verkerk A. O., Wilde A. A. Genetic control of sodium channel function. Cardiovasc Res 2003; 57: 961–73
  • Chen Q., Kirsch G. E., Zhang D., Brugada R., Brugada J., Brugada P., et al. Genetic basis and molecular mechanism for idiopathic ventricular fibrillation. Nature 1998; 392: 293–6
  • Schott J. J., Alshinawi C., Kyndt F., Probst V., Hoorntje T. M., Hulsbeek M., et al. Cardiac conduction defects associate with mutations in SCN5A. Nat Genet 1999; 23: 20–1
  • Benson D. W., Wang D. W., Dyment M., Knilans T. K., Fish F. A., Strieper M. J., et al. Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). J Clin Invest 2003; 112: 1019–28
  • Ackerman M. J., Siu B. L., Sturner W. Q., Tester D. J., Valdivia C. R., Makielski J. C., et al. Postmortem molecular analysis of SCN5A defects in sudden infant death syndrome. JAMA 2001; 286: 2264–9
  • McNair W. P., Ku L., Taylor M. R., Fain P. R., Dao D., Wolfel E., et al. SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia. Circulation 2004; 110: 2163–7
  • Groenewegen W. A., Firouzi M., Bezzina C. R., Vliex S., van L., I., Sandkuijl L., et al. A cardiac sodium channel mutation cosegregates with a rare connexin40 genotype in familial atrial standstill. Circ Res 2003; 92: 14–22
  • Wang Q., Shen J., Li Z., Timothy K., Vincent G. M., Priori S. G., et al. Cardiac sodium channel mutations in patients with long QT syndrome, an inherited cardiac arrhythmia. Hum Mol Genet 1995; 4: 1603–7
  • Bezzina C. R., Rook M. B., Wilde A. A. Cardiac sodium channel and inherited arrhythmia syndromes. Cardiovasc Res 2001; 49: 257–71
  • Bezzina C., Veldkamp M. W., van den Berg M. P., Postma A. V., Rook M. B., Viersma J. W., et al. A single Na(+) channel mutation causing both long‐QT and Brugada syndromes. Circ Res 1999; 85: 1206–13
  • Antzelevitch C., Brugada P., Borggrefe M., Brugada J., Brugada R., Corrado D., et al. Brugada syndrome: report of the second consensus conference. Heart Rhythm 2005; 2: 429–40
  • Meregalli P. G., Wilde A. A., Tan H. L. Pathophysiological mechanisms of Brugada syndrome: depolarization disorder, repolarization disorder, or more?. Cardiovasc Res 2005; 67: 367–78
  • Probst V., Allouis M., Sacher F., Pattier S., Babuty D., Mabo P., et al. Progressive Cardiac Conduction Defect is the Prevailing Phenotype in Carriers of a Brugada Syndrome SCN5A Mutation. J Cardiovasc Electrophysiol 2006; 17: 270–5
  • Smits J. P., Eckardt L., Probst V., Bezzina C. R., Schott J. J., Remme C. A., et al. Genotype‐phenotype relationship in Brugada syndrome: electrocardiographic features differentiate SCN5A‐related patients from non‐SCN5A‐related patients. J Am Coll Cardiol 2002; 40: 350–6
  • Bezzina C. R., Shimizu W., Yang P., Koopmann T. T., Tanck M. W., Miyamoto Y., et al. Common sodium channel promoter haplotype in asian subjects underlies variability in cardiac conduction. Circulation 2006; 113: 338–44
  • Bennett P. B., Yazawa K., Makita N., George A. L., Jr. Molecular mechanism for an inherited cardiac arrhythmia. Nature 1995; 376: 683–5
  • Veldkamp M. W., Wilders R., Baartscheer A., Zegers J. G., Bezzina C. R., Wilde A. A. Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in LQT3 families. Circ Res 2003; 92: 976–83
  • Domingo AM, Kaku T, Tester DJ, Torres PI, Itty A, Ye B, et al. Sodium Channel ß4 Subunit Mutation Causes Congenital Long QT Syndrome. Program no 16‐6 presented at the 27th Annual Scientific Sessions of the Heart Rhythm Society, 2006, May 18. Boston, MA, USA, http://www.heartrhythm2006.org/Heart_Rhythm_2006_Program.asp
  • Escayg A., MacDonald B. T., Meisler M. H., Baulac S., Huberfeld G., An‐Gourfinkel I., et al. Mutations of SCN1A, encoding a neuronal sodium channel, in two families with GEFS+2. Nat Genet 2000; 24: 343–5
  • Sugawara T., Mazaki‐Miyazaki E., Ito M., Nagafuji H., Fukuma G., Mitsudome A., et al. Nav1.1 mutations cause febrile seizures associated with afebrile partial seizures. Neurology 2001; 57: 703–5
  • Wallace R. H., Scheffer I. E., Barnett S., Richards M., Dibbens L., Desai R. R., et al. Neuronal sodium‐channel alpha1‐subunit mutations in generalized epilepsy with febrile seizures plus. Am J Hum Genet 2001; 68: 859–65
  • Nabbout R., Gennaro E., Dalla B. B., Dulac O., Madia F., Bertini E., et al. Spectrum of SCN1A mutations in severe myoclonic epilepsy of infancy. Neurology 2003; 60: 1961–7
  • Meisler M. H., Kearney J. A. Sodium channel mutations in epilepsy and other neurological disorders. J Clin Invest 2005; 115: 2010–7
  • Wallace R. H., Hodgson B. L., Grinton B. E., Gardiner R. M., Robinson R., Rodriguez‐Casero V., et al. Sodium channel alpha1‐subunit mutations in severe myoclonic epilepsy of infancy and infantile spasms. Neurology 2003; 61: 765–9
  • Kanai K., Hirose S., Oguni H., Fukuma G., Shirasaka Y., Miyajima T., et al. Effect of localization of missense mutations in SCN1A on epilepsy phenotype severity. Neurology 2004; 63: 329–34
  • Fujiwara T., Sugawara T., Mazaki‐Miyazaki E., Takahashi Y., Fukushima K., Watanabe M., et al. Mutations of sodium channel alpha subunit type 1 (SCN1A) in intractable childhood epilepsies with frequent generalized tonic‐clonic seizures. Brain 2003; 126: 531–46
  • Mantegazza M., Gambardella A., Rusconi R., Schiavon E., Annesi F., Cassulini R. R., et al. Identification of an Nav1.1 sodium channel (SCN1A) loss‐of‐function mutation associated with familial simple febrile seizures. Proc Natl Acad Sci U S A 2005; 102: 18177–82
  • Fukuma G., Oguni H., Shirasaka Y., Watanabe K., Miyajima T., Yasumoto S., et al. Mutations of neuronal voltage‐gated Na+ channel alpha 1 subunit gene SCN1A in core severe myoclonic epilepsy in infancy (SMEI) and in borderline SMEI (SMEB). Epilepsia 2004; 45: 140–8
  • Dichgans M., Freilinger T., Eckstein G., Babini E., Lorenz‐Depiereux B., Biskup S., et al. Mutation in the neuronal voltage‐gated sodium channel SCN1A in familial hemiplegic migraine. Lancet 2005; 366: 371–7
  • Weiss L. A., Escayg A., Kearney J. A., Trudeau M., MacDonald B. T., Mori M., et al. Sodium channels SCN1A, SCN2A and SCN3A in familial autism. Mol Psychiatry 2003; 8: 186–94
  • Kamiya K., Kaneda M., Sugawara T., Mazaki E., Okamura N., Montal M., et al. A nonsense mutation of the sodium channel gene SCN2A in a patient with intractable epilepsy and mental decline. J Neurosci 2004; 24: 2690–8
  • Sugawara T., Tsurubuchi Y., Agarwala K. L., Ito M., Fukuma G., Mazaki‐Miyazaki E., et al. A missense mutation of the Na+ channel alpha II subunit gene Na(v)1.2 in a patient with febrile and afebrile seizures causes channel dysfunction. Proc Natl Acad Sci U S A 2001; 98: 6384–9
  • Berkovic S. F., Heron S. E., Giordano L., Marini C., Guerrini R., Kaplan R. E., et al. Benign familial neonatal‐infantile seizures: characterization of a new sodium channelopathy. Ann Neurol 2004; 55: 550–7
  • Gastaldi M., Bartolomei F., Massacrier A., Planells R., Robaglia‐Schlupp A., Cau P. Increase in mRNAs encoding neonatal II and III sodium channel alpha‐isoforms during kainate‐induced seizures in adult rat hippocampus. Brain Res Mol Brain Res 1997; 44: 179–90
  • Whitaker W. R., Faull R. L., Dragunow M., Mee E. W., Emson P. C., Clare J. J. Changes in the mRNAs encoding voltage‐gated sodium channel types II and III in human epileptic hippocampus. Neuroscience 2001; 106: 275–85
  • Meisler M. H., Plummer N. W., Burgess D. L., Buchner D. A., Sprunger L. K. Allelic mutations of the sodium channel SCN8A reveal multiple cellular and physiological functions. Genetica 2004; 122: 37–45
  • Trudeau M. M., Dalton J. C., Day J. W., Ranum L. P., Meisler M. H. Heterozygosity for a protein truncation mutation of sodium channel SCN8A in a patient with cerebellar atrophy, ataxia and mental retardation. J Med Genet 2005; 43: 527–30
  • Wasserman D., Geijer T., Rozanov V., Wasserman J. Suicide attempt and basic mechanisms in neural conduction: relationships to the SCN8A and VAMP4 genes. Am J Med Genet B Neuropsychiatr Genet 2005; 133: 116–9
  • Wallace R. H., Wang D. W., Singh R., Scheffer I. E., George A. L., Jr., Phillips H. A., et al. Febrile seizures and generalized epilepsy associated with a mutation in the Na+‐channel beta1 subunit gene SCN1B. Nat Genet 1998; 19: 366–70
  • Wallace R. H., Scheffer I. E., Parasivam G., Barnett S., Wallace G. B., Sutherland G. R., et al. Generalized epilepsy with febrile seizures plus: mutation of the sodium channel subunit SCN1B. Neurology 2002; 58: 1426–9
  • Audenaert D., Claes L., Ceulemans B., Lofgren A., Van B. C., De J. P. A deletion in SCN1B is associated with febrile seizures and early‐onset absence epilepsy. Neurology 2003; 61: 854–6
  • Meadows L. S., Malhotra J., Loukas A., Thyagarajan V., Kazen‐Gillespie K. A., Koopman M. C., et al. Functional and biochemical analysis of a sodium channel beta1 subunit mutation responsible for generalized epilepsy with febrile seizures plus type 1. J Neurosci 2002; 22: 10699–709
  • Chen C., Westenbroek R. E., Xu X., Edwards C. A., Sorenson D. R., Chen Y., et al. Mice lacking sodium channel beta1 subunits display defects in neuronal excitability, sodium channel expression, and nodal architecture. J Neurosci 2004; 24: 4030–42
  • Yang Y., Wang Y., Li S., Xu Z., Li H., Ma L., et al. Mutations in SCN9A, encoding a sodium channel alpha subunit, in patients with primary erythermalgia. J Med Genet 2004; 41: 171–4
  • Cummins T. R., Dib‐Hajj S. D., Waxman S. G. Electrophysiological properties of mutant Nav1.7 sodium channels in a painful inherited neuropathy. J Neurosci 2004; 24: 8232–6
  • Dib‐Hajj S. D., Rush A. M., Cummins T. R., Hisama F. M., Novella S., Tyrrell L., et al. Gain‐of‐function mutation in Nav1.7 in familial erythromelalgia induces bursting of sensory neurons. Brain 2005; 128: 1847–54
  • Fertleman CR, Rees M, Parker KA, Barlow E, Gardiner RM. Identification of the gene underlying an inherited disorder of pain sensation. Program no 197 presented at the 54th Annual Meeting of the American Society of Human Genetics, 2004, October 30. Bethesda, MD, USA. http://www.ashg.org/genetics/ashg04s/index.shtml
  • Nassar M. A., Levato A., Stirling L. C., Wood J. N. Neuropathic pain develops normally in mice lacking both Nav1.7 and Nav1.8. Mol Pain 2005; 1: 24
  • Rabert D. K., Koch B. D., Ilnicka M., Obernolte R. A., Naylor S. L., Herman R. C., et al. A tetrodotoxin‐resistant voltage‐gated sodium channel from human dorsal root ganglia, hPN3/SCN10A. Pain 1998; 78: 107–14
  • Lai J., Gold M. S., Kim C. S., Bian D., Ossipov M. H., Hunter J. C., et al. Inhibition of neuropathic pain by decreased expression of the tetrodotoxin‐resistant sodium channel, NaV1.8. Pain 2002; 95: 143–52
  • Yoshimura N., Seki S., Novakovic S. D., Tzoumaka E., Erickson V. L., Erickson K. A., et al. The involvement of the tetrodotoxin‐resistant sodium channel Na(v)1.8 (PN3/SNS) in a rat model of visceral pain. J Neurosci 2001; 21: 8690–6
  • Dib‐Hajj S. D., Tyrrell L., Black J. A., Waxman S. G. NaN, a novel voltage‐gated Na channel, is expressed preferentially in peripheral sensory neurons and down‐regulated after axotomy. Proc Natl Acad Sci U S A 1998; 95: 8963–8
  • Tate S., Benn S., Hick C., Trezise D., John V., Mannion R. J., et al. Two sodium channels contribute to the TTX‐R sodium current in primary sensory neurons. Nat Neurosci 1998; 1: 653–5
  • Priest B. T., Murphy B. A., Lindia J. A., Diaz C., Abbadie C., Ritter A. M., et al. Contribution of the tetrodotoxin‐resistant voltage‐gated sodium channel NaV1.9 to sensory transmission and nociceptive behavior. Proc Natl Acad Sci U S A 2005; 102: 9382–7
  • Fontaine B., Khurana T. S., Hoffman E. P., Bruns G. A., Haines J. L., Trofatter J. A., et al. Hyperkalemic periodic paralysis and the adult muscle sodium channel alpha‐subunit gene. Science 1990; 250: 1000–2
  • Bulman D. E., Scoggan K. A., van Oene M. D., Nicolle M. W., Hahn A. F., Tollar L. L., et al. A novel sodium channel mutation in a family with hypokalemic periodic paralysis. Neurology 1999; 53: 1932–6
  • Ptacek L. J., Trimmer J. S., Agnew W. S., Roberts J. W., Petajan J. H., Leppert M. Paramyotonia congenita and hyperkalemic periodic paralysis map to the same sodium‐channel gene locus. Am J Hum Genet 1991; 49: 851–4
  • Ptacek L. J., Tawil R., Griggs R. C., Storvick D., Leppert M. Linkage of atypical myotonia congenita to a sodium channel locus. Neurology 1992; 42: 431–3
  • Chahine M., George A. L., Jr., Zhou M., Ji S., Sun W., Barchi R. L., et al. Sodium channel mutations in paramyotonia congenita uncouple inactivation from activation. Neuron 1994; 12: 281–94
  • Lerche H., Heine R., Pika U., George A. L., Jr., Mitrovic N., Browatzki M., et al. Human sodium channel myotonia: slowed channel inactivation due to substitutions for a glycine within the III‐IV linker. J Physiol 1993; 470: 13–22
  • Mitrovic N., George A. L., Jr., Lerche H., Wagner S., Fahlke C., Lehmann‐Horn F. Different effects on gating of three myotonia‐causing mutations in the inactivation gate of the human muscle sodium channel. J Physiol 1995; 487((Pt 1))107–14
  • Cannon S. C. Ion‐channel defects and aberrant excitability in myotonia and periodic paralysis. Trends Neurosci 1996; 19: 3–10
  • Cannon S. C., Strittmatter S. M. Functional expression of sodium channel mutations identified in families with periodic paralysis. Neuron 1993; 10: 317–26
  • Bendahhou S., Cummins T. R., Griggs R. C., Fu Y. H., Ptacek L. J. Sodium channel inactivation defects are associated with acetazolamide‐exacerbated hypokalemic periodic paralysis. Ann Neurol 2001; 50: 417–20
  • Jurkat‐Rott K., Mitrovic N., Hang C., Kouzmekine A., Iaizzo P., Herzog J., et al. Voltage‐sensor sodium channel mutations cause hypokalemic periodic paralysis type 2 by enhanced inactivation and reduced current. Proc Natl Acad Sci U S A 2000; 97: 9549–54
  • Kuzmenkin A., Muncan V., Jurkat‐Rott K., Hang C., Lerche H., Lehmann‐Horn F., et al. Enhanced inactivation and pH sensitivity of Na(+) channel mutations causing hypokalaemic periodic paralysis type II. Brain 2002; 125: 835–43
  • Priori S. G., Napolitano C., Gasparini M., Pappone C., Della B. P., Brignole M., et al. Clinical and genetic heterogeneity of right bundle branch block and ST‐segment elevation syndrome: A prospective evaluation of 52 families. Circulation 2000; 102: 2509–15
  • Lerche H., Weber Y. G., Baier H., Jurkat‐Rott K., Kraus de Camargo O., Ludolph A. C., et al. Generalized epilepsy with febrile seizures plus: further heterogeneity in a large family. Neurology 2001; 57: 1191–8
  • Buchner D. A., Trudeau M., Meisler M. H. SCNM1, a putative RNA splicing factor that modifies disease severity in mice. Science 2003; 301: 967–9
  • Kearney J. A., Yang Y., Beyer B., Bergren S. K., Claes L., Dejonghe P., et al. Severe epilepsy resulting from genetic interaction between Scn2a and Kcnq2. Hum Mol Genet 2006; 15: 1043–8
  • Bergren S. K., Chen S., Galecki A., Kearney J. A. Genetic modifiers affecting severity of epilepsy caused by mutation of sodium channel Scn2a. Mamm Genome 2005; 16: 683–90
  • Viswanathan P. C., Benson D. W., Balser J. R. A common SCN5A polymorphism modulates the biophysical effects of an SCN5A mutation. J Clin Invest 2003; 111: 341–6

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.