Skip main navigation

Phenotypical Manifestations of Mutations in the Genes Encoding Subunits of the Cardiac Sodium Channel

Originally publishedhttps://doi.org/10.1161/CIRCRESAHA.110.238469Circulation Research. 2011;108:884–897

    Abstract

    Variations in the gene encoding for the major sodium channel (Nav1.5) in the heart, SCN5A, has been shown to cause a number of arrhythmia syndromes (with or without structural changes in the myocardium), including the long-QT syndrome (type 3), Brugada syndrome, (progressive) cardiac conduction disease, sinus node dysfunction, atrial fibrillation, atrial standstill, and dilated cardiomyopathy. Of equal importance are variations in genes encoding for various subunits and regulatory proteins interacting with the α-subunit Nav1.5 and modifying its function. Based on detailed studies of genotype–phenotype relationships in these disease entities, on detailed studies of the basic electrophysiological phenotypes (heterologous expressed wild-type and mutant sodium channels and their interacting proteins), and on attempts to integrate the obtained knowledge, the past 15 years has witnessed an explosion of knowledge about these disease entities.

    Initially based on genetic studies in the mid-1990s, the SCN5A gene, encoding the major sodium channel (Nav1.5) in the heart, has been unmasked as an important player in a number of primary arrhythmia syndromes. Indeed, sodium channel dysfunction resulting from mutation-based alterations in channel function is associated with the long-QT syndrome (LQTS) (type 3),1 the syndrome of right precordial ST elevation, ie, Brugada syndrome (BrS),2 (progressive) cardiac conduction disease (CCD),2,3 sinus node dysfunction,4 atrial fibrillation (AF),5 atrial standstill,6,7 and dilated cardiomyopathy.8 Because many of these diseases associate with sudden death in young individuals with a structurally normal heart, potentially causal SCN5A variants have also been identified in sudden infant or sudden unexplained death syndrome (SIDS/SUDS).9,10 Detailed descriptions of the clinical (patients) and basic electrophysiological phenotypes (heterologous expressed wild-type and mutant sodium channels) have profoundly increased our knowledge about these disease entities and about the role of the sodium channel and its interaction with various subunits. Indeed, besides β-subunits (β1 to β4), the α-subunit Nav1.5 also interacts with several regulatory proteins, of which, more recently, malfunction resulting from disease-causing variants in their encoding genes has been shown to underlie comparable phenotypes.

    This article aims to review the existing knowledge of the basic and clinical aspects of sodium channel related diseases. It begins with a review of sodium channel function in normal and abnormal conditions that is followed by a description of the diverse clinical phenotypes.

    The Cardiac Sodium Channel

    The α-subunit of the cardiac isoform of the sodium channel is known as NaV1.5 and is encoded by the SCN5A gene. SCN5A maps to the short arm of chromosome 3 (3p21), contains 28 exons spanning ≈80 kb,11,12 and encodes the α-subunit of the voltage-gated cardiac sodium channel (designated Nav1.5), conducting the cardiac sodium current INa. Sodium channels are expressed in several tissues and share a common structure.1316 Hence, the cardiac sodium channel is a member of the voltage-dependent family of sodium channels, which consist of heteromeric assemblies of the pore-forming α-subunit. NaV1.5 contains 2016 residues and has an approximate molecular mass of 227 kDa. It consists of 4 homologous domains, known as DI to DIV, joined by so-called linkers. These 3 linkers, as well as the N terminus and the C terminus of the protein, are cytoplasmic. Each domain, DI to DIV, contains 6 transmembrane helices (namely S1 to S6) linked by intracellular or extracellular loops. S5 and S6 in each of the domains form the pore-lining helices (Figure 1).

    Figure 1.

    Figure 1. Schematic representation of the α-subunits of cardiac sodium channel, consisting of 4 serially linked homologous domains (DI–DIV), each containing 6 transmembrane segments (S1–S6) (A). In B, the interacting β-subunits and other regulatory proteins are depicted.

    Opening of the channel allows a rapid influx of positively charged Na ions (INa) that will depolarize within tenths of a millisecond the membrane potential of cardiac cells. As such, sodium channel activity plays a central role in cardiac excitability and conduction of the cardiac impulse. The S4 segments in each domain, containing a positively charged amino acid at every third position, are held responsible for voltage-dependent activation. At the end of the action potential upstroke (ie, phase 0), most channels are inactivated and do not further allow passage of ions. Furthermore, this inactivation process is voltage-dependent. It is generally accepted that inactivation is mediated mainly by the inactivation gate (DIII to DIV linker), which blocks the inside of the channel shortly after it has been activated. Furthermore, the C-terminal cytoplasmic domain plays an important role in inactivation. Channels can only be reactivated after recovery from inactivation during phase 4. However, a small percentage of channels remain available to conduct and may reopen during the plateau phase of the action potential (phases 2 and 3). This fraction of channels, less than 1% of total available Na+ channels, contributes a small “late” Na+ current (INaL) to total membrane current. Under normal conditions, INaL is very small and has little impact on action potential morphology but can be very important in pathological conditions involving sodium channel mutations, among others (see below).

    The cardiac Na+ channel is a multiprotein complex in which auxiliary proteins interact with the α-subunit (Nav1.5) to regulate its gating, cellular localization, intracellular transport, and degradation. By performing screening experiments such as coimmunoprecipitation and yeast 2-hybrid, interactions with Nav1.5 have been demonstrated for several proteins. In addition, mapping techniques have been used to discover the sites of interactions and the regulatory role of these auxiliary proteins have been studied with coexpression experiments in heterologous expression systems such as human embryonic kidney cells (HEK-293), Chinese hamster ovary cells (CHO), or Xenopus oocytes. Recently, genetic screening studies have linked mutations in genes encoding several of these subunits of the cardiac Na+ channel to phenotypes similar to those previously associated with mutations in SCN5A. Auxiliary proteins that interact with Nav1.5 may be classified as enzymes, adapter proteins, and regulatory proteins. Enzymes may directly interact with Nav1.5 and control its cellular localization by ubiquitylation and internalization (ubiquitin-protein ligases). Enzymes may also alter gating properties by phosphorylation (kinases) or dephosphorylation (phosphatase). Adapter proteins anchor Nav1.5 to the cytoskeleton and play a role in the intracellular transport and targeting of Nav1.5 to the sarcolemma. Regulatory proteins modulate gating properties of Nav1.5 on binding.16

    As indicated mutations in SCN5A lead to various phenotypes. In addition, mutations in 7 genes that encode for auxiliary proteins of the cardiac Na+ channel have been linked to arrhythmic phenotypes previously associated with these phenotypes (Table 1). Four of these genes (SCN1B, SCN2B, SCN3B, and SCN4B) encode for proteins with a uniform molecular structure (β1, β2, β3, and β4, respectively; Figure 1). These so-called β-subunits of the cardiac Na+ channel contain an extracellular amino terminus, a single transmembrane segment, and an intracellular carboxyl terminus. All 4 β-subunits are expressed in the heart, and are particularly localized at the T-tubules/Z-lines and intercalated disks in cardiac myocytes.17 The β-subunits play critical roles in the regulation of sarcolemmal expression and gating of Nav1.5, in the recruitment of cytoskeletal adapter proteins, enzymes, and signaling molecules to the channel and in the adhesion of Nav1.5 to the cytoskeletal framework and extracellular matrix.

    Table 1. Cardiac Sodium Channel Regulatory Proteins Linked to Arrhythmic Phenotypes

    Gene Locus Protein Phenotype(s) Effect of Mutation(s) on INa Reference
    SCN1B 19q13.1 β1-subunit Brugada syndrome Decrease in peak INa 48, 74
    Cardiac conduction disease
    Atrial fibrillation
    SCN2B 11q23 β2-subunit Atrial fibrillation Decrease in peak INa 48
    SCN3B 11q23.3 β3-subunit Brugada syndrome Decrease in peak INa 49, 79, 82
    Idiopathic ventricular fibrillation Increase in persistent INa
    Sudden infant death syndrome
    Atrial fibrillation
    SCN4B 11q23.3 β4-subunit Long QT syndrome type 10 Increase in persistent INa 84, 85
    Sudden infant death syndrome
    GPD1L 3p22.3 Glycerol 3 phosphate dehydrogenase 1-like (GPDL1) Brugada syndrome Decrease in peak INa 90, 93
    Sudden infant death syndrome
    CAV3 3p25 Caveolin-3 Long QT syndrome type 9 Increase in persistent INa 96, 100
    Sudden infant death syndrome
    SNTA1 20q11.2 α-1 syntrophin Long QT syndrome type 12 Increase in (persistent) INa 107, 108
    Sudden infant death syndrome

    For many of the identified putative causal variants in both SCN5A, as well as in all its subunits (see below), one should realize that almost always sound segregation data, linking the genetic variant to the clinical phenotype lack. Other evidence that the identified variants are disease-causing is based on “weaker criteria” that include absence in control alleles, conservation throughout species and functional data. Indeed, in a study designed to identify the presence of rare variants in the 3 main genes associated with LQTS, including SCN5A, in control individuals, these variants were particularly identified in the intra domain linker and the C terminus of SCN5A.18 All of these variants were altered amino acids considered well-conserved among species. This has led to the notion that the pretest probability of pathogenicity of SCN5A variants identified in these regions to cause LQTS3 (see below) is far less than 50%.19 With this caveat in mind, one should examine the existing literature and be very cautious in calling a rare genetic variant a novel arrhythmia syndrome–susceptibility mutation.20 This holds in particular for variants identified in single patients only once. Clearly, detailed functional studies demonstrating clear aberration from normal behavior increase the likelihood that certain variants are disease-causing.

    Clinical Phenotypes Related to SCN5A Mutations

    Brugada Syndrome

    BrS is a hereditary disease responsible for ventricular fibrillation and SCD in the young. The disease is characterized by the presence of right ventricular conduction abnormalities and coved-type ST-segment elevation in the anterior precordial leads (V1 to V3).21 The prevalence of BrS is estimated at approximately 3 to 5 in 10 000 people. Although the mean age of onset of events is approximately 40 years, SCD can affect individuals of any age, particularly men (75%). Of those, 20% to 50% have a family history of SCD. The disease exhibits an autosomal dominant pattern of transmission and variable penetrance. Most mutations occur in genes related to Na+ channels, although other channels may also be implicated in BrS.22,23

    Between 20% and 25% of patients affected by BrS have mutations in the SCN5A gene. To date, there are more than 200 known SCN5A BrS-related mutations described.24 These mutations lead to a loss of Na+ channel function through several mechanisms, including trafficking defects, generation of truncated proteins, faster channel inactivation, shift of voltage dependence of steady-state activation toward more depolarized membrane potentials (Figure 2), or slower recovery from inactivation. Recently, it has been shown that in several large SCN5A-related BrS families (5 of 13 families with 4 or more phenotypically affected) affected individuals did not carry the familial disease-causing mutation. These observations could be explained by phenocopies but could also explained by the presence of an anatomic substrate (an area with slow conduction in the RV outflow track) with a strong modifying role of SCN5A loss-of-function mutations. Apparently, in some individuals, the substrate is sufficient to provide the Brugada pattern spontaneously, ie, in the absence of an SCN5A mutation. In any case, these observations question a direct causal role of SCN5A mutations,25 and it stresses once more how careful one should be when investigating mutations.

    Figure 2.

    Figure 2. Two mechanisms leading to reduced peak inward current (A). B, The voltage dependence of normal activation is depicted by the yellow squares and activation at more depolarized potentials is depicted by the red squares. Shift of voltage dependence of steady-state activation toward more depolarized membrane potentials will lead to a decrease in the peak sodium current. Furthermore, a hyperpolarizing shift of the voltage dependence of steady-state inactivation curve will lead to a decrease in the peak sodium current, because, for a given resting membrane potential, fewer sodium channels are available. Both mechanisms, in addition to slower recovery from inactivation (not shown), have been described for BrS-associated SCN5A mutations that alter channel characteristics.

    Genotype–phenotype correlation studies in BrS have been limited because of the low prevalence of mutation carriers. A recently published study has proposed that the type of SCN5A mutation may be useful for risk stratification. In this study, a genotype-phenotype correlation was performed according to the type of SCN5A mutation (missense, truncated). Patients and relatives with a truncated protein had a more severe phenotype and more severe conduction disorders.26 At this point in time, such data are too preliminary to use in clinical decision making.

    Lev-Lenègre Syndrome

    Lev-Lenègre syndrome is a progressive cardiac conduction disease (characterized by either prolongation of the P wave, PR or QRS segment duration, complete AV block, and bundle branch blocks) probably associated with gradually developed ventricular fibrosis of the conduction system. The first locus for the disease was reported in 1995, on chromosome 19q13.2-13.3.27 In 1999, the first mutations causing the disease were described in SCN5A.3 Mutations in SCN5A lead to a reduction in Na+ current, reducing the velocity of the impulse conduction.28

    Long QT Syndrome

    LQTS is a cardiac channelopathy characterized by a QT interval prolongation, responsible for ventricular tachyarrhythmias with associated episodes of syncope and SCD.29,30 LQTS is one of the leading causes of SCD among young people. It can be congenital or acquired, generally in association with drugs and electrolyte imbalance (hypokalemia, hypocalcemia and hypomagnesemia). The clinical presentation can be variable, ranging from asymptomatic patients to episodes of syncope and SCD caused by ventricular tachyarrhythmias (torsades de pointes) in a structurally normal heart. Prolongation of the QT interval may arise because of a decrease in the K+ repolarization currents, an increase in Ca2+ entry or to a sustained entry of Na+ into the myocyte (ie, increase in INaL). An increase in INaL is caused by either impaired inactivation, ie, failure to inactivate completely (Figure 3A) or an increase in “window current.” The latter is generally attributable to a shift of voltage dependence of steady-state inactivation toward more depolarized membrane potentials leaving an increased “window” of inward sodium current (Figure 3B).

    Figure 3.

    Figure 3. The 2 mechanisms through which sodium channel mutations lead to an increase in INaL.A, Either impaired inactivation, ie, failure to inactivate completely (A) or an increase in “window current” caused by a depolarizing shift in inactivation (“increased availability”), leaving an increased “window” inward sodium current (B) causes a late sodium inward current. The latter current flows through already activated but not yet inactivated sodium channels (B).

    Until now, 13 different types of LQTS have been described, most of them associated to K+ channel disorders.31 However, a 10% to 15% of the LQTS cases are related to mutations in SCN5A (type 3 LQTS) or in Na current–associated proteins. Patients with type 3 LQTS present arrhythmias associated with bradycardia and symptoms at rest, especially during sleep.32 Whether β-blockers are as important in type 3 LQT patients as they are in LQT type 1 and 2 patients has been questioned. Data to support nonresponsiveness lack but, preliminary data on a large cohort of LQTS type 3 patients suggest a beneficial effect (A Wilde et al, unpublished data). The best therapy for type 3 LQT patients is currently unknown, but high risk patients are probably those who are symptomatic and those with the longest QT intervals (>500ms). In the latter category a prophylactic ICD might be considered. Gene-specific therapies include sodium channel blockers like flecainide and mexiletine3335 although detrimental attempts, with pathophysiological explanations based on sound experimental studies, have been published.36 Furthermore, ranolazine, a new drug with late sodium current blocking efficacy, has been shown to be able to shorten the QTc interval significantly.37 As many different mutations lead to an overlap syndrome, ie, families with both gain-of-function and loss-of-function phenotype38 the use of flecainide should be accompanied with serial ECGs because a BrS type ECG (with the proarrhythmogenic substrate) may be the result.33 For all of these drugs, long-term follow-up is not available.

    Dilated Cardiomyopathy

    Dilated cardiomyopathy is a cardiac structural disease characterized by decreased systolic function and ventricular dilatation. An average 20% of cases are thought to be familial, most associated with mutations in the cytoskeletal proteins. The identification of a mutation in SCN5A in a family with dilated cardiomyopathy provided the first evidence of the disease being caused by an alteration in a cardiac ion channel.8 Since then, other publications have confirmed this association.39 The phenotype is usually associated with arrhythmias (atrial and ventricular arrhythmias and conduction disease).

    Atrial Fibrillation

    AF is defined as an unpredictable activation of the atria, characterized by irregular fibrillatory ECG waves causing an irregular ventricular response, which manifests clinically as an irregular pulse. AF is one of the most common and yet challenging arrhythmias encountered in clinical practice. The prevalence of 1% among the general population increases to 10% among individuals older than 80 years, and it is responsible for more than one-third of all cardioembolic episodes.40

    Although familial forms had remained mostly unknown, the identification in 1997 of a genetic locus causing familial AF, which defined AF in a subset of patients as a genetic disease with an autosomal dominant pattern of inheritance, initiated further research into the understanding the pathophysiology of this inherited form of the arrhythmia.41 To date, most of the genes associated to AF encode K+ channel proteins (KCNQ1, KCNE2, KCNE3, KCNA5, KCNJ2, and KCNH2).42 The role of Na+ channel proteins in AF has become evident in these last few years. In 2008, several genetic variants in SCN5A, causing both loss and gain of INa, have been associated with lone AF.5,43 Most of these phenotypes do not show alterations in the QT interval or in the ST segment elevation.44,45 A family with AF with a combined phenotype of LQTS type 3 has been recently described.46 In addition, mutations in the β-subunits SCN1B, SCN2B, and SCN3B have been associated with AF by decreasing INa.4749 It is hypothesized that loss of INa function leads to altered atrial conduction parameters, increased fibrosis, and associated structural abnormalities (including increased diastolic ventricular pressure in the setting of dilated cardiomyopathy with subsequent enlargement of the atria), that all act together to cause AF.

    Sick Sinus Syndrome

    Diseases of the sinoatrial node (brady-tachy syndrome, sinus bradycardia, sinus arrest) have been associated with loss-of-function mutations in SCN5A, especially severe in patients with compound heterozygote mutations.4 Similar with other loss-of-function phenotypes, families may show important overlapping features, with family members displaying sick sinus syndrome, whereas others may display BrS or progressive cardiac conduction disease.45,50,51

    Atrial standstill has also been described in SCN5A families. A combined loss-of-function mutation with a loss-of-function connexin 43 polymorphism also leads to atrial standstill.52

    Genetic Modulators in Na Current and Sudden Cardiac Death

    Modulation of the phenotype may be related to the presence of environmental factors (ie, inducers like fever in BrS or medications in LQTS) or additional genetic factors. Among these genetic factors, it is well accepted that the presence of compound heterozygotes will confer a worse prognosis in family members. As such, compound heterozygotes involving genes associated with Na current have been described in LQTS,53 BrS,54 and sick sinus syndrome.4

    Polymorphisms have recently acquired more importance in the explanation of certain phenotypes of genetic diseases. Genetic variants in the SCN5A promoter region may have a pathophysiologic role in BrS. A haplotype of 6 polymorphisms in the SCN5A promoter has been identified and functionally linked to a reduced expression of the sodium current. This variant was found among patients of Asian origin, and it could play a role in modulating the expression of BrS in far eastern countries.55 In addition, the common H558R polymorphism has been shown to partially restore the sodium current impaired by other simultaneous mutations causing either cardiac conduction disturbances (T512I)56 or BrS (R282H).57 The common variant H558R seems to be a genetic modulator of BrS among carriers of an SCN5A mutation, in whom the presence of the less common allele makes BrS less severe (shorter QRS complex duration in lead II, lower J-point elevation in lead V2, and a trend toward fewer symptoms).58

    Finally, the presence of the genetic variant S1103Y, which accelerates sodium channel activation, has been associated with arrhythmias and sudden cardiac death in blacks.59,60 Of interest in experimental settings, the late sodium current was exacerbated by lowering intracellular pH, making it an important mechanism predisposing a Y1103 carrier to sudden infant death syndrome (SIDS). Indeed, this variant has been identified in black SIDS cases.61

    Proteins Associated With Arrhythmias

    Mutations in all 4 β-subunit–encoding genes are found in individuals with various arrhythmic phenotypes, reflecting their importance in normal cardiac electric activity. Interestingly, most such mutations are located in the extracellular amino terminus of the β-subunits (Table 2 and Figure 4), suggesting this domain may play an important role in the interaction and regulation of Nav1.5 by the β-subunits.

    Table 2. Mutations in Cardiac Sodium Channel β-Subunits Linked to Arrhythmic Phenotypes

    β-Subunit Nucleotide Change Variant Mutation Type Location Phenotype(s) Effect of INa Reference
    β1 536G>A W179X Nonsense C terminus Brugada syndrome and cardiac conduction disease Decrease in peak INa and positive shift of activation 74
    β1 537G>A W179X Nonsense C terminus Cardiac conduction disease Decrease in peak INa 74
    β1 259G>C E87Q Missense N terminus Cardiac conduction disease Decrease in peak INa and positive shift of activation 74
    β1 254G>A R85H Missense N terminus Atrial fibrillation Decrease in peak INa and positive shift of activation and inactivation 48
    β1 457G>A D153N Missense N terminus Atrial fibrillation Decrease in peak INa 48
    β2 82C>T R28W Missense N terminus Atrial fibrillation Decrease in peak INa and positive shift of activation 48
    β2 83G>A R28Q Missense N terminus Atrial fibrillation Decrease in peak INa and positive shift of activation and inactivation 48
    β3 17G>A R6K Missense N terminus Lone atrial fibrillation Decrease in peak INa 47
    β3 29T>C L10P Missense N terminus Brugada syndrome Decrease in peak INa and negative shift of inactivation 47, 82
    Lone atrial Fibrillation
    β3 161T>G V54G Missense N terminus Idiopathic ventricular fibrillation Decrease in peak INa and positive shift of activation and inactivation 79, 84
    Sudden infant death syndrome
    β3 106G>A V36M Missense N terminus Sudden infant death syndrome Decrease in peak INa and increase in persistent INa 84
    β3 389C>T A130V Missense N terminus Atrial fibrillation Decrease in peak INa 49
    β3 482T>C L161T Missense Transmembrane segment Lone atrial Fibrillation Decrease in peak INa 47
    β4 535C>T L179F Missense Transmembrane segment Long QT syndrome type 10 Increase in persistent INa and positive shift of inactivation 85
    β4 671G>A S206L Missense C terminus Sudden infant death syndrome Increase in persistent INa and positive shift of inactivation 84
    Figure 4.

    Figure 4. Schematic representation of the β-subunits of cardiac sodium channel (β1 [A], β2 [B], β3 [C], and β4 [D]), each containing 1 transmembrane segment. For each β-subunit, identified variants that have been linked to either syndrome (see color code below the schematic illustrations) are depicted.

    β1-Subunit

    Two splice variant transcripts of SCN1B, generating β1 and β1 B proteins, are expressed in the heart.62 The β1-subunit coimmunoprecipitates with Nav1.563 and is found to colocalize with Nav1.5 already within the endoplasmic reticulum, suggesting that they mediate the transport of the channel complex to the sarcolemma.64 The β1-subunit is believed to interact with Nav1.5 at multiple sites. An interaction between the extracellular domain of the β1-subunit with the extracellular loops between segments 5 and 6 of domains III and IV (S5-S6 loops of DIII and DIV) of Nav1.5 is required for the regulation of channel gating.65,66 In addition, an interaction may also exist between the carboxyl terminus of these 2 subunits.67

    Accordingly, the p.D1790G mutation in the carboxyl terminus of Nav1.5, discovered in a large family with LQT-3 has been shown to affect the inactivation of the cardiac Na+ channel by disrupting the interaction between Nav1.5 and β1-subunit.68 In heterologous expression systems, coexpression of SCN5A with SCN1B have demonstrated variable effects of β1-subunit on gating, depending on the expression system used. However, in most cases, an increase in peak INa amplitude,69,70 a shift of steady-state inactivation toward more positive potentials,63,71 and a decrease in persistent INa amplitude has been observed.71,72 Surprisingly, ventricular myocytes from SCN1B knockout mice showed increased expression of Nav1.5 on mRNA and protein level, an increase in peak INa amplitude, and no change in Nav1.5 gating.73 These findings are not in agreement with those observed in in vitro experiments and suggest that heterologous expression systems may not mimic the intracellular environment (in particular the protein–protein interactions) present in cardiac myocytes. However, the SCN1B knockout mice also displayed lower heart rates and repolarization delay (reflected by prolonged QTc and action potential durations), and this may result from an increase in persistent INa attributable to the absence of the suppressive effects of β1-subunit on this current. Accordingly, individuals with the previously mentioned D1790G mutation in Nav1.5, which disrupts the interaction of Nav1.5 with the β1-subunit, also experienced sinus bradycardia and QTc prolongation.68

    Mutations in SCN1B have been found in individuals with BrS, cardiac conduction disease (CCD), or AF. In a recent study by Bezzina and colleagues, 282 probands with BrS and 44 with CCD were screened for mutations in SCN1B.74 A nonsense mutation (c.536G>A resulting in p.W179X) was identified in 4 related individuals, of whom 3 displayed signs of BrS and CCD on their ECGs. This mutation is predicted to produce a prematurely truncated protein lacking the membrane-spanning segment and the carboxyl terminus. A different nonsense mutation (c.537G>A resulting in p.W179X) was identified in a father and his daughter, of whom only the daughter displayed signs of CCD (right bundle branch block and prolonged PR interval). In both individuals, BrS was unlikely, as established by a negative sodium channel drug challenge test. Finally, a missense mutation (c.259G>C resulting in p.E87Q), located in the amino terminus of the β1-subunit, was found in 3 family members, of whom 2 displayed signs of CCD on their ECGs (complete left bundle branch block, right bundle branch block, and left anterior hemiblock). It must be noted that the families in this study were too small to perform linkage analysis between the phenotypes and the mutations, and that the mutations were also found in individuals with no phenotype. In another study, Roden and colleagues screened 842 AF patients for mutations in SCN1B to SCN4B and identified a mutation in SCN1B in 2 unrelated individuals.48 The mutations (c.254G>A resulting in p.R85H, and c.457G>A resulting in p.D153N) were both located in the amino terminus of the β1-subunit. Coexpression of all of these mutant β1-subunits with Nav1.5 in CHO cells resulted in a decrease of peak INa density compared with coexpression of Nav1.5 with normal β1-subunits, suggesting that β1-subunits increase the sarcolemmal expression of Nav1.5 in cardiac myocytes.48,74 In addition, p.W179X, p.E87Q, and p.R85H also caused a shift of steady-state activation toward more positive potentials, indicating delayed activation of the channels. These findings are in line with the INa reduction found for virtually all SCN5A mutations related to BrS and CCD, and for few SCN5A mutations related to AF.13INa reduction is speculated to predispose to AF by slowing the conduction velocity of electric stimuli through the atria, thereby facilitating the maintenance of reentrant excitation waves.75

    β2-Subunit

    The β2-subunit, encoded by SCN2B, coimmunoprecipitates with Nav1.5 and colocalizes at the intercalated disks in cardiac myocytes.63,76 Coexpression of Nav1.5 with the β2-subunit resulted in a shift of the steady-state activation toward more positive potentials compared with Nav1.5 expression alone.48 A similar effect was observed when Nav1.5 was coexpressed with the intracellular domain of the β2-subunit, suggesting that this portion of the protein may interact with Nav1.5 to modulate the gating.77 A recent study revealed that the effect of the β2-subunit on Nav1.5 activation depends on the presence of sialic acid residues on the β2-subunit (ie, the extent of β2-sialylation). When completely sialylated, β2-subunit caused a reverse shift of the steady-state of activation, ie, toward more negative potentials.78

    In the above-mentioned study by Roden and colleagues, mutations in SCN2B were identified in 2 unrelated individuals with AF.48 The mutations (c.82C>T resulting in p.R28W, and c.83G>A resulting in p.R28Q) were located in the amino terminus of the β2-subunit, and their coexpression with Nav1.5 resulted in a decrease of peak INa density and a shift of steady-state activation toward more positive potentials compared with when Nav1.5 was coexpressed with normal β2-subunits.48 Both of these effects led to INa reduction, and, accordingly, the individuals carrying the mutations were reported to display electrocardiographic BrS-like changes. Although β2-subunits were previously shown not to colocalize with Nav1.5 in the endoplasmic reticulum, indicating they are transported separately to the sarcolemma,63 the decrease in peak INa density by the mutant β2-subunits suggests that β2-subunits play a role in controlling the sarcolemma expression of Nav1.5.

    β3-Subunit

    The β3-subunit, encoded by SCN3B, coimmunoprecipitates with Nav1.5.79 An initial study showed that the β3-subunit is highly expressed in the ventricles and Purkinje fibers but not in atrial tissue of sheep hearts.80 However, recently, the β3-subunit expression was demonstrated in mouse atrial and ventricular tissues.81 Compared with Nav1.5 alone, coexpression with the β3-subunit resulted in increase in peak INa amplitude and a shift of steady-state inactivation toward more positive potentials (in Xenopus oocytes),80 or no change in peak INa amplitude and a shift of steady-state inactivation toward more negative potentials (in CHO cells and in HEK-293 cells).72,79 These findings illustrate again that the regulatory effects of β-subunits on Nav1.5 may vary significantly between heterologous expression systems, probably because of differences in intracellular environments of various expression systems and conditions in which the cells are cultured (eg, lower culture temperature for Xenopus oocytes). Interestingly, coexpression of Nav1.5 with β1- and β3-subunits caused an additional shift of inactivation toward more negative potentials, suggesting that β-subunits may act synergistically to regulate gating.72 Compared with wild-type mice, SCN3B knockout mice displayed slower heart rates, longer P wave durations, and prolonged PR intervals on their ECGs, indicating slowing of electric conduction in the heart.81 In addition, electrophysiological studies demonstrated inducibility of atrial tachycardia, AF, and ventricular tachycardia in hearts of SCN3B knockout mice, indicating increased atrial and ventricular arrhythmogenesis.81 Ventricular myocytes isolated from SCN3B knockout mice hearts showed a decrease in peak INa density and a shift of steady-state inactivation toward more negative potentials compared with ventricular myocytes isolated from wild-type hearts.81 These data are in agreement with those obtained from coexpression studies of β3-subunit with Nav1.5 in Xenopus oocytes.80

    To date, 4 mutations in SCN3B have been linked to arrhythmic phenotypes in humans. Using a candidate gene approach, Antzelevitch and colleagues identified a missense mutation in SCN3B (c.29T>C resulting in p.L10P, located in the amino terminus of β3-subunit) in a patient displaying the ECG signs of BrS.82 By a similar approach, Makielski and colleagues identified a missense mutation (c.161T>G resulting in p.V54G, also located in the amino terminus of β3-subunit) in a patient with aborted cardiac arrest, attributable to documented ventricular fibrillation, in the absence of structural heart disease or any well-defined arrhythmic phenotype.79 Coexpression studies revealed that both these mutations cause a decrease in peak INa density and a shift in steady-state inactivation compared with coexpression of Nav1.5 with normal β3-subunits. Additionally, immunocytochemistry revealed that the mutant β3-subunits decreased the expression of Nav1.5 by disrupting its transport from the endoplasmic reticulum to the sarcolemma.79,82 It should be noted that in both cases, there is not much evidence for loss-of-function of cardiac sodium channel activity as even in the presence of procainamide conduction intervals are within normal limits, whereas with SCN5A loss-of-function mutations, conduction is generally prolonged at different levels.26,83 Recently, the p.V54G mutation was also linked to SIDS.84 Makielski and colleagues screened 292 SIDS cases for mutations in β-subunit–encoding genes and found a mutation in SCN3B in 2 cases (V54G, and c.106G>A resulting in p.V36M, located in the amino terminus). Coexpression of Nav1.5 with p.V36M β3-subunit resulted in a decrease in peak INa density but, interestingly, also in an increase of the persistent INa, compared with coexpression with normal β3-subunits, showing that mutations in β3-subunits may also cause INa gain of function.84 Taken together, findings from these translational studies are largely in line with those obtained from in vitro experiments and SCN3B knockout mice and indicate that β3-subunit plays an important regulatory role in the sarcolemmal expression and gating of Nav1.5. However, recently, Wang and colleagues sequenced 477 patients with AF for mutations in SCN3B and identified a missense mutation in SCN3B (c.389C>T resulting in p.A130V, located in the distal portion of the amino terminus) in one patient.49 When coexpressed with Nav1.5, the mutant β3-subunit resulted in a decrease in peak INa density and no gating changes. Surprisingly, Western blot analysis revealed that the mutation did not affect the surface expression of Nav1.5. Based on these data, the authors speculated that p.A30V may decrease INa by impairing the conductance, but not the intracellular trafficking, of Nav1.5. Although this is an interesting hypothesis, it requires further experimental investigation. Recently, 3 additional loss-of-function mutations in SCN3B have been associated with lone AF (R6K, L10P, M161T).47 It must be noted that all SCN3B mutations linked to arrhythmic phenotypes are located in the amino terminus of the β3-subunit, suggesting this portion of the protein is important for its regulatory role in the expression and gating of Nav1.5.

    β4-Subunit

    The β4-subunit is encoded by SCN4B and coimmunoprecipitates with Nav1.5,85 and its expression in the heart has been demonstrated on both mRNA and protein levels.86,87 Coexpression of Nav1.5 with the β4-subunit did not change peak INa amplitude but shifted the steady-state inactivation toward more negative potentials compared with expression of Nav1.5 alone.84,85 Recently, the β4-subunit has been proposed as a potential genetic modifier of conduction slowing in individuals with cardiac Na+ channel loss of function.86 This study aimed to assess the role of genetic modifiers on the variable severity of conduction slowing in patients with the SCN5A p.1795insD mutation. To investigate the effect of genetic background without the confounding influences of environmental factors, 2 different mouse strains (FVB/N and 129P2), both carrying the mouse homolog of the p.1795insD mutation, were studied. Similar to the human phenotype, both transgenic mice strains displayed bradycardia and conduction slowing. However, 129P2 mice displayed more severe conduction slowing. Pan-genomic mRNA expression profiling and subsequent Western blot analysis revealed a decrease in SCN4B mRNA and β4-subunit protein levels, respectively, in 129P2 mice compared with FVB/N mice. In addition, measurements of INa in ventricular myocytes did not find differences in peak INa but uncovered a shift of steady-state activation toward more positive potentials in 129P2 mice compared with FVB/N, indicating delayed opening of the Na+ channels. Finally, computer simulation models predicted that this delay in activation may be responsible for the differences in conduction slowing between the mouse strains.86 This report first recognizes a β-subunit of the cardiac Na+ channel as a genetic modifier of phenotype severity in individuals with a mutation in SCN5A.

    The importance of the β4-subunit is further reflected by recent studies linking mutations in SCN4B to LQTS (LQT10) and SIDS. Ackerman and colleagues identified a mutation (c.535C>T resulting in p.L179F, located in the transmembrane segment of the β4-subunit) in a LQTS family and a history of unexpected and unexplained sudden deaths.85 Coexpression of Nav1.5 with mutant β4-subunits resulted in an increase of persistent INa and a shift of steady-state inactivation toward more positive potentials compared with coexpression of Nav1.5 with normal β4-subunits. In the above-mentioned study on 272 SIDS cases, a mutation in SCN4B (c.671G>A resulting in p.S206L, located in the carboxyl terminus) was also identified.84 Coexpression experiments in HEK-293 cells showed that, similar to p.L179F, the p.S206L mutation also resulted in an increase in persistent INa and a shift of steady-state inactivation toward more positive potentials. Accordingly, adult rat ventricular myocytes, adenovirally transduced with the p.S206L mutant β4-subunits, displayed an increase in persistent INa and prolongation of action potential duration compared with myocytes transduced with wild-type β4-subunits.84 Taken together, effects of these 2 SCN4B mutations on the persistent INa and Nav1.5 gating are in line with effects previously described for SCN5A mutations related to LQTS.88

    Glycerol 3 Phosphate Dehydrogenase 1-Like Protein

    Approximately 20% of BrS cases have been linked to mutation in SCN5A (see below), and only in a few cases have mutations in β-subunit–encoding genes been found. In search for other genetic causes, London and colleagues linked a multigenerational BrS family to a locus at chromosome 3p22-24 and excluded a disease-causing mutation in SCN5A by linkage and direct sequencing.89 In a further study, the authors identified a missense mutation (c.899C>T resulting in p.A280V) in GPD1-L, located within the previously found locus, encoding the glycerol-3-phosphate dehydrogenase 1-like protein (GPD1-L).90 This protein displays 84% homology with the glycerol-3-phosphate dehydrogenase 1 (GPD1), a cytoplasmic enzyme that catalyzes the dehydrogenation of glycerol-3-phosphate (G3P) to dihydroxyacetone phosphate (DHAP) in the glycolytic pathway and the reduction of DHAP to G3P in the synthesis of triglycerides.91 GPD1-L was shown to be expressed abundantly in the heart. Compared with coexpression of Nav1.5 with wild-type GPD1-L, its coexpression with p.A280V GPD1-L resulted in a marked decrease of peak INa density and reduction of cell surface expression of Nav1.5.90 In another study, Ackerman and colleagues screened a large cohort of SIDS cases for mutations in GPD1-L.92 They found 3 novel missense mutations in GPD1-L (c.307G>A resulting in p.E83K, c.430A>G resulting in p.I124V, and c.877C>T resulting in R273C). Similar to the BrS-linked p.A280V mutation, coexpression of Nav1.5 with each of these mutations cells caused a decrease in peak INa density. Accordingly, adenovirus-mediated transduction of neonatal mouse myocytes with p.E83K GPD1-L reduced INa compared with when myocytes were transduced with wild-type GPD1-L.93 Based on these data, GPD1-L was suggested to play a role in the intracellular trafficking of Nav1.5 to the sarcolemma.9093

    More recently, Makielski and colleagues demonstrated in HEK-293 cells that, like GPD1, GPD1-L catalyzes the reaction of G3P to DHAP, and that the BrS-linked p.A280V mutation and the SIDS-linked p.E83K mutation reduce the enzymatic activity of GPD1-L.92 Next, they showed that reduced GPD1-L activity results in an increased phosphorylation of Nav1.5 at residue S1503, and this phosphorylation is previously recognized to decrease INa.94 The authors hypothesized that GPD1L loss of function increases the substrate G3P and feeds the protein kinase (PK)C-mediated phosphorylation of Nav1.5. Indeed, additional experiments, whereby PKC activity was stimulated by G3P or inhibited by staurosporine, indicated that Nav1.5 phosphorylation and subsequent INa decrease occur through a G3P-dependent PKC-mediated pathway.92 Although this is an interesting model, future studies confirming these findings in a more native environment are required. Finally, it is worth mentioning that, by using a GST pull-down assay, Makielski and colleagues also showed a previously unrecognized interaction between Nav1.5 and GPD1-L, which was not disrupted by mutations in GPD1-L. However, they did not determine the location of this interaction or whether it was direct or mediated by channel interacting proteins.

    Caveolin-3

    Caveolae are small narrow-necked invaginations of the plasma membrane of various types that are enriched with lipids (cholesterol and sphingolipids) and proteins (ion channels, receptors, transporters, and signaling molecules). Caveolins are responsible for the formation of caveolae and the maintenance of their structure. Caveolin-3 is the main isoform expressed in the heart. Coimmunoprecipitation of Nav1.5 with caveolin-3 has been shown in both rat cardiac myocytes and HEK-293 cells.95,96 Immunocytochemistry showed that Nav1.5 colocalizes with caveolin-3,95 particularly in the caveolae at the sarcolemma but not intracellularly.97 Interestingly, Shibata and colleagues have shown that β-adrenergic stimulation of rat ventricular myocytes with isoproterenol, in the presence of a PKA inhibitor, causes an increase in peak INa. Importantly, this effect could be mimicked by the application of the stimulatory G-protein α-subunit, Gαs, and abolished by the addition of antibodies against Gαs or caveolin-3.95,98 These findings suggest that Gαs and caveolin-3 act in conjunction to increase peak INa through a cAMP-independent pathway. The authors speculated that a reservoir of Nav1.5 may be located in caveolar invaginations, where the proteins are not exposed to the extracellular side of the myocyte. By β-adrenergic stimulation and through a Gαs- and caveolin-3–dependent pathway, the caveolar neck will open and allow Nav1.5 to become functional. In this line of research, it has also been demonstrated that a histidine residue at position 41 of Gαs is critical in the regulation of cAMP-independent increase in INa.99 The exact molecular mechanisms by which Gαs or caveolin-3 directly or indirectly (eg, through adaptor proteins such as dystrophin and ankyrin that are also present in the caveolae) interact with Nav1.5 and cause an increase in functional INa remain to be unraveled.

    Based on the preceding evidence, Vatta et al hypothesized that mutations in caveolin-3 may underlie LQTS.96 They screened 905 unrelated patients with LQTS for mutations in CAV3, the gene encoding caveolin-3, and identified 4 novel missense mutations (c.233C>T resulting in p.T78M in 3 patients, c.253G>A resulting in p.A85T in 1 patient, c.290T>G resulting in p.F97C in 1 patient, and c.423C>G resulting in p.S141R in 1 patient, LQT9). Three mutations are located in the intramembrane domain, and the p.S141R mutation is located in the cytoplasmic carboxyl terminus of caveolin-3. Compared with the coexpression of Nav1.5 with wild-type caveolin-3, its coexpression with mutant p.F97C or p.S141R caveolin-3 resulted in an increase in persistent INa with no changes in peak INa density and gating.96 This is in agreement with the effect of SCN5A mutations linked to LQTS type 3 on Nav1.5.13 Importantly, Vatta et al also displayed that both p.F97C and p.S141R caveolin-3 coimmunoprecipitate with Nav1.5, indicating that the increase in persistent INa is not attributable to loss of interaction between mutant caveolin-3 and Nav1.5. More recently, Ackerman and colleagues screened 134 SIDS cases for mutation in CAV3 and identified 3 distinct mutations in 3 cases (p.T78M previously linked to LQT9, c.40G>C resulting in p.V14L, located in the amino terminus, and c.236T>G resulting in p.L79R, located in the intramembrane domain).100 Similar to the LQTS-linked mutations in caveolin-3, coexpression of Nav1.5 with each of these 3 mutations resulted in a marked increase in persistent INa compared with coexpression with wild-type caevolin-3.

    Finally, it is noteworthy that the LQTS patient with the p.F97C mutation in caveolin-3 showed QT prolongation only but reproducibly during β-agonist inhaler therapy for asthma.96 This phenomenon is in line with the previous experimental data demonstrating that caveolin-3 plays a role in the functional increase in INa on β-adrenergic stimulation95,98,101 and suggests that the mutation-induced persistent INa may also increase during such conditions.

    α1-Syntrophin

    The SNTA1-encoded α1-syntrophin (SNTA1) is a member of the dystrophin-associated multiprotein complex that interacts directly with the PDZ domain–binding motif formed by the last 3 residues (serine-isoleucine-valine) of the Nav1.5 carboxyl terminus.102 SNTA1 contains 4 domains, including 2 pleckstrin homology domains (PH1 and PH2), a PDZ domain within the PH1 domain that interacts with Nav1.5, and a syntrophin unique (SU) domain. Previously, SNTA1 had been shown to interact also with the neuronal nitric oxide synthase (nNOS), which is constitutively expressed the heart, and the cardiac isoform of plasma membrane Ca2+/calmodulin-dependent ATPase (PMCA4b).103,104 Of note, in the presence of SNTA1, PMCA4b acts as a potent inhibitor of NO production by nNOS. Importantly, NO has been demonstrated to increase the amplitude of persistent INa in ventricular myocytes.105 Based on these findings, Makielski and colleagues hypothesized that mutations in SNTA1 may disrupt its interaction with PMCA4b, thereby nullifying the ability of PMCA4b to inhibit nNOS, resulting in increased NO synthesis and increased persistent INa.106 They screened 50 unrelated individuals with LQTS for mutations in SNTA1 and identified one missense mutation (p.A390V) in one single individual (LQT12). GST-pull-down assay revealed an interaction between Nav1.5 carboxyl terminus, nNOS, PMCA4b, and wild-type SNTA1. The p.A390V mutation selectively disrupted the interaction of PMCA4b with Nav1.5 and nNOS, and this was associated with increased nitrosylation of Nav1.5 as detected by S-nitrosylation biotin switch assay. Moreover, coexpression of Nav1.5, nNOS, PMCA4b, and p.A390V-SNTA1 in HEK293 cells increased the peak and persistent INa and shifted the steady-state inactivation toward more positive potentials resulting in larger inward window current compared with coexpression with wild-type SNTA1. These gain-of-function effects were partially abolished when cells were cultured in the presence of NG-monomethyl-l-arginine (L-NMMA), an NOS inhibitor. Moreover, adenoviral transduction of neonatal mouse myocytes with p.A390V-SNTA1 also increased the persistent INa compared with wild-type SNTA1.106 Taken together, these data indicate that, by acting as a scaffolding protein for Nav1.5, nNOS, and PMCA4b, SNTA1 plays a key role in regulating the level of Nav1.5 nitrosylation, and thereby the amplitude of peak and persistent INa.

    According to this study, Vatta and colleagues identified a missense mutation (p.A257G) in 3 of 39 individuals with LQTS.107 Compared with wild-type SNTA1, p.A257G-SNTA1 caused an increase in peak INa and a shift of steady-state activation toward more negative potentials resulting in larger inward window current in HEK-293 cells and neonatal rat myocytes. Although the p.A257G mutation did not alter the persistent INa, its gain-of-function effects are in line with those observed for LQTS type 3-related mutations in SCN5A.88 Finally, mutations in SNTA1 have also been linked to SIDS.108 Ackerman and colleagues screened 292 SIDS cases, and identified 6 distinct SNTA1 mutations (p.G54R, p.P56S, p.T262P, p.S287R, p.T372M, and p.G460S) in 8 cases. Interestingly, when coexpressed with Nav1.5 in HEK-293 cells, the p.G54R and p.P56S mutations did not affect INa compared with wild-type SNTA1. The p.S287R mutation increased only the peak INa, whereas p.S287R, p.T372M, and p.G460S caused similar gain-of-function effects as the LQTS-linked p.A390V mutation (ie, increase in peak and persistent INa and a shift of steady-state inactivation toward more positive potentials). Because the mutations causing an increase in both peak and persistent INa (p.A390V, p.S287R, p.T372M, and p.G460S) were all located in or near to the PH2 and SU domains of SNTA1, the authors speculated that these mutations disrupt the interaction of SNTA1 with PMCA4b, thereby abolishing its inhibitory effect on nNOS. In contrast, mutations located in the PH1 domain of SNTA1 (the LQTS-linked p.A257G and SIDS-linked p.G54R, p.P56S, p.T262P) may not possess the ability to disrupt this interaction.

    Proteins Not Associated With Arrhythmias But With Specific Interactions With SCN5A

    A number of other enzymes and regulatory proteins have been shown to interact with the Nav1.5, but, as yet, no disease-related variants in any of these proteins have been identified (Table III and references in the Online Data Supplement, available at http://circres.ahajournals.org). However, for a number of individual mutations in SCN5A, it has been shown that the interaction between Nav1.5 and the relevant protein is modified, giving rise to a distinct phenotype. Notable examples are the A1924T mutation in the IQ motif of Nav1.5 that abolishes the enhanced entry of the channels into slow inactivation induced by calmodulin, resulting in reduced cardiac excitability109; the BrS-linked E1053K in Nav1.5 mutation that abolishes ankyrin G binding to Nav1.5 and prevents normal channel trafficking to the intercalated discs110,111; and the above-mentioned LQTS type 3–linked D1790G that abolishes binding of fibroblast growth factor homologous factor 1B, thereby abolishing the negative shift of inactivation.112

    Conclusion

    A short 2 decades of research have raised enormous insight into the role of the cardiac sodium channels and their regulatory subunits in arrhythmogenesis and have led to the description of new disease entities. For some patients with monogenic diseases, tailored therapy has ensued; for others, nothing has changed or therapeutic choices have become even more complicated. Basic insight into the role of all of these proteins assembled in the multiprotein cardiac sodium channel complex has been greatly advanced. Yet, many details are still lacking, among which, for example, which β subunits associate with which α subunits. Is this only Nav1.5? Furthermore, the spatial, temporal, and developmental distribution of these interaction are only partly known. In addition, it is still incompletely understood why, eg, loss-of-function mutations in SCN5A lead to so many different phenotypes. Gender, age, differential expression throughout the heart, coexpression, eventual tissue specific, and expression of other genetic variants (eg, the connexin variant discussed above) all may play a role, and the interaction of SCN5A is complex but is being increasingly unraveled. In the end, this may lead to a better prospect for all patients with these rare diseases, as well as for those with far more common disease entities, among which are heart failure– or ischemia-related atrial and ventricular arrhythmias, all with a mutifactorial background.

    Non-standard Abbreviations and Acronyms

    AF

    atrial fibrillation

    CCD

    cardiac conduction disease

    LQTS

    long QT syndrome

    Nav1.5

    voltage-gated cardiac sodium channel

    nNOS

    neuronal nitric oxide synthase

    PK

    protein kinase

    SCD

    sudden cardiac death

    SCN5A

    α-subunit of the cardiac sodium channel

    SCN1B

    β1-subunit of the cardiac sodium channel

    SCN2B

    β2-subunit of the cardiac sodium channel

    SCN3B

    β3-subunit of the cardiac sodium channel

    SCN4B

    β4-subunit of the cardiac sodium channel

    SIDS

    sudden infant death syndrome

    SUDS

    sudden unexplained death syndrome

    Acknowledgments

    We thank Dr Ahmad Amin for help with the manuscript and preparation of the figures.

    Sources of Funding

    Support for A.A.M.W. is supported by the Leducq Foundation (CVD05, “Alliance against Sudden cardiac Death”). R.B. is supported by Fondos Investigacion Sanitarias (FIS) , Centro Nacional Investigaciones Cardiologicas (CNIC) , Leducq Foundation , Obra Social La Caixa .

    Disclosures

    A.A.M.W. is a member of the scientific advisory committee of PGxHealth. R.B. is a paid consultant for Gendia SL.

    Footnotes

    In December 2010, the average time from submission to first decision for all original research papers submitted to Circulation Research was 14.5 days.

    Correspondence to Arthur A.M. Wilde,
    Department of Clinical and Experimental Cardiology, Academic Medical Center, PO Box 22700, University Medical Center, University of Amsterdam, 1100 DE Amsterdam, The Netherlands
    . E-mail

    This Review is part of a thematic series on Inherited Arrhythmogenic Syndromes: The Molecular Revolution, which includes the following articles:

    The Fifteen Years That Shaped Molecular Electrophysiology: Time for Appraisal [Circ Res. 2010;107:451–456]

    Defining a New Paradigm for Human Arrhythmia Syndromes: Phenotypic Manifestations of Gene Mutations in Ion Channel– and Transporter-Associated Proteins [Circ Res. 2010;107:457–465]

    The Cardiac Desmosome and Arrhythmogenic Cardiomyopathies: From Gene to Disease [Circ Res. 2010;107:700–714]

    Phenotypical Manifestations of Mutations in the Genes Encoding Subunits of the Cardiac Voltage–Dependent L-Type Calcium Channel [Circ Res. 2010;107:607–618]

    Inherited Dysfunction of Sarcoplasmic Reticulum Ca2+ Handling and Arrhythmogenesis [Circ Res. 2010;107:871–883]

    Phenotypical Manifestations of Mutations in the Genes Encoding Subunits of the Cardiac Sodium Channel

    Phenotypical Manifestations of Mutations in Genes Encoding Subunits of Cardiac Potassium Channels

    References

    • 1. Wang Q, Shen J, Splawski I, Atkinson D, Li Z, Robinson JL, Moss AJ, Towbin JA, Keating MT . SCN5A mutations associated with an inherited cardiac arrhythmia, long QT syndrome. Cell. 1995; 80:805–811.CrossrefMedlineGoogle Scholar
    • 2. Probst V, Kyndt F, Potet F, Trochu JN, Mialet G, Demolombe S, Schott JJ, Baro I, Escande D, Le MH . Haploinsufficiency in combination with aging causes SCN5A-linked hereditary Lenegre disease. J Am Coll Cardiol. 2003; 41:643–652.CrossrefMedlineGoogle Scholar
    • 3. Schott JJ, Alshinawi C, Kyndt F, Probst V, Hoorntje TM, Hulsbeek M, Wilde AA, Escande D, Mannens MM, Le MH . Cardiac conduction defects associate with mutations in SCN5A. Nat Genet. 1999; 23:20–21.CrossrefMedlineGoogle Scholar
    • 4. Benson DW, Wang DW, Dyment M, Knilans TK, Fish FA, Strieper MJ, Rhodes TH, George AL . Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). J Clin Invest. 2003; 112:1019–1028.CrossrefMedlineGoogle Scholar
    • 5. Ellinor PT, Nam EG, Shea MA, Milan DJ, Ruskin JN, Macrae CA . Cardiac sodium channel mutation in atrial fibrillation. Heart Rhythm. 2008; 5:99–105.CrossrefMedlineGoogle Scholar
    • 6. Tan HL . Sodium channel variants in heart disease: expanding horizons. J Cardiovasc Electrophysiol. 2006; 17(Suppl 1): S151– S157. CrossrefMedlineGoogle Scholar
    • 7. Remme CA, Wilde AA, Bezzina CR . Cardiac sodium channel overlap syndromes: different faces of SCN5A mutations. Trends Cardiovasc Med. 2008; 18:78–87.CrossrefMedlineGoogle Scholar
    • 8. McNair WP, Ku L, Taylor MR, Fain PR, Dao D, Wolfel E, Mestroni L . SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia. Circulation. 2004; 110:2163–2167.LinkGoogle Scholar
    • 9. Schwartz PJ, Priori SG, Dumaine R, Napolitano C, Antzelevitch C, Stramba-Badiale M, Richard TA, Berti MR, Bloise R . A molecular link between the sudden infant death syndrome and the long-QT syndrome. N Engl J Med. 2000; 343:262–267.CrossrefMedlineGoogle Scholar
    • 10. Vatta M, Dumaine R, Varghese G, Richard TA, Shimizu W, Aihara N, Nademanee K, Brugada R, Brugada J, Veerakul G, Li H, Bowles NE, Brugada P, Antzelevitch C, Towbin JA . Genetic and biophysical basis of sudden unexplained nocturnal death syndrome (SUNDS), a disease allelic to Brugada syndrome. Hum Mol Genet. 2002; 11:337–345.CrossrefMedlineGoogle Scholar
    • 11. George AL, Varkony TA, Drabkin HA, Han J, Knops JF, Finley WH, Brown GB, Ward DC, Haas M . Assignment of the human heart tetrodotoxin-resistant voltage-gated Na+ channel alpha-subunit gene (SCN5A) to band 3p21. Cytogenet Cell Genet. 1995; 68:67–70.CrossrefMedlineGoogle Scholar
    • 12. Wang Q, Li Z, Shen J, Keating MT . Genomic organization of the human SCN5A gene encoding the cardiac sodium channel. Genomics. 1996; 34:9–16.CrossrefMedlineGoogle Scholar
    • 13. Koopmann TT, Bezzina CR, Wilde AAM . Voltage-gated sodium channels: action players with many faces. Ann Med. 2006; 38:472–482.CrossrefMedlineGoogle Scholar
    • 14. Zimmer T, Surber R . SCN5A channelopathies–an update on mutations and mechanisms. Prog Biophys Mol Biol. 2008; 98:120–136.CrossrefMedlineGoogle Scholar
    • 15. Grant AO . Molecular biology of sodium channels and their role in cardiac arrhythmias. Am J Med. 2001; 110:296–305.CrossrefMedlineGoogle Scholar
    • 16. Abriel H . Cardiac sodium channel Na(v)1.5 and interacting proteins: physiology and pathophysiology. J Mol Cell Cardiol. 2010; 48:2–11.CrossrefMedlineGoogle Scholar
    • 17. Meadows LS, Isom LL . Sodium channels as macromolecular complexes: implications for inherited arrhythmia syndromes. Cardiovasc Res. 2005; 67:448–458.CrossrefMedlineGoogle Scholar
    • 18. Ackerman MJ, Splawski I, Makielski JC, Tester DJ, Will ML, Timothy KW, Keating MT, Jones G, Chadha M, Burrow CR, Stephens JC, Xu C, Judson R, Curran ME . Spectrum and prevalence of cardiac sodium channel variants among black, white, Asian, and Hispanic individuals: implications for arrhythmogenic susceptibility and Brugada/long QT syndrome genetic testing. Heart Rhythm. 2004; 1:600–607.CrossrefMedlineGoogle Scholar
    • 19. Kapa S, Tester DJ, Salisbury BA, Harris-Kerr C, Pungliya MS, Alders M, Wilde AA, Ackerman MJ . Genetic testing for long-QT syndrome: distinguishing pathogenic mutations from benign variants. Circulation. 2009; 120:1752–1760.LinkGoogle Scholar
    • 20. Wilde AA, Ackerman MJ . Exercise extreme caution when calling rare genetic variants novel arrhythmia syndrome susceptibility mutations. Heart Rhythm. 2010; 7:1883–1885.CrossrefMedlineGoogle Scholar
    • 21. Brugada P, Brugada J . Right bundle branch block, persistent ST segment elevation and sudden cardiac death: a distinct clinical and electrocardiographic syndrome. A multicenter report. J Am Coll Cardiol. 1992; 20:1391–1396.CrossrefMedlineGoogle Scholar
    • 22. Wilde AA, Antzelevitch C, Borggrefe M, Brugada J, Brugada R, Brugada P, Corrado D, Hauer RN, Kass RS, Nademanee K, Priori SG, Towbin JA . Proposed diagnostic criteria for the Brugada syndrome: consensus report. Circulation. 2002; 106:2514–2519.LinkGoogle Scholar
    • 23. Campuzano O, Brugada R, Iglesias A . Genetics of Brugada syndrome. Curr Opin Cardiol. 2010Mar10[Epub ahead of print]. CrossrefMedlineGoogle Scholar
    • 24. Kapplinger JD, Tester DJ, Alders M, Benito B, Berthet M, Brugada J, Brugada P, Fressart V, Guerchicoff A, Harris-Kerr C, Kamakura S, Kyndt F, Koopmann TT, Miyamoto Y, Pfeiffer R, Pollevick GD, Probst V, Zumhagen S, Vatta M, Towbin JA, Shimizu W, Schulze-Bahr E, Antzelevitch C, Salisbury BA, Guicheney P, Wilde AA, Brugada R, Schott JJ, Ackerman MJ . An international compendium of mutations in the SCN5A-encoded cardiac sodium channel in patients referred for Brugada syndrome genetic testing. Heart Rhythm. 2010; 7:33–46.CrossrefMedlineGoogle Scholar
    • 25. Probst V, Wilde AA, Barc J, Sacher F, Babuty D, Mabo P, Mansourati J, Le SS, Kyndt F, Le CC, Guicheney P, Gouas L, Albuisson J, Meregalli PG, Le MH, Tan HL, Schott JJ . SCN5A mutations and the role of genetic background in the pathophysiology of Brugada syndrome. Circ Cardiovasc Genet. 2009; 2:552–557.LinkGoogle Scholar
    • 26. Meregalli PG, Tan HL, Probst V, Koopmann TT, Tanck MW, Bhuiyan ZA, Sacher F, Kyndt F, Schott JJ, Albuisson J, Mabo P, Bezzina CR, Le MH, Wilde AA . Type of SCN5A mutation determines clinical severity and degree of conduction slowing in loss-of-function sodium channelopathies. Heart Rhythm. 2009; 6:341–348.CrossrefMedlineGoogle Scholar
    • 27. Brink PA, Ferreira A, Moolman JC, Weymar HW, van der Merwe PL, Corfield VA . Gene for progressive familial heart block type I maps to chromosome 19q13. Circulation. 1995; 91:1633–1640.LinkGoogle Scholar
    • 28. Kyndt F, Probst V, Potet F, Demolombe S, Chevallier JC, Baro I, Moisan JP, Boisseau P, Schott JJ, Escande D, Le Marec H . Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family. Circulation. 2001; 104:3081–3086.LinkGoogle Scholar
    • 29. Dumaine R, Antzelevitch C . Molecular mechanisms underlying the long QT syndrome. Curr Opin Cardiol. 2002; 17:36–42.CrossrefMedlineGoogle Scholar
    • 30. Roden DM . Long QT syndrome: reduced repolarization reserve and the genetic link. J Intern Med. 2006; 259:59–69.CrossrefMedlineGoogle Scholar
    • 31. Campuzano O, Beltran-Alvarez P, Iglesias A, Scornik F, Perez G, Brugada R . Genetics and cardiac channelopathies. Genet Med. 2010; 12:260–267.CrossrefMedlineGoogle Scholar
    • 32. Eckardt L . LQT3: who is at risk for sudden cardiac death?Heart Rhythm. 2009; 6:121–122.CrossrefMedlineGoogle Scholar
    • 33. Priori SG, Napolitano C, Schwartz PJ, Bloise R, Crotti L, Ronchetti E . The elusive link between LQT3 and Brugada syndrome: the role of flecainide challenge. Circulation. 2000; 102:945–947.LinkGoogle Scholar
    • 34. Schwartz PJ, Priori SG, Locati EH, Napolitano C, Cantu F, Towbin JA, Keating MT, Hammoude H, Brown AM, Chen LS . Long QT syndrome patients with mutations of the SCN5A and HERG genes have differential responses to Na+ channel blockade and to increases in heart rate. Implications for gene-specific therapy. Circulation. 1995; 92:3381–3386.LinkGoogle Scholar
    • 35. Moss AJ, Windle JR, Hall WJ, Zareba W, Robinson JL, McNitt S, Severski P, Rosero S, Daubert JP, Qi M, Cieciorka M, Manalan AS . Safety and efficacy of flecainide in subjects with Long QT-3 syndrome (DeltaKPQ mutation): a randomized, double-blind, placebo-controlled clinical trial. Ann Noninvasive Electrocardiol. 2005; 10(Suppl): 59– 66. CrossrefMedlineGoogle Scholar
    • 36. Ruan Y, Denegri M, Liu N, Bachetti T, Seregni M, Morotti S, Severi S, Napolitano C, Priori SG . Trafficking defects and gating abnormalities of a novel SCN5A mutation question gene-specific therapy in long QT syndrome type 3. Circ Res. 2010; 106:1374–1383.LinkGoogle Scholar
    • 37. Moss AJ, Zareba W, Schwarz KQ, Rosero S, McNitt S, Robinson JL . Ranolazine shortens repolarization in patients with sustained inward sodium current due to type-3 long-QT syndrome. J Cardiovasc Electrophysiol. 2008; 19:1289–1293.CrossrefMedlineGoogle Scholar
    • 38. Bezzina C, Veldkamp MW, van den Berg MP, Postma AV, Rook MB, Viersma JW, van Langen IM, Tan-Sindhunata G, Bink-Boelkens MT, van der Hout AH, Mannens MM, Wilde AA . A single Na(+) channel mutation causing both long-QT and Brugada syndromes. Circ Res. 1999; 85:1206–1213.LinkGoogle Scholar
    • 39. Olson TM, Michels VV, Ballew JD, Reyna SP, Karst ML, Herron KJ, Horton SC, Rodeheffer RJ, Anderson JL . Sodium channel mutations and susceptibility to heart failure and atrial fibrillation. JAMA. 2005; 293:447–454.CrossrefMedlineGoogle Scholar
    • 40. Feinberg WM, Blackshear JL, Laupacis A, Kronmal R, Hart RG . Prevalence, age distribution, and gender of patients with atrial fibrillation. Analysis and implications. Arch Intern Med. 1995; 155:469–473.CrossrefMedlineGoogle Scholar
    • 41. Brugada R, Tapscott T, Czernuszewicz GZ, Marian AJ, Iglesias A, Mont L, Brugada J, Girona J, Domingo A, Bachinski LL, Roberts R . Identification of a genetic locus for familial atrial fibrillation. N Engl J Med. 1997; 336:905–911.CrossrefMedlineGoogle Scholar
    • 42. Campuzano O, Brugada R . Genetics of familial atrial fibrillation. Europace. 2009; 11:1267–1271.CrossrefMedlineGoogle Scholar
    • 43. Darbar D, Kannankeril PJ, Donahue BS, Kucera G, Stubblefield T, Haines JL, George AL, Roden DM . Cardiac sodium channel (SCN5A) variants associated with atrial fibrillation. Circulation. 2008; 117:1927–1935.LinkGoogle Scholar
    • 44. Li Q, Huang H, Liu G, Lam K, Rutberg J, Green MS, Birnie DH, Lemery R, Chahine M, Gollob MH . Gain-of-function mutation of Nav1.5 in atrial fibrillation enhances cellular excitability and lowers the threshold for action potential firing. Biochem Biophys Res Commun. 2009; 380:132–137.CrossrefMedlineGoogle Scholar
    • 45. Makiyama T, Akao M, Shizuta S, Doi T, Nishiyama K, Oka Y, Ohno S, Nishio Y, Tsuji K, Itoh H, Kimura T, Kita T, Horie M . A novel SCN5A gain-of-function mutation M1875T associated with familial atrial fibrillation. J Am Coll Cardiol. 2008; 52:1326–1334.CrossrefMedlineGoogle Scholar
    • 46. Benito B, Brugada R, Perich RM, Lizotte E, Cinca J, Mont L, Berruezo A, Tolosana JM, Freixa X, Brugada P, Brugada J . A mutation in the sodium channel is responsible for the association of long QT syndrome and familial atrial fibrillation. Heart Rhythm. 2008; 5:1434–1440.CrossrefMedlineGoogle Scholar
    • 47. Olesen MS, Jespersen T, Nielsen JB, Liang B, Moller DV, Hedley P, Christiansen M, Varro A, Olesen SP, Haunso S, Schmitt N, Svendsen JH . Mutations in sodium channel beta subunit SCN3B are associated with early onset lone Atrial Fibrillation. Cardiovasc Res. 2010Dec2[Epub ahead of print]. MedlineGoogle Scholar
    • 48. Watanabe H, Darbar D, Kaiser DW, Jiramongkolchai K, Chopra S, Donahue BS, Kannankeril PJ, Roden DM . Mutations in sodium channel beta1- and beta2-subunits associated with atrial fibrillation. Circ Arrhythm Electrophysiol. 2009; 2:268–275.LinkGoogle Scholar
    • 49. Wang P, Yang Q, Wu X, Yang Y, Shi L, Wang C, Wu G, Xia Y, Yang B, Zhang R, Xu C, Cheng X, Li S, Zhao Y, Fu F, Liao Y, Fang F, Chen Q, Tu X, Wang QK . Functional dominant-negative mutation of sodium channel subunit gene SCN3B associated with atrial fibrillation in a Chinese GeneID population. Biochem Biophys Res Commun. 2010; 398:98–104.CrossrefMedlineGoogle Scholar
    • 50. Grant AO, Carboni MP, Neplioueva V, Starmer CF, Memmi M, Napolitano C, Priori S . Long QT syndrome, Brugada syndrome, and conduction system disease are linked to a single sodium channel mutation. J Clin Invest. 2002; 110:1201–1209.CrossrefMedlineGoogle Scholar
    • 51. Smits JP, Koopmann TT, Wilders R, Veldkamp MW, Opthof T, Bhuiyan ZA, Mannens MM, Balser JR, Tan HL, Bezzina CR, Wilde AA . A mutation in the human cardiac sodium channel (E161K) contributes to sick sinus syndrome, conduction disease and Brugada syndrome in two families. J Mol Cell Cardiol. 2005; 38:969–981.CrossrefMedlineGoogle Scholar
    • 52. Groenewegen WA, Firouzi M, Bezzina CR, Vliex S, van Langen IM, Sandkuijl L, Smits JP, Hulsbeek M, Rook MB, Jongsma HJ, Wilde AA . A cardiac sodium channel mutation cosegregates with a rare connexin40 genotype in familial atrial standstill. Circ Res. 2003; 92:14–22.LinkGoogle Scholar
    • 53. Westenskow P, Splawski I, Timothy KW, Keating MT, Sanguinetti MC . Compound mutations: a common cause of severe long-QT syndrome. Circulation. 2004; 109:1834–1841.LinkGoogle Scholar
    • 54. Cordeiro JM, Barajas-Martinez H, Hong K, Burashnikov E, Pfeiffer R, Orsino AM, Wu YS, Hu D, Brugada J, Brugada P, Antzelevitch C, Dumaine R, Brugada R . Compound heterozygous mutations P336L and I1660V in the human cardiac sodium channel associated with the Brugada syndrome. Circulation. 2006; 114:2026–2033.LinkGoogle Scholar
    • 55. Bezzina CR, Shimizu W, Yang P, Koopmann TT, Tanck MW, Miyamoto Y, Kamakura S, Roden DM, Wilde AA . Common sodium channel promoter haplotype in Asian subjects underlies variability in cardiac conduction. Circulation. 2006; 113:338–344.LinkGoogle Scholar
    • 56. Viswanathan PC, Benson DW, Balser JR . A common SCN5A polymorphism modulates the biophysical effects of an SCN5A mutation. J Clin Invest. 2003; 111:341–346.CrossrefMedlineGoogle Scholar
    • 57. Poelzing S, Forleo C, Samodell M, Dudash L, Sorrentino S, Anaclerio M, Troccoli R, Iacoviello M, Romito R, Guida P, Chahine M, Pitzalis M, Deschenes I . SCN5A polymorphism restores trafficking of a Brugada syndrome mutation on a separate gene. Circulation. 2006; 114:368–376.LinkGoogle Scholar
    • 58. Lizotte E, Junttila MJ, Dube MP, Hong K, Benito B, DE ZM, Henkens S, Sarkozy A, Huikuri HV, Towbin J, Vatta M, Brugada P, Brugada J, Brugada R . Genetic modulation of Brugada syndrome by a common polymorphism. J Cardiovasc Electrophysiol. 2009; 20:1137–1141.CrossrefMedlineGoogle Scholar
    • 59. Splawski I, Timothy KW, Tateyama M, Clancy CE, Malhotra A, Beggs AH, Cappuccio FP, Sagnella GA, Kass RS, Keating MT . Variant of SCN5A sodium channel implicated in risk of cardiac arrhythmia. Science. 2002; 297:1333–1336.CrossrefMedlineGoogle Scholar
    • 60. Burke A, Creighton W, Mont E, Li L, Hogan S, Kutys R, Fowler D, Virmani R . Role of SCN5A Y1102 polymorphism in sudden cardiac death in blacks. Circulation. 2005; 112:798–802.LinkGoogle Scholar
    • 61. Plant LD, Bowers PN, Liu Q, Morgan T, Zhang T, State MW, Chen W, Kittles RA, Goldstein SA . A common cardiac sodium channel variant associated with sudden infant death in African Americans, SCN5A S1103Y. J Clin Invest. 2006; 116:430–435.CrossrefMedlineGoogle Scholar
    • 62. Qin N, D'Andrea MR, Lubin ML, Shafaee N, Codd EE, Correa AM . Molecular cloning and functional expression of the human sodium channel beta1B subunit, a novel splicing variant of the beta1 subunit. Eur J Biochem. 2003; 270:4762–4770.CrossrefMedlineGoogle Scholar
    • 63. Dhar MJ, Chen C, Rivolta I, Abriel H, Malhotra R, Mattei LN, Brosius FC, Kass RS, Isom LL . Characterization of sodium channel alpha- and beta-subunits in rat and mouse cardiac myocytes. Circulation. 2001; 103:1303–1310.LinkGoogle Scholar
    • 64. Zimmer T, Biskup C, Bollensdorff C, Benndorf K . The beta1 subunit but not the beta2 subunit colocalizes with the human heart Na+ channel (hH1) already within the endoplasmic reticulum. J Membr Biol. 2002; 186:13–21.CrossrefMedlineGoogle Scholar
    • 65. Makita N, Bennett PB, George AL . Molecular determinants of beta 1 subunit-induced gating modulation in voltage-dependent Na+ channels. J Neurosci. 1996; 16:7117–7127.CrossrefMedlineGoogle Scholar
    • 66. McCormick KA, Srinivasan J, White K, Scheuer T, Catterall WA . The extracellular domain of the beta1 subunit is both necessary and sufficient for beta1-like modulation of sodium channel gating. J Biol Chem. 1999; 274:32638–32646.CrossrefMedlineGoogle Scholar
    • 67. Meadows L, Malhotra JD, Stetzer A, Isom LL, Ragsdale DS . The intracellular segment of the sodium channel beta 1 subunit is required for its efficient association with the channel alpha subunit. J Neurochem. 2001; 76:1871–1878.CrossrefMedlineGoogle Scholar
    • 68. An RH, Wang XL, Kerem B, Benhorin J, Medina A, Goldmit M, Kass RS . Novel LQT-3 mutation affects Na+ channel activity through interactions between alpha- and beta1-subunits. Circ Res. 1998; 83:141–146.LinkGoogle Scholar
    • 69. Qu Y, Isom LL, Westenbroek RE, Rogers JC, Tanada TN, McCormick KA, Scheuer T, Catterall WA . Modulation of cardiac Na+ channel expression in Xenopus oocytes by beta 1 subunits. J Biol Chem. 1995; 270:25696–25701.CrossrefMedlineGoogle Scholar
    • 70. Bezzina CR, Rook MB, Groenewegen WA, Herfst LJ, van der Wal AC, Lam J, Jongsma HJ, Wilde AA, Mannens MM . Compound heterozygosity for mutations (W156X and R225W) in SCN5A associated with severe cardiac conduction disturbances and degenerative changes in the conduction system. Circ Res. 2003; 92:159–168.LinkGoogle Scholar
    • 71. Valdivia CR, Nagatomo T, Makielski JC . Late Na currents affected by alpha subunit isoform and beta1 subunit co-expression in HEK293 cells. J Mol Cell Cardiol. 2002; 34:1029–1039.CrossrefMedlineGoogle Scholar
    • 72. Ko SH, Lenkowski PW, Lee HC, Mounsey JP, Patel MK . Modulation of Na(v)1.5 by beta1– and beta3-subunit co-expression in mammalian cells. Pflugers Arch. 2005; 449:403–412.CrossrefMedlineGoogle Scholar
    • 73. Lopez-Santiago LF, Meadows LS, Ernst SJ, Chen C, Malhotra JD, McEwen DP, Speelman A, Noebels JL, Maier SK, Lopatin AN, Isom LL . Sodium channel Scn1b null mice exhibit prolonged QT and RR intervals. J Mol Cell Cardiol. 2007; 43:636–647.CrossrefMedlineGoogle Scholar
    • 74. Watanabe H, Koopmann TT, Le SS, Yang T, Ingram CR, Schott JJ, Demolombe S, Probst V, Anselme F, Escande D, Wiesfeld AC, Pfeufer A, Kaab S, Wichmann HE, Hasdemir C, Aizawa Y, Wilde AA, Roden DM, Bezzina CR . Sodium channel beta1 subunit mutations associated with Brugada syndrome and cardiac conduction disease in humans. J Clin Invest. 2008; 118:2260–2268.MedlineGoogle Scholar
    • 75. Nattel S, Maguy A, Le BS, Yeh YH . Arrhythmogenic ion-channel remodeling in the heart: heart failure, myocardial infarction, and atrial fibrillation. Physiol Rev. 2007; 87:425–456.CrossrefMedlineGoogle Scholar
    • 76. Maier SK, Westenbroek RE, McCormick KA, Curtis R, Scheuer T, Catterall WA . Distinct subcellular localization of different sodium channel alpha and beta subunits in single ventricular myocytes from mouse heart. Circulation. 2004; 109:1421–1427.LinkGoogle Scholar
    • 77. Zimmer T, Benndorf K . The intracellular domain of the beta 2 subunit modulates the gating of cardiac Na v 1.5 channels. Biophys J. 2007; 92:3885–3892.CrossrefMedlineGoogle Scholar
    • 78. Johnson D, Bennett ES . Isoform-specific effects of the beta2 subunit on voltage-gated sodium channel gating. J Biol Chem. 2006; 281:25875–25881.CrossrefMedlineGoogle Scholar
    • 79. Valdivia CR, Medeiros-Domingo A, Ye B, Shen WK, Algiers TJ, Ackerman MJ, Makielski JC . Loss-of-function mutation of the SCN3B-encoded sodium channel {beta}3 subunit associated with a case of idiopathic ventricular fibrillation. Cardiovasc Res. 2010; 86:392–400.CrossrefMedlineGoogle Scholar
    • 80. Fahmi AI, Patel M, Stevens EB, Fowden AL, John JE, Lee K, Pinnock R, Morgan K, Jackson AP, Vandenberg JI . The sodium channel beta-subunit SCN3b modulates the kinetics of SCN5a and is expressed heterogeneously in sheep heart. J Physiol. 2001; 537:693–700.CrossrefMedlineGoogle Scholar
    • 81. Hakim P, Thresher R, Grace AA, Huang CL . Effects of flecainide and quinidine on action potential and ventricular arrhythmogenic properties in Scn3b knockout mice. Clin Exp Pharmacol Physiol. 2010; 37:782–789.CrossrefMedlineGoogle Scholar
    • 82. Hu D, Barajas-Martinez H, Burashnikov E, Springer M, Wu Y, Varro A, Pfeiffer R, Koopmann TT, Cordeiro JM, Guerchicoff A, Pollevick GD, Antzelevitch C . A mutation in the beta 3 subunit of the cardiac sodium channel associated with Brugada ECG phenotype. Circ Cardiovasc Genet. 2009; 2:270–278.LinkGoogle Scholar
    • 83. Smits JP, Eckardt L, Probst V, Bezzina CR, Schott JJ, Remme CA, Haverkamp W, Breithardt G, Escande D, Schulze-Bahr E, LeMarec H, Wilde AA . Genotype-phenotype relationship in Brugada syndrome: electrocardiographic features differentiate SCN5A-related patients from non-SCN5A-related patients. J Am Coll Cardiol. 2002; 40:350–356.CrossrefMedlineGoogle Scholar
    • 84. Tan BH, Pundi KN, Van Norstrand DW, Valdivia CR, Tester DJ, Medeiros-Domingo A, Makielski JC, Ackerman MJ . Sudden infant death syndrome-associated mutations in the sodium channel beta subunits. Heart Rhythm. 2010; 7:771–778.CrossrefMedlineGoogle Scholar
    • 85. Medeiros-Domingo A, Kaku T, Tester DJ, Iturralde-Torres P, Itty A, Ye B, Valdivia C, Ueda K, Canizales-Quinteros S, Tusie-Luna MT, Makielski JC, Ackerman MJ . SCN4B-encoded sodium channel beta4 subunit in congenital long-QT syndrome. Circulation. 2007; 116:134–142.LinkGoogle Scholar
    • 86. Remme CA, Scicluna BP, Verkerk AO, Amin AS, van BS, Beekman L, Deneer VH, Chevalier C, Oyama F, Miyazaki H, Nukina N, Wilders R, Escande D, Houlgatte R, Wilde AA, Tan HL, Veldkamp MW, de Bakker JM, Bezzina CR . Genetically determined differences in sodium current characteristics modulate conduction disease severity in mice with cardiac sodium channelopathy. Circ Res. 2009; 104:1283–1292.LinkGoogle Scholar
    • 87. Yu FH, Westenbroek RE, Silos-Santiago I, McCormick KA, Lawson D, Ge P, Ferriera H, Lilly J, DiStefano PS, Catterall WA, Scheuer T, Curtis R . Sodium channel beta4, a new disulfide-linked auxiliary subunit with similarity to beta2. J Neurosci. 2003; 23:7577–7585.CrossrefMedlineGoogle Scholar
    • 88. Amin AS, Asghari-Roodsari A, Tan HL . Cardiac sodium channelopathies. Pflugers Arch. 2010; 460:223–237.CrossrefMedlineGoogle Scholar
    • 89. Weiss R, Barmada MM, Nguyen T, Seibel JS, Cavlovich D, Kornblit CA, Angelilli A, Villanueva F, McNamara DM, London B . Clinical and molecular heterogeneity in the Brugada syndrome: a novel gene locus on chromosome 3. Circulation. 2002; 105:707–713.LinkGoogle Scholar
    • 90. London B, Michalec M, Mehdi H, Zhu X, Kerchner L, Sanyal S, Viswanathan PC, Pfahnl AE, Shang LL, Madhusudanan M, Baty CJ, Lagana S, Aleong R, Gutmann R, Ackerman MJ, McNamara DM, Weiss R, Dudley SC . Mutation in glycerol-3-phosphate dehydrogenase 1 like gene (GPD1-L) decreases cardiac Na+ current and causes inherited arrhythmias. Circulation. 2007; 116:2260–2268.LinkGoogle Scholar
    • 91. Brisson D, Vohl MC, St-Pierre J, Hudson TJ, Gaudet D . Glycerol: a neglected variable in metabolic processes?Bioessays. 2001; 23:534–542.CrossrefMedlineGoogle Scholar
    • 92. Valdivia CR, Ueda K, Ackerman MJ, Makielski JC . GPD1L links redox state to cardiac excitability by PKC-dependent phosphorylation of the sodium channel SCN5A. Am J Physiol Heart Circ Physiol. 2009; 297:H1446–H1452.CrossrefMedlineGoogle Scholar
    • 93. Van Norstrand DW, Valdivia CR, Tester DJ, Ueda K, London B, Makielski JC, Ackerman MJ . Molecular and functional characterization of novel glycerol-3-phosphate dehydrogenase 1 like gene (GPD1-L) mutations in sudden infant death syndrome. Circulation. 2007; 116:2253–2259.LinkGoogle Scholar
    • 94. Murray KT, Hu NN, Daw JR, Shin HG, Watson MT, Mashburn AB, George AL . Functional effects of protein kinase C activation on the human cardiac Na+ channel. Circ Res. 1997; 80:370–376.LinkGoogle Scholar
    • 95. Yarbrough TL, Lu T, Lee HC, Shibata EF . Localization of cardiac sodium channels in caveolin-rich membrane domains: regulation of sodium current amplitude. Circ Res. 2002; 90:443–449.LinkGoogle Scholar
    • 96. Vatta M, Ackerman MJ, Ye B, Makielski JC, Ughanze EE, Taylor EW, Tester DJ, Balijepalli RC, Foell JD, Li Z, Kamp TJ, Towbin JA . Mutant caveolin-3 induces persistent late sodium current and is associated with long-QT syndrome. Circulation. 2006; 114:2104–2112.LinkGoogle Scholar
    • 97. Scriven DR, Klimek A, Asghari P, Bellve K, Moore ED . Caveolin-3 is adjacent to a group of extradyadic ryanodine receptors. Biophys J. 2005; 89:1893–1901.CrossrefMedlineGoogle Scholar
    • 98. Lu T, Lee HC, Kabat JA, Shibata EF . Modulation of rat cardiac sodium channel by the stimulatory G protein alpha subunit. J Physiol. 1999; 518:371–384.CrossrefMedlineGoogle Scholar
    • 99. Palygin OA, Pettus JM, Shibata EF . Regulation of caveolar cardiac sodium current by a single Gsalpha histidine residue. Am J Physiol Heart Circ Physiol. 2008; 294:H1693–H1699.CrossrefMedlineGoogle Scholar
    • 100. Cronk LB, Ye B, Kaku T, Tester DJ, Vatta M, Makielski JC, Ackerman MJ . Novel mechanism for sudden infant death syndrome: persistent late sodium current secondary to mutations in caveolin-3. Heart Rhythm. 2007; 4:161–166.CrossrefMedlineGoogle Scholar
    • 101. Palygin OA, Pettus JM, Shibata EF . Regulation of caveolar cardiac sodium current by a single Gsalpha histidine residue. Am J Physiol Heart Circ Physiol. 2008; 294:H1693–H1699.CrossrefMedlineGoogle Scholar
    • 102. Gavillet B, Rougier JS, Domenighetti AA, Behar R, Boixel C, Ruchat P, Lehr HA, Pedrazzini T, Abriel H . Cardiac sodium channel Nav1.5 is regulated by a multiprotein complex composed of syntrophins and dystrophin. Circ Res. 2006; 99:407–414.LinkGoogle Scholar
    • 103. Williams JC, Armesilla AL, Mohamed TM, Hagarty CL, McIntyre FH, Schomburg S, Zaki AO, Oceandy D, Cartwright EJ, Buch MH, Emerson M, Neyses L . The sarcolemmal calcium pump, alpha-1 syntrophin, and neuronal nitric-oxide synthase are parts of a macromolecular protein complex. J Biol Chem. 2006; 281:23341–23348.CrossrefMedlineGoogle Scholar
    • 104. Oceandy D, Cartwright EJ, Emerson M, Prehar S, Baudoin FM, Zi M, Alatwi N, Venetucci L, Schuh K, Williams JC, Armesilla AL, Neyses L . Neuronal nitric oxide synthase signaling in the heart is regulated by the sarcolemmal calcium pump 4b. Circulation. 2007; 115:483–492.LinkGoogle Scholar
    • 105. Ahern GP, Hsu SF, Klyachko VA, Jackson MB . Induction of persistent sodium current by exogenous and endogenous nitric oxide. J Biol Chem. 2000; 275:28810–28815.CrossrefMedlineGoogle Scholar
    • 106. Ueda K, Valdivia C, Medeiros-Domingo A, Tester DJ, Vatta M, Farrugia G, Ackerman MJ, Makielski JC . Syntrophin mutation associated with long QT syndrome through activation of the nNOS-SCN5A macromolecular complex. Proc Natl Acad Sci U S A. 2008; 105:9355–9360.CrossrefMedlineGoogle Scholar
    • 107. Wu G, Ai T, Kim JJ, Mohapatra B, Xi Y, Li Z, Abbasi S, Purevjav E, Samani K, Ackerman MJ, Qi M, Moss AJ, Shimizu W, Towbin JA, Cheng J, Vatta M . alpha-1-syntrophin mutation and the long-QT syndrome: a disease of sodium channel disruption. Circ Arrhythm Electrophysiol. 2008; 1:193–201.LinkGoogle Scholar
    • 108. Cheng J, Van Norstrand DW, Medeiros-Domingo A, Valdivia C, Tan BH, Ye B, Kroboth S, Vatta M, Tester DJ, January CT, Makielski JC, Ackerman MJ . Alpha1-syntrophin mutations identified in sudden infant death syndrome cause an increase in late cardiac sodium current. Circ Arrhythm Electrophysiol. 2009; 2:667–676.LinkGoogle Scholar
    • 109. Tan HL, Kupershmidt S, Zhang R, Stepanovic S, Roden DM, Wilde AA, Anderson ME, Balser JR . A calcium sensor in the sodium channel modulates cardiac excitability. Nature. 2002; 415:442–447.CrossrefMedlineGoogle Scholar
    • 110. Mohler PJ, Rivolta I, Napolitano C, LeMaillet G, Lambert S, Priori SG, Bennett V . Nav1.5 E1053K mutation causing Brugada syndrome blocks binding to ankyrin-G and expression of Nav1.5 on the surface of cardiomyocytes. Proc Natl Acad Sci U S A. 2004; 101:17533–17538.CrossrefMedlineGoogle Scholar
    • 111. Lowe JS, Palygin O, Bhasin N, Hund TJ, Boyden PA, Shibata E, Anderson ME, Mohler PJ . Voltage-gated Nav channel targeting in the heart requires an ankyrin-G dependent cellular pathway. J Cell Biol. 2008; 180:173–186.CrossrefMedlineGoogle Scholar
    • 112. Liu CJ, Dib-Hajj SD, Renganathan M, Cummins TR, Waxman SG . Modulation of the cardiac sodium channel Nav1.5 by fibroblast growth factor homologous factor 1B. J Biol Chem. 2003; 278:1029–1036.CrossrefMedlineGoogle Scholar

    eLetters(0)

    eLetters should relate to an article recently published in the journal and are not a forum for providing unpublished data. Comments are reviewed for appropriate use of tone and language. Comments are not peer-reviewed. Acceptable comments are posted to the journal website only. Comments are not published in an issue and are not indexed in PubMed. Comments should be no longer than 500 words and will only be posted online. References are limited to 10. Authors of the article cited in the comment will be invited to reply, as appropriate.

    Comments and feedback on AHA/ASA Scientific Statements and Guidelines should be directed to the AHA/ASA Manuscript Oversight Committee via its Correspondence page.