Entry - *600163 - SODIUM VOLTAGE-GATED CHANNEL, ALPHA SUBUNIT 5; SCN5A - OMIM
* 600163

SODIUM VOLTAGE-GATED CHANNEL, ALPHA SUBUNIT 5; SCN5A


Alternative titles; symbols

SODIUM CHANNEL, VOLTAGE-GATED, TYPE V, ALPHA SUBUNIT
NAV1.5


HGNC Approved Gene Symbol: SCN5A

Cytogenetic location: 3p22.2     Genomic coordinates (GRCh38): 3:38,548,062-38,649,687 (from NCBI)


Gene-Phenotype Relationships
Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
3p22.2 {Sudden infant death syndrome, susceptibility to} 272120 AR 3
Atrial fibrillation, familial, 10 614022 AD 3
Brugada syndrome 1 601144 AD 3
Cardiomyopathy, dilated, 1E 601154 AD 3
Heart block, nonprogressive 113900 AD 3
Heart block, progressive, type IA 113900 AD 3
Long QT syndrome 3 603830 AD 3
Sick sinus syndrome 1 608567 AR 3
Ventricular fibrillation, familial, 1 603829 3

TEXT

Cloning and Expression

Gellens et al. (1992) cloned and characterized the cardiac sodium channel gene SCN5A. The deduced 2,016-amino acid protein has a structure similar to that of previously characterized sodium channels (see 182392) and contains 4 homologous domains, each of which has 6 putative membrane-spanning regions.

Freyermuth et al. (2016) stated that alternative splicing creates fetal and adult isoforms of SCN5A that differ in inclusion of alternative exons 6a or 6b, respectively. Both exons 6 have 92 bp, but encode 7 different amino acids in the voltage-sensor region of SCN5A domain I.


Gene Structure

Wang et al. (1996) found that SCN5A consists of 28 exons spanning approximately 80 kb. They described the sequences of all intron/exon boundaries and a dinucleotide repeat polymorphism in intron 16.


Mapping

George et al. (1995) mapped the SCN5A gene to chromosome 3p21 by fluorescence in situ hybridization, thus making it an important candidate gene for long QT syndrome-3 (LQT3; 603830).

Gross (2019) mapped the SCN5A gene to chromosome 3p22.2 based on an alignment of the SCN5A sequence (GenBank BC051374) with the genomic sequence (GRCh38).


Gene Function

By immunoprecipitation and nano-liquid chromatography-mass spectroscopy/mass spectroscopy of transgenic mouse bone marrow macrophages expressing the human macrophage splice variant of SCN5A, followed by Western blot analysis, Jones et al. (2014) identified interaction of SCN5A with activating transcription factor-2 (ATF2; 123811). Microarray analysis of SCN5A-positive macrophages revealed increased expression of Sp100 (604585), an Atf2-regulated gene. Knockdown of Adcy8 (103070), the calcium-dependent isoform of adenylate cyclase, inhibited channel agonist-induced expression of Sp100-related genes. Activation of SCN5A increased expression of cAMP in macrophages. Treatment of macrophages with poly(I:C), a mimic of viral double-stranded RNA, activated the Adcy8 signaling pathway to regulate expression of Sp100-related genes and Ifnb (147640). Electrophysiologic analysis showed that the SCN5A variant mediated nonselective outward currents, as well as a small yet detectable inward current. Jones et al. (2014) proposed that human macrophage SCN5A initiates signaling in an innate immune pathway relevant to antiviral host defense, and that SCN5A is a pathogen sensor.

Myotonic dystrophy (see DM1, 160900) is caused by expression of mutant RNAs containing expanded CUG repeats. These repeats sequester muscleblind-like (MBNL; see MBNL1, 606516) splicing factors in nuclear RNA foci, resulting in changes in pre-mRNA splicing. Freyermuth et al. (2016) showed that MBNL1 specifically promoted inclusion of exon 6b in SCN5A pre-mRNA and expression of the adult SCN5A isoform. Freyermuth et al. (2016) found that left ventricle samples of 3 adult DM1 patients showed alternative splicing in a number of genes, including SCN5A. A portion of the SCN5A mRNA in these samples was the fetal isoform. When expressed in Xenopus oocytes, the fetal isoform of SCN5A showed reduced excitability compared with the adult SCN5A isoform. In mice, expression of fetal Scn5a promoted heart arrhythmia and cardiac-conduction delay, which are 2 predominant features of myotonic dystrophy.


Molecular Genetics

Missense mutations in the skeletal muscle sodium channel gene, SCN4A (603967), cause myotonia. Physiologic data show that these mutations affect sodium channel inactivation and lead to repetitive depolarizations, consistent with the myotonic phenotype. By analogy, similar mutations in the cardiac sodium channel gene might be expected to cause a phenotype like LQT. Indeed, Wang et al. (1995) found a mutation in the SCN5A gene in families with chromosome 3-linked LQT (see 600163.0001).

Bennett et al. (1995) determined the functional defect resulting from the 3-amino acid (KPQ) deletion (600163.0001) in the SCN5A protein. By expression of recombinant human heart sodium channels in Xenopus laevis oocytes, mutant channels showed a sustained inward current during membrane depolarization. Single-channel recordings indicated that mutant channels fluctuate between normal and noninactivating gating modes. Persistent inward sodium current explains prolongation of cardiac action potentials and provides a molecular mechanism for the chromosome 3-linked form of long QT syndrome.

Wang et al. (1995) identified SCN5A mutations in affected members of 4 additional families with chromosome 3-linked LQT. Two of the families had the same 9-bp deletion found earlier; the other families were found to have missense mutations affecting highly conserved amino acid residues (600163.0002 and 600163.0003). The location and character of the mutation suggested to the authors that this form of LQT results from a delay in cardiac sodium channel fast inactivation or altered voltage-dependence of inactivation.

Wang et al. (1996) determined the biophysical and functional characteristics of each of the 3 distinct mutations that had been identified in the cardiac sodium channel gene in patients with LQT3 to that time. For this they used heterologous expression of a recombinant human heart sodium channel in a mammalian cell line. Each mutation caused a sustained, noninactivating sodium current amounting to a few percent of the peak inward sodium current, observable during long (more than 50 msec) depolarizations. The voltage dependence and rate of inactivation were altered and the rate of recovery from inactivation was changed compared with wildtype channels. These mutations in diverse regions of the ion channel protein all produced a common defect in channel gating that can cause the long QT phenotype. The sustained inward current caused by these mutations would prolong the action potential. Furthermore, they might create conditions that promote arrhythmias due to prolonged depolarization and the altered recovery from inactivation.

Wang et al. (1997) explored the potential for targeted suppression of the defect in LQT3 by heterologous expression of mutant channels in cultured human cells. Channel behavior and inhibition by mexiletine were investigated by whole-cell patch-clamp methods. The investigators showed that late-opening LQT3 mutant channels were much more sensitive to inhibition by mexiletine than were wildtype sodium channels. The defective late openings were selectively suppressed more than the peak sodium current and these late openings could be suppressed by concentrations at the lower end of the therapeutic range.

Using a candidate gene approach, Chen et al. (1998) studied 6 small families and 2 sporadic patients with idiopathic ventricular fibrillation (IVF; 603829) using SSCP and DNA sequence analyses to identify mutations in known ion channel genes, including the cardiac sodium channel gene SCN5A. They identified several mutations in families with a distinct form of IVF known as Brugada syndrome (BRGDA1; 601144). In 1 family all affected members had 2 mutations: an arg1232-to-trp mutation in exon 21 of the gene in the extracellular loop between transmembrane segments S1 and S2 of domain III of the protein, and a thr1620-to-met mutation in exon 28 of the gene in the extracellular loop between S3 and S4 of domain IV of the protein (600163.0004). Additional SCN5A mutations were found in 2 IVF families: insertion of 2 nucleotides (AA) in the splice-donor sequence of intron 7 (600163.0005); and deletion of a single nucleotide (A) at codon 1397, resulting in an in-frame stop codon (600163.0006). The frameshift mutation caused the sodium channel to be nonfunctional.

Schott et al. (1999) reported a mutation in the SCN5A gene that segregated with progressive familial heart block (PFHB1A; 113900) in an autosomal dominant manner in a large French family. In a smaller Dutch family, another SCN5A mutation cosegregated with familial nonprogressive conduction defect (see 113900). The French family with PFHB1A was identified through a member with right bundle branch block (RBBB) and syncope; a brother had RBBB, and a sister had complete atrioventricular (AV) block and syncope. Clinical and electrocardiographic abnormalities were found in 15 members of the family; mean QRS duration was 135 +/- 7 ms. RBBB was present in 5, left bundle branch block (LBBB) in 2, left anterior or posterior hemiblock in 3, and long PR interval (more than 210 ms) in 8. None had a structural heart disease. Four members of earlier generations had received a pacemaker implantation because of syncope or complete AV block. Long-term follow-up of several affected members demonstrated that their conduction defect increased in severity with age. In the Dutch family, the proband presented after birth with an asymptomatic first-degree AV block associated with RBBB (PR interval and QRS duration, 200 and 120 ms, respectively). In the French family, Schott et al. (1999) excluded the chromosome 19 locus for this disorder (604559) by linkage studies, as well as other loci for inherited cardiac disorders associated with conduction defects. SCN5A was considered a candidate locus, and using markers flanking SCN5A, the authors demonstrated segregation of the disease with D3S1260 in every affected individual (maximum lod score of 6.03 at theta of 0.0). A donor splice site mutation in SCN5A was found in the French family (600163.0009), and a frameshift mutation was identified in the Dutch family (600163.0010). Clinical data and family histories indicated that none of the affected individuals in these 2 families had LQT3 or idiopathic ventricular fibrillation (Brugada syndrome). Therefore, PFHB1 represents a third cardiac disease linked to SCN5A.

Splawski et al. (2000) screened 262 unrelated individuals with LQT syndrome for mutations in the 5 defined genes (KCNQ1, 607542; KCNH2, 152427; SCN5A; KCNE1, 176261; and KCNE2, 603796) and identified mutations in 177 individuals (68%). KCNQ1 and KCNH2 accounted for 87% of mutations (42% and 45%, respectively), and SCN5A, KCNE1, and KCNE2 for the remaining 13% (8%, 3%, and 2%, respectively).

Tan et al. (2002) demonstrated that calmodulin (114180) binds to the carboxy terminal 'IQ' domain of the SCN5A in a calcium-dependent manner. This binding interaction significantly enhances slow inactivation, a channel-gating process linked to life-threatening idiopathic ventricular arrhythmias. Mutations targeted to the IQ domain disrupted calmodulin binding and eliminated calcium/calmodulin-dependent slow inactivation, whereas the gating effects of calcium/calmodulin were restored by intracellular application of a peptide modeled after the IQ domain. A naturally occurring mutation (A1924T; 600163.0012) in the IQ domain altered SCN5A function in a manner characteristic of the Brugada syndrome, but at the same time inhibited slow inactivation induced by calcium/calmodulin, yielding a clinically benign (arrhythmia-free) phenotype.

Splawski et al. (2002) identified a common variant of the SCN5A gene, ser1103 to tyr (S1103Y; 600163.0024), which is present in 13.2% of African Americans and is associated with accelerated channel activation, increasing the likelihood of abnormal cardiac repolarization and arrhythmia. Splawski et al. (2002) suggested that the S1103Y mutation in the African American population may be a useful molecular marker for the prediction of arrhythmia susceptibility in the context of additional acquired risk factors such as the use of certain medications or the presence of hypokalemia.

Rivolta et al. (2001) identified 2 mutations at the same codon of the SCN5A gene: a tyr1795-to-cys mutation (Y1795C; 600163.0029) in a patient with LQT3, and a Y1795H (600163.0030) mutation in a patient with Brugada syndrome. Functional analysis revealed marked and opposing effects on channel gating consistent with activity associated with the cellular basis of each clinical disorder: Y1795H accelerated and Y1795C slowed the onset of activation; Y1795H, but not Y1795C, caused a marked negative shift in the voltage dependence of inactivation; and neither affected the kinetics of the recovery from inactivation. However, both mutations increased the expression of sustained Na(+) channel activity compared with wildtype channels, although this effect was most pronounced for the Y1795C mutation, and both promoted entrance into an intermediate or slowly developing inactivated state. Rivolta et al. (2001) concluded that these data confirmed the key role of the C-terminal tail of the cardiac Na(+) channel in the control of channel gating and provided further evidence of the close interrelationship between Brugada syndrome and LQT3 at the molecular level.

Clancy et al. (2002) performed detailed kinetic analyses of the Y1795C mutant described by Rivolta et al. (2001). Theoretical entry and exit rates from the bursting mode of gating were derived from single channels. Computational analysis suggested that the amount of time mutant channels spend bursting (burst mode dwell time) is primarily responsible for rate-dependent changes in single-channel bursting and macroscopic inward sodium channel (I-sus), hence delaying repolarization and prolonging the QT interval. This prediction was experimentally confirmed by analysis of delta-KPQ mutant channels (600163.0001) for which the burst mode exit rate (determined by the burst mode dwell time) was found to be very similar to the derived rate for Y1795C channels. These results provided an explanation of the molecular mechanism for bradycardia-induced QT prolongation in patients carrying LQT3 mutations.

Veldkamp et al. (2003) studied the effect of the 1795insD SCN5A mutation (600163.0013), which causes LQT3 or Brugada syndrome, on sinoatrial (SA) pacemaking. Activity of 1795insD channels during SA node pacemaking was confirmed by action potential (AP) clamp experiments, and the previously characterized persistent inward current (I-pst) and negative shift were implemented into SA node (AP) models. The -10 mV shift decreased the sinus rate by decreasing the diastolic depolarization rate, whereas the I-pst decreased the sinus rate by AP prolongation, despite a concomitant increase in the diastolic depolarization rate. In combination, a moderate I-pst (1 to 2%) and the shift reduced the sinus rate by about 10%. Veldkamp et al. (2003) concluded that sodium channel mutations displaying an I-pst or a negative shift in inactivation may account for the bradycardia seen in LQT3 patients, whereas SA node pauses or arrest may result from failure of SA node cells to repolarize under conditions of extra net inward current.

Based on prior associations with disorders of cardiac rhythm and conduction, Benson et al. (2003) screened the SCN5A gene as a candidate gene in 10 pediatric patients from 7 families who were diagnosed with autosomal recessive congenital sick sinus syndrome (SSS1; 608567) during the first decade of life. Probands from 3 kindreds exhibited compound heterozygosity for 6 distinct SCN5A alleles (e.g., 600163.0025), 2 of which had previously been associated with dominant disorders of cardiac excitability. Biophysical characterization of the mutants using heterologously expressed recombinant human heart sodium channels demonstrated loss of function or significant impairment in channel gating that predicted reduced myocardial excitability. Thus Benson et al. (2003) provided a molecular basis for some forms of congenital SSS and defined a recessive disorder of a human heart voltage-gated sodium channel.

In a patient with Brugada syndrome, Mohler et al. (2004) identified an E1053K mutation (600163.0033) in the ankyrin-binding motif of Na(v)1.5. The mutation abolished binding of Na(v)1.5 to ankyrin-G (ANK3; 600465), and also prevented accumulation of Na(v)1.5 at cell surface sites in ventricular cardiomyocytes. Both ankyrin-G and Na(v)1.5 localized at intercalated disc and T-tubule membranes in cardiomyocytes, and Na(v)1.5 coimmunoprecipitated with the 190-kD ankyrin-G isoform from detergent-soluble lysates from rat heart. These data suggested that Na(v)1.5 associates with ankyrin-G and that ankyrin-G is required for Na(v)1.5 localization at excitable membranes in cardiomyocytes.

Miller et al. (2004) reported a case of repeated germline transmission of a severe form of LQT syndrome from an asymptomatic mother with somatic mosaicism for a mutation in the SCN5A gene (600163.0007).

Maekawa et al. (2005) sequenced the SCN5A gene in 166 Japanese patients with arrhythmia who were not diagnosed with LQT or Brugada syndrome and in 232 healthy controls, identifying 69 genetic variations including 66 SNPs. The frequency of a 703+130G-A SNP was significantly higher in patients than in controls (OR, 1.70), suggesting an association with an unknown risk factor for arrhythmia. Haplotype analysis revealed that the so-called GG haplotype with both the leu1988-to-arg and his558-to-arg (600163.0031) SNPs was significantly less frequent in patients than in controls (p = 0.018), suggesting a possible protective effect.

Tester et al. (2005) analyzed 5 LQTS-associated cardiac channel genes in 541 consecutive unrelated patients with LQT syndrome (average QTc, 482 ms). In 272 (50%) patients, they identified 211 different pathogenic mutations, including 88 in KCNQ1, 89 in KCNH2, 32 in SCN5A, and 1 each in KCNE1 and KCNE2. Mutations considered pathogenic were absent in more than 1,400 reference alleles. Among the mutation-positive patients, 29 (11%) had 2 LQTS-causing mutations, of which 16 (8%) were in 2 different LQTS genes (biallelic digenic). Tester et al. (2005) noted that patients with multiple mutations were younger at diagnosis, but they did not discern any genotype/phenotype correlations associated with location or type of mutation.

In 44 unrelated patients with LQT syndrome, Millat et al. (2006) used DHLP chromatography to analyze the KCNQ1, KCNH2, SCN5A, KCNE1, and KCNE2 genes for mutations and SNPs. Most of the patients (84%) showed a complex molecular pattern, with an identified mutation associated with 1 or more SNPs located in several LQTS genes; 4 of the patients also had a second mutation in a different LQTS gene (biallelic digenic inheritance; see, e.g., 600163.0007 and 603796.0005).

In affected members of the family reported by Greenlee et al. (1986) with a form of dilated cardiomyopathy (CMD1E; 601154), McNair et al. (2004) identified heterozygosity for a missense mutation (D1765N; 600163.0001) in the SCN5A gene. In affected members of a family with atrial standstill (ATRST1; 108770), Groenewegen et al. (2003) had identified coinheritance of the D1275N mutation in the SCN5A gene with polymorphisms in the atria-specific junction channel protein connexin-40 (GJA5; 121013). None of the patients with atrial standstill had dilated cardiomyopathy, leading Groenewegen and Wilde (2005) to question the relationship of the SCN5A mutation to dilated cardiomyopathy in the family reported by McNair et al. (2004). McNair et al. (2005) responded that the younger age of the affected members studied by Groenewegen et al. (2003) as well as additional or genetic environmental factors may account for the difference between the 2 families.

In a Japanese family in which an 11-year-old boy had sick sinus syndrome that progressed to atrial standstill, Makita et al. (2005) analyzed 3 cardiac ion channel genes previously associated with atrial standstill, atrial fibrillation, or sick sinus syndrome: SCN5A, HCN4 (605206), and GJA5. No mutations were found in HCN4, but the proband and his asymptomatic father were heterozygous for a missense mutation in SCN5A (L212P; 600163.0048). In addition, the proband and his unaffected mother and maternal grandmother were all heterozygous for the same 2 rare GJA5 polymorphisms identified by Groenewegen et al. (2003) in atrial standstill patients, -44A/+71G. Functional analysis with the L212P mutant channels demonstrated large hyperpolarizing shifts in both the voltage dependence of activation and inactivation and delayed recovery from inactivation compared to wildtype. Makita et al. (2005) suggested that defects in SCN5A underlie atrial standstill, and that coinheritance of GJA5 polymorphisms represents a possible genetic modifier of the clinical manifestations.

Olson et al. (2005) analyzed the SCN5A gene in 156 unrelated patients with dilated cardiomyopathy who were negative for mutations in the known CMD genes encoding cardiac actin (102540), alpha-tropomyosin (191010), and metavinculin (see 193065), and identified 5 heterozygous mutations in 5 probands, respectively (see, e.g., 600163.0027, 600163.0038-600163.0039). All of the mutations altered highly conserved residues in the transmembrane domains of SCN5A.

Albert et al. (2008) analyzed 5 cardiac ion channel genes, SCN5A, KCNQ1, KCNH2, KCNE1, and KCNE2, in 113 cases of sudden cardiac death. No mutations or rare variants were identified in any of the 53 male subjects, but in 6 (10%) of 60 female subjects, 5 rare missense variants in SCN5A were identified, 2 previously associated with long QT syndrome, 1 with sudden infant death syndrome, and 2 not previously reported in control populations. Functional studies showed that all of the variants resulted in significantly shorter recovery times from inactivation. Albert et al. (2008) concluded that functionally significant mutations and rare variants in the SCN5A gene may contribute to the risk of sudden cardiac death in women.

Makita et al. (2008) genotyped 66 members of 44 LQT3 families of multiple ethnicities and identified the E1784K mutation (600163.0008) in 41 individuals from 15 (34%) of the kindreds; the diagnoses in these individuals included LQT3 syndrome, Brugada syndrome, and/or sinus node dysfunction (see 608567). In vitro functional characterization of E1784K channels compared to properties reported for other LQT3 variants suggested that a negative shift of steady-state Na channel inactivation and enhanced tonic block in response to Na channel blockers confer an additional Brugada syndrome/sinus node dysfunction phenotype, and further indicated that class IC drugs should be avoided in patients with Na channels displaying these behaviors.

In a large Finnish family with atrial fibrillation (AF) and conduction defects (ATFB10; 614022), Laitinen-Forsblom et al. (2006) analyzed the SCN5A gene and identified a heterozygous missense mutation (600163.0034) that segregated with disease and was not found in more than 370 control chromosomes.

Ellinor et al. (2008) analyzed the SCN5A gene in 57 probands with a familial history of isolated or 'lone' atrial fibrillation and identified heterozygosity for a missense mutation (600163.0041) in a 45-year-old male proband and his affected father. The authors concluded that SCN5A gene was not a major cause of familial AF.

Darbar et al. (2008) analyzed the SCN5A gene in 375 probands with AF, including 118 with lone AF, which was defined as AF occurring in individuals less than 65 years of age who did not have hypertension, overt structural heart disease, or thyroid dysfunction. The authors identified 8 heterozygous variants in 10 probands that were not found in 360 age-, sex-, and ethnicity-matched controls (see, e.g., 600163.0042-600163.0045). In addition, 11 previously reported rare nonsynonymous coding region variants were identified in 12 probands (see, e.g., 600163.0033), and 3 known common nonsynonymous SCN5A polymorphisms were also identified in the AF cohort (see, e.g., 600163.0024 and 600163.0031). Darbar et al. (2008) stated that in their study, nearly 6% of AF probands carried heterozygous mutations or rare variants in the SCN5A gene.

In affected members of 2 unrelated families with CMD and conduction system disease, Hershberger et al. (2008) identified heterozygosity for 2 different missense mutations in the SCN5A gene, R222Q (600163.0046) and I1835T (600163.0047), respectively. Cheng et al. (2010) restudied the 2 families, noting that all affected individuals were also either homozygous or heterozygous for the SCN5A common polymorphism, H558R (600163.0031). Whole-cell voltage clamp studies in HEK293 cells using the Q1077del background, which is the more abundant alternatively spliced SCN5A transcript present in human hearts (65%), showed that sodium current densities of the R222Q and I1835T mutants were not different from wildtype, but the combined variants R222Q/H558R and I1835T/H558R caused approximately 35% and 30% reduction, respectively, and each showed slower recovery from inactivation than wildtype. With the Q1077del background, R222Q and R222Q/H558R variants also exhibited a significant negative shift in both activation and inactivation, whereas I1835T/H558R showed a significant negative shift in inactivation that tended to decrease window current. In contrast, expression in the Q1077 background showed no changes in peak sodium current densities, decay, or recovery from inactivation for R222Q/H558R or I1835T/H558R. Cheng et al. (2010) concluded that CMD-associated SCN5A rare variants perturb the SCN5A biophysical phenotype that is modulated by SCN5A common variants.

In 3 unrelated families with multifocal ectopic Purkinje-related premature contractions and dilated cardiomyopathy, Laurent et al. (2012) identified heterozygosity for the R222Q mutation in the SCN5A gene, which was fully penetrant and strictly segregated with the cardiac phenotype in each family. Laurent et al. (2012) stated that the R222Q effects that they observed on channel parameters were similar to those measured by Cheng et al. (2010); in addition, they noted that the effects were intermediate in the heterozygous state and also impaired the window current, which is crucial during the plateau phase of the action potential. In vitro studies recapitulated the normalization of the ventricular action potentials in the presence of quinidine.

In affected members of a 3-generation Canadian family with CMD and junctional escape ventricular capture bigeminy, Nair et al. (2012) identified the R222Q mutation in the SCN5A gene. Heterologous expression studies revealed a unique biophysical phenotype of R222Q channels in which an approximately 10-mV leftward shift in the sodium current steady-state activation curve occurs without corresponding shifts in steady-state inactivation at cardiomyocyte resting membrane-potential voltages. Nair et al. (2012) noted that the absence of H558R in these patients established that the H558R polymorphism is not required for the induction of cardiomyopathy in patients carrying the R222Q mutation.

In 16 affected members over 3 generations of a large kindred with CMD and multiple arrhythmias, including premature ventricular complexes (PVCs) of variable morphologies, Mann et al. (2012) identified heterozygosity for the R222Q mutation in the SCN5A gene. The mutation was also identified in 1 clinically unaffected family member, a 56-year-old man with a normal EKG and echocardiogram. None of the R222Q carriers had the common SCN5A variant, H558R.

O'Neill et al. (2022) studied the effects of 50 previously published, functionally characterized missense variants in the SCN5A gene. Based on their effects on peak currents, variants were divided into loss-of-function (less than 10% of wildtype peak current, 35 variants) and partial loss-of-function (10-50% of wildtype peak current, 15 variants). Using cell lines created to study the effects of the variants in heterozygous coexpression with wildtype SCN5A, the authors found that 32 of 35 loss-of-function variants and 6 of 15 partial loss-of-function variants showed a reduction to less than 75% of wildtype-alone peak current, demonstrating evidence of dominant-negative effects. Using data from a published consortia and gnomAD, they found that patients with dominant-negative variants were 2.7 times more likely to present with Brugada syndrome than individuals with putative haploinsufficient variants (p = 0.019).

Associations Pending Confirmation

For discussion of a possible association between variants in the SCN5A, SCN10A (604427), and HEY2 (604674) genes and Brugada syndrome, see 601144.


Genotype/Phenotype Correlations

Westenskow et al. (2004) analyzed the KCNQ1, KCNH2, SCN5A, KCNE1, and KCNE2 genes in 252 probands with long QT syndrome and identified 19 with biallelic mutations in LQTS genes, of whom 18 were either compound (monogenic) or double (digenic) heterozygotes and 1 was a homozygote. They also identified 1 patient who had triallelic digenic mutations (see 152427.0021). Compared with probands who had 1 or no identified mutation, probands with 2 mutations had longer QTc intervals (p less than 0.001) and were 3.5-fold more likely to undergo cardiac arrest (p less than 0.01). All 20 probands with 2 mutations had experienced cardiac events. Westenskow et al. (2004) concluded that biallelic mono- or digenic mutations (which the authors termed 'compound mutations') cause a severe phenotype and are relatively common in long QT syndrome. The authors noted that these findings support the concept of arrhythmia risk as a multi-hit process and suggested that genotype can be used to predict risk.

Niu et al. (2006) analyzed the SCN5A gene in 17 members of a 4-generation Han Chinese family with apparent autosomal dominant inheritance of cardiac arrhythmias and sudden death. All affected individuals were heterozygous for a nonsense mutation in the SCN5A gene (W1421X; 600163.0036), and 1 unaffected individual was compound heterozygous for the W1421X mutation and R1193Q (600163.0023). Niu et al. (2006) suggested that the R1193Q mutation, which results in a gain of sodium channel function, may compensate for the deleterious effects of W1421X.


Animal Model

Nuyens et al. (2001) reported that mice heterozygous for a knockin KPQ deletion (600163.0001) of the Scn5a gene showed the essential features of LQT3 and spontaneously developed life-threatening polymorphous ventricular arrhythmias. Sudden accelerations in heart rate or premature beats caused lengthening of the action potential with early after-depolarization and triggered arrhythmias in mice heterozygous for the deletion. Adrenergic agonists normalized the response to rate acceleration in vitro and suppressed arrhythmias upon premature stimulation in vivo. These results showed the possible risk of sudden heart rate accelerations. The heterozygous knockin mouse with its predisposition for pacing-induced arrhythmia might be a useful model for the development of new treatments for the LQT3 syndrome.

Papadatos et al. (2002) showed that disruption of the mouse Scn5a gene caused intrauterine lethality in homozygotes with severe defects in ventricular morphogenesis, whereas heterozygotes showed normal survival. Whole-cell patch-clamp analyses of isolated ventricular myocytes from adult Scn5a +/- mice demonstrated a reduction of approximately 50% in sodium conductance. Scn5a +/- hearts had several defects, including impaired atrioventricular conduction, delayed intramyocardial conduction, increased ventricular refractoriness, and ventricular tachycardia with characteristics of reentrant excitation. These findings reconciled reduced activity of the cardiac sodium channel leading to slowed conduction with several apparently diverse clinical phenotypes, providing a model for the detailed analysis of the pathophysiology of arrhythmias. Noble (2002) commented that detailed understanding of the mechanisms of cardiac arrhythmia at all relevant levels is important to the design of therapeutic programs, and cited the work of Papadatos et al. (2002) as an important step.


ALLELIC VARIANTS ( 48 Selected Examples):

.0001 LONG QT SYNDROME 3

SCN5A, 9-BP DEL, NT4661
  
RCV000009962...

In 2 apparently unrelated kindreds with chromosome 3-linked LQT syndrome (LQT3; 603830), Wang et al. (1995) found deletion of 9 basepairs beginning at nucleotide 4661 of their cDNA for SCN5A. The deletion, which was detected by sequencing an aberrant SSCP conformer, resulted in deletion of lys-pro-gln (KPQ), which are 3 conserved amino acids in the cytoplasmic linker between domains III and IV of the channel protein. The 3 amino acids involved in the in-frame deletion are lys1505, pro1506, and gln1507. The effect of this mutation on membrane depolarization was studied by Bennett et al. (1995).

Clancy and Rudy (1999) developed a model representative of the behavior of the sodium channel in heart muscle cells using a single-channel-based Markov model approach. They showed that the delta-KPQ mutant form of the sodium channel stays open for too long, causing an overlarge inward current of sodium which gives rise to arrhythmia. This model view was corroborated by experiments recording actual sodium currents in cardiac muscle cells.


.0002 LONG QT SYNDROME 3

SCN5A, ARG1644HIS
  
RCV000009963...

In a mother and son with the long QT syndrome (LQT3; 603830), Wang et al. (1995) demonstrated a CGC-to-CAC mutation in codon 1644, resulting in the substitution of a highly conserved arginine residue by histidine.


.0003 LONG QT SYNDROME 3

SCN5A, ASN1325SER
  
RCV000009964...

In a family in which members of 4 generations had been affected by the long QT syndrome (LQT3; 603830), Wang et al. (1995) found an AAT-to-AGT transition in codon 1325, predicted to cause substitution of a highly conserved asparagine residue by a serine residue.


.0004 BRUGADA SYNDROME 1

SCN5A, ARG1232TRP AND THR1620MET
  
RCV000009965...

In affected members of a family with Brugada syndrome (BRGDA1; 601144), a distinct form of idiopathic ventricular fibrillation, Chen et al. (1998) found an arg1232-to-trp (R1232W) and a thr1620-to-met (T1620M) mutation on the same chromosome with no mutation in the other chromosome, suggesting to them that IVF in this family was inherited as an autosomal dominant trait. The presence of both normal and mutated sodium channels in the same tissue would promote heterogeneity of the refractory period, a well established mechanism in arrhythmogenesis, and therefore may be the underlying molecular defect that causes re-entrant arrhythmia in this family. The potential contribution of R1232W and T1620M mutations to the mechanism of IVF was determined by heterologous expression in Xenopus oocytes. They found that sodium channels with the missense mutation recovered from inactivation more rapidly than normal, indicating that IVF with right bundle branch block (RBBB) and ST segment elevation is a defect distinct from long QT syndrome. When studied alone, the R1232W mutant behaved most like normal channels, whereas the T1620M mutant closely followed the kinetic pattern of the double mutant. This indicated that T1620M is the mutation probably responsible for the IVF phenotype in this kindred and that R1232W could be a rare polymorphism. In summary, biophysical analysis of the 2 missense mutations in SCN5A showed a shift in the voltage dependence of steady-state inactivation toward more positive potentials associated with a 25 to 30% acceleration in recovery time from inactivation at potentials near -80mV.

Commenting that studies of the thr1620-to-met mutant by Chen et al. (1998) revealed an abnormal electrophysiologic profile at room temperature that did not adequately explain the ECG signature of Brugada syndrome, Dumaine et al. (1999) undertook a more detailed electrophysiologic study of the thr1620-to-met mutant protein. Dumaine et al. (1999) expressed the mutant protein in a mammalian cell line and employed a patch-clamp technique to study current kinetics at 32 degrees C. The results indicated that current decay kinetics were faster in mutant than in wildtype channels at this temperature and that recovery from inactivation was slower, with a significant shift in steady-state activation. These findings provided an explanation for the ECG features of Brugada syndrome and represented the first illustration of a cardiac sodium channel mutation in which arrhythmogenicity is revealed only at temperatures approaching the physiologic range.

Voltage-gated sodium channels are multimeric structures consisting of a large, heavily glycosylated alpha subunit and 1 or 2 smaller beta subunits. The beta subunits are thought necessary for normal gating function. In brain and skeletal muscle, the beta-1 subunit (600235) accelerates sodium channel inactivation. Makita et al. (2000) characterized the functional roles of the auxiliary beta subunit by coexpression of the beta subunit with either wildtype SCN5A or SCN5A carrying the heterologously expressed T1620M mutation in Xenopus oocytes. The midpoint of steady-state inactivation was significantly shifted to positive potentials in the T1620M alpha/beta-1 channel, with an acceleration in recovery from inactivation when compared to other channels. Makita et al. (2000) therefore suggested that coexpression of T1620M alpha/beta-1 subunits exposed a significant electrophysiologic deficit that may predispose to ventricular fibrillation. Expression of both normal and mutant channels, as in the hearts of patients with Brugada syndrome, would promote heterogeneity of the refractory period in their myocardium, which serves as an ideal electrical substrate for reentrant arrhythmia.


.0005 BRUGADA SYNDROME 1

SCN5A, IVS7DS, 2-BP INS
  
RCV000009966

In affected members of a family with idiopathic ventricular fibrillation with right bundle branch block (RBBB) and elevated ST segments, a disorder known as Brugada syndrome (BRGDA1; 601144), Chen et al. (1998) found an insertion of 2 nucleotides, AA, after the first 4 nucleotides (gtaa) in the splice donor sequence of intron 7 of the SCN5A gene. The functional consequences of this splicing mutation were not established.


.0006 BRUGADA SYNDROME 1

SCN5A, 1-BP DEL, VAL1398TER
  
RCV000009967...

In affected members of a family with idiopathic ventricular fibrillation characterized by RBBB and elevated ST segments, a disorder known as Brugada syndrome (BRGDA1; 601144), Chen et al. (1998) found a deletion of a single nucleotide (A) from codon 1397 of the SCN5A gene. This deletion resulted in an in-frame stop at codon 1398 (normally val). The resulting truncation eliminated DIII/S6, DIV/S1-S6, and the C-terminal portion of the cardiac sodium channel.


.0007 LONG QT SYNDROME 3

LONG QT SYNDROME 3/6, DIGENIC, INCLUDED
SCN5A, ARG1623GLN
  
RCV000009970...

In an infant Japanese girl with a severe form of long QT syndrome (LQT3; 603830), Makita et al. (1998) identified a de novo missense mutation, arg1623 to gln (R1623Q), in the S4 segment of domain 4 of the SCN5A gene. When expressed in oocytes, mutant sodium channels exhibited only minor abnormalities in channel activation, but in contrast to 3 previously characterized LQT3 mutations, had significantly delayed macroscopic inactivation. Single channel analysis revealed that R1623Q channels had significantly prolonged open times with bursting behavior, suggesting a novel mechanism of pathophysiology in Na(+) channel-linked long QT syndrome.

Kambouris et al. (2000) reported that the R1623Q mutation imparts unusual lidocaine sensitivity to the sodium channel that is attributable to its altered functional behavior. Studies of lidocaine on individual R1623Q single-channel openings indicated that the open-time distribution was not changed, indicating the drug does not block the open pore as proposed previously. Rather, the mutant channels have a propensity to inactivate without ever opening ('closed-state inactivation'), and lidocaine augments this gating behavior. An allosteric gating model incorporating closed-state inactivation recapitulated the effects of lidocaine on the pathologic sodium current. These findings explained the unusual drug sensitivity of R1623Q and provided a general and unanticipated mechanism for understanding how sodium channel-blocking agents may suppress the pathologic, sustained sodium current induced by LQT3 mutations.

In a male infant diagnosed with ventricular arrhythmias and cardiac decompensation in utero at 28 weeks' gestation and with long QT syndrome at birth, Miller et al. (2004) identified heterozygosity for the R1623Q mutation. The mother had no ECG abnormalities, but a previous and a subsequent pregnancy both ended in stillbirth at 7 months. Initial studies detected no genetic abnormality, but a sensitive restriction enzyme-based assay revealed a small percentage (8 to 10%) of cells harboring the mutation in the mother's blood, skin, and buccal mucosa; R1623Q was also identified in cord blood from the third fetus. Miller et al. (2004) concluded that recurrent late-term fetal loss or sudden infant death can result from unsuspected parental mosaicism for LQT-associated mutations.

In a 1-month-old male infant who had syncope, torsade de pointes, cardiac arrest, and a QTc of 460 ms, Millat et al. (2006) identified biallelic digenic mutations: a 4868G-A transition in exon 28 of the SCN5A gene resulting in the R1623Q substitution; and a missense mutation in the KCNE2 gene (F60L; 603796.0005).


.0008 LONG QT SYNDROME 3

BRUGADA SYNDROME 1, INCLUDED
SINUS NODE DISEASE, INCLUDED
SCN5A, GLU1784LYS
  
RCV000009972...

Wei et al. (1999) described a family in which the 13-year-old proband died suddenly at rest with no antecedent illness and no significant findings at postmortem. Her father had sinus bradycardia with occasional sinus pauses and ventricular ectopy together with profound prolongation of his QT interval (QTc = 527 ms) (see LQT3; 603830). He experienced only occasional light-headedness. Other family members experienced occasional syncope and had sinus bradycardia and prolonged QT intervals on their ECGs. In those individuals with prolonged QT intervals, SSCP analysis detected an aberrant conformer in the coding region of the SCN5A gene corresponding to the C terminus. Nucleotide sequencing revealed a G-to-A transition at codon 1784, resulting in a glu-to-lys substitution. This mutation occurs at a highly conserved residue in most voltage-gated sodium channels in most animals, including invertebrates. When the mutation was expressed in Xenopus oocytes, a defect in channel inactivation was demonstrated in the form of a small residual steady state current throughout prolonged depolarization. Wei et al. (1999) explored this further by engineering SCN5A constructs with amino acid substitutions at other positions in the C terminus. All exhibited similar electrophysiologic phenotypes, suggesting that heterozygous charge-neutralizing amino acid substitution at this site causes an allosteric effect on sodium channel gating, resulting in delayed myocardial repolarization. This provided a novel mechanism for LQT3.

Makita et al. (2008) genotyped 66 members of 44 LQT3 families of multiple ethnicities and identified the E1784K mutation in 41 individuals from 15 (34%) of the kindreds, including the family previously reported by Wei et al. (1999); the diagnoses in these individuals included LQT3 syndrome, Brugada syndrome (BRGDA1; 601144), and/or sinus node disease (see 608567). Heterologously expressed E1784K channels showed a 15.0-mV negative shift in the voltage dependence of Na channel inactivation and a 7.5-fold increase in flecainide affinity for resting-state channels, properties also seen with other LQT3 mutations associated with a mixed clinical phenotype. Furthermore, these properties were absent in Na channels harboring the T1304M mutation, which is associated with LQT3 without a mixed clinical phenotype. Makita et al. (2008) suggested that a negative shift of steady-state Na channel inactivation and enhanced tonic block by class IC drugs represent common biophysical mechanisms underlying the phenotypic overlap of LQT3 and Brugada syndromes, and further indicated that class IC drugs should be avoided in patients with Na channels displaying these behaviors.


.0009 PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, IVS22DS, T-C, +2
  
RCV000009975...

In a large French family with progressive heart block (PFHB1A; 113900), Schott et al. (1999) identified a T-to-C transition in the highly conserved +2 donor splice site of intron 22 of the SCN5A gene. The abnormal transcript predicted in-frame skipping of exon 22 and an impaired gene product lacking the voltage-sensitive DIIIS4 segment.


.0010 HEART BLOCK, NONPROGRESSIVE

SCN5A, 1-BP DEL, 5280G
  
RCV000009976

In a Dutch family with asymptomatic first-degree atrioventricular block associated with right bundle branch block from birth, without apparent progression (see 113900), Schott et al. (1999) identified a 1-bp deletion (G) at nucleotide 5280 of the SCN5A gene, resulting in a frameshift predicted to cause a premature stop codon.


.0011 BRUGADA SYNDROME 1

SCN5A, ARG1512TRP
  
RCV000009977...

In the screening of SCN5A in 6 individuals with Brugada syndrome (BRGDA1; 601144), Rook et al. (1999) found missense mutations in the coding region of the gene in 2: arg1512 to trp (R1512W) in the DIII-DIV cytoplasmic linker, and ala1924 to thr (A1924T; 600163.0012) in the C-terminal cytoplasmic domain. In 2 other patients mutations were detected near intron/exon junctions. To assess the functional consequences of the R1512W and A1924T mutations, wildtype and mutant sodium channel proteins were expressed in Xenopus oocytes. Both missense mutations affected channel function and seemed to be associated with an increase in inward sodium current during the action potential upstroke.


.0012 BRUGADA SYNDROME 1

SCN5A, ALA1924THR
  
RCV000009978...

For discussion of the ala1924-to-thr (A1924T) substitution in the SCN5A gene that was found in compound heterozygous state in 2 patients with Brugada syndrome (BRGDA1; 601144) by Rook et al. (1999), see 600163.0011.


.0013 LONG QT SYNDROME 3

BRUGADA SYNDROME 1, INCLUDED
SCN5A, 3-BP INS, 5537TGA
  
RCV000009979...

In a large Dutch family with electrocardiographic features both of long QT syndrome (LQT3; 603830) and Brugada syndrome (BRGDA1; 601144), Bezzina et al. (1999) demonstrated a 3-bp insertion at nucleotide position 5537 of the SCN5A gene, predicted to cause insertion of an aspartic acid residue at amino acid position 1795 (1795insD) in the C-terminal domain of the protein. Expression of this mutant channel protein in Xenopus oocytes permitted characterization of defects in channel activation and inactivation when compared to a wildtype control. These defects were predicted to cause a reduction in sodium flux during the upstroke of the cardiac action potential.

The co-occurrence of Brugada syndrome and long QT syndrome in this family was paradoxical, since LQT3 is associated with activating SCN5A mutations and Brugada syndrome with inactivating mutations. Clancy and Rudy (2002) modeled the cellular effects of the 1795insD mutation in a virtual transgenic cell. Since ion channel proteins are expressed nonuniformly throughout the myocardium, there is an intrinsic electrophysiologic heterogeneity. The authors demonstrated that the interplay between this underlying myocardial electrophysiologic heterogeneity and the mutation-induced changes in cardiac sodium channel function provided the substrate for both ST segment elevation (in Brugada syndrome) and QT prolongation (LQT3) in a rate-dependent manner.


.0014 VENTRICULAR FIBRILLATION, PAROXYSMAL FAMILIAL, 1 (1 patient)

SCN5A, SER1710LEU
  
RCV000009981...

Akai et al. (2000) screened 25 Japanese patients with idiopathic ventricular fibrillation (VF1; 603829). The diagnosis was based on the occurrence of at least one episode of syncope and/or cardiac arrest and documentation of ventricular fibrillation. Structural heart disorders were excluded. Eighteen patients were diagnosed as Brugada syndrome. The authors identified a heterozygous ser1710-to-leu missense mutation of the SCN5A gene in a 39-year-old man who was admitted to the hospital for recurrent syncope and suffered an episode of spontaneous ventricular fibrillation while hospitalized. An implanted cardiac defibrillator was successful in preventing further attacks of palpitation or syncope. Brugada syndrome was not present. The paternal grandfather and a paternal uncle had died suddenly in their sixth decade of unknown cause; the parents and sibs were asymptomatic.


.0015 LONG QT SYNDROME 3

SCN5A, SER941ASN
  
RCV000009982...

Schwartz et al. (2000) described an infant who nearly died of SIDS (272120), whose parents had normal QT intervals and in whom the long QT syndrome (LQT3; 603830) was diagnosed with identification of a spontaneous mutation of the SCN5A gene: a change of codon 941 from TCC (serine) to AAC (asparagine). The patient had all the classic features of near-SIDS. Before the episode, the infant appeared to be in perfect health. His age at the time of the episode (7 weeks) was within the age range of 5 to 12 weeks during which the incidence of SIDS peaks. The parents found him cyanotic, apneic, and pulseless. Ventricular fibrillation was documented in an emergency room; this point is important given the frequent statements that ventricular arrhythmias have not been recorded in infants at risk for SIDS. Had the infant died--an outcome that was almost a certainty in the absence of cardioversion--the absence of an electrocardiogram and the normal QT intervals of both parents would have eliminated suspicion of the long QT syndrome and would have prompted a diagnosis of SIDS.


.0016 CARDIAC CONDUCTION DEFECT, NONPROGRESSIVE

SCN5A, GLY514CYS
  
RCV000009984...

Tan et al. (2001) studied a family who came to medical attention when the proband, a 3-year-old girl, experienced episodes of fainting during a febrile illness. Her 12-lead ECG showed characteristics of slow conduction throughout the atria and ventricles, including broad P waves, PR interval prolongation, and a wide QRS complex (see 113900). Continuous monitoring revealed episodes of severe bradycardia (25 beats/minute). During these slow periods the cardiac rhythm was maintained by infrequent atrioventricular nodal 'escape' impulses. Conduction disturbance persisted after the febrile illness, but there was no evidence of structural heart disease or systemic diseases associated with conduction defects in children. Therapeutic intervention with a dual-chamber pacemaker was initially limited by inability to pace the atrium (maximal stimulus: 10 V, 1 ms); however, this difficulty resolved with 1 week of empiric steroid treatment. During the 4 years following diagnosis, the patient continuously required dual-chamber pacing. The proband's 6-year-old sister was similarly affected and required pacemaker implantation, with episodes of noncapture that reproducibly resolved with corticosteroid therapy. Three other family members with no structural heart disease had ECG evidence of conduction slowing (prolonged PR and QRS intervals), but did not experience bradycardia or require pacemaker implantation. All affected family members had a G-to-T transition in the first nucleotide of codon 514 in exon 12 of the SCN5A gene resulting in the replacement of glycine by cysteine (G514C). Biophysical characterization of the mutant channel showed that there were abnormalities in voltage-dependent gating behavior that could be partially corrected by dexamethasone, consistent with the salutary effects of glucocorticoids on the clinical phenotype. Computational analysis predicts that the gating defects of G514C selectively slow myocardial conduction, but do not provoke the rapid cardiac arrhythmias associated previously with SCN5A mutations.


.0017 PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, ASP1595ASN
  
RCV000009983...

Wang et al. (2002) reported a family in which the proband had presented with first-degree atrioventricular block at the age of 9, progressing to complete AV block by the age of 20 (PFHB1A; 113900). The proband's sister and father had electrocardiographic evidence of right bundle branch block and left axis deviation with normal PR intervals. The corrected QT interval was normal (less than 420 ms) in all 3 individuals. Sequencing of the coding region of SCN5A revealed a G-to-A mutation at nucleotide position 4783, which replaced an aspartic acid residue at amino acid position 1595 with asparagine (D1595N). The G4783A mutation was engineered into a recombinant human heart sodium channel and transiently coexpressed with human sodium channel beta-1 subunit (600760) in a cultured mammalian cell line (tsA201). Functional characterization using a patch-clamp technique revealed a significant defect in the kinetics of fast-channel inactivation distinct from those of SCN5A mutations reported in LQT3 (603830). The authors considered this a plausible mechanism for the observed conduction system disease in this family.


.0018 PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, GLN298SER
  
RCV000009985...

Wang et al. (2002) reported a child in whom second-degree atrioventricular block had been diagnosed at the age of 6, progressing to complete atrioventricular block by the age of 12 (PFHB1A; 113900). The child's mother had a normal electrocardiogram and the father declined testing. There was no family history of sudden death. Sequencing of the coding region of SCN5A revealed a G-to-A mutation at nucleotide position 892 that replaced a glycine residue at amino acid position 298 with serine (G298S). The G892A mutation was engineered into a recombinant human heart sodium channel and transiently coexpressed with human sodium channel beta-1 subunit (600760) in a cultured mammalian cell line (tsA201). Functional characterization using a patch-clamp technique revealed a significant defect in the kinetics of fast-channel inactivation distinct from those of SCN5A mutations reported in LQT3 (603830). The authors considered this a plausible mechanism for the observed conduction system disease in this family.


.0019 LONG QT SYNDROME 3

SCN5A, ALA997SER
  
RCV000009986...

In a 6-week-old male infant who died of SIDS (272120), Ackerman et al. (2001) identified a heterozygous G-to-T transversion in the SCN5A gene, resulting in an ala997-to-ser substitution. The mutation was not detected in 800 control alleles. Ackerman et al. (2001) determined that amino acid 997 is located in the cytoplasmic connector between the second and third domains of the sodium channel and is highly conserved across species. They demonstrated that the mutant SCN5A channel expressed a sodium current characterized by slower decay and a 2- to 3-fold increase in late sodium current.


.0020 LONG QT SYNDROME 3

SCN5A, ARG1826HIS
  
RCV000009987...

In a 42-day-old male infant who died of possible SIDS (272120), Ackerman et al. (2001) identified a heterozygous G-to-A replacement in the SCN5A gene, resulting in an arg1826-to-his substitution. The mutation was not detected in 800 control alleles. Ackerman et al. (2001) determined that amino acid 1826 is located in the cytoplasmic C-terminal region of the sodium channel and is highly conserved. They demonstrated that the SCN5A mutant channel expressed a sodium current characterized by slower decay and a 2- to 3-fold increase in late sodium current.


.0021 BRUGADA SYNDROME 1

SCN5A, ARG367HIS
  
RCV000009988...

Sudden unexplained nocturnal death syndrome (SUNDS), a disorder found in southeast Asia, is characterized by an abnormal electrocardiogram with ST segment elevation in leads V1 to V3 and sudden death due to ventricular fibrillation, identical to that seen in Brugada syndrome (BRGDA1; 601144). Vatta et al. (2002) found mutations in the SCN5A gene in 3 of 10 Asian SUNDS patients. In a sporadic Asian SUNDS patient, the authors identified a 1100G-A transition in SCN5A. The mutation is predicted to result in an arg367-to-his (R367H) substitution, which lies in the first P segment of the pore-lining region between the DIS5 and DIS6 transmembrane segments. In transfected Xenopus oocytes, the R367H mutant channel did not express any current. The authors hypothesized that the likely effect of this mutation is to depress peak current due to the loss of one functional allele.


.0022 BRUGADA SYNDROME 1

SCN5A, ALA735VAL
  
RCV000009989...

In a family with SUNDS, a disorder identical to Brugada syndrome (BRGDA1; 601144), that exhibited autosomal dominant inheritance, Vatta et al. (2002) identified among affected members a 2204C-T transition, which is predicted to result in an ala735-to-val (A735V) substitution. The mutation lies in the first transmembrane segment of domain II, (DIIS1), close to the first extracellular loop between DIIS1 and DIIS2. In transfected Xenopus oocytes, the A735V mutant expressed currents with steady-state activation voltage shifted to more positive potentials and exhibited reduced sodium channel current at the end of phase I of the action potential.


.0023 BRUGADA SYNDROME 1

LONG QT SYNDROME 3, ACQUIRED, SUSCEPTIBILITY TO, INCLUDED
SCN5A, ARG1193GLN
  
RCV000009990...

In a pair of Japanese dizygotic twins, one of whom died at 4 months of SUNDS, a disorder identical to Brugada syndrome (BRGDA1; 601144), Vatta et al. (2002) identified a 3575G-A transition in exon 20 of the SCN5A gene, predicted to result in an arg1192-to-gln (R1192Q) substitution in Domain III. In transfected Xenopus oocytes, the mutation accelerated the inactivation of the sodium channel current and exhibited reduced sodium channel current at the end of phase I of the action potential. Wang (2005) stated that this variant was mislabeled in the Vatta et al. (2002) report and should be designated R1993Q.

In an 82-year-old Caucasian male who developed long QT syndrome after the administration of D-sotolol or quinidine (see LQT3, 603830), Wang et al. (2004) identified heterozygosity for the R1993Q mutation in the SCN5A gene. The mutation was found in 4 of 2,087 predominantly Caucasian controls (0.2%). Electrophysiologic studies showed that mutant R1193Q channels destabilize inactivation gating and generate a persistent, nonactivating current that is expected to prolong the cardiac action potential duration, leading to LQT syndrome; single channel recording revealed the molecular mechanism to be frequent, dispersed reopening of the channels. The patient also carried the H558R SCN5A variant (600163.0031), but due to a lack of family members, it could not be determined whether H558R was in cis or trans with R1993Q.

Hwang et al. (2005) found the R1993Q mutation in 11 of 94 (12%) randomly selected Han Chinese individuals and concluded that the variant is a common polymorphism in this population. None of the carriers had electrocardiographic signs of Brugada syndrome, although 1 had a prolonged QTc interval (472 ms) and another, who was homozygous for the mutation, had a borderline long QTc (437 ms).

In an asymptomatic 73-year-old male member of a 4-generation Han Chinese family with autosomal dominant cardiac arrhythmias and sudden death, Niu et al. (2006) identified compound heterozygosity for R1193Q and a nonsense mutation in the SCN5A gene (W1421X; 600163.0036). Niu et al. (2006) suggested that the R1193Q mutation, which results in a gain of sodium channel function, may compensate for the deleterious effects of W1421X. Haplotype analysis of an asymptomatic daughter-in-law and 2 asymptomatic grandchildren who also carried the R1193Q mutation revealed that the children inherited the mutation from their mother rather than their grandfather.


.0024 LONG QT SYNDROME 3, ACQUIRED, SUSCEPTIBILITY TO

SUDDEN INFANT DEATH SYNDROME, INCLUDED
SCN5A, SER1103TYR
  
RCV000009992...

Splawski et al. (2002) screened DNA samples from individuals with nonfamilial cardiac arrhythmias and identified a C-to-A transversion in the SCN5A gene leading to a ser1103-to-tyr (S1103Y) substitution 1 patient. Ackerman et al. (2004) noted that the variant was originally published as SER1102TYR from numbering based on the 2,015 amino acid alternatively spliced transcript. Subsequently, numbering was revised to account for the full-length 2,016 amino acid transcript. Serine-1103 is a conserved residue located in the intracellular sequences that link domains II and III of the channel. The proband had idiopathic dilated cardiomyopathy and hypokalemia and developed prolonged QT and torsade de pointes ventricular tachycardia while on amiodarone. Splawski et al. (2002) determined that the Y1103 allele is present in 19.2% of West Africans and Caribbeans and in 13.2% of African Americans. The Y1103 allele was not found in 511 Caucasians or 578 Asians. Splawski et al. (2002) studied 22 African Americans with acquired arrhythmia and 100 population-matched controls. The Y1103 allele was overrepresented among arrhythmia patients, being found in 56.5% of cases and among 13% of controls. The likelihood of displaying signs of arrhythmia in a Y1103 carrier heterozygote or homozygote yielded an odds ratio of 8.7 (95% CI 3.2 to 23.9). The odds ratio was not significantly altered after controlling for age or gender. To determine whether this mutation is an inherited risk factor for arrhythmias, Splawski et al. (2002) examined the extended family of 1 proband. They ascertained and phenotypically characterized 23 members of this kindred. Phenotypic analysis revealed that 11 members of the family had prolonged QT and/or a history of syncope. All 11 phenotypically affected members of this family carried the Y1103 allele (6 were homozygotes and 5 were heterozygotes). Physiologic analysis of the effect of this mutation recorded a small but significant negative shift in the voltage dependence of activation. Splawski et al. (2002) concluded that the Y1103 allele is a common SCN5A variant in Africans and African Americans and causes a small but inherent chronic risk of acquired arrhythmia. In the setting of additional acquired risk factors, including medications, hypokalemia, or structural heart disease, individuals carrying this allele are at increased risk of arrhythmia.

In 3 white sisters and their father, Chen et al. (2002) identified the S1103Y mutation, thus demonstrating that this mutation does exist in the white population. The mutation was associated with a considerable risk of syncope, ventricular arrhythmia, ventricular fibrillation, and sudden death, Each of the 3 sibs was genotyped for 31 'ancestry informative markers' to provide an estimation of biogeographic ancestry on 3 axes: Native American, West African, and European. The maximum likelihood point estimates for each of the sibs were 100% European, 0.0% African, and 0.0% Native American. The proband had a baseline QTc of 520 ms, and developed 2 episodes of syncope at age 49 years. The first episode was triggered by emotion and excitement. The second episode occurred in the setting of amiodarone and low serum potassium, and progressed to ventricular fibrillation and cardiac arrest. She was resuscitated by cardioversion. The second sister had a QTc of 431 ms, and died suddenly at age 44 years when awakening from sleep. The third sister had a QTc of 452 ms, developed 1 episode of syncope at the age of 33 years, and had complained of palpitations all her life. The father died suddenly in his sleep at age 50 years. Family members without S1103Y had a normal QTc.

Plant et al. (2006) screened DNA samples from 133 African American autopsy-confirmed cases of sudden infant death syndrome (SIDS; 272120) and identified 3 that were homozygous for the S1103Y variant. Among 1,056 African American controls, 120 were carriers of the heterozygous genotype, suggesting that infants with 2 copies of S1103Y have a 24-fold increased risk for SIDS. Variant Y1103 channels were found to operate normally under baseline conditions in vitro. Because risk factors for SIDS include apnea and respiratory acidosis, Y1103 and wildtype channels were subjected to lowered intracellular pH; only Y1103 channels developed abnormal function, with late reopenings suppressible by the drug mexiletine. Plant et al. (2006) suggested that the Y1103 variant confers susceptibility to acidosis-induced arrhythmia, a gene-environment interaction.

Darbar et al. (2008) stated that the S1103Y variant was a known common nonsynonymous polymorphism in the SCN5A gene; they detected S1103Y in 1 patient with lone atrial fibrillation and in 5 patients with atrial fibrillation associated with other heart disease, as well as in 15 of 720 control chromosomes, for a minor allele frequency of 0.7%.


.0025 SICK SINUS SYNDROME 1

SCN5A, PRO1298LEU
  
RCV000009994...

In 3 sibs with congenital sick sinus syndrome (SSS1; 608567), Benson et al. (2003) identified compound heterozygosity for 2 mutations in the SCN5A gene. The maternal allele carried a 3893C-T transition, resulting in a pro1298-to-leu (P1298L) change; the paternal allele carried a gly1408-to-arg substitution (600163.0026).


.0026 SICK SINUS SYNDROME 1

BRUGADA SYNDROME 1, INCLUDED
CARDIAC CONDUCTION DEFECT, NONSPECIFIC, INCLUDED
SCN5A, GLY1408ARG
  
RCV000009995...

In 3 sibs with congenital sick sinus syndrome (SSS1; 608567) with compound heterozygosity for mutation in the SCN5A gene, Benson et al. (2003) found on the paternal allele a 4222G-A transition, resulting in a gly1408-to-arg substitution (G1408R). The maternal allele carried a pro1298-to-leu substitution (600163.0025).

Kyndt et al. (2001) reported the G1408R mutation, which they designated GLY1406ARG, in heterozygous state in a large French family segregating both isolated cardiac conduction defect (see 601144) and Brugada syndrome (BRGDA1; 601144).


.0027 SICK SINUS SYNDROME 1

CARDIOMYOPATHY, DILATED, 1E, INCLUDED
SCN5A, THR220ILE
  
RCV000009998...

In a child with congenital sick sinus syndrome (SSS1; 608567), Benson et al. (2003) identified compound heterozygosity for 2 mutations in the SCN5A gene: the paternal allele carried a 659C-T transition, resulting in a thr220-to-ile (T220I) mutation, and the maternal allele carried a 4867C-T transition, resulting in an arg1623-to-ter mutation (R1623X; 600163.0028). The authors noted that an R1623Q mutation (600163.0007) resulting in congenital long QT syndrome-3 (603830) had previously been described.

In a 54-year-old man with dilated cardiomyopathy (CMD1E; 601154), atrial fibrillation, and heart block, Olson et al. (2005) identified heterozygosity for a 659C-T transition in exon 6 of the SCN5A gene, resulting in a thr220-to-ile (T220I) substitution at a highly conserved residue in the transmembrane domain. Coronary artery disease was excluded by angiography; cardiac biopsy showed moderate myocyte hypertrophy and marked interstitial fibrosis. He died 13 years later in severe congestive heart failure. A female first cousin once removed who also carried the mutation was diagnosed at 55 years of age with dilated cardiomyopathy (ejection fraction, 10%) and incomplete bundle branch block; she died 2 years later, also in severe congestive heart failure. Other relatives were reported to have enlarged hearts, but were unavailable for evaluation.


.0028 SICK SINUS SYNDROME 1

SCN5A, ARG1623TER
  
RCV000009968...

For discussion of the arg1623-to-ter (R1623X) mutation that was found in compound heterozygous state in a child with congenital sick sinus syndrome (SSS1; 608567) by Benson et al. (2003), see 600163.0027.


.0029 LONG QT SYNDROME 3

SCN5A, TYR1795CYS
  
RCV000009969...

In a patient with long QT syndrome-3 (LQT3; 603830), Rivolta et al. (2001) identified a tyr1795-to-cys (Y1795C) mutation in the SCN5A gene. The mutation slowed the onset of activation, but did not cause a marked negative shift in the voltage dependence of inactivation or affect the kinetics of the recovery from inactivation. The mutation increased the expression of sustained Na(+) channel activity compared with wildtype channels and promoted entrance into an intermediate or slowly developing inactivated state.


.0030 BRUGADA SYNDROME 1

SCN5A, TYR1795HIS
  
RCV000009999...

In a patient with Brugada syndrome (BRGDA1; 601144), Rivolta et al. (2001) identified a tyr1795-to-his (Y1795H) mutation in the SCN5A gene. The mutation accelerated the onset of activation and caused a marked negative shift in the voltage dependence of inactivation. It did not affect the kinetics of the recovery from inactivation. The mutation increased the expression of sustained Na(+) channel activity compared with wildtype channels and promoted entrance into an intermediate or slowly developing inactivated state.


.0031 PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, THR512ILE AND HIS558ARG
  
RCV000010000...

In a 2-year-old boy with second-degree atrioventricular conduction block (PFHB1A; 113900) necessitating a pacemaker, Viswanathan et al. (2003) identified a heterozygous 1535C-T transition in the SCN5A gene, resulting in a thr512-to-ile (T512I) substitution. In addition, there was a homozygous 1673A-G transition, resulting in a his558-to-arg (H558R) substitution. H558R (rs1805124) is a polymorphism present in 20% of the population (Yang et al., 2002). One of the patient's alleles contained both T512I and H558R. The patient's father was heterozygous for the H558R substitution, the asymptomatic mother was compound heterozygous for the T512I and H558R substitutions, and 2 sibs were heterozygous for the H558R substitution. Functional expression studies showed that activation and inactivation of wildtype and H558R channels were similar. By contrast, voltage-dependent activation and inactivation of the T512I channel was shifted negatively by 8 to 9 mV and had enhanced slow activation and slower recovery from inactivation compared to the wildtype channel. Studies of the double H558R/T512I channel showed that H558R eliminated the negative shift induced by T512I, but only partially restored the kinetic abnormalities. Viswanathan et al. (2003) suggested that enhanced slow inactivation disproportionately affected Purkinje cells, which have a longer action potential duration and smaller diastolic interval, resulting in slowed atrioventricular conduction.

Darbar et al. (2008) stated that the H558R variant was a known common nonsynonymous polymorphism in the SCN5A gene; they detected H558R in 59 patients with lone atrial fibrillation and in 130 patients with atrial fibrillation associated with other heart disease, as well as in 128 of 720 control chromosomes, for a minor allele frequency of approximately 25%.


.0032 BRUGADA SYNDROME 1

SCN5A, GLY1262SER
  
RCV000010001...

Shin et al. (2004) studied a family with 9 members as well as 12 unrelated sporadic cases, all Koreans, diagnosed with Brugada syndrome (BRGDA1; 601144). They identified a novel missense mutation associated with Brugada syndrome in the family: a single-nucleotide substitution of G to A at nucleotide position 3934 in exon 21 of the SCN5A gene that changed glycine-1262 to serine (G1262S) in segment 2 of domain III of the SCN5A protein. Four individuals in the family carried the identical mutation, but none of the 12 sporadic patients did. The mutation was not found in 150 unrelated normal individuals.


.0033 BRUGADA SYNDROME 1

ATRIAL FIBRILLATION, FAMILIAL, 10, INCLUDED
SCN5A, GLU1053LYS
  
RCV000010002...

In a patient with Brugada syndrome (BRGDA1; 601144), Mohler et al. (2004) identified a 3157G-A transition in the SCN5A gene resulting in a glu1053-to-lys (E1053K) mutation in the ankyrin-binding motif of the cardiac sodium channel. The mutation abolished binding of Na(v)1.5 to ankyrin-G (600465) and also prevented accumulation of Na(v)1.5 at cell surface sites in ventricular cardiomyocytes.

In a patient with lone atrial fibrillation (ATFB10; 614022), Darbar et al. (2008) identified heterozygosity for the E1053K mutation in the SCN5A gene. The mutation was not found in 720 control alleles.


.0034 CARDIOMYOPATHY, DILATED, 1E

ATRIAL STANDSTILL 1, DIGENIC, INCLUDED
ATRIAL FIBRILLATION, FAMILIAL, 10, INCLUDED
SCN5A, ASP1275ASN
  
RCV000010003...

In a large family reported by Greenlee et al. (1986) with dilated cardiomyopathy with conduction disorder and arrhythmia (CMD1E; 601154), McNair et al. (2004) identified heterozygosity for a 3823G-A mutation in exon 21 of the SCN5A gene, resulting in an asp1275-to-asn (D1275N) substitution and predicting a change of charge within the S2 segment of domain III. The mutation was present in 22 affected family members and was not found in 300 control chromosomes.

Groenewegen et al. (2003) reported the D1275N mutation, coinherited with polymorphisms in the atrial-specific gap junction channel protein connexin-40 (GJA5; 121013), in affected members of a family with atrial standstill (ATRST1; 108770). No member of this family had dilated cardiomyopathy, leading Groenewegen and Wilde (2005) to question whether the D1275N mutation was the primary cause of dilated cardiomyopathy as reported by McNair et al. (2004).

In affected members of a large Finnish family with atrial fibrillation and conduction defects (ATFB10; 614022), Laitinen-Forsblom et al. (2006) identified heterozygosity for the D1275N mutation in the SCN5A gene. The mutation was not found in more than 370 control chromosomes. Echocardiography revealed an enlarged left ventricle with an increased end-diastolic left ventricular diameter in 1 affected individual, and the right ventricle was slightly enlarged in 3 other affected individuals.


.0035 LONG QT SYNDROME 2/3, DIGENIC

SCN5A, ASP1819ASN
  
RCV000010005...

In a 41-year-old female who had cardiac arrest due to torsade de pointes triggered by exercise and leading to ventricular fibrillation, and a QTc of 520 ms (see 603830), Millat et al. (2006) identified biallelic digenic mutations: a 5455G-A transition in exon 28 of the SCN5A gene, resulting in an asp1819-to-asn (D1819N) substitution; and a missense mutation in the KCNH2 gene (R100G; 152427.0023).


.0036 BRUGADA SYNDROME 1

SCN5A, TRP1421TER
  
RCV000010006

In affected members of a 4-generation Han Chinese family with autosomal dominant cardiac arrhythmias and sudden death (BRGDA1; 601144), Niu et al. (2006) identified heterozygosity for a G-A transition in exon 24 of the SCN5A gene, resulting in a trp1421-to-ter (W1421X) substitution. The mutation was not found in 95 control subjects. An asymptomatic 73-year-old male family member was found to be compound heterozygous for W1421X and the R1993Q mutation (600163.0023). Niu et al. (2006) suggested that R1193Q, which results in a gain of sodium channel function, may compensate for the deleterious effects of W1421X.


.0037 MOVED TO 600163.0027


.0038 CARDIOMYOPATHY, DILATED, 1E

SCN5A, 2-BP INS, NT2550
  
RCV000010008...

In a man with dilated cardiomyopathy (CMD1E; 601154) and monomorphic ventricular tachycardia who later developed third-degree heart block requiring pacemaker implantation, Olson et al. (2005) identified heterozygosity for a 2-bp insertion (2550insTG) in exon 17 of the SCN5A gene, resulting in a premature stop codon and a truncated protein. Cardiac biopsy was normal. His father, who carried the mutation, was diagnosed with CMD (ejection fraction, 30%), left bundle branch block, and monomorphic ventricular tachycardia at age 67 years. The mutation was also present in a paternal uncle who had sinus bradycardia, first-degree heart block, and complete left bundle branch block; his paternal grandfather developed congestive heart failure at 50 years of age and died 6 years later, but DNA was unavailable for evaluation.


.0039 CARDIOMYOPATHY, DILATED, 1E

SCN5A, ASP1595HIS
  
RCV000010009...

In a 7-year-old boy with early manifestations of dilated cardiomyopathy (CMD1E; 601154) including sinus bradycardia, left ventricular dilation, and normal contractile function, Olson et al. (2005) identified heterozygosity for a 4783G-C transversion in exon 27 of the SCN5A gene, resulting in an asp1595-to-his (D1595H) substitution at a highly conserved residue in the transmembrane domain. The mutation was found in DNA from postmortem tissue of a brother who died at 34 years of age with an autopsy diagnosis of cardiomyopathy and only mild coronary artery disease. The mutation was also identified in 2 sibs and a paternal uncle, all of whom had sinus bradycardia, and a paternal aunt with borderline left atrial enlargement. His father, an obligate mutation carrier, had atrial fibrillation and died at 49 years of age from pulmonary embolism; his paternal grandfather, a presumed mutation carrier, developed congestive heart failure at 70 years of age.


.0040 BRUGADA SYNDROME 1

SCN5A, VAL232ILE and LEU1308PHE
  
RCV000010010...

In a 45-year-old black man with no history of cardiac disease who developed monomorphic wide-complex ventricular tachycardia with right precordial ST segment elevation consistent with Brugada syndrome (BRGDA1; 601144) after the administration of lidocaine, Barajas-Martinez et al. (2008) identified 2 mutations in the SCN5A gene, a G-to-A transition in exon 6 of the SCN5A gene, resulting in a val232-to-ile (V232I) substitution in the C terminus of the transmembrane segment S4 of domain I, and a C-to-T transition in exon 22, resulting in a leu1308-to-phe (L1308F) substitution, in the C terminus of transmembrane segment S4 of domain III. Although L1308F had previously been identified as a polymorphism found mostly in Americans of African descent (Ackerman et al., 2004), Barajas-Martinez et al. (2008) did not find either mutation in over 400 alleles from 200 ethnically matched controls. The patient's parents were unavailable for study, but given the severity of his clinical manifestations, the authors strongly suspected that both mutations were on the same allele (Dumaine, 2009). Using patch-clamp techniques in mammalian TSA201 cells, Barajas-Martinez et al. (2008) observed use-dependent inhibition of I(Na) by lidocaine that was more pronounced in double-mutant channels than in wildtype; the individual mutations produced a much less accentuated effect. The authors concluded that the double mutation in SCN5A alters the affinity of the cardiac sodium channel for lidocaine such that the drug assumes class IC characteristics with potent use-dependent block of the sodium channel.


.0041 ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, ASN1986LYS
  
RCV000022947...

In a father and son with atrial fibrillation (ATFB10; 614022), Ellinor et al. (2008) identified heterozygosity for a 5958C-A transversion in the SCN5A gene, resulting in an asn1986-to-lys (N1986K) substitution in the C-terminal region of the protein. The mutation was not found in more than 600 ethnically and racially matched control chromosomes. Expression of the N1986K mutant in Xenopus oocytes revealed a hyperpolarizing shift in channel steady-state inactivation.


.0042 ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, HIS445ASP
  
RCV000022948...

In white male proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 39 years of age, Darbar et al. (2008) identified heterozygosity for a G-to-C transversion in the SCN5A gene, resulting in a his445-to-asp (H445D) substitution at a highly conserved residue that was predicted to perturb cardiac sodium channel function. The proband had left atrial enlargement and an ejection fraction of 60% by transthoracic echocardiography. The mutation was also detected in his affected father and brother, but was not found in an unaffected sister or in 720 control alleles.


.0043 ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, ASN470LYS
  
RCV000022949...

In a black male proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 17 years of age, Darbar et al. (2008) identified heterozygosity for a G-to-C transversion in the SCN5A gene, resulting in an asn470-to-lys (N470K) substitution at a highly conserved residue and predicted to perturb cardiac sodium channel function. The proband had left atrial enlargement with an ejection fraction of 60% by transthoracic echocardiography. The mutation was also detected in his affected mother and maternal grandmother, but was not found in an unaffected maternal aunt or in 720 control alleles.


.0044 ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, GLU428LYS
  
RCV000022950...

In a white male proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 52 years of age, Darbar et al. (2008) identified heterozygosity for a A-to-G transition in the SCN5A gene, resulting in an glu428-to-lys (E428K) substitution at a highly conserved residue and predicted to perturb cardiac sodium channel function. The proband had left atrial enlargement with an ejection fraction of 58% by transthoracic echocardiography. The mutation was also detected in his affected daughter and granddaughter, but was not found in an unaffected daughter and granddaughter or in 720 control alleles.


.0045 ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, GLU655LYS
  
RCV000022951...

In a white female proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 37 years of age, Darbar et al. (2008) identified heterozygosity for an A-to-G transition in the SCN5A gene, resulting in a glu655-to-lys (E655K) substitution at a highly conserved residue that was predicted to perturb cardiac sodium channel function. The proband had a normal-sized left atrium and ventricle with an ejection fraction of 55% by transthoracic echocardiography. The mutation was also detected in her affected daughter and maternal grandmother, but was not found in 720 control alleles.


.0046 CARDIOMYOPATHY, DILATED, 1E

SCN5A, ARG222GLN (rs45546039)
  
RCV000032639...

In 6 affected members over 3 generations of a non-Hispanic white family with cardiomyopathy and conduction system disease (CMD1E; 601154), Hershberger et al. (2008) identified heterozygosity for a 36683G-A (numbering per SeattleSNP) transition in exon 6 of the SCN5A gene, resulting in an arg222-to-gln (R222Q) substitution at a conserved residue. The mutation was not found in unaffected family members or in 253 controls.

In 19 affected individuals from 3 unrelated 3-generation families with multifocal ectopic Purkinje-related premature contractions and dilated cardiomyopathy, Laurent et al. (2012) identified heterozygosity for the 665G-A (R222Q) mutation in the SCN5A gene, located in the voltage-sensing S4 segment of domain I. The mutation, which was fully penetrant and strictly segregated with the cardiac phenotype in each family, was not found in 600 control chromosomes; haplotype analysis showed that a founder effect for these 3 families was very unlikely. In vitro studies recapitulated the normalization of the ventricular action potentials in the presence of quinidine. Because only 6 of the 19 patients carrying the R222Q mutation had CMD, and the cardiomyopathy recovered at least partially with antiarrhythmia treatment and a reduction in the number of premature ventricular contractions, Laurent et al. (2012) suggested that CMD might be a consequence of the arrhythmia and not directly linked to the mutation.

In affected members of a 3-generation Canadian family with CMD and junctional escape ventricular capture bigeminy, Nair et al. (2012) identified the R222Q mutation in the SCN5A gene. Heterologous expression studies in Chinese hamster ovary K1 cells revealed a unique biophysical phenotype of R222Q channels in which an approximately 10-mV leftward shift in the sodium current steady-state activation curve occurs without corresponding shifts in steady-state inactivation at cardiomyocyte resting membrane-potential voltages. The activation and inactivation of cells expressing equimolar combinations of wildtype and R222Q channels showed properties intermediate between those seen in cells expressing either wildtype or mutant channels alone. The changes in mutant channel properties were predicted to produce hyperexcitability of R222Q sodium channels.

In 16 affected members over 3 generations of a large kindred with CMD and multiple arrhythmias, including premature ventricular complexes (PVCs) of variable morphology, Mann et al. (2012) identified heterozygosity for the R222Q mutation in the SCN5A gene. The mutation was also identified in 1 clinically unaffected family member, a 56-year-old man with a normal EKG and echocardiogram, but was not found in 200 control chromosomes. Patch-clamp studies showed that the R222Q mutation did not alter sodium channel current density, but did shift steady-state parameters of activation and inactivation to the left. Using a voltage ramp protocol, normalized current responses of mutant channels were of earlier onset and greater magnitude than wildtype. Action potential modeling using Purkinje fiber and ventricular cell models suggested that rate-dependent ectopy of Purkinje fiber origin is the predominant ventricular effect of the R222Q variant; this was supported by the clinical observation that PVC frequency increased during periods of low heart rate at rest and at night, and was reduced by high heart rates during exercise. Patients responded to sodium-channel blocking drugs with early and substantial reductions in PVCs followed by normalization of CMD over time.


.0047 CARDIOMYOPATHY, DILATED, 1E

SCN5A, ILE1835THR
  
RCV000032640...

In 3 affected members over 2 generations of an African American family with cardiomyopathy and conduction system disease (CMD1E; 601154), Hershberger et al. (2008) identified heterozygosity for a 99599T-C transition (numbering per SeattleSNP) in exon 28 of the SCN5A gene, resulting in an ile1835-to-thr (I1835T) substitution at a conserved residue. The mutation was not found in unaffected family members or in 253 controls.


.0048 ATRIAL STANDSTILL 1, DIGENIC

SCN5A, LEU212PRO
  
RCV000058830...

In a Japanese boy with atrial standstill (ATRST1; 108770), Makita et al. (2005) identified coinheritance of a heterozygous c.635C-T transition in exon 6 of the SCN5A gene, resulting in a leu212-to-pro (L212P) substitution in the extracellular loop connecting transmembrane segments 3 and 4 of domain 1 of the Nav1.5 cardiac sodium channel, and heterozygous rare polymorphisms in the GJA5 gene (121013). The L212P mutation, which was also present in the proband's asymptomatic father, was not found in 400 control chromosomes. Functional analysis with the L212P mutant channels demonstrated large hyperpolarizing shifts in both the voltage dependence of activation and inactivation and delayed recovery from inactivation compared to wildtype. The asymptomatic father did not carry the rare polymorphisms in the GJA5 gene; the GJA5 polymorphisms were, however, present in heterozygosity in the proband's unaffected mother and maternal grandmother, who did not carry the L212P mutation.


REFERENCES

  1. Ackerman, M. J., Siu, B. L., Sturner, W. Q., Tester, D. J., Valdivia, C. R., Makielski, J. C., Towbin, J. A. Postmortem molecular analysis of SCN5A defects in sudden infant death syndrome. JAMA 286: 2264-2269, 2001. [PubMed: 11710892, related citations] [Full Text]

  2. Ackerman, M. J., Splawski, I., Makielski, J. C., Tester, D. J., Will, M. L., Timothy, K. W., Keating, M. T., Jones, G., Chadha, M., Burrow, C. R., Stephens, J. C., Xu, C., Judson, R., Curran, M. E. 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 1: 600-607, 2004. [PubMed: 15851227, related citations] [Full Text]

  3. Akai, J., Makita, N., Sakurada, H., Shirai, N., Ueda, K., Kitabatake, A., Nakazawa, K., Kimura, A., Hiraoka, M. A novel SCN5A mutation associated with idiopathic ventricular fibrillation without typical ECG findings of Brugada syndrome. FEBS Lett. 479: 29-34, 2000. [PubMed: 10940383, related citations] [Full Text]

  4. Albert, C. M., Nam, E. G., Rimm, E. B., Jin, H. W., Hajjar, R. J., Hunter, D. J., MacRae, C. A., Ellinor, P. T. Cardiac sodium channel gene variants and sudden cardiac death in women. Circulation 117: 16-23, 2008. [PubMed: 18071069, related citations] [Full Text]

  5. Barajas-Martinez, H. M., Hu, D., Cordeiro, J. M., Wu, Y., Kovacs, R. J., Meltser, H., Kui, H., Elena, B., Brugada, R., Antzelevitch, C., Dumaine, R. Lidocaine-induced Brugada syndrome phenotype linked to a novel double mutation in the cardiac sodium channel. Circ. Res. 103: 396-404, 2008. [PubMed: 18599870, images, related citations] [Full Text]

  6. Bennett, P. B., Yazawa, K., Makita, N., George, A. L., Jr. Molecular mechanism for an inherited cardiac arrhythmia. Nature 376: 683-685, 1995. [PubMed: 7651517, related citations] [Full Text]

  7. Benson, D. W., Wang, D. W., Dyment, M., Knilans, T. K., Fish, F. A., Strieper, M. J., Rhodes, T. H., George, A. L., Jr. Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). J. Clin. Invest. 112: 1019-1028, 2003. [PubMed: 14523039, images, related citations] [Full Text]

  8. Bezzina, C., Veldkamp, M. W., van den Berg, M. P., Postma, A. V., Rook, M. B., Viersma, J.-W., van Langen, I. M., Tan-Sindhunata, G., Bink-Boelkens, M. T. E., van der Hout, A. H., Mannens, M. M. A. M., Wilde, A. A. M. A single Na+ channel mutation causing both long-QT and Brugada syndromes. Circ. Res. 85: 1206-1213, 1999. [PubMed: 10590249, related citations] [Full Text]

  9. Chen, Q., Kirsch, G. E., Zhang, D., Brugada, R., Brugada, J., Brugada, P., Potenza, D., Moya, A., Borggrefe, M., Breithardt, G., Ortiz-Lopez, R., Wang, Z., Antzelevitch, C., O'Brien, R. E., Schulze-Bahr, E., Keating, M. T., Towbin, J. A., Wang, Q. Genetic basis and molecular mechanism for idiopathic ventricular fibrillation. Nature 392: 293-295, 1998. [PubMed: 9521325, related citations] [Full Text]

  10. Chen, S., Chung, M. K., Martin, D., Rozich, R., Tchou, P. J., Wang, Q. SNP S1103Y in the cardiac sodium channel gene SCN5A is associated with cardiac arrhythmias and sudden death in a white family. J. Med. Genet. 39: 913-915, 2002. [PubMed: 12471205, related citations] [Full Text]

  11. Cheng, J., Morales, A., Siegfried, J. D., Li, D., Norton, N., Song, J., Gonzalez-Quintana, J., Makielski, J. C., Hershberger, R. E. SCN5A rare variants in familial dilated cardiomyopathy decrease peak sodium current depending on the common polymorphism H558R and common splice variant Q1077del. Clin. Transl. Sci. 3: 287-294, 2010. [PubMed: 21167004, images, related citations] [Full Text]

  12. Clancy, C. E., Rudy, Y. Linking a genetic defect to its cellular phenotype in a cardiac arrhythmia. Nature 400: 566-569, 1999. [PubMed: 10448858, related citations] [Full Text]

  13. Clancy, C. E., Rudy, Y. Na+ channel mutation that causes both Brugada and long-QT syndrome phenotypes: a simulation study of mechanism. Circulation 105: 1208-1213, 2002. [PubMed: 11889015, images, related citations] [Full Text]

  14. Clancy, C. E., Tateyama, M., Kass, R. S. Insights into the molecular mechanisms of bradycardia-triggered arrhythmias in long QT-3 syndrome. J. Clin. Invest. 110: 1251-1262, 2002. [PubMed: 12417563, images, related citations] [Full Text]

  15. Darbar, D., Kannankeril, P. J., Donahue, B. S., Kucera, G., Stubblefield, T., Haines, J. L., George, A. L., Jr., Roden, D. M. Cardiac sodium channel (SCN5A) variants associated with atrial fibrillation. Circulation 117: 1927-1935, 2008. [PubMed: 18378609, images, related citations] [Full Text]

  16. Dumaine, R., Towbin, J. A., Brugada, P., Vatta, M., Nesterenko, D. V., Nesterenko, V. V., Brugada, J., Brugada, R., Antzelevitch, C. Ionic mechanisms responsible for the electrocardiographic phenotype of the Brugada syndrome are temperature dependent. Circ. Res. 85: 803-809, 1999. [PubMed: 10532948, related citations] [Full Text]

  17. Dumaine, R. Personal Communication. Quebec, Canada 6/2009.

  18. Ellinor, P. T., Nam, E. G., Shea, M. A., Milan, D. J., Ruskin, J. N., MacRae, C. A. Cardiac sodium channel mutation in atrial fibrillation. Heart Rhythm 5: 99-105, 2008. [PubMed: 18088563, related citations] [Full Text]

  19. Freyermuth, F., Rau, F., Kokunai, Y., Linke, T., Sellier, C., Nakamori, M., Kino, Y., Arandel, L., Jollet, A., Thibault, C., Philipps, M., Vicaire, S., and 31 others. Splicing misregulation of SCN5A contributes to cardiac-conduction delay and heart arrhythmia in myotonic dystrophy. Nature Commun. 7: 11067, 2016. Note: Electronic Article. [PubMed: 27063795, images, related citations] [Full Text]

  20. Gellens, M. E., George, A. L., Jr., Chen, L., Chahine, M., Horn, R., Barchi, R. L., Kallen, R. G. Primary structure and functional expression of the human cardiac tetrodotoxin-insensitive voltage-dependent sodium channel. Proc. Nat. Acad. Sci. 89: 554-558, 1992. [PubMed: 1309946, related citations] [Full Text]

  21. George, A. L., Jr., Varkony, T. A., Drabkin, H. A., Han, J., Knops, J. F., Finley, W. H., Brown, G. B., Ward, D. C., Haas, M. Assignment of the human heart tetrodotoxin-resistant voltage-gated Na(+) channel alpha-subunit gene (SCN5A) to band 3p21. Cytogenet. Cell Genet. 68: 67-70, 1995. [PubMed: 7956363, related citations] [Full Text]

  22. Greenlee, P. R., Anderson, J. L., Lutz, J. R., Lindsay, A. E., Hagan, A. D. Familial automaticity-conduction disorder with associated cardiomyopathy. West. J. Med. 144: 33-41, 1986. [PubMed: 3953067, related citations]

  23. Groenewegen, W. A., Firouzi, M., Bezzina, C. R., Vliex, S., van Langen, I. M., Sandkuijl, L., Smits, J. P. P., Hulsbeek, M., Rook, M. B., Jongsma, H. J., Wilde, A. A. M. A cardiac sodium channel mutation cosegregates with a rare connexin40 genotype in familial atrial standstill. Circ. Res. 92: 14-22, 2003. [PubMed: 12522116, related citations] [Full Text]

  24. Groenewegen, W. A., Wilde, A. A. M. Letter regarding article by McNair et al, 'SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia'. (Letter) Circulation 112: e9, 2005. Note: Electronic Article. [PubMed: 15998690, related citations] [Full Text]

  25. Gross, M. B. Personal Communication. Baltimore, Md. 4/23/2019.

  26. Hershberger, R. E., Parks, S. B., Kushner, J. D., Li, D., Ludwigsen, S., Jakobs, P., Nauman, D., Burgess, D., Partain, J., Litt, M. Coding sequence mutations identified in MYH7, TNNT2, SCN5A, CSRP3, LBD3 (sic), and TCAP from 313 patients with familial or idiopathic dilated cardiomyopathy. Clin. Transl. Sci. 1: 21-26, 2008. [PubMed: 19412328, related citations] [Full Text]

  27. Hwang, H. W., Chen, J. J., Lin, Y. J., Shieh, R. C., Lee, M. T., Hung, S. I., Wu, J. Y., Chen, Y. T., Niu, D. M., Hwang, B. T., Chen, Y. T. R1193Q of SCN5A, a Brugada and long QT mutation, is a common polymorphism in Han Chinese. (Letter) J. Med. Genet. 42: e7, 2005. Note: Electronic Article. [PubMed: 15689442, related citations] [Full Text]

  28. Jones, A., Kainz, D., Khan, F., Lee, C., Carrithers, M. D. Human macrophage SCN5A activates an innate immune signaling pathway for antiviral host defense. J. Biol. Chem. 289: 35326-35340, 2014. [PubMed: 25368329, images, related citations] [Full Text]

  29. Kambouris, N. G., Nuss, H. B., Johns, D. C., Marban, E., Tomaselli, G. F., Balser, J. R. A revised view of cardiac sodium channel 'blockade' in the long-QT syndrome. J. Clin. Invest. 105: 1133-1140, 2000. [PubMed: 10772658, images, related citations] [Full Text]

  30. Kyndt, F., Probst, V., Potet, F., Demolombe, S., Chevallier, J.-C., Baro, I., Moisan, J.-P., Boisseau, P., Schott, J.-J., Escande, D., Le Marec, H. Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family. Circulation 104: 3081-3086, 2001. [PubMed: 11748104, related citations] [Full Text]

  31. Laitinen-Forsblom, P. J., Makynen, P., Makynen, H., Yli-Mayry, S., Virtanen, V., Kontula, K., Aalto-Setala, K. SCN5A mutation associated with cardiac conduction defect and atrial arrhythmias. J. Cardiovasc. Electrophysiol. 17: 480-485, 2006. [PubMed: 16684018, related citations] [Full Text]

  32. Laurent, G., Saal, S., Amarouch, M. Y., Beziau, D. M., Marsman, R. F. J., Faivre, L., Barc, J., Dina, C., Bertaux, G., Barthez, O., Thauvin-Robinet, C., Charron, P., and 15 others. Multifocal ectopic Purkinje-related premature contractions. J. Am. Coll. Cardiol. 60: 144-156, 2012. [PubMed: 22766342, related citations] [Full Text]

  33. Maekawa, K., Saito, Y., Ozawa, S., Adachi-Akahane, S., Kawamoto, M., Komamura, K., Shimizu, W., Ueno, K., Kamakura, S., Kamatani, N., Kitakaze, M., Sawada, J. Genetic polymorphisms and haplotypes of the human cardiac sodium channel alpha subunit gene (SCN5A) in Japanese and their association with arrhythmia. Ann. Hum. Genet. 69: 413-428, 2005. [PubMed: 15996170, related citations] [Full Text]

  34. Makita, N., Behr, E., Shimizu, W., Horie, M., Sunami, A., Crotti, L., Schulze-Bahr, E., Fukuhara, S., Mochizuki, N., Makiyama, T., Itoh, H., Christiansen, M., McKeown, P., Miyamoto, K., Kamakura, S., Tsutsui, H., Schwartz, P. J., George, A. L., Jr., Roden, D. M. The E1784K mutation in SCN5A is associated with mixed clinical phenotype of type 3 long QT syndrome. J. Clin. Invest. 118: 2219-2229, 2008. [PubMed: 18451998, images, related citations] [Full Text]

  35. Makita, N., Sasaki, K., Groenewegen, W. A., Yokota, T., Yokoshiki, H., Murakami, T., Tsutsui, H. Congenital atrial standstill associated with coinheritance of a novel SCN5A mutation and connexin 40 polymorphisms. Heart Rhythm 2: 1128-1134, 2005. [PubMed: 16188595, related citations] [Full Text]

  36. Makita, N., Shirai, N., Nagashima, M., Matsuoka, R., Yamada, Y., Tohse, N., Kitabatake, A. A de novo missense mutation of human cardiac Na(+) channel exhibiting novel molecular mechanisms of long QT syndrome. FEBS Lett. 423: 5-9, 1998. [PubMed: 9506831, related citations] [Full Text]

  37. Makita, N., Shirai, N., Wang, D. W., Sasaki, K., George, A. L., Kanno, M., Kitabatake, A. Cardiac Na+ channel dysfunction in Brugada syndrome is aggravated by beta(1)-subunit. Circulation 101: 54-60, 2000. [PubMed: 10618304, related citations] [Full Text]

  38. Mann, S. A., Castro, M. L., Ohanian, M., Guo, G., Zodgekar, P., Sheu, A., Stockhammer, K., Thompson, T., Playford, D., Subbiah, R., Kuchar, D., Aggarwal, A., Vandenberg, J. I., Fatkin, D. R222Q SCN5A mutation is associated with reversible ventricular ectopy and dilated cardiomyopathy. J. Am. Coll. Cardiol. 60: 1566-1573, 2012. [PubMed: 22999724, related citations] [Full Text]

  39. McNair, W. P., Ku, L., Taylor, M. R. G., Fain, P. R., Dao, D., Wolfel, E., Mestroni, L., Familial Cardiomyopathy Registry Research Group. SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia. Circulation 110: 2163-2167, 2004. [PubMed: 15466643, related citations] [Full Text]

  40. McNair, W. P., Ku, L., Taylor, M. R. G., Fain, P. R., Wolfel, E., Mestroni, L. Response to letter regarding article by McNair et al., 'SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia'. (Letter) Circulation 112: e9, 2005. Note: Electronic Article.

  41. Millat, G., Chevalier, P., Restier-Miron, L., Da Costa, A., Bouvagnet, P., Kugener, B., Fayol, L., Gonzalez Armengod, C., Oddou, B., Chanavat, V., Froidefond, E., Perraudin, R., Rousson, R., Rodriguez-Lafrasse, C. Spectrum of pathogenic mutations and associated polymorphisms in a cohort of 44 unrelated patients with long QT syndrome. Clin. Genet. 70: 214-227, 2006. [PubMed: 16922724, related citations] [Full Text]

  42. Miller, T. E., Estrella, E., Myerburg, R. J., Garcia de Viera, J., Moreno, N., Rusconi, P., Ahearn, M. E., Baumbach, L., Kurlansky, P., Wolff, G., Bishopric, N. H. Recurrent third-trimester fetal loss and maternal mosaicism for long-QT syndrome. Circulation 109: 3029-3034, 2004. [PubMed: 15184283, related citations] [Full Text]

  43. Mohler, P. J., Rivolta, I., Napolitano, C., LeMaillet, G., Lambert, S., Priori, S. G., Bennett, V. Na(v)1.5 E1053K mutation causing Brugada syndrome blocks binding to ankyrin-G and expression of Na(v)1.5 on the surface of cardiomyocytes. Proc. Nat. Acad. Sci. 101: 17533-17538, 2004. [PubMed: 15579534, images, related citations] [Full Text]

  44. Nair, K., Pekhletski, R., Harris, L., Care, M., Morel, C., Farid, T., Backx, P. H., Szabo, E., Nanthakumar, K. Escape capture bigeminy: phenotypic marker of cardiac sodium channel voltage sensor mutation R222Q. Heart Rhythm 9: 1681-1688, 2012. [PubMed: 22710484, related citations] [Full Text]

  45. Niu, D.-M., Hwang, B., Hwang, H.-W., Wang, N. H., Wu, J.-Y., Lee, P.-C., Chien, J.-C., Shieh, R.-C., Chen, Y.-T. A common SCN5A polymorphism attenuates a severe cardiac phenotype caused by a nonsense SCN5A mutation in a Chinese family with an inherited cardiac conduction defect. J. Med. Genet. 43: 817-821, 2006. [PubMed: 16707561, images, related citations] [Full Text]

  46. Noble, D. Unraveling the genetics and mechanisms of cardiac arrhythmia. (Commentary) Proc. Nat. Acad. Sci. 99: 5755-5756, 2002. [PubMed: 11983875, related citations] [Full Text]

  47. Nuyens, D., Stengl, M., Dugarmaa, S., Rossenbacker, T., Compernolle, V., Rudy, Y., Smits, J. F., Flameng, W., Clancy, C. E., Moons, L., Vos, M. A., Dewerchin, M., Benndorf, K., Collen, D., Carmeliet, E., Carmeliet, P. Abrupt rate accelerations or premature beats cause life-threatening arrhythmias in mice with long-QT3 syndrome. Nature Med. 7: 1021-1027, 2001. [PubMed: 11533705, related citations] [Full Text]

  48. O'Neill, M. J., Muhammad, A., Li, B., Wada, Y., Hall, L., Solus, J. F., Short, L., Roden, D. M., Glazer, A. M. Dominant negative effects of SCN5A missense variants. Genet. Med. 24: 1238-1248, 2022. [PubMed: 35305865, images, related citations] [Full Text]

  49. Olson, T. M., Michels, V. V., Ballew, J. D., Reyna, S. P., Karst, M. L., Herron, K. I., Horton, S. C., Rodeheffer, R. J., Anderson, J. L. Sodium channel mutations and susceptibility of heart failure and atrial fibrillation. JAMA 293: 447-454, 2005. [PubMed: 15671429, images, related citations] [Full Text]

  50. Papadatos, G. A., Wallerstein, P. M. R., Head, C. E. G., Ratcliff, R., Brady, P. A., Benndorf, K., Saumarez, R. C., Trezise, A. E. O., Huang, C. L.-H., Vandenberg, J. I., Colledge, W. H., Grace, A. A. Slowed conduction and ventricular tachycardia after targeted disruption of the cardiac sodium channel gene Scn5a. Proc. Nat. Acad. Sci. 99: 6210-6215, 2002. [PubMed: 11972032, images, related citations] [Full Text]

  51. Plant, L. D., Bowers, P. N., Liu, Q., Morgan, T., Zhang, T., State, M. W., Chen, W., Kittles, R. A., Goldstein, S. A. N. A common cardiac sodium channel variant associated with sudden infant death in African Americans, SCN5A S1103Y. J. Clin. Invest. 116: 430-435, 2006. [PubMed: 16453024, images, related citations] [Full Text]

  52. Rivolta, I., Abriel, H., Tateyama, M., Liu, H., Memmi, M., Vardas, P., Napolitano, C., Priori, S. G., Kass, R. S. Inherited Brugada and long QT-3 syndrome mutations of a single channel residue of the cardiac sodium channel confer distinct channel and clinical phenotypes. J. Biol. Chem. 276: 30623-30630, 2001. [PubMed: 11410597, related citations] [Full Text]

  53. Rook, M. B., Alshinawi, C. B., Groenewegen, W. A., van Gelder, I. C., van Ginneken, A. C. G., Jongsma, H. J., Mannens, M. M. A. M., Wilde, A. A. M. Human SCN5A gene mutations alter cardiac sodium channel kinetics and are associated with the Brugada syndrome. Cardiovasc. Res. 44: 507-517, 1999. [PubMed: 10690282, related citations] [Full Text]

  54. Schott, J.-J., Alshinawi, C., Kyndt, F., Probst, V., Hoorntje, T. M., Hulsbeek, M., Wilde, A. A. M., Escande, D., Mannens, M. M. A. M., Le Marec, H. Cardiac conduction defects associate with mutations in SCN5A. (Letter) Nature Genet. 23: 20-21, 1999. [PubMed: 10471492, related citations] [Full Text]

  55. Schwartz, P. J., Priori, S. G., Dumaine, R., Napolitano, C., Antzelevitch, C., Stramba-Badiale, M., Richard, T. A., Berti, M. R., Bloise, R. A molecular link between the sudden infant death syndrome and the long-QT syndrome. New Eng. J. Med. 343: 262-267, 2000. [PubMed: 10911008, related citations] [Full Text]

  56. Shin, D.-J., Jang, Y., Park, H.-Y., Lee, J. E., Yang, K., Kim, E., Bae, Y., Kim, J., Kim, J., Kim, S. S., Lee, M. H., Chahine, M., Yoon, S. K. Genetic analysis of the cardiac sodium channel gene SCN5A in Koreans with Brugada syndrome. J. Hum. Genet. 49: 573-578, 2004. [PubMed: 15338453, related citations] [Full Text]

  57. Splawski, I., Shen, J., Timothy, K. W., Lehmann, M. H., Priori, S., Robinson, J. L., Moss, A. J., Schwartz, P. J., Towbin, J. A., Vincent, G. M., Keating, M. T. Spectrum of mutations in long-QT syndrome genes: KVLQT1, HERG, SCN5A, KCNE1, and KCNE2. Circulation 102: 1178-1185, 2000. [PubMed: 10973849, related citations] [Full Text]

  58. Splawski, I., Timothy, K. W., Tateyama, M., Clancy, C. E., Malhotra, A., Beggs, A. H., Cappuccio, F. P., Sagnella, G. A., Kass, R. S., Keating, M. T. Variant of SCN5A sodium channel implicated in risk of cardiac arrhythmia. Science 297: 1333-1336, 2002. [PubMed: 12193783, related citations] [Full Text]

  59. Tan, H. L., Bink-Boelkens, M. T. E., Bezzina, C. R., Viswanathan, P. C., Beaufort-Krol, G. C. M., van Tintelen, P. J., van den Berg, M. P., Wilde, A. A. M., Balser, J. R. A sodium-channel mutation causes isolated cardiac conduction disease. Nature 409: 1043-1047, 2001. [PubMed: 11234013, related citations] [Full Text]

  60. Tan, H. L., Kupershmidt, S., Zhang, R., Stepanovic, S., Roden, D. M., Wilde, A. A. M., Anderson, M. E., Balser, J. R. A calcium sensor in the sodium channel modulates cardiac excitability. Nature 415: 442-447, 2002. [PubMed: 11807557, related citations] [Full Text]

  61. Tester, D. J., Will, M. L., Haglund, C. M., Ackerman, M. J. Compendium of cardiac channel mutations in 541 consecutive unrelated patients referred for long QT syndrome genetic testing. Heart Rhythm 2: 507-517, 2005. [PubMed: 15840476, related citations] [Full Text]

  62. Vatta, M., Dumaine, R., Varghese, G., Richard, T. A., Shimizu, W., Aihara, N., Nademanee, K., Brugada, R., Brugada, J., Veerakul, G., Li, H., Bowles, N. E., Brugada, P., Antzelevitch, C., Towbin, J. A. Genetic and biophysical basis of sudden unexplained nocturnal death syndrome (SUNDS), a disease allelic to Brugada syndrome. Hum. Molec. Genet. 11: 337-345, 2002. [PubMed: 11823453, related citations] [Full Text]

  63. Veldkamp, M. W., Wilders, R., Baartscheer, A., Zegers, J. G., Bezzina, C. R., Wilde, A. A. M. Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in LQT3 families. Circ. Res. 92: 976-983, 2003. [PubMed: 12676817, related citations] [Full Text]

  64. Viswanathan, P. C., Benson, D. W., Balser, J. R. A common SCN5A polymorphism modulates the biophysical effects of an SCN5A mutation. J. Clin. Invest. 111: 341-346, 2003. [PubMed: 12569159, images, related citations] [Full Text]

  65. Wang, D. W., Viswanathan, P. C., Balser, J. R., George, A. L., Jr., Benson, W. Clinical, genetic and biophysical characterisation of SCN5A mutations associated with atrioventricular block. Circulation 105: 341-346, 2002. [PubMed: 11804990, related citations] [Full Text]

  66. Wang, D. W., Yazawa, K., George, A. L., Jr., Bennett, P. B. Characterization of human cardiac Na(+) channel mutations in the congenital long QT syndrome. Proc. Nat. Acad. Sci. 93: 13200-13205, 1996. [PubMed: 8917568, images, related citations] [Full Text]

  67. Wang, D. W., Yazawa, K., Makita, N., George, A. L., Jr., Bennett, P. B. Pharmacological targeting of long QT mutant sodium channels. J. Clin. Invest. 99: 1714-1720, 1997. [PubMed: 9120016, related citations] [Full Text]

  68. Wang, Q., Chen, S., Chen, Q., Wan, X., Shen, J., Hoeltge, G. A., Timur, A. A., Keating, M. T., Kirsch, G. E. The common SCN5A mutation R1193Q causes LQTS-type electrophysiological alterations of the cardiac sodium channel. J. Med. Genet. 41: e66, 2004. Note: Electronic Article. [PubMed: 15121794, related citations] [Full Text]

  69. Wang, Q., Li, Z., Shen, J., Keating, M. T. Genomic organization of the human SCN5A gene encoding the cardiac sodium channel. Genomics 34: 9-16, 1996. [PubMed: 8661019, related citations] [Full Text]

  70. Wang, Q., Shen, J., Li, Z., Timothy, K., Vincent, G. M., Priori, S. G., Schwartz, P. J., Keating, M. T. Cardiac sodium channel mutations in patients with long QT syndrome, an inherited cardiac arrhythmia. Hum. Molec. Genet. 4: 1603-1607, 1995. [PubMed: 8541846, related citations] [Full Text]

  71. Wang, Q., Shen, J., Splawski, I., Atkinson, D., Li, Z., Robinson, J. L., Moss, A. J., Towbin, J. A., Keating, M. T. SCN5A mutations associated with an inherited cardiac arrhythmia, long QT syndrome. Cell 80: 805-811, 1995. [PubMed: 7889574, related citations] [Full Text]

  72. Wang, Q. Author's reply: link of SCN5A SNP R1193Q to long QT syndrome. (Letter) J. Med. Genet. 42: e8, 2005. Note: Electronic Article.

  73. Wei, J., Wang, D. W., Alings, M., Fish, F., Wathen, M., Roden, D. M., George, A. L., Jr. Congenital long-QT syndrome caused by a novel mutation in a conserved acidic domain of the cardiac Na(+) channel. Circulation 99: 3165-3171, 1999. [PubMed: 10377081, related citations] [Full Text]

  74. Westenskow, P., Splawski, I., Timothy, K. W., Keating, M. T., Sanguinetti, M. C. Compound mutations: a common cause of severe long-QT syndrome. Circulation 109: 1834-1841, 2004. [PubMed: 15051636, related citations] [Full Text]

  75. Yang, P., Kanki, H., Drolet, B., Yang, T., Wei, J., Viswanathan, P. C., Hohnloser, S. H., Shimizu, W., Schwartz, P. J., Stanton, M., Murray, K. T., Norris, K., George, A. L., Jr., Roden, D. M. Allelic variants in long-QT disease genes in patients with drug-associated torsades de pointes. Circulation 105: 1943-1948, 2002. [PubMed: 11997281, related citations] [Full Text]


Sonja A. Rasmussen - updated : 09/12/2022
Matthew B. Gross - updated : 04/23/2019
Patricia A. Hartz - updated : 11/16/2016
Paul J. Converse - updated : 1/8/2015
Marla J. F. O'Neill - updated : 4/29/2014
Marla J. F. O'Neill - updated : 1/29/2013
Marla J. F. O'Neill - updated : 6/1/2011
Marla J. F. O'Neill - updated : 6/8/2009
Marla J. F. O'Neill - updated : 12/23/2008
Marla J. F. O'Neill - updated : 5/14/2008
Marla J. F. O'Neill - updated : 3/6/2008
Marla J. F. O'Neill - updated : 2/12/2008
Marla J. F. O'Neill - updated : 1/12/2007
Marla J. F. O'Neill - updated : 11/9/2006
Marla J. F. O'Neill - updated : 7/10/2006
Victor A. McKusick - updated : 2/20/2006
Marla J. F. O'Neill - updated : 1/31/2006
Marla J. F. O'Neill - updated : 10/11/2005
Victor A. McKusick - updated : 1/27/2005
Victor A. McKusick - updated : 1/3/2005
Cassandra L. Kniffin - updated : 10/26/2004
Marla J. F. O'Neill - updated : 2/18/2004
Victor A. McKusick - updated : 11/18/2003
Victor A. McKusick - updated : 6/30/2003
Denise L. M. Goh - updated : 1/6/2003
Ada Hamosh - updated : 10/18/2002
George E. Tiller - updated : 9/23/2002
Deborah L. Stone - updated : 6/26/2002
Victor A. McKusick - updated : 6/6/2002
Paul Brennan - updated : 3/27/2002
Paul Brennan - updated : 3/8/2002
Ada Hamosh - updated : 1/22/2002
Victor A. McKusick - updated : 11/6/2001
Ada Hamosh - updated : 2/27/2001
Victor A. McKusick - updated : 9/27/2000
Victor A. McKusick - updated : 9/15/2000
Victor A. McKusick - updated : 6/1/2000
Paul Brennan - updated : 4/12/2000
Paul Brennan - updated : 4/3/2000
Paul Brennan - updated : 4/3/2000
Victor A. McKusick - updated : 2/24/2000
Victor A. McKusick - updated : 1/12/2000
Paul Brennan - updated : 8/31/1999
Ada Hamosh - updated : 8/4/1999
Ada Hamosh - updated : 5/25/1999
Victor A. McKusick - updated : 10/2/1998
Victor A. McKusick - updated : 5/12/1998
Victor A. McKusick - updated : 3/17/1998
Victor A. McKusick - updated : 5/27/1997
Creation Date:
Victor A. McKusick : 10/26/1994
carol : 01/11/2023
carol : 10/03/2022
alopez : 09/30/2022
carol : 09/25/2022
carol : 09/13/2022
carol : 09/12/2022
carol : 09/01/2020
carol : 05/29/2019
mgross : 04/23/2019
alopez : 11/07/2018
carol : 01/05/2018
carol : 10/05/2017
alopez : 09/14/2017
carol : 01/12/2017
alopez : 11/16/2016
alopez : 11/09/2016
alopez : 11/09/2016
carol : 05/24/2016
carol : 5/23/2016
mgross : 1/30/2015
carol : 1/14/2015
mcolton : 1/8/2015
alopez : 11/12/2014
carol : 4/29/2014
mcolton : 4/28/2014
carol : 4/17/2014
carol : 3/19/2014
carol : 8/27/2013
carol : 3/8/2013
alopez : 1/30/2013
alopez : 1/29/2013
carol : 12/15/2011
carol : 11/23/2011
terry : 11/4/2011
carol : 11/3/2011
carol : 7/15/2011
wwang : 6/3/2011
wwang : 6/3/2011
terry : 6/1/2011
wwang : 6/30/2009
terry : 6/8/2009
carol : 6/2/2009
carol : 12/24/2008
terry : 12/23/2008
carol : 12/22/2008
carol : 5/14/2008
carol : 3/6/2008
wwang : 3/5/2008
wwang : 2/26/2008
terry : 2/12/2008
joanna : 2/7/2008
carol : 11/15/2007
alopez : 10/4/2007
carol : 9/10/2007
carol : 1/19/2007
terry : 1/12/2007
carol : 12/8/2006
carol : 11/16/2006
carol : 11/9/2006
carol : 11/9/2006
carol : 10/4/2006
terry : 8/24/2006
wwang : 7/11/2006
terry : 7/10/2006
joanna : 6/2/2006
carol : 2/22/2006
carol : 2/22/2006
terry : 2/20/2006
wwang : 2/20/2006
wwang : 2/3/2006
terry : 1/31/2006
wwang : 10/14/2005
terry : 10/11/2005
wwang : 2/10/2005
wwang : 2/8/2005
terry : 1/27/2005
wwang : 1/6/2005
wwang : 1/6/2005
terry : 1/3/2005
tkritzer : 10/27/2004
ckniffin : 10/26/2004
carol : 10/26/2004
carol : 10/12/2004
joanna : 9/10/2004
ckniffin : 4/30/2004
carol : 4/30/2004
ckniffin : 4/14/2004
tkritzer : 3/18/2004
tkritzer : 3/16/2004
carol : 2/18/2004
alopez : 11/25/2003
tkritzer : 11/20/2003
terry : 11/18/2003
carol : 7/14/2003
tkritzer : 7/8/2003
terry : 6/30/2003
carol : 2/5/2003
carol : 1/6/2003
carol : 1/6/2003
alopez : 10/23/2002
terry : 10/18/2002
cwells : 9/23/2002
carol : 6/26/2002
mgross : 6/10/2002
terry : 6/6/2002
alopez : 3/27/2002
carol : 3/14/2002
alopez : 3/8/2002
alopez : 1/23/2002
terry : 1/22/2002
carol : 11/8/2001
carol : 11/8/2001
mcapotos : 11/6/2001
alopez : 3/7/2001
alopez : 3/6/2001
terry : 2/27/2001
mcapotos : 10/13/2000
mcapotos : 10/11/2000
terry : 9/27/2000
carol : 9/25/2000
terry : 9/22/2000
terry : 9/15/2000
mcapotos : 6/15/2000
mcapotos : 6/14/2000
terry : 6/1/2000
alopez : 4/12/2000
alopez : 4/3/2000
alopez : 4/3/2000
mcapotos : 3/17/2000
mcapotos : 3/8/2000
terry : 2/24/2000
mgross : 2/17/2000
terry : 1/12/2000
carol : 11/4/1999
mgross : 8/31/1999
alopez : 8/4/1999
terry : 8/4/1999
kayiaros : 7/8/1999
carol : 7/7/1999
carol : 5/25/1999
carol : 5/11/1999
carol : 10/7/1998
terry : 10/2/1998
terry : 6/4/1998
carol : 5/21/1998
terry : 5/12/1998
alopez : 3/18/1998
terry : 3/17/1998
jenny : 5/30/1997
terry : 5/27/1997
terry : 12/10/1996
terry : 12/5/1996
terry : 6/5/1996
terry : 6/3/1996
joanna : 12/29/1995
mimadm : 9/23/1995
mark : 9/22/1995
terry : 4/20/1995
mark : 3/30/1995
terry : 10/26/1994

* 600163

SODIUM VOLTAGE-GATED CHANNEL, ALPHA SUBUNIT 5; SCN5A


Alternative titles; symbols

SODIUM CHANNEL, VOLTAGE-GATED, TYPE V, ALPHA SUBUNIT
NAV1.5


HGNC Approved Gene Symbol: SCN5A

SNOMEDCT: 283645003, 51178009, 60423000, 698249005;   ICD10CM: I49.8;   ICD9CM: 427.81, 798.0;  


Cytogenetic location: 3p22.2     Genomic coordinates (GRCh38): 3:38,548,062-38,649,687 (from NCBI)


Gene-Phenotype Relationships

Location Phenotype Phenotype
MIM number
Inheritance Phenotype
mapping key
3p22.2 {Sudden infant death syndrome, susceptibility to} 272120 Autosomal recessive 3
Atrial fibrillation, familial, 10 614022 Autosomal dominant 3
Brugada syndrome 1 601144 Autosomal dominant 3
Cardiomyopathy, dilated, 1E 601154 Autosomal dominant 3
Heart block, nonprogressive 113900 Autosomal dominant 3
Heart block, progressive, type IA 113900 Autosomal dominant 3
Long QT syndrome 3 603830 Autosomal dominant 3
Sick sinus syndrome 1 608567 Autosomal recessive 3
Ventricular fibrillation, familial, 1 603829 3

TEXT

Cloning and Expression

Gellens et al. (1992) cloned and characterized the cardiac sodium channel gene SCN5A. The deduced 2,016-amino acid protein has a structure similar to that of previously characterized sodium channels (see 182392) and contains 4 homologous domains, each of which has 6 putative membrane-spanning regions.

Freyermuth et al. (2016) stated that alternative splicing creates fetal and adult isoforms of SCN5A that differ in inclusion of alternative exons 6a or 6b, respectively. Both exons 6 have 92 bp, but encode 7 different amino acids in the voltage-sensor region of SCN5A domain I.


Gene Structure

Wang et al. (1996) found that SCN5A consists of 28 exons spanning approximately 80 kb. They described the sequences of all intron/exon boundaries and a dinucleotide repeat polymorphism in intron 16.


Mapping

George et al. (1995) mapped the SCN5A gene to chromosome 3p21 by fluorescence in situ hybridization, thus making it an important candidate gene for long QT syndrome-3 (LQT3; 603830).

Gross (2019) mapped the SCN5A gene to chromosome 3p22.2 based on an alignment of the SCN5A sequence (GenBank BC051374) with the genomic sequence (GRCh38).


Gene Function

By immunoprecipitation and nano-liquid chromatography-mass spectroscopy/mass spectroscopy of transgenic mouse bone marrow macrophages expressing the human macrophage splice variant of SCN5A, followed by Western blot analysis, Jones et al. (2014) identified interaction of SCN5A with activating transcription factor-2 (ATF2; 123811). Microarray analysis of SCN5A-positive macrophages revealed increased expression of Sp100 (604585), an Atf2-regulated gene. Knockdown of Adcy8 (103070), the calcium-dependent isoform of adenylate cyclase, inhibited channel agonist-induced expression of Sp100-related genes. Activation of SCN5A increased expression of cAMP in macrophages. Treatment of macrophages with poly(I:C), a mimic of viral double-stranded RNA, activated the Adcy8 signaling pathway to regulate expression of Sp100-related genes and Ifnb (147640). Electrophysiologic analysis showed that the SCN5A variant mediated nonselective outward currents, as well as a small yet detectable inward current. Jones et al. (2014) proposed that human macrophage SCN5A initiates signaling in an innate immune pathway relevant to antiviral host defense, and that SCN5A is a pathogen sensor.

Myotonic dystrophy (see DM1, 160900) is caused by expression of mutant RNAs containing expanded CUG repeats. These repeats sequester muscleblind-like (MBNL; see MBNL1, 606516) splicing factors in nuclear RNA foci, resulting in changes in pre-mRNA splicing. Freyermuth et al. (2016) showed that MBNL1 specifically promoted inclusion of exon 6b in SCN5A pre-mRNA and expression of the adult SCN5A isoform. Freyermuth et al. (2016) found that left ventricle samples of 3 adult DM1 patients showed alternative splicing in a number of genes, including SCN5A. A portion of the SCN5A mRNA in these samples was the fetal isoform. When expressed in Xenopus oocytes, the fetal isoform of SCN5A showed reduced excitability compared with the adult SCN5A isoform. In mice, expression of fetal Scn5a promoted heart arrhythmia and cardiac-conduction delay, which are 2 predominant features of myotonic dystrophy.


Molecular Genetics

Missense mutations in the skeletal muscle sodium channel gene, SCN4A (603967), cause myotonia. Physiologic data show that these mutations affect sodium channel inactivation and lead to repetitive depolarizations, consistent with the myotonic phenotype. By analogy, similar mutations in the cardiac sodium channel gene might be expected to cause a phenotype like LQT. Indeed, Wang et al. (1995) found a mutation in the SCN5A gene in families with chromosome 3-linked LQT (see 600163.0001).

Bennett et al. (1995) determined the functional defect resulting from the 3-amino acid (KPQ) deletion (600163.0001) in the SCN5A protein. By expression of recombinant human heart sodium channels in Xenopus laevis oocytes, mutant channels showed a sustained inward current during membrane depolarization. Single-channel recordings indicated that mutant channels fluctuate between normal and noninactivating gating modes. Persistent inward sodium current explains prolongation of cardiac action potentials and provides a molecular mechanism for the chromosome 3-linked form of long QT syndrome.

Wang et al. (1995) identified SCN5A mutations in affected members of 4 additional families with chromosome 3-linked LQT. Two of the families had the same 9-bp deletion found earlier; the other families were found to have missense mutations affecting highly conserved amino acid residues (600163.0002 and 600163.0003). The location and character of the mutation suggested to the authors that this form of LQT results from a delay in cardiac sodium channel fast inactivation or altered voltage-dependence of inactivation.

Wang et al. (1996) determined the biophysical and functional characteristics of each of the 3 distinct mutations that had been identified in the cardiac sodium channel gene in patients with LQT3 to that time. For this they used heterologous expression of a recombinant human heart sodium channel in a mammalian cell line. Each mutation caused a sustained, noninactivating sodium current amounting to a few percent of the peak inward sodium current, observable during long (more than 50 msec) depolarizations. The voltage dependence and rate of inactivation were altered and the rate of recovery from inactivation was changed compared with wildtype channels. These mutations in diverse regions of the ion channel protein all produced a common defect in channel gating that can cause the long QT phenotype. The sustained inward current caused by these mutations would prolong the action potential. Furthermore, they might create conditions that promote arrhythmias due to prolonged depolarization and the altered recovery from inactivation.

Wang et al. (1997) explored the potential for targeted suppression of the defect in LQT3 by heterologous expression of mutant channels in cultured human cells. Channel behavior and inhibition by mexiletine were investigated by whole-cell patch-clamp methods. The investigators showed that late-opening LQT3 mutant channels were much more sensitive to inhibition by mexiletine than were wildtype sodium channels. The defective late openings were selectively suppressed more than the peak sodium current and these late openings could be suppressed by concentrations at the lower end of the therapeutic range.

Using a candidate gene approach, Chen et al. (1998) studied 6 small families and 2 sporadic patients with idiopathic ventricular fibrillation (IVF; 603829) using SSCP and DNA sequence analyses to identify mutations in known ion channel genes, including the cardiac sodium channel gene SCN5A. They identified several mutations in families with a distinct form of IVF known as Brugada syndrome (BRGDA1; 601144). In 1 family all affected members had 2 mutations: an arg1232-to-trp mutation in exon 21 of the gene in the extracellular loop between transmembrane segments S1 and S2 of domain III of the protein, and a thr1620-to-met mutation in exon 28 of the gene in the extracellular loop between S3 and S4 of domain IV of the protein (600163.0004). Additional SCN5A mutations were found in 2 IVF families: insertion of 2 nucleotides (AA) in the splice-donor sequence of intron 7 (600163.0005); and deletion of a single nucleotide (A) at codon 1397, resulting in an in-frame stop codon (600163.0006). The frameshift mutation caused the sodium channel to be nonfunctional.

Schott et al. (1999) reported a mutation in the SCN5A gene that segregated with progressive familial heart block (PFHB1A; 113900) in an autosomal dominant manner in a large French family. In a smaller Dutch family, another SCN5A mutation cosegregated with familial nonprogressive conduction defect (see 113900). The French family with PFHB1A was identified through a member with right bundle branch block (RBBB) and syncope; a brother had RBBB, and a sister had complete atrioventricular (AV) block and syncope. Clinical and electrocardiographic abnormalities were found in 15 members of the family; mean QRS duration was 135 +/- 7 ms. RBBB was present in 5, left bundle branch block (LBBB) in 2, left anterior or posterior hemiblock in 3, and long PR interval (more than 210 ms) in 8. None had a structural heart disease. Four members of earlier generations had received a pacemaker implantation because of syncope or complete AV block. Long-term follow-up of several affected members demonstrated that their conduction defect increased in severity with age. In the Dutch family, the proband presented after birth with an asymptomatic first-degree AV block associated with RBBB (PR interval and QRS duration, 200 and 120 ms, respectively). In the French family, Schott et al. (1999) excluded the chromosome 19 locus for this disorder (604559) by linkage studies, as well as other loci for inherited cardiac disorders associated with conduction defects. SCN5A was considered a candidate locus, and using markers flanking SCN5A, the authors demonstrated segregation of the disease with D3S1260 in every affected individual (maximum lod score of 6.03 at theta of 0.0). A donor splice site mutation in SCN5A was found in the French family (600163.0009), and a frameshift mutation was identified in the Dutch family (600163.0010). Clinical data and family histories indicated that none of the affected individuals in these 2 families had LQT3 or idiopathic ventricular fibrillation (Brugada syndrome). Therefore, PFHB1 represents a third cardiac disease linked to SCN5A.

Splawski et al. (2000) screened 262 unrelated individuals with LQT syndrome for mutations in the 5 defined genes (KCNQ1, 607542; KCNH2, 152427; SCN5A; KCNE1, 176261; and KCNE2, 603796) and identified mutations in 177 individuals (68%). KCNQ1 and KCNH2 accounted for 87% of mutations (42% and 45%, respectively), and SCN5A, KCNE1, and KCNE2 for the remaining 13% (8%, 3%, and 2%, respectively).

Tan et al. (2002) demonstrated that calmodulin (114180) binds to the carboxy terminal 'IQ' domain of the SCN5A in a calcium-dependent manner. This binding interaction significantly enhances slow inactivation, a channel-gating process linked to life-threatening idiopathic ventricular arrhythmias. Mutations targeted to the IQ domain disrupted calmodulin binding and eliminated calcium/calmodulin-dependent slow inactivation, whereas the gating effects of calcium/calmodulin were restored by intracellular application of a peptide modeled after the IQ domain. A naturally occurring mutation (A1924T; 600163.0012) in the IQ domain altered SCN5A function in a manner characteristic of the Brugada syndrome, but at the same time inhibited slow inactivation induced by calcium/calmodulin, yielding a clinically benign (arrhythmia-free) phenotype.

Splawski et al. (2002) identified a common variant of the SCN5A gene, ser1103 to tyr (S1103Y; 600163.0024), which is present in 13.2% of African Americans and is associated with accelerated channel activation, increasing the likelihood of abnormal cardiac repolarization and arrhythmia. Splawski et al. (2002) suggested that the S1103Y mutation in the African American population may be a useful molecular marker for the prediction of arrhythmia susceptibility in the context of additional acquired risk factors such as the use of certain medications or the presence of hypokalemia.

Rivolta et al. (2001) identified 2 mutations at the same codon of the SCN5A gene: a tyr1795-to-cys mutation (Y1795C; 600163.0029) in a patient with LQT3, and a Y1795H (600163.0030) mutation in a patient with Brugada syndrome. Functional analysis revealed marked and opposing effects on channel gating consistent with activity associated with the cellular basis of each clinical disorder: Y1795H accelerated and Y1795C slowed the onset of activation; Y1795H, but not Y1795C, caused a marked negative shift in the voltage dependence of inactivation; and neither affected the kinetics of the recovery from inactivation. However, both mutations increased the expression of sustained Na(+) channel activity compared with wildtype channels, although this effect was most pronounced for the Y1795C mutation, and both promoted entrance into an intermediate or slowly developing inactivated state. Rivolta et al. (2001) concluded that these data confirmed the key role of the C-terminal tail of the cardiac Na(+) channel in the control of channel gating and provided further evidence of the close interrelationship between Brugada syndrome and LQT3 at the molecular level.

Clancy et al. (2002) performed detailed kinetic analyses of the Y1795C mutant described by Rivolta et al. (2001). Theoretical entry and exit rates from the bursting mode of gating were derived from single channels. Computational analysis suggested that the amount of time mutant channels spend bursting (burst mode dwell time) is primarily responsible for rate-dependent changes in single-channel bursting and macroscopic inward sodium channel (I-sus), hence delaying repolarization and prolonging the QT interval. This prediction was experimentally confirmed by analysis of delta-KPQ mutant channels (600163.0001) for which the burst mode exit rate (determined by the burst mode dwell time) was found to be very similar to the derived rate for Y1795C channels. These results provided an explanation of the molecular mechanism for bradycardia-induced QT prolongation in patients carrying LQT3 mutations.

Veldkamp et al. (2003) studied the effect of the 1795insD SCN5A mutation (600163.0013), which causes LQT3 or Brugada syndrome, on sinoatrial (SA) pacemaking. Activity of 1795insD channels during SA node pacemaking was confirmed by action potential (AP) clamp experiments, and the previously characterized persistent inward current (I-pst) and negative shift were implemented into SA node (AP) models. The -10 mV shift decreased the sinus rate by decreasing the diastolic depolarization rate, whereas the I-pst decreased the sinus rate by AP prolongation, despite a concomitant increase in the diastolic depolarization rate. In combination, a moderate I-pst (1 to 2%) and the shift reduced the sinus rate by about 10%. Veldkamp et al. (2003) concluded that sodium channel mutations displaying an I-pst or a negative shift in inactivation may account for the bradycardia seen in LQT3 patients, whereas SA node pauses or arrest may result from failure of SA node cells to repolarize under conditions of extra net inward current.

Based on prior associations with disorders of cardiac rhythm and conduction, Benson et al. (2003) screened the SCN5A gene as a candidate gene in 10 pediatric patients from 7 families who were diagnosed with autosomal recessive congenital sick sinus syndrome (SSS1; 608567) during the first decade of life. Probands from 3 kindreds exhibited compound heterozygosity for 6 distinct SCN5A alleles (e.g., 600163.0025), 2 of which had previously been associated with dominant disorders of cardiac excitability. Biophysical characterization of the mutants using heterologously expressed recombinant human heart sodium channels demonstrated loss of function or significant impairment in channel gating that predicted reduced myocardial excitability. Thus Benson et al. (2003) provided a molecular basis for some forms of congenital SSS and defined a recessive disorder of a human heart voltage-gated sodium channel.

In a patient with Brugada syndrome, Mohler et al. (2004) identified an E1053K mutation (600163.0033) in the ankyrin-binding motif of Na(v)1.5. The mutation abolished binding of Na(v)1.5 to ankyrin-G (ANK3; 600465), and also prevented accumulation of Na(v)1.5 at cell surface sites in ventricular cardiomyocytes. Both ankyrin-G and Na(v)1.5 localized at intercalated disc and T-tubule membranes in cardiomyocytes, and Na(v)1.5 coimmunoprecipitated with the 190-kD ankyrin-G isoform from detergent-soluble lysates from rat heart. These data suggested that Na(v)1.5 associates with ankyrin-G and that ankyrin-G is required for Na(v)1.5 localization at excitable membranes in cardiomyocytes.

Miller et al. (2004) reported a case of repeated germline transmission of a severe form of LQT syndrome from an asymptomatic mother with somatic mosaicism for a mutation in the SCN5A gene (600163.0007).

Maekawa et al. (2005) sequenced the SCN5A gene in 166 Japanese patients with arrhythmia who were not diagnosed with LQT or Brugada syndrome and in 232 healthy controls, identifying 69 genetic variations including 66 SNPs. The frequency of a 703+130G-A SNP was significantly higher in patients than in controls (OR, 1.70), suggesting an association with an unknown risk factor for arrhythmia. Haplotype analysis revealed that the so-called GG haplotype with both the leu1988-to-arg and his558-to-arg (600163.0031) SNPs was significantly less frequent in patients than in controls (p = 0.018), suggesting a possible protective effect.

Tester et al. (2005) analyzed 5 LQTS-associated cardiac channel genes in 541 consecutive unrelated patients with LQT syndrome (average QTc, 482 ms). In 272 (50%) patients, they identified 211 different pathogenic mutations, including 88 in KCNQ1, 89 in KCNH2, 32 in SCN5A, and 1 each in KCNE1 and KCNE2. Mutations considered pathogenic were absent in more than 1,400 reference alleles. Among the mutation-positive patients, 29 (11%) had 2 LQTS-causing mutations, of which 16 (8%) were in 2 different LQTS genes (biallelic digenic). Tester et al. (2005) noted that patients with multiple mutations were younger at diagnosis, but they did not discern any genotype/phenotype correlations associated with location or type of mutation.

In 44 unrelated patients with LQT syndrome, Millat et al. (2006) used DHLP chromatography to analyze the KCNQ1, KCNH2, SCN5A, KCNE1, and KCNE2 genes for mutations and SNPs. Most of the patients (84%) showed a complex molecular pattern, with an identified mutation associated with 1 or more SNPs located in several LQTS genes; 4 of the patients also had a second mutation in a different LQTS gene (biallelic digenic inheritance; see, e.g., 600163.0007 and 603796.0005).

In affected members of the family reported by Greenlee et al. (1986) with a form of dilated cardiomyopathy (CMD1E; 601154), McNair et al. (2004) identified heterozygosity for a missense mutation (D1765N; 600163.0001) in the SCN5A gene. In affected members of a family with atrial standstill (ATRST1; 108770), Groenewegen et al. (2003) had identified coinheritance of the D1275N mutation in the SCN5A gene with polymorphisms in the atria-specific junction channel protein connexin-40 (GJA5; 121013). None of the patients with atrial standstill had dilated cardiomyopathy, leading Groenewegen and Wilde (2005) to question the relationship of the SCN5A mutation to dilated cardiomyopathy in the family reported by McNair et al. (2004). McNair et al. (2005) responded that the younger age of the affected members studied by Groenewegen et al. (2003) as well as additional or genetic environmental factors may account for the difference between the 2 families.

In a Japanese family in which an 11-year-old boy had sick sinus syndrome that progressed to atrial standstill, Makita et al. (2005) analyzed 3 cardiac ion channel genes previously associated with atrial standstill, atrial fibrillation, or sick sinus syndrome: SCN5A, HCN4 (605206), and GJA5. No mutations were found in HCN4, but the proband and his asymptomatic father were heterozygous for a missense mutation in SCN5A (L212P; 600163.0048). In addition, the proband and his unaffected mother and maternal grandmother were all heterozygous for the same 2 rare GJA5 polymorphisms identified by Groenewegen et al. (2003) in atrial standstill patients, -44A/+71G. Functional analysis with the L212P mutant channels demonstrated large hyperpolarizing shifts in both the voltage dependence of activation and inactivation and delayed recovery from inactivation compared to wildtype. Makita et al. (2005) suggested that defects in SCN5A underlie atrial standstill, and that coinheritance of GJA5 polymorphisms represents a possible genetic modifier of the clinical manifestations.

Olson et al. (2005) analyzed the SCN5A gene in 156 unrelated patients with dilated cardiomyopathy who were negative for mutations in the known CMD genes encoding cardiac actin (102540), alpha-tropomyosin (191010), and metavinculin (see 193065), and identified 5 heterozygous mutations in 5 probands, respectively (see, e.g., 600163.0027, 600163.0038-600163.0039). All of the mutations altered highly conserved residues in the transmembrane domains of SCN5A.

Albert et al. (2008) analyzed 5 cardiac ion channel genes, SCN5A, KCNQ1, KCNH2, KCNE1, and KCNE2, in 113 cases of sudden cardiac death. No mutations or rare variants were identified in any of the 53 male subjects, but in 6 (10%) of 60 female subjects, 5 rare missense variants in SCN5A were identified, 2 previously associated with long QT syndrome, 1 with sudden infant death syndrome, and 2 not previously reported in control populations. Functional studies showed that all of the variants resulted in significantly shorter recovery times from inactivation. Albert et al. (2008) concluded that functionally significant mutations and rare variants in the SCN5A gene may contribute to the risk of sudden cardiac death in women.

Makita et al. (2008) genotyped 66 members of 44 LQT3 families of multiple ethnicities and identified the E1784K mutation (600163.0008) in 41 individuals from 15 (34%) of the kindreds; the diagnoses in these individuals included LQT3 syndrome, Brugada syndrome, and/or sinus node dysfunction (see 608567). In vitro functional characterization of E1784K channels compared to properties reported for other LQT3 variants suggested that a negative shift of steady-state Na channel inactivation and enhanced tonic block in response to Na channel blockers confer an additional Brugada syndrome/sinus node dysfunction phenotype, and further indicated that class IC drugs should be avoided in patients with Na channels displaying these behaviors.

In a large Finnish family with atrial fibrillation (AF) and conduction defects (ATFB10; 614022), Laitinen-Forsblom et al. (2006) analyzed the SCN5A gene and identified a heterozygous missense mutation (600163.0034) that segregated with disease and was not found in more than 370 control chromosomes.

Ellinor et al. (2008) analyzed the SCN5A gene in 57 probands with a familial history of isolated or 'lone' atrial fibrillation and identified heterozygosity for a missense mutation (600163.0041) in a 45-year-old male proband and his affected father. The authors concluded that SCN5A gene was not a major cause of familial AF.

Darbar et al. (2008) analyzed the SCN5A gene in 375 probands with AF, including 118 with lone AF, which was defined as AF occurring in individuals less than 65 years of age who did not have hypertension, overt structural heart disease, or thyroid dysfunction. The authors identified 8 heterozygous variants in 10 probands that were not found in 360 age-, sex-, and ethnicity-matched controls (see, e.g., 600163.0042-600163.0045). In addition, 11 previously reported rare nonsynonymous coding region variants were identified in 12 probands (see, e.g., 600163.0033), and 3 known common nonsynonymous SCN5A polymorphisms were also identified in the AF cohort (see, e.g., 600163.0024 and 600163.0031). Darbar et al. (2008) stated that in their study, nearly 6% of AF probands carried heterozygous mutations or rare variants in the SCN5A gene.

In affected members of 2 unrelated families with CMD and conduction system disease, Hershberger et al. (2008) identified heterozygosity for 2 different missense mutations in the SCN5A gene, R222Q (600163.0046) and I1835T (600163.0047), respectively. Cheng et al. (2010) restudied the 2 families, noting that all affected individuals were also either homozygous or heterozygous for the SCN5A common polymorphism, H558R (600163.0031). Whole-cell voltage clamp studies in HEK293 cells using the Q1077del background, which is the more abundant alternatively spliced SCN5A transcript present in human hearts (65%), showed that sodium current densities of the R222Q and I1835T mutants were not different from wildtype, but the combined variants R222Q/H558R and I1835T/H558R caused approximately 35% and 30% reduction, respectively, and each showed slower recovery from inactivation than wildtype. With the Q1077del background, R222Q and R222Q/H558R variants also exhibited a significant negative shift in both activation and inactivation, whereas I1835T/H558R showed a significant negative shift in inactivation that tended to decrease window current. In contrast, expression in the Q1077 background showed no changes in peak sodium current densities, decay, or recovery from inactivation for R222Q/H558R or I1835T/H558R. Cheng et al. (2010) concluded that CMD-associated SCN5A rare variants perturb the SCN5A biophysical phenotype that is modulated by SCN5A common variants.

In 3 unrelated families with multifocal ectopic Purkinje-related premature contractions and dilated cardiomyopathy, Laurent et al. (2012) identified heterozygosity for the R222Q mutation in the SCN5A gene, which was fully penetrant and strictly segregated with the cardiac phenotype in each family. Laurent et al. (2012) stated that the R222Q effects that they observed on channel parameters were similar to those measured by Cheng et al. (2010); in addition, they noted that the effects were intermediate in the heterozygous state and also impaired the window current, which is crucial during the plateau phase of the action potential. In vitro studies recapitulated the normalization of the ventricular action potentials in the presence of quinidine.

In affected members of a 3-generation Canadian family with CMD and junctional escape ventricular capture bigeminy, Nair et al. (2012) identified the R222Q mutation in the SCN5A gene. Heterologous expression studies revealed a unique biophysical phenotype of R222Q channels in which an approximately 10-mV leftward shift in the sodium current steady-state activation curve occurs without corresponding shifts in steady-state inactivation at cardiomyocyte resting membrane-potential voltages. Nair et al. (2012) noted that the absence of H558R in these patients established that the H558R polymorphism is not required for the induction of cardiomyopathy in patients carrying the R222Q mutation.

In 16 affected members over 3 generations of a large kindred with CMD and multiple arrhythmias, including premature ventricular complexes (PVCs) of variable morphologies, Mann et al. (2012) identified heterozygosity for the R222Q mutation in the SCN5A gene. The mutation was also identified in 1 clinically unaffected family member, a 56-year-old man with a normal EKG and echocardiogram. None of the R222Q carriers had the common SCN5A variant, H558R.

O'Neill et al. (2022) studied the effects of 50 previously published, functionally characterized missense variants in the SCN5A gene. Based on their effects on peak currents, variants were divided into loss-of-function (less than 10% of wildtype peak current, 35 variants) and partial loss-of-function (10-50% of wildtype peak current, 15 variants). Using cell lines created to study the effects of the variants in heterozygous coexpression with wildtype SCN5A, the authors found that 32 of 35 loss-of-function variants and 6 of 15 partial loss-of-function variants showed a reduction to less than 75% of wildtype-alone peak current, demonstrating evidence of dominant-negative effects. Using data from a published consortia and gnomAD, they found that patients with dominant-negative variants were 2.7 times more likely to present with Brugada syndrome than individuals with putative haploinsufficient variants (p = 0.019).

Associations Pending Confirmation

For discussion of a possible association between variants in the SCN5A, SCN10A (604427), and HEY2 (604674) genes and Brugada syndrome, see 601144.


Genotype/Phenotype Correlations

Westenskow et al. (2004) analyzed the KCNQ1, KCNH2, SCN5A, KCNE1, and KCNE2 genes in 252 probands with long QT syndrome and identified 19 with biallelic mutations in LQTS genes, of whom 18 were either compound (monogenic) or double (digenic) heterozygotes and 1 was a homozygote. They also identified 1 patient who had triallelic digenic mutations (see 152427.0021). Compared with probands who had 1 or no identified mutation, probands with 2 mutations had longer QTc intervals (p less than 0.001) and were 3.5-fold more likely to undergo cardiac arrest (p less than 0.01). All 20 probands with 2 mutations had experienced cardiac events. Westenskow et al. (2004) concluded that biallelic mono- or digenic mutations (which the authors termed 'compound mutations') cause a severe phenotype and are relatively common in long QT syndrome. The authors noted that these findings support the concept of arrhythmia risk as a multi-hit process and suggested that genotype can be used to predict risk.

Niu et al. (2006) analyzed the SCN5A gene in 17 members of a 4-generation Han Chinese family with apparent autosomal dominant inheritance of cardiac arrhythmias and sudden death. All affected individuals were heterozygous for a nonsense mutation in the SCN5A gene (W1421X; 600163.0036), and 1 unaffected individual was compound heterozygous for the W1421X mutation and R1193Q (600163.0023). Niu et al. (2006) suggested that the R1193Q mutation, which results in a gain of sodium channel function, may compensate for the deleterious effects of W1421X.


Animal Model

Nuyens et al. (2001) reported that mice heterozygous for a knockin KPQ deletion (600163.0001) of the Scn5a gene showed the essential features of LQT3 and spontaneously developed life-threatening polymorphous ventricular arrhythmias. Sudden accelerations in heart rate or premature beats caused lengthening of the action potential with early after-depolarization and triggered arrhythmias in mice heterozygous for the deletion. Adrenergic agonists normalized the response to rate acceleration in vitro and suppressed arrhythmias upon premature stimulation in vivo. These results showed the possible risk of sudden heart rate accelerations. The heterozygous knockin mouse with its predisposition for pacing-induced arrhythmia might be a useful model for the development of new treatments for the LQT3 syndrome.

Papadatos et al. (2002) showed that disruption of the mouse Scn5a gene caused intrauterine lethality in homozygotes with severe defects in ventricular morphogenesis, whereas heterozygotes showed normal survival. Whole-cell patch-clamp analyses of isolated ventricular myocytes from adult Scn5a +/- mice demonstrated a reduction of approximately 50% in sodium conductance. Scn5a +/- hearts had several defects, including impaired atrioventricular conduction, delayed intramyocardial conduction, increased ventricular refractoriness, and ventricular tachycardia with characteristics of reentrant excitation. These findings reconciled reduced activity of the cardiac sodium channel leading to slowed conduction with several apparently diverse clinical phenotypes, providing a model for the detailed analysis of the pathophysiology of arrhythmias. Noble (2002) commented that detailed understanding of the mechanisms of cardiac arrhythmia at all relevant levels is important to the design of therapeutic programs, and cited the work of Papadatos et al. (2002) as an important step.


ALLELIC VARIANTS 48 Selected Examples):

.0001   LONG QT SYNDROME 3

SCN5A, 9-BP DEL, NT4661
SNP: rs397514251, ClinVar: RCV000009962, RCV000183165, RCV002336463, RCV003318368

In 2 apparently unrelated kindreds with chromosome 3-linked LQT syndrome (LQT3; 603830), Wang et al. (1995) found deletion of 9 basepairs beginning at nucleotide 4661 of their cDNA for SCN5A. The deletion, which was detected by sequencing an aberrant SSCP conformer, resulted in deletion of lys-pro-gln (KPQ), which are 3 conserved amino acids in the cytoplasmic linker between domains III and IV of the channel protein. The 3 amino acids involved in the in-frame deletion are lys1505, pro1506, and gln1507. The effect of this mutation on membrane depolarization was studied by Bennett et al. (1995).

Clancy and Rudy (1999) developed a model representative of the behavior of the sodium channel in heart muscle cells using a single-channel-based Markov model approach. They showed that the delta-KPQ mutant form of the sodium channel stays open for too long, causing an overlarge inward current of sodium which gives rise to arrhythmia. This model view was corroborated by experiments recording actual sodium currents in cardiac muscle cells.


.0002   LONG QT SYNDROME 3

SCN5A, ARG1644HIS
SNP: rs28937316, gnomAD: rs28937316, ClinVar: RCV000009963, RCV000058726, RCV000183090, RCV000246905, RCV002307360, RCV003335025, RCV003591624

In a mother and son with the long QT syndrome (LQT3; 603830), Wang et al. (1995) demonstrated a CGC-to-CAC mutation in codon 1644, resulting in the substitution of a highly conserved arginine residue by histidine.


.0003   LONG QT SYNDROME 3

SCN5A, ASN1325SER
SNP: rs28937317, ClinVar: RCV000009964, RCV000058618, RCV002354154, RCV003234898

In a family in which members of 4 generations had been affected by the long QT syndrome (LQT3; 603830), Wang et al. (1995) found an AAT-to-AGT transition in codon 1325, predicted to cause substitution of a highly conserved asparagine residue by a serine residue.


.0004   BRUGADA SYNDROME 1

SCN5A, ARG1232TRP AND THR1620MET
SNP: rs199473207, rs199473282, gnomAD: rs199473282, ClinVar: RCV000009965, RCV000058588, RCV000058715, RCV000144030, RCV000144031, RCV000183042, RCV001836727, RCV001842342, RCV001842371, RCV002345369, RCV002477202

In affected members of a family with Brugada syndrome (BRGDA1; 601144), a distinct form of idiopathic ventricular fibrillation, Chen et al. (1998) found an arg1232-to-trp (R1232W) and a thr1620-to-met (T1620M) mutation on the same chromosome with no mutation in the other chromosome, suggesting to them that IVF in this family was inherited as an autosomal dominant trait. The presence of both normal and mutated sodium channels in the same tissue would promote heterogeneity of the refractory period, a well established mechanism in arrhythmogenesis, and therefore may be the underlying molecular defect that causes re-entrant arrhythmia in this family. The potential contribution of R1232W and T1620M mutations to the mechanism of IVF was determined by heterologous expression in Xenopus oocytes. They found that sodium channels with the missense mutation recovered from inactivation more rapidly than normal, indicating that IVF with right bundle branch block (RBBB) and ST segment elevation is a defect distinct from long QT syndrome. When studied alone, the R1232W mutant behaved most like normal channels, whereas the T1620M mutant closely followed the kinetic pattern of the double mutant. This indicated that T1620M is the mutation probably responsible for the IVF phenotype in this kindred and that R1232W could be a rare polymorphism. In summary, biophysical analysis of the 2 missense mutations in SCN5A showed a shift in the voltage dependence of steady-state inactivation toward more positive potentials associated with a 25 to 30% acceleration in recovery time from inactivation at potentials near -80mV.

Commenting that studies of the thr1620-to-met mutant by Chen et al. (1998) revealed an abnormal electrophysiologic profile at room temperature that did not adequately explain the ECG signature of Brugada syndrome, Dumaine et al. (1999) undertook a more detailed electrophysiologic study of the thr1620-to-met mutant protein. Dumaine et al. (1999) expressed the mutant protein in a mammalian cell line and employed a patch-clamp technique to study current kinetics at 32 degrees C. The results indicated that current decay kinetics were faster in mutant than in wildtype channels at this temperature and that recovery from inactivation was slower, with a significant shift in steady-state activation. These findings provided an explanation for the ECG features of Brugada syndrome and represented the first illustration of a cardiac sodium channel mutation in which arrhythmogenicity is revealed only at temperatures approaching the physiologic range.

Voltage-gated sodium channels are multimeric structures consisting of a large, heavily glycosylated alpha subunit and 1 or 2 smaller beta subunits. The beta subunits are thought necessary for normal gating function. In brain and skeletal muscle, the beta-1 subunit (600235) accelerates sodium channel inactivation. Makita et al. (2000) characterized the functional roles of the auxiliary beta subunit by coexpression of the beta subunit with either wildtype SCN5A or SCN5A carrying the heterologously expressed T1620M mutation in Xenopus oocytes. The midpoint of steady-state inactivation was significantly shifted to positive potentials in the T1620M alpha/beta-1 channel, with an acceleration in recovery from inactivation when compared to other channels. Makita et al. (2000) therefore suggested that coexpression of T1620M alpha/beta-1 subunits exposed a significant electrophysiologic deficit that may predispose to ventricular fibrillation. Expression of both normal and mutant channels, as in the hearts of patients with Brugada syndrome, would promote heterogeneity of the refractory period in their myocardium, which serves as an ideal electrical substrate for reentrant arrhythmia.


.0005   BRUGADA SYNDROME 1

SCN5A, IVS7DS, 2-BP INS
SNP: rs397514252, ClinVar: RCV000009966

In affected members of a family with idiopathic ventricular fibrillation with right bundle branch block (RBBB) and elevated ST segments, a disorder known as Brugada syndrome (BRGDA1; 601144), Chen et al. (1998) found an insertion of 2 nucleotides, AA, after the first 4 nucleotides (gtaa) in the splice donor sequence of intron 7 of the SCN5A gene. The functional consequences of this splicing mutation were not established.


.0006   BRUGADA SYNDROME 1

SCN5A, 1-BP DEL, VAL1398TER
SNP: rs397514446, ClinVar: RCV000009967, RCV003542270

In affected members of a family with idiopathic ventricular fibrillation characterized by RBBB and elevated ST segments, a disorder known as Brugada syndrome (BRGDA1; 601144), Chen et al. (1998) found a deletion of a single nucleotide (A) from codon 1397 of the SCN5A gene. This deletion resulted in an in-frame stop at codon 1398 (normally val). The resulting truncation eliminated DIII/S6, DIV/S1-S6, and the C-terminal portion of the cardiac sodium channel.


.0007   LONG QT SYNDROME 3

LONG QT SYNDROME 3/6, DIGENIC, INCLUDED
SCN5A, ARG1623GLN
SNP: rs137854600, ClinVar: RCV000009970, RCV000009971, RCV000058716, RCV001588806

In an infant Japanese girl with a severe form of long QT syndrome (LQT3; 603830), Makita et al. (1998) identified a de novo missense mutation, arg1623 to gln (R1623Q), in the S4 segment of domain 4 of the SCN5A gene. When expressed in oocytes, mutant sodium channels exhibited only minor abnormalities in channel activation, but in contrast to 3 previously characterized LQT3 mutations, had significantly delayed macroscopic inactivation. Single channel analysis revealed that R1623Q channels had significantly prolonged open times with bursting behavior, suggesting a novel mechanism of pathophysiology in Na(+) channel-linked long QT syndrome.

Kambouris et al. (2000) reported that the R1623Q mutation imparts unusual lidocaine sensitivity to the sodium channel that is attributable to its altered functional behavior. Studies of lidocaine on individual R1623Q single-channel openings indicated that the open-time distribution was not changed, indicating the drug does not block the open pore as proposed previously. Rather, the mutant channels have a propensity to inactivate without ever opening ('closed-state inactivation'), and lidocaine augments this gating behavior. An allosteric gating model incorporating closed-state inactivation recapitulated the effects of lidocaine on the pathologic sodium current. These findings explained the unusual drug sensitivity of R1623Q and provided a general and unanticipated mechanism for understanding how sodium channel-blocking agents may suppress the pathologic, sustained sodium current induced by LQT3 mutations.

In a male infant diagnosed with ventricular arrhythmias and cardiac decompensation in utero at 28 weeks' gestation and with long QT syndrome at birth, Miller et al. (2004) identified heterozygosity for the R1623Q mutation. The mother had no ECG abnormalities, but a previous and a subsequent pregnancy both ended in stillbirth at 7 months. Initial studies detected no genetic abnormality, but a sensitive restriction enzyme-based assay revealed a small percentage (8 to 10%) of cells harboring the mutation in the mother's blood, skin, and buccal mucosa; R1623Q was also identified in cord blood from the third fetus. Miller et al. (2004) concluded that recurrent late-term fetal loss or sudden infant death can result from unsuspected parental mosaicism for LQT-associated mutations.

In a 1-month-old male infant who had syncope, torsade de pointes, cardiac arrest, and a QTc of 460 ms, Millat et al. (2006) identified biallelic digenic mutations: a 4868G-A transition in exon 28 of the SCN5A gene resulting in the R1623Q substitution; and a missense mutation in the KCNE2 gene (F60L; 603796.0005).


.0008   LONG QT SYNDROME 3

BRUGADA SYNDROME 1, INCLUDED
SINUS NODE DISEASE, INCLUDED
SCN5A, GLU1784LYS
SNP: rs137854601, gnomAD: rs137854601, ClinVar: RCV000009972, RCV000009973, RCV000009974, RCV000058773, RCV000183117, RCV000208193, RCV000245905, RCV000588022, RCV000824758, RCV001813738, RCV003591625

Wei et al. (1999) described a family in which the 13-year-old proband died suddenly at rest with no antecedent illness and no significant findings at postmortem. Her father had sinus bradycardia with occasional sinus pauses and ventricular ectopy together with profound prolongation of his QT interval (QTc = 527 ms) (see LQT3; 603830). He experienced only occasional light-headedness. Other family members experienced occasional syncope and had sinus bradycardia and prolonged QT intervals on their ECGs. In those individuals with prolonged QT intervals, SSCP analysis detected an aberrant conformer in the coding region of the SCN5A gene corresponding to the C terminus. Nucleotide sequencing revealed a G-to-A transition at codon 1784, resulting in a glu-to-lys substitution. This mutation occurs at a highly conserved residue in most voltage-gated sodium channels in most animals, including invertebrates. When the mutation was expressed in Xenopus oocytes, a defect in channel inactivation was demonstrated in the form of a small residual steady state current throughout prolonged depolarization. Wei et al. (1999) explored this further by engineering SCN5A constructs with amino acid substitutions at other positions in the C terminus. All exhibited similar electrophysiologic phenotypes, suggesting that heterozygous charge-neutralizing amino acid substitution at this site causes an allosteric effect on sodium channel gating, resulting in delayed myocardial repolarization. This provided a novel mechanism for LQT3.

Makita et al. (2008) genotyped 66 members of 44 LQT3 families of multiple ethnicities and identified the E1784K mutation in 41 individuals from 15 (34%) of the kindreds, including the family previously reported by Wei et al. (1999); the diagnoses in these individuals included LQT3 syndrome, Brugada syndrome (BRGDA1; 601144), and/or sinus node disease (see 608567). Heterologously expressed E1784K channels showed a 15.0-mV negative shift in the voltage dependence of Na channel inactivation and a 7.5-fold increase in flecainide affinity for resting-state channels, properties also seen with other LQT3 mutations associated with a mixed clinical phenotype. Furthermore, these properties were absent in Na channels harboring the T1304M mutation, which is associated with LQT3 without a mixed clinical phenotype. Makita et al. (2008) suggested that a negative shift of steady-state Na channel inactivation and enhanced tonic block by class IC drugs represent common biophysical mechanisms underlying the phenotypic overlap of LQT3 and Brugada syndromes, and further indicated that class IC drugs should be avoided in patients with Na channels displaying these behaviors.


.0009   PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, IVS22DS, T-C, +2
SNP: rs397514447, ClinVar: RCV000009975, RCV003541536

In a large French family with progressive heart block (PFHB1A; 113900), Schott et al. (1999) identified a T-to-C transition in the highly conserved +2 donor splice site of intron 22 of the SCN5A gene. The abnormal transcript predicted in-frame skipping of exon 22 and an impaired gene product lacking the voltage-sensitive DIIIS4 segment.


.0010   HEART BLOCK, NONPROGRESSIVE

SCN5A, 1-BP DEL, 5280G
SNP: rs397514448, ClinVar: RCV000009976

In a Dutch family with asymptomatic first-degree atrioventricular block associated with right bundle branch block from birth, without apparent progression (see 113900), Schott et al. (1999) identified a 1-bp deletion (G) at nucleotide 5280 of the SCN5A gene, resulting in a frameshift predicted to cause a premature stop codon.


.0011   BRUGADA SYNDROME 1

SCN5A, ARG1512TRP
SNP: rs137854602, gnomAD: rs137854602, ClinVar: RCV000009977, RCV000058688, RCV000157490, RCV000222521, RCV001841232

In the screening of SCN5A in 6 individuals with Brugada syndrome (BRGDA1; 601144), Rook et al. (1999) found missense mutations in the coding region of the gene in 2: arg1512 to trp (R1512W) in the DIII-DIV cytoplasmic linker, and ala1924 to thr (A1924T; 600163.0012) in the C-terminal cytoplasmic domain. In 2 other patients mutations were detected near intron/exon junctions. To assess the functional consequences of the R1512W and A1924T mutations, wildtype and mutant sodium channel proteins were expressed in Xenopus oocytes. Both missense mutations affected channel function and seemed to be associated with an increase in inward sodium current during the action potential upstroke.


.0012   BRUGADA SYNDROME 1

SCN5A, ALA1924THR
SNP: rs137854603, gnomAD: rs137854603, ClinVar: RCV000009978, RCV000058806, RCV000420298, RCV001841233, RCV002251424

For discussion of the ala1924-to-thr (A1924T) substitution in the SCN5A gene that was found in compound heterozygous state in 2 patients with Brugada syndrome (BRGDA1; 601144) by Rook et al. (1999), see 600163.0011.


.0013   LONG QT SYNDROME 3

BRUGADA SYNDROME 1, INCLUDED
SCN5A, 3-BP INS, 5537TGA
SNP: rs397514449, ClinVar: RCV000009979, RCV000009980, RCV001530164

In a large Dutch family with electrocardiographic features both of long QT syndrome (LQT3; 603830) and Brugada syndrome (BRGDA1; 601144), Bezzina et al. (1999) demonstrated a 3-bp insertion at nucleotide position 5537 of the SCN5A gene, predicted to cause insertion of an aspartic acid residue at amino acid position 1795 (1795insD) in the C-terminal domain of the protein. Expression of this mutant channel protein in Xenopus oocytes permitted characterization of defects in channel activation and inactivation when compared to a wildtype control. These defects were predicted to cause a reduction in sodium flux during the upstroke of the cardiac action potential.

The co-occurrence of Brugada syndrome and long QT syndrome in this family was paradoxical, since LQT3 is associated with activating SCN5A mutations and Brugada syndrome with inactivating mutations. Clancy and Rudy (2002) modeled the cellular effects of the 1795insD mutation in a virtual transgenic cell. Since ion channel proteins are expressed nonuniformly throughout the myocardium, there is an intrinsic electrophysiologic heterogeneity. The authors demonstrated that the interplay between this underlying myocardial electrophysiologic heterogeneity and the mutation-induced changes in cardiac sodium channel function provided the substrate for both ST segment elevation (in Brugada syndrome) and QT prolongation (LQT3) in a rate-dependent manner.


.0014   VENTRICULAR FIBRILLATION, PAROXYSMAL FAMILIAL, 1 (1 patient)

SCN5A, SER1710LEU
SNP: rs137854604, gnomAD: rs137854604, ClinVar: RCV000009981, RCV000058743, RCV000183102, RCV000197520, RCV000246596, RCV001841234, RCV002504774

Akai et al. (2000) screened 25 Japanese patients with idiopathic ventricular fibrillation (VF1; 603829). The diagnosis was based on the occurrence of at least one episode of syncope and/or cardiac arrest and documentation of ventricular fibrillation. Structural heart disorders were excluded. Eighteen patients were diagnosed as Brugada syndrome. The authors identified a heterozygous ser1710-to-leu missense mutation of the SCN5A gene in a 39-year-old man who was admitted to the hospital for recurrent syncope and suffered an episode of spontaneous ventricular fibrillation while hospitalized. An implanted cardiac defibrillator was successful in preventing further attacks of palpitation or syncope. Brugada syndrome was not present. The paternal grandfather and a paternal uncle had died suddenly in their sixth decade of unknown cause; the parents and sibs were asymptomatic.


.0015   LONG QT SYNDROME 3

SCN5A, SER941ASN
SNP: rs137854605, ClinVar: RCV000009982, RCV003542271

Schwartz et al. (2000) described an infant who nearly died of SIDS (272120), whose parents had normal QT intervals and in whom the long QT syndrome (LQT3; 603830) was diagnosed with identification of a spontaneous mutation of the SCN5A gene: a change of codon 941 from TCC (serine) to AAC (asparagine). The patient had all the classic features of near-SIDS. Before the episode, the infant appeared to be in perfect health. His age at the time of the episode (7 weeks) was within the age range of 5 to 12 weeks during which the incidence of SIDS peaks. The parents found him cyanotic, apneic, and pulseless. Ventricular fibrillation was documented in an emergency room; this point is important given the frequent statements that ventricular arrhythmias have not been recorded in infants at risk for SIDS. Had the infant died--an outcome that was almost a certainty in the absence of cardioversion--the absence of an electrocardiogram and the normal QT intervals of both parents would have eliminated suspicion of the long QT syndrome and would have prompted a diagnosis of SIDS.


.0016   CARDIAC CONDUCTION DEFECT, NONPROGRESSIVE

SCN5A, GLY514CYS
SNP: rs137854606, ClinVar: RCV000009984, RCV000058427

Tan et al. (2001) studied a family who came to medical attention when the proband, a 3-year-old girl, experienced episodes of fainting during a febrile illness. Her 12-lead ECG showed characteristics of slow conduction throughout the atria and ventricles, including broad P waves, PR interval prolongation, and a wide QRS complex (see 113900). Continuous monitoring revealed episodes of severe bradycardia (25 beats/minute). During these slow periods the cardiac rhythm was maintained by infrequent atrioventricular nodal 'escape' impulses. Conduction disturbance persisted after the febrile illness, but there was no evidence of structural heart disease or systemic diseases associated with conduction defects in children. Therapeutic intervention with a dual-chamber pacemaker was initially limited by inability to pace the atrium (maximal stimulus: 10 V, 1 ms); however, this difficulty resolved with 1 week of empiric steroid treatment. During the 4 years following diagnosis, the patient continuously required dual-chamber pacing. The proband's 6-year-old sister was similarly affected and required pacemaker implantation, with episodes of noncapture that reproducibly resolved with corticosteroid therapy. Three other family members with no structural heart disease had ECG evidence of conduction slowing (prolonged PR and QRS intervals), but did not experience bradycardia or require pacemaker implantation. All affected family members had a G-to-T transition in the first nucleotide of codon 514 in exon 12 of the SCN5A gene resulting in the replacement of glycine by cysteine (G514C). Biophysical characterization of the mutant channel showed that there were abnormalities in voltage-dependent gating behavior that could be partially corrected by dexamethasone, consistent with the salutary effects of glucocorticoids on the clinical phenotype. Computational analysis predicts that the gating defects of G514C selectively slow myocardial conduction, but do not provoke the rapid cardiac arrhythmias associated previously with SCN5A mutations.


.0017   PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, ASP1595ASN
SNP: rs137854607, ClinVar: RCV000009983, RCV000058705, RCV000183084, RCV001329632, RCV003407312

Wang et al. (2002) reported a family in which the proband had presented with first-degree atrioventricular block at the age of 9, progressing to complete AV block by the age of 20 (PFHB1A; 113900). The proband's sister and father had electrocardiographic evidence of right bundle branch block and left axis deviation with normal PR intervals. The corrected QT interval was normal (less than 420 ms) in all 3 individuals. Sequencing of the coding region of SCN5A revealed a G-to-A mutation at nucleotide position 4783, which replaced an aspartic acid residue at amino acid position 1595 with asparagine (D1595N). The G4783A mutation was engineered into a recombinant human heart sodium channel and transiently coexpressed with human sodium channel beta-1 subunit (600760) in a cultured mammalian cell line (tsA201). Functional characterization using a patch-clamp technique revealed a significant defect in the kinetics of fast-channel inactivation distinct from those of SCN5A mutations reported in LQT3 (603830). The authors considered this a plausible mechanism for the observed conduction system disease in this family.


.0018   PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, GLN298SER
SNP: rs137854608, gnomAD: rs137854608, ClinVar: RCV000009985, RCV000058858, RCV000151803, RCV000415287, RCV001149035, RCV001149036, RCV001149037, RCV001149038, RCV001149039, RCV001841235, RCV002482852, RCV003137510

Wang et al. (2002) reported a child in whom second-degree atrioventricular block had been diagnosed at the age of 6, progressing to complete atrioventricular block by the age of 12 (PFHB1A; 113900). The child's mother had a normal electrocardiogram and the father declined testing. There was no family history of sudden death. Sequencing of the coding region of SCN5A revealed a G-to-A mutation at nucleotide position 892 that replaced a glycine residue at amino acid position 298 with serine (G298S). The G892A mutation was engineered into a recombinant human heart sodium channel and transiently coexpressed with human sodium channel beta-1 subunit (600760) in a cultured mammalian cell line (tsA201). Functional characterization using a patch-clamp technique revealed a significant defect in the kinetics of fast-channel inactivation distinct from those of SCN5A mutations reported in LQT3 (603830). The authors considered this a plausible mechanism for the observed conduction system disease in this family.


.0019   LONG QT SYNDROME 3

SCN5A, ALA997SER
SNP: rs137854609, gnomAD: rs137854609, ClinVar: RCV000009986, RCV000058542, RCV000183020, RCV002504775, RCV003415682, RCV003591626

In a 6-week-old male infant who died of SIDS (272120), Ackerman et al. (2001) identified a heterozygous G-to-T transversion in the SCN5A gene, resulting in an ala997-to-ser substitution. The mutation was not detected in 800 control alleles. Ackerman et al. (2001) determined that amino acid 997 is located in the cytoplasmic connector between the second and third domains of the sodium channel and is highly conserved across species. They demonstrated that the mutant SCN5A channel expressed a sodium current characterized by slower decay and a 2- to 3-fold increase in late sodium current.


.0020   LONG QT SYNDROME 3

SCN5A, ARG1826HIS
SNP: rs137854610, gnomAD: rs137854610, ClinVar: RCV000009987, RCV000058786, RCV000148848, RCV000154827, RCV000619902, RCV000766811, RCV001841236, RCV002476953

In a 42-day-old male infant who died of possible SIDS (272120), Ackerman et al. (2001) identified a heterozygous G-to-A replacement in the SCN5A gene, resulting in an arg1826-to-his substitution. The mutation was not detected in 800 control alleles. Ackerman et al. (2001) determined that amino acid 1826 is located in the cytoplasmic C-terminal region of the sodium channel and is highly conserved. They demonstrated that the SCN5A mutant channel expressed a sodium current characterized by slower decay and a 2- to 3-fold increase in late sodium current.


.0021   BRUGADA SYNDROME 1

SCN5A, ARG367HIS
SNP: rs28937318, ClinVar: RCV000009988, RCV000058390, RCV001841237, RCV002426498, RCV003654174

Sudden unexplained nocturnal death syndrome (SUNDS), a disorder found in southeast Asia, is characterized by an abnormal electrocardiogram with ST segment elevation in leads V1 to V3 and sudden death due to ventricular fibrillation, identical to that seen in Brugada syndrome (BRGDA1; 601144). Vatta et al. (2002) found mutations in the SCN5A gene in 3 of 10 Asian SUNDS patients. In a sporadic Asian SUNDS patient, the authors identified a 1100G-A transition in SCN5A. The mutation is predicted to result in an arg367-to-his (R367H) substitution, which lies in the first P segment of the pore-lining region between the DIS5 and DIS6 transmembrane segments. In transfected Xenopus oocytes, the R367H mutant channel did not express any current. The authors hypothesized that the likely effect of this mutation is to depress peak current due to the loss of one functional allele.


.0022   BRUGADA SYNDROME 1

SCN5A, ALA735VAL
SNP: rs137854611, gnomAD: rs137854611, ClinVar: RCV000009989, RCV000058488, RCV003591627, RCV003654175

In a family with SUNDS, a disorder identical to Brugada syndrome (BRGDA1; 601144), that exhibited autosomal dominant inheritance, Vatta et al. (2002) identified among affected members a 2204C-T transition, which is predicted to result in an ala735-to-val (A735V) substitution. The mutation lies in the first transmembrane segment of domain II, (DIIS1), close to the first extracellular loop between DIIS1 and DIIS2. In transfected Xenopus oocytes, the A735V mutant expressed currents with steady-state activation voltage shifted to more positive potentials and exhibited reduced sodium channel current at the end of phase I of the action potential.


.0023   BRUGADA SYNDROME 1

LONG QT SYNDROME 3, ACQUIRED, SUSCEPTIBILITY TO, INCLUDED
SCN5A, ARG1193GLN
SNP: rs41261344, gnomAD: rs41261344, ClinVar: RCV000009990, RCV000009991, RCV000058578, RCV000154828, RCV000157488, RCV000171819, RCV000252422, RCV000755697, RCV001147624, RCV001147625, RCV001147626, RCV001147627, RCV001841238, RCV002476954, RCV003149566

In a pair of Japanese dizygotic twins, one of whom died at 4 months of SUNDS, a disorder identical to Brugada syndrome (BRGDA1; 601144), Vatta et al. (2002) identified a 3575G-A transition in exon 20 of the SCN5A gene, predicted to result in an arg1192-to-gln (R1192Q) substitution in Domain III. In transfected Xenopus oocytes, the mutation accelerated the inactivation of the sodium channel current and exhibited reduced sodium channel current at the end of phase I of the action potential. Wang (2005) stated that this variant was mislabeled in the Vatta et al. (2002) report and should be designated R1993Q.

In an 82-year-old Caucasian male who developed long QT syndrome after the administration of D-sotolol or quinidine (see LQT3, 603830), Wang et al. (2004) identified heterozygosity for the R1993Q mutation in the SCN5A gene. The mutation was found in 4 of 2,087 predominantly Caucasian controls (0.2%). Electrophysiologic studies showed that mutant R1193Q channels destabilize inactivation gating and generate a persistent, nonactivating current that is expected to prolong the cardiac action potential duration, leading to LQT syndrome; single channel recording revealed the molecular mechanism to be frequent, dispersed reopening of the channels. The patient also carried the H558R SCN5A variant (600163.0031), but due to a lack of family members, it could not be determined whether H558R was in cis or trans with R1993Q.

Hwang et al. (2005) found the R1993Q mutation in 11 of 94 (12%) randomly selected Han Chinese individuals and concluded that the variant is a common polymorphism in this population. None of the carriers had electrocardiographic signs of Brugada syndrome, although 1 had a prolonged QTc interval (472 ms) and another, who was homozygous for the mutation, had a borderline long QTc (437 ms).

In an asymptomatic 73-year-old male member of a 4-generation Han Chinese family with autosomal dominant cardiac arrhythmias and sudden death, Niu et al. (2006) identified compound heterozygosity for R1193Q and a nonsense mutation in the SCN5A gene (W1421X; 600163.0036). Niu et al. (2006) suggested that the R1193Q mutation, which results in a gain of sodium channel function, may compensate for the deleterious effects of W1421X. Haplotype analysis of an asymptomatic daughter-in-law and 2 asymptomatic grandchildren who also carried the R1193Q mutation revealed that the children inherited the mutation from their mother rather than their grandfather.


.0024   LONG QT SYNDROME 3, ACQUIRED, SUSCEPTIBILITY TO

SUDDEN INFANT DEATH SYNDROME, INCLUDED
SCN5A, SER1103TYR
SNP: rs7626962, gnomAD: rs7626962, ClinVar: RCV000009992, RCV000009993, RCV000041615, RCV000058563, RCV000204216, RCV000274325, RCV000304064, RCV000363449, RCV000368908, RCV000396768, RCV000621429, RCV000755696, RCV001094834, RCV001841239, RCV002504776, RCV003125829, RCV003149567

Splawski et al. (2002) screened DNA samples from individuals with nonfamilial cardiac arrhythmias and identified a C-to-A transversion in the SCN5A gene leading to a ser1103-to-tyr (S1103Y) substitution 1 patient. Ackerman et al. (2004) noted that the variant was originally published as SER1102TYR from numbering based on the 2,015 amino acid alternatively spliced transcript. Subsequently, numbering was revised to account for the full-length 2,016 amino acid transcript. Serine-1103 is a conserved residue located in the intracellular sequences that link domains II and III of the channel. The proband had idiopathic dilated cardiomyopathy and hypokalemia and developed prolonged QT and torsade de pointes ventricular tachycardia while on amiodarone. Splawski et al. (2002) determined that the Y1103 allele is present in 19.2% of West Africans and Caribbeans and in 13.2% of African Americans. The Y1103 allele was not found in 511 Caucasians or 578 Asians. Splawski et al. (2002) studied 22 African Americans with acquired arrhythmia and 100 population-matched controls. The Y1103 allele was overrepresented among arrhythmia patients, being found in 56.5% of cases and among 13% of controls. The likelihood of displaying signs of arrhythmia in a Y1103 carrier heterozygote or homozygote yielded an odds ratio of 8.7 (95% CI 3.2 to 23.9). The odds ratio was not significantly altered after controlling for age or gender. To determine whether this mutation is an inherited risk factor for arrhythmias, Splawski et al. (2002) examined the extended family of 1 proband. They ascertained and phenotypically characterized 23 members of this kindred. Phenotypic analysis revealed that 11 members of the family had prolonged QT and/or a history of syncope. All 11 phenotypically affected members of this family carried the Y1103 allele (6 were homozygotes and 5 were heterozygotes). Physiologic analysis of the effect of this mutation recorded a small but significant negative shift in the voltage dependence of activation. Splawski et al. (2002) concluded that the Y1103 allele is a common SCN5A variant in Africans and African Americans and causes a small but inherent chronic risk of acquired arrhythmia. In the setting of additional acquired risk factors, including medications, hypokalemia, or structural heart disease, individuals carrying this allele are at increased risk of arrhythmia.

In 3 white sisters and their father, Chen et al. (2002) identified the S1103Y mutation, thus demonstrating that this mutation does exist in the white population. The mutation was associated with a considerable risk of syncope, ventricular arrhythmia, ventricular fibrillation, and sudden death, Each of the 3 sibs was genotyped for 31 'ancestry informative markers' to provide an estimation of biogeographic ancestry on 3 axes: Native American, West African, and European. The maximum likelihood point estimates for each of the sibs were 100% European, 0.0% African, and 0.0% Native American. The proband had a baseline QTc of 520 ms, and developed 2 episodes of syncope at age 49 years. The first episode was triggered by emotion and excitement. The second episode occurred in the setting of amiodarone and low serum potassium, and progressed to ventricular fibrillation and cardiac arrest. She was resuscitated by cardioversion. The second sister had a QTc of 431 ms, and died suddenly at age 44 years when awakening from sleep. The third sister had a QTc of 452 ms, developed 1 episode of syncope at the age of 33 years, and had complained of palpitations all her life. The father died suddenly in his sleep at age 50 years. Family members without S1103Y had a normal QTc.

Plant et al. (2006) screened DNA samples from 133 African American autopsy-confirmed cases of sudden infant death syndrome (SIDS; 272120) and identified 3 that were homozygous for the S1103Y variant. Among 1,056 African American controls, 120 were carriers of the heterozygous genotype, suggesting that infants with 2 copies of S1103Y have a 24-fold increased risk for SIDS. Variant Y1103 channels were found to operate normally under baseline conditions in vitro. Because risk factors for SIDS include apnea and respiratory acidosis, Y1103 and wildtype channels were subjected to lowered intracellular pH; only Y1103 channels developed abnormal function, with late reopenings suppressible by the drug mexiletine. Plant et al. (2006) suggested that the Y1103 variant confers susceptibility to acidosis-induced arrhythmia, a gene-environment interaction.

Darbar et al. (2008) stated that the S1103Y variant was a known common nonsynonymous polymorphism in the SCN5A gene; they detected S1103Y in 1 patient with lone atrial fibrillation and in 5 patients with atrial fibrillation associated with other heart disease, as well as in 15 of 720 control chromosomes, for a minor allele frequency of 0.7%.


.0025   SICK SINUS SYNDROME 1

SCN5A, PRO1298LEU
SNP: rs28937319, ClinVar: RCV000009994, RCV000058612, RCV001841240

In 3 sibs with congenital sick sinus syndrome (SSS1; 608567), Benson et al. (2003) identified compound heterozygosity for 2 mutations in the SCN5A gene. The maternal allele carried a 3893C-T transition, resulting in a pro1298-to-leu (P1298L) change; the paternal allele carried a gly1408-to-arg substitution (600163.0026).


.0026   SICK SINUS SYNDROME 1

BRUGADA SYNDROME 1, INCLUDED
CARDIAC CONDUCTION DEFECT, NONSPECIFIC, INCLUDED
SCN5A, GLY1408ARG
SNP: rs137854612, ClinVar: RCV000009995, RCV000009996, RCV000009997, RCV000058649, RCV000183190, RCV002326672, RCV002496318

In 3 sibs with congenital sick sinus syndrome (SSS1; 608567) with compound heterozygosity for mutation in the SCN5A gene, Benson et al. (2003) found on the paternal allele a 4222G-A transition, resulting in a gly1408-to-arg substitution (G1408R). The maternal allele carried a pro1298-to-leu substitution (600163.0025).

Kyndt et al. (2001) reported the G1408R mutation, which they designated GLY1406ARG, in heterozygous state in a large French family segregating both isolated cardiac conduction defect (see 601144) and Brugada syndrome (BRGDA1; 601144).


.0027   SICK SINUS SYNDROME 1

CARDIOMYOPATHY, DILATED, 1E, INCLUDED
SCN5A, THR220ILE
SNP: rs45620037, gnomAD: rs45620037, ClinVar: RCV000009998, RCV000058832, RCV000148857, RCV000151804, RCV000251727, RCV000258831, RCV000586618, RCV000622951, RCV000678935, RCV001146372, RCV001841241

In a child with congenital sick sinus syndrome (SSS1; 608567), Benson et al. (2003) identified compound heterozygosity for 2 mutations in the SCN5A gene: the paternal allele carried a 659C-T transition, resulting in a thr220-to-ile (T220I) mutation, and the maternal allele carried a 4867C-T transition, resulting in an arg1623-to-ter mutation (R1623X; 600163.0028). The authors noted that an R1623Q mutation (600163.0007) resulting in congenital long QT syndrome-3 (603830) had previously been described.

In a 54-year-old man with dilated cardiomyopathy (CMD1E; 601154), atrial fibrillation, and heart block, Olson et al. (2005) identified heterozygosity for a 659C-T transition in exon 6 of the SCN5A gene, resulting in a thr220-to-ile (T220I) substitution at a highly conserved residue in the transmembrane domain. Coronary artery disease was excluded by angiography; cardiac biopsy showed moderate myocyte hypertrophy and marked interstitial fibrosis. He died 13 years later in severe congestive heart failure. A female first cousin once removed who also carried the mutation was diagnosed at 55 years of age with dilated cardiomyopathy (ejection fraction, 10%) and incomplete bundle branch block; she died 2 years later, also in severe congestive heart failure. Other relatives were reported to have enlarged hearts, but were unavailable for evaluation.


.0028   SICK SINUS SYNDROME 1

SCN5A, ARG1623TER
SNP: rs137854613, gnomAD: rs137854613, ClinVar: RCV000009968, RCV000183087, RCV000465149, RCV000477950, RCV000622049, RCV001841231, RCV002496317, RCV003314549

For discussion of the arg1623-to-ter (R1623X) mutation that was found in compound heterozygous state in a child with congenital sick sinus syndrome (SSS1; 608567) by Benson et al. (2003), see 600163.0027.


.0029   LONG QT SYNDROME 3

SCN5A, TYR1795CYS
SNP: rs137854614, ClinVar: RCV000009969, RCV000058778, RCV001561910, RCV002345237

In a patient with long QT syndrome-3 (LQT3; 603830), Rivolta et al. (2001) identified a tyr1795-to-cys (Y1795C) mutation in the SCN5A gene. The mutation slowed the onset of activation, but did not cause a marked negative shift in the voltage dependence of inactivation or affect the kinetics of the recovery from inactivation. The mutation increased the expression of sustained Na(+) channel activity compared with wildtype channels and promoted entrance into an intermediate or slowly developing inactivated state.


.0030   BRUGADA SYNDROME 1

SCN5A, TYR1795HIS
SNP: rs137854615, ClinVar: RCV000009999, RCV000058777

In a patient with Brugada syndrome (BRGDA1; 601144), Rivolta et al. (2001) identified a tyr1795-to-his (Y1795H) mutation in the SCN5A gene. The mutation accelerated the onset of activation and caused a marked negative shift in the voltage dependence of inactivation. It did not affect the kinetics of the recovery from inactivation. The mutation increased the expression of sustained Na(+) channel activity compared with wildtype channels and promoted entrance into an intermediate or slowly developing inactivated state.


.0031   PROGRESSIVE FAMILIAL HEART BLOCK, TYPE IA

SCN5A, THR512ILE AND HIS558ARG
SNP: rs1805124, rs199473118, gnomAD: rs1805124, rs199473118, ClinVar: RCV000010000, RCV000041604, RCV000058426, RCV000058440, RCV000144029, RCV000251327, RCV000300603, RCV000304709, RCV000339196, RCV000361696, RCV000405409, RCV000406777, RCV000588264, RCV000987225, RCV001841593, RCV002496658, RCV002498340, RCV003125879

In a 2-year-old boy with second-degree atrioventricular conduction block (PFHB1A; 113900) necessitating a pacemaker, Viswanathan et al. (2003) identified a heterozygous 1535C-T transition in the SCN5A gene, resulting in a thr512-to-ile (T512I) substitution. In addition, there was a homozygous 1673A-G transition, resulting in a his558-to-arg (H558R) substitution. H558R (rs1805124) is a polymorphism present in 20% of the population (Yang et al., 2002). One of the patient's alleles contained both T512I and H558R. The patient's father was heterozygous for the H558R substitution, the asymptomatic mother was compound heterozygous for the T512I and H558R substitutions, and 2 sibs were heterozygous for the H558R substitution. Functional expression studies showed that activation and inactivation of wildtype and H558R channels were similar. By contrast, voltage-dependent activation and inactivation of the T512I channel was shifted negatively by 8 to 9 mV and had enhanced slow activation and slower recovery from inactivation compared to the wildtype channel. Studies of the double H558R/T512I channel showed that H558R eliminated the negative shift induced by T512I, but only partially restored the kinetic abnormalities. Viswanathan et al. (2003) suggested that enhanced slow inactivation disproportionately affected Purkinje cells, which have a longer action potential duration and smaller diastolic interval, resulting in slowed atrioventricular conduction.

Darbar et al. (2008) stated that the H558R variant was a known common nonsynonymous polymorphism in the SCN5A gene; they detected H558R in 59 patients with lone atrial fibrillation and in 130 patients with atrial fibrillation associated with other heart disease, as well as in 128 of 720 control chromosomes, for a minor allele frequency of approximately 25%.


.0032   BRUGADA SYNDROME 1

SCN5A, GLY1262SER
SNP: rs137854616, gnomAD: rs137854616, ClinVar: RCV000010001, RCV000058602, RCV000755698, RCV000779407, RCV001146725, RCV001146726, RCV001146727, RCV001753410

Shin et al. (2004) studied a family with 9 members as well as 12 unrelated sporadic cases, all Koreans, diagnosed with Brugada syndrome (BRGDA1; 601144). They identified a novel missense mutation associated with Brugada syndrome in the family: a single-nucleotide substitution of G to A at nucleotide position 3934 in exon 21 of the SCN5A gene that changed glycine-1262 to serine (G1262S) in segment 2 of domain III of the SCN5A protein. Four individuals in the family carried the identical mutation, but none of the 12 sporadic patients did. The mutation was not found in 150 unrelated normal individuals.


.0033   BRUGADA SYNDROME 1

ATRIAL FIBRILLATION, FAMILIAL, 10, INCLUDED
SCN5A, GLU1053LYS
SNP: rs137854617, gnomAD: rs137854617, ClinVar: RCV000010002, RCV000022945, RCV000058552, RCV000469648, RCV000755695, RCV001528558, RCV001841242, RCV002321478

In a patient with Brugada syndrome (BRGDA1; 601144), Mohler et al. (2004) identified a 3157G-A transition in the SCN5A gene resulting in a glu1053-to-lys (E1053K) mutation in the ankyrin-binding motif of the cardiac sodium channel. The mutation abolished binding of Na(v)1.5 to ankyrin-G (600465) and also prevented accumulation of Na(v)1.5 at cell surface sites in ventricular cardiomyocytes.

In a patient with lone atrial fibrillation (ATFB10; 614022), Darbar et al. (2008) identified heterozygosity for the E1053K mutation in the SCN5A gene. The mutation was not found in 720 control alleles.


.0034   CARDIOMYOPATHY, DILATED, 1E

ATRIAL STANDSTILL 1, DIGENIC, INCLUDED
ATRIAL FIBRILLATION, FAMILIAL, 10, INCLUDED
SCN5A, ASP1275ASN
SNP: rs137854618, gnomAD: rs137854618, ClinVar: RCV000010003, RCV000022946, RCV000058604, RCV000114992, RCV000183045, RCV000617238, RCV000656563, RCV002222347, RCV003894798

In a large family reported by Greenlee et al. (1986) with dilated cardiomyopathy with conduction disorder and arrhythmia (CMD1E; 601154), McNair et al. (2004) identified heterozygosity for a 3823G-A mutation in exon 21 of the SCN5A gene, resulting in an asp1275-to-asn (D1275N) substitution and predicting a change of charge within the S2 segment of domain III. The mutation was present in 22 affected family members and was not found in 300 control chromosomes.

Groenewegen et al. (2003) reported the D1275N mutation, coinherited with polymorphisms in the atrial-specific gap junction channel protein connexin-40 (GJA5; 121013), in affected members of a family with atrial standstill (ATRST1; 108770). No member of this family had dilated cardiomyopathy, leading Groenewegen and Wilde (2005) to question whether the D1275N mutation was the primary cause of dilated cardiomyopathy as reported by McNair et al. (2004).

In affected members of a large Finnish family with atrial fibrillation and conduction defects (ATFB10; 614022), Laitinen-Forsblom et al. (2006) identified heterozygosity for the D1275N mutation in the SCN5A gene. The mutation was not found in more than 370 control chromosomes. Echocardiography revealed an enlarged left ventricle with an increased end-diastolic left ventricular diameter in 1 affected individual, and the right ventricle was slightly enlarged in 3 other affected individuals.


.0035   LONG QT SYNDROME 2/3, DIGENIC

SCN5A, ASP1819ASN
SNP: rs137854619, gnomAD: rs137854619, ClinVar: RCV000010005, RCV000058782, RCV000171695, RCV000183199, RCV000987198, RCV001507624, RCV001841243, RCV002345238

In a 41-year-old female who had cardiac arrest due to torsade de pointes triggered by exercise and leading to ventricular fibrillation, and a QTc of 520 ms (see 603830), Millat et al. (2006) identified biallelic digenic mutations: a 5455G-A transition in exon 28 of the SCN5A gene, resulting in an asp1819-to-asn (D1819N) substitution; and a missense mutation in the KCNH2 gene (R100G; 152427.0023).


.0036   BRUGADA SYNDROME 1

SCN5A, TRP1421TER
SNP: rs137854620, ClinVar: RCV000010006

In affected members of a 4-generation Han Chinese family with autosomal dominant cardiac arrhythmias and sudden death (BRGDA1; 601144), Niu et al. (2006) identified heterozygosity for a G-A transition in exon 24 of the SCN5A gene, resulting in a trp1421-to-ter (W1421X) substitution. The mutation was not found in 95 control subjects. An asymptomatic 73-year-old male family member was found to be compound heterozygous for W1421X and the R1993Q mutation (600163.0023). Niu et al. (2006) suggested that R1193Q, which results in a gain of sodium channel function, may compensate for the deleterious effects of W1421X.


.0037   MOVED TO 600163.0027


.0038   CARDIOMYOPATHY, DILATED, 1E

SCN5A, 2-BP INS, NT2550
SNP: rs397514450, ClinVar: RCV000010008, RCV000183154, RCV001842933, RCV002433811, RCV002500544

In a man with dilated cardiomyopathy (CMD1E; 601154) and monomorphic ventricular tachycardia who later developed third-degree heart block requiring pacemaker implantation, Olson et al. (2005) identified heterozygosity for a 2-bp insertion (2550insTG) in exon 17 of the SCN5A gene, resulting in a premature stop codon and a truncated protein. Cardiac biopsy was normal. His father, who carried the mutation, was diagnosed with CMD (ejection fraction, 30%), left bundle branch block, and monomorphic ventricular tachycardia at age 67 years. The mutation was also present in a paternal uncle who had sinus bradycardia, first-degree heart block, and complete left bundle branch block; his paternal grandfather developed congestive heart failure at 50 years of age and died 6 years later, but DNA was unavailable for evaluation.


.0039   CARDIOMYOPATHY, DILATED, 1E

SCN5A, ASP1595HIS
SNP: rs137854607, ClinVar: RCV000010009, RCV000058706, RCV001258074, RCV003156212

In a 7-year-old boy with early manifestations of dilated cardiomyopathy (CMD1E; 601154) including sinus bradycardia, left ventricular dilation, and normal contractile function, Olson et al. (2005) identified heterozygosity for a 4783G-C transversion in exon 27 of the SCN5A gene, resulting in an asp1595-to-his (D1595H) substitution at a highly conserved residue in the transmembrane domain. The mutation was found in DNA from postmortem tissue of a brother who died at 34 years of age with an autopsy diagnosis of cardiomyopathy and only mild coronary artery disease. The mutation was also identified in 2 sibs and a paternal uncle, all of whom had sinus bradycardia, and a paternal aunt with borderline left atrial enlargement. His father, an obligate mutation carrier, had atrial fibrillation and died at 49 years of age from pulmonary embolism; his paternal grandfather, a presumed mutation carrier, developed congestive heart failure at 70 years of age.


.0040   BRUGADA SYNDROME 1

SCN5A, VAL232ILE and LEU1308PHE
SNP: rs41313031, rs45471994, gnomAD: rs41313031, rs45471994, ClinVar: RCV000010010, RCV000058614, RCV000058840, RCV000148841, RCV000148856, RCV000176338, RCV000243761, RCV000246365, RCV000724673, RCV000987205, RCV000987235, RCV001842355, RCV001842410

In a 45-year-old black man with no history of cardiac disease who developed monomorphic wide-complex ventricular tachycardia with right precordial ST segment elevation consistent with Brugada syndrome (BRGDA1; 601144) after the administration of lidocaine, Barajas-Martinez et al. (2008) identified 2 mutations in the SCN5A gene, a G-to-A transition in exon 6 of the SCN5A gene, resulting in a val232-to-ile (V232I) substitution in the C terminus of the transmembrane segment S4 of domain I, and a C-to-T transition in exon 22, resulting in a leu1308-to-phe (L1308F) substitution, in the C terminus of transmembrane segment S4 of domain III. Although L1308F had previously been identified as a polymorphism found mostly in Americans of African descent (Ackerman et al., 2004), Barajas-Martinez et al. (2008) did not find either mutation in over 400 alleles from 200 ethnically matched controls. The patient's parents were unavailable for study, but given the severity of his clinical manifestations, the authors strongly suspected that both mutations were on the same allele (Dumaine, 2009). Using patch-clamp techniques in mammalian TSA201 cells, Barajas-Martinez et al. (2008) observed use-dependent inhibition of I(Na) by lidocaine that was more pronounced in double-mutant channels than in wildtype; the individual mutations produced a much less accentuated effect. The authors concluded that the double mutation in SCN5A alters the affinity of the cardiac sodium channel for lidocaine such that the drug assumes class IC characteristics with potent use-dependent block of the sodium channel.


.0041   ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, ASN1986LYS
SNP: rs199473335, gnomAD: rs199473335, ClinVar: RCV000022947, RCV000148858, RCV000154830, RCV000688881, RCV000756620, RCV001841252

In a father and son with atrial fibrillation (ATFB10; 614022), Ellinor et al. (2008) identified heterozygosity for a 5958C-A transversion in the SCN5A gene, resulting in an asn1986-to-lys (N1986K) substitution in the C-terminal region of the protein. The mutation was not found in more than 600 ethnically and racially matched control chromosomes. Expression of the N1986K mutant in Xenopus oocytes revealed a hyperpolarizing shift in channel steady-state inactivation.


.0042   ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, HIS445ASP
SNP: rs199473112, gnomAD: rs199473112, ClinVar: RCV000022948, RCV000058414, RCV000418451, RCV000458775, RCV000991041, RCV001841253, RCV002381260, RCV002504819

In white male proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 39 years of age, Darbar et al. (2008) identified heterozygosity for a G-to-C transversion in the SCN5A gene, resulting in a his445-to-asp (H445D) substitution at a highly conserved residue that was predicted to perturb cardiac sodium channel function. The proband had left atrial enlargement and an ejection fraction of 60% by transthoracic echocardiography. The mutation was also detected in his affected father and brother, but was not found in an unaffected sister or in 720 control alleles.


.0043   ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, ASN470LYS
SNP: rs199473115, gnomAD: rs199473115, ClinVar: RCV000022949, RCV000058421, RCV002482899, RCV003541160

In a black male proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 17 years of age, Darbar et al. (2008) identified heterozygosity for a G-to-C transversion in the SCN5A gene, resulting in an asn470-to-lys (N470K) substitution at a highly conserved residue and predicted to perturb cardiac sodium channel function. The proband had left atrial enlargement with an ejection fraction of 60% by transthoracic echocardiography. The mutation was also detected in his affected mother and maternal grandmother, but was not found in an unaffected maternal aunt or in 720 control alleles.


.0044   ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, GLU428LYS
SNP: rs199473111, gnomAD: rs199473111, ClinVar: RCV000022950, RCV000148855, RCV000182967, RCV000765740, RCV001841512, RCV002371781, RCV003944837

In a white male proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 52 years of age, Darbar et al. (2008) identified heterozygosity for a A-to-G transition in the SCN5A gene, resulting in an glu428-to-lys (E428K) substitution at a highly conserved residue and predicted to perturb cardiac sodium channel function. The proband had left atrial enlargement with an ejection fraction of 58% by transthoracic echocardiography. The mutation was also detected in his affected daughter and granddaughter, but was not found in an unaffected daughter and granddaughter or in 720 control alleles.


.0045   ATRIAL FIBRILLATION, FAMILIAL, 10

SCN5A, GLU655LYS
SNP: rs199473579, gnomAD: rs199473579, ClinVar: RCV000022951, RCV000058468, RCV000485732, RCV002482900

In a white female proband who was diagnosed with paroxysmal lone atrial fibrillation (ATFB10; 614022) at 37 years of age, Darbar et al. (2008) identified heterozygosity for an A-to-G transition in the SCN5A gene, resulting in a glu655-to-lys (E655K) substitution at a highly conserved residue that was predicted to perturb cardiac sodium channel function. The proband had a normal-sized left atrium and ventricle with an ejection fraction of 55% by transthoracic echocardiography. The mutation was also detected in her affected daughter and maternal grandmother, but was not found in 720 control alleles.


.0046   CARDIOMYOPATHY, DILATED, 1E

SCN5A, ARG222GLN ({dbSNP rs45546039})
SNP: rs45546039, gnomAD: rs45546039, ClinVar: RCV000032639, RCV000058833, RCV000182941, RCV000211852, RCV000678965, RCV000763109

In 6 affected members over 3 generations of a non-Hispanic white family with cardiomyopathy and conduction system disease (CMD1E; 601154), Hershberger et al. (2008) identified heterozygosity for a 36683G-A (numbering per SeattleSNP) transition in exon 6 of the SCN5A gene, resulting in an arg222-to-gln (R222Q) substitution at a conserved residue. The mutation was not found in unaffected family members or in 253 controls.

In 19 affected individuals from 3 unrelated 3-generation families with multifocal ectopic Purkinje-related premature contractions and dilated cardiomyopathy, Laurent et al. (2012) identified heterozygosity for the 665G-A (R222Q) mutation in the SCN5A gene, located in the voltage-sensing S4 segment of domain I. The mutation, which was fully penetrant and strictly segregated with the cardiac phenotype in each family, was not found in 600 control chromosomes; haplotype analysis showed that a founder effect for these 3 families was very unlikely. In vitro studies recapitulated the normalization of the ventricular action potentials in the presence of quinidine. Because only 6 of the 19 patients carrying the R222Q mutation had CMD, and the cardiomyopathy recovered at least partially with antiarrhythmia treatment and a reduction in the number of premature ventricular contractions, Laurent et al. (2012) suggested that CMD might be a consequence of the arrhythmia and not directly linked to the mutation.

In affected members of a 3-generation Canadian family with CMD and junctional escape ventricular capture bigeminy, Nair et al. (2012) identified the R222Q mutation in the SCN5A gene. Heterologous expression studies in Chinese hamster ovary K1 cells revealed a unique biophysical phenotype of R222Q channels in which an approximately 10-mV leftward shift in the sodium current steady-state activation curve occurs without corresponding shifts in steady-state inactivation at cardiomyocyte resting membrane-potential voltages. The activation and inactivation of cells expressing equimolar combinations of wildtype and R222Q channels showed properties intermediate between those seen in cells expressing either wildtype or mutant channels alone. The changes in mutant channel properties were predicted to produce hyperexcitability of R222Q sodium channels.

In 16 affected members over 3 generations of a large kindred with CMD and multiple arrhythmias, including premature ventricular complexes (PVCs) of variable morphology, Mann et al. (2012) identified heterozygosity for the R222Q mutation in the SCN5A gene. The mutation was also identified in 1 clinically unaffected family member, a 56-year-old man with a normal EKG and echocardiogram, but was not found in 200 control chromosomes. Patch-clamp studies showed that the R222Q mutation did not alter sodium channel current density, but did shift steady-state parameters of activation and inactivation to the left. Using a voltage ramp protocol, normalized current responses of mutant channels were of earlier onset and greater magnitude than wildtype. Action potential modeling using Purkinje fiber and ventricular cell models suggested that rate-dependent ectopy of Purkinje fiber origin is the predominant ventricular effect of the R222Q variant; this was supported by the clinical observation that PVC frequency increased during periods of low heart rate at rest and at night, and was reduced by high heart rates during exercise. Patients responded to sodium-channel blocking drugs with early and substantial reductions in PVCs followed by normalization of CMD over time.


.0047   CARDIOMYOPATHY, DILATED, 1E

SCN5A, ILE1835THR
SNP: rs45563942, gnomAD: rs45563942, ClinVar: RCV000032640, RCV000058788, RCV000148847, RCV000212993, RCV000621032, RCV001841554, RCV003964828

In 3 affected members over 2 generations of an African American family with cardiomyopathy and conduction system disease (CMD1E; 601154), Hershberger et al. (2008) identified heterozygosity for a 99599T-C transition (numbering per SeattleSNP) in exon 28 of the SCN5A gene, resulting in an ile1835-to-thr (I1835T) substitution at a conserved residue. The mutation was not found in unaffected family members or in 253 controls.


.0048   ATRIAL STANDSTILL 1, DIGENIC

SCN5A, LEU212PRO
SNP: rs199473070, ClinVar: RCV000058830, RCV000114993, RCV003539788

In a Japanese boy with atrial standstill (ATRST1; 108770), Makita et al. (2005) identified coinheritance of a heterozygous c.635C-T transition in exon 6 of the SCN5A gene, resulting in a leu212-to-pro (L212P) substitution in the extracellular loop connecting transmembrane segments 3 and 4 of domain 1 of the Nav1.5 cardiac sodium channel, and heterozygous rare polymorphisms in the GJA5 gene (121013). The L212P mutation, which was also present in the proband's asymptomatic father, was not found in 400 control chromosomes. Functional analysis with the L212P mutant channels demonstrated large hyperpolarizing shifts in both the voltage dependence of activation and inactivation and delayed recovery from inactivation compared to wildtype. The asymptomatic father did not carry the rare polymorphisms in the GJA5 gene; the GJA5 polymorphisms were, however, present in heterozygosity in the proband's unaffected mother and maternal grandmother, who did not carry the L212P mutation.


REFERENCES

  1. Ackerman, M. J., Siu, B. L., Sturner, W. Q., Tester, D. J., Valdivia, C. R., Makielski, J. C., Towbin, J. A. Postmortem molecular analysis of SCN5A defects in sudden infant death syndrome. JAMA 286: 2264-2269, 2001. [PubMed: 11710892] [Full Text: https://doi.org/10.1001/jama.286.18.2264]

  2. Ackerman, M. J., Splawski, I., Makielski, J. C., Tester, D. J., Will, M. L., Timothy, K. W., Keating, M. T., Jones, G., Chadha, M., Burrow, C. R., Stephens, J. C., Xu, C., Judson, R., Curran, M. E. 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 1: 600-607, 2004. [PubMed: 15851227] [Full Text: https://doi.org/10.1016/j.hrthm.2004.07.013]

  3. Akai, J., Makita, N., Sakurada, H., Shirai, N., Ueda, K., Kitabatake, A., Nakazawa, K., Kimura, A., Hiraoka, M. A novel SCN5A mutation associated with idiopathic ventricular fibrillation without typical ECG findings of Brugada syndrome. FEBS Lett. 479: 29-34, 2000. [PubMed: 10940383] [Full Text: https://doi.org/10.1016/s0014-5793(00)01875-5]

  4. Albert, C. M., Nam, E. G., Rimm, E. B., Jin, H. W., Hajjar, R. J., Hunter, D. J., MacRae, C. A., Ellinor, P. T. Cardiac sodium channel gene variants and sudden cardiac death in women. Circulation 117: 16-23, 2008. [PubMed: 18071069] [Full Text: https://doi.org/10.1161/CIRCULATIONAHA.107.736330]

  5. Barajas-Martinez, H. M., Hu, D., Cordeiro, J. M., Wu, Y., Kovacs, R. J., Meltser, H., Kui, H., Elena, B., Brugada, R., Antzelevitch, C., Dumaine, R. Lidocaine-induced Brugada syndrome phenotype linked to a novel double mutation in the cardiac sodium channel. Circ. Res. 103: 396-404, 2008. [PubMed: 18599870] [Full Text: https://doi.org/10.1161/CIRCRESAHA.108.172619]

  6. Bennett, P. B., Yazawa, K., Makita, N., George, A. L., Jr. Molecular mechanism for an inherited cardiac arrhythmia. Nature 376: 683-685, 1995. [PubMed: 7651517] [Full Text: https://doi.org/10.1038/376683a0]

  7. Benson, D. W., Wang, D. W., Dyment, M., Knilans, T. K., Fish, F. A., Strieper, M. J., Rhodes, T. H., George, A. L., Jr. Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). J. Clin. Invest. 112: 1019-1028, 2003. [PubMed: 14523039] [Full Text: https://doi.org/10.1172/JCI18062]

  8. Bezzina, C., Veldkamp, M. W., van den Berg, M. P., Postma, A. V., Rook, M. B., Viersma, J.-W., van Langen, I. M., Tan-Sindhunata, G., Bink-Boelkens, M. T. E., van der Hout, A. H., Mannens, M. M. A. M., Wilde, A. A. M. A single Na+ channel mutation causing both long-QT and Brugada syndromes. Circ. Res. 85: 1206-1213, 1999. [PubMed: 10590249] [Full Text: https://doi.org/10.1161/01.res.85.12.1206]

  9. Chen, Q., Kirsch, G. E., Zhang, D., Brugada, R., Brugada, J., Brugada, P., Potenza, D., Moya, A., Borggrefe, M., Breithardt, G., Ortiz-Lopez, R., Wang, Z., Antzelevitch, C., O'Brien, R. E., Schulze-Bahr, E., Keating, M. T., Towbin, J. A., Wang, Q. Genetic basis and molecular mechanism for idiopathic ventricular fibrillation. Nature 392: 293-295, 1998. [PubMed: 9521325] [Full Text: https://doi.org/10.1038/32675]

  10. Chen, S., Chung, M. K., Martin, D., Rozich, R., Tchou, P. J., Wang, Q. SNP S1103Y in the cardiac sodium channel gene SCN5A is associated with cardiac arrhythmias and sudden death in a white family. J. Med. Genet. 39: 913-915, 2002. [PubMed: 12471205] [Full Text: https://doi.org/10.1136/jmg.39.12.913]

  11. Cheng, J., Morales, A., Siegfried, J. D., Li, D., Norton, N., Song, J., Gonzalez-Quintana, J., Makielski, J. C., Hershberger, R. E. SCN5A rare variants in familial dilated cardiomyopathy decrease peak sodium current depending on the common polymorphism H558R and common splice variant Q1077del. Clin. Transl. Sci. 3: 287-294, 2010. [PubMed: 21167004] [Full Text: https://doi.org/10.1111/j.1752-8062.2010.00249.x]

  12. Clancy, C. E., Rudy, Y. Linking a genetic defect to its cellular phenotype in a cardiac arrhythmia. Nature 400: 566-569, 1999. [PubMed: 10448858] [Full Text: https://doi.org/10.1038/23034]

  13. Clancy, C. E., Rudy, Y. Na+ channel mutation that causes both Brugada and long-QT syndrome phenotypes: a simulation study of mechanism. Circulation 105: 1208-1213, 2002. [PubMed: 11889015] [Full Text: https://doi.org/10.1161/hc1002.105183]

  14. Clancy, C. E., Tateyama, M., Kass, R. S. Insights into the molecular mechanisms of bradycardia-triggered arrhythmias in long QT-3 syndrome. J. Clin. Invest. 110: 1251-1262, 2002. [PubMed: 12417563] [Full Text: https://doi.org/10.1172/JCI15928]

  15. Darbar, D., Kannankeril, P. J., Donahue, B. S., Kucera, G., Stubblefield, T., Haines, J. L., George, A. L., Jr., Roden, D. M. Cardiac sodium channel (SCN5A) variants associated with atrial fibrillation. Circulation 117: 1927-1935, 2008. [PubMed: 18378609] [Full Text: https://doi.org/10.1161/CIRCULATIONAHA.107.757955]

  16. Dumaine, R., Towbin, J. A., Brugada, P., Vatta, M., Nesterenko, D. V., Nesterenko, V. V., Brugada, J., Brugada, R., Antzelevitch, C. Ionic mechanisms responsible for the electrocardiographic phenotype of the Brugada syndrome are temperature dependent. Circ. Res. 85: 803-809, 1999. [PubMed: 10532948] [Full Text: https://doi.org/10.1161/01.res.85.9.803]

  17. Dumaine, R. Personal Communication. Quebec, Canada 6/2009.

  18. Ellinor, P. T., Nam, E. G., Shea, M. A., Milan, D. J., Ruskin, J. N., MacRae, C. A. Cardiac sodium channel mutation in atrial fibrillation. Heart Rhythm 5: 99-105, 2008. [PubMed: 18088563] [Full Text: https://doi.org/10.1016/j.hrthm.2007.09.015]

  19. Freyermuth, F., Rau, F., Kokunai, Y., Linke, T., Sellier, C., Nakamori, M., Kino, Y., Arandel, L., Jollet, A., Thibault, C., Philipps, M., Vicaire, S., and 31 others. Splicing misregulation of SCN5A contributes to cardiac-conduction delay and heart arrhythmia in myotonic dystrophy. Nature Commun. 7: 11067, 2016. Note: Electronic Article. [PubMed: 27063795] [Full Text: https://doi.org/10.1038/ncomms11067]

  20. Gellens, M. E., George, A. L., Jr., Chen, L., Chahine, M., Horn, R., Barchi, R. L., Kallen, R. G. Primary structure and functional expression of the human cardiac tetrodotoxin-insensitive voltage-dependent sodium channel. Proc. Nat. Acad. Sci. 89: 554-558, 1992. [PubMed: 1309946] [Full Text: https://doi.org/10.1073/pnas.89.2.554]

  21. George, A. L., Jr., Varkony, T. A., Drabkin, H. A., Han, J., Knops, J. F., Finley, W. H., Brown, G. B., Ward, D. C., Haas, M. Assignment of the human heart tetrodotoxin-resistant voltage-gated Na(+) channel alpha-subunit gene (SCN5A) to band 3p21. Cytogenet. Cell Genet. 68: 67-70, 1995. [PubMed: 7956363] [Full Text: https://doi.org/10.1159/000133892]

  22. Greenlee, P. R., Anderson, J. L., Lutz, J. R., Lindsay, A. E., Hagan, A. D. Familial automaticity-conduction disorder with associated cardiomyopathy. West. J. Med. 144: 33-41, 1986. [PubMed: 3953067]

  23. Groenewegen, W. A., Firouzi, M., Bezzina, C. R., Vliex, S., van Langen, I. M., Sandkuijl, L., Smits, J. P. P., Hulsbeek, M., Rook, M. B., Jongsma, H. J., Wilde, A. A. M. A cardiac sodium channel mutation cosegregates with a rare connexin40 genotype in familial atrial standstill. Circ. Res. 92: 14-22, 2003. [PubMed: 12522116] [Full Text: https://doi.org/10.1161/01.res.0000050585.07097.d7]

  24. Groenewegen, W. A., Wilde, A. A. M. Letter regarding article by McNair et al, 'SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia'. (Letter) Circulation 112: e9, 2005. Note: Electronic Article. [PubMed: 15998690] [Full Text: https://doi.org/10.1161/CIRCULATIONAHA.104.531475]

  25. Gross, M. B. Personal Communication. Baltimore, Md. 4/23/2019.

  26. Hershberger, R. E., Parks, S. B., Kushner, J. D., Li, D., Ludwigsen, S., Jakobs, P., Nauman, D., Burgess, D., Partain, J., Litt, M. Coding sequence mutations identified in MYH7, TNNT2, SCN5A, CSRP3, LBD3 (sic), and TCAP from 313 patients with familial or idiopathic dilated cardiomyopathy. Clin. Transl. Sci. 1: 21-26, 2008. [PubMed: 19412328] [Full Text: https://doi.org/10.1111/j.1752-8062.2008.00017.x]

  27. Hwang, H. W., Chen, J. J., Lin, Y. J., Shieh, R. C., Lee, M. T., Hung, S. I., Wu, J. Y., Chen, Y. T., Niu, D. M., Hwang, B. T., Chen, Y. T. R1193Q of SCN5A, a Brugada and long QT mutation, is a common polymorphism in Han Chinese. (Letter) J. Med. Genet. 42: e7, 2005. Note: Electronic Article. [PubMed: 15689442] [Full Text: https://doi.org/10.1136/jmg.2004.027995]

  28. Jones, A., Kainz, D., Khan, F., Lee, C., Carrithers, M. D. Human macrophage SCN5A activates an innate immune signaling pathway for antiviral host defense. J. Biol. Chem. 289: 35326-35340, 2014. [PubMed: 25368329] [Full Text: https://doi.org/10.1074/jbc.M114.611962]

  29. Kambouris, N. G., Nuss, H. B., Johns, D. C., Marban, E., Tomaselli, G. F., Balser, J. R. A revised view of cardiac sodium channel 'blockade' in the long-QT syndrome. J. Clin. Invest. 105: 1133-1140, 2000. [PubMed: 10772658] [Full Text: https://doi.org/10.1172/JCI9212]

  30. Kyndt, F., Probst, V., Potet, F., Demolombe, S., Chevallier, J.-C., Baro, I., Moisan, J.-P., Boisseau, P., Schott, J.-J., Escande, D., Le Marec, H. Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family. Circulation 104: 3081-3086, 2001. [PubMed: 11748104] [Full Text: https://doi.org/10.1161/hc5001.100834]

  31. Laitinen-Forsblom, P. J., Makynen, P., Makynen, H., Yli-Mayry, S., Virtanen, V., Kontula, K., Aalto-Setala, K. SCN5A mutation associated with cardiac conduction defect and atrial arrhythmias. J. Cardiovasc. Electrophysiol. 17: 480-485, 2006. [PubMed: 16684018] [Full Text: https://doi.org/10.1111/j.1540-8167.2006.00411.x]

  32. Laurent, G., Saal, S., Amarouch, M. Y., Beziau, D. M., Marsman, R. F. J., Faivre, L., Barc, J., Dina, C., Bertaux, G., Barthez, O., Thauvin-Robinet, C., Charron, P., and 15 others. Multifocal ectopic Purkinje-related premature contractions. J. Am. Coll. Cardiol. 60: 144-156, 2012. [PubMed: 22766342] [Full Text: https://doi.org/10.1016/j.jacc.2012.02.052]

  33. Maekawa, K., Saito, Y., Ozawa, S., Adachi-Akahane, S., Kawamoto, M., Komamura, K., Shimizu, W., Ueno, K., Kamakura, S., Kamatani, N., Kitakaze, M., Sawada, J. Genetic polymorphisms and haplotypes of the human cardiac sodium channel alpha subunit gene (SCN5A) in Japanese and their association with arrhythmia. Ann. Hum. Genet. 69: 413-428, 2005. [PubMed: 15996170] [Full Text: https://doi.org/10.1046/j.1529-8817.2005.00167.x]

  34. Makita, N., Behr, E., Shimizu, W., Horie, M., Sunami, A., Crotti, L., Schulze-Bahr, E., Fukuhara, S., Mochizuki, N., Makiyama, T., Itoh, H., Christiansen, M., McKeown, P., Miyamoto, K., Kamakura, S., Tsutsui, H., Schwartz, P. J., George, A. L., Jr., Roden, D. M. The E1784K mutation in SCN5A is associated with mixed clinical phenotype of type 3 long QT syndrome. J. Clin. Invest. 118: 2219-2229, 2008. [PubMed: 18451998] [Full Text: https://doi.org/10.1172/JCI34057]

  35. Makita, N., Sasaki, K., Groenewegen, W. A., Yokota, T., Yokoshiki, H., Murakami, T., Tsutsui, H. Congenital atrial standstill associated with coinheritance of a novel SCN5A mutation and connexin 40 polymorphisms. Heart Rhythm 2: 1128-1134, 2005. [PubMed: 16188595] [Full Text: https://doi.org/10.1016/j.hrthm.2005.06.032]

  36. Makita, N., Shirai, N., Nagashima, M., Matsuoka, R., Yamada, Y., Tohse, N., Kitabatake, A. A de novo missense mutation of human cardiac Na(+) channel exhibiting novel molecular mechanisms of long QT syndrome. FEBS Lett. 423: 5-9, 1998. [PubMed: 9506831] [Full Text: https://doi.org/10.1016/s0014-5793(98)00033-7]

  37. Makita, N., Shirai, N., Wang, D. W., Sasaki, K., George, A. L., Kanno, M., Kitabatake, A. Cardiac Na+ channel dysfunction in Brugada syndrome is aggravated by beta(1)-subunit. Circulation 101: 54-60, 2000. [PubMed: 10618304] [Full Text: https://doi.org/10.1161/01.cir.101.1.54]

  38. Mann, S. A., Castro, M. L., Ohanian, M., Guo, G., Zodgekar, P., Sheu, A., Stockhammer, K., Thompson, T., Playford, D., Subbiah, R., Kuchar, D., Aggarwal, A., Vandenberg, J. I., Fatkin, D. R222Q SCN5A mutation is associated with reversible ventricular ectopy and dilated cardiomyopathy. J. Am. Coll. Cardiol. 60: 1566-1573, 2012. [PubMed: 22999724] [Full Text: https://doi.org/10.1016/j.jacc.2012.05.050]

  39. McNair, W. P., Ku, L., Taylor, M. R. G., Fain, P. R., Dao, D., Wolfel, E., Mestroni, L., Familial Cardiomyopathy Registry Research Group. SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia. Circulation 110: 2163-2167, 2004. [PubMed: 15466643] [Full Text: https://doi.org/10.1161/01.CIR.0000144458.58660.BB]

  40. McNair, W. P., Ku, L., Taylor, M. R. G., Fain, P. R., Wolfel, E., Mestroni, L. Response to letter regarding article by McNair et al., 'SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia'. (Letter) Circulation 112: e9, 2005. Note: Electronic Article.

  41. Millat, G., Chevalier, P., Restier-Miron, L., Da Costa, A., Bouvagnet, P., Kugener, B., Fayol, L., Gonzalez Armengod, C., Oddou, B., Chanavat, V., Froidefond, E., Perraudin, R., Rousson, R., Rodriguez-Lafrasse, C. Spectrum of pathogenic mutations and associated polymorphisms in a cohort of 44 unrelated patients with long QT syndrome. Clin. Genet. 70: 214-227, 2006. [PubMed: 16922724] [Full Text: https://doi.org/10.1111/j.1399-0004.2006.00671.x]

  42. Miller, T. E., Estrella, E., Myerburg, R. J., Garcia de Viera, J., Moreno, N., Rusconi, P., Ahearn, M. E., Baumbach, L., Kurlansky, P., Wolff, G., Bishopric, N. H. Recurrent third-trimester fetal loss and maternal mosaicism for long-QT syndrome. Circulation 109: 3029-3034, 2004. [PubMed: 15184283] [Full Text: https://doi.org/10.1161/01.CIR.0000130666.81539.9E]

  43. Mohler, P. J., Rivolta, I., Napolitano, C., LeMaillet, G., Lambert, S., Priori, S. G., Bennett, V. Na(v)1.5 E1053K mutation causing Brugada syndrome blocks binding to ankyrin-G and expression of Na(v)1.5 on the surface of cardiomyocytes. Proc. Nat. Acad. Sci. 101: 17533-17538, 2004. [PubMed: 15579534] [Full Text: https://doi.org/10.1073/pnas.0403711101]

  44. Nair, K., Pekhletski, R., Harris, L., Care, M., Morel, C., Farid, T., Backx, P. H., Szabo, E., Nanthakumar, K. Escape capture bigeminy: phenotypic marker of cardiac sodium channel voltage sensor mutation R222Q. Heart Rhythm 9: 1681-1688, 2012. [PubMed: 22710484] [Full Text: https://doi.org/10.1016/j.hrthm.2012.06.029]

  45. Niu, D.-M., Hwang, B., Hwang, H.-W., Wang, N. H., Wu, J.-Y., Lee, P.-C., Chien, J.-C., Shieh, R.-C., Chen, Y.-T. A common SCN5A polymorphism attenuates a severe cardiac phenotype caused by a nonsense SCN5A mutation in a Chinese family with an inherited cardiac conduction defect. J. Med. Genet. 43: 817-821, 2006. [PubMed: 16707561] [Full Text: https://doi.org/10.1136/jmg.2006.042192]

  46. Noble, D. Unraveling the genetics and mechanisms of cardiac arrhythmia. (Commentary) Proc. Nat. Acad. Sci. 99: 5755-5756, 2002. [PubMed: 11983875] [Full Text: https://doi.org/10.1073/pnas.102171699]

  47. Nuyens, D., Stengl, M., Dugarmaa, S., Rossenbacker, T., Compernolle, V., Rudy, Y., Smits, J. F., Flameng, W., Clancy, C. E., Moons, L., Vos, M. A., Dewerchin, M., Benndorf, K., Collen, D., Carmeliet, E., Carmeliet, P. Abrupt rate accelerations or premature beats cause life-threatening arrhythmias in mice with long-QT3 syndrome. Nature Med. 7: 1021-1027, 2001. [PubMed: 11533705] [Full Text: https://doi.org/10.1038/nm0901-1021]

  48. O'Neill, M. J., Muhammad, A., Li, B., Wada, Y., Hall, L., Solus, J. F., Short, L., Roden, D. M., Glazer, A. M. Dominant negative effects of SCN5A missense variants. Genet. Med. 24: 1238-1248, 2022. [PubMed: 35305865] [Full Text: https://doi.org/10.1016/j.gim.2022.02.010]

  49. Olson, T. M., Michels, V. V., Ballew, J. D., Reyna, S. P., Karst, M. L., Herron, K. I., Horton, S. C., Rodeheffer, R. J., Anderson, J. L. Sodium channel mutations and susceptibility of heart failure and atrial fibrillation. JAMA 293: 447-454, 2005. [PubMed: 15671429] [Full Text: https://doi.org/10.1001/jama.293.4.447]

  50. Papadatos, G. A., Wallerstein, P. M. R., Head, C. E. G., Ratcliff, R., Brady, P. A., Benndorf, K., Saumarez, R. C., Trezise, A. E. O., Huang, C. L.-H., Vandenberg, J. I., Colledge, W. H., Grace, A. A. Slowed conduction and ventricular tachycardia after targeted disruption of the cardiac sodium channel gene Scn5a. Proc. Nat. Acad. Sci. 99: 6210-6215, 2002. [PubMed: 11972032] [Full Text: https://doi.org/10.1073/pnas.082121299]

  51. Plant, L. D., Bowers, P. N., Liu, Q., Morgan, T., Zhang, T., State, M. W., Chen, W., Kittles, R. A., Goldstein, S. A. N. A common cardiac sodium channel variant associated with sudden infant death in African Americans, SCN5A S1103Y. J. Clin. Invest. 116: 430-435, 2006. [PubMed: 16453024] [Full Text: https://doi.org/10.1172/JCI25618]

  52. Rivolta, I., Abriel, H., Tateyama, M., Liu, H., Memmi, M., Vardas, P., Napolitano, C., Priori, S. G., Kass, R. S. Inherited Brugada and long QT-3 syndrome mutations of a single channel residue of the cardiac sodium channel confer distinct channel and clinical phenotypes. J. Biol. Chem. 276: 30623-30630, 2001. [PubMed: 11410597] [Full Text: https://doi.org/10.1074/jbc.M104471200]

  53. Rook, M. B., Alshinawi, C. B., Groenewegen, W. A., van Gelder, I. C., van Ginneken, A. C. G., Jongsma, H. J., Mannens, M. M. A. M., Wilde, A. A. M. Human SCN5A gene mutations alter cardiac sodium channel kinetics and are associated with the Brugada syndrome. Cardiovasc. Res. 44: 507-517, 1999. [PubMed: 10690282] [Full Text: https://doi.org/10.1016/s0008-6363(99)00350-8]

  54. Schott, J.-J., Alshinawi, C., Kyndt, F., Probst, V., Hoorntje, T. M., Hulsbeek, M., Wilde, A. A. M., Escande, D., Mannens, M. M. A. M., Le Marec, H. Cardiac conduction defects associate with mutations in SCN5A. (Letter) Nature Genet. 23: 20-21, 1999. [PubMed: 10471492] [Full Text: https://doi.org/10.1038/12618]

  55. Schwartz, P. J., Priori, S. G., Dumaine, R., Napolitano, C., Antzelevitch, C., Stramba-Badiale, M., Richard, T. A., Berti, M. R., Bloise, R. A molecular link between the sudden infant death syndrome and the long-QT syndrome. New Eng. J. Med. 343: 262-267, 2000. [PubMed: 10911008] [Full Text: https://doi.org/10.1056/NEJM200007273430405]

  56. Shin, D.-J., Jang, Y., Park, H.-Y., Lee, J. E., Yang, K., Kim, E., Bae, Y., Kim, J., Kim, J., Kim, S. S., Lee, M. H., Chahine, M., Yoon, S. K. Genetic analysis of the cardiac sodium channel gene SCN5A in Koreans with Brugada syndrome. J. Hum. Genet. 49: 573-578, 2004. [PubMed: 15338453] [Full Text: https://doi.org/10.1007/s10038-004-0182-z]

  57. Splawski, I., Shen, J., Timothy, K. W., Lehmann, M. H., Priori, S., Robinson, J. L., Moss, A. J., Schwartz, P. J., Towbin, J. A., Vincent, G. M., Keating, M. T. Spectrum of mutations in long-QT syndrome genes: KVLQT1, HERG, SCN5A, KCNE1, and KCNE2. Circulation 102: 1178-1185, 2000. [PubMed: 10973849] [Full Text: https://doi.org/10.1161/01.cir.102.10.1178]

  58. Splawski, I., Timothy, K. W., Tateyama, M., Clancy, C. E., Malhotra, A., Beggs, A. H., Cappuccio, F. P., Sagnella, G. A., Kass, R. S., Keating, M. T. Variant of SCN5A sodium channel implicated in risk of cardiac arrhythmia. Science 297: 1333-1336, 2002. [PubMed: 12193783] [Full Text: https://doi.org/10.1126/science.1073569]

  59. Tan, H. L., Bink-Boelkens, M. T. E., Bezzina, C. R., Viswanathan, P. C., Beaufort-Krol, G. C. M., van Tintelen, P. J., van den Berg, M. P., Wilde, A. A. M., Balser, J. R. A sodium-channel mutation causes isolated cardiac conduction disease. Nature 409: 1043-1047, 2001. [PubMed: 11234013] [Full Text: https://doi.org/10.1038/35059090]

  60. Tan, H. L., Kupershmidt, S., Zhang, R., Stepanovic, S., Roden, D. M., Wilde, A. A. M., Anderson, M. E., Balser, J. R. A calcium sensor in the sodium channel modulates cardiac excitability. Nature 415: 442-447, 2002. [PubMed: 11807557] [Full Text: https://doi.org/10.1038/415442a]

  61. Tester, D. J., Will, M. L., Haglund, C. M., Ackerman, M. J. Compendium of cardiac channel mutations in 541 consecutive unrelated patients referred for long QT syndrome genetic testing. Heart Rhythm 2: 507-517, 2005. [PubMed: 15840476] [Full Text: https://doi.org/10.1016/j.hrthm.2005.01.020]

  62. Vatta, M., Dumaine, R., Varghese, G., Richard, T. A., Shimizu, W., Aihara, N., Nademanee, K., Brugada, R., Brugada, J., Veerakul, G., Li, H., Bowles, N. E., Brugada, P., Antzelevitch, C., Towbin, J. A. Genetic and biophysical basis of sudden unexplained nocturnal death syndrome (SUNDS), a disease allelic to Brugada syndrome. Hum. Molec. Genet. 11: 337-345, 2002. [PubMed: 11823453] [Full Text: https://doi.org/10.1093/hmg/11.3.337]

  63. Veldkamp, M. W., Wilders, R., Baartscheer, A., Zegers, J. G., Bezzina, C. R., Wilde, A. A. M. Contribution of sodium channel mutations to bradycardia and sinus node dysfunction in LQT3 families. Circ. Res. 92: 976-983, 2003. [PubMed: 12676817] [Full Text: https://doi.org/10.1161/01.RES.0000069689.09869.A8]

  64. Viswanathan, P. C., Benson, D. W., Balser, J. R. A common SCN5A polymorphism modulates the biophysical effects of an SCN5A mutation. J. Clin. Invest. 111: 341-346, 2003. [PubMed: 12569159] [Full Text: https://doi.org/10.1172/JCI16879]

  65. Wang, D. W., Viswanathan, P. C., Balser, J. R., George, A. L., Jr., Benson, W. Clinical, genetic and biophysical characterisation of SCN5A mutations associated with atrioventricular block. Circulation 105: 341-346, 2002. [PubMed: 11804990] [Full Text: https://doi.org/10.1161/hc0302.102592]

  66. Wang, D. W., Yazawa, K., George, A. L., Jr., Bennett, P. B. Characterization of human cardiac Na(+) channel mutations in the congenital long QT syndrome. Proc. Nat. Acad. Sci. 93: 13200-13205, 1996. [PubMed: 8917568] [Full Text: https://doi.org/10.1073/pnas.93.23.13200]

  67. Wang, D. W., Yazawa, K., Makita, N., George, A. L., Jr., Bennett, P. B. Pharmacological targeting of long QT mutant sodium channels. J. Clin. Invest. 99: 1714-1720, 1997. [PubMed: 9120016] [Full Text: https://doi.org/10.1172/JCI119335]

  68. Wang, Q., Chen, S., Chen, Q., Wan, X., Shen, J., Hoeltge, G. A., Timur, A. A., Keating, M. T., Kirsch, G. E. The common SCN5A mutation R1193Q causes LQTS-type electrophysiological alterations of the cardiac sodium channel. J. Med. Genet. 41: e66, 2004. Note: Electronic Article. [PubMed: 15121794] [Full Text: https://doi.org/10.1136/jmg.2003.013300]

  69. Wang, Q., Li, Z., Shen, J., Keating, M. T. Genomic organization of the human SCN5A gene encoding the cardiac sodium channel. Genomics 34: 9-16, 1996. [PubMed: 8661019] [Full Text: https://doi.org/10.1006/geno.1996.0236]

  70. Wang, Q., Shen, J., Li, Z., Timothy, K., Vincent, G. M., Priori, S. G., Schwartz, P. J., Keating, M. T. Cardiac sodium channel mutations in patients with long QT syndrome, an inherited cardiac arrhythmia. Hum. Molec. Genet. 4: 1603-1607, 1995. [PubMed: 8541846] [Full Text: https://doi.org/10.1093/hmg/4.9.1603]

  71. Wang, Q., Shen, J., Splawski, I., Atkinson, D., Li, Z., Robinson, J. L., Moss, A. J., Towbin, J. A., Keating, M. T. SCN5A mutations associated with an inherited cardiac arrhythmia, long QT syndrome. Cell 80: 805-811, 1995. [PubMed: 7889574] [Full Text: https://doi.org/10.1016/0092-8674(95)90359-3]

  72. Wang, Q. Author's reply: link of SCN5A SNP R1193Q to long QT syndrome. (Letter) J. Med. Genet. 42: e8, 2005. Note: Electronic Article.

  73. Wei, J., Wang, D. W., Alings, M., Fish, F., Wathen, M., Roden, D. M., George, A. L., Jr. Congenital long-QT syndrome caused by a novel mutation in a conserved acidic domain of the cardiac Na(+) channel. Circulation 99: 3165-3171, 1999. [PubMed: 10377081] [Full Text: https://doi.org/10.1161/01.cir.99.24.3165]

  74. Westenskow, P., Splawski, I., Timothy, K. W., Keating, M. T., Sanguinetti, M. C. Compound mutations: a common cause of severe long-QT syndrome. Circulation 109: 1834-1841, 2004. [PubMed: 15051636] [Full Text: https://doi.org/10.1161/01.CIR.0000125524.34234.13]

  75. Yang, P., Kanki, H., Drolet, B., Yang, T., Wei, J., Viswanathan, P. C., Hohnloser, S. H., Shimizu, W., Schwartz, P. J., Stanton, M., Murray, K. T., Norris, K., George, A. L., Jr., Roden, D. M. Allelic variants in long-QT disease genes in patients with drug-associated torsades de pointes. Circulation 105: 1943-1948, 2002. [PubMed: 11997281] [Full Text: https://doi.org/10.1161/01.cir.0000014448.19052.4c]


Contributors:
Sonja A. Rasmussen - updated : 09/12/2022
Matthew B. Gross - updated : 04/23/2019
Patricia A. Hartz - updated : 11/16/2016
Paul J. Converse - updated : 1/8/2015
Marla J. F. O'Neill - updated : 4/29/2014
Marla J. F. O'Neill - updated : 1/29/2013
Marla J. F. O'Neill - updated : 6/1/2011
Marla J. F. O'Neill - updated : 6/8/2009
Marla J. F. O'Neill - updated : 12/23/2008
Marla J. F. O'Neill - updated : 5/14/2008
Marla J. F. O'Neill - updated : 3/6/2008
Marla J. F. O'Neill - updated : 2/12/2008
Marla J. F. O'Neill - updated : 1/12/2007
Marla J. F. O'Neill - updated : 11/9/2006
Marla J. F. O'Neill - updated : 7/10/2006
Victor A. McKusick - updated : 2/20/2006
Marla J. F. O'Neill - updated : 1/31/2006
Marla J. F. O'Neill - updated : 10/11/2005
Victor A. McKusick - updated : 1/27/2005
Victor A. McKusick - updated : 1/3/2005
Cassandra L. Kniffin - updated : 10/26/2004
Marla J. F. O'Neill - updated : 2/18/2004
Victor A. McKusick - updated : 11/18/2003
Victor A. McKusick - updated : 6/30/2003
Denise L. M. Goh - updated : 1/6/2003
Ada Hamosh - updated : 10/18/2002
George E. Tiller - updated : 9/23/2002
Deborah L. Stone - updated : 6/26/2002
Victor A. McKusick - updated : 6/6/2002
Paul Brennan - updated : 3/27/2002
Paul Brennan - updated : 3/8/2002
Ada Hamosh - updated : 1/22/2002
Victor A. McKusick - updated : 11/6/2001
Ada Hamosh - updated : 2/27/2001
Victor A. McKusick - updated : 9/27/2000
Victor A. McKusick - updated : 9/15/2000
Victor A. McKusick - updated : 6/1/2000
Paul Brennan - updated : 4/12/2000
Paul Brennan - updated : 4/3/2000
Paul Brennan - updated : 4/3/2000
Victor A. McKusick - updated : 2/24/2000
Victor A. McKusick - updated : 1/12/2000
Paul Brennan - updated : 8/31/1999
Ada Hamosh - updated : 8/4/1999
Ada Hamosh - updated : 5/25/1999
Victor A. McKusick - updated : 10/2/1998
Victor A. McKusick - updated : 5/12/1998
Victor A. McKusick - updated : 3/17/1998
Victor A. McKusick - updated : 5/27/1997

Creation Date:
Victor A. McKusick : 10/26/1994

Edit History:
carol : 01/11/2023
carol : 10/03/2022
alopez : 09/30/2022
carol : 09/25/2022
carol : 09/13/2022
carol : 09/12/2022
carol : 09/01/2020
carol : 05/29/2019
mgross : 04/23/2019
alopez : 11/07/2018
carol : 01/05/2018
carol : 10/05/2017
alopez : 09/14/2017
carol : 01/12/2017
alopez : 11/16/2016
alopez : 11/09/2016
alopez : 11/09/2016
carol : 05/24/2016
carol : 5/23/2016
mgross : 1/30/2015
carol : 1/14/2015
mcolton : 1/8/2015
alopez : 11/12/2014
carol : 4/29/2014
mcolton : 4/28/2014
carol : 4/17/2014
carol : 3/19/2014
carol : 8/27/2013
carol : 3/8/2013
alopez : 1/30/2013
alopez : 1/29/2013
carol : 12/15/2011
carol : 11/23/2011
terry : 11/4/2011
carol : 11/3/2011
carol : 7/15/2011
wwang : 6/3/2011
wwang : 6/3/2011
terry : 6/1/2011
wwang : 6/30/2009
terry : 6/8/2009
carol : 6/2/2009
carol : 12/24/2008
terry : 12/23/2008
carol : 12/22/2008
carol : 5/14/2008
carol : 3/6/2008
wwang : 3/5/2008
wwang : 2/26/2008
terry : 2/12/2008
joanna : 2/7/2008
carol : 11/15/2007
alopez : 10/4/2007
carol : 9/10/2007
carol : 1/19/2007
terry : 1/12/2007
carol : 12/8/2006
carol : 11/16/2006
carol : 11/9/2006
carol : 11/9/2006
carol : 10/4/2006
terry : 8/24/2006
wwang : 7/11/2006
terry : 7/10/2006
joanna : 6/2/2006
carol : 2/22/2006
carol : 2/22/2006
terry : 2/20/2006
wwang : 2/20/2006
wwang : 2/3/2006
terry : 1/31/2006
wwang : 10/14/2005
terry : 10/11/2005
wwang : 2/10/2005
wwang : 2/8/2005
terry : 1/27/2005
wwang : 1/6/2005
wwang : 1/6/2005
terry : 1/3/2005
tkritzer : 10/27/2004
ckniffin : 10/26/2004
carol : 10/26/2004
carol : 10/12/2004
joanna : 9/10/2004
ckniffin : 4/30/2004
carol : 4/30/2004
ckniffin : 4/14/2004
tkritzer : 3/18/2004
tkritzer : 3/16/2004
carol : 2/18/2004
alopez : 11/25/2003
tkritzer : 11/20/2003
terry : 11/18/2003
carol : 7/14/2003
tkritzer : 7/8/2003
terry : 6/30/2003
carol : 2/5/2003
carol : 1/6/2003
carol : 1/6/2003
alopez : 10/23/2002
terry : 10/18/2002
cwells : 9/23/2002
carol : 6/26/2002
mgross : 6/10/2002
terry : 6/6/2002
alopez : 3/27/2002
carol : 3/14/2002
alopez : 3/8/2002
alopez : 1/23/2002
terry : 1/22/2002
carol : 11/8/2001
carol : 11/8/2001
mcapotos : 11/6/2001
alopez : 3/7/2001
alopez : 3/6/2001
terry : 2/27/2001
mcapotos : 10/13/2000
mcapotos : 10/11/2000
terry : 9/27/2000
carol : 9/25/2000
terry : 9/22/2000
terry : 9/15/2000
mcapotos : 6/15/2000
mcapotos : 6/14/2000
terry : 6/1/2000
alopez : 4/12/2000
alopez : 4/3/2000
alopez : 4/3/2000
mcapotos : 3/17/2000
mcapotos : 3/8/2000
terry : 2/24/2000
mgross : 2/17/2000
terry : 1/12/2000
carol : 11/4/1999
mgross : 8/31/1999
alopez : 8/4/1999
terry : 8/4/1999
kayiaros : 7/8/1999
carol : 7/7/1999
carol : 5/25/1999
carol : 5/11/1999
carol : 10/7/1998
terry : 10/2/1998
terry : 6/4/1998
carol : 5/21/1998
terry : 5/12/1998
alopez : 3/18/1998
terry : 3/17/1998
jenny : 5/30/1997
terry : 5/27/1997
terry : 12/10/1996
terry : 12/5/1996
terry : 6/5/1996
terry : 6/3/1996
joanna : 12/29/1995
mimadm : 9/23/1995
mark : 9/22/1995
terry : 4/20/1995
mark : 3/30/1995
terry : 10/26/1994