Next Article in Journal
The Role of Small Extracellular Vesicles in the Progression of Colorectal Cancer and Its Clinical Applications
Next Article in Special Issue
Veratridine Can Bind to a Site at the Mouth of the Channel Pore at Human Cardiac Sodium Channel NaV1.5
Previous Article in Journal
Human Chorionic Gonadotropin and Early Embryogenesis: Review
Previous Article in Special Issue
Zfhx3 Transcription Factor Represses the Expression of SCN5A Gene and Decreases Sodium Current Density (INa)
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Genomic and Non-Genomic Regulatory Mechanisms of the Cardiac Sodium Channel in Cardiac Arrhythmias

by
Houria Daimi
1,*,
Estefanía Lozano-Velasco
2,3,
Amelia Aranega
2,3 and
Diego Franco
2,3
1
Biochemistry and Molecular Biology Laboratory, Faculty of Pharmacy, University of Monastir, Monastir 5000, Tunisia
2
Department of Experimental Biology, University of Jaen, 23071 Jaen, Spain
3
Medina Foundation, Technology Park of Health Sciences, Av. del Conocimiento, 34, 18016 Granada, Spain
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2022, 23(3), 1381; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23031381
Submission received: 2 December 2021 / Revised: 30 December 2021 / Accepted: 6 January 2022 / Published: 26 January 2022
(This article belongs to the Special Issue New Insights into Cardiac Ion Channel Regulation 2.0)

Abstract

:
Nav1.5 is the predominant cardiac sodium channel subtype, encoded by the SCN5A gene, which is involved in the initiation and conduction of action potentials throughout the heart. Along its biosynthesis process, Nav1.5 undergoes strict genomic and non-genomic regulatory and quality control steps that allow only newly synthesized channels to reach their final membrane destination and carry out their electrophysiological role. These regulatory pathways are ensured by distinct interacting proteins that accompany the nascent Nav1.5 protein along with different subcellular organelles. Defects on a large number of these pathways have a tremendous impact on Nav1.5 functionality and are thus intimately linked to cardiac arrhythmias. In the present review, we provide current state-of-the-art information on the molecular events that regulate SCN5A/Nav1.5 and the cardiac channelopathies associated with defects in these pathways.

1. Introduction

The upstroke phase of the cardiac action potential (AP) is mainly coordinated by cardiac sodium channels, which are immediately activated and generate a fast Na+ inward current, through the membrane, after membrane depolarization [1]. In atrial and ventricular myocytes, the sodium current (INa) is principally governed by cardiac voltage-gated sodium channel 1.5 (Nav1.5) with a tinny contribution of NaV1.8 [2]. The human Nav1.5 channel is composed of a pore-forming α-subunit (227-kDa) and one or more auxiliary β-subunit (30-kDa) [3]. SCN5A gene with 80 kb length is located on chromosome 3p21 and consists of 28 exons which encode a protein of 2016 amino acid, the α-subunit of Nav1.5 channel [4]. This protein contains four homologous sites (DI–DIV), each composed of six transmembrane segments organized into two functional modules. Segments from one to four (S1–S4) generate the voltage-sensing module (VS), and segments five and six (S5–S6) jointly with P-loop create the pore module (PM). Finally, we can find an α-helical S4–S5 linker, whose function is to bind these two structures, the voltage-sensing and the pore modules. Moreover, there are intracellular linkers that are in charge of DI–DII, DII–DIII, and DIII–DIV binding, and more concretely, the DIII–DIV linker is the controller of pore closing, acting as a fast inactivation gate [5]. The VS and PM modules of the Nav1.5 constitute preferred therapeutic targets for the treatment of several cardiac sodium channelopathies. Particularly, flecainide, as well as other class IC antiarrhythmic drugs, bind to the central cavity of the pore and block sodium permeation directly [5]. The class IA antiarrhythmic drugs (e.g., procainamide) and the class IB antiarrhythmic drugs (e.g., lidocaine) might act on a smaller surface of the central cavity of the pore as well [5]. However, polyunsaturated fatty acids (PUFAs) and PUFA analogs have been shown to be antiarrhythmic by inhibiting Nav1.5 channel currents, probably through acting on the voltage-sensing S4 segments that control inactivation in these channels [6,7].
Since SCN5A is transcribed to mRNA until Nav1.5 is assembled into the plasma membrane to exert its function, there are several steps with different proteins involved. SCN5A mRNA is translocated from the nucleus to the cytoplasm by the Nucleoporin 107 (Nup107) protein. It has been demonstrated that Nup107 is increased whenever a hypoxic and oxidative situation has a place in the heart tissue [8]. Once in the cytoplasm, SCN5A mRNA is translated to protein and assembled to β1–β4 subunits in the rough endoplasmic reticulum (ER), and subsequently, Nav1.5 protein is exported to the Golgi by the nuclear import protein, RAN guanine nucleotide release factor (MOG1), and Protein Kinase A [9,10,11]. It has been previously reported that the defect of MOG1-Nav1.5 interaction causes Brugada syndrome [9]. After protein glycosylation in Golgi, Nav1.5 is ready for anchoring into the intercalated discs (ID) for the AP transmission between cardiomyocytes, being responsible for the electro-mechanical coupling through gap junctions, adherens junctions, and desmosomes [12,13]. Moreover, AP is conducted through the lateral membrane (LM) due to localized Nav1.5 channels into T-tubules and in the focal adhesion complex or costameres that link adjacent myocytes in the myocardium through extracellular matrix interaction [12,13]. This process is characterized by the presence of specific proteins that control Nav1.5 trafficking and anchoring [13].
At the beginning of cardiac AP, the cell membrane is depolarized, and Nav1.5 channels are activated due to positively charged arginine or lysine residues in S4 segments. All four charged S4 segments change their position in the cell membrane with an outward movement, which leads to the opening of the channel pore, conducting to an inward Na+ current [14]. Few milliseconds later, the depolarization process generates “fast inactivation”, where the Nav1.5 channel is closed and channel opening does not occur, due to an outward movement of S4 segments in domain III and IV, again. This dual role of the S4 segment is tightly integrated [15,16,17,18,19,20]. After a prolonged depolarization, P-loop and S5–S6 linker change their position with reference to the membrane, and Nav1.5 channels enter in a “slow inactivation”, leading to the termination of the Na+ current flow [15,21].
At the end of its life cycle, Nav1.5 is degraded by proteasome and autophagic degradation pathways [11].

2. Genomic Regulation of the Cardiac Sodium Channel

2.1. Genetic Code of SCN5A

Nav1.5 channel expression and function may be impaired due to variations in the genomic sequence of SCN5A, including missense, nonsense, splice-altering, and frame shift truncation [22,23]. These variations cause different cardiac diseases because of a loss- or gain-of-function and occasionally both, generating overlapped phenotypes [24]. For example, Brugada Syndrome (BrS) [24,25,26], progressive cardiac conduction disease (Lev-Lenegre disease) [27,28], and sick sinus syndrome [29,30] are some diseases caused by loss-of-function mutations in SCN5A. However, long QT syndrome type 3 (LQTS3) [24,31] and multifocal ectopic Purkinje-related premature contractions (MEPPC) [32,33,34] are due to gain-of-function mutations in SCN5A. Finally, a combination of gain- and loss-of-function mutations are associated with atrial fibrillation (AF) [35,36] and dilated cardiomyopathy (DCM) [34,36,37,38]. Some of these channelopathies are widely described in Section 4 of this review. Additionally, several SCN5A missenses can generate dominant-negative variants affecting Nav1.5 trafficking or gating at the cell surface [39,40,41]. A recent study demonstrated that most of SCN5A missense that generate loss-of-function variants exert a dominant-negative effect that confers a high burden of BrS [42]. Finally, post-translational modifications affecting Nav1.5 have an impact on the use of antiarrhythmic drugs, making these data quite relevant for future drug design [43].

2.2. Regulation of SCN5A Transcription

2.2.1. Epigenetic Regulation of SCN5A

Regulation of SCN5A by Distinct Regulatory Elements and Histones

Gene transcriptional activation is not only modulated by transcription factors; in this process, the role of distinct regulatory elements (RE) is also important, as well as how these REs interact with chromatin, depending on DNA accessibility. Several authors have identified different roles of an enhancer cluster in the SCN5A-SCN10A locus, which modulate SCN5A gene expression [44,45,46]. RE-1 and RE-5 are located in an SCN10A and SCN5A intron, respectively, and RE-6, located downstream of SCN5A, contains genetic variants associated with PR intervals and QRS duration [44,45,47,48]. Moreover, Christoffels’ lab [49] has recently demonstrated that there are several downstream SCN5A REs acting as cardiac-specific “super enhancers”, concretely the intergenic region composed by RE6-9, which possess an extensive association with Histone H3 lysine (K) 27 acetylation (H3K27ac) [50]. RE6-9 has the ability to fine-tune Scn5a-Scn10a chromatin architecture modulating Scn5a expression. In addition, it has been identified that some single-nucleotide polymorphisms (SNPs) located in an enhancer region are able to regulate transcription factor binding and modulate gene expression. In particular, major alleles of rs6801957 and rs10428132 lead to SCN5A gene expression, while minor alleles cannot due to a loss of a T-box protein binding site [51,52]. Furthermore, an enrichment of H3K27ac and Histone H3 lysine (K) 4 trimethylation (H3K4me3) near of SCN5A promoter region in striated muscles regulates normal expression of Scn5a and improve the re-expression of SCN5A in denervated muscle [49,53,54]. Moreover, Lamin A/C (encoded by LMNA) is a component of the nuclear lamina, and its K219T mutation has been described to trigger a change in the distribution of the histone marks. Concretely, H3K9me and H3K27me, which are transcriptional repressive histone marks, and H3K4me3, which acts as transcriptional active histone mark, generate cardiac conduction defects through SCN5A inhibition and reduced INa density [55].

Regulation of SCN5A by Transcription Factors

During biosynthesis, SCN5A transcription is regulated by several transcription factors. Sometimes this transcription step can be enhanced or decreased, i.e., TBX5 has a binding site downstream of the SCN5A gene, and several authors have demonstrated that TBX5 knockout presents a decreased density of Nav1.5 that leads to arrhythmias and eventually sudden cardiac death [56,57,58]. Additionally, GATA4 and GATA5 have their binding site in the SCN5A promoter and intron 1 region. These transcription factors activate the SCN5A gene in human left ventricles, whereas heterozygous mutants for GATA4+/ show short PR intervals [59,60,61]. Moreover, MEF2C has its binding site in the SCN5A promoter region and enhances SCN5A transcription [62,63]. Finally, IRX3 gain-of-function upregulates SCN5A mRNA levels [64,65], whereas, on the contrary, FOXO1 and Snail negatively regulate SCN5A mRNA levels [66,67,68,69,70] (Figure 1).

2.2.2. Post-Transcriptional Regulation of SCN5A

Regulation of SCN5A by Alternative Splicing

After transcription, precursor mRNA copes with splicing and post-transcriptional modification to generate mature mRNA and finally translation into protein. Alternative splicing generate multiple functional (Nav1.5a, Nav1.5d, Nav1.5e, and Nav1.5c) and non-functional (Nav1.5b, Nav1.5f, and truncated) Nav1.5 variants [11]. Nav1.5a isoform is characterized by the deletion of exon 18. This isoform is only present in small rodents and, compared with full-length Nav1.5, leads to altered electrophysiological kinetics properties. There is no evidence of Nav1.5a expression in human cardiac cells [11,71]. Another alternative spliced variant of SCN5A generates Nav1.5c isoform, which has been identified as the most abundant isoform in humans. Nav1.5c is characterized by a 5′-trinucleotide deletion in exon 18, concretely a CAG—Glu (Q) in 1077 position, affected by the splicing machinery and generating a Nav1.5 variant that contains 2015 polypeptides instead of 2016. It has been identified that the electrophysiological properties of Nav1.5 and Nav1.5c are indistinguishable [71,72]. Nav1.5d is another Nav1.5 variant, where 120 bp fragment is deleted from exon 17. This Nav1.5d isoform is present in the fetal and adult human heart and has altered channel kinetics due to a reduction of open channel probability [71,73,74]. Finally, the last functional Nav1.5 variant is Nav1.5e and is generated by alternative splicing on exon 6. It can be found Nav1.5e with 5′-exon 6 in neonatal (exon 6a) or 3′-exon 6 (exon 6b) in any adult mammalian heart [4,75]. Nav1.5e contains a K211 residue, instead of D211 residue in Nav1.5, being responsible for slower kinetics of the channel [76]. Navd1.5b is a non-functional Nav1.5 variant and is generated by the deletion of exon 17 and exon 18. Heterologous expression reveals that exon 17 encodes an essential Nav1.5 region that confers functionality to the channel [74,77]. This splice variant is present in mouse hearts, but there is no evidence of this variant in other mammals’ hearts [74]. On the other hand, deletion of exon 24 of Nav1.5 generates Nav1.5f variant; this isoform is highly detected in rat heart and human brain but not in the human heart [78,79]. Electrophysiological experiments evidenced that Nav1.5f is a non-functional variant [71]. Finally, it has been identified three C-terminal truncated spliced variants, E28B, E28C, and E28D, that generate reduced protein levels and no functional Na+ currents in the normal fetal and adult human heart [80]. In another layer of complexity, in a very recent study, it has been evidenced that minor introns modulate gene families at a post-transcriptional level. Concretely, U6actac, which is a minor spliceosome component, modulates Nav1.5 and Cav1.2 protein levels through the removal of minor introns in Scn5a and Cacna1c, regulating electrophysiological properties of cardiomyocytes [81].

Regulation of SCN5A by Non-Coding RNAs

SCN5A mRNA degradation or translational repression can be mediated by microRNAs. It has been widely demonstrated that microRNAs exert their function controlling the development of several cardiac arrhythmogenic diseases; however, there is not much information about the SCN5A/Nav1.5 post-transcriptional control by microRNAs despite being an important modulator of cardiac arrhythmias [82,83,84,85]. A few years ago, data from our laboratory demonstrated that miR-98, miR-106, miR-200, and miR-219 directly regulated human SCN5A, while miR-125 and miR-153 regulate it indirectly, being evidenced that miR219 and miR200 modulate Scn5a expression in an opposite way in HL1 cardiomyocytes. Concretely, miR-200 decreases, while miR-219 increases Scn5a expression. Moreover, miR-219 is able to increase Nav1.5 protein levels leading to the subsequent rise in INa [86]. In another study, we elucidated a complex regulatory network where Pitx2 transcription factor and microRNAs are involved, being able to modulate ion channel expression in atrial fibrillation [87,88,89]. Additionally, Zhao et al. [90] demonstrated that miR-192-5p exerts its function, reducing Nav1.5 expression and INa density in humans. It has also been evidenced that SCN5A expression can be indirectly modulated through ERG1 mediated regulation by miR-143 [91]. Furthermore, a miR-24 binding site is generated in the coding sequence of SCN5A by the SNP rs1805126, decreasing SCN5A expression and INa density [92]. Finally, it has been evidenced that some SNPs can generate new miRNAs binding sites at the SCN5A 3′UTR region. Previous work in our lab demonstrated that rs4073797 and rs4073796 polymorphisms create a new miR-1270 binding site in SCN5A 3′UTR, which imbalance Scn5a expression. Moreover, rs107822C localized upstream of miR-219a precursor impairs miR-219a expression deregulating SCN5A/Nav1.5 levels [93]. Although INa is the most relevant during cardiac AP, the functional role of microRNAs controlling SCN5A expression is poorly described. Quite recently, a new class of RNAs has emerged, i.e., long non-coding RNAs (lncRNAs), which are long RNA molecules with >200 nt length and scarce protein-coding properties, but being able to modulate several biological processes and thus participating in pathogenesis [94,95]. It has been demonstrated that some lncRNAs modulate cardiac AP regulating the slow component of the delayed rectifier potassium current (IKS) or transient outward potassium current (Ito), concretely, lncRNA-Kcna2as and lncRNA-MALAT1, respectively [96,97], and more recently, it has been elucidated that lncDACH1 modulate Nav1.5 protein distribution through dystrophin binding which determines Nav1.5 membrane anchoring [98]. Studies about how cardiac AP can be modulated by lncRNA need to be further developed, as these molecular mechanisms could reveal a potential therapeutic work line.

3. Non-Genomic Regulation of the Cardiac Sodium Channel

Being an ion channel, Nav1.5 is first synthesized as a primary protein chain that is subsequently folded in order to acquire the pore-forming three-dimensional conformation [11]. This tertiary structure is then assembled with its beta subunits, most likely (β1), and trafficked through the Golgi apparatus to be targeted to the corresponding cell membrane compartments [99]. Along this whole process, Nav1.5 went through distinct non-genomic regulatory modifications and quality control steps conferring its unique conformational and functional identity as a voltage-gated sodium channel [11,100]. These steps are ensured by a growing set of regulatory proteins that have been demonstrated to covalently or non-covalently interact with Nav1.5 [101]. In addition to the interacting proteins, Nav1.5 function has been demonstrated to be influenced by wider intracellular (oxidative stress, metabolic stress, electrolyte homeostasis, etc.) and extracellular (pH, temperature, hormones, etc.) factors. All these factors are discussed subsequently in this review.

3.1. Regulation of Nav1.5 Biosynthesis and Post-Translational Modifications

3.1.1. Regulation of Nav1.5 Translation and ER Retention

The translation of Nav1.5 starts in the cytosol and then pursue into the endoplasmic reticulum (ER). Anchoring the ribosome with the elongating Nav1.5 polypeptide chain to the ER occurs when a signal peptide is recognized by the signal recognition particle (SRP) that targets the active ribosome to the rough endoplasmic reticulum (ER) membrane. Unlike cytosolic proteins, which have their signal peptide generally within the amino terminal, ion channels contain numerous signal sequences that are not restricted to the amino terminal [102]. Although the signal sequences of some ion channels such as Kv1.3 and CFTR have been already mapped to the second transmembrane spanning domain, almost 200 amino acids downstream from the NH2 terminal [102], that of Nav1.5 are not yet identified. Once anchored, the ribosome translocates the elongating polypeptide chain into the ER lumen [103]. As a transmembrane protein, the nascent Nav1.5 is soon pushed to the ER membrane, where it is anchored and retained [103]. The ER retention is thought to occur when specific ER retention motifs embedded in the elongating Nav1.5 polypeptide (most likely in the DI-DII linker of the sodium channel [104]) binds a cytosolic signal recognition particle (endoplasmic reticulum recognition particle, ERRP), that then directs the Nav1.5-ERRP complex to receptors within the ER membrane [105,106,107]. The complex Nav1.5-ERRP is then trapped within the ER, ensuring that the newly formed channel does not leave the ER membrane before finishing the folding and assembly steps [102]. At this level, several regulatory proteins and residues are reported to bind to this complex and facilitate the folding and maturation of the nascent protein [11,102,105,106,107,108,109,110] (Figure 2).
The correct folding of the newly synthesized Nav1.5 channels is commonly thought to be a condition for their forward trafficking to the cell membrane and their proper gating function. This notion has been tested by the exploration of Nav1.5 trafficking-deficient mutants such as R282H, A124D, and V1378M that, due to folding defects, they failed to exit the ER and thus to reach the cell membrane [111,112]. Although the importance of this step in the life cycle of any ion channel, very scarce information is currently available about the mechanism of Nav1.5 folding and its regulation. Nonetheless, it is currently established that one of the prerequisites for proper Nav1.5 folding is core-glycosylation, as will be discussed later in this review [113]. In addition, molecular chaperone proteins such as protein disulfide isomerases (PDI), ER oxidoreductases (ERO), 70 kDa heat shock proteins (Hsp70), 90 kDa heat shock proteins (Hsp90), as well as calnexin and calreticulin, have been demonstrated to regulate the folding of the nascent proteins and the ER-associated degradation of the misfolded proteins [114,115,116,117,118]. However, their implication in the Nav1.5 folding process has not been specifically studied yet. Interestingly, Nav1.5 H558R polymorphism has been shown to have a corrective effect on the misfolded R282H mutant by restoring its trafficking to the cell membrane and thus limiting the misfolding of the mutant through a physical interaction [111,119].
Some antiarrhythmic drugs such as mexiletine, quinidine, and flecainide proved their efficiency rescuing the trafficking of some misfolded Nav1.5 variants, thus playing the role of pharmacological chaperones [112,120]. In addition, curcumin, a major constituent of turmeric known to block the ER calcium pump, has also been reported as effective in rescuing the INa current of L325R misfolded Nav1.5 channels [40]. Low temperature has also been demonstrated to trigger the rescue of misfolded Nav1.5 mutants [120], probably through slowing the folding process, which prevents protein misfolding and aggregation [121].

3.1.2. Co-Translational and Post-Translational Regulation of Nav1.5

N-Linked Glycosylation of Nascent Nav1.5

One of the earliest modifications that the Nav1.5 undergoes co-translationally once inserted into the ER is the N-glycosylation [99,113]. This quality control step has been first evidenced in the rat heart by Cohen and Levitt, who have found that glycosylation increases Nav1.5 mass by only 5%, compared to 25–30% increases observed in other voltage-gated sodium channel isoforms [122]. Glycosylation initiates in the ER and terminates in the Golgi [113,123]. In the ER, glycosylation initiates when glycan (Glc3Man9GlcNAc2) is dissociated from a lipid derivative by oligosaccharyl transferase (OST) and bind to the amide nitrogen of asparagine (N) localized in the extracellular side of the nascent Nav1.5 protein [100,124]. Although no validated “map” of the N-glycosylation sites has been published yet for Nav1.5, 13 potential external N-glycosylation sites have been identified in human Nav1.5 [125], and at least 14 putative N-linked glycosylation sites have been reported in the rat cardiac sodium channel [122]. The N-glycosylation of the newly formed cardiac sodium channel has been reported to be a prerequisite for proper Nav1.5 folding and subsequent surface expression as well as an assembly with its β subunits [99,113,126]. According to Arakel et al., Nav1.5 maturation strongly depends on the presence of the auxiliary β1 that binds to the pore-forming α subunit and promotes its glycosylation and its trafficking to the cell membrane [127].
In this context, N-glycosylated Nav1.5 is thought to undergo subsequent serial de-glucosylation steps and extreme quality controls involving the ER-resident chaperones, which will ensure that only correctly folded and fully glycosylated channels can be trafficked [113,123,128]. Interestingly, Mercier et al. found that early N-glycosylated Nav1.5 channels generated in the ER could reach the cell membrane through an unconventional trafficking pathway bypassing the Golgi stacks while functional channels are trafficked through the conventional pathway that is Golgi-dependent [113]. In addition, ER-resident chaperones such as Calnexin and Calreticulin have been reported to play a crucial role in controlling ion channels folding and efficient export to the Golgi [129,130,131]. However, there is no evidence of physical interaction of Calnexin and Nav1.5 despite their proven co-localization in the ER [132,133]. While properly folded Nav1.5 are trafficked forward to the cis-Golgi where they will be fully maturated, misfolded Nav1.5 are retained in the ER to be later degraded, most likely through the activation of the unfolded protein response (UPR) pathway and/or ER-associated degradation (ERAD) pathway that is linked to the cytoplasmic ubiquitin-proteasome pathway [134,135,136].

Phosphorylation and Dephosphorylation of Nav1.5

In addition to N-linked glycosylation, Nav1.5 undergoes phosphorylation as a post-translational modification [137]. Thirty years ago, Shubert et al. brought the first evidence of Nav1.5 phosphorylation by protein kinase A (PKA) through the activation of the β-adrenergic system by isoproterenol, which led to an increased level of cAMP, which in turn reduced Na+ current (INa) [138]. These findings were further confirmed by a subsequent study by Frohnwieser and his co-worker, who showed that combined cytosolic injection of cAMP and a PKA activator increased INa suggesting a modulatory effect of PKA on human Nav1.5 [139]. The same study demonstrated that this modulatory effect of PKA is conferred by the DI–DII intracellular linker of Nav1.5. In this regard, it has been reported that the rat Nav1.5 protein sequence harbors two distinct sites for PKA phosphorylation that were mapped to serine positions S526 (525 in human) and S529 (528 in human) [100,140,141,142]. These sites are localized in the cytosolic loop interconnecting DI and DII of Nav1.5, where the three putative RXR-type (R479KR481, R533RR535, and R659QR661) ER retention motifs have been localized too [104,137,143]. Zhou et al. have previously demonstrated that PKA activation promotes trafficking of channels to the plasma membrane [143]. In the same context, Scott et al. have shown that a PKA-PKC mediated phosphorylation of NMDA receptor masks its ER retention motifs leading thus to its release from the ER and exportation to the cell membrane [144]. Taken together, these findings suggest a similar mechanism where the phosphorylation of Nav1.5 at S525 and S528 by PKA leads to changes in the Nav1.5 conformation that masks the ER retention signals and eases the trafficking of the channel to the cell membrane [142,145]. This is consistent with the idea that proper folding of Nav1.5 unmasks its ER retention motifs and facilitates its forward trafficking to the Golgi apparatus [106].
In an antagonistic way to PKA, Nav1.5 is downregulated by protein kinase C (PKC)-mediated phosphorylation which leads to a reduced channel density at the cell surface and INa decay [146]. Although ten different PKC isoforms have been identified in human ventricular myocytes and in different animal species [147], isoform-specific activation/inhibition studies suggested εPKC isoform as the key player in the PKC-mediated regulation of Nav1.5 and INa [148,149]. Nonetheless, PKCδ-mediated Nav1.5/INa downregulation either directly through phosphorylation at S1503 or indirectly through elevated mitoROS production has been reported [150]. In addition, a minor role of αPKC reducing INa through angiotensin II has also been described [151]. As a direct mechanism, the PKC (particularly εPKC) effect on Nav1.5 and INa has been partially attributed to the phosphorylation of a conserved serine S1503 of the DIII-DIV cytosolic linker of Nav1.5 [152,153]. However, intracellular metabolic changes have been described as a mediator of PKC activation and PKC-mediated phosphorylation of Nav1.5 [150]. In this regard, high intracellular levels of NADH have been described as triggers of PKC, thus leading to overproduction of mitochondrial reactive oxygen species (mitoROS) and INa decay [154,155,156]. This effect has been demonstrated to be mediated by glycerol 3-phosphate dehydrogenase 1 (GPD1L) [157] and could be reversed by NAD+-mediated PKA activation [154,158,159,160]. Interestingly, Fouda et al. have demonstrated that PKA and PKC phosphorylation pathways could be activated by Cannabidiol and Estradiol and that this activation could rescue the high glucose-induced changes in Nav1.5 properties [161,162].
Importantly, not far from the PKA phosphorylation sites in Nav1.5 DI–DII linker, there is a Ca2+/Calmodulin-dependent Protein Kinase II (CaMKII) phosphorylation site as well, which was mapped to S516 [163]. This CaMKII phosphorylation site is not the only one in Nav1.5 since Ashpole et al. have identified four extra potential sites; all of them are localized in DI–DII linker, suggesting linker I as a hotspot for Nav1.5 phosphorylation [163]. However, a recent study by Herren et al. identified 23 sites along Nav1.5 intracellular regions that could be phosphorylated by CaMKII in human Nav1.5 [164]. More recently, Burel et al. identified two further CaMKII phosphorylation sites localized in the C-terminal region of Nav1.5 [165]. Several studies have shown that Nav1.5 is regulated by CaMKII and that activation of this kinase increases the so-called pathogenic late cardiac sodium current INaL [166]. Interestingly, El Refaey et al. demonstrated that INaL could also be regulated by B56α, the key regulatory subunit of the PP (protein phosphatase) 2A holoenzyme [167]. This phosphatase is targeted by ankyrin-G to the Nav1.5-CaMKII -βIV spectrin axis at the ID where it is thought to dephosphorylate Nav1.5 at S571 in the DI-DII linker via B56α balancing, thus the CaMKII-dependent phosphorylation of the cardiac sodium channel. According to a study by Deschênes et al., inhibition of CaMKII slowed Nav1.5 channel current decay, produced a depolarizing shift in fast inactivation, and slowed entry into inactivated states [168].
Nav1.5 is also phosphorylated by Tyrosine kinases. In this regard, phosphorylation of Nav1.5 by the Src family Tyrosine kinase Fyn has been first reported by Ahern and co-workers, who have demonstrated that this kinase acts by increasing the rates of recovery from fast-inactivated states, thus impairing the steady-state inactivation of Nav1.5 [169]. Fyn kinase acts most likely on Tyr1495 of Nav1.5 not far from the Ile-Phe-Met (IFM) motif of DIII–DIV linker that is known to modulate the rapid inactivation process of the channel [5]. In the heart, Fyn tyrosine kinases are reported to co-localize with Nav1.5 channels at adherens junctions, where they modulate electrical coupling and propagation of action potential [170,171]. Iqbal et al. found that the major Nav1.5 splice variants Q1077 and delQ1077 are differentially phosphorylated by Fyn kinase, which results in coordinated steady-state rapid inactivation kinetics for smooth electrical activity of the heart [172]. The same researchers suggested a multistep mechanism by which Fyn kinases bind and modulate Nav1.5. This mechanism starts by the association of Fyn kinase to proline-rich regions in the DI–DII linker and C-terminal region of Nav1.5, which activates the phosphorylation of neighboring tyrosine residues in the N-terminal region (Y68, Y87, and Y112), DIII–DIV linker (Y1494, Y1495), and C-terminal region (Y1811, Y1889) [169,172,173]. Particularly, Y1494 and Y1495 of the DIII–IV linker have been demonstrated to play an essential role in the anchoring of Ca2+/Calmodulin to the Nav1.5 inactivation gate, and thus Fyn-mediated phosphorylation of the two Tyrosine residues has been suggested to reduce or abolish calmodulin binding and to impair the interaction of the side chain with the inactivation gate receptor [174].
Additionally, Nav1.5 has been reported to be dephosphorylated by the protein tyrosine phosphatase 1 (PTPH1), which interacts with the Nav1.5 PDZ domain binding site at the C-terminal region [101]. PTPH1–mediated dephosphorylation of Nav1.5 modulates its gating by shifting the steady-state inactivation towards hyperpolarized potentials [175].

Arginine Methylation

Beltran-Alvarez and co-workers evidenced for the first time that Nav1.5 is post-translationally modified by arginine methylation at three residues (R513, R526, and R680) within the Nav1.5 DI–DII linker [176]. This modification is catalyzed by arginine methyltransferases (PRMT) PRMT-3 and PMRT-5 and leads to an increased expression of Nav1.5 in cell surface [177]. Studying the PTMs of Nav1.5 in end-stage heart failure patients, the same team demonstrated that methylation of R526 is the major quality control step of any Nav1.5 arginine or lysine residue [178].

N-Terminal and Lysine Acetylation

Another PTM during the Nav1.5 life cycle is the acetylation process. Two types of acetylation have been reported so far: reversible and irreversible. The first type is mediated by histone acetyltransferases (HATs) which exert N-terminal acetylation of a Nav1.5 lysine residue leading to enhanced trafficking of Nav1.5 and therefore to an increased INa current [179], whereas the second type of acetylation is mediated by N-terminal acetyltransferases (NATs), where a Nav1.5 alanine residue is acetylated and has been reported as a Nav1.5 degradation signal [177]. Interestingly, native Nav1.5 channels purified from end-stage heart failure patients were reported to lack the initiation of methionine and be acetylated at the resulting initial alanine residue [178]. Recently, Vikram et al. showed that Nav1.5 undergoes reversible lysine acetylation. For instance, sirtuin 1 deacetylase (Sirt1), an NAD+-dependent lysine deacetylase, has been demonstrated to regulate Nav1.5 channels by deacetylating lysine residue 1479 (K1479) in the DIII–DIV linker, which promotes Nav1.5 cell surface expression and increases INa [180]. Interestingly, the murine model of cardiac Sirt1 deficiency presents fatal cardiac conduction defects as a result of K1479 hyperacetylation, which decreases Nav1.5 cell surface expression and reduces INa. These arrhythmogenic substrates are similar to those characterizing human Nav1.5 loss-of-function cardiac arrhythmias suggesting that Nav1.5 Sirt1-mediated deacetylation is crucial for the proper function of the cardiac sodium channel. It is noteworthy that the authors of this study raised an interesting point regarding the role of the functional interaction and interplay between different PTMs fine-tune regulating the Nav1.5 channel expression and function. In this regard, it has been suggested that Nav1.5 is regulated by Sirt1-mediated interaction between lysine acetylation and the ubiquitination in one hand and NAD+ dependent interplay between PKC-mediated phosphorylation and Sirt1-mediated deacetylation in another hand [180].

SUMOylation

Although more than 25 years have passed since the discovery of SUMOylation, a post-translational modification conjugating a small ubiquitin-like modifier (SUMO) molecule to a lysine residue in the substrate protein [181], very scarce information are currently available about the regulation of Nav1.5 by SUMOylation. For instance, only one study, that of Plant et al., has reported that one of the mechanisms underlying INaL elevation in response to acute cardiac hypoxia is the quick SUMOylation of Nav1.5 channels at the cell surface [182]. Particularly, SUMOylation of K442 residue has been reported to contribute to the pathological increasing of INaL and action potential prolongation through activation of Nav1.5 channels when they should normally be inactivated.

S-Nitrosylation

S-nitrosylation, a PTM consisting of the covalent binding of a nitrogen monoxide (NO) moiety to the thiol side chain of cysteine in the target protein, has recently gained progressive attention as a crucial quality control step that is required for the proper function of a given protein [183]. In the cardiomyocytes, NO is produced by neuronal nitric oxide synthase (nNOS) [184]. nNOS mediated S-nitrosylation of Nav1.5 has been demonstrated to maintain INaL [185]. Interestingly, nNOS has been shown to interact with Nav1.5 via its regulating protein α1-syntrophin, which acts as a scaffolding protein bringing together Nav1.5 with nNOS and plasma membrane Ca-ATPase (PMCA4b) (an inhibitor of nNOS activity) [186]. Therefore, LQTS-associated α1-syntrophin mutation has been demonstrated to break the SNTA1- PMCA4b association neutralizing, thus the nNOS inhibition and increasing Nav1.5 S-nitrosylation, which in turn increase INaL currents [186]. A similar effect has been observed with a decreased caveolin 3(Cav3) expression, which has been shown to enhance S-nitrosylation of Nav1.5 through increasing the nNOS activity, which increased INaL in cardiomyocytes [187]. However, a very recent study by Wang and co-workers suggested an indirect mechanism by which S-nitrosylation modulates the cardiac sodium channel expression and function. For instance, NO has been demonstrated to down-regulate SCN5A expression and Nav1.5 function through S-nitrosylation of regulatory transcription factor FOXO1 [188]. These findings increase our current understanding of the role of redox and free radicals in the regulation of Nav1.5 function (see [100] for further review).

Lipoxidation

Lipoxidation refers to the establishment of covalent adducts between reactive products of lipid peroxidation and macromolecules such as proteins, phospholipids, and DNA [189]. Recently, lipoxidation gained interest as a post-translational modification of the cardiac sodium channel that gives further evidence on the regulation of Nav1.5 by oxidative stress [190]. Nonetheless, little information is currently available about the mechanism of Nav1.5 regulation by lipoxidation. In this respect, in vitro data by Nakajima and co-worker provided the first evidence that Nav1.5 is post-translationally modified by lipoxidation during oxidant injury and that sodium channel dysfunction evoked by lipid peroxidation could be prevented by scavenging Isoketals (IsoKs), which are the most reactive products of lipoxidation [191].

Methionine Oxidation

A previous study by Quiñonez et al. demonstrated that skeletal Nav1.4 fast inactivation could be impaired by oxidizing at least two methionine residues in the channel [192]. These findings have been supported in cardiac Nav1.5 as well, where oxidative modification of the methionine within the IFM motif has been shown to lead to a drastic loss of Nav1.5 inactivation [193]. Interestingly, Nav1.5 channels and INa currents have been reported to be indirectly modulated by CaMKII, the activation of which depends on the oxidation of its own methionine residues [194].

Palmitoylation

Palmitoylation (also called S-acylation) is the PTM of protein cysteines with saturated fatty acids that modify protein hydrophobicity and thereby influence their function [195]. Palmitoylation has been reported to regulate ion channel’s function, most likely through controlling their trafficking and cell membrane expression [99,196]. An early study by Schmidt et al. showed that Nav1.5 is subject to palmitoylation [99]. However, palmitoylation has been demonstrated to slightly influence cell surface expression of Nav1.5 and rather significantly impact channel availability by regulating the voltage dependence of steady-state inactivation in both HEK293 cells and cardiomyocytes [197]. Additionally, cysteine residues predicted to be palmitoylated in Nav1.5 are mapped to the DII–DIII linker of the channel by prediction algorithms [197].

3.1.3. Regulation of the ER-to-Golgi Trafficking

Well folded and assembled proteins are supposed to cross the ER-Golgi space in vesicle budding guided by cytoskeletal proteins [198]. Studying the subcellular distribution of the cardiac sodium channel Nav1.5 in HEK293 Cells and canine cardiac myocytes, Zimmer et al. noticed an accumulation of the intracellular channels within the ER and a lower channel density in the Golgi apparatus. Thereby, they proposed that ER plays the role of an intracellular reservoir where sodium channels are transiently stored [199]. As discussed previously, stimulation of PKA likely results in the activation of the ER-to-Golgi trafficking, which in turn leads to a rapid increase of the channel density in the cell membrane [104]. However, the whole mechanisms underlying the ER exit of Nav1.5 to the Golgi is not yet fully deciphered, and current advances in this topic show that not only the PKA-mediated phosphorylation of the Nav1.5 ER retention sites is what facilitates its ER-Golgi exportation. That is, several proteins and enzymes have been reported to bind to Nav1.5 once retained to the ER and enhance its release. In this context, Wu et al. have identified the Ran-guanine nucleotide release factor (RANGRF or MOG1) as a cofactor of Nav1.5, which by binding to its intracellular loop DII–DIII facilitates its cell surface expression [200]. Using the DII–DIII linker of Nav1.5, in yeast two-hybrid analyses, the team demonstrated that MOG1 is crucial for the optimal expression of Nav1.5 and promotes its ER export and intracellular trafficking to the plasma membrane [200]. These findings are consistent with Chakrabarti et al. study, which showed that silencing of MOG1 expression by small interfering RNAs caused retention of Nav1.5 in the ER, reduced Nav1.5 plasma membrane expression, and disrupted the Nav1.5 targeting to the cell surface, in particular, to the caveolin-enriched microdomains (caveolae) [201]. A subsequent mutational study performed by Yu et al. further revealed that mutations in the amino acids E83, D148, R150, and S151 of MOG1 disrupt its interaction with Nav1.5 and significantly reduce the cardiac sodium channel trafficking to the cell surface, suggesting that these amino acids are important for the MOG1-Nav1.5 binding and interaction [9]. The same team found that MOG1-mediated trafficking and function of Nav1.5 requires the interaction of MOG1 with two small GTPases SAR1A and SAR1B and that the knockdown of both enzymes abolishes the function of MOG1 [202]. Furthermore, it has been demonstrated that activation of SAR1 leads to the recruitment and internalization of Nav1.5 cargo into the coated transition vesicle COPII-coated vesicles that will ensure its ER-to-Golgi trafficking [202]. The Nav1.5 ER export is also controlled by Dynamitin as demonstrated by Chatin et al., who have proved, using a yeast two-hybrid system, that Dynamitin (C-terminal domain), interacted with the Nav1.5 DI–DII linker between amino acids 417 and 444 and that this interaction is crucial for the Nav1.5 cell-surface density probably through controlling the ER-to-Golgi trafficking [203].

3.1.4. Regulation of Nav1.5 Maturation and Golgi Export

Once in the Golgi, N-glycosylated Nav1.5 undergoes additional mannose trimming and terminal glycosylation where acetyl-glucosamine, oligosaccharides, and finally sialic acid residues are sequentially added as the protein crosses the distinct Golgi cisternae. It has been demonstrated that glycosylation regulates voltage-gated sodium channels (including Nav1.5) gating, inactivation, and recovery process during cardiac AP by interfering with the electric field near the gating sensors [204,205,206,207,208]. Hence, it has been suggested that extracellular sialic acid residues, which are negatively charged at physiological pH, modulate the sensitivity of the Nav1.5 voltage sensor domains to the transmembrane electrical potential fluctuation [209]. Particularly, sialic acid residues localized to DI S5-S6 have been demonstrated to regulate the sialic acid-dependent gating of Nav1.5 [125].
Mature Nav1.5 (fully glycosylated) are exported from the Golgi apparatus, which acts as a major secretory sorting hub that targets newly synthesized proteins to their final subcellular destinations [210]. Although the current knowledge on the exact mechanisms regulating the Nav1.5 export from the Golgi and trafficking to the cell membrane is still limited, a recent study by Ponce-Balbuena and co-workers reported that Nav1.5 Golgi export is driven by a trafficking signal localized in its terminal COOH region. This signal corresponds to a binding site of the adaptor protein complex 1 (AP1) mapped to Nav1.5’s Y1810 residue. AP1-marked Nav1.5 will be then incorporated into clathrin-coated vesicles that will migrate to the cell membrane where the channel will be anchored [211]. The same team showed that the Nav1.5 cross the Golgi-cell membrane space by a common anterograde trafficking pathway as Kir2.1. These findings support previous studies demonstrating that both ion channels form a channelosome that shares common trafficking, targeting, anchoring, recycling, and degradation pathways [212,213].

3.1.5. Regulation of the Nav1.5 Targeting to the Cell Membrane

Over the last few years, it became widely accepted that not all the Nav1.5 proteins synthesized in one cardiomyocyte undergo the same regulatory steps till reaching their final localization in the cell membrane [214]. After years of debate and controversial studies about the subcellular distribution of the cardiac sodium channel, the new cellular imaging techniques excluded the idea of an exclusive expression of Nav1.5 at the ID [171,215] and gave way to a more conceivable model that suggests a multi-pool aggregation of Nav1.5 along with the cellular membrane compartments including the LM and the T-tubules [216,217,218]. Being in one membrane domain or the other put the Nav1.5 in distinct microenvironments composed of different interacting proteins that regulate its gating function and biophysical properties. Above all these interacting proteins, beta subunits are without doubt the ones that most gained interest in this field over the last few decades as their presence and function are dependent on the presence of the pore-forming α-subunit (Figure 2 and Figure 3).

Regulation of Nav1.5 by β-Subunits

The β subunit family consists of four different proteins β1–4 encoded by four genes, SCN1B–SCN4B, respectively, with β1 alternatively spliced into two isoforms, β1A and β1B [219]. As mentioned earlier in this review, the β-subunits, most likely β1-subunits, assemble with Nav1.5 at the endoplasmic reticulum and influence its maturation and trafficking to the plasma membrane [127,220]. Alpha-beta subunits assembly is either covalent (β2 or β4) or non-covalent (β1 or β3) [221]. Particularly, β 4-Nav1.5 covalent association is ensured by an extracellular cysteine–cysteine single disulfide bond [222,223], while β2 does not form a disulfide linkage at this position with Nav1.5 as recently specified [5], whereas β1 and β3 non-covalently interact with Nav1.5 through the channels DIV and DIII voltage gating domain respectively [224].
Despite the structural similarities between β2/β4 on one hand and β1/β3 on the other hand, their expression differs from one cellular sub-domain to another. Inside the cardiomyocyte, β3 are expressed at the T-tubules and β4 at the ID, while β1 and β2 are found at both locations [215,225,226]. Zimmer et al. have suggested that, unlike β2, β1 associates to Nav1.5 early at the ER, and both α and β1 subunits are trafficked together to their final destination at the cell membrane [227]. Subsequent studies revealed that β1-subunits enhance the α-subunits dimerization and promote the dominant-negative effect of trafficking defective mutants [228]. β2 has been reported to promote surface localization of Nav1.5 [229]. Importantly, β3 subunits have been demonstrated to bind to Nav1.5 in multiple sites and promote the formation of α subunit oligomers, including trimers [230]. However, β4 has been reported as a modulator of Nav1.5 kinetic and gating properties by increasing INa [231]. Taken together, these findings are consistent with the idea that the distinct sodium channel β subunits provide support for the pore-forming subunit, facilitate the trafficking of the mature channel to the different membrane domains, and modulate the gating function of Nav1.5 by increasing the INa [232,233,234,235,236]. More details regarding the regulation of Nav1.5 by β subunits in the context of sodium channelopathies are discussed in Section 4 of this review.

The Nav1.5 and the Intercalated Disc Interactome

As suggested by the Delmar research team, several evidence point to the fact that the ID is not a hub of proteins playing independent functions within the cardiomyocyte, but rather a network of molecules interacting together in order to fulfill a specific function (AP propagation, cell-to-cell coupling, cardiac excitability, etc.) that cannot be accomplished if this “interactome” is impaired [237]. As a component of the ID proteins, Nav1.5 has been demonstrated to be in the heart of this interactome by physically and functionally associating to several proteins belonging to this macromolecular complex.
In this context, it is currently well known that Nav1.5 targeted to the ID are “tagged” with synapse-associated protein 97 (SAP97), a scaffolding MAGUK ((membrane-associated guanylate kinase) protein that is abundantly expressed in human and rat ventricular myocardium [238]. SAP97 has been introduced as the determinant of the Nav1.5 ID pool as it plays an important role in targeting Nav1.5 along with Kir2.1 to this cell membrane domain [238,239]. Both channels were structurally evidenced to co-assemble to SAP97 by their C-terminal domains [238,240]. For Nav1.5, it is assumed that the last three amino-acids (serine–isoleucine–valine or SIV motif) of the C-terminal region form a PDZ (postsynaptic density protein (PSD95), Drosophila disc large tumor suppressor (Dlg1), and zonula occludens-1 protein (zo-1) domain binding motif) that interacts with the syntrophin–dystrophin complex at the cardiomyocyte LM and PDZ domains of SAP97 at the ID [218]. In the absence of the PDZ-domain-binding motif of Nav1.5 or SAP97, Nav1.5 expression at the cell surface decreased, thus leading to a reduction in the cardiac INa in vitro [241]. However, a subsequent study by the same team demonstrated that in vivo ablation of SAP97 did not change Nav1.5 localization and function, but it did decrease the cardiac potassium currents [242]. The authors of these studies justified this discrepancy by the fact that SAP97 silencing in vitro is induced in adult cardiomyocytes while in vivo, it is a constitutive ablation present early in development, which may impact protein expression and interactions.
In addition, the Nav1.5-SAP97-Kir2.1 complex has been demonstrated to reach the ID through the microtubule highway [133,238,239,243]. Although the exact mechanism by which Nav1.5 is targeted to the ID is not yet fully discovered, part of it is already elucidated. A few years ago, Agullo-Pascual et al. proved for the first time that the microtubule plus-end tracking protein “end-binding 1” (EB1) is captured to the IDs by connexin 43 (cx43), which facilitates the cargo delivery, including Nav1.5 [244]. These findings are consistent with Marchal and co-workers’ recent study in which they have further proved that EB1 modulates Nav1.5 trafficking to the IDs and that loss of EB1 function leads to reduced INa and conduction slowing [245]. Moreover, EB1 has been previously demonstrated to bind directly to CLASP2 (cytoplasmic linker associated protein 2) and form a complex at the microtubule plus-end, promoting thus microtubule polymerization and stabilization [246]. Interestingly, inhibiting the GSK3β (glycogen synthase kinase 3β)-mediated phosphorylation of CLASP2 enhanced the EB1–CLASP2 interaction, which in turn led to an increased Nav1.5 delivery at the ID of cardiomyocytes and an increased INa [245]. Furthermore, Rhett et al. have shown that in addition to its known localization at the gap junction where it interacts with zonula occludens-1 (ZO-1) [247,248], Cx43 also co-localizes with ZO-1 in the zone surrounding the gap junction, conventionally termed as perinexus and that Cx43 but not ZO-1 interact with Nav1.5 at this zone in physiological conditions [249]. In vivo and in vitro assays show that Nav1.5 expression and function are reduced as a result of Cx43 expression/function decrease, thus giving more evidence that Cx43 is required for a proper Nav1.5 function at the ID [250].
Importantly, Nav1.5 and Cx43 interaction at the perinexus is thought to be mediated by scaffolding proteins SAP97 and Ankyrin G (AnkG) as their interaction has been reported [241,251]. In the cardiovascular system, ankyrins are critical components of ion channels and transporter signaling complexes, and their dysfunction has been linked with abnormal ion channel and transporter membrane organization and fatal human arrhythmias [252]. Although both ankyrin-B (AnkB, encoded by ANK2) and ankyrin-G (ANK3) have been found to be expressed in the myocardium, only ankyrin-G has been shown to interact with Nav1.5 [253]. Specifically, AnkG is necessary for normal expression of Nav1.5 and acts as a coordinating signaling center, functionally coupling Nav1.5 gating with upstream kinase and phosphatase enzymes and downstream cytoskeletal proteins [110,254]. AnkG is primarily expressed at the ID membrane and T tubules, where it co-localizes with Nav1.5 [142]. In vitro, it has been demonstrated that AnkG binds to Nav1.5 and that AnkG downregulation impaired the subcellular localization of Nav1.5 and reduced the INa current amplitude [255,256]. In vivo, Makara and his collaborators have demonstrated that AnkG plays an indispensable role in directing Nav1.5 and its regulatory protein CaMKII to the ID [254,257]. Mutational studies have further confirmed that disrupting the binding of AnkG to Nav1.5 impairs AnkG dependent targeting of the Na+ channel to the ID leading thus to a reduction in INa density and cardiac arrhythmias [253,254,258]. A recent study performed by Yang et al. has demonstrated that AnkG, but not AnkB, are expressed at the IDs and that masking Nav1.5 binding sites in AnkG using competitive peptides caused a decrease in sodium channel current (INa) and targeting defects of the Na+ channels to the ID, but not to LM [213]. However, a more recent study by Cavus and collaborators specified that only canonical AnkG isoforms have this regulatory effect on Nav1.5 and that noncanonical (giant) AnkG isoforms mediated electrical dysfunction is independent of Nav1.5 [259].
Furthermore, AnkG is thought to mediate the interaction between Cx43 and PKP2, thus connecting desmosomal proteins with the molecular complex that captures the microtubule plus-end at the ID, thus allowing for delivery of Nav1.5 [244,256,260]. This is consistent with the fact that loss of desmosomal integrity impacts cardiac conduction and leads to cardiac arrhythmias [260,261,262]. Accordingly, loss of Plakophilin-2 (PKP2), a crucial component of the cardiac desmosome, has been demonstrated to decrease INa in cardiac myocytes [263]. Similarly, loss of PKP2 expression in HL1 cells and in induced pluripotent stem cell-derived cardiomyocytes (iPSC-CMs) from a patient with PKP2 deficiency reduced INa amplitude [261,264]. Likewise, Rizzo et al. have demonstrated that desmoglein-2 (Dsg2), another desmosome protein, physically interacts with Nav1.5 at the ID of mouse cardiomyocytes in vivo [265]. They showed that mice models over-expressing a desmoglein-2 mutation present a wider intercellular space at the level of the ID, longer ventricular activation time, lower conduction velocity, lower upstroke velocity, and lower INa amplitude compared to wild type. Although no evidence of direct interaction between Desmoplakin (DSP) and Nav1.5 has been reported, RNAi-based Desmoplakin silencing in vitro resulted in a reduction in Nav1.5 expression at the ID of cardiomyocytes, an abnormal sub-cellular distribution of Cx43 and Nav1.5, INa decay, and slowed conduction velocity suggesting that DSP regulates Nav1.5 [266].
Similarly, AnkG is established as an adaptor protein that organizes, transports, and anchors Nav1.5 to the actin/spectrin cytoskeleton [267,268,269]. In fact, the AnkyrinG-Nav1.5 complex is believed to connect with the actin/α-spectrin cytoskeleton through CaMKII-βIV-spectrin interaction where the latter acts as a CaMKII-anchoring protein and thereby orchestrating the whole macromolecular complex; however, no evidence of direct interaction between Nav1.5 and βIV-spectrin has been found yet [255]. On the other hand, βIV-spectrin is assumed to control the CaMKII-dependent regulation of Nav1.5 at the ID, and loss of βIV-spectrin/CaMKII interaction precludes CaMKII-dependent phosphorylation of Nav1.5 at Serine 571 in the DI–DII linker and abolishes the stress-induced activation of the pathogenic INa,L [270,271].
Remme’s team [272] has recently demonstrated that ID Nav1.5 physically interacts with coxsackie and adenovirus receptor (CAR), a single-pass transmembrane cell adhesion molecule (CAM) [273]. Furthermore, they have demonstrated that CAR haploinsufficiency decreased INa amplitude at the ID, which in turn reduced sodium channel availability at this cell membrane compartment. Nav1.5–CAR interaction is only beginning to be understood, and thus, mechanisms underlying this interaction are still to be studied.
Our current understanding regarding the Nav1.5 auto-regulation is still limited. Over the last decades, several controversial studies emerged regarding the sodium channel α-α-subunits interaction and dimerization. However, Clatot and co-workers settled this controversy by demonstrating for the first time that trafficking-defective Nav1.5 exerts a dominant-negative effect on non-defective ones through α-α-subunits physical interaction at their N-terminal regions, precluding thus their cell surface expression [274]. Building on these findings, the team further evidenced that cardiac sodium channel α-subunits assemble as dimers with coupled gating and that this dimerization is mediated through an interaction site found within the DI-II linker of Nav1.5, between amino acids 493 and 517 [275]. Curiously, earlier studies have shown that 14-3-3 protein, a member of highly conserved cytosolic acidic proteins, physically interacts with the DI-II linker of Nav1.5 (between amino acid 417 and 467) at the ID and that this interaction facilitates the dimerization of cardiac sodium channels [276]. Strikingly, Clatot et al. identified a second 14-3-3 protein-Nav1.5 interaction site between amino acid 517–555 and demonstrated that co-operative gating behavior but not dimerization of α-subunits is dependent on 14-3-3-Nav1.5 interaction [275].

Nav1.5 and the Lateral Membrane’s Interactome

Nav1.5 targeting to the LM has been demonstrated to be mediated by the syntrophin–dystrophin complex [3,241]; however, a sub-pool of Nav1.5 at the LM, which is independent of syntrophin, has been recently characterized as well [277]. Dystrophin is known to indirectly mediate Nav1.5 expression at the LM through binding to Syntrophin adapter protein which physically associates to the PDZ domain-binding motif at the C-terminal region of Nav1.5 [3,241,278,279,280]. Interestingly, Matamoros et al. demonstrated that α1-syntrophin also interacts with the N-terminal region of Nav1.5 through an “internal” PDZ-like binding domain localized at this region which acts as “chaperone-like” domain that increases Nav1.5 density at the LM and INa [281]. The same mechanism has been validated for Kir2.1 and Kir2.2 that were demonstrated to reciprocally interact with Nav1.5 channels and modulate each other’s trafficking and expression [281,282].
Interestingly, Nav1.5 has been demonstrated to interact with CASK (calcium/calmodulin-dependent serine kinase), a member of the MAGUK protein family [283]. In several ways, CASK is considered an unconventional Nav1.5 regulator since it is the only MAGUK protein that is lateral membrane-specific and also the only Nav1.5 interacting protein that exerts a repressive effect on the functional expression of Nav1.5, most likely by preventing its early trafficking to the LM. In this regard, CASK has been demonstrated to decrease INa when the former is over-expressed and to increase INa when CASK is inhibited in vivo and in vitro [283].
In addition, Nav1.5 has been evidenced to interact with members of the Z-line scaffolding protein complex, such as α-actinin-2 and telethonin. While α-actinin-2 is currently known to physically interact with Nav1.5 through the channel DIII–DIV linker [284], the telethonin interaction site on Nav1.5 has not yet been identified [101]. α-actinin-2 is thought to positively regulate Nav1.5 by increasing its cell surface expression, most likely through promoting its anchoring to the contact zones between T-tubules and Z-lines and connecting the channel to the actin cytoskeleton network [284]. However, scarce information is available regarding the mechanism of Nav1.5 regulation by telethonin, although physical interaction between TCAP and Nav1.5 was evidenced by co-immunoprecipitation methods and mutations in the telethonin coding gene (TCAP) has been found to alter the channel-gating properties of Nav1.5 in patients with abnormal gut motility and Brugada syndrome [285,286].
Moreover, the role of fibroblast growth factor homologous factors (FHFs), a subset of the fibroblast growth factor (FGF) family [287], has been well elucidated modulating the neuron voltage-gated sodium channels [288]. However, their role in controlling cardiac sodium channel function is still poorly understood and subject to debate. In this respect, fibroblast growth factor homologous factor 1B (FHF1B), also known as FGF12B, has been reported to regulate the biophysical properties and kinetics of Nav1.5 through its physical interaction (amino acids 1773–1832) with the Nav1.5 C terminal region [289]. Both in vitro data show that FHF1B interacts with Nav1.5, and this interaction results in hyperpolarizing shift in steady-state inactivation of this channel [289]. However, the opposite effect has also been reported where a depolarizing shift in the V1/2 of steady-state inactivation has been attributed to the FHF1B-Nav1.5 interaction [290]. Furthermore, FGF13 (FHF2), which is the major FHFs in adult mouse hearts, has been identified as a Nav1.5 interacting protein [290]. In the cardiomyocyte, FHF2 co-localizes with distinct Nav1.5 pools, i.e., the LM and ID suggesting an important role for FHF2 modulating Nav1.5 cell surface expression and function [291]. Like FGF12B, FGF13 physically binds to Nav1.5 through the channel’s C terminus region. In vivo, FGF13 knockdown altered Nav1.5 function resulting in a decreased INa current density, reduced Nav1.5 channel availability, slowed Nav1.5, and reduced INa current recovery from inactivation [290]. This effect of FGF13 is isoform-specific [292]. FHFs have also been implicated in voltage-gated sodium channel trafficking control. In this context, FGF14 has been reported as a modulator of Nav1.5 current densities in neurons and in the heart by impairing their biophysical properties or by controlling channel trafficking and cell surface expression in vitro [293].
Furthermore, calmodulin (CaM), a ubiquitous Ca2+-sensing protein, has been reported to interact with Nav1.5 N- and C-terminal regions [294,295,296,297] and the DIII–IV linker [174,295]. This interaction has been demonstrated to enhance slow inactivation and modulate Nav1.5 gating [296], while disruption of CaM binding to Nav1.5 decreases channel activity and enhances the propensity for persistent Na+ current, all resulting from a switch in the NaV inactivation mechanism [297]. Nav1.5–CaM interaction has been further studied in a mutational context related to cardiac sodium channelopathies (See Section 4).
Finally, dipeptidyl peptidase-like protein-10 (DPP10), previously reported as a modulator of Kv4.3-current kinetics [298], has recently emerged as a new regulator of Nav1.5 [299]. In vivo, DPP10 has been reported to modulate Nav1.5 current kinetics as well by altering voltage dependence of Na+ current and upstroke velocity of the action potential [299].

The Caveolar Nav1.5

Cardiac sodium channels have also been localized to cardiomyocyte caveolae, which are specialized subsarcolemmal membrane compartments enriched in lipids and play a crucial role in vesicular trafficking and protein targeting to the cell surface [300,301]. Caveolar Nav1.5 is exposed to a very rich macromolecular complex encompassing fatty acids, ion channels (pacemaker channels, potassium channels, calcium channels, etc.), and signaling complexes (G-protein-coupled receptors, protein kinases, etc.). This microenvironment has been reported to regulate Nav1.5 function and membrane expression in a multilayers fashion [301].
The first layer is related to the biochemical properties of caveolae itself as a specialized lipid raft rich in fatty acids. In this regard, previous reports demonstrated that Nav1.5 is blocked by polyunsaturated fatty acids (PUFAs), suggesting that interaction of Nav1.5 with the caveolar lipids that also include PUFAs might have the same effect [302,303]. Nonetheless, the mechanism by which caveolar lipid rafts regulate Nav1.5 is not yet fully understood.
The second layer of caveolar Nav1.5 regulation is mediated by caveolins which are the major proteins of caveolae [301]. This mechanism was first reported by the Shibata group, which demonstrated that in addition to the indirect β-adrenergic regulation of Nav1.5, which is PKA-dependent, stimulation of the β-adrenergic pathway in the presence of a PKA inhibitor, activates G-protein (Gsα) cascade, which in turn leads to a rapid increase of INa [300]. A subsequent study by the same group suggested that caveolar Nav1.5 channels are stored at caveolae invaginations and that PKA-independent Gsα-dependant stimulation of the β-adrenergic pathway leads to the opening of caveolae, the exposition of Nav1.5 channels to the extracellular environment, which in turn increase INa [304]. This mechanism has been completely neutralized by anti-caveolin 3 antibodies dialyzed into the myocytes suggesting that caveolar Nav1.5 function is dependent on the Gsα-Caveolin 3 (Cav3) interaction [304]. Although Nav1.5 has been confirmed to interact with caveolin 3 in rodent and human cardiomyocytes [300,305], it is not yet clear if this interaction is direct or indirect. Several reports suggested that Cav3 modulates Nav1.5 function indirectly through inhibiting the nNOS, which is a part of the Nav1.5-SNTA1-PMCA4b macromolecular complex [305,306]. As mentioned earlier in this review, a decay in Cav3 expression has been demonstrated to activate S-nitrosylation of Nav1.5 through increasing the local NO production, which increased INa,L in cardiomyocytes [187].

3.1.6. Regulation of Nav1.5 Degradation

Maintaining the balance between protein synthesis and degradation is crucial for the fine-tune regulation of Nav1.5 levels [307]. In fact, it is currently well established that internalization and degradation of Nav1.5 are regulated either by ubiquitination, covalent attachment of ubiquitin moieties [308], or autophagy [309]. The first mechanism is mediated by the interaction of C-terminus PY motifs of Nav1.5 with the fourth tryptophan-rich domain (WW) of E3 ubiquitin ligase NEDD4-2, which leads to the labeling of Nav1.5 by ubiquitin residues that will be later recognized by the degradation machine [100,310,311]. Interestingly, yeast two-hybrid data demonstrated that the interaction between Nav1.5/αβ-Cristallin from one hand and αβ-Cristallin/Nedd4-2 from another hand reduced internalization of cell surface Nav1.5 and ubiquitination of Nav1.5 [312]. Similarly, serum and glucocorticoid inducible kinase (SGK) has been reported to regulate Nav1.5 degradation by phosphorylating and inhibiting Nedd4-2 [313], whereas UBC9, a SUMO-conjugating enzyme, has been shown to promote Nav1.5 ubiquitination [314]. A very recent study by Liu et al. demonstrated that Nav1.5 ubiquitination would be downregulated by the association of FAT10, a small ubiquitin-like modifier, to the C-terminal lysine residues of Nav1.5, thus decreasing the binding of Nav1.5 to the Nedd4-2 and preventing its degradation [315].
Nedd4-2 has been reported as a direct target of AMP-activated protein kinase (AMPK) in epithelial cells [316]. However, a recent report by Liu X et al. attributed a Nedd4-2 independent Nav1.5 degradation mechanism to AMPK [309]. AMPK, through phosphorylating Nav1.5 T101 residue, facilitates the association of the channel to the autophagic adapter protein and microtubule-associated protein 1 light chain 3 (LC3) and exposes the complex to the autophagic degradation machinery [309].

3.1.7. Effect of Gonadal Hormones on Nav1.5 Expression and Function

The male predominance of some sodium channelopathies such as Brugada syndrome has been extensively studied over the last few years, thus questioning a possible link between sex hormones and Nav1.5 [317,318]. However, comparing the expression levels of Nav1.5 between normal male and female human hearts showed no difference [319]. In addition, concentration-related block of Nav1.5 by estradiol showed that estradiol could not reduce the current of Nav1.5 [320], although a slight reduction in INa currents has been observed at a high concentration of estradiol in vitro [321]. Yang et al. have recently studied the expression levels and function of Nav1.5 in HEK293 cells co-expressing SCN5A (wild-type or BrS mutants R878C and R104W) and sex hormone receptors. They whereby showed that sex hormones have no effects on the expression level of SCN5A (either WT or mutant) and INa currents [322]. Similarly, gonadal hormones testosterone and estrogen showed no effect on fast INa in a canine model [323]. However, a recent study by Hu et al. demonstrated that estrogen through its rapid signal receptor GPER ameliorated the damaging effects of stress in human induced pluripotent stem cell-derived cardiomyocytes (hiPSC-CMs) model mimicking β-adrenergic overstimulation [324]. Taken together, these findings demonstrate that our current knowledge on the regulation of Nav1.5 by sex hormones is still limited and that further studies in this regard are necessary.

3.1.8. Effect of Temperature and pH on Nav1.5 Expression and Function

Febrile states and acidosis are two environmental factors that have been extensively studied as non-genomic modulators of Nav1.5 function in health and disease. It is currently well known that Nav1.5 kinetic is temperature and pH-sensitive [325]. In this context, mild hypothermia has been described as an antiarrhythmic factor that maintains myocardial conduction during prolonged ischemia by sustaining Nav1.5 and Cx43 function [326], whereas hyperthermia has been described as a proarrhythmic factor, especially in combination with SCN5A mutations as is the case in Brugada syndrome [327,328,329,330]. Two mechanisms have been suggested so far for the temperature-dependent regulation of Nav1.5. The first one is a direct mechanism by which temperature accelerates the inactivation of only the wild-type Nav1.5 channels in heterozygous patients, which results in the misbalance between depolarization and repolarization currents and thus may lead to fever-induced arrhythmias [40,331]. The second mechanism is indirect by which temperature modulates the function of Nav1.5 interacting proteins, which in turn modulate Nav1.5 function as is the case of FGF13 [332].
Similarly, fluctuation of the extracellular pH has been demonstrated to influence the Nav1.5 function. For instance, acidic extracellular pH has been shown to modify wild-type Nav1.5 kinetics by destabilizing both the fast inactivated and the slow inactivated states of Nav1.5 [333]. In addition, it has been reported that extracellular protons disrupt charge immobilization which leads to the destabilization of the Nav1.5 fast-inactivation through direct interaction with outer ring carboxylates of the Nav1.5 DIII or DIV [334]. Particularly, His-880 and Cys-373 were identified as the key mediator of Nav1.5 sensitivity to pH fluctuation, where Cys-373 is responsible for isoform-specific proton modulation of use-dependent inactivation of Nav1.5 [335].

4. Cardiac Sodium Channelopathies

SCN5A dysfunction has been extensively reported in distinct cardiac channelopathies such as long QT syndrome, Brugada syndrome, atrial fibrillation, ventricular fibrillation, sick sinus syndrome, and sudden infant death syndrome, as well as in complex electrophysiological disorders that combine several of the previously mentioned channelopathies. In the next paragraphs, we provide an updated view of the mutational, genomic, and non-genomic contribution to SCN5A dysfunction on each of these channelopathies (Table 1).

4.1. Long QT Syndrome

Seminal studies by Wang et al. [336] determined a causative link between SCN5A mutations and long QT syndrome (LQT3). These authors also determined the biophysical and functional characteristics of the novel identified mutations that displayed non-inactivating INa amounting to a few percent of the peak inward INa, as well as impairment on voltage dependence and rate of inactivation and the rate of recovery from inactivation [337]. Since these early studies, more than 80 different mutations have been identified in SCN5A associated with LQT, accounting for 5–10% of the cases (see for recent reviews [14,338,339,340,341]). Functional analyses of several of the identified LQT SCN5A variants displayed gain-of-function either by increasing the late phase of the INa or increasing the window current or both conditions simultaneously, yet our current understanding of the functional roles of most SCN5A mutations described remains elusive.
In addition to mutations on SCN5A associated with long QT syndrome, several SCN5A/Nav1.5 interacting proteins have also been associated in this context. Mutations in the sodium channel ancillary protein SCN1B are associated with LQT, leading to increased INaL [342], while mutations in SCN4B have also been identified [231], yet its plausible implication in LQT is still disputed [343]. Besides these ancillary subunits, a large number of proteins have been reported to interact with SCN5A/Nav1.5, as recently reviewed by Abriel Hugues [101] and detailed in the previous subchapters of this manuscript. In this context, mutations in syntrophin (SNTA1) and caveolin (Cav3) are associated with long QT and disrupt sodium channel function, increasing the INaL [186,187,344,345,346,347]. On the other hand, mutations in BAG3, ankyrin B (ANKB), and α-actinin (ACTN1) leads to LQT, but it remains unclear how such mutations affect the INa [294,348,349], while calmodulin (CaM) and telethonin (TCAP) have been implicated in SCN5A/Nav1.5 interaction and function [163,285] yet to date no mutations have been reported in the context of LQT syndrome.
The genomic contribution to long QT syndrome in transcriptional regulators of SCN5A expression is still incipient since only variants on TBX5 have been reported but not in any of the other transcriptional modulators. Variants affecting transcriptional regulators influencing SCN5A expression have been recently linked to long QT syndrome. In particular, Markunas et al. [350] described a TBX5 variant that co-segregated with prolonged QT interval in a family with otherwise genotype-negative LQTS and demonstrated that such variant impaired the transactivation capacity of this transcription factor. Nieto-Marín et al. [351] reported two additional TBX5 variants that co-segregated with LQT and BrS patients and electrophysiologically impaired INa currents in vitro.
Several SCN5A mutations leading to splice donor variants have been associated with LQT [339,352,353,354,355,356,357]; however, to date, the functional relationship between the distinct SCN5A isoforms and this syndrome remains to be elucidated. While our current understanding of distinct non-coding RNAs in cardiovascular pathology is increasingly emerging, no data are available regarding the functional contribution of microRNAs and/or lncRNAs to long QT syndrome physiopathology. Similarly, the contribution of post-transcriptional SCN5A/Nav1.5 regulation by phosphorylation, glycosylation, acetylation, and/or methylation to QT syndrome has not been reported, and the contribution of ubiquitination is currently controversial [358,359]. Importantly, acetylation in KCNH2, i.e., another ion channel that, if mutated, contributes to long QT syndrome, has been reported [360], opening the possibility that SCN5A/Nav1.5 post-transcriptional regulation also contributes to LQT.

4.2. Brugada Syndrome

SCN5A mutations were associated for the first time to the right bundle branch and ST-elevation syndrome, i.e., Brugada syndrome by Chen et al. [25]. These authors identified a two-nucleotide insertion after the first four nucleotides of the splice donor sequence in intron 7, leading to impaired SCN5A splicing. They also identified a deletion of a single nucleotide in the SCN5A gene that resulted in the elimination of two transmembrane domains and the C-terminal portion of the cardiac sodium channel. Since then, an increasing number of SCN5A mutations have been reported to be associated with Brugada syndrome. To date, more than a hundred mutations have been reported in BrS. Functional analyses of several of these SCN5A reported mutations to demonstrate in most cases, a loss-of-function is achieved, either by decreased Nav1.5 expression in the sarcolemma, because the channels are non-functional or because there is an impaired gating off the channel that results in decreased INa current. However, only a relatively limited number of SCN5A mutations associated with BrS have been fully electrophysiologically characterized.
Importantly, single point mutations have also been associated with both long QT and BrS phenotypes [26,361], and several of these have been electrophysiologically characterized [24,362,363,364,365], yet the precise molecular mechanism of their dual action remains enigmatic.
As previously reported in the context of LQT, in addition to mutations on SCN5A, several SCN5A/Nav1.5 interacting proteins have also been associated with Brugada syndrome. Mutations in the sodium channel ancillary proteins SCN1B [232,356,359,366,367,368,369,370], SCN2B [229], and SCN3B [356] are associated with Brugada syndrome, leading in several cases to decrease INa current [366] and/or cellular trafficking defects [229,356]. Besides the role of the sodium channel ancillary subunits, additional SCN5A/Nav1.5 interacting proteins have been reported in BrS. Mutations in Plakophilin (PKP2) [260,371] MOG1 [9,201,372,373,374,375], FGF13 [376], syntrophin (SNTA1) [377], NEDD4 [378,379], Tmem168 [379] and telethonin [285,286] are identified in BrS patients and their implication to sodium channel function has been reported. Distinct Nav1.5 interacting protein mutants lead to INa deficit [261,286,375,377] while others influence Nav trafficking and thus subcellular localization [201,379].
On the other hand, mutations in desmoglein and SAP97 have been described in BrS patients [380,381], but their influence on SCN5A/Nav1.5 remains to be elucidated, while calmodulin, CamKII, and GPD1L modulation of SCN5A/Nav1.5 is well-established [157,163,382,383], but to date, no mutation in these genes have been reported in the context of Brugada syndrome.
At the transcriptional level, Nieto-Marín [351] identified TBX5 variants associated with BrS and LQT, as previously mentioned. Furthermore, additional evidence of the transcriptional modulation of SCN5A expression in Brugada syndrome, including GATA4 [384] and IRX3 [385] variants, have been reported. Transcriptional contributions by mutations in the SCN5A promoter are also linked to Brugada syndrome [384], causing decreased SCN5A/Nav1.5 and INa, thus loss-of-function. Importantly, they have also been associated with distinct arrhythmogenic diseases [24,386]. More recently, [387] reported that a common SCN5A polymorphism, i.e., H558R, modulates the SCN5A promoter methylation and thus the clinical phenotype of Brugada syndrome patients. However, it remains unclear which are the molecular mechanisms underlying such association.
The importance of non-coding RNAs in the context of Brugada syndrome has been recently explored using an integrative omics approach [388]. These authors identified several microRNAs that are distinctly upregulated in BrS, such as miR-92a-3p and miR-320b, or down-regulated such as miR-425-5p and established their plausible links as molecular determinants of Brugada syndrome, yet functional evidence is missing. More recently, Matsumura et al. [93] described SCN5A coding variants that lack genotype-phenotype concordance, and additionally, SCN5A 3’UTR variants were identified that impaired microRNA binding sites but similarly failed to properly segregate with BrS phenotype [93], suggesting that a combination of multiple genetic factors, rather than a single variant is the cause of BrS onset [93]. Thus, to date, the plausible role of ncRNAs in BrS remains almost unexplored.
Our current understanding of the mechanistic links between SCN5A/Nav1.5 post-transcriptional modifications and the onset of Brugada syndrome is scarce. Aiba et al. [389] described that R526H and S528A SCN5A mutations, identified in a BrS family, impaired Nav1.5 phosphorylation, leading to reduced peak current densities, yet steady-state activation, inactivation, and recovery for inactivation were not modified. Besides this study, no additional links for glycosylation, acetylation, and/or methylation have been reported in the context of BrS.

4.3. Atrial Fibrillation

Atrial fibrillation has also been associated with mutations in SCN5A. Ellinor et al. reported missense mutation in a 45-year-old male proband and his affected father among a series of 57 probands with a familial history of isolated or ‘lone’ atrial fibrillation [384]. More recently, Darbar et al. [390] identified eight heterozygous variants in ten probands that were not found in age-, sex-, and ethnicity-matched controls. In addition, rare nonsynonymous coding region variants previously reported were also demonstrating that in their study, nearly 6% of AF probands carried heterozygous mutations or rare SCN5A variants. Thus, the causal contribution of SCN5A to familial AF is limited.
Curiously, an extensive array of SCN5A mutations have been associated with atrial fibrillation and other electrophysiological disorders such as BrS [391], cardiac conduction defects [392], LQT [393], and even a spectrum of atrial flutter, conduction diseases, and BrS [394].
Mutations in all sodium channel ancillary subunits have been reported in atrial fibrillation patients [395,396,397,398], while only mutations in Mog1 [375,399], ankyrin [400,401], alpha-actinin [402], and caveolin [403] have been associated with atrial fibrillation among those other SCN5A/Nav1.5 interacting proteins. Curiously, no evidence of their functional impact on SCN5A/Nav1.5 function is reported to date.

4.4. Ventricular Fibrillation

Mutations in SCN5A have been associated with ventricular fibrillation [404,405,406] and idiopathic ventricular fibrillation [407,408,409]. However, a large array of SCN5A mutations linked to ventricular fibrillation are reported in the context of additional cardiac impairments, such as Brugada syndrome [410,411,412,413], atrial fibrillation [414], acute myocardial infarction [415,416,417], sudden cardiac death [418,419], or Graves’ disease [420]. Surprisingly, while abundant information is available about the genetic determinants of ventricular fibrillation, scarce information is available about the molecular mechanisms leading to ventricular fibrillation.
Genetic screening of sodium channel ancillary subunits has revealed only mutation in SCN3B associated with ventricular fibrillation [421]. Although no direct evidence of mutations in desmosomal proteins such as desmoglein or plakoglobin has been reported in ventricular fibrillation, there is compiling evidence that such mutations in the context of arrhythmogenic right ventricular cardiomyopathy (ARVD) predispose to an early onset of ventricular fibrillation [422,423,424,425]. In addition to those SCN5A/Nav1.5 interacting proteins, mutations in calmodulin [426,427] and alpha-actinin [428] have also been reported in ventricular fibrillation patients, yet their plausible implications deregulating SCN5A/Nav1.5 is not reported to date.
Within the currently described transcriptional regulators of SCN5A expression, only mutations in IRX3 have been associated with ventricular fibrillation [65]. Interestingly, those IRX3 mutations impaired SCN5A expression in in vitro experimental assays, yet the causal contribution to ventricular fibrillation remains unclear. On the other hand, no evidence has been reported so far on SCN5A/Nav1.5 post-transcriptional modifications in the context of ventricular fibrillation.

4.5. Sick Sinus Syndrome

Ample evidence has been reported on SCN5A mutations leading to sick sinus syndrome [29,429,430,431,432,433,434,435,436] alone or in combination with other electrophysiological disorders such as atrial fibrillation [437] and Brugada syndrome with conduction diseases [30]. However, no mutations in sodium channels ancillary subunits or SCN5A/Nav1.5 interacting proteins have been reported. Similarly, no link between SCN5A transcriptional and post-transcriptional regulators and sick sinus syndrome is reported to date.

4.6. Sudden Infant Death Syndrome

Mutations in SCN5A are also causative of sudden infant death syndrome [438,439,440,441,442,443,444] as well as in channel ancillary subunits [445,446,447,448]. It is assumed that, as in the case of Brugada syndrome, an SCN5A/Nav1.5 loss-of-function underlies sudden infant death. In this context, Tan et al. [446] described SCN3B and SCNB4 mutations that decreased peak INa current but increased INaL, whereas Neubauer et al. [412] and Denti et al. [447] identified a novel SCN1B mutations that decreased INa density, providing thus plausible electrophysiological mechanisms underlying sudden infant death syndrome. Importantly SCN5A mutations in SIDS have also been identified in conjunction with other electrophysiological disorders, such as Brugada syndrome [356].
In addition to those mutations, several SCN5A/Nav1.5 interacting proteins have also been associated with sudden infant cardiac syndromes, such as calmodulin [383,449], syntrophin [186,346,450], and caveolin [187,451]. Cheng et al. reported that syntrophin mutation leads to increased peak INa current and overlap between activation and inactivation curves increasing thus the window current [186]. Similar findings on the increased peak INa current were also reported by Wu et al. [346] and Cheng et al. [450], a phenotype that was rescued by another calmodulin variant identified in SIDS. Caveolin mutations seem to distinctly contribute to sodium regulation, either by suppressing the INaL by inhibiting nNOS-dependent S-nitrosylation of SCN5A [187] or by the persistence of the INaL [451].
On the other hand, no association between SCN5A transcriptional and post-transcriptional modulators and sudden infant death syndrome has been reported to date.

4.7. Other Channelopathies

SCN5A mutations also play a pivotal role in complex electrophysiological disorders that present a combination of alterations, particularly on conduction disorders associated with sick sinus syndrome, atrial fibrillation and ventricular tachycardia [435]; LQT and BrS [452]; atrial fibrillation, BrS, and sudden cardiac death [394]; BrS [453]; BrS and AF [454]; or atrial arrhythmias [392]. Similarly, increasing evidence is also reported in cases of sudden cardiac death, alone [455] or in combination with LQT and dilated cardiomyopathy [452], ventricular fibrillation and BrS [456], or atrial fibrillation [433]. Our current understanding of the molecular mechanisms providing such diversity of phenotypic manifestations is still limited. It is important nonetheless to highlight that mutations in SCN5A/Nav1.5 interacting proteins [9,201,401] and/or transcriptional regulators [457] have also been recently reported to be associated with these complex electrophysiological disorders.
Table 1. Summary of the Nav1.5-interacting protein-coding genes involved in cardiac sodium channelopathies.
Table 1. Summary of the Nav1.5-interacting protein-coding genes involved in cardiac sodium channelopathies.
Altered Gene/AlterationEffect on INa CurrentMode of ActionReferences
Long QT syndrome
SCN5AIncreased INaL,
increased window current
NA[14,336,338,339,340,341]
SCN1BIncreased INaLAncillary subunit[342]
SCN4BUndeterminedAncillary subunit[231,343]
SNTA1Increased INaLActivation of nNOS–SCN5A macromolecular complex[186,187,344,345,346,347]
Cav3Increased INaLnNOS-dependent S-nitrosylation of SCN5A[187]
BAG3UndeterminedUndetermined[348]
ANK BUndeterminedUndetermined[258]
ACTN1UndeterminedUndetermined[349]
TBX5Increased INaLTransactivation impairment[350,351]
Brugada syndrome
SCN5ADecreased INaNA[24,25,26,361,362,363,364,365]
SCN1BDecreased INaAncillary subunit; cellular trafficking defects[232,356,359,366,367,368,369,370]
SCN2BDecreased INaAncillary subunit; cellular trafficking defects[229]
SCN3BDecreased INaAncillary subunit; cellular trafficking defects[356]
PKP2Decreased INaINa deficit[260,371]
MOG1Decreased INaINa deficit; subcellular trafficking[9,201,372,373,374,375]
FGF13Decreased INaEnhanced Nav1.5 inactivation[376]
SNTA1Decreased INaDefective Nav1.5 protein interaction[377]
TMEM168Decreased INaReduced Nav1-5 expression[378,379]
TCAPDecreased INaDefective Nav1.5 protein interaction[285,286]
SAP97UndeterminedUndetermined[380,381]
TBX5UndeterminedTransactivation impairment[351]
GATA4UndeterminedUndetermined[384]
IRX3UndeterminedUndetermined[385]
SCN5A promoter methylationUndeterminedDecreased SCN5A promoter methylation[387]
SCN5A promoter mutationsUndeterminedTransactivation impairment[384]
Nav1.5 phosphorylationReduced peak INa densityImpaired PKA stimulation[389]
Atrial fibrillation
SCN5AUndeterminedNA[384,390]
SCN1BReduced INaAncillary subunit; altered channel gating[395]
SCN2BReduced INaAncillary subunit; altered channel gating[395]
SCN3BUndeterminedAncillary subunit[398]
SCN4BUndeterminedAncillary subunit[396,397]
MOG1Reduced INaUndetermined[375,399]
ANKUndeterminedUndetermined[400,401]
ACTN1UndeterminedUndetermined[402]
Cav3UndeterminedUndetermined[403]
Ventricular fibrillation
SCN5AReduced INaNA[404,405,406]
SCN3BReduced peak INaAncillary subunit; impaired trafficking[421]
CaMUndeterminedUndetermined[426,427]
ACTN1UndeterminedUndetermined[428]
IRX3UndeterminedImpaired transactivation[65]
Sick sinus syndrome
SCN5AReduced INaNA[29,429,430,431,432,433,434,435,436]
Sudden infant death syndrome
SCN5AIncreased INaLImpaired Nav1.5 inactivation[438,439,440,441,442,443,444]
SCN1BDecreased INa densityAncillary subunit[447]
SCN3BDecreased INa,
Increased INaL
Ancillary subunit[446,448]
SCN4BDecreased INa,
Increased INaL
Ancillary subunit[446]
CaMUndeterminedUndetermined[383,449]
SNTA1Increased peak INa and window currentUndetermined[186,346,450]
Cav3Suppression of INaLInhibiting nNOS-dependent S-nitrosylation of Nav1.5[187,451]

5. Conclusions and Perspectives

Understanding the complex cellular and molecular mechanisms by which the cardiac sodium channel is formed is essential for dissecting and eventually repairing the culprits of an important array of cardiac arrhythmogenic defects with a large impact on society, such as sudden death. In this review, we have provided current state-of-the-art information of the molecular events that emanate with the transcription and transcriptional regulation of the SCN5A gene, the post-transcriptional modifications that the SCN5A transcript undergoes leading to the configuration of a large array of distinct alternative spliced variants and being subjected to several types of epigenetic modulations along the way. Subsequently, distinct interacting proteins accompany the nascent Nav1.5 protein along with different subcellular organelles until it reaches its final membrane destination, accumulating a large number of diverse post-translational modifications, which eventually allows to carry out a fundamental electrophysiological role on the configuration of the cardiac action potential. Importantly, defects on a large number of these pathways have a tremendous impact on Nav1.5 functionality and are thus intimately linked to cardiac arrhythmias. While the advancement of the knowledge of the cardiac sodium channel biology and electrophysiology has greatly increased over the last two decades, understanding of the functional contribution of multiple SCN5A mutations as well as mutations in SCN5A/Nav1.5 interacting proteins is still incipient. In this context, it is particularly challenging to deeply understand the molecular mechanisms driving the association of single mutations with distinct impaired electrophysiological entities, such as Brugada and LQT syndrome, or in cases of even more complex phenotypical manifestations, such as sick sinus syndrome, atrial fibrillation, and ventricular tachycardia. In the coming years, insights into distinct molecular mechanisms involved in SCN5A/Nav1.5 formation and maturation will be more deeply explored, providing us with an increased understanding of the molecular and cellular routes that, if impaired, lead to distinct cardiac electrophysiological pathophysiologies, particularly on our currently poorly explored such as ventricular fibrillation, sick sinus syndrome, and sudden infant cardiac death. We also envision that the impact of non-coding biology into the modulation of SCN5A/Nav1.5 will progressively emerge, providing novel therapeutic tools to repair or at least modulate the pathological consequences of severe arrhythmogenic defects such as long QT, Brugada syndrome, and/or atrial fibrillation.

Author Contributions

All authors contributed equally to the writing, review and editing of this work. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

APAction potential
VSVoltage-sensing module
PMPore module
PUFAPolyunsaturated fatty acids
EREndoplasmic reticulum
RERegulatory elements
UPR Unfolded protein response
ERADER-associated degradation pathway
PTMsPost-translational modifications
IDIntercalated discs
LMLateral membrane
IksSlow component of the delayed rectifier potassium current
ItoTransient outward potassium current

References

  1. Catterall, W.A.; Maier, S. Voltage-Gated Sodium Channels and Electrical Excitability of the Heart. In Cardiac Electrophysiology: From Cell to Bedside, 7th ed.; WB Saunders: Philadelphia, PA, USA, 2015; pp. 1–11. [Google Scholar] [CrossRef]
  2. Fozzard, H.A.; Hanck, D.A. Structure and function of voltage-dependent sodium channels: Comparison of brain II and cardiac isoforms. Physiol. Rev. 1996, 76, 887–926. [Google Scholar] [CrossRef]
  3. Gavillet, B.; Rougier, J.S.; Domenighetti, A.A.; Behar, R.; Boixel, C.; Ruchat, P.; Lehr, H.A.; Pedrazzini, T.; Abriel, H. Cardiac sodium channel Nav1.5 is regulated by a multiprotein complex composed of syntrophins and dystrophin. Circ. Res. 2006, 99, 407–414. [Google Scholar] [CrossRef] [Green Version]
  4. Gellens, M.E.; George, A.L.; 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. Natl. Acad. Sci. USA 1992, 89, 554–558. [Google Scholar] [CrossRef] [Green Version]
  5. Jiang, D.; Shi, H.; Tonggu, L.; Gamal El-Din, T.M.; Lenaeus, M.J.; Zhao, Y.; Yoshioka, C.; Zheng, N.; Catterall, W.A. Structure of the Cardiac Sodium Channel. Cell 2020, 180, 122–134.e10. [Google Scholar] [CrossRef]
  6. Bohannon, B.M.; de la Cruz, A.; Wu, X.; Jowais, J.J.; Perez, M.E.; Dykxhoorn, D.M.; Liin, S.I.; Larsson, H.P. Erratum: Correction: Polyunsaturated fatty acid analogues differentially affect cardiac NaV, CaV, and KV channels through unique mechanisms. eLife 2020, 9, e60141. [Google Scholar] [CrossRef]
  7. Kang, J.X.; Leaf, A. Evidence that free polyunsaturated fatty acids modify Na+ channels by directly binding to the channel proteins. Proc. Natl. Acad. Sci. USA 1996, 93, 3542–3546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Guan, Y.; Gao, X.; Tang, Q.; Huang, L.; Gao, S.; Yu, S.; Huang, J.; Li, J.; Zhou, D.; Zhang, Y.; et al. Nucleoporin 107 facilitates the nuclear export of Scn5a mRNA to regulate cardiac bioelectricity. J. Cell. Mol. Med. 2019, 23, 1448–1457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Yu, G.; Liu, Y.; Qin, J.; Wang, Z.; Hu, Y.; Wang, F.; Li, Y.; Chakrabarti, S.; Chen, Q.; Wang, Q.K. Mechanistic insights into the interaction of the MOG1 protein with the cardiac sodium channel Nav1.5 clarify the molecular basis of Brugada syndrome. J. Biol. Chem. 2018, 293, 18207–18217. [Google Scholar] [CrossRef] [Green Version]
  10. Kurokawa, K.; Nakano, A. The ER exit sites are specialized ER zones for the transport of cargo proteins from the ER to the Golgi apparatus. J. Biochem. 2019, 165, 109–114. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  11. Dong, C.; Wang, Y.; Ma, A.; Wang, T. Life Cycle of the Cardiac Voltage-Gated Sodium Channel Nav1.5. Front. Physiol. 2020, 11, 609733. [Google Scholar] [CrossRef]
  12. Clark, K.A.; McElhinny, A.S.; Beckerle, M.C.; Gregorio, C.C. Striated muscle cytoarchitecture: An intricate web of form and function. Annu. Rev. Cell Dev. Biol. 2002, 18, 637–706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Balse, E.; Eichel, C. The Cardiac Sodium Channel and Its Protein Partners. In Handbook of Experimental Pharmacology; Springer: Cham, Switzerland, 2017; pp. 251–263. [Google Scholar] [CrossRef]
  14. Song, W.; Shou, W. Cardiac sodium channel Nav1.5 mutations and cardiac arrhythmia. Pediatr. Cardiol. 2012, 33, 943–949. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Balser, J.R. The cardiac sodium channel: Gating function and molecular pharmacology. J. Mol. Cell. Cardiol. 2001, 33, 599–613. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chahine, M.; George, A.L.; Zhou, M.; Ji, S.; Sun, W.; Barchi, R.L.; Horn, R. Sodium channel mutations in paramyotonia congenita uncouple inactivation from activation. Neuron 1994, 12, 281–294. [Google Scholar] [CrossRef]
  17. Yang, N.; Horn, R. Evidence for voltage-dependent S4 movement in sodium channels. Neuron 1995, 15, 213–218. [Google Scholar] [CrossRef] [Green Version]
  18. Albert, C.; Ruben, P.C.; George, A.L.; Fujimoto, E.; Bezanilla, F. Voltage sensors in domains III and IV, but not I and II, are immobilized by Na+ channel fast inactivation. Neuron 1999, 22, 73–87. [Google Scholar] [CrossRef] [Green Version]
  19. Armstrong, C.M.; Bezanilla, F. Inactivation of the sodium channel: II. gating current experiments. J. Gen. Physiol. 1977, 70, 567–590. [Google Scholar] [CrossRef]
  20. Armstrong, C.M.; Wm Gilly, F. Fast and slow steps in the activation of sodium channels. J. Gen. Physiol. 1979, 74, 691–711. [Google Scholar] [CrossRef] [Green Version]
  21. Hodgkin, A.L.; Huxley, A.F. Currents carried by sodium and potassium ions through the membrane of the giant axon of Loligo. J. Physiol. 1952, 116, 449–472. [Google Scholar] [CrossRef]
  22. Kapplinger, J.D.; Tester, D.J.; Alders, M.; Benito, B.; Berthet, M.; Brugada, J.; Brugada, P.; Fressart, V.; Guerchicoff, A.; Harris-Kerr, C.; et al. An international compendium of mutations in the SCN5A-encoded cardiac sodium channel in patients referred for Brugada syndrome genetic testing. Heart Rhythm 2010, 7, 33–46. [Google Scholar] [CrossRef] [Green Version]
  23. Kroncke, B.M.; Glazer, A.M.; Smith, D.K.; Blume, J.D.; Roden, D.M. SCN5A (Nav1.5) Variant Functional Perturbation and Clinical Presentation: Variants of a Certain Significance. Circ. Genom. Precis. Med. 2018, 11, e002095. [Google Scholar] [CrossRef] [Green Version]
  24. Remme, C.A.; Verkerk, A.O.; Nuyens, D.; van Ginneken, A.C.G.; van Brunschot, S.; Belterman, C.N.W.; Wilders, R.; van Roon, M.A.; Tan, H.L.; Wilde, A.A.M.; et al. Overlap Syndrome of Cardiac Sodium Channel Disease in Mice Carrying the Equivalent Mutation of Human SCN5A -1795insD. Circulation 2006, 114, 2584–2594. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Chen, Q.; Kirsch, G.E.; Zhang, D.; Brugada, R.; Brugada, J.; Brugada, P.; Potenza, D.; Moya, A.; Borggrefe, M.; Breithardt, G.; et al. Genetic basis and molecular mechanism for idiopathic ventricular fibrillation. Nature 1998, 392, 293–296. [Google Scholar] [CrossRef] [PubMed]
  26. 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.; et al. A Single Na+ Channel Mutation Causing Both Long-QT and Brugada Syndromes. Circ. Res. 1999, 85, 1206–1213. [Google Scholar] [CrossRef] [Green Version]
  27. Schott, 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. Nat. Genet. 1999, 23, 20–21. [Google Scholar] [CrossRef]
  28. 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 2001, 409, 1043–1047. [Google Scholar] [CrossRef]
  29. Benson, D.W.; Wang, D.W.; Dyment, M.; Knilans, T.K.; Fish, F.A.; Strieper, M.J.; Rhodes, T.H.; George, A.L. Congenital sick sinus syndrome caused by recessive mutations in the cardiac sodium channel gene (SCN5A). J. Clin. Investig. 2003, 112, 1019–1028. [Google Scholar] [CrossRef] [Green Version]
  30. Smits, J.P.P.; Koopmann, T.T.; Wilders, R.; Veldkamp, M.W.; Opthof, T.; Bhuiyan, Z.A.; Mannens, M.M.A.M.; Balser, J.R.; Tan, H.L.; Bezzina, C.R.; et al. A mutation in the human cardiac sodium channel (E161K) contributes to sick sinus syndrome, conduction disease and Brugada syndrome in two families. J. Mol. Cell. Cardiol. 2005, 38, 969–981. [Google Scholar] [CrossRef]
  31. 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 1995, 80, 805–811. [Google Scholar] [CrossRef] [Green Version]
  32. Mann, S.A.; Castro, M.L.; Ohanian, M.; Guo, G.; Zodgekar, P.; Sheu, A.; Stockhammer, K.; Thompson, T.; Playford, D.; Subbiah, R.; et al. R222Q SCN5A mutation is associated with reversible ventricular ectopy and dilated cardiomyopathy. J. Am. Coll. Cardiol. 2012, 60, 1566–1573. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Swan, H.; Amarouch, M.Y.; Leinonen, J.; Marjamaa, A.; Kucera, J.P.; Laitinen-Forsblom, P.J.; Lahtinen, A.M.; Palotie, A.; Kontula, K.; Toivonen, L.; et al. Gain-of-function mutation of the SCN5A gene causes exercise-induced polymorphic ventricular arrhythmias. Circ. Cardiovasc. Genet. 2014, 7, 771–781. [Google Scholar] [CrossRef] [Green Version]
  34. Laurent, G.; Saal, S.; Amarouch, M.Y.; Béziau, D.M.; Marsman, R.F.J.; Faivre, L.; Barc, J.; Dina, C.; Bertaux, G.; Barthez, O.; et al. Multifocal ectopic Purkinje-related premature contractions: A new SCN5A-related cardiac channelopathy. J. Am. Coll. Cardiol. 2012, 60, 144–156. [Google Scholar] [CrossRef]
  35. Makiyama, T.; Akao, M.; Shizuta, S.; Doi, T.; Nishiyama, K.; Oka, Y.; Ohno, S.; Nishio, Y.; Tsuji, K.; Itoh, H.; et al. A novel SCN5A gain-of-function mutation M1875T associated with familial atrial fibrillation. J. Am. Coll. Cardiol. 2008, 52, 1326–1334. [Google Scholar] [CrossRef] [Green Version]
  36. Olson, T.M.; Michels, V.V.; Ballew, J.D.; Reyna, S.P.; Karst, M.L.; Herron, K.J.; Horton, S.C.; Rodeheffer, R.J.; Anderson, J.L. Sodium channel mutations and susceptibility to heart failure and atrial fibrillation. JAMA 2005, 293, 447–454. [Google Scholar] [CrossRef]
  37. McNair, W.P.; Ku, L.; Taylor, M.R.G.; Fain, P.R.; Dao, D.; Wolfel, E.; Mestroni, L. SCN5A mutation associated with dilated cardiomyopathy, conduction disorder, and arrhythmia. Circulation 2004, 110, 2163–2167. [Google Scholar] [CrossRef] [Green Version]
  38. Bezzina, C.R.; Rook, M.B.; Groenewegen, W.A.; Herfst, L.J.; Van der Wal, A.C.; Lam, J.; Jongsma, H.J.; Wilde, A.A.M.; Mannens, M.M.A.M. Compound heterozygosity for mutations (W156X and R225W) in SCN5A associated with severe cardiac conduction disturbances and degenerative changes in the conduction system. Circ. Res. 2003, 92, 159–168. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Clatot, J.; Zheng, Y.; Girardeau, A.; Liu, H.; Laurita, K.R.; Marionneau, C.; Deschênes, I. Mutant voltage-gated Na+ channels can exert a dominant negative effect through coupled gating. Am. J. Physiol.—Heart Circ. Physiol. 2018, 315, H1250–H1257. [Google Scholar] [CrossRef] [PubMed]
  40. Keller, D.I.; Kucera, J.P.; Benammar, N. Brugada syndrome and fever: Genetic and molecular characterization of patients carrying SCN5A mutations. Cardiovasc. Res. 2005, 67, 510–519. [Google Scholar] [CrossRef] [Green Version]
  41. Doisne, N.; Grauso, M.; Mougenot, N.; Clergue, M.; Souil, C.; Coulombe, A.; Guicheney, P.; Neyroud, N. In vivo Dominant-Negative Effect of an SCN5A Brugada Syndrome Variant. Front. Physiol. 2021, 12, 1–13. [Google Scholar] [CrossRef] [PubMed]
  42. Neill, M.J.O.; 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. bioRxiv 2021. [Google Scholar] [CrossRef]
  43. Galleano, I.; Harms, H.; Choudhury, K.; Khoo, K.; Delemotte, L.; Pless, S.A. Functional cross-talk between phosphorylation and disease-causing mutations in the cardiac sodium channel Nav1.5. Proc. Natl. Acad. Sci. USA 2021, 118, e2025320118. [Google Scholar] [CrossRef]
  44. Van der Harst, P.; van Setten, J.; Verweij, N.; Vogler, G.; Franke, L.; Maurano, M.T.; Wang, X.; Mateo Leach, I.; Eijgelsheim, M.; Sotoodehnia, N.; et al. 52 Genetic Loci Influencing Myocardial Mass A. J. Am. Coll. Cardiol. 2016, 68, 1435–1448. [Google Scholar] [CrossRef]
  45. Van Den Boogaard, M.; Smemo, S.; Burnicka-Turek, O.; Arnolds, D.E.; Van De Werken, H.J.G.; Klous, P.; McKean, D.; Muehlschlegel, J.D.; Moosmann, J.; Toka, O.; et al. A common genetic variant within SCN10A modulates cardiac SCN5A expression. J. Clin. Investig. 2014, 124, 1844–1852. [Google Scholar] [CrossRef] [Green Version]
  46. Van Den Boogaard, M.; Barnett, P.; Vincent, M.; Van Den Boogaard, M.; Wong, L.Y.E.; Tessadori, F.; Bakker, M.L. Genetic variation in T-box binding element functionally affects SCN5A/SCN10A enhancer Find the latest version: Genetic variation in T-box binding element functionally affects SCN5A/SCN10A enhancer. J. Clin. Investig. 2012, 122, 2519–2530. [Google Scholar] [CrossRef] [Green Version]
  47. Verweij, N.; Leach, I.M.; Van Den Boogaard, M.; Van Veldhuisen, D.J.; Christoffels, V.M.; Hillege, H.L.; Van Gilst, W.H.; Barnett, P.; De Boer, R.A.; Van Der Harst, P. Genetic Determinants of P Wave Duration and PR Segment. Circ. Cardiovasc. Genet. 2014, 7, 475–481. [Google Scholar] [CrossRef] [Green Version]
  48. van Setten, J.; Brody, J.A.; Jamshidi, Y.; Swenson, B.R.; Butler, A.M.; Campbell, H.; Del Greco, F.M.; Evans, D.S.; Gibson, Q.; Gudbjartsson, D.F.; et al. PR interval genome-wide association meta-analysis identifies 50 loci associated with atrial and atrioventricular electrical activity. Nat. Commun. 2018, 9, 1–11. [Google Scholar] [CrossRef] [Green Version]
  49. Man, J.C.; Mohan, R.A.; van den Boogaard, M.; Hilvering, C.R.; Jenkins, C.; Wakker, V.; Bianchi, V.; de Laat, W.; Barnett, P.; Boukens, B.J.; et al. An enhancer cluster controls gene activity and topology of the SCN5A-SCN10A locus in vivo. Nat. Commun. 2019, 10, 4943. [Google Scholar] [CrossRef] [Green Version]
  50. Lister, R.; Pelizzola, M.; Dowen, R.H.; Hawkins, R.D.; Hon, G.; Tonti-Filippini, J.; Nery, J.R.; Lee, L.; Ye, Z.; Ngo, Q.M.; et al. Human DNA methylomes at base resolution show widespread epigenomic differences. Nature 2009, 462, 315–322. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  51. Postma, A.V.; Bezzina, C.R.; Christoffels, V.M. Genetics of congenital heart disease: The contribution of the noncoding regulatory genome. J. Hum. Genet. 2016, 61, 13–19. [Google Scholar] [CrossRef] [PubMed]
  52. Bezzina, C.R.; Barc, J.; Mizusawa, Y.; Remme, C.A.; Gourraud, J.B.; Simonet, F.; Verkerk, A.O.; Schwartz, P.J.; Crotti, L.; Dagradi, F.; et al. Common variants at SCN5A-SCN10A and HEY2 are associated with Brugada syndrome, a rare disease with high risk of sudden cardiac death. Nat Genet. 2013, 45, 1044–1049. [Google Scholar] [CrossRef] [PubMed]
  53. Gilsbach, R.; Preissl, S.; Grüning, B.A.; Schnick, T.; Burger, L.; Benes, V.; Würch, A.; Bönisch, U.; Günther, S.; Backofen, R.; et al. Dynamic DNA methylation orchestrates cardiomyocyte development, maturation and disease. Nat. Commun. 2014, 5, 5288. [Google Scholar] [CrossRef] [Green Version]
  54. Carreras, D.; Martinez-Moreno, R.; Pinsach-Abuin, M.L.; Santafe, M.M.; Gomà, P.; Brugada, R.; Scornik, F.S.; Pérez, G.J.; Pagans, S. Epigenetic changes governing Scn5a expression in denervated skeletal muscle. Int. J. Mol. Sci. 2021, 22, 2755. [Google Scholar] [CrossRef]
  55. Salvarani, N.; Crasto, S.; Miragoli, M.; Bertero, A.; Paulis, M.; Kunderfranco, P.; Serio, S.; Forni, A.; Lucarelli, C.; Dal Ferro, M.; et al. The K219T-Lamin mutation induces conduction defects through epigenetic inhibition of SCN5A in human cardiac laminopathy. Nat. Commun. 2019, 10, 1–16. [Google Scholar] [CrossRef] [Green Version]
  56. Arnolds, D.E.; Liu, F.; Fahrenbach, J.P.; Kim, G.H.; Schillinger, K.J.; Smemo, S.; McNally, E.M.; Nobrega, M.A.; Patel, V.V.; Moskowitz, I.P. TBX5 drives Scn5a expression to regulate cardiac conduction system function. J. Clin. Investig. 2012, 122, 2509–2518. [Google Scholar] [CrossRef] [Green Version]
  57. Moskowitz, I.P.G.; Pizard, A.; Patel, V.V.; Bruneau, B.G.; Kim, J.B.; Kupershmidt, S.; Roden, D.; Berul, C.I.; Seidman, C.E.; Seidman, J.G. The T-Box transcription factor Tbx5 is required for the patterning and maturation of the murine cardiac conduction system. Development 2004, 131, 4107–4116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Steimle, J.D.; Moskowitz, I.P. TBX5: A Key Regulator of Heart Development. Curr. Top. Dev. Biol. 2017, 122, 195–221. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  59. Brewer, A.; Pizzey, J. GATA factors in vertebrate heart development and disease. Expert Rev. Mol. Med. 2006, 8, 1–20. [Google Scholar] [CrossRef]
  60. Garg, V.; Kathiriya, I.S.; Barnes, R.; Schluterman, M.K.; King, I.N.; Butler, C.A.; Rothrock, C.R.; Eapen, R.S.; Hirayama-Yamada, K.; Joo, K.; et al. GATA4 mutations cause human congenital heart defects and reveal an interaction with TBX5. Nature 2003, 424, 443–447. [Google Scholar] [CrossRef]
  61. Tarradas, A.; Pinsach-Abuin, M.; Mackintosh, C.; Llora-Batlle, O.; Perez-Serra, A.; Batlle, M.; Pérez-Villa, F.; Zimmer, T.; Garcia-Bassets, I.; Brugada, R.; et al. Transcriptional regulation of the sodium channel gene (SCN5A) by GATA4 in human heart. J. Mol. Cell. Cardiol. 2017, 102, 74–82. [Google Scholar] [CrossRef] [Green Version]
  62. Zhou, A.; Shi, G.; Kang, G.J.; Xie, A.; Liu, H.; Jiang, N.; Liu, M.; Jeong, E.M.; Dudley, S.C. RNA binding protein, HuR, regulates SCN5A expression through stabilizing MEF2C transcription factor mRNA. J. Am. Heart Assoc. 2018, 7, 1–12. [Google Scholar] [CrossRef] [PubMed]
  63. Zhou, A.; Xie, A.; Kim, T.; Liu, H.; Shi, G.; Kang, G.; Jiang, N.; Liu, M.; Jeong, E.; Choi, B.; et al. HuR-mediated SCN5A mRNA stability reduces arrhythmic risk in heart failure. Heart Rhythm 2018, 15, 1072–1080. [Google Scholar] [CrossRef]
  64. Gaborit, N.; Sakuma, R.; Wylie, J.N.; Kim, K.H.; Zhang, S.S.; Hui, C.C.; Bruneau, B.G. Cooperative and antagonistic roles for Irx3 and Irx5 in cardiac morphogenesis and postnatal physiology. Development 2012, 139, 4007–4019. [Google Scholar] [CrossRef] [Green Version]
  65. Koizumi, A.; Sasano, T.; Kimura, W.; Miyamoto, Y.; Aiba, T.; Ishikawa, T.; Nogami, A.; Fukamizu, S.; Sakurada, H.; Takahashi, Y.; et al. Genetic defects in a His-Purkinje system transcription factor, IRX3, cause lethal cardiac arrhythmias. Eur. Heart J. 2016, 37, 1469–1475. [Google Scholar] [CrossRef] [Green Version]
  66. Cai, B.; Wang, N.; Mao, W.; You, T.; Lu, Y.; Li, X.; Ye, B.; Li, F.; Xu, H. Deletion of FoxO1 leads to shortening of QRS by increasing Na+ channel activity through enhanced expression of both cardiac Nav1.5 and β3 subunit. J. Mol. Cell. Cardiol. 2014, 74, 297–306. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Mao, W.; You, T.; Ye, B.; Li, X.; Dong, H.H.; Hill, J.A.; Li, F.; Xu, H. Reactive oxygen species suppress cardiac Nav1.5 expression through Foxo1. PLoS ONE 2012, 7, e32738. [Google Scholar] [CrossRef] [Green Version]
  68. Atack, T.C.; Stroud, D.M.; Watanabe, H.; Yang, T.; Hall, L.; Susan, B.; Lowe, J.S.; Leake, B.; Magnuson, M.A.; Yang, P.; et al. Informatic and Functional Approaches to Identifying a Regulatory Region for the Cardiac Sodium Channel. Circ. Res. 2012, 109, 38–46. [Google Scholar] [CrossRef] [Green Version]
  69. Hesse, M.; Kondo, C.S.; Clark, R.B.; Su, L.; Allen, F.L.; Geary-Joo, C.T.M.; Kunnathu, S.; Severson, D.L.; Nygren, A.; Giles, W.R.; et al. Dilated cardiomyopathy is associated with reduced expression of the cardiac sodium channel Scn5a. Cardiovasc. Res. 2007, 75, 498–509. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  70. Nieto, M.A. The snail superfamily of zinc-finger transcription factors. Nat. Rev. Mol. Cell Biol. 2002, 3, 155–166. [Google Scholar] [CrossRef] [PubMed]
  71. Schroeter, A.; Walzik, S.; Blechschmidt, S.; Haufe, V.; Benndorf, K.; Zimmer, T. Structure and function of splice variants of the cardiac voltage-gated sodium channel Nav1.5. J. Mol. Cell. Cardiol. 2010, 49, 16–24. [Google Scholar] [CrossRef] [PubMed]
  72. Makielski, J.C.; Ye, B.; Valdivia, C.R.; Pagel, M.D.; Pu, J.; Tester, D.J.; Ackerman, M.J. A Ubiquitous Splice Variant and a Common Polymorphism Affect Heterologous Expression of Recombinant Human SCN5A Heart Sodium Channels. Circ. Res. 2003, 93, 821–828. [Google Scholar] [CrossRef] [Green Version]
  73. Camacho, J.A.; Hensellek, S.; Rougier, J.S.; Blechschmidt, S.; Abriel, H.; Benndorf, K.; Zimmer, T. Modulation of Nav1.5 channel function by an alternatively spliced sequence in the DII/DIII linker region. J. Biol. Chem. 2006, 281, 9498–9506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Blechschmidt, S.; Haufe, V.; Benndorf, K.; Zimmer, T. Voltage-gated Na+ channel transcript patterns in the mammalian heart are species-dependent. Prog. Biophys. Mol. Biol. 2008, 98, 309–318. [Google Scholar] [CrossRef]
  75. Chioni, A.M.; Fraser, S.P.; Pani, F.; Foran, P.; Wilkin, G.P.; Diss, J.K.J.; Djamgoz, M.B.A. A novel polyclonal antibody specific for the Nav1.5 voltage-gated Na+ channel “neonatal” splice form. J. Neurosci. Methods 2005, 147, 88–98. [Google Scholar] [CrossRef]
  76. Onkal, R.; Mattis, J.H.; Fraser, S.P.; Diss, J.K.J.; Shao, D.; Okuse, K.; Djamgoz, M.B.A. Alternative splicing of Nav1.5: An electrophysiological comparison of “neonatal” and “adult” isoforms and critical involvement of a lysine residue. J. Cell. Physiol. 2008, 216, 716–726. [Google Scholar] [CrossRef]
  77. Zimmer, T.; Bollensdorff, C.; Haufe, V.; Birch-Hirschfeld, E.; Benndorf, K. Mouse heart Na+ channels: Primary structure and function of two isoforms and alternatively spliced variants. Am. J. Physiol.—Heart Circ. Physiol. 2002, 282, 1007–1017. [Google Scholar] [CrossRef] [Green Version]
  78. Wang, J.; Ou, S.W.; Wang, Y.J.; Zong, Z.H.; Lin, L.; Kameyama, M.; Kameyama, A. New variants of Nav1.5/SCN5A encode Na+ channels in the brain. J. Neurogenet. 2008, 22, 57–75. [Google Scholar] [CrossRef] [PubMed]
  79. Wang, J.; Ou, S.W.; Wang, Y.J.; Kameyama, M.; Kameyama, A.; Zong, Z.H. Analysis of four novel variants of Nav1.5/SCN5A cloned from the brain. Neurosci. Res. 2009, 64, 339–347. [Google Scholar] [CrossRef] [PubMed]
  80. Shang, L.L.; Pfahnl, A.E.; Sanyal, S.; Jiao, Z.; Allen, J.; Banach, K.; Fahrenbach, J.; Weiss, D.; Taylor, W.R.; Zafari, A.M.; et al. Human Heart Failure Is Associated With Abnormal C-Terminal Splicing Variants in the Cardiac Sodium Channel. Physiology 2007, 101, 1146–1154. [Google Scholar] [CrossRef] [Green Version]
  81. Montañés-Agudo, P.; Casini, S.; Aufiero, S.; Ernault, A.C.; van der Made, I.; Pinto, Y.M.; Remme, C.A.; Creemers, E.E. Inhibition of minor intron splicing reduces Na+ and Ca2+ channel expression and function in cardiomyocytes. J. Cell Sci. 2021, 135, 259191. [Google Scholar] [CrossRef]
  82. Wang, Z.; Luo, X.; Lu, Y.; Yang, B. miRNAs at the heart of the matter. J. Mol. Med. 2008, 86, 771–783. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Yang, B.; Lu, Y.; Wang, Z. Control of cardiac excitability by microRNAs. Cardiovasc. Res. 2008, 79, 571–580. [Google Scholar] [CrossRef] [Green Version]
  84. Remme, C.A.; Bezzina, C.R. Sodium channel (Dys)function and cardiac arrhythmias. Cardiovasc. Ther. 2010, 28, 287–294. [Google Scholar] [CrossRef] [PubMed]
  85. Boštjančič, E.; Zidar, N.; Štajer, D.; Glavač, D. MicroRNAs miR-1, miR-133a, miR-133b and miR-208 are dysregulated in human myocardial infarction. Cardiology 2010, 115, 163–169. [Google Scholar] [CrossRef]
  86. Daimi, H.; Lozano-Velasco, E.; Haj Khelil, A.; Chibani, J.B.E.; Barana, A.; Amorós, I.; González De La Fuente, M.; Caballero, R.; Aranega, A.; Franco, D. Regulation of SCN5A by microRNAs: MiR-219 modulates SCN5A transcript expression and the effects of flecainide intoxication in mice. Heart Rhythm 2015, 12, 1333–1342. [Google Scholar] [CrossRef] [PubMed]
  87. Lozano-Velasco, E.; Hernández-Torres, F.; Daimi, H.; Serra, S.A.S.A.; Herraiz, A.; Hove-Madsen, L.; Aránega, A.; Franco, D. Pitx2 impairs calcium handling in a dose-dependent manner by modulating Wnt signalling. Cardiovasc. Res. 2016, 109, 55–66. [Google Scholar] [CrossRef] [Green Version]
  88. Lozano-Velasco, E.; Wangensteen, R.; Quesada, A.; Garcia-Padilla, C.; Osorio, J.A.; Ruiz-Torres, M.D.; Aranega, A.; Franco, D. Hyperthyroidism, but not hypertension, impairs PITX2 expression leading to Wnt-microRNA-ion channel remodeling. PLoS ONE 2017, 12, e0188473. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  89. Chinchilla, A.; Daimi, H.; Lozano-velasco, E.; Dominguez, J.N.; Caballero, R.; Delpo, E.; Tamargo, J.; Cinca, J.; Hove-madsen, L.; Aranega, A.E.; et al. PITX2 Insufficiency Leads to Atrial Electrical and Structural Remodeling Linked to Arrhythmogenesis. Circ. Cardiovasc. Genet. 2011, 4, 269–279. [Google Scholar] [CrossRef] [Green Version]
  90. Zhao, Y.; Huang, Y.; Li, W.; Wang, Z.; Zhan, S.; Zhou, M.; Yao, Y.; Zeng, Z.; Hou, Y.; Chen, Q.; et al. Post-transcriptional regulation of cardiac sodium channel gene SCN5A expression and function by miR-192-5p. Biochim. Biophys. Acta—Mol. Basis Dis. 2015, 1852, 2024–2034. [Google Scholar] [CrossRef] [Green Version]
  91. Li, J.; Xu, C.; Liu, Y.; Li, Y.; Du, S.; Zhang, R.; Sun, Y.; Zhang, R.; Wang, Y.; Xue, H.; et al. Fibroblast growth factor 21 inhibited ischemic arrhythmias via targeting miR-143/EGR1 axis. Basic Res. Cardiol. 2020, 115, 1–17. [Google Scholar] [CrossRef]
  92. Zhang, X.; Yoon, J.Y.; Morley, M.; McLendon, J.M.; Mapuskar, K.A.; Gutmann, R.; Mehdi, H.; Bloom, H.L.; Dudley, S.C.; Ellinor, P.T.; et al. A common variant alters SCN5A-miR-24 interaction and associates with heart failure mortality. J. Clin. Investig. 2018, 128, 1154–1163. [Google Scholar] [CrossRef] [Green Version]
  93. Daimi, H.; Khelil, A.H.; Neji, A.; Ben Hamda, K.; Maaoui, S.; Aranega, A.; Chibani, J.B.; Franco, D. Role of SCN5A coding and non-coding sequences in Brugada syndrome onset: What’s behind the scenes? Biomed. J. 2019, 42, 252–260. [Google Scholar] [CrossRef] [PubMed]
  94. Quinn, J.J.; Chang, H.Y. Unique features of long non-coding RNA biogenesis and function. Nat. Rev. Genet. 2016, 17, 47–62. [Google Scholar] [CrossRef] [PubMed]
  95. Bär, C.; Chatterjee, S.; Thum, T. Long Noncoding RNAs in Cardiovascular Pathology, Diagnosis, and Therapy. Circulation 2016, 134, 1484–1499. [Google Scholar] [CrossRef]
  96. Long, Q.Q.; Wang, H.; Gao, W.; Fan, Y.; Li, Y.F.; Ma, Y.; Yang, Y.; Shi, H.J.; Chen, B.R.; Meng, H.Y.; et al. Long noncoding RNA Kcna2 antisense RNA contributes to ventricular arrhythmias via silencing Kcna2 in rats with congestive heart failure. J. Am. Heart Assoc. 2017, 6, e005965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  97. Zhu, P.; Yang, M.; Ren, H.; Shen, G.; Chen, J.; Zhang, J.; Liu, J.; Sun, C. Long noncoding RNA MALAT1 downregulates cardiac transient outward potassium current by regulating miR-200c/HMGB1 pathway. J. Cell. Biochem. 2018, 119, 10239–10249. [Google Scholar] [CrossRef] [PubMed]
  98. Pan, Z. Binding of LncRNA-DACH1 to dystrophin impairs the membrane tra cking of Nav1.5 protein and increases ventricular arrhythmia susceptibility. Preprint 2021, 1–17. [Google Scholar] [CrossRef]
  99. Schmidt, J.W.; Catterall, W.A. Palmitylation, sulfation, and glycosylation of the alpha subunit of the sodium channel. Role of post-translational modifications in channel assembly. J. Biol. Chem. 1987, 262, 13713–13723. [Google Scholar] [CrossRef]
  100. Marionneau, C.; Abriel, H. Regulation of the cardiac Na+ channel Nav1.5 by post-translational modifications. J. Mol. Cell. Cardiol. 2015, 82, 36–47. [Google Scholar] [CrossRef]
  101. Abriel, H. Cardiac sodium channel Nav1.5 and interacting proteins: Physiology and pathophysiology. J. Mol. Cell. Cardiol. 2010, 48, 2–11. [Google Scholar] [CrossRef]
  102. Delisle, B.P.; Anson, B.D.; Rajamani, S.; January, C.T. Biology of Cardiac Arrhythmias Ion Channel Protein Trafficking. Circ. Res. 2004, 94, 1418–1428. [Google Scholar] [CrossRef] [Green Version]
  103. Jan, C.H.; Williams, C.C.; Weissman, J.S. Principles of ER cotranslational translocation revealed by proximity-specific ribosome profiling. Science 2014, 346, 716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Zhou, J.; Shin, H.G.; Yi, J.; Shen, W.; Williams, C.P.; Murray, K.T. Phosphorylation and putative ER retention signals are required for protein kinase A-mediated potentiation of cardiac sodium current. Circ. Res. 2002, 91, 540–546. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Abriel, H. Roles and regulation of the cardiac sodium channel Nav1.5: Recent insights from experimental studies. Cardiovasc. Res. 2007, 76, 381–389. [Google Scholar] [CrossRef] [Green Version]
  106. Fiona, S.; Cusdin, J.J.C.; Jackson, A.P. Trafficking and Cellular Distribution of Voltage-Gated Sodium Channels. Traffic 2008, 9, 17–26. [Google Scholar] [CrossRef]
  107. Herfst, L.J.; Rook, M.B.; Jongsma, H.J. Trafficking and functional expression of cardiac Na+ channels. J. Mol. Cell. Cardiol. 2004, 36, 185–193. [Google Scholar] [CrossRef]
  108. Ranek, M.J.; Stachowski, M.J.; Kirk, J.A.; Willis, M.S. The role of heat shock proteins and co-chaperones in heart failure. Phil. Trans. R. Soc. B 2017, 373, 20160530. [Google Scholar] [CrossRef] [Green Version]
  109. Remme, C.A.; Wilde, A.A. Targeting sodium channels in cardiac arrhythmia. Curr. Opin. Pharmacol. 2014, 15, 53–60. [Google Scholar] [CrossRef]
  110. Abriel, H.; Kass, R.S. Regulation of the voltage-gated cardiac sodium channel Nav1.5 by interacting proteins. Trends Cardiovasc. Med. 2005, 15, 35–40. [Google Scholar] [CrossRef]
  111. Shinlapawittayatorn, K.; Dudash, L.A.; Du, X.X.; Heller, L.; Poelzing, S.; Ficker, E.; Deschênes, I. A novel strategy using cardiac sodium channel polymorphic fragments to rescue trafficking-deficient SCN5A Mutations. Circ. Cardiovasc. Genet. 2011, 4, 500–509. [Google Scholar] [CrossRef] [Green Version]
  112. Moreau, A.; Keller, D.I.; Huang, H.; Fressart, V.; Schmied, C.; Timour, Q.; Chahine, M. Mexiletine differentially restores the trafficking defects caused by two Brugada syndrome mutations. Front. Pharmacol. 2012, 3, 1–8. [Google Scholar] [CrossRef] [Green Version]
  113. Mercier, A.; Clément, R.; Harnois, T.; Bourmeyster, N.; Bois, P.; Chatelier, A. Nav1.5 channels can reach the plasma membrane through distinct N-glycosylation states. Biochim. Biophys. Acta—Gen. Subj. 2015, 1850, 1215–1223. [Google Scholar] [CrossRef]
  114. Hartl, F.U.; Bracher, A.; Hayer-Hartl, M. Molecular chaperones in protein folding and proteostasis. Nature 2011, 475, 324–332. [Google Scholar] [CrossRef] [PubMed]
  115. Young, J.C. The role of the cytosolic HSP70 chaperone system in diseases caused by misfolding and aberrant trafficking of ion channels. DMM Dis. Models Mech. 2014, 7, 319–329. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  116. Li, K.; Jiang, Q.; Bai, X.; Yang, Y.F.; Ruan, M.Y.; Cai, S.Q. Tetrameric Assembly of K+ Channels Requires ER-Located Chaperone Proteins. Mol. Cell 2017, 65, 52–65. [Google Scholar] [CrossRef] [Green Version]
  117. Tarone, G.; Brancaccio, M. Keep your heart in shape: Molecular chaperone networks for treating heart disease. Cardiovasc. Res. 2014, 102, 346–361. [Google Scholar] [CrossRef] [Green Version]
  118. Penna, C.; Sorge, M.; Femminò, S.; Pagliaro, P.; Brancaccio, M. Redox aspects of chaperones in cardiac function. Front. Physiol. 2018, 9, 1–15. [Google Scholar] [CrossRef]
  119. Poelzing, S.; Forleo, C.; Samodell, M.; Dudash, L.; Sorrentino, S.; Anaclerio, M.; Troccoli, R.; Iacoviello, M.; Romito, R.; Guida, P.; et al. SCN5A polymorphism restores trafficking of a Brugada syndrome mutation on a separate gene. Circulation 2006, 114, 368–376. [Google Scholar] [CrossRef] [Green Version]
  120. Valdivia, C.R.; Tester, D.J.; Rok, B.A.; Porter, C.-B.J.; Munger, T.M.; Jahangir, A.; Makielski, J.C.; Ackerman, M.J. A trafficking defective, Brugada syndrome-causing SCN5A mutation rescued by drugs. Cardiovasc. Res. 2004, 62, 53–62. [Google Scholar] [CrossRef]
  121. Ulloa-Aguirre, A.; Janovick, J.A.; Brothers, S.P.; Conn, P.M. Pharmacologic rescue of conformationally-defective proteins: Implications for the treatment of human disease. Traffic 2004, 5, 821–837. [Google Scholar] [CrossRef]
  122. Cohen, S.A.; Levitt, L.K. Partial characterization of the rH1 sodium channel protein from rat heart using subtype-specific antibodies. Circ. Res. 1993, 73, 735–742. [Google Scholar] [CrossRef] [Green Version]
  123. Caramelo, J.J.; Parodi, A.J. A sweet code for glycoprotein folding. FEBS Lett. 2015, 589, 3379–3387. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Verges, M. Trafficking of Cardiac Ion Channels; MDPI AG: Basel, Switzerland, 2021; ISBN 9783039434725. [Google Scholar]
  125. Bennett, E.S. Isoform-specific effects of sialic acid on voltage-dependent Na+ channel gating: Functional sialic acids are localized to the S5–S6 loop of domain I. J. Physiol. 2002, 538, 675–690. [Google Scholar] [CrossRef] [PubMed]
  126. Stocker, P.J.; Bennett, E.S. Differential sialylation modulates voltage-gated Na+ channel gating throughout the developing myocardium. J. Gen. Physiol. 2006, 127, 253–265. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Arakel, E.C.; Brandenburg, S.; Uchida, K.; Zhang, H.; Lin, Y.W.; Kohl, T.; Schrul, B.; Sulkin, M.S.; Efimov, I.R.; Nichols, C.G.; et al. Tuning the electrical properties of the heart by differential trafficking of KATP ion channel complexes. J. Cell Sci. 2014, 127, 2106–2119. [Google Scholar] [CrossRef] [Green Version]
  128. Roth, J.; Zuber, C.; Park, S.; Jang, I.; Lee, Y.; Kysela, K.G.; Le Fourn, V.; Santimaria, R.; Guhl, B.; Cho, J.W. Protein N-glycosylation, protein folding and protein quality control. Mol. Cells 2010, 30, 497–506. [Google Scholar] [CrossRef]
  129. Xiao, X.; Chen, C.; Yu, T.M.; Ou, J.; Rui, M.; Zhai, Y.; He, Y.; Xue, L.; Ho, M.S. Molecular chaperone calnexin regulates the function of Drosophila sodium channel paralytic. Front. Mol. Neurosci. 2017, 10, 57. [Google Scholar] [CrossRef] [Green Version]
  130. Khanna, R.; Lee, E.J.; Papazian, D.M. Transient calnexin interaction confers long-term stability on folded K+ channel protein in the ER. J. Cell Sci. 2004, 117, 2897–2908. [Google Scholar] [CrossRef] [Green Version]
  131. Proft, J.; Rzhepetskyy, Y.; Lazniewska, J.; Zhang, F.X.; Cain, S.M.; Snutch, T.P.; Zamponi, G.W.; Weiss, N. The Cacna1h mutation in the GAERS model of absence epilepsy enhances T-type Ca2+ currents by altering calnexin-dependent trafficking of Cav3.2 channels. Sci. Rep. 2017, 7, 11513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Zumhagen, S.; Veldkamp, M.W.; Stallmeyer, B.; Baartscheer, A.; Eckardt, L.; Paul, M.; Remme, C.A.; Bhuiyan, Z.A.; Bezzina, C.R.; Schulze-Bahr, E. A Heterozygous Deletion Mutation in the Cardiac Sodium Channel Gene SCN5A with Loss- and Gain-of-Function Characteristics Manifests as Isolated Conduction Disease, without Signs of Brugada or Long QT Syndrome. PLoS ONE 2013, 8, e67963. [Google Scholar] [CrossRef]
  133. Casini, S.; Tan, H.L.; Demirayak, I.; Remme, C.A.; Amin, A.S.; Scicluna, B.P.; Chatyan, H.; Ruijter, J.M.; Bezzina, C.R.; Van Ginneken, A.C.G.; et al. Tubulin polymerization modifies cardiac sodium channel expression and gating. Cardiovasc. Res. 2010, 85, 691–700. [Google Scholar] [CrossRef] [Green Version]
  134. Abriel, H.; Sottas, V. Unexpected α-α interactions with nav1.5 genetic variants in brugada syndrome. Circ. Cardiovasc. Genet. 2014, 7, 97–99. [Google Scholar] [CrossRef] [Green Version]
  135. Gao, G.; Xie, A.; Zhang, J.; Herman, A.M.; Jeong, E.M.; Gu, L.; Liu, M.; Yang, K.C.; Kamp, T.J.; Dudley, S.C. Unfolded protein response regulates cardiac sodium current in systolic human heart failure. Circ. Arrhythmia Electrophysiol. 2013, 6, 1018–1024. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Ren, J.; Bi, Y.; Sowers, J.R.; Hetz, C.; Zhang, Y. Endoplasmic reticulum stress and unfolded protein response in cardiovascular diseases. Nat. Rev. Cardiol. 2021, 18, 499–521. [Google Scholar] [CrossRef] [PubMed]
  137. Iqbal, S.M.; Lemmens-Gruber, R. Phosphorylation of cardiac voltage-gated sodium channel: Potential players with multiple dimensions. Acta Physiol. 2019, 225, 1–18. [Google Scholar] [CrossRef]
  138. Schubert, B.; Vandongen, A.M.J.; Kirsch, G.E.; Brown, A.M. Inhibition of cardiac Na+ currents by isoproterenol. Am. J. Physiol.—Heart Circ. Physiol. 1990, 258. [Google Scholar] [CrossRef] [Green Version]
  139. Frohnwieser, B.; Chen, L.Q.; Schreibmayer, W.; Kallen, R.G. Modulation of the human cardiac sodium channel α-subunit by cAMP-dependent protein kinase and the responsible sequence domain. J. Physiol. 1997, 498, 309–318. [Google Scholar] [CrossRef] [Green Version]
  140. Murphy, B.J.; Rogers, J.; Perdichizzi, A.P.; Colvin, A.A.; Catterall, W.A. cAMP-dependent phosphorylation of two sites in the α subunit of the cardiac sodium channel. J. Biol. Chem. 1996, 271, 28837–28843. [Google Scholar] [CrossRef] [Green Version]
  141. Tateyama, M.; Rivolta, I.; Clancy, C.E.; Kass, R.S. Modulation of Cardiac Sodium Channel Gating by Protein Kinase A Can Be Altered by Disease-linked Mutation. J. Biol. Chem. 2003, 278, 46718–46726. [Google Scholar] [CrossRef] [Green Version]
  142. Rook, M.B.; Evers, M.M.; Vos, M.A.; Bierhuizen, M.F. Biology of cardiac sodium channel Nav1.5 expression. Cardiovasc. Res. 2012, 93, 12–23. [Google Scholar] [CrossRef] [Green Version]
  143. Zhou, J.; Yi, J.; Hu, N.N.; George, A.L.; Murray, K.T. Activation of protein kinase A modulates trafficking of the human cardiac sodium channel in Xenopus oocytes. Circ. Res. 2000, 87, 33–38. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Scott, D.B.; Blanpied, T.A.; Ehlers, M.D. Coordinated PKA and PKC phosphorylation suppresses RXR-mediated ER retention and regulates the surface delivery of NMDA receptors. Neuropharmacology 2003, 45, 755–767. [Google Scholar] [CrossRef]
  145. Hallaqa, H.; Yanga, Z.; Viswanathanb, P.C.; Fukudab, K.; Shena, W.; Wanga, D.W.; Wellsc, K.S.; Zhoua, J.; Yia, J.; Hallaqa, K.T.M.; et al. Quantitation of protein kinase a mediated trafficking of cardiac sodium channels in living cells. Cardiovasc. Res. 2006, 72, 250–261. [Google Scholar] [CrossRef] [PubMed]
  146. Qu, Y.; Rogers, J.; Tanada, T.; Scheuer, T.; Catterall, W.A. Modulation of cardiac Na+ channels expressed in a mammalian cell line and in ventricular myocytes by protein kinase C. Proc. Natl. Acad. Sci. USA 1994, 91, 3289–3293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Shin, H.G.; Murray, K.T. Conventional protein kinase C isoforms and cross-activation of protein kinase A regulate cardiac Na+ current. FEBS Lett. 2001, 495, 154–158. [Google Scholar] [CrossRef] [Green Version]
  148. Xiao, G.Q.; Qu, Y.; Sun, Z.Q.; Mochly-Rosen, D.; Boutjdir, M. Evidence for functional role of εPKC isozyme in the regulation of cardiac Na+ channels. Am. J. Physiol.—Cell Physiol. 2001, 281, C1477–C1486. [Google Scholar] [CrossRef]
  149. Ferreira, J.C.B.; Mochly-Rosen, D.; Boutjdir, M. Regulation of cardiac excitability by protein kinase C isozymes. Front. Biosci.—Sch. 2012, 4, 532–546. [Google Scholar] [CrossRef]
  150. Liu, M.; Shi, G.; Yang, K.C.; Gu, L.; Kanthasamy, A.G.; Anantharam, V.; Dudley, S.C., Jr. The role of protein kinase C in the metabolic regulation of the cardiac Na+ channel. Heart Rhythm 2017, 14, 440–447. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Mathieu, S.; El Khoury, N.; Rivard, K.; Gélinas, R.; Goyette, P.; Paradis, P.; Nemer, M.; Fiset, C. Reduction in Na+ current by angiotensin II is mediated by PKCα in mouse and human-induced pluripotent stem cell-derived cardiomyocytes. Heart Rhythm 2016, 13, 1346–1354. [Google Scholar] [CrossRef] [PubMed]
  152. Qu, Y.; Rogers, J.C.; Tanada, T.N.; Catterall, W.A.; Scheuer, T. Phosphorylation of S1505 in the cardiac Na+ channel inactivation gate is required for modulation by protein kinase C. J. Gen. Physiol. 1996, 108, 375–379. [Google Scholar] [CrossRef] [PubMed]
  153. Murray, K.T.; Hu, N.; Daw, J.R.; Shin, H.G.; Watson, M.T.; Mashburn, A.B.; George, A.L. Functional effects of protein kinase C activation on the human cardiac Na+ channel. Circ. Res. 1997, 80, 370–376. [Google Scholar] [CrossRef]
  154. Liu, M.; Sanyal, S.; Gao, G.; Gurung, I.S.; Zhu, X.; Gaconnet, G.; Kerchner, L.J.; Shang, L.L.; Huang, C.L.H.; Grace, A.; et al. Cardiac Na+ Current regulation by pyridine nucleotides. Circ. Res. 2009, 105, 737–745. [Google Scholar] [CrossRef] [Green Version]
  155. Strickland, M.; Yacoubi-Loueslati, B.; Bouhaouala-Zahar, B.; Pender, S.L.F.; Larbi, A. Relationships between ion channels, mitochondrial functions and inflammation in human aging. Front. Physiol. 2019, 10, 1–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  156. Hu, Y.F.; Wu, C.H.; Lai, T.C.; Chang, Y.C.; Hwang, M.J.; Chang, T.Y.; Weng, C.H.; Chang, P.M.H.; Chen, C.H.; Mochly-Rosen, D.; et al. ALDH2 deficiency induces atrial fibrillation through dysregulated cardiac sodium channel and mitochondrial bioenergetics: A multi-omics analysis. Biochim. Biophys. Acta—Mol. Basis Dis. 2021, 1867, 166088. [Google Scholar] [CrossRef] [PubMed]
  157. Valdivia, C.R.; Ueda, K.; Ackerman, M.J.; Makielski, J.C. GPD1L links redox state to cardiac excitability by PKC-dependentphosphorylation of the sodium channel SCN5A. Am. J. Physiol. Heart Circ. Physiol. 2009, 297, H1446–H1452. [Google Scholar] [CrossRef] [PubMed]
  158. Liu, M.; Liu, H.; Dudley, S.C. Reactive oxygen species originating from mitochondria regulate the cardiac sodium channel. Circ. Res. 2010, 107, 967–974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  159. Liu, M.; Gu, L.; Sulkin, M.S.; Liu, H.; Jeong, E.M.; Greener, I.; Xie, A.; Efimov, I.R.; Dudley, S.C. Mitochondrial dysfunction causing cardiac sodium channel downregulation in cardiomyopathy. J. Mol. Cell. Cardiol. 2013, 54, 25–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Matasic, D.S.; Yoon, J.Y.; McLendon, J.M.; Mehdi, H.; Schmidt, M.S.; Greiner, A.M.; Quinones, P.; Morgan, G.M.; Boudreau, R.L.; Irani, K.; et al. Modulation of the cardiac sodium channel Nav1.5 peak and late currents by NAD+ precursors. J. Mol. Cell. Cardiol. 2020, 141, 70–81. [Google Scholar] [CrossRef]
  161. Fouda, M.A.; Ghovanloo, M.R.; Ruben, P.C. Cannabidiol protects against high glucose-induced oxidative stress and cytotoxicity in cardiac voltage-gated sodium channels. Br. J. Pharmacol. 2020, 177, 2932–2946. [Google Scholar] [CrossRef]
  162. Fouda, M.A.; Ruben, P.C. Protein kinases mediate anti-inflammatory effects of cannabidiol and estradiol against high glucose in cardiac sodium channels. Front. Pharmacol. 2021, 12, 1–17. [Google Scholar] [CrossRef]
  163. Ashpole, N.M.; Herren, A.W.; Ginsburg, K.S.; Brogan, J.D.; Johnson, D.E.; Cummins, T.R.; Bers, D.M.; Hudmon, A. Ca2+/calmodulin-dependent protein kinase II (CaMKII) regulates cardiac sodium channel Nav1.5 gating by multiple phosphorylation sites. J. Biol. Chem. 2012, 287, 19856–19869. [Google Scholar] [CrossRef] [Green Version]
  164. Herren, A.W.; Weber, D.M.; Rigor, R.R.; Margulies, K.B.; Phinney, B.S.; Bers, D.M. CaMKII Phosphorylation of Nav1.5: Novel in Vitro Sites Identified by Mass Spectrometry and Reduced S516 Phosphorylation in Human Heart Failure. J. Proteome Res. 2015, 14, 2298–2311. [Google Scholar] [CrossRef] [Green Version]
  165. Burel, S.; Coyan, F.C.; Lorenzini, M.; Meyer, M.R.; Lichti, C.F.; Brown, J.H.; Loussouarn, G.; Charpentier, F.; Nerbonne, J.M.; Townsend, R.R.; et al. C-terminal phosphorylation of Nav1.5 impairs FGF13-dependent regulation of channel inactivation. J. Biol. Chem. 2017, 292, 17431–17448. [Google Scholar] [CrossRef] [Green Version]
  166. Wagner, S.; Dybkova, N.; Rasenack, E.C.L.; Jacobshagen, C.; Fabritz, L.; Kirchhof, P.; Maier, S.K.G.; Zhang, T.; Hasenfuss, G.; Brown, J.H.; et al. Ca2+/calmodulin-dependent protein kinase II regulates cardiac Na+ channels. J. Clin. Investig. 2006, 116, 3127–3138. [Google Scholar] [CrossRef] [PubMed]
  167. El Refaey, M.; Musa, H.; Murphy, N.P.; Lubbers, E.R.; Skaf, M.; Han, M.; Cavus, O.; Koenig, S.N.; Wallace, M.J.; Gratz, D.; et al. Protein Phosphatase 2A Regulates Cardiac Na+ Channels. Circ. Res. 2019, 124, 737–746. [Google Scholar] [CrossRef] [PubMed]
  168. Deschênes, I.; Neyroud, N.; DiSilvestre, D.; Marbán, E.; Yue, D.T.; Tomaselli, G.F. Isoform-specific modulation of voltage-gated Na+ channels by calmodulin. Circ. Res. 2002, 90, 49–57. [Google Scholar] [CrossRef] [Green Version]
  169. Ahern, C.A.; Zhang, J.F.; Wookalis, M.J.; Horn, R. Modulation of the cardiac sodium channel Nav1.5 by Fyn, a Src family tyrosine kinase. Circ. Res. 2005, 96, 991–998. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  170. Razani, B.; Woodman, S.E.; Lisanti, M.P. Caveolae: From cell biology to animal physiology. Pharmacol. Rev. 2002, 54, 431–467. [Google Scholar] [CrossRef]
  171. Maier, S.K.; Westenbroek, R.E.; Schenkman, K.A.; Feigl, E.O.; Scheuer, T.; Catterall, W.A. An unexpected role for brain type sodium channels in coupling of cell surface depolarization to contraction in the heart. Proc. Natl. Acad. Sci. USA 2002, 99, 4073–4078. [Google Scholar] [CrossRef] [Green Version]
  172. Iqbal, S.M.; Andavan, G.S.B.; Lemmens-Gruber, R. Differential modulation of fast inactivation in cardiac sodium channel splice variants by fyn tyrosine kinase. Cell. Physiol. Biochem. 2015, 37, 825–837. [Google Scholar] [CrossRef]
  173. Iqbal, S.M.; Aufy, M.; Shabbir, W.; Lemmens-Gruber, R. Identification of phosphorylation sites and binding pockets for modulation of Nav1.5 channel by Fyn tyrosine kinase. FEBS J. 2018, 285, 2520–2530. [Google Scholar] [CrossRef] [Green Version]
  174. Sarhan, M.F.; Van Petegem, F.; Ahern, C.A. A double tyrosine motif in the cardiac sodium channel domain III-IV linker couples calcium-dependent calmodulin binding to inactivation gating. J. Biol. Chem. 2009, 284, 33265–33274. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Jespersen, T.; Gavillet, B.; van Bemmelen, M.X.; Cordonier, S.; Thomas, M.A.; Staub, O.; Abriel, H. Cardiac sodium channel Nav1.5 interacts with and is regulated by the protein tyrosine phosphatase PTPH1. Biochem. Biophys. Res. Commun. 2006, 348, 1455–1462. [Google Scholar] [CrossRef] [PubMed]
  176. Beltran-Alvarez, P.; Pagans, S.; Brugada, R. The cardiac sodium channel is post-translationally modified by arginine methylation. J. Proteome Res. 2011, 10, 3712–3719. [Google Scholar] [CrossRef] [PubMed]
  177. Beltran-Alvarez, P.; Espejo, A.; Schmauder, R.; Beltran, C.; Mrowka, R.; Linke, T.; Batlle, M.; Pérez-Villa, F.; Pérez, G.J.; Scornik, F.S.; et al. Protein arginine methyl transferases-3 and -5 increase cell surface expression of cardiac sodium channel. FEBS Lett. 2013, 587, 3159–3165. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  178. Beltran-Alvarez, P.; Tarradas, A.; Chiva, C.; Pérez-Serra, A.; Batlle, M.; Pérez-Villa, F.; Schulte, U.; Sabidó, E.; Brugada, R.; Pagans, S. Identification of N-terminal protein acetylation and arginine methylation of the voltage-gated sodium channel in end-stage heart failure human heart. Curr. Ther. Res.—Clin. Exp. 2014, 76, 126–129. [Google Scholar] [CrossRef]
  179. Yoon, J.Y.; Vikram, A.; London, B.; Irani, K. Reversible lysine acetylation: Another layer of post-translational regulation of the cardiac sodium channel. Channels 2017, 11, 360–361. [Google Scholar] [CrossRef] [Green Version]
  180. Vikram, A.; Lewarchik, C.M.; Yoon, J.-Y.; Naqvi, A.; Kumar, S.; Morgan, G.M.; Jacobs, J.S.; Li, Q.; Kim, Y.-R.; Kassan, M.; et al. Sirtuin 1 regulates cardiac electrical activity by deacetylating the cardiac sodium channel. Nat. Med. 2017, 23, 361–370. [Google Scholar] [CrossRef]
  181. Henley, J.M.; Craig, T.J.; Wilkinson, K.A. Neuronal Sumoylation: Mechanisms, Dysfunction. Physiol. Rev. 2014, 94, 1249–1285. [Google Scholar] [CrossRef]
  182. Plant, L.D.; Xiong, D.; Romero, J.; Dai, H.; Goldstein, S.A.N. Hypoxia Produces Pro-arrhythmic Late Sodium Current in Cardiac Myocytes by SUMOylation of Nav1.5 Channels. Cell Rep. 2020, 30, 2225–2236.e4. [Google Scholar] [CrossRef] [Green Version]
  183. Hess, D.T.; Matsumoto, A.; Kim, S.O.; Marshall, H.E.; Stamler, J.S. Protein S-nitrosylation: Purview and parameters. Nat. Rev. Mol. Cell Biol. 2005, 6, 150–166. [Google Scholar] [CrossRef]
  184. Nathan, C.; Xie, Q. wen Nitric oxide synthases: Roles, tolls, and controls. Cell 1994, 78, 915–918. [Google Scholar] [CrossRef]
  185. Ahern, G.P.; Hsu, S.F.; Klyachko, V.A.; Jackson, M.B. Induction of persistent sodium current by exogenous and endogenous nitric oxide. J. Biol. Chem. 2000, 275, 28810–28815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Cheng, J.; Van Norstrand, D.W.; Medeiros-Domingo, A.; Valdivia, C.; Tan, B.; Ye, B.; Kroboth, S.; Vatta, M.; Tester, D.J.; January, C.T.; et al. α1-Syntrophin Mutations Identified in Sudden Infant Death Syndrome Cause an Increase in Late Cardiac Sodium Current. Circ. Arrhythmia Electrophysiol. 2009, 2, 667–676. [Google Scholar] [CrossRef] [Green Version]
  187. Cheng, J.; Valdivia, C.R.; Vaidyanathan, R.; Balijepalli, R.C.; Ackerman, M.J.; Makielski, J.C. Caveolin-3 suppresses late sodium current by inhibiting nNOS-dependent S-nitrosylation of SCN5A. J. Mol. Cell. Cardiol. 2013, 61, 102–110. [Google Scholar] [CrossRef] [Green Version]
  188. Wang, P.; Wei, M.; Zhu, X.; Liu, Y.; Yoshimura, K.; Zheng, M.; Liu, G.; Kume, S.; Morishima, M.; Kurokawa, T.; et al. Nitric oxide down-regulates voltage-gated Na+ channel in cardiomyocytes possibly through S-nitrosylation-mediated signaling. Sci. Rep. 2021, 11, 11273. [Google Scholar] [CrossRef] [PubMed]
  189. Krüger, E.; Hahn, K. Lipoxidation products as biomarkers of oxidative damage to proteins during lipid peroxidation reactions. Nephrol.—Dial.—Transplant. 2010, 11, 1–191. [Google Scholar]
  190. Herren, A.W.; Bers, D.M.; Grandi, E. Post-translational modifications of the cardiac Na channel: Contribution of CaMKII-dependent phosphorylation to acquired arrhythmias. Am. J. Physiol.—Heart Circ. Physiol. 2013, 305, H431–H445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  191. Nakajima, T.; Davies, S.S.; Matafonova, E.; Potet, F.; Amarnath, V.; Tallman, K.A.; Serwa, R.A.; Porter, N.A.; Balser, J.R.; Kupershmidt, S.; et al. Selective γ-ketoaldehyde scavengers protect Nav1.5 from oxidant-induced inactivation. J. Mol. Cell. Cardiol. 2010, 48, 352–359. [Google Scholar] [CrossRef] [Green Version]
  192. Quiñonez, M.; DiFranco, M.; González, F. Involvement of methionine residues in the fast inactivation mechanism of the sodium current from toad skeletal muscle fibers. J. Membr. Biol. 1999, 169, 83–90. [Google Scholar] [CrossRef]
  193. Kassmann, M.; Hansel, A.; Leipold, E.; Birkenbeil, J.; Lu, S.Q.; Hoshi, T.; Heinemann, S.H. Oxidation of multiple methionine residues impairs rapidsodium channel inactivation. Pflugers Arch.—Eur. J. Physiol. 2008, 456, 1085–1095. [Google Scholar] [CrossRef] [Green Version]
  194. Erickson, J.R.; Mei-ling, A.J.; Guan, X.; Kutschke, W.; Yang, J.; Oddis, C.V.; Bartlett, R.K.; Lowe, J.S.; O’Donnell, S.E.; Aykin-Burns, N.; et al. A Dynamic Pathway for Calcium-Independent Activation of CaMKII by Methionine Oxidation. Cell 2008, 133, 462–474. [Google Scholar] [CrossRef] [Green Version]
  195. Chamberlain, L.H.; Shipston, M.J. The physiology of protein s-acylation. Physiol. Rev. 2015, 95, 341–376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  196. Chen, L.; Bi, D.; Tian, L.; McClafferty, H.; Steeb, F.; Ruth, P.; Knaus, H.G.; Shipston, M.J. Palmitoylation of the β4-subunit regulates surface expression of large conductance calcium-activated potassium channel splice variants. J. Biol. Chem. 2013, 288, 13136–13144. [Google Scholar] [CrossRef] [Green Version]
  197. Pei, Z.; Xiao, Y.; Meng, J.; Hudmon, A.; Cummins, T.R. Cardiac sodium channel palmitoylation regulates channel availability and myocyte excitability with implications for arrhythmia generation. Nat. Commun. 2016, 7, 12035. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  198. Kahl, G. Protein truncation test (PTT; protein truncation assay, PTA). Dict. Genom. Transcr. Proteom. 2015, 1. [Google Scholar] [CrossRef]
  199. Zimmer, T.; Biskup, C.; Dugarmaa, S.; Vogel, F.; Steinbis, M.; Böhle, T.; Wu, Y.S.; Dumaine, R.; Benndorf, K. Functional expression of GFP-linked human heart sodium channel (hH1) and subcellular localization of the α subunit in HEK293 cells and dog cardiac myocytes. J. Membr. Biol. 2002, 186, 1–12. [Google Scholar] [CrossRef]
  200. Wu, L.; Yong, S.L.; Fan, C.; Ni, Y.; Yoo, S.; Zhang, T.; Zhang, X.; Obejero-Paz, C.A.; Rho, H.J.; Ke, T.; et al. Identification of a new co-factor, MOG1, required for the full function of cardiac sodium channel Nav1.5. J. Biol. Chem. 2008, 283, 6968–6978. [Google Scholar] [CrossRef] [Green Version]
  201. Chakrabarti, S.; Wu, X.; Yang, Z.; Wu, L.; Yong, S.L.; Zhang, C.; Hu, K.; Wang, Q.K.; Chen, Q. MOG1 Rescues Defective Trafficking of Nav1.5 Mutations in Brugada Syndrome and Sick Sinus Syndrome. Circ. Arrhythmia Electrophysiol. 2013, 6, 392–401. [Google Scholar] [CrossRef] [Green Version]
  202. Wang, Z.; Yu, G.; Liu, Y.; Liu, S.; Aridor, M.; Huang, Y.; Hu, Y.; Wang, L.; Li, S.; Xiong, H.; et al. Small GTPases SAR1A and SAR1B regulate the trafficking of the cardiac sodium channel Nav1.5. Biochim. Biophys. Acta—Mol. Basis Dis. 2018, 1864, 3672–3684. [Google Scholar] [CrossRef]
  203. Chatin, B.; Colombier, P.; Gamblin, A.L.; Allouis, M.; Le Bouffant, F. Dynamitin affects cell-surface expression of voltage-gated sodium channel Nav1.5. Biochem. J. 2014, 463, 339–349. [Google Scholar] [CrossRef]
  204. Ednie, A.R.; Horton, K.K.; Wu, J.; Bennett, E.S. Expression of the sialyltransferase, ST3Gal4, impacts cardiac voltage-gated sodium channel activity, refractory period and ventricular conduction. J. Mol. Cell. Cardiol. 2013, 59, 117–127. [Google Scholar] [CrossRef] [PubMed]
  205. Baroni, D.; Picco, C.; Moran, O. A mutation of SCN1B associated with GEFS+ causes functional and maturation defects of the voltage-dependent sodium channel. Hum. Mutat. 2018, 39, 1402–1415. [Google Scholar] [CrossRef] [PubMed]
  206. Rosnoblet, C.; Peanne, R.; Legrand, D.; Foulquier, F. Glycosylation disorders of membrane trafficking. Glycoconj. J. 2013, 30, 23–31. [Google Scholar] [CrossRef] [PubMed]
  207. Bennett, E.; Urcan, M.S.; Tinkle, S.S.; Koszowski, A.G.; Levinson, S.R. Contribution of sialic acid to the voltage dependence of sodium channel gating: A possible electrostatic mechanism. J. Gen. Physiol. 1997, 109, 327–343. [Google Scholar] [CrossRef] [Green Version]
  208. Ufret-Vincenty, C.A.; Baro, D.J.; Lederer, W.J.; Rockman, H.A.; Quiñones, L.E.; Santana, L.F. Role of Sodium Channel Deglycosylation in the Genesis of Cardiac Arrhythmias in Heart Failure. J. Biol. Chem. 2001, 276, 28197–28203. [Google Scholar] [CrossRef] [Green Version]
  209. Ednie, A.R.; Bennett, E.S. Modulation of voltage-gated ion channels by sialylation. Compr. Physiol. 2012, 2, 1269–1301. [Google Scholar] [CrossRef]
  210. Gu, F.; Crump, C.M.; Thomas, G. Trans-golgi network sorting. Cell. Mol. Life Sci. 2001, 58, 1067–1084. [Google Scholar] [CrossRef] [Green Version]
  211. Ponce-Balbuena, D.; Guerrero-Serna, G.; Valdivia, C.R.; Caballero, R.; Diez-Guerra, F.J.; Jiménez-Vázquez, E.N.; Ramírez, R.J.; Monteiro Da Rocha, A.; Herron, T.J.; Campbell, K.F.; et al. Cardiac Kir2.1 and Nav1.5 Channels Traffic Together to the Sarcolemma to Control Excitability. Circ. Res. 2018, 122, 1501–1516. [Google Scholar] [CrossRef]
  212. Willis, B.C.; Ponce-Balbuena, D.; Jalife, J. Protein assemblies of sodium and inward rectifier potassium channels control cardiac excitability and arrhythmogenesis. Am. J. Physiol.—Heart Circ. Physiol. 2015, 308, H1463–H1473. [Google Scholar] [CrossRef] [Green Version]
  213. Yang, H.-Q.; Perez-Hernandez, M.; Sanchez-Alonso, J.; Shevchuk, A.; Gorelik, J.; Rothenberg, E.; Delmar, M.; Coetzee, W.A. Ankyrin-G mediates targeting of both Na+ and K ATP channels to the rat cardiac intercalated disc. Elife 2020, 9, 1–23. [Google Scholar] [CrossRef]
  214. Shy, D.; Gillet, L.; Abriel, H. Targeting the sodium channel Nav1.5 to specific membrane compartments of cardiac cells not a simple task! Circ. Res. 2014, 115, 901–903. [Google Scholar] [CrossRef] [Green Version]
  215. Maier, S.K.G.; Westenbroek, R.E.; McCormick, K.A.; Curtis, R.; Scheuer, T.; Catterall, W.A. Distinct Subcellular Localization of Different Sodium Channel α and β Subunits in Single Ventricular Myocytes from Mouse Heart. Circulation 2004, 109, 1421–1427. [Google Scholar] [CrossRef] [Green Version]
  216. Domínguez, J.N.; De La Rosa, Á.; Navarro, F.; Franco, D.; Aránega, A.E. Tissue distribution and subcellular localization of the cardiac sodium channel during mouse heart development. Cardiovasc. Res. 2008, 78, 45–52. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Cohen, S.A. Immunocytochemical localization of rh1 sodium channel in adult rat heart atria and ventricle: Presence in terminal intercalated disks. Circulation 1996, 94, 3083–3086. [Google Scholar] [CrossRef] [PubMed]
  218. Shy, D.; Gillet, L.; Abriel, H. Cardiac sodium channel Nav1.5 distribution in myocytes via interacting proteins: The multiple pool model. Biochim. Biophys. Acta—Mol. Cell Res. 2013, 1833, 886–894. [Google Scholar] [CrossRef] [Green Version]
  219. Peeters, U.; Scornik, F.; Riuró, H.; Pérez, G.; Komurcu-Bayrak, E.; Van Malderen, S.; Pappaert, G.; Tarradas, A.; Pagans, S.; Daneels, D.; et al. Contribution of cardiac sodium channel β-subunit variants to brugada syndrome. Circ. J. 2015, 79, 2118–2129. [Google Scholar] [CrossRef] [Green Version]
  220. Salvage, S.C.; Rees, J.S.; McStea, A.; Hirsch, M.; Wang, L.; Tynan, C.J.; Reed, M.W.; Irons, J.R.; Butler, R.; Thompson, A.J.; et al. Supramolecular clustering of the cardiac sodium channel Nav1.5 in HEK293F cells, with and without the auxiliary β3-subunit. FASEB J. 2020, 34, 3537–3553. [Google Scholar] [CrossRef]
  221. Catterall, W.A. Voltage-gated sodium channels at 60_ structure, function and pathophysiology. J. Physiol. 2012, 590, 2577–2589. [Google Scholar] [CrossRef]
  222. Chen, C.; Calhoun, J.D.; Zhang, Y.; Lopez-Santiago, L.; Zhou, N.; Davis, T.H.; Salzer, J.L.; Isom, L.L. Identification of the cysteine residue responsible for disulfide linkage of Na+ channel alpha and beta2. J. Biol. Chem. 2012, 287, 39061–39069. [Google Scholar] [CrossRef] [Green Version]
  223. Yu, F.H.; Westenbroek, R.E.; Silos-santiago, I.; Mccormick, K.A.; Lawson, D.; Ge, P.; Ferriera, H.; Lilly, J.; Distefano, P.S.; Catterall, W.A.; et al. Sodium Channel β4, a New Disulfide-Linked Auxiliary Subunit with Similarity to β2. J. Neurosci. 2003, 23, 7577–7585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  224. Zhu, W.; Voelker, T.L.; Varga, Z.; Schubert, A.R.; Nerbonne, J.M.; Silva, J.R. Mechanisms of noncovalent β subunit regulation of Nav channel gating. J. Gen. Physiol. 2017, 149, 813–831. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  225. Malhotra, J.D.; Kazen-Gillespie, K.; Hortsch, M.; Isom, L.L. Sodium channel β subunits mediate homophilic cell adhesion and recruit ankyrin to points of cell-cell contact. J. Biol. Chem. 2000, 275, 11383–11388. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  226. Domínguez, J.N.; Navarro, F.; Franco, D.; Thompson, R.P.; Aránega, A.E. Temporal and spatial expression pattern ofh1 sodium channel subunitduring heart development. Cardiovasc. Res. 2004, 65, 842–850. [Google Scholar] [CrossRef] [Green Version]
  227. Zimmer, T.; Biskup, C.; Bollensdorff, C.; Benndorf, K. The β1 Subunit but not the β2 Subunit Colocalizes with the Human Heart Na+ Channel (hHl) already within the Endoplasmic Reticulum. J. Membr. Biol. 2002, 186, 13–21. [Google Scholar] [CrossRef]
  228. Mercier, A.; Clement, R.; Harnois, T.; Bourmeyster, N.; Faivre, J.F.; Findlay, I.; Chahine, M.; Bois, P.; Chatelier, A. The β1-Subunit of Nav1.5 Cardiac Sodium Channel Is Required for a Dominant Negative Effect through α-α Interaction. PLoS ONE 2012, 7, e48690. [Google Scholar]
  229. Dulsat, G.; Palomeras, S.; Cortada, E.; Riuró, H.; Brugada, R.; Vergés, M. Trafficking and localisation to the plasma membrane of Na v 1.5 promoted by the β2 subunit is defective due to a β2 mutation associated with Brugada syndrome. Biol. Cell 2017, 109, 273–291. [Google Scholar] [CrossRef]
  230. Namadurai, S.; Balasuriya, D.; Rajappa, R.; Wiemhöfer, M.; Stott, K.; Klingauf, J.; Edwardson, J.M.; Chirgadze, D.Y.; Jackson, A.P. Crystal structure and molecular imaging of the Nav channelβ3 subunit indicates a trimeric assembly. J. Biol. Chem. 2014, 289, 10797–10811. [Google Scholar] [CrossRef] [Green Version]
  231. Medeiros-Domingo, A.; Kaku, T.; Tester, D.J.; Iturralde-Torres, P.; Itty, A.; Ye, B.; Valdivia, C.; Ueda, K.; Canizales-Quinteros, S.; Tusié-Luna, M.T.; et al. SCN4B-Encoded Sodium Channel β4 Subunit in Congenital Long-QT Syndrome. Circulation 2007, 116, 134–142. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Watanabe, H.; Koopmann, T.T.; Le Scouarnec, S.; Yang, T.; Ingram, C.R.; Schott, J.-J.; Demolombe, S.; Probst, V.; Anselme, F.; Escande, D.; et al. Sodium channel β1 subunit mutations associated with Brugada syndrome and cardiac conduction disease in humans. J. Clin. Investig. 2008, 118, 2260–2268. [Google Scholar] [CrossRef] [Green Version]
  233. Wilde, A.A.M.; Brugada, R. Phenotypical manifestations of mutations in the genes encoding subunits of the cardiac sodium channel. Circ. Res. 2011, 108, 884–897. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  234. O’Malley, H.A.; Isom, L.L. Sodium channel β subunits: Emerging targets in channelopathies. Annu. Rev. Physiol. 2015, 77, 481–504. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  235. Qu, Y.; Curtis, R.; Lawson, D.; Gilbride, K.; Ge, P.; DiStefano, P.S.; Silos-Santiago, I.; Catterall, W.A.; Scheuer, T. Differential Modulation of Sodium Channel Gating and Persistent Sodium Currents by the β1, β2, and β3 Subunits. Mol. Cell. Neurosci. 2001, 18, 570–580. [Google Scholar] [CrossRef] [PubMed]
  236. Morgan, K.; Stevens, E.B.; Shah, B.; Cox, P.J.; Dixon, A.K.; Lee, K.; Pinnock, R.D.; Hughes, J.; Richardson, P.J.; Mizuguchi, K.; et al. β3: An additional auxiliary subunit of the voltage-sensitive sodium channel that modulates channel gating with distinct kinetics. Proc. Natl. Acad. Sci. USA 2000, 97, 2308–2313. [Google Scholar] [CrossRef] [Green Version]
  237. Delmar, M. Connexin43 regulates sodium current; Ankyrin-G modulates gap junctions: The intercalated disc exchanger. Cardiovasc. Res. 2012, 93, 220–222. [Google Scholar] [CrossRef]
  238. Abriel, H.; Rougier, J.S.; Jalife, J. Ion Channel Macromolecular Complexes in Cardiomyocytes: Roles in Sudden Cardiac Death. Circ. Res. 2015, 116, 1971–1988. [Google Scholar] [CrossRef] [Green Version]
  239. Godreau, D.; Vranckx, R.; Maguy, A.; Goyenvalle, C.; Hatem, S.N. Different Isoforms of Synapse-associated Protein, SAP97, Are Expressed in the Heart and Have Distinct Effects on the Voltage-gated K+ Channel Kv1.5. J. Biol. Chem. 2003, 278, 47046–47052. [Google Scholar] [CrossRef] [Green Version]
  240. Milstein, M.L.; Musa, H.; Balbuena, D.P.; Anumonwo, J.M.B.; Auerbach, D.S.; Furspan, P.B.; Hou, L.; Hu, B.; Schumacher, S.M.; Vaidyanathan, R.; et al. Dynamic reciprocity of sodium and potassium channel expression in a macromolecular complex controls cardiac excitability and arrhythmia. Proc. Natl. Acad. Sci. USA 2012, 109, E2134–E2143. [Google Scholar] [CrossRef] [Green Version]
  241. Petitprez, S.; Zmoos, A.F.; Ogrodnik, J.; Balse, E.; Raad, N.; El-Haou, S.; Albesa, M.; Bittihn, P.; Luther, S.; Lehnart, S.E.; et al. SAP97 and dystrophin macromolecular complexes determine two pools of cardiac sodium channels Nav1.5 in cardiomyocytes. Circ. Res. 2011, 108, 294–304. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Gillet, L.; Rougier, J.S.; Shy, D.; Sonntag, S.; Mougenot, N.; Essers, M.; Shmerling, D.; Balse, E.; Hatem, S.N.; Abriel, H. Cardiac-specific ablation of synapse-associated protein SAP97 in mice decreases potassium currents but not sodium current. Heart Rhythm 2015, 12, 181–192. [Google Scholar] [CrossRef] [PubMed]
  243. Mohler, P.J. Editorial: Sodium channel traffic on the cardiac microtubule highway. Cardiovasc. Res. 2010, 85, 645–646. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  244. Agullo-Pascual, E.; Lin, X.; Leo-Macias, A.; Zhang, M.; Liang, F.; Li, Z.; Pfenniger, A.; Bkemeier, I.L.; Keegan, S.; Fenyo, D.; et al. Super-resolution imaging reveals that loss of theC-terminus of connexin43 limits microtubuleplus-end capture and Nav1.5 localization at theintercalated disc. Cardiovasc. Res. 2014, 104, 371–381. [Google Scholar] [CrossRef] [PubMed]
  245. Marchal, G.A.; Jouni, M.; Chiang, D.Y.; Pérez-Hernández, M.; Podliesna, S.; Yu, N.; Casini, S.; Potet, F.; Veerman, C.C.; Klerk, M.; et al. Targeting the Microtubule EB1-CLASP2 Complex Modulates Nav1.5 at Intercalated Discs. Circ. Res. 2021, 129, 349–365. [Google Scholar] [CrossRef] [PubMed]
  246. Mimori-Kiyosue, Y.; Grigoriev, I.; Lansbergen, G.; Sasaki, H.; Matsui, C.; Severin, F.; Galjart, N.; Grosveld, F.; Vorobjev, I.; Tsukita, S.; et al. CLASP1 and CLASP2 bind to EB1 and regulate microtubule plus-end dynamics at the cell cortex. J. Cell Biol. 2005, 168, 141–153. [Google Scholar] [CrossRef]
  247. Desplantez, T. Cardiac Cx43, Cx40 and Cx45 co-assembling: Involvement of connexins epitopes in formation of hemichannels and Gap junction channels. BMC Cell Biol. 2017, 18, 1–13. [Google Scholar] [CrossRef] [Green Version]
  248. Hunter, A.W.; Barker, R.J.; Zhu, C.; Gourdie, R.G. Zonula occludens-1 alters connexin43 gap junction size and organization by influencing channel accretion. Mol. Biol. Cell 2005, 16, 5686–5698. [Google Scholar] [CrossRef] [Green Version]
  249. Rhett, J.M.; Ongstad, E.L.; Jourdan, J.; Gourdie, R.G. Cx43 Associates with Nav1.5 in the Cardiomyocyte Perinexus. J. Membr. Biol. 2012, 245, 411–422. [Google Scholar]
  250. Noorman, M.; Hakim, S.; Kessler, E.; Groeneweg, J.; Cox, M.G.P.J.; Asimaki, A.; van Rijen, H.V.M.; van Stuijvenberg, L.; Chkourko, H.; van der Heyden, M.A.G.; et al. Remodeling of the cardiac sodium channel, Connexin43 andPlakoglobin at the intercalated disk in patients witharrhythmogenic cardiomyopathy. Heart Rhythm 2013, 10, 412–419. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  251. Rhett, J.M.; Veeraraghavan, R.; Poelzing, S.; Gourdie, R.G. The perinexus: Sign-post on the path to a new model of cardiac conduction? Trends Cardiovasc. Med. 2013, 23, 222–228. [Google Scholar] [CrossRef] [Green Version]
  252. Cunha, S.R.; Mohler, P.J. Cardiac ankyrins: Essential components for development and maintenance of excitable membrane domains in heart. Cardiovasc. Res. 2006, 71, 22–29. [Google Scholar] [CrossRef] [Green Version]
  253. Hashemi, S.M.; Hund, T.J.; Mohler, P.J. Cardiac ankyrins in health and disease. J. Mol. Cell. Cardiol. 2009, 47, 203–209. [Google Scholar] [CrossRef] [Green Version]
  254. Makara, M.A.; Curran, J.; Little, S.C.; Musa, H.; Polina, I.; Smith, S.A.; Wright, P.J.; Unudurthi, S.D.; Snyder, J.; Bennett, V.; et al. Ankyrin-G Coordinates Intercalated Disc Signaling Platform to Regulate Cardiac Excitability In Vivo. Circ Res. 2014, 115, 929–938. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  255. Hund, T.J.; Koval, O.M.; Li, J.; Wright, P.J.; Qian, L.; Snyder, J.S.; Gudmundsson, H.; Kline, C.F.; Davidson, N.P.; Cardona, N.; et al. A β IV -spectrin/CaMKII signaling complex is essential for membrane excitability in mice. J. Clin. Investig. 2010, 120, 3508–3519. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  256. Sato, P.Y.; Coombs, W.; Lin, X.; Nekrasova, O.; Green, K.J.; Isom, L.L.; Taffet, S.M.; Delmar, M. Interactions Between Ankyrin-G, Plakophilin-2, and Connexin43 at the Cardiac Intercalated Disc. Circ Res. 2011, 109, 193–201. [Google Scholar] [CrossRef] [Green Version]
  257. Makara, M.A.; Curran, J.; Lubbers, E.R.; Murphy, N.P.; Little, S.C.; Musa, H.; Smith, S.A.; Unudurthi, S.D.; Rajaram, M.V.S.; Janssen, P.M.L.; et al. Novel Mechanistic Roles for Ankyrin-G in Cardiac Remodeling and Heart Failure. JACC Basic Transl. Sci. 2018, 3, 675–689. [Google Scholar] [CrossRef] [PubMed]
  258. Mohler, P.J.; Rivolta, I.; Napolitano, C.; Lemaillet, G.; Lambert, S.; Priori, S.G.; Bennett, V. Nav1.5 E1053K mutation causing Brugada syndrome blocks binding to ankyrin-G and expression of Nav1.5 on the surface of cardiomyocytes. Proc. Natl. Acad. Sci. USA 2004, 101, 17533–17538. [Google Scholar] [CrossRef] [Green Version]
  259. Cavus, O.; Williams, J.; Musa, H.; el Refaey, M.; Gratz, D.; Shaheen, R.; Schwieterman, N.A.; Koenig, S.; Antwi-Boasiako, S.; Young, L.J.; et al. Giant ankyrin-G regulates cardiac function. J. Biol. Chem. 2021, 296, 1–13. [Google Scholar] [CrossRef]
  260. Cerrone, M.; Delmar, M. Desmosomes and the sodium channel complex: Implications for arrhythmogenic cardiomyopathy and Brugada syndrome. Trends Cardiovasc. Med. 2014, 24, 184–190. [Google Scholar] [CrossRef] [Green Version]
  261. Cerrone, M.; Lin, X.; Zhang, M.; Agullo-Pascual, E.; Pfenniger, A.; Chkourko Gusky, H.; Novelli, V.; Kim, C.; Tirasawadichai, T.; Judge, D.P.; et al. Missense mutations in plakophilin-2 cause sodium current deficit and associate with a brugada syndrome phenotype. Circulation 2014, 129, 1092–1103. [Google Scholar] [CrossRef] [Green Version]
  262. Cerrone, M.; Lin, X.; Zhang, M.; Agullo-Pascual, E.; Pfenniger, A.; Chkourko Gusky, H.; Novelli, V.; Kim, C.; Tirasawadischai, T.; Judge, D.P.; et al. Missense Mutations in Plakophilin-2 Can Cause Brugada Syndrome Phenotype By Decreasing Sodium Current and Nav1.5 Membrane Localization. Heart Rhythm 2013, 10, 1743. [Google Scholar] [CrossRef]
  263. Sato, P.Y.; Musa, H.; Coombs, W.; Guerrero-Serna, G.; Patiño, G.A.; Taffet, S.M.; Isom, L.L.; Delmar, M. Loss of plakophilin-2 expression leads to decreased sodium current and slower conduction velocity in cultured cardiac myocytes. Circ. Res. 2009, 105, 523–526. [Google Scholar] [CrossRef] [Green Version]
  264. Kim, C.; Wong, J.; Wen, J.; Wang, S.; Wang, C.; Spiering, S.; Kan, N.G.; Forcales, S.; Puri, P.L.; Leone, T.C.; et al. Studying arrhythmogenic right ventricular dysplasia with patient-specific iPSCs. Nature 2013, 494, 105–110. [Google Scholar] [CrossRef] [Green Version]
  265. Rizzo, S.; Lodder, E.M.; Verkerk, A.O.; Wolswinkel, R.; Beekman, L.; Pilichou, K.; Basso, C.; Remme, C.A.; Thiene, G.; Bezzina, C.R. Intercalated disc abnormalities, reduced Na+ current density, and conduction slowing in desmoglein-2 mutant mice prior to cardiomyopathic changes. Cardiovasc. Res. 2012, 95, 409–418. [Google Scholar] [CrossRef] [Green Version]
  266. Zhang, Q.; Deng, C.; Rao, F.; Modi, R.M.; Zhu, J.; Liu, X.; Mai, L.; Tan, H.; Yu, X.; Lin, Q.; et al. Silencing of desmoplakin decreases connexin43/Nav1.5 expression and sodium current in HL-1 cardiomyocytes. Mol. Med. Rep. 2013, 8, 780–786. [Google Scholar] [CrossRef]
  267. Cunha, S.R.; Mohler, P.J. Ankyrin protein networks in membrane formation and stabilization. J. Cell. Mol. Med. 2009, 13, 4364–4376. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  268. Nassal, D.; Yu, J.; Min, D.; Lane, C.; Shaheen, R.; Gratz, D.; Hund, T.J. Regulation of cardiac conduction and arrhythmias by ankyrin/spectrin-based macromolecular complexes. J. Cardiovasc. Dev. Dis. 2021, 8, 48. [Google Scholar] [CrossRef] [PubMed]
  269. Bennett, V.; Chen, L. Ankyrins and cellular targeting of diverse membrane proteins to physiological sites. Curr. Opin. Cell Biol. 2001, 13, 61–67. [Google Scholar] [CrossRef]
  270. Glynn, P.; Musa, H.; Wu, X.; Unudurthi, S.D.; Little, S.; Qian, L.; Wright, P.J.; Radwanski, P.B.; Gyorke, S.; Mohler, P.J.; et al. Voltage-Gated Sodium Channel Phosphorylation at Ser571 Regulates Late Current, Arrhythmia, and Cardiac Function in Vivo. Circulation 2015, 132, 567–577. [Google Scholar] [CrossRef] [Green Version]
  271. Greer-Short, A.; Musa, H.; Alsina, K.M.; Ni, L.; Word, T.A.; Reynolds, J.O.; Gratz, D.; Lane, C.; El-Refaey, M.; Unudurthi, S.; et al. Calmodulin kinase II regulates atrial myocyte late sodium current, calcium handling, and atrial arrhythmia. Heart Rhythm 2020, 17, 503–511. [Google Scholar] [CrossRef]
  272. Marsman, R.F.; Bezzina, C.R.; Freiberg, F.; Verkerk, A.O.; Adriaens, M.E.; Podliesna, S.; Chen, C.; Purfürst, B.; Spallek, B.; Koopmann, T.T.; et al. Coxsackie and adenovirus receptor (CAR) is a modifier of cardiac conduction and arrhythmia vulnerability in the setting of myocardial ischemia. J. Am. Coll. Cardiol. 2015, 63, 549–559. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  273. Noutsias, M.; Fechner, H.; De Jonge, H.; Wang, X.; Dekkers, D.; Houtsmuller, A.B.; Pauschinger, M.; Bergelson, J.; Warraich, R.; Yacoub, M.; et al. Human Coxsackie-Adenovirus Receptor Is Colocalized With Integrins avb3 and avb5 on the Cardiomyocyte Sarcolemma and Upregulated in Dilated Cardiomyopathy Implications for Cardiotropic Viral Infections. Heart 2001, 104, 275–280. [Google Scholar]
  274. Clatot, J.; Ziyadeh-Isleem, A.; Maugenre, S.; Denjoy, I.; Liu, H.; Dilanian, G.; Hatem, S.N.; Deschenes, I.; Coulombe, A.; Guicheney, P.; et al. Dominant-negative effect of SCN5A N-terminal mutations through the interaction of Nav1.5 alpha subunits. Cardiovasc. Res. 2012, 96, 53–56. [Google Scholar] [CrossRef] [Green Version]
  275. Clatot, J.; Hoshi, M.; Wan, X.; Liu, H.; Jain, A.; Shinlapawittayatorn, K.; Marionneau, C.; Ficker, E.; Ha, T.; Deschênes, I. Voltage-gated sodium channels assemble and gate as dimers. Nat. Commun. 2017, 8, 2077. [Google Scholar] [CrossRef] [Green Version]
  276. Allouis, M.; Le Bouffant, F.; Wilders, R.; Péroz, D.; Schott, J.J.; Noireaud, J.; Le Marec, H.; Mérot, J.; Escande, D.; Baró, I. 14-3-3 Is a regulator of the cardiac voltage-gated sodium channel Nav1.5. Circ. Res. 2006, 98, 1538–1546. [Google Scholar] [CrossRef] [Green Version]
  277. Rougier, J.S.; Essers, M.C.; Gillet, L.; Guichard, S.; Sonntag, S.; Shmerling, D.; Abriel, H. A Distinct Pool of Nav1.5 Channels at the Lateral Membrane of Murine Ventricular Cardiomyocytes. Front. Physiol. 2019, 10, 1–20. [Google Scholar] [CrossRef] [Green Version]
  278. Gee, S.H.; Madhavan, R.; Levinson, S.R.; Caldwell, J.H.; Sealock, R.; Froehner, S.C. Interaction of muscle and brain sodium channels with multiple members of the syntrophin family of dystrophin-associated proteins. J. Neurosci. 1998, 18, 128–137. [Google Scholar] [CrossRef]
  279. Ou, Y.; Strege, P.; Miller, S.M.; Makielski, J.; Ackerman, M.; Gibbons, S.J.; Farrugia, G. Syntrophin γ2 regulates SCN5A gating by a PDZ domain-mediated interaction. J. Biol. Chem. 2003, 278, 1915–1923. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  280. Schultz, J.; Hoffmuller, U.; Krause, G.; Ashurst, J.; Macias, M.J.; Schmieder, P.; Schneider-Mergener, J.; Oschkinat, H. Specific interactions between the syntrophin PDZ domain and voltagegated sodium channels. Nat. Struct. Biol. 1998, 5, 19–24. [Google Scholar] [CrossRef] [PubMed]
  281. Matamoros, M.; Perez-Hernandez, M.; Guerrero-Serna, G.; Amoros, I.; Barana, A.; Nunez, M.; Ponce-Balbuena, D.; Sacristan, S.; Gomez, R.; Tamargo, J.; et al. Nav1.5 N-terminal domain binding to alpha1 syntrophin incraeses membrane density of human Kir2.1 Kir2.2 and Nav1.5. Cardiovasc. Res. 2016, 110, 279–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  282. Pérez-hernández, M.; Caballero, R.; Delpón, E.; Pérez-hernández, M.; Matamoros, M.; Alfayate, S.; Nieto-marín, P.; Utrilla, R.G.; Tinaquero, D.; De Andrés, R.; et al. channels can trap cardiac Kir2.1/2.2 channels. JCI Insights 2018, 3, e96291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  283. Eichel, C.A.; Beuriot, A.; Chevalier, M.Y.E.; Rougier, J.S.; Louault, F.; Dilanian, G.; Amour, J.; Coulombe, A.; Abriel, H.; Hatem, S.N.; et al. Lateral Membrane-Specific MAGUK CASK Down-Regulates Nav1.5 Channel in Cardiac Myocytes. Circ. Res. 2016, 119, 544–556. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  284. Ziane, R.; Huang, H.; Moghadaszadeh, B.; Beggs, A.H.; Levesque, G.; Chahine, M. Cell membrane expression of cardiac sodium channel Nav1.5 is modulated by α-actinin-2 interaction. Biochemistry 2010, 49, 166–178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  285. Mazzone, A.; Strege, P.R.; Tester, D.J.; Bernard, C.E.; Faulkner, G.; De Giorgio, R.; Makielski, J.C.; Stanghellini, V.; Gibbons, S.J.; Ackerman, M.J.; et al. A Mutation in Telethonin Alters Nav1.5 Function. J. Biol. Chem. 2008, 283, 16537–16544. [Google Scholar] [CrossRef] [Green Version]
  286. Turker, I.; Makiyama, T.; Ueyama, T.; Shimizu, A.; Yamakawa, M.; Chen, P.S.; Vatta, M.; Horie, M.; Ai, T. Telethonin Variants Found in Brugada Syndrome, J-Wave Pattern ECG, and ARVC Reduce Peak Nav1.5 Currents in HEK-293 Cells. PACE—Pacing Clin. Electrophysiol. 2020, 43, 838–846. [Google Scholar] [CrossRef]
  287. Goldfarb, M. Fibroblast growth factor homologous factors evolution, structure, and function. Cytokine Growth Factor Rev. 2005, 16, 215–220. [Google Scholar] [CrossRef] [Green Version]
  288. Rush, A.M.; Wittmack, E.K.; Tyrrell, L.; Black, J.A.; Dib-hajj, S.D.; Waxman, S.G. Differential modulation of sodium channel Nav1.6 by two members of the fibroblast growth factor homologous factor 2 subfamily. Eur. J. Neurosci. 2006, 23, 2551–2562. [Google Scholar] [CrossRef]
  289. Liu, C.; Dib-Hajj, S.D.; Renganathan, M.; Cummins, T.R.; Waxman, S.G. Modulation of the cardiac sodium channel Nav1.5 by fibroblast growth factor homologous factor 1B. J. Biol. Chem. 2003, 278, 1029–1036. [Google Scholar] [CrossRef] [Green Version]
  290. Wang, C.; Hennessey, J.A.; Kirkton, R.D.; Wang, C.; Graham, V.; Puranam, R.S.; Rosenberg, P.B.; Bursac, N.; Pitt, G.S. Fibroblast Growth Factor Homologous Factor 13 Regulates Na+ Channels and Conduction Velocity in Murine Hearts. Circ. Res. 2011, 109, 775–782. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  291. Wang, C.; Wang, C.; Hoch, E.G.; Pitt, G.S. Identification of Novel Interaction Sites that Determine Specificity between Fibroblast Growth Factor Homologous Factors and Voltage-gated Sodium Channels. J. Biol. Chem. 2011, 286, 24253–24263. [Google Scholar] [CrossRef] [Green Version]
  292. Yang, J.; Wang, Z.; Sinden, D.S.; Wang, X.; Shan, B.; Yu, X.; Zhang, H.; Pitt, G.S.; Wang, C. FGF13 modulates the gating properties of the cardiac sodium channel Nav1.5 in an isoform-specific maner. Channels 2016, 10, 410–420. [Google Scholar] [CrossRef] [Green Version]
  293. Lou, J.Y.; Laezza, F.; Gerber, B.R.; Xiao, M.; Yamada, K.A.; Hartmann, H.; Craig, A.M.; Nerbonne, J.M.; Ornitz, D.M. Fibroblast growth factor 14 is an intracellular modulator of voltage-gated sodium channels. J. Physiol. 2005, 569, 179–193. [Google Scholar] [CrossRef] [PubMed]
  294. Mohler, P.J.; Splawski, I.; Napolitano, C.; Bottelli, G.; Sharpe, L.; Timothy, K.; Priori, S.G.; Keating, M.T.; Bennett, V. A cardiac arrhythmia syndrome caused by loss of ankyrin-B function. Proc. Natl. Acad. Sci. USA 2004, 101, 9137–9142. [Google Scholar] [CrossRef] [Green Version]
  295. Wang, C.; Chung, B.C.; Yan, H.; Wang, H.G.; Lee, S.Y.; Pitt, G.S. Structural analyses of Ca2+/CaM interaction with NaV channel C-termini reveal mechanisms of calcium-dependent regulation. Nat. Commun. 2014, 5, 4896. [Google Scholar] [CrossRef] [Green Version]
  296. Urbauer, J.L. Direct Sodium Channel Regulation by Calmodulin. Structure 2018, 26, 677–678. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  297. Kang, P.W.; Chakouri, N.; Diaz, J.; Tomaselli, G.F.; Yue, D.T.; Ben-Johny, M. Elementary mechanisms of calmodulin regulation of Nav1.5 producing divergent arrhythmogenic phenotypes. Proc. Natl. Acad. Sci. USA 2021, 118, e2025085118. [Google Scholar] [CrossRef] [PubMed]
  298. Turnow, K.; Metzner, K.; Cotella, D.; Morales, M.J.; Schaefer, M.; Christ, T.; Ravens, U.; Wettwer, E.; Kämmerer, S. Interaction of DPP10a with Kv4.3 channel complex results in a sustained current component of human transient outward current Ito. Basic Res. Cardiol. 2015, 110, 5. [Google Scholar] [CrossRef]
  299. Belau, F.; Metzner, K.; Christ, T.; Ravens, U.; Schaefer, M.; Künzel, S.; Li, W.; Wettwer, E.; Dobrev, D.; El-Armouche, A.; et al. DPP10 is a new regulator of Nav1.5 channels in human heart. Int. J. Cardiol. 2019, 284, 68–73. [Google Scholar] [CrossRef]
  300. Yarbrough, T.L.; Lu, T.; Lee, H.-C.; Shibata, E.F. Localization of Cardiac Sodium Channels in Caveolin-Rich Membrane Domains. Circ. Res. 2002, 90, 443–449. [Google Scholar] [CrossRef] [Green Version]
  301. Balijepalli, R.C.; Kamp, T.J. Caveolae, Ion Channels and Cardiac Arrhythmias. Prog. Biophys. Mol. Biol. 2008, 98, 149–160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  302. Xiao, Y.F.; Kang, J.X.; Morgan, J.P.; Leaf, A. Blocking effects of polyunsaturated fatty acids on Na+ channels of neonatal rat ventricular myocytes. Proc. Natl. Acad. Sci. USA 1995, 92, 11000–11004. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  303. Kang, J.X.; Xiao, Y.F.; Leaf, A. Free, long-chain, polyunsaturated fatty acids reduce membrane electrical excitability in neonatal rat cardiac myocytes. Proc. Natl. Acad. Sci. USA 1995, 92, 3997–4001. [Google Scholar] [CrossRef] [Green Version]
  304. Shibata, E.F.; Brown, T.L.Y.; Washburn, Z.W.; Bai, J.; Revak, T.J.; Butters, C.A. Autonomic Regulation of Voltage-Gated Cardiac Ion Channels. J. Cardiovasc. Electrophysiol. 2006, 17, 34–42. [Google Scholar] [CrossRef]
  305. Vatta, M.; Ackerman, M.J.; Ye, B.; Makielski, J.C.; Ughanze, E.E.; Taylor, E.W.; Tester, D.J.; Balijepalli, R.C.; Foell, J.D.; Li, Z.; et al. Mutant Caveolin-3 Induces Persistent Late Sodium Current and Is Associated With Long-QT Syndrome. Circulation 2006, 114, 2104–2112. [Google Scholar] [CrossRef] [Green Version]
  306. Vaidyanathan, R.; Reilly, L.; Eckhardt, L.L. Caveolin-3 Microdomain: Arrhythmia Implications for Potassium Inward Rectifier and Cardiac Sodium Channel. Front. Physiol. 2018, 9, 1–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  307. Ehrlich, J.R. Taking ion channel degradation to heart. Cardiovasc. Res. 2007, 74, 6–7. [Google Scholar] [CrossRef]
  308. Abriel, H.; Staub, O. Ubiquitylation of ion channels. Physiology 2005, 20, 398–407. [Google Scholar] [CrossRef] [PubMed]
  309. Liu, X.; Chen, Z.; Han, Z.; Liu, Y.; Wu, X.; Peng, Y.; Di, W.; Lan, R.; Sun, B.; Xu, B.; et al. AMPK-mediated degradation of Nav1.5 through autophagy. FASEB J. 2019, 33, 5366–5376. [Google Scholar] [CrossRef] [PubMed]
  310. Van Bemmelen, M.X.; Rougier, J.S.; Gavillet, B.; Apothéloz, F.; Daidié, D.; Tateyama, M.; Rivolta, I.; Thomas, M.A.; Kass, R.S.; Staub, O.; et al. Cardiac voltage-gated sodium channel Nav1.5 is regulated by Nedd4-2 mediated ubiquitination. Circ. Res. 2004, 95, 284–291. [Google Scholar] [CrossRef] [Green Version]
  311. Rougier, J.S.; Van Bemmelen, M.X.; Bruce, M.C.; Jespersen, T.; Gavillet, B.; Apothéloz, F.; Cordonier, S.; Staub, O.; Rotin, D.; Abriel, H. Molecular determinants of voltage-gated sodium channel regulation by the Nedd4/Nedd4-like proteins. Am. J. Physiol.—Cell Physiol. 2005, 288, 692–701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  312. Huang, Y.; Wang, Z.; Liu, Y.; Xiong, H.; Zhao, Y.; Wu, L.; Yuan, C.; Wang, L.; Hou, Y.; Yu, G.; et al. αB-crystallin interacts with Nav1.5 and regulates ubiquitination and internalization of cell surface Nav1.5. J. Biol. Chem. 2016, 291, 11030–11041. [Google Scholar] [CrossRef] [Green Version]
  313. Boehmer, C.; Wilhelm, V.; Palmada, M.; Wallisch, S.; Henke, G.; Brinkmeier, H.; Cohen, P.; Pieske, B.; Lang, F. Serum and glucocorticoid inducible kinases in the regulation of the cardiac sodium channel SCN5A. Cardiovasc. Res. 2003, 57, 1079–1084. [Google Scholar] [CrossRef] [Green Version]
  314. Tang, B.; Hu, Y.; Wang, Z.; Cheng, C.; Wang, P.; Liang, L.; Xiong, H.; Luo, C.; Xu, C.; Chen, Q.; et al. UBC9 regulates cardiac sodium channel Nav1.5 ubiquitination, degradation and sodium current density. J. Mol. Cell Cardiol. 2019, 129, 79–91. [Google Scholar] [CrossRef] [PubMed]
  315. Liu, X.; Ge, J.; Chen, C.; Shen, Y.; Xie, J.; Zhu, X.; Liu, M.; Hu, J.; Chen, L.; Guo, L.; et al. FAT10 protects against ischemia-induced ventricular arrhythmia by decreasing Nedd4-2_Nav1.5 complex formation. Cell Death Dis. 2021, 12, 1–12. [Google Scholar]
  316. Bhalla, V.; Oyster, N.M.; Fitch, A.C.; Wijngaarden, M.A.; Neumann, D.; Schlattner, U.; Pearce, D.; Hallows, K.R. AMP-activated kinase inhibits the epithelial Na+ channel through functional regulation of the ubiquitin ligase Nedd4-2. J. Biol. Chem. 2006, 281, 26159–26169. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  317. Tadros, R.; Ton, A.T.; Fiset, C.; Nattel, S. Sex differences in cardiac electrophysiology and clinical arrhythmias: Epidemiology, therapeutics, and mechanisms. Can. J. Cardiol. 2014, 30, 783–792. [Google Scholar] [CrossRef]
  318. Costa, S.; Saguner, A.M.; Gasperetti, A.; Akdis, D.; Brunckhorst, C.; Duru, F. The Link Between Sex Hormones and Susceptibility to Cardiac Arrhythmias: From Molecular Basis to Clinical Implications. Front. Cardiovasc. Med. 2021, 8, 85. [Google Scholar] [CrossRef]
  319. Gaborit, N.; Varro, A.; Le Bouter, S.; Szuts, V.; Escande, D.; Nattel, S.; Demolombe, S. Gender-related differences in ion-channel and transporter subunit expression in non-diseased human hearts. J. Mol. Cell. Cardiol. 2010, 49, 639–646. [Google Scholar] [CrossRef]
  320. Furukawa, T.; Kurokawa, J. Regulation of cardiac ion channels via non-genomic action of sex steroid hormones: Implication for the gender difference in cardiac arrhythmias. Pharmacol. Ther. 2007, 115, 106–115. [Google Scholar] [CrossRef]
  321. Clemens Moller, R.N. Effects of estradiol on cardiac ion channel currents. Eur. J. Pharmacol. 2006, 532, 44–49. [Google Scholar] [CrossRef]
  322. Yang, G.; Liu, J.; Wang, Y.; Du, Y.; Ma, A.; Wang, T. Lack of influence of sex hormones on Brugada syndrome-associated mutant Nav1.5 sodium channel. J. Electrocardiol. 2019, 52, 82–87. [Google Scholar] [CrossRef] [PubMed]
  323. Peto, K.; Miko, I.; Ba, T.; To, I.B. Effects of sex hormones on ECG parameters and expression of cardiac ion channels in dogs. Acta Physiol. 2006, 188, 163–171. [Google Scholar] [CrossRef]
  324. Hu, X.; Fu, L.; Zhao, M.; Wang, D.; Zhang, H.; Gong, Z.; Ma, T.; Zhang, Y.; Machuki, J.; Pan, X.; et al. Sex hormones ameliorated sodium channel dysfunction induced by β-adrenergic overstimulation: The role of estrogen and G protein-coupled estrogen receptor. Authorea Prepr. 2020, 1–12. [Google Scholar] [CrossRef]
  325. Mangold, K.E.; Brumback, B.D.; Angsutararux, P.; Voelker, T.L.; Zhu, W.; Kang, P.W.; Moreno, J.D.; Silva, J.R. Mechanisms and models of cardiac sodium channel inactivation. Channels 2017, 11, 517–533. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  326. Nassal, M.M.J.; Wan, X.; Dale, Z.; Deschênes, I.; Wilson, L.D.; Piktel, J.S. Mild hypothermia preserves myocardial conduction during ischemia by maintaining gap junction intracellular communication and Na+ channel function. Am. J. Physiol. Heart Circ. Physiol. 2021, 312, 886–895. [Google Scholar] [CrossRef] [Green Version]
  327. Porres, J.M.; Brugada, J.; Urbistondo, V.; García, F.; Reviejo, K.; Marco, P. Fever unmasking the Brugada syndrome. Pacing Clin. Electrophysiol. 2002, 25, 1646–1648. [Google Scholar] [CrossRef] [PubMed]
  328. Abdelsayed, M.; Peters, C.H.; Ruben, P.C. Differential thermosensitivity in mixed syndrome cardiac sodium channel mutants. J. Physiol. 2015, 593, 4201–4223. [Google Scholar] [CrossRef]
  329. 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. 1999, 85, 803–809. [Google Scholar] [CrossRef] [Green Version]
  330. Peters, C.H.; Abdelsayed, M.; Ruben, P.C. Triggers for arrhythmogenesis in the Brugada and long QT 3 syndromes. Prog. Biophys. Mol. Biol. 2016, 120, 77–88. [Google Scholar] [CrossRef]
  331. Kazutaka Gima and Yoram Rudy Ionic current basis of electrocardiographic waveforms: A model study. Circ. Res. 2007, 90, 889–896.
  332. Park, D.S.; Shekhar, A.; Marra, C.; Lin, X.; Vasquez, C.; Solinas, S.; Kelley, K.; Morley, G.; Goldfarb, M.; Fishman, G.I. Fhf2 gene deletion causes temperature-sensitive cardiac conduction failure. Nat. Commun. 2016, 7, 1–10. [Google Scholar] [CrossRef] [Green Version]
  333. Jones, D.K.; Peters, C.H.; Tolhurst, S.A.; Claydon, T.W.; Ruben, P.C. Extracellular proton modulation of the cardiac voltage-gated sodium channel, Nav1.5. Biophys. J. 2011, 101, 2147–2156. [Google Scholar] [CrossRef] [Green Version]
  334. Jones, D.K.; Claydon, T.W.; Ruben, P.C. Extracellular protons inhibit charge immobilization in the cardiac voltage-gated sodium channel. Biophys. J. 2013, 105, 101–107. [Google Scholar] [CrossRef] [Green Version]
  335. Jones, D.K.; Peters, C.H.; Allard, C.R.; Claydon, T.W.; Ruben, P.C. Proton Sensors in the Pore Domain of the Cardiac Voltage-gated Sodium Channel. J. Biol. Chem. 2013, 288, 4782–4791. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  336. 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. Mol. Genet. 1995, 4, 1603–1607. [Google Scholar] [CrossRef]
  337. Wang, D.W.; Yazawa, K.; George, A.L.; Bennett, P.B. Characterization of human cardiac Na+ channel mutations in the congenital long QT syndrome. Proc. Natl. Acad. Sci. USA 1996, 93, 13200–13205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  338. Wilde, A.A.M.; Amin, A.S. Clinical Spectrum of SCN5A Mutations. JACC Clin. Electrophysiol. 2018, 4, 569–579. [Google Scholar] [CrossRef]
  339. Tester, D.J.; Ackerman, M.J. Genetics of Long QT Syndrome. Methodist Debakey Cardiovasc. J. 2014, 10, 29. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  340. Pérez-Riera, A.R.; Barbosa-Barros, R.; Daminello Raimundo, R.; da Costa de Rezende Barbosa, M.P.; Esposito Sorpreso, I.C.; de Abreu, L.C. The congenital long QT syndrome Type 3: An update. Indian Pacing Electrophysiol. J. 2018, 18, 25–35. [Google Scholar] [CrossRef]
  341. Remme, C.A. Cardiac sodium channelopathy associated with SCN5A mutations: Electrophysiological, molecular and genetic aspects. J. Physiol. 2013, 591, 4099–4116. [Google Scholar] [CrossRef]
  342. Riuró, H.; Campuzano, O.; Arbelo, E.; Iglesias, A.; Batlle, M.; Pérez-Villa, F.; Brugada, J.; Pérez, G.J.; Scornik, F.S.; Brugada, R. A missense mutation in the sodium channel β1b subunit reveals SCN1B as a susceptibility gene underlying long QT syndrome. Heart Rhythm 2014, 11, 1202–1209. [Google Scholar] [CrossRef]
  343. Adler, A.; Novelli, V.; Amin, A.S.; Abiusi, E.; Care, M.; Nannenberg, E.A.; Feilotter, H.; Amenta, S.; Mazza, D.; Bikker, H.; et al. An International, Multicentered, Evidence-Based Reappraisal of Genes Reported to Cause Congenital Long QT Syndrome. Circulation 2020, 141, 418–428. [Google Scholar] [CrossRef]
  344. Choi, J.-I.; Wang, C.; Thomas, M.J.; Pitt, G.S. α1-Syntrophin Variant Identified in Drug-Induced Long QT Syndrome Increases Late Sodium Current. PLoS ONE 2016, 11, e0152355. [Google Scholar] [CrossRef]
  345. Ueda, K.; Valdivia, C.; Medeiros-Domingo, A.; Tester, D.J.; Vatta, M.; Farrugia, G.; Ackerman, M.J.; Makielski, J.C. Syntrophin mutation associated with long QT syndrome through activation of the nNOS-SCN5A macromolecular complex. Proc. Natl. Acad. Sci. USA 2008, 105, 9355–9360. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  346. Wu, G.; Ai, T.; Kim, J.J.; Mohapatra, B.; Xi, Y.; Li, Z.; Abbasi, S.; Purevjav, E.; Samani, K.; Ackerman, M.J.; et al. α-1-Syntrophin Mutation and the Long-QT Syndrome. Circ. Arrhythmia Electrophysiol. 2008, 1, 193–201. [Google Scholar] [CrossRef] [Green Version]
  347. Hu, R.-M.; Tan, B.-H.; Orland, K.M.; Valdivia, C.R.; Peterson, A.; Pu, J.; Makielski, J.C. Digenic inheritance novel mutations in SCN5A and SNTA1 increase late INa contributing to LQT syndrome. Am. J. Physiol. Circ. Physiol. 2013, 304, H994–H1001. [Google Scholar] [CrossRef] [Green Version]
  348. Kostera-Pruszczyk, A.; Suszek, M.; Płoski, R.; Franaszczyk, M.; Potulska-Chromik, A.; Pruszczyk, P.; Sadurska, E.; Karolczak, J.; Kamińska, A.M.; Rędowicz, M.J. BAG3-related myopathy, polyneuropathy and cardiomyopathy with long QT syndrome. J. Muscle Res. Cell Motil. 2015, 36, 423–432. [Google Scholar] [CrossRef] [Green Version]
  349. Prondzynski, M.; Lemoine, M.D.; Zech, A.T.; Horváth, A.; Di Mauro, V.; Koivumäki, J.T.; Kresin, N.; Busch, J.; Krause, T.; Krämer, E.; et al. Disease modeling of a mutation in α-actinin 2 guides clinical therapy in hypertrophic cardiomyopathy. EMBO Mol. Med. 2019, 11, e11115. [Google Scholar] [CrossRef]
  350. Markunas, A.M.; Manivannan, P.K.R.; Ezekian, J.E.; Agarwal, A.; Eisner, W.; Alsina, K.; Allen, H.D.; Wray, G.A.; Kim, J.J.; Wehrens, X.H.T.; et al. TBX5-encoded T-box transcription factor 5 variant T223M is associated with long QT syndrome and pediatric sudden cardiac death. Am. J. Med. Genet. Part A 2021, 185, 923–929. [Google Scholar] [CrossRef] [PubMed]
  351. Nieto-Marín, P.; Tinaquero, D.; Utrilla, R.G.; Cebrián, J.; González-Guerra, A.; Crespo-García, T.; Cámara-Checa, A.; Rubio-Alarcón, M.; Dago, M.; Alfayate, S.; et al. Tbx5 variants disrupt Nav1.5 function differently in patients diagnosed with Brugada or Long QT Syndrome. Cardiovasc. Res. 2021, cvab045. [Google Scholar] [CrossRef]
  352. 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.; et al. Spectrum of Mutations in Long-QT Syndrome Genes. Circulation 2000, 102, 1178–1185. [Google Scholar] [CrossRef] [Green Version]
  353. Chung, S.-K.; MacCormick, J.M.; McCulley, C.H.; Crawford, J.; Eddy, C.-A.; Mitchell, E.A.; Shelling, A.N.; French, J.K.; Skinner, J.R.; Rees, M.I. Long QT and Brugada syndrome gene mutations in New Zealand. Heart Rhythm 2007, 4, 1306–1314. [Google Scholar] [CrossRef]
  354. Chang, R.-K.R.; Lan, Y.-T.; Silka, M.J.; Morrow, H.; Kwong, A.; Smith-Lang, J.; Wallerstein, R.; Lin, H.J. Genetic Variants for Long QT Syndrome among Infants and Children from a Statewide Newborn Hearing Screening Program Cohort. J. Pediatr. 2014, 164, 590–595.e3. [Google Scholar] [CrossRef] [Green Version]
  355. Murphy, L.L.; Moon-Grady, A.J.; Cuneo, B.F.; Wakai, R.T.; Yu, S.; Kunic, J.D.; Benson, D.W.; George, A.L. Developmentally regulated SCN5A splice variant potentiates dysfunction of a novel mutation associated with severe fetal arrhythmia. Heart Rhythm 2012, 9, 590–597. [Google Scholar] [CrossRef] [Green Version]
  356. Hu, D.; Barajas-Martínez, H.; Medeiros-Domingo, A.; Crotti, L.; Veltmann, C.; Schimpf, R.; Urrutia, J.; Alday, A.; Casis, O.; Pfeiffer, R.; et al. A novel rare variant in SCN1Bb linked to Brugada syndrome and SIDS by combined modulation of Nav1.5 and Kv4.3 channel currents. Heart Rhythm 2012, 9, 760–769. [Google Scholar] [CrossRef] [Green Version]
  357. Rossenbacker, T. Unconventional intronic splice site mutation in SCN5A associates with cardiac sodium channelopathy. J. Med. Genet. 2005, 42, e29. [Google Scholar] [CrossRef] [Green Version]
  358. Casini, S.; Albesa, M.; Wang, Z.; Portero, V.; Ross-Kaschitza, D.; Rougier, J.-S.; Marchal, G.A.; Chung, W.K.; Bezzina, C.R.; Abriel, H.; et al. Functional Consequences of the SCN5A-p.Y1977N Mutation within the PY Ubiquitylation Motif: Discrepancy between HEK293 Cells and Transgenic Mice. Int. J. Mol. Sci. 2019, 20, 5033. [Google Scholar] [CrossRef] [Green Version]
  359. Wang, L.; Han, Z.; Dai, J.; Cao, K. Brugada Syndrome Caused by Sodium Channel Dysfunction Associated with a SCN1B Variant A197V. Arch. Med. Res. 2020, 51, 245–253. [Google Scholar] [CrossRef]
  360. Liu, L.; Hayashi, K.; Kaneda, T.; Ino, H.; Fujino, N.; Uchiyama, K.; Konno, T.; Tsuda, T.; Kawashiri, M.; Ueda, K.; et al. A novel mutation in the transmembrane nonpore region of the KCNH2 gene causes severe clinical manifestations of long QT syndrome. Heart Rhythm 2013, 10, 61–67. [Google Scholar] [CrossRef]
  361. 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 Residue of the Cardiac Sodium Channel Confer Distinct Channel and Clinical Phenotypes. J. Biol. Chem. 2001, 276, 30623–30630. [Google Scholar] [CrossRef] [Green Version]
  362. Liu, H.; Tateyama, M.; Clancy, C.E.; Abriel, H.; Kass, R.S. Channel Openings Are Necessary but not Sufficient for Use-dependent Block of Cardiac Na+ Channels by Flecainide. J. Gen. Physiol. 2002, 120, 39–51. [Google Scholar] [CrossRef] [Green Version]
  363. Davis, R.P.; Casini, S.; van den Berg, C.W.; Hoekstra, M.; Remme, C.A.; Dambrot, C.; Salvatori, D.; Oostwaard, D.W.; Wilde, A.A.M.; Bezzina, C.R.; et al. Cardiomyocytes Derived From Pluripotent Stem Cells Recapitulate Electrophysiological Characteristics of an Overlap Syndrome of Cardiac Sodium Channel Disease. Circulation 2012, 125, 3079–3091. [Google Scholar] [CrossRef]
  364. Zhang, J.; Chen, Y.; Yang, J.; Xu, B.; Wen, Y.; Xiang, G.; Wei, G.; Zhu, C.; Xing, Y.; Li, Y. Electrophysiological and trafficking defects of the SCN5A T353I mutation in Brugada syndrome are rescued by alpha-allocryptopine. Eur. J. Pharmacol. 2015, 746, 333–343. [Google Scholar] [CrossRef] [PubMed]
  365. Ortiz-Bonnin, B.; Rinné, S.; Moss, R.; Streit, A.K.; Scharf, M.; Richter, K.; Stöber, A.; Pfeufer, A.; Seemann, G.; Kääb, S.; et al. Electrophysiological characterization of a large set of novel variants in the SCN5A-gene: Identification of novel LQTS3 and BrS mutations. Pflügers Arch.—Eur. J. Physiol. 2016, 468, 1375–1387. [Google Scholar] [CrossRef]
  366. El-Battrawy, I.; Müller, J.; Zhao, Z.; Cyganek, L.; Zhong, R.; Zhang, F.; Kleinsorge, M.; Lan, H.; Li, X.; Xu, Q.; et al. Studying Brugada Syndrome With an SCN1B Variants in Human-Induced Pluripotent Stem Cell-Derived Cardiomyocytes. Front. Cell Dev. Biol. 2019, 7, 261. [Google Scholar] [CrossRef] [PubMed]
  367. Gray, B.; Hasdemir, C.; Ingles, J.; Aiba, T.; Makita, N.; Probst, V.; Wilde, A.A.M.; Newbury-Ecob, R.; Sheppard, M.N.; Semsarian, C.; et al. Lack of genotype-phenotype correlation in Brugada Syndrome and Sudden Arrhythmic Death Syndrome families with reported pathogenic SCN1B variants. Heart Rhythm 2018, 15, 1051–1057. [Google Scholar] [CrossRef] [Green Version]
  368. Ricci, M.T.; Menegon, S.; Vatrano, S.; Mandrile, G.; Cerrato, N.; Carvalho, P.; De Marchi, M.; Gaita, F.; Giustetto, C.; Giachino, D.F. SCN1B gene variants in Brugada Syndrome: A study of 145 SCN5A-negative patients. Sci. Rep. 2015, 4, 6470. [Google Scholar] [CrossRef]
  369. Ogawa, R.; Kishi, R.; Takagi, A.; Sakaue, I.; Takahashi, H.; Matsumoto, N.; Masuhara, K.; Nakazawa, K.; Kobayashi, S.; Miyake, F.; et al. A novel microsatellite polymorphism of sodium channel beta1-subunit gene (SCN1B) may underlie abnormal cardiac excitation manifested by coved-type ST-elevation compatible with Brugada syndrome in Japanese. Int. J. Clin. Pharmacol. Ther. 2010, 48, 109–119. [Google Scholar] [CrossRef]
  370. Olesen, M.S.; Holst, A.G.; Svendsen, J.H.; Haunsø, S.; Tfelt-Hansen, J. SCN1Bb R214Q found in 3 patients: 1 with Brugada syndrome and 2 with lone atrial fibrillation. Heart Rhythm 2012, 9, 770–773. [Google Scholar] [CrossRef]
  371. Peters, S. Arrhythmogenic cardiomyopathy and provocable Brugada ECG in a patient caused by missense mutation in plakophilin-2. Int. J. Cardiol. 2014, 173, 317–318. [Google Scholar] [CrossRef]
  372. Forkmann, M.; Tomala, J.; Huo, Y.; Mayer, J.; Christoph, M.; Wunderlich, C.; Salmas, J.; Gaspar, T.; Piorkowski, C. Epicardial Ventricular Tachycardia Ablation in a Patient With Brugada ECG Pattern and Mutation of PKP2 and DSP Genes. Circ. Arrhythmia Electrophysiol. 2015, 8, 505–507. [Google Scholar] [CrossRef] [Green Version]
  373. Kattygnarath, D.; Maugenre, S.; Neyroud, N.; Balse, E.; Ichai, C.; Denjoy, I.; Dilanian, G.; Martins, R.P.; Fressart, V.; Berthet, M.; et al. MOG1. Circ. Cardiovasc. Genet. 2011, 4, 261–268. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  374. Zhou, J.; Wang, L.; Zuo, M.; Wang, X.; Ahmed, A.S.I.; Chen, Q.; Wang, Q.K. Cardiac sodium channel regulator MOG1 regulates cardiac morphogenesis and rhythm. Sci. Rep. 2016, 6, 21538. [Google Scholar] [CrossRef] [Green Version]
  375. Olesen, M.S.; Jensen, N.F.; Holst, A.G.; Nielsen, J.B.; Tfelt-Hansen, J.; Jespersen, T.; Sajadieh, A.; Haunsø, S.; Lund, J.T.; Calloe, K.; et al. A Novel Nonsense Variant in Nav1.5 Cofactor MOG1 Eliminates Its Sodium Current Increasing Effect and May Increase the Risk of Arrhythmias. Can. J. Cardiol. 2011, 27, 523.e17–523.e23. [Google Scholar] [CrossRef] [PubMed]
  376. Park, D.S.; Shekhar, A.; Santucci, J.; Redel-Traub, G.; Solinas, S.; Mintz, S.; Lin, X.; Chang, E.W.; Narke, D.; Xia, Y.; et al. Ionic Mechanisms of Impulse Propagation Failure in the FHF2-Deficient Heart. Circ. Res. 2020, 127, 1536–1548. [Google Scholar] [CrossRef] [PubMed]
  377. Schmidt, C.; Wiedmann, F.; El-Battrawy, I.; Fritz, M.; Ratte, A.; Beller, C.J.; Lang, S.; Rudic, B.; Schimpf, R.; Akin, I.; et al. Reduced Na+ Current in Native Cardiomyocytes of a Brugada Syndrome Patient Associated With β-2-Syntrophin Mutation. Circ. Genom. Precis. Med. 2018, 11, e002263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  378. Shimizu, A.; Zankov, D.P.; Sato, A.; Komeno, M.; Toyoda, F.; Yamazaki, S.; Makita, T.; Noda, T.; Ikawa, M.; Asano, Y.; et al. Identification of transmembrane protein 168 mutation in familial Brugada syndrome. FASEB J. 2020, 34, 6399–6417. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  379. Nguyen, L.K.C.; Shimizu, A.; Soh, J.E.C.; Komeno, M.; Sato, A.; Ogita, H. Transmembrane protein 168 mutation reduces cardiomyocyte cell surface expression of Nav1.5 through αB-crystallin intracellular dynamics. J. Biochem. 2021, 170, mvab066. [Google Scholar] [CrossRef]
  380. Di Resta, C.; Pietrelli, A.; Sala, S.; Della Bella, P.; De Bellis, G.; Ferrari, M.; Bordoni, R.; Benedetti, S. High-throughput genetic characterization of a cohort of Brugada syndrome patients. Hum. Mol. Genet. 2015, 24, 5828–5835. [Google Scholar] [CrossRef] [Green Version]
  381. Musa, H.; Marcou, C.A.; Herron, T.J.; Makara, M.A.; Tester, D.J.; O’Connell, R.P.; Rosinski, B.; Guerrero-Serna, G.; Milstein, M.L.; Monteiro da Rocha, A.; et al. Abnormal myocardial expression of SAP97 is associated with arrhythmogenic risk. Am. J. Physiol. Circ. Physiol. 2020, 318, H1357–H1370. [Google Scholar] [CrossRef]
  382. Wang, Z.; Vermij, S.H.; Sottas, V.; Shestak, A.; Ross-Kaschitza, D.; Zaklyazminskaya, E.V.; Hudmon, A.; Pitt, G.S.; Rougier, J.S.; Abriel, H. Calmodulin binds to the N-terminal domain of the cardiac sodium channel Nav1.5. Channels 2020, 14, 268–286. [Google Scholar] [CrossRef]
  383. Jiménez-Jáimez, J.; Palomino Doza, J.; Ortega, Á.; Macías-Ruiz, R.; Perin, F.; Rodríguez-Vázquez del Rey, M.M.; Ortiz-Genga, M.; Monserrat, L.; Barriales-Villa, R.; Blanca, E.; et al. Calmodulin 2 Mutation N98S Is Associated with Unexplained Cardiac Arrest in Infants Due to Low Clinical Penetrance Electrical Disorders. PLoS ONE 2016, 11, e0153851. [Google Scholar] [CrossRef] [PubMed]
  384. Yagihara, N.; Watanabe, H.; Barnett, P.; Duboscq-Bidot, L.; Thomas, A.C.; Yang, P.; Ohno, S.; Hasegawa, K.; Kuwano, R.; Chatel, S.; et al. Variants in the SCN5A Promoter Associated with Various Arrhythmia Phenotypes. J. Am. Heart Assoc. 2016, 5, 1–10. [Google Scholar] [CrossRef] [Green Version]
  385. Kimura, Y.; Aiba, T.; Sasano, T.; Furukawa, T.; Kusano, K.; Shimizu, W. IRX3 variant as a modifier of Brugada syndrome with frequent ventricular fibrillation. Heart Case Rep. 2016, 2, 465–468. [Google Scholar] [CrossRef] [Green Version]
  386. Yang, P.; Koopmann, T.T.; Pfeufer, A.; Jalilzadeh, S.; Schulze-Bahr, E.; Kääb, S.; Wilde, A.A.; Roden, D.M.; Bezzina, C.R. Polymorphisms in the cardiac sodium channel promoter displaying variant in vitro expression activity. Eur. J. Hum. Genet. 2008, 16, 350–357. [Google Scholar] [CrossRef] [Green Version]
  387. Matsumura, H.; Nakano, Y.; Ochi, H.; Onohara, Y.; Sairaku, A.; Tokuyama, T.; Tomomori, S.; Motoda, C.; Amioka, M.; Hironobe, N.; et al. H558R, a common SCN5A polymorphism, modifies the clinical phenotype of Brugada syndrome by modulating DNA methylation of SCN5A promoters. J. Biomed. Sci. 2017, 24, 91. [Google Scholar] [CrossRef]
  388. Scumaci, D.; Oliva, A.; Concolino, A.; Curcio, A.; Fiumara, C.V.; Tammè, L.; Campuzano, O.; Pascali, V.L.; Coll, M.; Iglesias, A.; et al. Integration of “Omics” Strategies for Biomarkers Discovery and for the Elucidation of Molecular Mechanisms Underlying Brugada Syndrome. Proteom.—Clin. Appl. 2018, 12, 1800065. [Google Scholar] [CrossRef]
  389. Aiba, T.; Farinelli, F.; Kostecki, G.; Hesketh, G.G.; Edwards, D.; Biswas, S.; Tung, L.; Tomaselli, G.F. A Mutation Causing Brugada Syndrome Identifies a Mechanism for Altered Autonomic and Oxidant Regulation of Cardiac Sodium Currents. Circ. Cardiovasc. Genet. 2014, 7, 249–256. [Google Scholar] [CrossRef] [Green Version]
  390. Darbar, D.; Kannankeril, P.J.; Donahue, B.S.; Kucera, G.; Stubblefield, T.; Haines, J.L.; George, A.L.J.; Roden, D.M. Cardiac sodium channel (SCN5A) variants associated with atrial fibrillation. Circulation 2008, 117, 1927–1935. [Google Scholar] [CrossRef] [Green Version]
  391. Takehara, N.; Makita, N.; Kawabe, J.; Sato, N.; Kawamura, Y.; Kitabatake, A.; Kikuchi, K. A cardiac sodium channel mutation identified in Brugada syndrome associated with atrial standstill. J. Intern. Med. 2004, 255, 137–142. [Google Scholar] [CrossRef] [Green Version]
  392. 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. 2006, 17, 480–485. [Google Scholar] [CrossRef]
  393. Benito, B.; Brugada, R.; Perich, R.M.; Lizotte, E.; Cinca, J.; Mont, L.; Berruezo, A.; Tolosana, J.M.; Freixa, X.; Brugada, P.; et al. A mutation in the sodium channel is responsible for the association of long QT syndrome and familial atrial fibrillation. Heart Rhythm 2008, 5, 1434–1440. [Google Scholar] [CrossRef]
  394. Rossenbacker, T.; Carroll, S.J.; Liu, H.; Kuipéri, C.; de Ravel, T.J.L.; Devriendt, K.; Carmeliet, P.; Kass, R.S.; Heidbüchel, H. Novel pore mutation in SCN5A manifests as a spectrum of phenotypes ranging from atrial flutter, conduction disease, and Brugada syndrome to sudden cardiac death. Heart Rhythm 2004, 1, 610–615. [Google Scholar] [CrossRef]
  395. Watanabe, H.; Darbar, D.; Kaiser, D.W.; Jiramongkolchai, K.; Chopra, S.; Donahue, B.S.; Kannankeril, P.J.; Roden, D.M. Mutations in sodium channel β1- and β2-subunits associated with atrial fibrillation. Circ. Arrhythmia Electrophysiol. 2009, 2, 268–275. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  396. Xiong, H.; Yang, Q.; Zhang, X.; Wang, P.; Chen, F.; Liu, Y.; Wang, P.; Zhao, Y.; Li, S.; Huang, Y.; et al. Significant association of rare variant p.Gly8Ser in cardiac sodium channel β4-subunit SCN4B with atrial fibrillation. Ann. Hum. Genet. 2019, 83, 239–248. [Google Scholar] [CrossRef] [PubMed]
  397. Li, R.-G.; Wang, Q.; Xu, Y.-J.; Zhang, M.; Qu, X.-K.; Liu, X.; Fang, W.-Y.; Yang, Y.-Q. Mutations of the SCN4B-encoded sodium channel β4 subunit in familial atrial fibrillation. Int. J. Mol. Med. 2013, 32, 144–150. [Google Scholar] [CrossRef] [Green Version]
  398. Wang, P.; Yang, Q.; Wu, X.; Yang, Y.; Shi, L.; Wang, C.; Wu, G.; Xia, Y.; Yang, B.; Zhang, R.; et al. Functional dominant-negative mutation of sodium channel subunit gene SCN3B associated with atrial fibrillation in a Chinese GeneID population. Biochem. Biophys. Res. Commun. 2010, 398, 98–104. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  399. Husser, D.; Ueberham, L.; Hindricks, G.; Büttner, P.; Ingram, C.; Weeke, P.; Shoemaker, M.B.; Adams, V.; Arya, A.; Sommer, P.; et al. Rare variants in genes encoding the cardiac sodium channel and associated compounds and their impact on outcome of catheter ablation of atrial fibrillation. PLoS ONE 2017, 12, e0183690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  400. Cunha, S.R.; Hund, T.J.; Hashemi, S.; Voigt, N.; Li, N.; Wright, P.; Koval, O.; Li, J.; Gudmundsson, H.; Gumina, R.J.; et al. Defects in Ankyrin-Based Membrane Protein Targeting Pathways Underlie Atrial Fibrillation. Circulation 2011, 124, 1212–1222. [Google Scholar] [CrossRef] [Green Version]
  401. Robaei, D.; Ford, T.; Ooi, S.-Y. Ankyrin-B Syndrome: A Case of Sinus Node Dysfunction, Atrial Fibrillation and Prolonged QT in a Young Adult. Heart Lung Circ. 2015, 24, e31–e34. [Google Scholar] [CrossRef]
  402. Girolami, F.; Iascone, M.; Tomberli, B.; Bardi, S.; Benelli, M.; Marseglia, G.; Pescucci, C.; Pezzoli, L.; Sana, M.E.; Basso, C.; et al. Novel α-Actinin 2 Variant Associated With Familial Hypertrophic Cardiomyopathy and Juvenile Atrial Arrhythmias. Circ. Cardiovasc. Genet. 2014, 7, 741–750. [Google Scholar] [CrossRef]
  403. LIU, Y.; NI, B.; LIN, Y.; CHEN, X.-G.; CHEN, M.; HU, Z.; ZHANG, F. The rs3807989 G/A Polymorphism in CAV1 is Associated with the Risk of Atrial Fibrillation in Chinese Han Populations. Pacing Clin. Electrophysiol. 2015, 38, 164–170. [Google Scholar] [CrossRef]
  404. Watanabe, H.; Nogami, A.; Ohkubo, K.; Kawata, H.; Hayashi, Y.; Ishikawa, T.; Makiyama, T.; Nagao, S.; Yagihara, N.; Takehara, N.; et al. Electrocardiographic Characteristics and SCN5A Mutations in Idiopathic Ventricular Fibrillation Associated with Early Repolarization. Circ. Arrhythmia Electrophysiol. 2011, 4, 874–881. [Google Scholar] [CrossRef] [Green Version]
  405. Watanabe, H.; Ohkubo, K.; Watanabe, I.; Matsuyama, T.; Ishibashi-Ueda, H.; Yagihara, N.; Shimizu, W.; Horie, M.; Minamino, T.; Makita, N. SCN5A mutation associated with ventricular fibrillation, early repolarization, and concealed myocardial abnormalities. Int. J. Cardiol. 2013, 165, e21–e23. [Google Scholar] [CrossRef]
  406. Reithmann, C.; Beckmann, B.-M.; Kääb, S. Purkinje-related ventricular fibrillation associated with a homozygous H558R polymorphism in the sodium channel SCN5A gene. Europace 2016, 18, 896. [Google Scholar] [CrossRef] [PubMed]
  407. Vilin, Y.; Fujimoto, E.; Ruben, P. A novel mechanism associated with idiopathic ventricular fibrillation (IVF) mutations R1232W and T1620M in human cardiac sodium channels. Pflügers Arch. Eur. J. Physiol. 2001, 442, 204–211. [Google Scholar] [CrossRef] [PubMed]
  408. 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. 2000, 479, 29–34. [Google Scholar] [CrossRef] [Green Version]
  409. Wan, X.; Chen, S.; Sadeghpour, A.; Wang, Q.; Kirsch, G.E. Accelerated inactivation in a mutant Na+ channel associated with idiopathic ventricular fibrillation. Am. J. Physiol.—Heart Circ. Physiol. 2001, 280, 354–360. [Google Scholar] [CrossRef] [PubMed]
  410. Raharjo, S.B.; Maulana, R.; Maghfirah, I.; Alzahra, F.; Putrinarita, A.D.; Hanafy, D.A.; Yuniadi, Y. SCN 5A gene mutations and the risk of ventricular fibrillation and syncope in Brugada syndrome patients: A meta-analysis. J. Arrhythmia 2018, 34, 473–477. [Google Scholar] [CrossRef]
  411. Nishii, N.; Ogawa, M.; Morita, H.; Nakamura, K.; Banba, K.; Miura, D.; Kumagai, N.; Matsunaga, A.; Kawamura, H.; Urakawa, S.; et al. SCN5A Mutation Is Associated With Early and Frequent Recurrence of Ventricular Fibrillation in Patients With Brugada Syndrome. Circ. J. 2010, 74, 2572–2578. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  412. KELLER, D.; HUANG, H.; ZHAO, J.; FRANK, R.; SUAREZ, V.; DELACRETAZ, E.; BRINK, M.; OSSWALD, S.; SCHWICK, N.; CHAHINE, M. A novel SCN5A mutation, F1344S, identified in a patient with Brugada syndrome and fever-induced ventricular fibrillation. Cardiovasc. Res. 2006, 70, 521–529. [Google Scholar] [CrossRef] [Green Version]
  413. Kawamura, M.; Ozawa, T.; Yao, T.; Ashihara, T.; Sugimoto, Y.; Yagi, T.; Itoh, H.; Ito, M.; Makiyama, T.; Horie, M. Dynamic Change in ST-Segment and Spontaneous Occurrence of Ventricular Fibrillation in Brugada Syndrome With a Novel Nonsense Mutation in the SCN5A Gene During Long-Term Follow-up. Circ. J. 2009, 73, 584–588. [Google Scholar] [CrossRef] [Green Version]
  414. Lieve, K.V.; Verkerk, A.O.; Podliesna, S.; van der Werf, C.; Tanck, M.W.; Hofman, N.; van Bergen, P.F.; Beekman, L.; Bezzina, C.R.; Wilde, A.A.M.; et al. Gain-of-function mutation in SCN5A causes ventricular arrhythmias and early onset atrial fibrillation. Int. J. Cardiol. 2017, 236, 187–193. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  415. Boehringer, T.; Bugert, P.; Borggrefe, M.; Elmas, E. SCN5A mutations and polymorphisms in patients with ventricular fibrillation during acute myocardial infarction. Mol. Med. Rep. 2014, 10, 2039–2044. [Google Scholar] [CrossRef] [Green Version]
  416. Wang, F.; Liu, Y.; Liao, H.; Xue, Y.; Zhan, X.; Fang, X.; Liang, Y.; Wei, W.; Rao, F.; Zhang, Q.; et al. Genetic Variants on SCN5A, KCNQ1, and KCNH2 in Patients with Ventricular Arrhythmias during Acute Myocardial Infarction in a Chinese Population. Cardiology 2020, 145, 38–45. [Google Scholar] [CrossRef]
  417. Hu, D.; Viskin, S.; Oliva, A.; Carrier, T.; Cordeiro, J.M.; Barajas-Martinez, H.; Wu, Y.; Burashnikov, E.; Sicouri, S.; Brugada, R.; et al. Novel mutation in the SCN5A gene associated with arrhythmic storm development during acute myocardial infarction. Heart Rhythm 2007, 4, 1072–1080. [Google Scholar] [CrossRef] [Green Version]
  418. Ter Bekke, R.M.A.; Isaacs, A.; Barysenka, A.; Hoos, M.B.; Jongbloed, J.D.H.; Hoorntje, J.C.A.; Patelski, A.S.M.; Helderman-van den Enden, A.T.J.M.; van den Wijngaard, A.; Stoll, M.; et al. Heritability in a SCN5A -mutation founder population with increased female susceptibility to non-nocturnal ventricular tachyarrhythmia and sudden cardiac death. Heart Rhythm 2017, 14, 1873–1881. [Google Scholar] [CrossRef] [PubMed]
  419. Son, M.K.; Ki, C.-S.; Park, S.-J.; Huh, J.; Kim, J.S.; On, Y.K. Genetic Mutation in Korean Patients of Sudden Cardiac Arrest as a Surrogating Marker of Idiopathic Ventricular Arrhythmia. J. Korean Med. Sci. 2013, 28, 1021. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  420. Stawiarski, K.; Clarke, J.-R.D.; Pollack, A.; Winslow, R.; Majumdar, S. Ventricular fibrillation in Graves disease reveals a rare SCN5A mutation with W1191X variant associated with Brugada syndrome. Heart Case Rep. 2021, 7, 95–99. [Google Scholar] [CrossRef]
  421. Valdivia, C.R.; Medeiros-Domingo, A.; Ye, B.; Shen, W.K.; Algiers, T.J.; Ackerman, M.J.; Makielski, J.C. Loss-of-function mutation of the SCN3B-encoded sodium channel 3 subunit associated with a case of idiopathic ventricular fibrillation. Cardiovasc. Res. 2010, 86, 392–400. [Google Scholar] [CrossRef] [Green Version]
  422. Lasam, G.; Oaks, J.B. A Rare Desmoglein-2 Gene Mutation in Arrhythmogenic Right Ventricular Cardiomyopathy Inciting Incessant Ventricular Fibrillation. Cureus 2018, 10, e3388. [Google Scholar] [CrossRef] [Green Version]
  423. Rigato, I.; Bauce, B.; Rampazzo, A.; Zorzi, A.; Pilichou, K.; Mazzotti, E.; Migliore, F.; Marra, M.P.; Lorenzon, A.; De Bortoli, M.; et al. Compound and Digenic Heterozygosity Predicts Lifetime Arrhythmic Outcome and Sudden Cardiac Death in Desmosomal Gene–Related Arrhythmogenic Right Ventricular Cardiomyopathy. Circ. Cardiovasc. Genet. 2013, 6, 533–542. [Google Scholar] [CrossRef] [Green Version]
  424. Bhonsale, A.; Groeneweg, J.A.; James, C.A.; Dooijes, D.; Tichnell, C.; Jongbloed, J.D.H.; Murray, B.; te Riele, A.S.J.M.; van den Berg, M.P.; Bikker, H.; et al. Impact of genotype on clinical course in arrhythmogenic right ventricular dysplasia/cardiomyopathy-associated mutation carriers. Eur. Heart J. 2015, 36, 847–855. [Google Scholar] [CrossRef] [PubMed]
  425. Cox, M.G.P.J.; van der Zwaag, P.A.; van der Werf, C.; van der Smagt, J.J.; Noorman, M.; Bhuiyan, Z.A.; Wiesfeld, A.C.P.; Volders, P.G.A.; van Langen, I.M.; Atsma, D.E.; et al. Arrhythmogenic Right Ventricular Dysplasia/Cardiomyopathy. Circulation 2011, 123, 2690–2700. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  426. Marsman, R.F.; Barc, J.; Beekman, L.; Alders, M.; Dooijes, D.; van den Wijngaard, A.; Ratbi, I.; Sefiani, A.; Bhuiyan, Z.A.; Wilde, A.A.M.; et al. A Mutation in CALM1 Encoding Calmodulin in Familial Idiopathic Ventricular Fibrillation in Childhood and Adolescence. J. Am. Coll. Cardiol. 2014, 63, 259–266. [Google Scholar] [CrossRef] [Green Version]
  427. Nomikos, M.; Thanassoulas, A.; Beck, K.; Vassilakopoulou, V.; Hu, H.; Calver, B.L.; Theodoridou, M.; Kashir, J.; Blayney, L.; Livaniou, E.; et al. Altered RyR2 regulation by the calmodulin F90L mutation associated with idiopathic ventricular fibrillation and early sudden cardiac death. FEBS Lett. 2014, 588, 2898–2902. [Google Scholar] [CrossRef] [Green Version]
  428. Bagnall, R.D.; Molloy, L.K.; Kalman, J.M.; Semsarian, C. Exome sequencing identifies a mutation in the ACTN2 gene in a family with idiopathic ventricular fibrillation, left ventricular noncompaction, and sudden death. BMC Med. Genet. 2014, 15, 99. [Google Scholar] [CrossRef] [Green Version]
  429. Butters, T.D.; Aslanidi, O.V.; Inada, S.; Boyett, M.R.; Hancox, J.C.; Lei, M.; Zhang, H. Mechanistic Links Between Na+ Channel (SCN5A) Mutations and Impaired Cardiac Pacemaking in Sick Sinus Syndrome. Circ. Res. 2010, 107, 126–137. [Google Scholar] [CrossRef] [Green Version]
  430. GUI, J.; WANG, T.; TRUMP, D.; ZIMMER, T.; LEI, M. Mutation-Specific Effects of Polymorphism H558R in SCN5A-Related Sick Sinus Syndrome. J. Cardiovasc. Electrophysiol. 2010, 21, 564–573. [Google Scholar] [CrossRef] [PubMed]
  431. Abe, K.; Machida, T.; Sumitomo, N.; Yamamoto, H.; Ohkubo, K.; Watanabe, I.; Makiyama, T.; Fukae, S.; Kohno, M.; Harrell, D.T.; et al. Sodium Channelopathy Underlying Familial Sick Sinus Syndrome With Early Onset and Predominantly Male Characteristics. Circ. Arrhythmia Electrophysiol. 2014, 7, 511–517. [Google Scholar] [CrossRef] [Green Version]
  432. Nakajima, S.; Makiyama, T.; Hanazawa, K.; Kaitani, K.; Amano, M.; Hayama, Y.; Onishi, N.; Tamaki, Y.; Miyake, M.; Tamura, T.; et al. A Novel SCN5A Mutation Demonstrating a Variety of Clinical Phenotypes in Familial Sick Sinus Syndrome. Intern. Med. 2013, 52, 1805–1808. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  433. Tan, B.; Iturraldetorres, P.; Medeirosdomingo, A.; Nava, S.; Tester, D.; Valdivia, C.; Tusieluna, T.; Ackerman, M.; Makielski, J. A novel C-terminal truncation SCN5A mutation from a patient with sick sinus syndrome, conduction disorder and ventricular tachycardia. Cardiovasc. Res. 2007, 76, 409–417. [Google Scholar] [CrossRef] [Green Version]
  434. Kodama, T.; Serio, A.; Disertori, M.; Bronzetti, G.; Diegoli, M.; Narula, N.; Grasso, M.; Mazzola, S.; Arbustini, E. Autosomal recessive paediatric sick sinus syndrome associated with novel compound mutations in SCN5A. Int. J. Cardiol. 2013, 167, 3078–3080. [Google Scholar] [CrossRef] [PubMed]
  435. Holst, A.G.; Liang, B.; Jespersen, T.; Bundgaard, H.; Haunso, S.; Svendsen, J.H.; Tfelt-Hansen, J. Sick Sinus Syndrome, Progressive Cardiac Conduction Disease, Atrial Flutter and Ventricular Tachycardia Caused by a Novel SCN5A Mutation. Cardiology 2010, 115, 311–316. [Google Scholar] [CrossRef] [PubMed]
  436. Alkorashy, M.; Al-Ghamdi, B.; Tulbah, S.; Al-Numair, N.S.; Alhadeq, F.; Takroni, S.A.; Al-Hassnan, Z.N. A novel homozygous SCN5A variant detected in sick sinus syndrome. Pacing Clin. Electrophysiol. 2021, 44, 380–384. [Google Scholar] [CrossRef] [PubMed]
  437. Ziyadeh-Isleem, A.; Clatot, J.; Duchatelet, S.; Gandjbakhch, E.; Denjoy, I.; Hidden-Lucet, F.; Hatem, S.; Deschênes, I.; Coulombe, A.; Neyroud, N.; et al. A truncating SCN5A mutation combined with genetic variability causes sick sinus syndrome and early atrial fibrillation. Heart Rhythm 2014, 11, 1015–1023. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  438. Ackerman, M.J. Postmortem Molecular Analysis of SCN5A Defects in Sudden Infant Death Syndrome. JAMA 2001, 286, 2264. [Google Scholar] [CrossRef]
  439. Wang, D.W.; Desai, R.R.; Crotti, L.; Arnestad, M.; Insolia, R.; Pedrazzini, M.; Ferrandi, C.; Vege, A.; Rognum, T.; Schwartz, P.J.; et al. Cardiac Sodium Channel Dysfunction in Sudden Infant Death Syndrome. Circulation 2007, 115, 368–376. [Google Scholar] [CrossRef] [Green Version]
  440. Turillazzi, E.; La Rocca, G.; Anzalone, R.; Corrao, S.; Neri, M.; Pomara, C.; Riezzo, I.; Karch, S.B.; Fineschi, V. Heterozygous nonsense SCN5A mutation W822X explains a simultaneous sudden infant death syndrome. Virchows Arch. 2008, 453, 209–216. [Google Scholar] [CrossRef]
  441. Van Norstrand, D.W.; Tester, D.J.; Ackerman, M.J. Overrepresentation of the proarrhythmic, sudden death predisposing sodium channel polymorphism S1103Y in a population-based cohort of African-American sudden infant death syndrome. Heart Rhythm 2008, 5, 712–715. [Google Scholar] [CrossRef] [Green Version]
  442. Wedekind, H.; Smits, J.P.P.; Schulze-Bahr, E.; Arnold, R.; Veldkamp, M.W.; Bajanowski, T.; Borggrefe, M.; Brinkmann, B.; Warnecke, I.; Funke, H.; et al. De Novo Mutation in the SCN5A Gene Associated With Early Onset of Sudden Infant Death. Circulation 2001, 104, 1158–1164. [Google Scholar] [CrossRef] [Green Version]
  443. Plant, L.D. A common cardiac sodium channel variant associated with sudden infant death in African Americans, SCN5A S1103Y. J. Clin. Investig. 2006, 116, 430–435. [Google Scholar] [CrossRef] [Green Version]
  444. Todd, S.J.; Campbell, M.J.; Roden, D.M.; Kannankeril, P.J. Novel Brugada SCN5A mutation causing sudden death in children. Heart Rhythm 2005, 2, 540–543. [Google Scholar] [CrossRef] [PubMed]
  445. Neubauer, J.; Rougier, J.-S.; Abriel, H.; Haas, C. Functional implications of a rare variant in the sodium channel β1B subunit (SCN1B) in a 5-month-old male sudden infant death syndrome case. Heart Case Rep. 2018, 4, 187–190. [Google Scholar] [CrossRef] [Green Version]
  446. Tan, B.-H.; Pundi, K.N.; Van Norstrand, D.W.; Valdivia, C.R.; Tester, D.J.; Medeiros-Domingo, A.; Makielski, J.C.; Ackerman, M.J. Sudden infant death syndrome–associated mutations in the sodium channel beta subunits. Heart Rhythm 2010, 7, 771–778. [Google Scholar] [CrossRef] [Green Version]
  447. Denti, F.; Bentzen, B.H.; Wojciak, J.; Thomsen, N.M.; Scheinman, M.; Schmitt, N. Multiple genetic variations in sodium channel subunits in a case of sudden infant death syndrome. Pacing Clin. Electrophysiol. 2018, 41, 620–626. [Google Scholar] [CrossRef]
  448. Winkel, B.G.; Yuan, L.; Olesen, M.S.; Sadjadieh, G.; Wang, Y.; Risgaard, B.; Jabbari, R.; Haunsø, S.; Holst, A.G.; Hollegaard, M.V.; et al. The role of the sodium current complex in a nonreferred nationwide cohort of sudden infant death syndrome. Heart Rhythm 2015, 12, 1241–1249. [Google Scholar] [CrossRef]
  449. Makita, N.; Yagihara, N.; Crotti, L.; Johnson, C.N.; Beckmann, B.-M.; Roh, M.S.; Shigemizu, D.; Lichtner, P.; Ishikawa, T.; Aiba, T.; et al. Novel Calmodulin Mutations Associated With Congenital Arrhythmia Susceptibility. Circ. Cardiovasc. Genet. 2014, 7, 466–474. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  450. Cheng, J.; Van Norstrand, D.W.; Medeiros-Domingo, A.; Tester, D.J.; Valdivia, C.R.; Tan, B.-H.; Vatta, M.; Makielski, J.C.; Ackerman, M.J. LQTS-associated mutation A257G in α1-syntrophin interacts with the intragenic variant P74L to modify its biophysical phenotype. Cardiogenetics 2011, 1, 13. [Google Scholar] [CrossRef] [PubMed]
  451. Cronk, L.B.; Ye, B.; Kaku, T.; Tester, D.J.; Vatta, M.; Makielski, J.C.; Ackerman, M.J. Novel mechanism for sudden infant death syndrome: Persistent late sodium current secondary to mutations in caveolin-3. Heart Rhythm 2007, 4, 161–166. [Google Scholar] [CrossRef] [Green Version]
  452. Veltmann, C.; Barajas-Martinez, H.; Wolpert, C.; Borggrefe, M.; Schimpf, R.; Pfeiffer, R.; Cáceres, G.; Burashnikov, E.; Antzelevitch, C.; Hu, D. Further Insights in the Most Common SCN5A Mutation Causing Overlapping Phenotype of Long QT Syndrome, Brugada Syndrome, and Conduction Defect. J. Am. Heart Assoc. 2016, 5, 1–15. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  453. Kyndt, F.; Probst, V.; Potet, F.; Demolombe, S.; Chevallier, J.C.; Baro, I.; Moisan, J.P.; Boisseau, P.; Schott, J.J.; Escande, D.; et al. Novel SCN5A mutation leading either to isolated cardiac conduction defect or Brugada syndrome in a large French family. Circulation 2001, 104, 3081–3086. [Google Scholar] [CrossRef] [Green Version]
  454. Hothi, S.S.; Ara, F.; Timperley, J. p.Y1449C SCN5A Mutation Associated with Overlap Disorder Comprising Conduction Disease, Brugada Syndrome, and Atrial Flutter. J. Cardiovasc. Electrophysiol. 2015, 26, 93–97. [Google Scholar] [CrossRef]
  455. Chen, S. SNP S1103Y in the cardiac sodium channel gene SCN5A is associated with cardiac arrhythmias and sudden death in a white family. J. Med. Genet. 2002, 39, 913–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  456. Márquez, M.F.; Cruz-Robles, D.; Inés-Real, S.; Gallardo, G.J.; Gonzlez-Hermosillo, A.; Cárdenas, M.; Vargas-Alarcón, G. A novel SCN5A deletion mutation in a child with ventricular tachycardia, recurrent aborted sudden death, and Brugada electrocardiographic pattern. Arch. Cardiol. Mex. 2007, 77, 284–287. [Google Scholar] [PubMed]
  457. Patel, C.; Silcock, L.; McMullan, D.; Brueton, L.; Cox, H. TBX5 intragenic duplication: A family with an atypical Holt–Oram syndrome phenotype. Eur. J. Hum. Genet. 2012, 20, 863–869. [Google Scholar] [CrossRef] [Green Version]
Figure 1. SCN5A biosynthesis: chromosomal localization, gene transcriptional activation modulated by regulatory elements, transcription factors, histones, and SNPs. Functional, non-functional, and truncated isoforms derived from mRNA splicing, mechanism of the U6-type intron retention, and post-transcriptional regulation mediated by ncRNAs. Alternative exon sequences, intronic or exonic sequences outside the open reading frame (squared), and stop codons (asterisks) are indicated.
Figure 1. SCN5A biosynthesis: chromosomal localization, gene transcriptional activation modulated by regulatory elements, transcription factors, histones, and SNPs. Functional, non-functional, and truncated isoforms derived from mRNA splicing, mechanism of the U6-type intron retention, and post-transcriptional regulation mediated by ncRNAs. Alternative exon sequences, intronic or exonic sequences outside the open reading frame (squared), and stop codons (asterisks) are indicated.
Ijms 23 01381 g001
Figure 2. Biosynthesis and degradation pathways of Nav1.5. Only one of the possible scenarios where Nav1.5 assembles with one or more β-subunits early at the ER is depicted here. Furthermore, one possible scenario where ERAD-dependent degradation exclusively affects α subunit rather than α-β assembly is shown here since no information is currently available about the detailed process. ERRP—endoplasmic reticulum retention protein; β—beta subunit; PKA—protein kinase A; MOG1—RAN guanine nucleotide release factor; COPII—coat protein complex II; ERAD—ER-associated degradation; Cx43—connexin 43; PKP2—plakophilin 2; DSG—desmoglein; DSC—desmocollin; EB1—end-binding 1; Cav3—caveolin 3; PTMs—post-translational modifications.
Figure 2. Biosynthesis and degradation pathways of Nav1.5. Only one of the possible scenarios where Nav1.5 assembles with one or more β-subunits early at the ER is depicted here. Furthermore, one possible scenario where ERAD-dependent degradation exclusively affects α subunit rather than α-β assembly is shown here since no information is currently available about the detailed process. ERRP—endoplasmic reticulum retention protein; β—beta subunit; PKA—protein kinase A; MOG1—RAN guanine nucleotide release factor; COPII—coat protein complex II; ERAD—ER-associated degradation; Cx43—connexin 43; PKP2—plakophilin 2; DSG—desmoglein; DSC—desmocollin; EB1—end-binding 1; Cav3—caveolin 3; PTMs—post-translational modifications.
Ijms 23 01381 g002
Figure 3. (A) Schematic representation of the Nav1.5 secondary structure, (B) the intracellular view of the channel, and (C) the tertiary structure along with the interacting proteins. Only proteins with known binding sites in Nav1.5 are represented here. Mechanism of Nav1.5 regulation by the mitochondrial reactive oxygen species ROS is represented as well.
Figure 3. (A) Schematic representation of the Nav1.5 secondary structure, (B) the intracellular view of the channel, and (C) the tertiary structure along with the interacting proteins. Only proteins with known binding sites in Nav1.5 are represented here. Mechanism of Nav1.5 regulation by the mitochondrial reactive oxygen species ROS is represented as well.
Ijms 23 01381 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Daimi, H.; Lozano-Velasco, E.; Aranega, A.; Franco, D. Genomic and Non-Genomic Regulatory Mechanisms of the Cardiac Sodium Channel in Cardiac Arrhythmias. Int. J. Mol. Sci. 2022, 23, 1381. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23031381

AMA Style

Daimi H, Lozano-Velasco E, Aranega A, Franco D. Genomic and Non-Genomic Regulatory Mechanisms of the Cardiac Sodium Channel in Cardiac Arrhythmias. International Journal of Molecular Sciences. 2022; 23(3):1381. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23031381

Chicago/Turabian Style

Daimi, Houria, Estefanía Lozano-Velasco, Amelia Aranega, and Diego Franco. 2022. "Genomic and Non-Genomic Regulatory Mechanisms of the Cardiac Sodium Channel in Cardiac Arrhythmias" International Journal of Molecular Sciences 23, no. 3: 1381. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms23031381

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop