949
Views
4
CrossRef citations to date
0
Altmetric
Editorial

Late sodium current dysregulation as a causal factor in arrhythmia

&
Pages 545-547 | Received 04 Jan 2016, Accepted 15 Feb 2016, Published online: 16 Mar 2016

Voltage-gated Na+ channels (Nav) are essential for normal cardiac function. Defects in the function of Nav1.5, the primary Nav alpha subunit in heart (encoded by SCN5A), have been linked to a host of congenital and acquired cardiac arrhythmia syndromes. Specifically, there has been growing appreciation in the field for the importance of Nav1.5 defects that produce a unique gating mode characterized by an increase in inappropriate persistent ‘late’ current (INa,L). Despite the considerable effort to understand the role of aberrant INa,L in disease, there remain fundamental unanswered questions about the pathways responsible for regulation of INa,L in health and disease, and, more importantly, whether these pathways may be exploited for therapeutic benefit in human patients.

Despite important advances in anti-arrhythmia therapy over the past half century, cardiac arrhythmia remains a major source of morbidity and mortality [Citation1,Citation2]. Much focus has been given to ion channels as targets for anti-arrhythmia agents, but unintended pro-arrhythmia and off-target effects associated with ion channel blocking drugs motivate the continued pursuit of new and improved anti-arrhythmia agents [Citation3]. Exciting recent work in the field suggests that a novel anti-arrhythmia approach with great potential involves targeting a pathogenic component of voltage-gated Na+ current (‘late’ Na+ current, INa,L) [Citation4Citation8].

Nav generate the cardiac action potential (AP) upstroke through a tightly controlled gating process involving rapid channel activation followed immediately by inactivation, which sets the stage for smaller conductance/slower ion channels (mostly voltage-gated Ca2+ and K+ channels) to repolarize the membrane in preparation for the next heartbeat. While Nav current (INa) is mostly terminated by the rapid voltage-dependent inactivation process within milliseconds, a small percentage of channels (<1%) remain available throughout the AP, giving rise to a persistent (‘late’) current (INa,L) even under normal conditions [Citation9]. While the physiological role of INa,L is undetermined, it is possible that INa,L supports contractility at baseline and in response to acute stress (i.e. ‘fight-or-flight’ response). Regardless of the physiological function, increased INa,L has been observed in congenital gain-of-function arrhythmia syndromes (e.g. long QT 3), and in acquired forms of disease (e.g. heart failure, atrial fibrillation (AF)) [Citation9Citation13]. INa,L likely promotes arrhythmogenesis by (1) increasing a depolarizing current that prolongs the AP and increases susceptibility to secondary depolarizations (afterdepolarizations) that serve as potential arrhythmia triggers and (2) increasing intracellular Na+ that promotes intracellular Ca2+ accumulation via altered activity of Na+/Ca2+ exchanger [Citation2,Citation6]. Consistent with pro-arrhythmia of INa,L, drugs that block INa,L have been shown in pre-clinical and clinical trials to reduce arrhythmia incidence, although the precise mode of action is not resolved [Citation4,Citation7,Citation8,Citation14Citation16]. Despite this important work, questions remain about the mechanisms for dysregulation of INa,L in disease and whether a therapeutic strategy focused on INa,L is likely the most beneficial in our effort to reduce arrhythmia burden in human patients.

How is INa,L regulated in health and disease?

The cardiac cell has evolved elaborate signaling networks for tuning of INa,L via post-translational modification of Nav1.5 (e.g. phosphorylation, nitrosylation, glycosylation) [Citation17]. Among the most well-studied signaling pathways for phosphorylation of Nav1.5 is the multifunctional Ca2+/calmodulin-dependent protein kinase II (CaMKII), which is known to be activated in a wide range of cardiac disease states [Citation18]. CaMKII directly phosphorylates Nav1.5 to alter several INa properties, including the magnitude of INa,L [Citation19Citation23]. Although several phosphorylation sites for CaMKII have been identified in the Nav1.5 Domain I-Domain II (DI-DII) linker [Citation19,Citation20,Citation24,Citation25], mounting data from our group and others point to an important regulatory locus at Ser571 [Citation5,Citation10,Citation19,Citation21,Citation24]. Specifically, increased phosphorylation of Ser571 occurs under stress conditions in vitro and in animal models of disease and in human heart failure [Citation19,Citation21]. We recently developed two knock-in mice models to study the in vivo role of Ser571 phosphorylation: (1) the S571A mouse, which lacks the Ser571 phosphorylation site due to replacement of the serine with an alanine and (2) the S571E mouse, which substitutes a glutamic acid at position 571 to mimic constitutive phosphorylation (phospho-mimetic) [Citation5]. Using these mouse models, we demonstrated that phosphorylation of Ser571 specifically regulates INa,L without affecting other INa properties (peak current, steady-steady inactivation) Furthermore, ablation of the Ser571 phosphorylation site in S571A mice conferred resistance to maladaptive remodeling and arrhythmias under chronic pressure-overload conditions (transaortic constriction) [Citation5]. While these studies point to an important role for Ser571 in regulating INa,L, they do not explain the full range of CaMKII-dependent effects on INa. Additional in vivo studies with relevant animal models will be required to assess the role of other phosphorylation sites (e.g. S516) in controlling the final phenotype. We expect that phosphorylation of additional sites act in concert with S571 to mediate CaMKII-dependent control of other channel properties (e.g. channel inactivation). Going forward, systems-based approaches to analyze post-translational modification in an unbiased way may be necessary to fully understand how Nav1.5 is regulated [Citation26]. Also, it is important to note that while CaMKII-dependent phosphorylation of Nav1.5 is likely a critical pathway for dysregulation of INa,L in disease, there are multiple potential avenues for enhanced INa,L downstream of pathogenic stimuli (e.g. adrenergic imbalance, oxidative stress).

Aside from common forms of acquired arrhythmia, increased INa,L has been shown to underlie dozens of gain-of-function mutations in SCN5A linked to long-QT syndrome type 3 [Citation12]. Most of these mutations are missense mutations that destabilize or slow the rapid inactivation process. For example, the first inherited LQT3 mutation to be identified, ΔKPQ, results in a 3-amino acid deletion in the Domain III-Domain IV (DIII-DIV) linker of Nav1.5, which promotes channel reopenings at depolarized potentials [Citation27,Citation28]. Interestingly, in recent studies, we found that inherited arrhythmia mutations involving phospho-mimetic mutations at sites adjacent to the Ser571 site (A572D and Q573E) conferred arrhythmia susceptibility, in part, by mimicking CaMKII-dependent phosphorylation of Nav1.5 [Citation21]. These studies support the importance of Ser571 in modulation of INa,L and suggest a novel mechanism for dysregulation of INa,L in inherited disease.

Is INa,L an optimal target for preventing arrhythmia?

Based on the strong link between increased INa,L and inherited and acquired forms of arrhythmia, it is not surprising that INa,L has emerged as a potential target for reducing risk of arrhythmia. A wealth of pre-clinical data support the utility of INa,L blockers in preventing a wide range of arrhythmias [Citation2,Citation4,Citation8,Citation14Citation16]. Moreover, recent clinical trials demonstrate efficacy of the INa,L blocker ranolazine for reducing incidence of supraventricular arrhythmias and new onset AF in myocardial infarction patients, as well as termination of paroxysmal AF and prevention of post-operative AF [Citation6,Citation7]. These studies all point to the anti-arrhythmia efficacy of INa,L blocking drugs. While there is clearly a basis for optimism, it is important to note limitations, including the nonspecific nature of these compounds with effects on other INa properties (i.e. peak current) and other channels, and effects that may vary across regions of the heart [Citation29]. Next-generation compounds such as GS-458967 attempt to address some of these limitations by increasing selectivity for INa,L with promising anti-arrhythmic results in pre-clinical studies [Citation30]. Furthermore, neuronal Nav isoforms (e.g. Nav1.1, Nav1.2), in addition to Nav1.5, likely make an important contribution to INa,L and may complicate therapeutic targeting of late current [Citation31]. At the same time, while studies from our group and others point to INa,L as an important gateway for arrhythmogenic imbalance of intracellular ion homeostasis, it is clearly not the only source. Pathogenic signals upstream of INa,L (e.g. CaMKII) alter the function of other key substrates important for normal cell excitability and ion homeostasis. For example, sarcoplasmic reticulum Ca2+ release channels are also phosphorylated by CaMKII in stress conditions and promote arrhythmia by increasing likelihood of inappropriate spontaneous Ca2+ release from the sarcoplasmic reticulum [Citation32]. Thus, while the data support the potential of an anti-arrhythmic strategy involving suppression of INa,L, it remains to be determined whether INa,L block alone will be necessary/sufficient or whether it should be considered in the context of an adjuvant therapy.

Financial and competing interests disclosure

The authors have received support from the National Institutes of Health (grant number HL114893 to TJ Hund) and from the James S. McDonnell Foundation (to TJ Hund). The authors have no other relevant affiliations or financial involvement with any organization or entity with a financial interest in or financial conflict with the subject matter or materials discussed in the manuscript apart from those disclosed.

References

  • Nattel S, et al. New directions in cardiac arrhythmia management: present challenges and future solutions. Can J Cardiol. 2014;30(12 Suppl): S420–30. doi:10.1016/j.cjca.2014.09.027.
  • Frommeyer G, et al. Late sodium current inhibition: the most promising antiarrhythmic principle in the near future? Curr Med Chem. 2014;21(11):1271–1280.
  • Hund TJ, Mohler PJ. Nav channel complex heterogeneity: new targets for the treatment of arrhythmia?. Circulation. 2014;130:132–134. doi:10.1161/CIRCULATIONAHA.114.010867.
  • Coppini R, et al. Late sodium current inhibition reverses electromechanical dysfunction in human hypertrophic cardiomyopathy. Circulation. 2013;127(5):575–584. doi:10.1161/CIRCULATIONAHA.112.134932.
  • Glynn P, et al. Voltage-gated sodium channel phosphorylation at Ser571 regulates late current, arrhythmia, and cardiac function in vivo. Circulation. 2015;132(7):567–577. doi:10.1161/CIRCULATIONAHA.114.015218.
  • Burashnikov A, Antzelevitch C. Role of late sodium channel current block in the management of atrial fibrillation. Cardiovasc Drugs Ther. 2013;27(1):79–89. doi:10.1007/s10557-012-6421-1.
  • Scirica BM, et al. Effect of ranolazine, an antianginal agent with novel electrophysiological properties, on the incidence of arrhythmias in patients with non ST-segment elevation acute coronary syndrome: results from the Metabolic Efficiency With Ranolazine for Less Ischemia in Non ST-Elevation Acute Coronary Syndrome Thrombolysis in Myocardial Infarction 36 (MERLIN-TIMI 36) randomized controlled trial. Circulation. 2007;116(15):1647–1652. doi:10.1161/CIRCULATIONAHA.107.724880.
  • Sossalla S, et al. Ranolazine improves diastolic dysfunction in isolated myocardium from failing human hearts–role of late sodium current and intracellular ion accumulation. J Mol Cell Cardiol. 2008;45(1):32–43. doi:10.1016/j.yjmcc.2008.03.006.
  • Maltsev VA, et al. Novel, ultraslow inactivating sodium current in human ventricular cardiomyocytes. Circulation. 1998;98(23):2545–2552.
  • Toischer K, et al. Role of late sodium current as a potential arrhythmogenic mechanism in the progression of pressure-induced heart disease. J Mol Cell Cardiol. 2013;61:111–122. doi:10.1016/j.yjmcc.2013.03.021.
  • Valdivia CR, et al. Increased late sodium current in myocytes from a canine heart failure model and from failing human heart. J Mol Cell Cardiol. 2005;38(3):475–483. doi:10.1016/j.yjmcc.2004.12.012.
  • Belardinelli L, et al. Cardiac late Na(+) current: proarrhythmic effects, roles in long QT syndromes, and pathological relationship to CaMKII and oxidative stress. Heart Rhythm. 2015;12(2):440–448. doi:10.1016/j.hrthm.2014.11.009.
  • Shryock JC, et al. The arrhythmogenic consequences of increasing late INa in the cardiomyocyte. Cardiovasc Res. 2013;99(4):600–611. doi:10.1093/cvr/cvt145.
  • Song Y, et al. Antagonism by ranolazine of the pro-arrhythmic effects of increasing late INa in guinea pig ventricular myocytes. J Cardiovasc Pharmacol. 2004;44(2):192–199.
  • Undrovinas AI, et al. Ranolazine improves abnormal repolarization and contraction in left ventricular myocytes of dogs with heart failure by inhibiting late sodium current. J Cardiovasc Electrophysiol. 2006;17(Suppl 1): S169–S177. doi:10.1111/j.1540-8167.2006.00401.x.
  • Burashnikov A, et al. Atrium-selective sodium channel block as a strategy for suppression of atrial fibrillation: differences in sodium channel inactivation between atria and ventricles and the role of ranolazine. Circulation. 2007;116(13):1449–1457. doi:10.1161/CIRCULATIONAHA.107.704890.
  • Marionneau C, Abriel H. Regulation of the cardiac Na+ channel NaV1.5 by post-translational modifications. J Mol Cell Cardiol. 2015;82:36–47. doi:10.1016/j.yjmcc.2015.02.013.
  • Swaminathan PD, et al. Calmodulin-dependent protein kinase II: linking heart failure and arrhythmias. Circ Res. 2012;110(12):1661–1677. doi:10.1161/CIRCRESAHA.111.243956.
  • Hund TJ, et al. A betaIV spectrin/CaMKII signaling complex is essential for membrane excitability in mice. J Clin Invest. 2010;120:3508–3519. doi:10.1172/JCI43621.
  • Ashpole NM, et al. 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. doi:10.1074/jbc.M111.322537.
  • Koval OM, et al. Ca2+/calmodulin-dependent protein kinase II-based regulation of voltage-gated Na+ channel in cardiac disease. Circulation. 2012;126(17):2084–2094. doi:10.1161/CIRCULATIONAHA.112.105320.
  • Aiba T, et al. Na+ channel regulation by Ca2+/calmodulin and Ca2+/calmodulin-dependent protein kinase II in guinea-pig ventricular myocytes. Cardiovasc Res. 2010;85:454–463. doi:10.1093/cvr/cvp324.
  • Wagner S, et al. Ca/calmodulin-dependent protein kinase II regulates cardiac Na channels. J Clin Invest. 2006;116(12):3127–3138. doi:10.1172/JCI26620.
  • Marionneau C, et al. Mass spectrometry-based identification of native cardiac Nav1.5 channel alpha subunit phosphorylation sites. J Proteome Res. 2012;11(12):5994–6007. doi:10.1021/pr300702c.
  • Herren AW, et al. 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(5):2298–2311. doi:10.1021/acs.jproteome.5b00107.
  • Hennrich ML, Gavin AC. Quantitative mass spectrometry of posttranslational modifications: keys to confidence. Sci Signal. 2015;8(371):re5. doi:10.1126/scisignal.aaa6466.
  • Wang Q, et al. SCN5A mutations associated with an inherited cardiac arrhythmia, long QT syndrome. Cell. 1995;80(5):805–811.
  • Bennett P, et al. Molecular mechanism for an inherited cardiac arrhythmia. Nature. 1995;376(Aug 24):683–685. doi:10.1038/376683a0.
  • Morotti S, et al. Atrial-selective targeting of arrhythmogenic phase-3 early afterdepolarizations in human myocytes. J Mol Cell Cardiol. 2015. doi:10.1016/j.yjmcc.2015.07.030.
  • Belardinelli L, et al. A novel, potent, and selective inhibitor of cardiac late sodium current suppresses experimental arrhythmias. J Pharmacol Exp Ther. 2013;344(1):23–32. doi:10.1124/jpet.112.198887.
  • Mishra S, et al. Contribution of sodium channel neuronal isoform Nav1.1 to late sodium current in ventricular myocytes from failing hearts. J Physiol. 2015;593(6):1409–1427. doi:10.1113/jphysiol.2014.278259.
  • Van Oort RJ, et al. Ryanodine receptor phosphorylation by calcium/calmodulin-dependent protein kinase II promotes life-threatening ventricular arrhythmias in mice with heart failure. Circulation. 2010;122(25):2669–2679. doi:10.1161/CIRCULATIONAHA.110.982298.

Reprints and Corporate Permissions

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

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

Academic Permissions

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

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

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