1,725
Views
35
CrossRef citations to date
0
Altmetric
Review Article

Intercellular Electrical Communication in the Heart: A New, Active Role for the Intercalated Disk

, &
Pages 161-167 | Received 04 Mar 2014, Accepted 16 Mar 2014, Published online: 15 Apr 2014

Abstract

Cardiac conduction is the propagation of electrical excitation through the heart and is responsible for triggering individual myocytes to contract in synchrony. Canonically, this process has been thought to occur electrotonically, by means of direct flow of ions from cell to cell. The intercalated disk (ID), the site of contact between adjacent myocytes, has been viewed as a structure composed of mechanical junctions that stabilize the apposition of cell membranes and gap junctions which constitute low resistance pathways between cells. However, emerging evidence suggests a more active role for structures within the ID in mediating intercellular electrical communication by means of non-canonical ephaptic mechanisms. This review will discuss the role of the ID in the context of the canonical, electrotonic view of conduction and highlight new, emerging possibilities of its playing a more active role in ephaptic coupling between cardiac myocytes.

Cardiac conduction is the propagation of electrical excitation through the heart, triggering synchronous contraction of individual cardiac myocytes in order to produce the pulsatile pumping action of the heart. The phenomenon was first demonstrated in preparations derived from frog hearts by Theodor Engelmann in 1874 (CitationEngelmann, 1875). Since then, significant advances have been made in the study of conduction motivated by the role played by abnormal conduction in mediating life-threatening arrhythmias in a variety of disease states.

WHAT IS CONDUCTION?

Cardiac conduction is the electrical phenomenon that integrates the function of individual myocytes to produce the function of the organ as a whole. The cardiac myocyte, which is an electrically active, mechanically contractile cell, is the structural and functional unit of cardiac muscle. At rest, the interior of the cardiac myocyte is electronegative with respect to the outside due to the heterogeneous distribution of ions across its membrane, resulting in a negative resting membrane potential. The cell can become electrically excited via rapid influx of positive charge in the form of sodium and calcium ions before being returned to resting potential by the efflux of positive charge in the form of potassium ions. This transient change in membrane potential comprising depolarization and repolarization is termed as action potential. Importantly, electrical activation of the myocytes elicits the release of calcium from intracellular stores, triggering mechanical contraction of the cell. Likewise, repolarization is accompanied by a decline in cytosolic calcium levels, resulting in relaxation of the myocyte. Thus, the individual myocyte functions as an electrically actuated contractile element. In the heart, electrical impulses originating in the pacemaking tissues of the sinoatrial node are communicated through the working myocardium in order to trigger coordinated contraction of individual myocytes, thus achieving the pumping action of the organ as a whole.

Here, we will focus on the mechanisms underlying the propagation of action potentials through the ventricular myocardium. Historically, ventricular conduction has been considered to have two separate aspects: cellular excitability, or the electrical response of a myocyte to a given excitatory impulse, and intercellular coupling. The intercalated disk (ID), which is the primary site of both mechanical and electrical contact has canonically been viewed as being critical to the latter with gap junctions located at the ID thought to afford direct electrical contact between adjacent myocytes. Here, we will review recent structural and functional insights that suggest a much more active role for the ID in conduction including potentially mediating long-theorized alternative modes of action potential propagation from myocyte to myocyte.

ELECTROTONIC CONDUCTION: THE CANONICAL VIEW

Historical perspective

After Engelmann's discovery of the phenomenon in 1874, a major milestone in the study of conduction at the tissue level came in 1914 when Lewis and colleagues reported the first measurements of conduction velocity (CitationLewis et al., 1914). A major breakthrough at the cellular level followed in 1959 when Hodgkin and Horowicz demonstrated that sodium entry underlay cardiac myocyte excitability (CitationHodgkin & Horowicz, 1959). They demonstrated a rapid increase in sodium conductance through the cell membrane during the depolarizing phase of the action potential, prompting Hodgkin and Huxley to hypothesize the existence of specialized “pores” or channels that convey Na+ across the cell membrane (CitationHodgkin & Huxley, 1952). These “pores” have since then been identified as voltage-gated sodium channels (Nav1.5) and their role in conduction extensively studied. While these advances occurred in the understanding of cellular excitability, similar strides were being made in explaining intercellular electrical communication. It was theorized early on that the cytoplasm of adjacent myocytes may be contiguous; however, electron microscopy revealed that myocytes are bounded by a membrane (CitationSjöstrand & Andersson, 1954). Shortly thereafter, Lloyd Barr published electron micrographs of fused membrane structures, which he labeled ‘the nexus’, connecting adjacent myocytes (CitationBarr et al., 1965; CitationDewey & Barr, 1962) while Van Der Kloot and Dane proposed that myocytes may be electrically coupled at the ID (CitationVanderkloot & Dane, 1964). Subsequently, work Revel and Karnovsky demonstrated that ‘the nexus’ consists of the membranes of adjacent myocytes very closely apposed but nonetheless separated by a narrow gap and coined the term ‘gap junction’ to describe this structure (CitationRevel & Karnovsky, 1967).

Cable theory

Even as inroads were made into understanding the structural basis of conduction, a conceptual framework began to develop with Silvio Weidman's application of continuous cable theory to describe the electrical properties of cardiac Purkinje fibers (CitationWeidmann, 1952). This approach, later extended to conduction through the ventricular myocardium, envisaged myocytes, electrically connected end-to-end and bounded by highly resistive lipid membranes, as analogous to a jacketed electrical cable consisting of a highly conductive core enveloped in an insulating sheath. Per this model, conduction occurs electrotonically by the movement of ions from cell to cell and the newly identified gap junctions offered an explanation of how. Early on, gap junctions were thought to be of sufficiently low resistance as to render the cytoplasms of coupled cells electrically contiguous and thereby to render cardiac muscle an electrical syncitium. Based on these assumptions, the model predicted an inverse square relationship between conduction velocity and the axial resistance of tissue and experimental measurements made in the late 1950s were consistent with this prediction (CitationDraper & Mya-Tu, 1959; CitationSano et al., 1959).

Also by the mid twentieth century, cardiac muscle was recognized as being anisotropic composed of brick-like myocytes, 100–150 μm long and 10–20 μm wide, organized in end-to-end fashion (CitationDe Mello, 1972). This structural picture fits well with Woddbury and Crill's work in the 1950s demonstrating direction-dependent current decay in cardiac muscle. By extending continuous cable theory to 2- and 3-dimensional models of anisotropic tissue, (CitationJack et al., 1975; CitationPeskoff, 1979a, Citation1979b) it was predicted that cardiac conduction velocity in tissue should be anisotropic. Experimental measurements in trabecular muscle as well as myocardial tissue demonstrated faster conduction parallel to fibers than transverse to them, as predicted by cable theory (CitationDraper & Mya-Tu, 1959; CitationSano et al., 1959; CitationClerc, 1976).

Gap junctions and conduction

Since first being identified, gap junctions have been the subject of intense study. Composed of proteins from the connexin family, gap junctions not only electrically couple myocytes but also allow the passage of molecules up to 1000 Daltons in size between them. Each gap-junction channel consists of two hemichannels, one from each of the abutting cells, docked together at the site of intercellular contact. In turn, each hemichannel or connexon is a hexamer composed of connexin proteins (CitationSöhl & Willecke, 2004). Three connexin isoforms 40, 43 and 45, so named after their molecular weights and possessing different conductances and permeabilities, are expressed in the heart (CitationBeyer, 1990; CitationBeyer et al., 1987; CitationKanter et al., 1992; CitationVozzi et al., 1999) with connexin43 (Cx43) being the principal gap-junction protein in the ventricular myocardium (CitationCoppen et al., 1998; CitationGourdie et al., 1993). Moreover, in the ventricle of humans, and many other mammals, Cx43 is almost exclusively localized to the IDs (CitationGourdie et al., 1991). Thus, gap junctions would appear to afford the type of intercellular electrical conductivity envisioned by cable theory.

A major evolution in our understanding of conduction occurred in the 1970s, when Spach and colleagues performed elegant microscopic measurements of conduction velocity and action potential upstroke demonstrating conduction to be discontinuous. This suggested that gap junctions may represent discontinuities in axial resistance, that is high resistance electrical pathways between myocytes, rather than rendering them syncitial. Indeed, direct measurements revealed the resistance at the ID to be comparable to the axial resistance of the rest of the myocyte (CitationRohr, 2004). Importantly, cable theory, when updated to include discontinuous axial resistance suggested that transverse conduction might be slower and also more discontinuous than longitudinal conduction and experimental studies proved consistent with this prediction (CitationSpach, 2003; CitationSpach et al., 1981).

For a more detailed discussion of electrotonic conduction, the reader is referred to our recent review (CitationVeeraraghavan et al., 2014) as well as the exhaustive review on the topic by CitationKléber and Rudy (2004).

BEYOND ELECTROTONIC CONDUCTION

Open questions

Motivation for research into the role played by gap junctions in conduction has stemmed primarily from their being implicated in arrhythmogenic conduction defects in a wide range of diseases. One of the most important open questions in this context is the precise correlation between the degree of loss of gap-junction coupling and the degree of conduction slowing resulting from it. While pharmacological gap-junction uncoupling has been uniformly correlated with slowed conduction, (CitationRohr, 2004; CitationKojodjojo et al., 2006; CitationRohr et al., 1998) the impact of genetically reduced functional expression and pathophysiological remodeling of gap junctions remains to be understood. Among the host of studies conducted on the same transgenic mouse lineage with a 50% loss of Cx43 levels, some groups observed significantly slowed conduction in the mutant mice relative to their wild-type littermates, (CitationEloff et al., 2001; CitationGuerrero et al., 1997) while others did not (CitationBeauchamp et al., 2004; CitationDanik et al., 2004; CitationMorley et al., 1999; CitationStein et al., 2009; CitationThomas et al., 2003; CitationVaidya et al., 2001; Citationvan Rijen et al., 2004). Likewise, reduced Cx43 expression was correlated with arrhythmogenic conduction slowing in a pacing-induced canine heart-failure model, (CitationPoelzing et al., 2004) whereas in a similar model, gap-junction remodeling and loss of Cx43 preceded conduction slowing (CitationAkar et al., 2007).

On the one hand, these conflicting studies hint at the complexity of the physiology underlying the catch-all term ‘gap-junction uncoupling’. Because of the short half-life of Cx43, its functional expression is regulated in a highly dynamic fashion. Further, modulation of Cx43 function is accomplished by post-translational modification, primarily phosphorylation/dephosphorylation of various residues on the molecule's C-terminus. The mechanisms by which Cx43 expression and function are regulated have been the subject of extensive research and are summarized in reviews by CitationSolan and Lampe (2007, Citation2014).

Another, more interesting possibility is that the seemingly disparate results vis-à-vis the relationship between gap-junction function and conduction are reflective of non-electrotonic mechanisms involved in cardiac conduction. The impetus to consider this hypothesis derives from investigations of conduction dependence on interstitial volume. If we consider conduction as that occurring via the electrotonic flow of current from cell to cell, then there must be a return path for the current via the interstitial space in order to complete the circuit. Cable theory predicts that the larger the interstitial space, the smaller its axial resistance and therefore, the faster conduction should be (CitationPlonsey & Barr, 2007; CitationSpach et al., 2004). This direct proportionality between interstitial volume and conduction velocity has been experimentally supported in the cable-like papillary muscle (CitationFleischhauer et al., 1995). However, in a recent experimental study in ventricular myocardium, we demonstrated conduction slowing and increased sensitivity to gap-junction uncoupling in response to an acute increase in interstitial volume (CitationVeeraraghavan et al., 2012). These results suggest that a purely electrotonic model may be insufficient to fully describe cardiac conduction.

Ephaptic coupling

As early as the 1960s, the hypothesis was put forth by Sperelakis and others that electrical excitation could be communicated between cardiac myocytes by means of an electric field or transient ion depletion/accumulation within the interstitial space (CitationSperelakis, 1969). These mechanisms, whereby propagation occurs without the electrotonic transfer of charge between myocytes via gap junctions, have been collectively termed ‘ephaptic coupling’. Although initially proposed as a complete alternative to gap-junction coupling, interest in the ideas has been revived by modeling studies that have suggested the possibility of ‘mixed mode’ coupling wherein both mechanisms operate in tandem (CitationKucera et al., 2002; CitationLin & Keener, 2010; Citation2014; CitationMori et al., 2008). These models envision ephaptic coupling as occurring by means of transient depletion of sodium from the interstitial space during depolarization. Briefly, depolarization of an upstream myocyte results in its sodium channels withdrawing positive charge from the interstitium, thereby driving the local extracellular potential to more negative levels. Consequently, the potential difference across the downstream myocyte's membrane is elevated, activating its sodium channels and causing it to depolarize. Based on this picture of ephaptic coupling, these models identify two principal requirements for ephaptic coupling to occur: (a) close apposition (< 10 nm) of adjacent cell membranes, and (b) a high density of Nav1.5 on the apposed membranes (CitationKucera et al., 2002; CitationLin & Keener, 2010, Citation2014; CitationMori et al., 2008; CitationCopene & Keener, 2008; CitationSperelakis, 2002; Citation2003; CitationSperelakis et al., 1983; CitationSperelakis & McConnell, 2002). Based on these criteria, the observation of conduction slowing in response to increased interstitial volume (CitationVeeraraghavan et al., 2012) would be expected if ephaptic coupling were involved in cardiac conduction. Further, the models also suggest that ephaptic coupling may sustain conduction when gap-junction coupling is reduced (CitationLin & Keener, 2013; Citation2014; CitationMori et al., 2008). This is consistent with the observation that increased interstitial volume unmasked conduction sensitivity to low levels of gap-junction uncoupling which on their own did not have a measurable impact (CitationVeeraraghavan et al., 2012). While these results are certainly encouraging, the key to elucidate the role, if any, of ephaptic coupling in cardiac conduction will be the identification of the cardiac ephapse, a structure that can serve as the functional unit of ephaptic coupling. Viewed in this light, recent insights into the structure of the ID raise some very interesting possibilities.

THE ID: THE MACHINERY OF CARDIAC CONDUCTION?

Given the close apposition between adjacent cell membranes required for ephaptic coupling, the ID is the logical place to start the search for the cardiac ephapse. Add to this the profusion of evidence that various ion channels, including the cardiac isoform of the Nav1.5, are localized at the ID, (CitationKucera et al., 2002; CitationCerrone et al., 2013; CitationCohen, 1996; CitationHong et al., 2012; CitationLin et al., 2011; CitationMaier et al., 2004; CitationMalhotra et al., 2004; CitationMilstein et al., 2012; CitationPetitprez et al., 2011; CitationRasmussen et al., 2004) and it is highly likely that the ID is the most likely location for structures conducive to ephaptic coupling. The emerging picture is one of macromolecular complex(es) localized at the ID that contain not just ion channels but also adapter proteins (e.g. ankyrin-G; AnkG) and components previously identified with electrical (e.g. Cx43) and mechanical (e.g. plakophilin-2; PKP2) junctions (CitationCerrone et al., 2013; CitationGeisler et al., 2010).

Sodium channels

Beyond just localizing to the ID, Nav1.5 also interacts with various disk-localized proteins. It was demonstrated to co-immunoprecipitate with Cx43 (CitationMalhotra et al., 2004) as well as to colocalize with it at the ID (CitationPetitprez et al., 2011) in mouse hearts. On the structural side, it has been demonstrated that Nav1.5 channels are preferentially localized at the ID compared with the lateral membrane of cardiac myocytes and are trafficked to these membrane sub-domains via distinct pathways (CitationPetitprez et al., 2011). Functionally, this translates into elevated sodium current density at the ID compared with the lateral membrane (CitationLin et al., 2011) and in silico results have gone as far as suggesting that disk-localized sodium channels may be primarily responsible for depolarization while those on the lateral membrane help maintain the stability of conduction (CitationTsumoto et al., 2011).

Important work by Delmar and colleagues in recent years has shed light on the degree to which Nav1.5 channels are integrated into the macromolecular complex at the ID. They have demonstrated that localization of mechanical junction proteins, mainly PKP2 (a desmosomal protein), precedes recruitment of Cx43 gap junctions and AnkG (a sub-membrane adapter protein involved in localizing Nav1.5 in the membrane) to sites of cell-to-cell contact in neonatal rat ventricular myocytes (CitationGeisler et al., 2010). Functionally, they have demonstrated that both regulation of the sodium current PKP2 and AnkG and linked mutations in both proteins to Brugada syndrome, an inherited arrhythmia syndrome resulting from loss of sodium current density (CitationCerrone et al., 2013; CitationHashemi et al., 2009). Additionally, they have reported reduced membrane levels of Nav1.5, particularly at sites devoid of Cx43, in conditional Cx43 knockout mice (CitationJansen et al., 2012) and provided data suggesting that Cx43 may be involved in recruiting Nav1.5 channels into the membrane (CitationAgullo-Pascual et al., 2013; CitationDelmar, 2012). Their observation of reduced Nav1.5 membrane levels without accompanying loss of gap-junction coupling in mice with the last five C-terminal amino acids of Cx43 deleted lends further credence to this hypothesis (CitationLübkemeier et al., 2013).

A key next step toward definitively the cardiac ephapse will be to locate microdomains within the ID that are rich in sodium channels and possess the requisite close apposition of membranes. In this context, a strong candidate that emerges is the perinexus—a membrane region located in the periphery of gap junctions where undocked connexon hemichannels localize (CitationRhett & Gourdie, 2012; CitationRhett et al., 2011a). Importantly, experiments using proximity ligation assays revealed fluorescent signal corresponding to the Cx43-Nav1.5 interaction localized at the perinexus (CitationRhett et al., 2012; CitationRhett et al., 2011b). Furthermore, the membranes of adjacent myocytes are very closely apposed at the perinexus by virtue of its location adjacent to the gap junction (CitationRhett et al., 2013). Likewise, work from the Delmar group has suggested the possibility of Nav1.5 channels localized near mechanical junctions, specifically desmosomes, (CitationCerrone et al., 2013; CitationNoorman et al., 2013) which could potentially constitute yet another microdomain conducive to ephaptic coupling. Additionally, these developments suggest novel scaffolding roles for proteins such as Cx43 which have been predominantly associated with their junctional functions.

Overall, perijunctional regions within the ID possess the necessary close apposition of adjacent cell membranes for ephaptic coupling and the localization of Nav1.5 to these regions is a strong, albeit suggestive, clue. Research into ephaptic coupling in the heart should focus on these microdomains as the most logical candidate structures to function as the cardiac ephapse. And in a somewhat ironic turn of events, it might very well turn out that ephaptic coupling, which was initially viewed as an alternative to gap-junction coupling, might not only occur in tandem with it, but also be rooted at the edge of the gap junctions. In this light, even as we rethink the mechanisms of cardiac conduction, so too must we now rethink the roles of electrical and mechanical junction proteins: rather than merely being the site of intercellular electrical coupling, the ID might represent the active machinery of cardiac conduction.

Declaration of interest: The authors report no declarations of interest. The authors alone are responsible for the content and writing of the paper.

This work was supported in part by grants from the National Institutes of Health (R01 HL102298-01A1 to SP, R01 HL56728-10A2 to RGG, R01-1DE019355-01 RGG subcontract), and American Heart Association Post-Doctoral Fellowship to RV. RGG holds stock in FirstString Research.

REFERENCES

  • Agullo-Pascual E, Lin X, Pfenniger A, Lübkemeier I, Willecke K, Rothenberg E, Delmar M (2013). A novel noncanonical role of cx43 in the heart: ensuring the arrival of Nav1.5 to the intercalated disk. Heart Rhythm. 10: 1742.
  • Akar FG, Nass RD, Hahn S, Cingolani E, Shah M, Hesketh GG, DiSilvestre D, Tunin RS, Kass DA, Tomaselli GF (2007). Dynamic changes in conduction velocity and gap junction properties during development of pacing-induced heart failure. Am J Physiol Heart Circ Physiol. 293: H1223–H1230.
  • Barr L, Dewey MM, Berger W (1965). Propagation of action potentials and the structure of the nexus in cardiac muscle. J Gen Physiol. 48: 797–823.
  • Beauchamp P, Choby C, Desplantez T, de Peyer K, Green K, Yamada KA, Weingart R, Saffitz JE, Kléber AG (2004). Electrical propagation in synthetic ventricular myocyte strands from germline connexin43 knockout mice. Circ Res. 95: 170–178.
  • Beyer EC (1990). Molecular cloning and developmental expression of two chick embryo gap junction proteins. J Biol Chem. 265: 14439–14443.
  • Beyer EC, Paul DL, Goodenough DA (1987). Connexin43: a protein from rat heart homologous to a gap junction protein from liver. J Cell Biol. 105: 2621–2629.
  • Cerrone M, Lin X, Zhang M, Agullo-Pascual E, Pfenniger A, Chkourko-Gusky H, Novelli V, Kim C, Tirasawadichai T, Judge DP, Rothenberg E, Chen HS, Napolitano C, Priori S, Delmar M (2013). Missense mutations in plakophilin-2 cause sodium current deficit and associate with a brugada syndrome phenotype. Circulation.
  • Clerc L (1976). Directional differences of impulse spread in trabecular muscle from mammalian heart. J Physiol. 255: 335–346.
  • Cohen SA (1996). Immunocytochemical localization of rH1 sodium channel in adult rat heart atria and ventricle. Presence in terminal intercalated disks. Circulation. 94: 3083–3086.
  • Copene ED, Keener JP (2008). Ephaptic coupling of cardiac cells through the junctional electric potential. J Math Biol. 57: 265–284.
  • Coppen SR, Dupont E, Rothery S, Severs NJ (1998). Connexin45 expression is preferentially associated with the ventricular conduction system in mouse and rat heart. Circ Res. 82: 232–243.
  • Danik SB, Liu F, Zhang J, Suk HJ, Morley GE, Fishman GI, Gutstein DE (2004). Modulation of cardiac gap junction expression and arrhythmic susceptibility. Circ Res. 95: 1035–1041.
  • De Mello WC (1972). Electrical Phenomena in the Heart. New York: Academic Press.
  • Delmar M (2012). Connexin43 regulates sodium current; ankyrin-G modulates gap junctions: the intercalated disc exchanger. Cardiovasc Res. 93: 220–222.
  • Dewey MM, Barr L (1962). Intercellular connection between smooth muscle cells: the nexus. Science. 137: 670–672.
  • Draper MH, Mya-Tu M (1959). A comparison of the conduction velocity in cardiac tissues of various mammals. Q J Exp Physiol Cogn Med Sci. 44: 91–109.
  • Eloff BC, Lerner DL, Yamada KA, Schuessler RB, Saffitz JE, Rosenbaum DS (2001). High resolution optical mapping reveals conduction slowing in connexin43 deficient mice. Cardiovasc Res. 51: 681–690.
  • Engelmann TW. (1875). Ueber die Leitung der Erregung im Herzmuskel. Pflüger Arch. 11: 465–480.
  • Fleischhauer J, Lehmann L, Kléber AG (1995). Electrical resistances of interstitial and microvascular space as determinants of the extracellular electrical field and velocity of propagation in ventricular myocardium. Circulation. 92: 587–594.
  • Geisler SB, Green KJ, Isom LL, Meshinchi S, Martens JR, Delmar M, Russell MW (2010). Ordered assembly of the adhesive and electrochemical connections within newly formed intercalated disks in primary cultures of adult rat cardiomyocytes. J Biomed Biotechnol. 2010: 624719.
  • Gourdie RG, Green CR, Severs NJ (1991). Gap junction distribution in adult mammalian myocardium revealed by an anti-peptide antibody and laser scanning confocal microscopy. J Cell Sci. 99: 41–55.
  • Gourdie RG, Green CR, Severs NJ, Anderson RH, Thompson RP (1993). Evidence for a distinct gap-junctional phenotype in ventricular conduction tissues of the developing and mature avian heart. Circ Res. 72: 278–289.
  • Guerrero PA, Schuessler RB, Davis LM, Beyer EC, Johnson CM, Yamada KA, Saffitz JE (1997). Slow ventricular conduction in mice heterozygous for a connexin43 null mutation. J Clin Invest. 99: 1991–1998.
  • Hashemi SM, Hund TJ, Mohler PJ (2009). Cardiac ankyrins in health and disease. J Mol Cell Cardiol. 47: 203–209.
  • Hodgkin AL, Horowicz P (1959). Movements of Na and K in single muscle fibres. J Physiol. 145: 405–432.
  • Hodgkin AL, Huxley AF (1952). A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol. 117: 500–544.
  • Hong M, Bao L, Kefaloyianni E, Agullo-Pascual E, Chkourko H, Foster M, Taskin E, Zhandre M, Reid DA, Rothenberg E, Delmar M, Coetzee WA (2012). Heterogeneity of ATP-sensitive K+ channels in cardiac myocytes: enrichment at the intercalated disk. J Biol Chem. 287: 41258–41267.
  • Jack JJB, Noble D, Tsien RW (1975). Electric Current Flow in Excitable Cells. Oxford: Clarendon Press.
  • Jansen JA, Noorman M, Musa H, Stein M, de Jong S, van der Nagel R, Hund TJ, Mohler PJ, Vos MA, van Veen TA, de Bakker JM, Delmar M, van Rijen HV (2012). Reduced heterogeneous expression of Cx43 results in decreased Nav1.5 expression and reduced sodium current that accounts for arrhythmia vulnerability in conditional Cx43 knockout mice. Heart Rhythm. 9: 600–607.
  • Kanter HL, Saffitz JE, Beyer EC (1992). Cardiac myocytes express multiple gap junction proteins. Circ Res. 70: 438–444.
  • Kléber AG, Rudy Y (2004). Basic mechanisms of cardiac impulse propagation and associated arrhythmias. Physiol Rev. 84: 431–488.
  • Kojodjojo P, Kanagaratnam P, Segal OR, Hussain W, Peters NS (2006). The effects of carbenoxolone on human myocardial conduction: a tool to investigate the role of gap junctional uncoupling in human arrhythmogenesis. J Am Coll Cardiol. 48: 1242–1249.
  • Kucera JP, Rohr S, Rudy Y (2002). Localization of sodium channels in intercalated disks modulates cardiac conduction. Circ Res. 91: 1176–1182.
  • Lewis T, Meakins J, White PD (1914). Part I. The Auricles. The Excitatory Process in the Dog's Heart. Philosophical Transactions of the Royal Society of London. Series B, Containing Papers of a Biological Character. Royal Society of London, 205: pp. 375–420.
  • Lin J, Keener JP (2010). Modeling electrical activity of myocardial cells incorporating the effects of ephaptic coupling. Proc Natl Acad Sci U S A. 107: 20935–20940.
  • Lin J, Keener JP (2013). Ephaptic coupling in cardiac myocytes. IEEE Trans Biomed Eng. 60: 576–582.
  • Lin J, Keener JP (2014). Microdomain effects on transverse cardiac propagation. Biophys J. 106: 925–931.
  • Lin X, Liu N, Lu J, Zhang J, Anumonwo JM, Isom LL, Fishman GI, Delmar M (2011). Subcellular heterogeneity of sodium current properties in adult cardiac ventricular myocytes. Heart Rhythm. 8: 1923–1930.
  • Lübkemeier I, Requardt RP, Lin X, Sasse P, Andrié R, Schrickel JW, Chkourko H, Bukauskas FF, Kim JS, Frank M, Malan D, Zhang J, Wirth A, Dobrowolski R, Mohler PJ, Offermanns S, Fleischmann BK, Delmar M, Willecke K (2013). Deletion of the last five C-terminal amino acid residues of connexin43 leads to lethal ventricular arrhythmias in mice without affecting coupling via gap junction channels. Basic Res Cardiol. 108: 348.
  • Maier SK, Westenbroek RE, McCormick KA, Curtis R, Scheuer T, Catterall WA (2004). Distinct subcellular localization of different sodium channel alpha and beta subunits in single ventricular myocytes from mouse heart. Circulation. 109: 1421–1427.
  • Malhotra JD, Thyagarajan V, Chen C, Isom LL (2004). Tyrosine-phosphorylated and nonphosphorylated sodium channel beta1 subunits are differentially localized in cardiac myocytes. J Biol Chem. 279: 40748–40754.
  • Milstein ML, Musa H, Balbuena DP, Anumonwo JM, Auerbach DS, Furspan PB, Hou L, Hu B, Schumacher SM, Vaidyanathan R, Martens JR, Jalife J (2012). Dynamic reciprocity of sodium and potassium channel expression in a macromolecular complex controls cardiac excitability and arrhythmia. Proc Natl Acad Sci U S A. 109: E2134–E2143.
  • Mori Y, Fishman GI, Peskin CS (2008). Ephaptic conduction in a cardiac strand model with 3D electrodiffusion. Proc Natl Acad Sci U S A. 105: 6463–6468.
  • Morley GE, Vaidya D, Samie FH, Lo C, Delmar M, Jalife J (1999). Characterization of conduction in the ventricles of normal and heterozygous Cx43 knockout mice using optical mapping. J Cardiovasc Electrophysiol. 10: 1361–1375.
  • Noorman M, Hakim S, Kessler E, Groeneweg JA, Cox MG, Asimaki A, van Rijen HV, van Stuijvenberg L, Chkourko H, van der Heyden MA, Vos MA, de Jonge N, van der Smagt JJ, Dooijes D, Vink A, de Weger RA, Varro A, de Bakker JM, Saffitz JE, Hund TJ, Mohler PJ, Delmar M, Hauer RN, van Veen TA (2013). Remodeling of the cardiac sodium channel, connexin43, and plakoglobin at the intercalated disk in patients with arrhythmogenic cardiomyopathy. Heart Rhythm. 10: 412–419.
  • Peskoff A. (1979a). Electric potential in cylindrical syncytia and muscle fibers. Bull Math Biol. 41: 183–192.
  • Peskoff A. (1979b). Electric potential in three-dimensional electrically syncytial tissues. Bull Math Biol. 41: 163–181.
  • Petitprez S, Zmoos AF, Ogrodnik J, Balse E, Raad N, El-Haou S, Albesa M, Bittihn P, Luther S, Lehnart SE, Hatem SN, Coulombe A, Abriel H (2011). SAP97 and dystrophin macromolecular complexes determine two pools of cardiac sodium channels Nav1.5 in cardiomyocytes. Circ Res. 108: 294–304.
  • Plonsey R, Barr RC. (2007). Bioelectricity: A Quantitative Approach. New York: Springer.
  • Poelzing S, Akar FG Baron E, Rosenbaum DS (2004). Heterogeneous connexin43 expression produces electrophysiological heterogeneities across ventricular wall. Am J Physiol Heart Circ Physiol. 286: H2001–H2009.
  • Rasmussen HB, Møller M, Knaus HG, Jensen BS, Olesen SP, Jørgensen NK (2004). Subcellular localization of the delayed rectifier K(+) channels KCNQ1 and ERG1 in the rat heart. Am J Physiol Heart Circ Physiol. 286: H1300–H1309.
  • Revel JP, Karnovsky MJ (1967). Hexagonal array of subunits in intercellular junctions of the mouse heart and liver. J Cell Biol. 33: C7–C12.
  • Rhett JM, Gourdie RG (2012). The perinexus: a new feature of Cx43 gap junction organization. Heart rhythm. 9: 619–623.
  • Rhett JM, Jourdan J, Gourdie RG (2011a). Connexin 43 connexon to gap junction transition is regulated by zonula occludens-1. Mol Biol Cell. 22: 1516–1528.
  • Rhett JM, Ongstad EL, Jourdan J, Gourdie RG (2012). Cx43 Associates with Na(v)1.5 in the cardiomyocyte perinexus. J Membr Biol. 245: 411–422.
  • Rhett JM, Palatinus JA Jourdan JA, Gourdie RG (2011b). Connexin43 interacts with voltage-gated sodium channel 1.5 in the perinexus. Circulation. 124: A9561.
  • Rhett JM, Veeraraghavan R, Poelzing S, Gourdie RG (2013). The perinexus: sign-post on the path to a new model of cardiac conduction?Trends Cardiovasc Med. 23: 222–228.
  • Rohr S (2004). Role of gap junctions in the propagation of the cardiac action potential. Cardiovasc Res. 62: 309–322.
  • Rohr S, Kucera JP, Kléber AG (1998). Slow conduction in cardiac tissue, I: effects of a reduction of excitability versus a reduction of electrical coupling on microconduction. Circ Res. 83: 781–794.
  • Sano T, Takayama N, Shimamoto T (1959). Directional difference of conduction velocity in the cardiac ventricular syncytium studied by microelectrodes. Circ Res. 7: 262–267.
  • Sjöstrand FS, Andersson E (1954). Electron microscopy of the intercalated discs of cardiac muscle tissue. Experientia. 10: 369–370.
  • Söhl G, Willecke K (2004). Gap junctions and the connexin protein family. Cardiovasc Res. 62: 228–232.
  • Solan JL, Lampe PD (2007). Key connexin 43 phosphorylation events regulate the gap junction life cycle. J Membr Biol. 217: 35–41.
  • Solan JL, Lampe PD (2014). Specific Cx43 phosphorylation events regulate gap junction turnover in vivo. FEBS lett. pii: S0014–5793(14)00091–X.
  • Spach MS (2003). Transition from a continuous to discontinuous understanding of cardiac conduction. Circ Res. 92: 125–126.
  • Spach MS, Heidlage JF, Barr RC, Dolber PC (2004). Cell size and communication: role in structural and electrical development and remodeling of the heart. Heart Rhythm. 1: 500–515.
  • Spach MS, Miller WT 3rd, Geselowitz DB, Barr RC, Kootsey JM, Johnson EA (1981). The discontinuous nature of propagation in normal canine cardiac muscle. Evidence for recurrent discontinuities of intracellular resistance that affect the membrane currents. Circ Res. 48: 39–54.
  • Sperelakis N (1969). Lack of electrical coupling between contiguous myocardial cells in vertebrate hearts. Experientia Suppl. 15: 135–165.
  • Sperelakis N (2002). An electric field mechanism for transmission of excitation between myocardial cells. Circ Res. 91: 985–987.
  • Sperelakis N (2003). Combined electric field and gap junctions on propagation of action potentials in cardiac muscle and smooth muscle in PSpice simulation. J Electrocardiol. 36: 279–293.
  • Sperelakis N, Marschall RA, Mann JE (1983). Propagation down a chain of excitable cells by electric field interactions in the junctional clefts: effect of variation in extracellular resistances, including a “sucrose gap” simulation. IEEE Trans Biomed Eng. 30: 658–664.
  • Sperelakis N, McConnell K (2002). Electric field interactions between closely abutting excitable cells. IEEE Eng Med Biol Mag. 21: 77–89.
  • Stein M, van Veen TA, Remme CA, Boulaksil M, Noorman M, van Stuijvenberg L, van der Nagel R, Bezzina CR, Hauer RN, de Bakker JM, van Rijen HV (2009). Combined reduction of intercellular coupling and membrane excitability differentially affects transverse and longitudinal cardiac conduction. Cardiovasc Res. 83: 52–60.
  • Thomas SP, Kucera JP, Bircher-Lehmann L, Rudy Y, Saffitz JE, Kléber AG (2003). Impulse propagation in synthetic strands of neonatal cardiac myocytes with genetically reduced levels of connexin43. Circ Res. 92: 1209–1216.
  • Tsumoto K, Ashihara T, Haraguchi R, Nakazawa K, Kurachi Y (2011). Roles of subcellular Na+ channel distributions in the mechanism of cardiac conduction. Biophys J. 100: 554–563.
  • Vaidya D, Tamaddon HS, Lo CW, Taffet SM, Delmar M, Morley GE, Jalife J (2001). Null mutation of connexin43 causes slow propagation of ventricular activation in the late stages of mouse embryonic development. Circ Res. 88: 1196–1202.
  • van Rijen HV, Eckardt D, Degen J, Theis M, Ott T, Willecke K, Jongsma HJ, Opthof T, de Bakker JM (2004). Slow conduction and enhanced anisotropy increase the propensity for ventricular tachyarrhythmias in adult mice with induced deletion of connexin43. Circulation. 109: 1048–1055.
  • Vanderkloot WG, Dane B (1964). Conduction of the action potential in the frog ventricle. Science. 146: 74–75.
  • Veeraraghavan R, Gourdie R, Poelzing S (2014). Mechanisms of cardiac conduction: a history of revisions. Am J Physiol Heart Circ Physiol. 306: H619–H627.
  • Veeraraghavan R, Salama ME, Poelzing S (2012). Interstitial volume modulates the conduction velocity-gap junction relationship. Am J Physiol Heart Circ Physiol. 302: H278–H286.
  • Vozzi C, Dupont E, Coppen SR, Yeh HI, Severs NJ (1999). Chamber-related differences in connexin expression in the human heart. J Mol Cell Cardiol. 31: 991–1003.
  • Weidmann S. (1952). The electrical constants of Purkinje fibres. J Physiol. 118: 348–360.

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.