Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2010 Mar 1.
Published in final edited form as: IUBMB Life. 2009 Mar;61(3):215–221. doi: 10.1002/iub.163

Na+ Transport in Cardiac Myocytes; Implications for Excitation-Contraction Coupling

Donald M Bers 1, Sanda Despa 1
PMCID: PMC2669704  NIHMSID: NIHMS100840  PMID: 19243007

Abstract

Intracellular Na+ concentration ([Na+]i) is very important in modulating the contractile and electrical activity of the heart. Upon electrical excitation of the myocardium, voltage-dependent Na+ channels open, triggering the upstroke of the action potential (AP). During the AP, Ca2+ enters the myocytes via L-type Ca2+ channels. This triggers Ca2+ release from the sarcoplasmic reticulum (SR) and thus activates contraction. Relaxation occurs when cytosolic Ca2+ declines, mainly due to re-uptake into the SR via SR Ca2+-ATPase and extrusion from the cell via the Na+/Ca2+ exchanger (NCX). NCX extrudes one Ca2+ ion in exchange for three Na+ ions and its activity is critically regulated by [Na+]i. Thus, via NCX, [Na+]i is centrally involved in the regulation of intracellular [Ca2+] and contractility. Na+ brought in by Na+ channels, NCX and other Na+ entry pathways is extruded by the Na+/K+ pump (NKA) to keep [Na+]i low. NKA is regulated by phospholemman, a small sarcolemmal protein that associates with NKA. Unphosphorylated phospholemman inhibits NKA by decreasing the pump affinity for internal Na+ and this inhibition is relieved upon phosphorylation. Here we discuss the main characteristics of the Na+ transport pathways in cardiac myocytes and their physiological and pathophysiological relevance.

Keywords: Na+/K+ ATPase, Na+/Ca2+ exchanger, Na+ channels, phospholemman

[Na+]i during cardiac cycle and its role in excitation-contraction coupling

Mammalian cells maintain a large electrochemical gradient of Na+ across the plasma membrane. Intracellular Na+ concentration ([Na+]i) is kept low (4-14 mM) by the Na+/K+-ATPase (NKA) at the expense of metabolic energy, while extracellular [Na+] is ∼140 mM. In neurons and muscles, the energy stored in the transmembrane [Na+] gradient provides the basis for fast electrical signaling, i.e. propagation of action potentials. Many cells utilize the electrochemical Na+ gradient to couple energetically unfavorable transmembrane solute flow to Na+ transport (secondary active transport). Examples include Na+/neurotransmitter, Na+/glucose and Na+/amino acids cotransporters as well as various Na+/ion co- and counter-transporters.

In the heart, [Na+]i is very important in modulating intracellular Ca2+, and thus contractility, through the Na+/Ca2+ exchanger (NCX). The exchanger is the main pathway for Ca2+ extrusion from cardiac myocytes and therefore is essential in excitation-contraction coupling. NCX can operate in both Ca2+ efflux and Ca2+ influx (or reverse) mode, depending on the internal and external concentration of both Na+ and Ca2+, as well as on the membrane potential. High [Ca2+]i favors Ca2+ efflux whereas positive membrane potential and high [Na+]i favor Ca2+ influx. In normal myocytes, under physiological conditions NCX works almost exclusively in the Ca2+ extrusion mode, driven mostly by the high subsarcolemmal Ca2+ transient (1). However, elevated [Na+]i would shift the balance of fluxes through NCX to favor more Ca2+ influx and less Ca2+ efflux, resulting in larger Ca2+ transients and therefore enhanced contractility. Indeed, the inotropic effect of increasing [Na+]i is well known and is the basic mechanism behind cardiac-glycosides induced inotropy. In cardiac Purkinje fibers, the developed tension doubles with a 1 mM increase in Na+ activity (2). The effect is smaller but still prominent in ventricular myocytes (3).

During electrical excitation of the heart, voltage-dependent Na+ channels open, triggering the upstroke of the action potential (AP; Fig. 1). Na+ channels inactivate rapidly (within a few milliseconds) at positive potentials, which limits the gain in intracellular Na+ ([Na+]i) to 6-15 μmol/L (4). During the AP, Ca2+ enters the myocytes via L-type Ca2+ channels, which triggers Ca2+ release from the sarcoplasmic reticulum (SR). This raises the free intracellular [Ca2+] ([Ca2+]i), allowing Ca2+ to bind to the myofilaments and trigger contraction. For relaxation to occur, Ca2+ must be taken out of the cytosol, mainly via the SR Ca2+-ATPase (SERCA), which takes Ca2+ back into the SR, and NCX (Fig. 1). NCX transports one Ca2+ ion in exchange for three Na+ ions, thus Ca2+ extrusion leads to an increase in [Na+]i by ∼32 μmol/L during each AP (Fig. 1). This is required to extrude the ∼10 μmol/L Ca2+ that enters via L-type Ca2+ channels. The Na+/H+ exchanger brings in ∼2 μmol/L Na+ at physiological intracellular pH and ∼16 μmol/L during intracellular acidosis (see 4). Other transporters, including the Na+/HCO3- cotransporter, the Na+/K+/2Cl cotransporter and the Na+/Mg2+ exchanger bring in smaller amounts of Na+. At steady-state, the excess Na+ (∼40-45 μmol/L) is then extruded by NKA to keep [Na+]i constant (Fig. 1).

Figure 1.

Figure 1

Na+ and Ca2+ transport during cardiac cycle in ventricular myocytes. Inset shows the integrated Na flux through NCX, voltage-gated Na+ channels and NKA during an action potential (AP). The corresponding Ca2+ transient is also shown. NCX - Na+/Ca2+ exchanger; ATP - ATPase; PLB - phospholamban; PLM – phospholemman; SR - sarcoplasmic reticulum; NHE – Na/H exchanger. Modified from Bers (2001).

Resting [Na+]i in cardiac myocytes is in the range of 4-8 mM for most mammalian species, including human, and higher (10-15 mM) in rat and mouse (4). [Na+]i increases during stimulation in a frequency-dependent manner, because, as detailed above, Na+ influx is larger due to more frequent activation of Na+ channels and NCX. [Na+]i is elevated in hypertrophy and heart failure (HF) (3,5,6) and this may limit the contractile dysfunction observed in these pathophysiological conditions.

Voltage-gated Na+ channels

The cardiac voltage-gated Na+ channel is composed of a pore-forming α subunit Nav1.5 (∼240 kDa) and auxiliary β subunits (∼30-35 kDa, β1-β4 subunits; (7)). The α subunit consists of four repeat domains (I-IV), each containing six transmembrane segments (S1-S6) and one membrane reentrant domain, connected by internal and external polypeptides loops (7). The S4 segments contain 4-8 positively charged residues, which serve as voltage sensors. An outward movement of these charges under depolarization underlies the activation of the channel (8). Inactivation is mediated by the short intracellular loop connecting domains III and IV. One α subunit is associated with one or two auxiliary β subunits. These auxiliary subunits modulate channel gating, interact with extracellular matrix and play a role as cell adhesion molecules (9). Nav1.5 is much less sensitive to inhibition by tetrodotoxin (K0.5 ∼ 2 μM) and is more sensitive to Cd2+ block than neuronal or skeletal muscle Na+ channels. There may also be neuronal Na+ channels in cardiac myocytes (10) but their contribution to the total Na+ current is small (<15%) and their role remains controversial.

Most cardiac Na+ channels inactivate within a few milliseconds. However, a small percentage of channels either do not close, or close and then reopen, generating a slowly inactivating, persistent Na+ current, INa,slow (11,12). This persistent Na+ current is more sensitive to block by tetrodotoxin, is activated at more negative potentials and has an amplitude up to 0.5% of the peak transient INa. Despite the small amplitude, INa,slow may prolong the AP and generate a substantial Na+ influx. Some mutations in the Na+ channel gene SCN5A result in an increased INa,slow, which leads to long QT syndrome (LQT3). An increased INa,slow was found in ventricular myocytes from a canine ischemic model of HF (13) and more recently in dogs with pacing-induced HF (14). The latter study also reports an increased slowly inactivating Na+ current in human HF. Thus, INa,slow might be partly responsible for the prolongation of the AP and the rise in [Na+]i that occur in HF. We have found that a tetrodotoxin-sensitive pathway is responsible for the increased [Na+]i in a rabbit HF model (5). Genetic (15) or pharmacologically-induced (16) increases in INa,slow were shown to result in delayed afterdepolarizations and triggered arrhythmias. Moreover, ranolazine, an INa,slow inhibitor, reduced [Na+]i and diastolic Ca2+ overload and improved diastolic function in failing human hearts (17). Interestingly, Wagner et al. (18) found that overexpression of CaMKII slows fast INa inactivation, enhances INa,slow and increases [Na+]i in cardiac myocytes. CaMKII also slows recovery from inactivation and reduces steady-state channel availability at physiological diastolic membrane potential (18). Thus, CaMKII causes both gain and loss of Na+ channels function in a manner similar to that produced by a human mutant Na+ channel (Asp insertion at 1795 in the C-terminus), which causes both long-QT (at low heart rates) and Brugada syndrome (at high heart rates) (19). CaMKII expression and activity is augmented in HF (20,21). Thus, the acquired Na+ channel dysfunction may affect millions of patients with heart failure.

In addition to normal biophysical properties, Na+ channels activity requires proper sarcolemmal localization. Immunofluorescence measurements revealed that Nav1.5 channels are located both in the T-tubules (TT) and external sarcolemma (ESL) (22,23), although in some studies they were found mostly at the intercalated disks (23). Functional studies indicated a relatively uniform Nav1.5 distribution between the TT and ESL (24). It has been recently shown that ankyrin-G, an adapter protein that links membrane proteins to the cytoskeleton, plays an important role in the proper membrane targeting of Na+ channels. Lowe et al. (25) showed that reduced ankyrin-G expression decreases total cellular Nav1.5 expression and its efficient membrane localization, as well as INa. Moreover, a mutation in the ankyrin-binding motif of Nav1.5 that abolishes the binding of ankyrin-G has been shown to induce Brugada syndrome (26).

Na+/Ca2+ exchanger

NCX1 is the only NCX isoform present in the heart and consists of 970 amino acids (∼110 kDa). Topological data suggest that NCX1 has nine transmembrane segments (27,28) and a large intracellular loop (∼550 amino acids) between TMS 5-6, with the N- and C-termini located on the external and internal sides, respectively. There are also two regions of internal repeats (α1 and α2) facing opposite sides of the membrane. The α1 repeat comprises part of the TMS 2 and 3 and the reentrant loop between them while the α2 repeat contains part of the TMS 7, the cytosolic loop connecting it to the TMS 8 and part of the TMS 8. Cross-linking experiments indicate that the α repeats are near one another in the folded protein, suggesting that they may contribute to the ion conduction pathway (29,30). The large cytosolic loop contains two separate Ca2+ binding domains CBD1 and CBD2 (also known as β-repeats) that are essential for the allosteric regulation by Ca2+ (see below and Refs. (31-33)). There is also a charged segment of 20 residues (219-238) at the N-end of the cytosolic loop (XIP) which resembles a calmodulin binding domain and may be autoinhibitory (since a peptide based on this sequence inhibits NCX activity).

Besides being transport substrates, intracellular Ca2+ and Na+ exert important modulatory effects on NCX activity. Increases in cytosolic Ca2+ allow Ca2+ binding to the CBD domains, which increases the activity of the exchanger (allosteric [Ca2+]i-dependent activation). Weber et al. (34) found a physiologically relevant Km(Ca) of 125 nM in intact ferret ventricular myocyte during dynamic [Ca2+]i changes. Allosteric Ca2+-activation develops rapidly (within tens of ms). However, once activated, NCX does not revert to an inactivated state on return of [Ca2+]i to rest levels for tens of seconds (35,36). This suggests that Ca2+-dependent activation may only gradually change during alterations in heart rate, but it is undoubtedly substantially activated during basal heart rates. High [Na+]i inhibits NCX (Na+-dependent inactivation) in a time and [Na+]i-dependent manner. Because of the high [Na+]i (> 30 mM) required, Na+-dependent inactivation may not be very important under normal physiological conditions. However, it could prevent excess Ca2+ influx and cellular Ca2+ overload under pathophysiological conditions of high [Na+]i, such as ischemia/ reperfusion, where net Ca2+ influx via NCX might be strongly favored. [Na+]i-dependent inactivation is eliminated by an increase in membrane PIP2, which interacts with the XIP region of the exchanger (37). Recent evidence indicates that PIP2 levels may also modulate NCX by affecting its membrane trafficking (38).

NCX is concentrated at the T-Tubules (39-41), however it is not known how much is at the junctions with the SR. Functional data indicate that Ca2+ entry via reverse-mode NCX can trigger Ca2+ release from the SR at least under certain conditions (42-44). However, only a small NCX fraction seems to co-localize with proteins specific to the dyadic cleft, i.e. L-type Ca2+ channels and RyRs (45).

Na+/K+-ATPase

NKA extrudes three Na+ ions in exchange for two K+ ions using the energy derived from the hydrolysis of one ATP molecule, and thus moves out one net charge per cycle. NKA has two essential subunits: α and β. The NKA-α subunit (∼ 110 kDa) contains the binding sites for Na+, K+, ATP and cardiac glycosides and its crystal structure has been recently uncovered (46). NKA-α has 10 transmembrane domains with the N- and C- termini located on the cytosolic side of the membrane. The extracellular loops are short, with the exception of the TM7 - TM8 loop. NKA-α has a large intracellular loop between TM4 and TM5, composed of about 430 amino acid residues, a long N-terminal tail of about 90 amino acid residues, and an intracellular loop of about 120 residues between TM2 and TM3. The smaller β subunit (∼50 kDa, one transmembrane domain) has a role in the processing and in the proper membrane insertion of the pump.

There are three α (α1-α3) and three β (β1-β3) NKA subunit isoforms in the heart, with any αβ combination resulting in a functional pump. NKA-α1 is present in the heart of all species studied, while the expression of NKA-α2 and NKA-α3 differs significantly between species (47). For instance, the fetal and neonatal rodent hearts express NKA-α3, and this is replaced by NKA-α2 early in development (48). Interestingly, the reverse switch occurs in hyperthrophied or failing rat heart (49,50). All three NKA-α isoforms can be detected in the human heart (51).

It has been suggested that different NKA isoforms may function differently in the cell, depending on specific localization in the membrane. Thus, it was proposed that NKA-α2 and NKA-α3 are located mainly in the T-tubules, at the junctions with the SR, where they could regulate local NCX and [Ca2+]i. There is rather convincing evidence supporting such a model in smooth muscle (52). However, things are less clear in the heart. James et al. (53) showed that mouse hearts with genetically reduced levels of NKA-α2 were hypercontractile as a result of increased Ca2+ transients. In contrast, hearts with reduced levels of NKA-α1 were hypocontractile. Moreover, inhibition of NKA-α2 with ouabain increased the contractility of heterozygous NKA-α1 hearts. These data indirectly implicate a selective involvement of NKA-α2 in cardiac myocyte Ca2+ regulation. Further support for this idea came from the observation that increased expression of NKA-α2 (at the expense of NKA-α1) decreased the reverse-mode Na+/Ca2+ exchange current and Ca2+ transients in mouse ventricular myocytes (54). Dostanic et al. (55) showed that reducing the ouabain affinity of mouse NKA-α2 leads to a loss of glycoside-induced inotropy, which suggests that NKA-α2 regulates local [Ca2+]i near the Na+/Ca2+ exchange, as in smooth muscle. However, the same authors later found that NKA-α1 is also functionally and physically coupled with NCX in cardiac myocytes (56).

The differential NKA-α1/NKA-α2 localization (TT vs. ESL) in cardiac myocytes is controversial. Using immunofluorescence, McDonough et al. (47) found that NKA-α1 is preferentially distributed in the TT, while NKA-α2 was uniformly distributed in the TT and ESL in the rat. Conversely, Silverman et al. (57) showed that in guinea-pig ventricular myocytes NKA-α2 is mainly in TT and NKA-α1 is predominantly on the ESL. Recent studies using formamide-induced detubulation of ventricular myocytes and measurements of dose-dependent inhibition of NKA current by ouabain showed that the functional density of NKA-α2 is significantly higher (about 4-fold) in the TT (vs. ESL) while NKA-α1 is more uniformly distributed in cardiac myocytes from both rats (58,59) and mice (60). Moreover, we (58) found that the amount of NKA-α1 and NKA-α2 in the TT is comparable. This raises the intriguing possibility that NKA-α2 is indeed localized to sarcolemmal junctions with the SR (48% of the TT area and ∼8% of ESL in rat ventricle, as measured by electron microscopy; Ref (61)) and could be important in regulating local cleft [Na+]i and [Ca2+]i and therefore excitation-contraction coupling as in smooth muscle. However, as discussed above this hypothesis requires further structural and functional tests in adult ventricular myocytes.

NKA regulation by phospholemman

Phospholemman (PLM) is a small (72 amino acids), single membrane-spanning sarcolemmal protein that has long been known as a major phosphorylation target for PKA and PKC in the heart (62,63). However, its physiological role was poorly understood. Initially it was thought that PLM forms ion channels selective for taurine and thus might be involved in cell volume regulation (64). More insight came with the finding that PLM belongs to a family of proteins that associate with and modulate NKA in various tissues (65). The protein family, named FXYD after a conserved Pro-Phe-X-Tyr-Asp motif in the extracellular N-terminus domain, also includes the NKA γ-subunit found in the kidney. PLM is the only FXYD protein highly expressed in the heart and is also unique in the family in having multiple phosphorylation sites at its cytosolic carboxyl terminus.

Recent evidence shows that PLM modulates NKA the same way phospholamban (PLB) modulates SERCA, a P-type pump closely related to NKA. That is, unphosphorylated PLM inhibits NKA, mostly by reducing its affinity for internal Na+ (57,66-68) and PLM phosphorylation relieves this inhibition (67-68). β-adrenergic agonists had no effect on NKA function in myocytes from PLM-KO mice (67). This indicates that phosphorylation of PLM, rather than direct NKA phosphorylation, mediates the enhancement of NKA function during sympathetic stimulation of the heart. This is in agreement with data showing that NKA-α subunit can be phosphorylated by PKA, but only in the presence of detergents, while in situ the phosphorylation site may be inaccessible to the kinase (69).

Interestingly, the physical association between PLM and NKA-α, as determined by co-immunoprecipitation experiments, appears to be unaffected by PKA phosphorylation (70,71). However, when fluorescence resonance energy transfer (FRET) was used to look for more subtle changes in the NKA-PLM association, robust FRET was observed at baseline and the FRET was dramatically decreased upon PLM phosphorylation by either PKA or PKC (72). Thus, PLM phosphorylation changes the PLM-NKA interaction but does not result in a complete dissociation. In the analogous PLB-SERCA system it was long thought that PLB phosphorylation caused PLB to dissociate from SERCA, but recently this has been challenged (73).

We have shown recently (74) that NKA activation, mediated by phosphorylation of PLM, limits the rise in [Na+]i and Ca2+ transient amplitude during β-adrenergic stimulation in cardiac myocytes. Enhancement of NKA activity may thus be an integral part of the sympathetic fight or flight response of the heart, by enhancing Na+ extrusion to better keep up with the higher level of Na+ influx (that is caused by the combined inotropic and chronotropic effects on the heart). We found (74) that the inability of NKA to be activated by β-adrenergic stimulation in the PLM-KO mouse leads to excessive elevation of [Na+]i during sympathetic activation, which has both the benefits and the risks associated with NKA inhibition by cardiac glycosides (inotropy, but enhanced arrhythmogenesis). Thus, the physiological role of PLM may be to prevent Ca2+ overload and triggered arrhythmias by limiting the rise of [Na+]i during sympathetic stimulation of the heart.

NKA and NCX regulation by ankyrin-B

Ankyrin-B is another member of the ankyrin family of adaptor proteins responsible for the proper localization of various membrane proteins at specialized membrane domains. In humans, loss-of-function mutations in ankyrin-B are associated with cardiac phenotypes including QT interval prolongation (LQT4 syndrome), bradycardia, idiopathic ventricular fibrillation and catecholaminergic polymorphic ventricular tachycardia (for a review, see Ref 75). Ankyrin-B has long been known to bind and target NKA to the membrane (76). More recently, it has been shown that the proper membrane localization and post-translational stability of NCX also require direct interaction with ankyrin-B (77). A decrease in ankyrin-B function resulted in reduced NKA and NCX targeting to the T-Tubules as well as reduced overall NKA and NCX protein level (78) in cardiac myocytes. This led to altered Ca2+ signaling and aftercontractions, which may cause the arrhythmias observed in vivo. However, whether ankyrin-B brings NCX and NKA close to each other in the sarcolemma, how alterations in ankyrin-B level and activity affect [Na+]i, cellular and SR Ca2+ cycling and the molecular mechanism leading to arrhythmias are still largely unknown.

Conclusion

The numerous Na+ transport pathways are critical individually in regulating cardiac [Na+]i. However, the integrated cellular handling of Na+ must be kept in mind in order to understand and predict how alterations in these pathways affect [Ca2+]i, contractility and electrical activity of cardiac myocytes.

Acknowledgments

Supported by grants from the National Institutes of Health HL-81526 and HL-64724 (DMB) and American Heart Association (0735084N to SD).

References

  • 1.Bers DM. Excitation-contraction coupling and cardiac contractile force. 2nd. Dordrecht, The Netherlands: Kluwer Academic Press; 2001. p. 427. [Google Scholar]
  • 2.Lee CO, Dagostino M. Effect of strophanthidin on intracellular Na+ ion activity and twitch tension on constantly driven canine cardiac Purkinje fibers. Biophys J. 1982;40:185–198. doi: 10.1016/S0006-3495(82)84474-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Pieske B, Maier LS, Piacentino V, 3rd, Weisser J, Hasenfuss G, Houser S. Rate dependence of [Na+]i and contractility in nonfailing and failing human myocardium. Circulation. 2002;106:447–453. doi: 10.1161/01.cir.0000023042.50192.f4. [DOI] [PubMed] [Google Scholar]
  • 4.Bers DM, Barry WH, Despa S. Intracellular Na+ regulation in cardiac myocytes. Cardiovasc Res. 2003;57:897–912. doi: 10.1016/s0008-6363(02)00656-9. [DOI] [PubMed] [Google Scholar]
  • 5.Despa S, Islam MA, Weber CR, Pogwizd SM, Bers DM. Intracellular Na+ Concentration is Elevated in Heart Failure, but Na/K-Pump Function is Unchanged. Circulation. 2002;105:2543–2548. doi: 10.1161/01.cir.0000016701.85760.97. [DOI] [PubMed] [Google Scholar]
  • 6.Baartscheer A, Schumacher CA, van Borren MM, Belterman CN, Coronel R, Fiolet JW. Increased Na+/H+-exchange activity is the cause of increased [Na+]i and underlies disturbed calcium handling in the rabbit pressure and volume overload heart failure model. Cardiovasc Res. 2003;57:1015–1024. doi: 10.1016/s0008-6363(02)00809-x. [DOI] [PubMed] [Google Scholar]
  • 7.Abriel H, Kass RS. Regulation of the voltage-gated cardiac sodium channel Nav1.5 by interacting proteins. TCM. 2005;15:35–40. doi: 10.1016/j.tcm.2005.01.001. [DOI] [PubMed] [Google Scholar]
  • 8.Stühmer W, Conti F, Suzuki H, Wang XD, Noda M, Yahagi N, Kubo H, Numa S. Structural parts involved in activation and inactivation of the sodium channel. Nature. 1989;339:597–603. doi: 10.1038/339597a0. [DOI] [PubMed] [Google Scholar]
  • 9.Nerbonne JM, Kass RS. Molecular physiology of cardiac repolarization. Physiol Rev. 2005;85:1205–1253. doi: 10.1152/physrev.00002.2005. [DOI] [PubMed] [Google Scholar]
  • 10.Haufe V, Chamberland C, Dumaine R. The promiscuous nature of the cardiac sodium current. J Mol Cell Cardiol. 2007;42:469–477. doi: 10.1016/j.yjmcc.2006.12.005. [DOI] [PubMed] [Google Scholar]
  • 11.Saint DA, Ju YK, Gage PW. A persistent sodium current in rat ventricular myocytes. J Physiol. 1992;453:219–231. doi: 10.1113/jphysiol.1992.sp019225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Maltsev VA, Sabbah HN, Higgins RS, Silverman N, Lesch M, Undrovinas AI. Novel, ultraslow inactivating sodium current in human ventricular cardiomyocytes. Circulation. 1998;98:2545–2552. doi: 10.1161/01.cir.98.23.2545. [DOI] [PubMed] [Google Scholar]
  • 13.Undrovinas AI, Maltsev VA, Sabbah HN. Repolarization abnormalities in cardiomyocytes of dogs with chronic heart failure: role of sustained inward current. Cell Mol Life Sci. 1999;55:494–505. doi: 10.1007/s000180050306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Valdivia CR, Chu WW, Pu J, Foell JD, Haworth RA, Wolff MR, Kamp TJ, Makielski JC. Increased late sodium current in myocytes from a canine heart failure model and from failing human heart. J Mol Cell Cardiol. 2005;38:475–483. doi: 10.1016/j.yjmcc.2004.12.012. [DOI] [PubMed] [Google Scholar]
  • 15.Fredj S, Lindegger N, Sampson KJ, Carmeliet P, Kass RS. Altered Na+ channels promote pause-induced spontaneous diastolic activity in long QT syndrome type 3 myocytes. Circ Res. 2006;99:1225–1232. doi: 10.1161/01.RES.0000251305.25604.b0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.Song Y, Shryock JC, Belardinelli L. An increase of late sodium current induces delayed afterdepolarizations and sustained triggered activity in atrial myocytes. Am J Physiol Heart Circ Physiol. 2008;294:H2031–H2039. doi: 10.1152/ajpheart.01357.2007. [DOI] [PubMed] [Google Scholar]
  • 17.Sossalla S, Wagner S, Rasenack ECL, Ruff H, Weber SL, Schondube FA, Tirilomis T, Tenderich G, Hasenfuss G, Belardinelli L, Maier LS. 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;45:32–43. doi: 10.1016/j.yjmcc.2008.03.006. [DOI] [PubMed] [Google Scholar]
  • 18.Wagner S, Dybkova N, Rasenack ECL, Jacobshagen C, Fabritz L, Kirchhof P, Maier SKG, Zhang T, Hasenfuss G, Brown JH, Bers DM, Maier LS. Ca2+/calmodulin-dependent protein kinase II regulates cardiac Na+ channels. J Clin Invest. 2006;116:3127–3138. doi: 10.1172/JCI26620. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Bezzina C, Veldkamp MW, van Den Berg MP, Postma AV, Rook MB, Viersma JW, van Langen IM, Tan-Sindhunata G, Bink-Boelkens MT, van Der Hout AH, Mannens MM, Wilde AA. A single Na+ channel mutation causing both long-QT and Brugada syndromes. Circ Res. 1999;85:1206–1213. doi: 10.1161/01.res.85.12.1206. [DOI] [PubMed] [Google Scholar]
  • 20.Kirchhefer U, Schmitz W, Scholz H, Neumann J. Activity of cAMP-dependent protein kinase and Ca2+/calmodulin-dependent protein kinase in failing and nonfailing human hearts. Cardiovasc Res. 1999;42:254–261. doi: 10.1016/s0008-6363(98)00296-x. [DOI] [PubMed] [Google Scholar]
  • 21.Ai X, Curran JW, Shannon TR, Bers DM, Pogwizd SM. Ca2+/Calmodulin-Dependent Protein Kinase Modulates Cardiac Ryanodine Receptor Phosphorylation and Sarcoplasmic Reticulum Ca2+ Leak in Heart Failure. Circ Res. 2005;97:1314–1322. doi: 10.1161/01.RES.0000194329.41863.89. [DOI] [PubMed] [Google Scholar]
  • 22.Scriven DR, Dan P, Moore ED. Distribution of proteins implicated in excitation-contraction coupling in rat ventricular myocytes. Biophys J. 2000;79:2682–2691. doi: 10.1016/S0006-3495(00)76506-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Kucera JP, Rohr S, Rudy Y. Localization of sodium channels in intercalated disks modulates cardiac conduction. Circ Res. 2002;91:1176–1182. doi: 10.1161/01.res.0000046237.54156.0a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Brette F, Orchard CH. Density and sub-cellular distribution of cardiac and neuronal sodium channel isoforms in rat ventricular myocytes. Biochem Biophys Res Commun. 2006;348:1163–1166. doi: 10.1016/j.bbrc.2006.07.189. [DOI] [PubMed] [Google Scholar]
  • 25.Lowe JS, Palygin O, Bhasin N, Hund TJ, Boyden PA, Shibata E, Anderson ME, Mohler PJ. Voltage-gated Nav channel targeting in the heart requires an ankyrin-G dependent cellular pathway. J Cell Biol. 2008;180:173–186. doi: 10.1083/jcb.200710107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Priori SG, Napolitano C, Gasparini M, Pappone C, Della Bella P, Brignole M, Giordano U, Giovannini T, Menozzi C, Bloise R, Crotti L, Terreni L, Schwartz PJ. Clinical and genetic heterogeneity of right bundle branch block and ST-segment elevation syndrome: A prospective evaluation of 52 families. Circulation. 2000;102:2509–2515. doi: 10.1161/01.cir.102.20.2509. [DOI] [PubMed] [Google Scholar]
  • 27.Nicoll DA, Ottolia M, Lu L, Lu Y, Philipson KD. A new topological model of the cardiac sarcolemmal Na+-Ca2+ exchanger. J Biol Chem. 1999;274:910–917. doi: 10.1074/jbc.274.2.910. [DOI] [PubMed] [Google Scholar]
  • 28.Nicoll DA, Ren X, Ottolia M, Phillips M, Paredes AR, Abramson J, Philipson KD. What we know about the structure of NCX1 and how it relates to its function. Ann N Y Acad Sci. 2007;1099:1–6. doi: 10.1196/annals.1387.014. [DOI] [PubMed] [Google Scholar]
  • 29.Qiu Z, Nicoll DA, Philipson KD. Helix packing of functionally important regions of the cardiac Na+-Ca2+ exchanger. J Biol Chem. 2001;276:194–199. doi: 10.1074/jbc.M005571200. [DOI] [PubMed] [Google Scholar]
  • 30.Ren X, Nicoll DA, Philipson KD. Helix packing of the cardiac Na+-Ca2+ exchanger: proximity of transmembrane segments 1, 2, and 6. J Biol Chem. 2006;281:22808–22814. doi: 10.1074/jbc.M604753200. [DOI] [PubMed] [Google Scholar]
  • 31.Levitsky DO, Nicoll DA, Philipson KD. Identification of the high affinity Ca2+-binding domain of the cardiac Na+-Ca2+ exchanger. J Biol Chem. 1994;269:22847–22852. [PubMed] [Google Scholar]
  • 32.Matsuoka S, Nicoll DA, Hrysko LV, Levitsky DO, Weiss JN, Philipson KD. Regulation of the cardiac Na+-Ca2+ exchanger by Ca2+. Mutational analysis of the Ca2+-binding domain. J Gen Physiol. 1995;105:403–420. doi: 10.1085/jgp.105.3.403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Mercado Besserer G, Ottolia M, Nicoll DA, Chaptal V, Cascio D, Philipson KD, Abramson J. The second Ca2+ - bonding domain of the Na+-Ca2+ exchanger is essential for regulation: Crystal structures and mutational analysis. Proc Natl Acad Sci. 2007;104:18467–18472. doi: 10.1073/pnas.0707417104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Weber CR, Ginsburg KS, Philipson KD, Shannon TR, Bers DM. Allosteric regulation of Na/Ca exchange current by cytosolic Ca in intact cardiac myocytes. J Gen Physiol. 2001;117:119–131. doi: 10.1085/jgp.117.2.119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Reeves JP, Condrescu M. Allosteric activation of sodium-calcium exchnage activity by calcium: persistence at low calcium concnetrations. J Gen Physiol. 2003;122:621–639. doi: 10.1085/jgp.200308915. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Bers DM, Ginsburg KS. Na:Ca stoichiometry and cytosolic Ca-dependent activation of NCX in intact cardiomyocytes. Ann NY Acad Sci. 2007;1099:326–338. doi: 10.1196/annals.1387.060. [DOI] [PubMed] [Google Scholar]
  • 37.He Z, Feng S, Tong Q, Hilgemann DW, Philipson KD. Interaction of PIP2 with the XIP region of the cardiac Na/Ca exchanger. Am J Physiol – Cell Physiol. 2000;278:C661–C666. doi: 10.1152/ajpcell.2000.278.4.C661. [DOI] [PubMed] [Google Scholar]
  • 38.Yaradanakul A, Feng S, Shen C, Lariccia V, Lin MJ, Yang J, Kang TM, Dong P, Yin HL, Albanesi JP, Hilgemann DW. Dual control of cardiac Na+/Ca2+ exchange by PIP2: electrophysiological analysis of direct and indirect mechanisms. J Physiol. 2007;582:991–1010. doi: 10.1113/jphysiol.2007.132712. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Frank HS, Mottino J, Reid D, Molday RS, Philipson KD. Distribution of the Na+-Ca2+ exchanger protein in mammalian cardiac myocytes; An immunofluorescence and immunoco-lloidal glod-labeling study. J Biol Chem. 1992;117:337–345. doi: 10.1083/jcb.117.2.337. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Yang Z, Pascarel C, Steele DS, Komukai K, Brette F, Orchard CH. Na+-Ca2+ exchange activity is localized in the T-tubules of rat ventricular myocytes. Circ Res. 2002;91:315–322. doi: 10.1161/01.res.0000030180.06028.23. [DOI] [PubMed] [Google Scholar]
  • 41.Despa S, Brette F, Orchard CH, Bers DM. Na/Ca exchange and Na/K-ATPase function are equally concentrated in transverse tubules of rat ventricular myocytes. Biophys J. 2003;85:3388–3396. doi: 10.1016/S0006-3495(03)74758-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Litwin SE, Li J, Bridge JH. Na-Ca exchange and the trigger for sarcoplasmic reticulum Ca release: studies in adult rabbit ventricular myocytes. Biophys J. 1998;75:359–371. doi: 10.1016/S0006-3495(98)77520-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Wasserstrom JA, Vites Am. The role of Na+-Ca2+ exchange in activation of excitation-contraction coupling in rat ventricular myocytes. J Physiol. 1996;493:529–542. doi: 10.1113/jphysiol.1996.sp021401. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Sipido KR, Maes M, Van de Werf F. Low efficiency of Ca2+ entry through the Na+-Ca2+ exchanger as trigger for Ca2+ release from the sarcoplasmic reticulum. A comparison between L-type Ca2+ current and reverse-mode Na+-Ca2+ exchange. Circ Res. 1997;81:1034–1044. doi: 10.1161/01.res.81.6.1034. [DOI] [PubMed] [Google Scholar]
  • 45.Scriven DRL, Dan P, Moore EDW. Distribution of proteins implicated in excitation-contraction coupling in rat ventricular myocytes. Biophys J. 2000;79:2682–2691. doi: 10.1016/S0006-3495(00)76506-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Morth JP, Pedersen BP, Toustrup-Jensen MS, Sorensen TLM, Petersen J, Andersen JP, Vilsen B, Nissen P. Crystal structure of the sodium-potassium pump. Nature. 2007;450:1043–1049. doi: 10.1038/nature06419. [DOI] [PubMed] [Google Scholar]
  • 47.McDonough AA, Zhang Y, Shin V, Frank JS. Subcellular distribution of Na pump isform subunits in mammalian cardiac myocytes. Am J Physiol. 1996;270:C1221–C1227. doi: 10.1152/ajpcell.1996.270.4.C1221. [DOI] [PubMed] [Google Scholar]
  • 48.Lucchesi PA, Sweadner KJ. Postnatal changes in Na,K-ATPase isoform expression in rat cardiac ventricle. J Biol Chem. 1991;266:9327–9331. [PubMed] [Google Scholar]
  • 49.Charlemagne D, Orlowski J, Oliviero P, Rannou F, Sainte Beuve C, Swynghedauw B, Lane LK. Alteration of Na,K-ATPase subunit mRNA and protein levels in hypertrophied rat heart. J Biol Chem. 1994;269:1541–1547. [PubMed] [Google Scholar]
  • 50.Semb SO, Lunde PK, Holt E, Tonnessen T, Christensen G, Sejersted OM. Reduced myocardial Na-K pump capacity in congestive heart failure following myocardial infarction in rats. J Mol Cell Cardiol. 1998;30:1311–1328. doi: 10.1006/jmcc.1998.0696. [DOI] [PubMed] [Google Scholar]
  • 51.Sweadner KJ, Herrera VLM, Amato S, Moellmann A, Gibbons DK, Repke KRH. Immunologic identification of Na,K-ATPase isoforms in myocardium. Circ Res. 1994;74:669–678. doi: 10.1161/01.res.74.4.669. [DOI] [PubMed] [Google Scholar]
  • 52.Juhaszova M, Blaustein MP. Na+ pump low and high ouabain affinity alpha subunit isoforms are differently distributed in cells. Proc Natl Acad Sci U S A. 1997;94:1800–1805. doi: 10.1073/pnas.94.5.1800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.James PF, Grupp IL, Grupp G, Woo AL, Askew GR, Croyle ML, Walsh RA, Lingrel JB. Identification of a specific role for the Na,K-ATPase α2 isoform as a regulator of calcium in the heart. Mol Cell. 1999;3:555–5563. doi: 10.1016/s1097-2765(00)80349-4. [DOI] [PubMed] [Google Scholar]
  • 54.Yamamoto T, Su Z, Moseley AE, Kadono T, Zhang J, Cougnon M, Li F, Lingrel JB, Barry WH. Relative abundance of α2 Na+ pump isoform influences Na+-Ca2+ exchanger currents and Ca2+ transients in mouse ventricular myocytes. J Mol Cell Cardiol. 2005;39:113–120. doi: 10.1016/j.yjmcc.2005.03.023. [DOI] [PubMed] [Google Scholar]
  • 55.Dostanic I, Lorenz JN, Schultz Jel J, Grupp IL, Neumann JC, Wani MA, Lingrell JB. The α2 isoform of Na,K-ATPase mediates ouabain-induced cardiac inotropy in mice. J Biol Chem. 2003;278:53026–53034. doi: 10.1074/jbc.M308547200. [DOI] [PubMed] [Google Scholar]
  • 56.Dostanic I, Schultz Jel J, Lorenz JN, Lingrell JB. The α1 isoform of Na,K-ATPase regulates cardiac contractility and functionally interacts and co-localizes with the Na/Ca exchanger in heart. J Biol Chem. 2004;279:54053–54061. doi: 10.1074/jbc.M410737200. [DOI] [PubMed] [Google Scholar]
  • 57.Silverman B, Fuller W, Eaton P, Deng J, Moorman JR, Cheung JY, James AF, Shattock MJ. Serine 68 phosphorylation of phospholemman: acute isoform-specific activation of cardiac Na/K ATPase. Cardiovasc Res. 2005;65:93–103. doi: 10.1016/j.cardiores.2004.09.005. [DOI] [PubMed] [Google Scholar]
  • 58.Despa S, Bers DM. Functional analysis of Na+/K+-ATPase isoform distribution in rat ventricular myocytes. Am J Physiol Cell Physiol. 2007;293:C321–C327. doi: 10.1152/ajpcell.00597.2006. [DOI] [PubMed] [Google Scholar]
  • 59.Swift F, Tovsrud N, Enger UH, Sjaastad I, Sejersted OM. The Na+,K+-ATPase α2-isoform regulates cardiac contractility in rat cardiomyocytes. Cardiovasc Res. 2007;75:109–117. doi: 10.1016/j.cardiores.2007.03.017. [DOI] [PubMed] [Google Scholar]
  • 60.Berry RG, Despa S, Fuller W, Bers DM, Shattock MJ. Differential distribution and regulation of mouse Na+/K+-ATPase α1 and α2-subunits in T-tubule and surface sarcolemmal membranes. Cardiovasc Res. 2007;73:92–100. doi: 10.1016/j.cardiores.2006.11.006. [DOI] [PubMed] [Google Scholar]
  • 61.Page E, Surdyk-Droske M. Distribution, surface density, and membrane area of diadic junctional contacts between plasma membrane and terminal cisterns in mammalian ventricle. Circ Res. 1979;45:260–267. doi: 10.1161/01.res.45.2.260. [DOI] [PubMed] [Google Scholar]
  • 62.Presti CF, Jones LR, Lindemann JP. Isoproterenol-induced phosphorylation of a 15-kilodalton sarcolemmal protein in intact myocardium. J Biol Chem. 1985;260:3860–3867. [PubMed] [Google Scholar]
  • 63.Palmer CJ, Scott BT, Jones LR. Purification and complete sequence determination of the major plasma membrane substrate for cAMP-dependent protein kinase and protein kinase C in myocardium. J Biol Chem. 1991;266:11126–11130. [PubMed] [Google Scholar]
  • 64.Moorman JR, Ackerman SJ, Kowdley GC, Griffin MP, Mounsey JP, Chen Z, Cala SE, O'Brian JJ, Szabo G, Jones LR. Unitary anion currents through phospholemman channel molecules. Nature. 1995;377:737–740. doi: 10.1038/377737a0. [DOI] [PubMed] [Google Scholar]
  • 65.Sweadner KJ, Rael E. The FXYD gene family of small ion transport regulators or channels: cDNA sequence, protein signature sequence, and expression. Genomics. 2000;68:41–56. doi: 10.1006/geno.2000.6274. [DOI] [PubMed] [Google Scholar]
  • 66.Crambert G, Fuzesi M, Garty H, Karlish S, Geering K. Phospholemman (FXYD1) associates with Na+,K+-ATPase and regulates its transport properties. Proc Natl Acad Sci U S A. 2002;99:11476–11481. doi: 10.1073/pnas.182267299. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Despa S, Bossuyt J, Han F, Ginsburg KS, Jia LG, Kutchai H, Tucker AL, Bers DM. Phospholemman -phosphorylation mediates the β-adrenergic effects on Na/K pump function in cardiac myocytes. Circ Res. 2005;97:252–259. doi: 10.1161/01.RES.0000176532.97731.e5. [DOI] [PubMed] [Google Scholar]
  • 68.Bibert S, Roy S, Schaer D, Horisberger JD, Geering K. Phosphorylation of phospholemman (FXYD1) by protein kinases A and C modulates distinct Na,K-ATPase isozymes. J Biol Chem. 2008;283:476–486. doi: 10.1074/jbc.M705830200. [DOI] [PubMed] [Google Scholar]
  • 69.Sweadner KJ, Feschenko MS. Predicted location and limited accessibility of protein kinase A phosphorylation site on Na+-K-ATPase. Am J Physiol. 2001;280:C1017–C1026. doi: 10.1152/ajpcell.2001.280.4.C1017. [DOI] [PubMed] [Google Scholar]
  • 70.Bossuyt J, Ai X, Moorman RJ, Pogwizd SM, Bers DM. Expression and phosphorylation the Na+-pump regulatory subunit phospholemman in heart failure. Circ Res. 2005;97:558–565. doi: 10.1161/01.RES.0000181172.27931.c3. [DOI] [PubMed] [Google Scholar]
  • 71.Fuller W, Eaton P, Bell JR, Shattock MJ. Ischemia-induced phosphorylation of phospholemman directly activates rat cardiac Na+/K+-ATPase. FASEB J. 2004;18:197–199. doi: 10.1096/fj.03-0213fje. [DOI] [PubMed] [Google Scholar]
  • 72.Bossuyt J, Despa S, Martin JL, Bers DM. Phospholemman Phosphorylation alters its association with the Na-pump as assessed by FRET. J Biol Chem. 2006;281:32765–32773. doi: 10.1074/jbc.M606254200. [DOI] [PubMed] [Google Scholar]
  • 73.Karim CB, Zhang Z, Howard EC, Torgersen KD, Thomas DD. Phosphorylation-dependent conformational switch in spin-labeled phospholamban bound to SERCA. J Mol Biol. 2006;358:1032–1040. doi: 10.1016/j.jmb.2006.02.051. [DOI] [PubMed] [Google Scholar]
  • 74.Despa S, Tucker AL, Bers DM. PLM-mediated activation of Na/K – ATPase limits [Na]i and inotropic state during β-adrenergic stimulation in mouse ventricular myocytes. Circulation. 2008;117:1849–1855. doi: 10.1161/CIRCULATIONAHA.107.754051. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Cunha SR, Mohler PJ. Cardiac ankyrins: Essential components for development and maintenance of excitable membrane domains in heart. Cardiovasc Res. 2006;71:22–29. doi: 10.1016/j.cardiores.2006.03.018. [DOI] [PubMed] [Google Scholar]
  • 76.Nelson WJ, Veshnock PJ. Ankyrin binding to Na+-K+ ATPase and implications for the organization of membrane domains in polarized cells. Nature. 1987;328:533–536. doi: 10.1038/328533a0. [DOI] [PubMed] [Google Scholar]
  • 77.Cunha SR, Bhasin N, Mohler PJ. Targeting and stability of Na/Ca exchanger 1 in cardiomyocytes requires direct interaction with the membrane adaptor protein ankyrin-B. J Biol Chem. 2007;282:4875–4883. doi: 10.1074/jbc.M607096200. [DOI] [PubMed] [Google Scholar]
  • 78.Mohler PJ, Schott JJ, Gramolini AO, Dilly KW, Guatimosin S, duBell WH, Song LS, Haurogne K, Kyndt F, Ali ME, Rogers TB, Jederer WJ, Escande D, Le Marec H, Bennett V. Ankyrin-B mutation causes type 4 long-QT cardiac arrhythmia and sudden cardiac death. Nature. 2003;421:634–639. doi: 10.1038/nature01335. [DOI] [PubMed] [Google Scholar]

RESOURCES