Elsevier

Progress in Neurobiology

Volume 82, Issue 6, August 2007, Pages 287-347
Progress in Neurobiology

Electrogenic Na+/Ca2+-exchange of nerve and muscle cells

This work is dedicated to Professor Tibor Török, Sc.D. (1914–1999).
https://doi.org/10.1016/j.pneurobio.2007.06.003Get rights and content

Abstract

The plasma membrane Na+/Ca2+-exchanger is a bi-directional electrogenic (3Na+:1Ca2+) and voltage-sensitive ion transport mechanism, which is mainly responsible for Ca2+-extrusion. The Na+-gradient, required for normal mode operation, is created by the Na+-pump, which is also electrogenic (3Na+:2K+) and voltage-sensitive. The Na+/Ca2+-exchanger operational modes are very similar to those of the Na+-pump, except that the uncoupled flux (Na+-influx or -efflux?) is missing. The reversal potential of the exchanger is around −40 mV; therefore, during the upstroke of the AP it is probably transiently activated, leading to Ca2+-influx. The Na+/Ca2+-exchange is regulated by transported and non-transported external and internal cations, and shows ATPi-, pH- and temperature-dependence. The main problem in determining the role of Na+/Ca2+-exchange in excitation-secretion/contraction coupling is the lack of specific (mode-selective) blockers. During recent years, evidence has been accumulated for co-localisation of the Na+-pump, and the Na+/Ca2+-exchanger and their possible functional interaction in the “restricted” or “fuzzy space.” In cardiac failure, the Na+-pump is down-regulated, while the exchanger is up-regulated. If the exchanger is working in normal mode (Ca2+-extrusion) during most of the cardiac cycle, upregulation of the exchanger may result in SR Ca2+-store depletion and further impairment in contractility. If so, a normal mode selective Na+/Ca2+-exchange inhibitor would be useful therapy for decompensation, and unlike CGs would not increase internal Na+. In peripheral sympathetic nerves, pre-synaptic α2-receptors may regulate not only the VSCCs but possibly the reverse Na+/Ca2+-exchange as well.

Introduction

The electrogenic and voltage-sensitive Na+/Ca2+-exchange was discovered nearly 40 years ago in nerve (Baker et al., 1967a, Baker et al., 1967b, Baker et al., 1969a, Baker and Blaustein, 1968) and muscle (Reuter and Seitz, 1967, Reuter and Seitz, 1968).

Earlier Lüttgau and Niedergerke (1958) had shown that a reduction of external Na+ ([Na+]o; Li+, choline+ or sucrose substitution) increased the contractility of frog ventricle in an external Ca2+-([Ca2+]o)-dependent manner and suggested a competition between the two ions for an extracellular binding site. They also suggested that the carrier (if any) is voltage-dependent. Indeed, flux measurements later demonstrated the interaction between [Na+]o and [Ca2+]o in frog ventricle muscle cells (Niedergerke, 1963).

The primary role of Na+/Ca2+-exchange is to extrude Ca2+ after excitation. For Ca2+-extrusion the energy of the Na+-gradient is required which is created by the Na+-pump. However, the exchanger can also work in reverse mode as well, i.e. internal Na+ ([Na+]i) can exchange for [Ca2+]o. Thus, if the Na+-pump is inhibited, the elevated level of Na+-inside may increase the influx of Ca2+ through the exchanger.

The Na+/Ca2+-exchange protein is expressed in many cells and during the past years several excellent reviews (Baker, 1972, Baker, 1986, Baker and Allen, 1984, Baker and Reuter, 1975, Blaustein, 1974, Blaustein, 1979, Blaustein, 1989, Requena and Mullins, 1979; Blaustein et al., 1991a, Blaustein et al., 1991b, Blaustein and Lederer, 1999, DiPolo and Beaugé, 1983, DiPolo and Beaugé, 1990, DiPolo and Beaugé, 1993, Philipson, 1985, Philipson, 1990, Philipson, 1999, Philipson, 2002, Philipson and Nicoll, 1992, Philipson and Nicoll, 1993, Philipson and Nicoll, 2000, Philipson et al., 1993, Reeves, 1985, Reeves, 1992, Reeves, 1998, Carafoli, 1985, Eisner and Lederer, 1985, Hilgemann, 1997, Khananshvilli, 1998, Schnetkamp, 1989, Lagnado and McNaughton, 1990, Bers, 2000, Annunziato et al., 2004, DiPolo and Beaugé, 2006) and Proceedings of International Conferences (Allen et al., 1989, Blaustein et al., 1991a, Hilgemann et al., 1996) have been published on Na+/Ca2+-exchange.

This current work concentrates mainly on the exchanger of nerve and muscle cells. Since the literature contains more than 2000 papers about Na+/Ca2+-exchange, I apologise to all of the authors whose significant and important work has not been cited here.

Section snippets

Stoichiometry and voltage-sensitivity of Na+/Ca2+-exchange

Table 1 summarizes the coupling-ratios of the exchanger published in the literature. In squid giant axons, a 3:1 stoichiometry has been determined by several laboratories (but see: Jundt et al., 1975). Cardiac muscle cells from different species, including dogs, rabbits, guinea-pigs, cows and rats also have a coupling ratio of 3:1 (but see: Ledvora and Hegyvary, 1983, Matsuoka, 2002, Dong et al., 2002) and, arterial smooth muscle cells, pancreatic beta-cells, barnacle muscle fibres and

Inward and outward Na+/Ca2+-exchange currents under physiological and pathological conditions

Since the exchanger is electrogenic, it generates a current through the membrane in the direction of the net charge carried by Na+. In normal mode of operation, it generates an inward current, thus it may prolong the duration of the AP. In reverse mode however, it produces an outward current, thereby repolarising the membrane and may shorten the duration of the AP. In this latter case however, the entering Ca2+ may produce activation of nerve and muscle cells in spite of the shortened AP.

Table 2

Number of Na+/Ca2+-exchange molecules (density of exchanger)

Since the Na+/Ca2+-exchanger and the Na+-pump are co-localised and functionally interact (see Section 9.1), it is worthwhile to make a comparison between them. The ATP-driven Ca2+-pump shows a uniform distribution in the plasma membrane and has a high affinity for Ca2+i (about 10 times higher to that of exchanger; DiPolo, 1979, Baker and DiPolo, 1984), but its turnover rate is relatively slow (around 50–150 s−1; Caroni and Carafoli, 1981, Schatzmann, 1983, Carafoli, 1983, Carafoli, 1987,

Turnover rate of Na+/Ca2+-exchange and other ion transporters

Most of the data published in the literature suggest that the Na+/Ca2+ exchanger has a much higher turnover rate than other membrane ion-extrusion transporters (e.g. Na+-pump: 80–200 Na+ s−1, Stein, 1986, Gadsby and Nakao, 1989, Friedrich et al., 1996; Ca2+-pump: ∼102 Ca2+ s−1, Stein, 1986).

Table 4 summarizes the values found in the literature.

In cardiac sarcolemmal vesicles Cheon and Reeves (1988) have determined a value around ∼1000 s−1 (see also: Reeves, 1989, Reeves, 1998).

Hilgemann et al.

Reversal potential (ENa/Ca) and driving force (ΔVNa/Ca) of Na+/Ca2+-exchange

The thermodynamic driving force for Na+/Ca2+-exchange can be expressed in terms of its reversal potential, i.e. the membrane potential at which the exchange system is in equilibrium:ENa/Ca=3ENa2ECa=RTF1ln{([Ca2+]o/[Ca2+]i)([Na+]i/[Na+]o)3},where ENa/Ca is the reversal potential, ENa and ECa are the equilibrium potentials for Na+ and Ca2+ determined by the Nernst equation, R the gas constant (8.314(1) J/kmol), T the absolute temperature (273 °C); (at 37 °C: 310 °C), and F the Faraday's constant

Pharmacology of Na+/Ca2+-exchange; putative inhibitors

The main problem for studying the physiological role of reverse Na+/Ca2+-exchange in “excitation-secretion/contraction coupling” is that no selective blocker is yet available (Siegl et al., 1984, Slaughter et al., 1988, Egger and Niggli, 1999; cf. Blaustein and Lederer, 1999). Mode-selective blockers (forward- and reverse-mode) are required. Currently used pharmacological tools will be described below.

Tetrodotoxin-sensitive Na+-influx-induced outward Na+/Ca2+-exchange current contributes to [Ca2+]i-transient and SR Ca2+-release (CICR) when ICa(L) is blocked

The data cited in this section differ, presumably depending on the species and the experimental conditions used. The fact that the exchanger, in comparison with other ion transporters, can readily reverse, supports the possibility that the reverse mode of operation may pertain under certain conditions in vivo. In fact, a crucial question is, the localisation of the Na+-channel and the Na+/Ca2+-exchanger. If Na+-channels are located close to the exchanger, the chance of exchanger mediated Ca2+

Na+-pump and Na+/Ca2+-exchange

The primary transporter, the Na+-pump is an electrogenic (3Na+:2K+) and voltage-sensitive ion transporter similar in these respects to the Na+/Ca2+-exchanger (cf. Baker, 1966, Baker and Connelly, 1966, Rang and Ritchie, 1968, Baker et al., 1969b, Kerkut and York, 1971, Thomas, 1972, Cavieres and Ellory, 1974, Glynn and Karlish, 1975, Akera, 1977, Akera and Brody, 1978, Glitsch, 1982, Glitsch, 2001, Hansen, 1984, De Weer, 1984, De Weer and Rakowski, 1984, Gadsby, 1984, Gadsby et al., 1985, Nakao

High voltage-activated Ca2+-channels involved in neurotransmitter release

Ca2+-entry through voltage-sensitive Ca2+-channels (VSCCs) triggers neurotransmitter release from presynaptic nerve terminals (Katz, 1969, Katz and Miledi, 1969, Nachshen and Blaustein, 1982, Augustine et al., 1985, Augustine et al., 1987, Augustine et al., 1991, Augustine and Charlton, 1986, Tsien et al., 1988, Almers, 1990, Mulkey and Zucker, 1991, Llinás et al., 1992, Randall and Tsien, 1995, Dunlap et al., 1995, Stanley, 1997, Bennett, 1999).

The “N-type” Ca2+-channel, first described in

Na+-pump and Na+/Ca2+-exchange regulation by adrenoceptors

Na+,K+-ATPase can be stimulated by catecholamines both in peripheral and central neurones (cf. Phillis and Wu, 1981).

In the heart, several studies have suggested that β-receptor activation can increase the activity of the Na+-pump (Wasserstrom et al., 1982, Lee and Vassalle, 1983, Désilets and Baumgarten, 1986, Sheu et al., 1991, Gao et al., 1992, Clausen, 1996, Overgaard et al., 1999). Other studies however, have failed to find evidence for stimulation (Glitsch et al., 1989, Ishizuka and

Conclusions

Selective blockers, even mode-selective blockers would be required to decide the physiological role of Na+/Ca2+-exchange in “excitation-secretion/contraction coupling”. In nerve terminals, e.g. in peripheral sympathetic nerves, the vesicle-docking sites and the exchanger proteins may localise close to each other. Since the nerve varicosities are so small, reverse Na+/Ca2+-exchange activation may occur physiologically, especially when AP frequency is high. Calculations based on the reversal

Acknowledgements

I thank Professor A.F. Brading (Department of Pharmacology, University of Oxford) for critical reading and discussion of the manuscript.

This work was supported, in part by The National Scientific Research Foundation (OTKA, Grant nos. 1020; T017749; TS040736; T042595), The Health Science Council (ETT, Grant nos. 237/96; 141/2003), Semmelweis University, (ETK-SOTE, Grant no. G6) and by a Grant from The Hungarian Academy of Sciences provided to the “Neurochemical Research Group of Semmelweis

References (902)

  • D.M. Bers et al.

    Sodium–calcium exchange in sidedness of isolated cardiac sarcolemmal vesicles

    Biochim. Biophys. Acta

    (1980)
  • M.M. Bersohn et al.

    Effect of temperature on sodium–calcium exchange in sarcolemma from mammalian and amphibian hearts

    Biochim. Biophys. Acta

    (1991)
  • D.J. Beuckelmann et al.

    Characteristic of calcium-current in isolated human ventricular myocytes from patients with terminal heart failure

    J. Mol. Cell. Cardiol.

    (1991)
  • F.V. Bielen et al.

    Changes of the subsarcolemmal Na+ concentration in internally perfused cardiac cells

    Biochim. Biophys. Acta

    (1991)
  • M.P. Blaustein

    Effects of internal and external cations and ATP on sodium–calcium and calcium–calcium exchage in squid axons

    Biophys. J.

    (1977)
  • M.P. Blaustein

    The role of calcium in catecholamine release from adrenergic nerve terminals

  • M.P. Blaustein

    The cellular basis of cardiotonic steroid action

    Trends Pharmacol. Sci.

    (1985)
  • M.P. Blaustein

    Sodium–calcium exchange in cardiac, smooth, and skeletal muscles: key to control of contractility. Curr. Top. Membr. Transp. vol. 34

  • M.P. Blaustein et al.

    Carrier-mediated, sodium-dependent, and calcium-dependent calcium-efflux from pinched-off presynaptic nerve terminals (synaptosomes) in vitro

    Biochim. Biophys. Acta

    (1976)
  • C. Aalkjaer et al.

    Effect of ouabain on tone, membrane potential and sodium efflux compared with [3H]ouabain binding in rat resistance vessels

    J. Physiol. (Lond.)

    (1985)
  • S. Adachi-Akahane et al.

    Calcium signaling in transgenic mice overexpressing Na+–Ca2+-exchanger

    J. Gen. Physiol.

    (1997)
  • C.C. Aickin et al.

    An investigation of sodium-calcium exchange in the smooth muscle of guinea-pig ureter

    J. Physiol. (Lond.)

    (1987)
  • T. Akasu et al.

    Reduction of N-type calcium current by noradrenaline in neurones of rabbit vesical parasympathetic ganglia

    J. Physiol. (Lond.)

    (1990)
  • T. Akera

    Membrane adenosine-triphosphatase: a digitalis receptor?

    Science N. Y.

    (1977)
  • T. Akera et al.

    The role of Na+–K+-ATPase in the inotropic action of digitalis

    Pharmacol. Rev.

    (1978)
  • D.G. Allen et al.

    The effects of low sodium solutions on intracellular calcium concentration in ferret ventricular muscle

    J. Physiol. (Lond.)

    (1983)
  • D.G. Allen et al.

    Sodium pump: birthday present for digitalis

    Nature

    (1985)
  • T.J.A. Allen et al.

    Intracellular Ca indicator Quin-2 inhibits Ca2+ inflow via Nai–Cao exchange in squid axon

    Nature

    (1985)
  • T.J.A. Allen et al.

    Comparison of the effects of potassium and membrane potential on the calcium-dependent sodium efflux in squid axons

    J. Physiol. (Lond.)

    (1986)
  • T.J.A. Allen et al.

    Influence of membrane potential on calcium efflux from giant axons of loligo

    J. Physiol. (Lond.)

    (1986)
  • W. Almers

    Exocytosis

    Annu. Rev. Physiol.

    (1990)
  • E. Alnaes et al.

    On the role of mitochondria in transmitter release from motor nerve terminals

    J. Physiol. (Lond.)

    (1975)
  • J. Altamirano et al.

    The inotropic effect of cardioactive glycosides in ventricular myocytes requires Na+-Ca2+-exchanger function

    J. Physiol. (Lond.)

    (2006)
  • M.S. Amran et al.

    Pharmacology of KB-R7943: a Na+–Ca2+ exchange inhibitor

    Cardiovasc. Drug Rev.

    (2003)
  • L. Annunziato et al.

    Pharmacology of brain Na+/Ca2+ exchanger: from molecular biology to therapeutic perspectives

    Pharmacol. Rev.

    (2004)
  • M. Arai et al.

    Alterations in sarcoplasmic reticulum gene expression in human heart failure. A possible mechanism for alterations in systolic and diastolic properties for the failing myocardium

    Circ. Res.

    (1993)
  • M. Arai et al.

    Sarcoplasmic reticulum gene expression in cardiac hypertrophy and heart failure

    Circ. Res.

    (1994)
  • E. Armon et al.

    Light activated electrogenic Na+–Ca2+ exchange in fly photoreceptors: modulation by Na+/K+-pump activity

    Biophys. Struct. Mech.

    (1983)
  • A.A. Armoundas et al.

    Role of sodium–calcium exchanger in modulating the action potential of ventricular myocytes from normal and failing hearts

    Circ. Res.

    (2003)
  • A. Arnon et al.

    Ouabain augments Ca2+-transients in arterial smooth muscle without raising cytosolic Na+

    Am. J. Physiol., Heart Circ. Physiol.

    (2000)
  • M. Artman et al.

    Na+/Ca2+-exchange current density in cardiac myocytes from rabbits and guinea pigs during postnatal development

    Am. J. Physiol.

    (1995)
  • E. Arystarkova et al.

    Na,K-ATPase extracellular surface probed with a monoclonal antibody enhances ouabain binding

    J. Biol. Chem.

    (1992)
  • C.C. Ashley et al.

    Calcium movements in single crustacean muscle fibres

    J. Physiol. (Lond.)

    (1974)
  • G.J. Augustine et al.

    The calcium signal for transmitter secretion from presynaptic nerve terminals. (In: Calcium entry and action at the presynaptic nerve terminal)

    Ann. N. Y. Acad. Sci.

    (1991)
  • G.J. Augustine et al.

    Calcium dependence of pre-synaptic calcium current and post-synaptic response at the giant squid synapse

    J. Physiol. (Lond.)

    (1986)
  • G.J. Augustine et al.

    Calcium entry and transmitter release at voltage-clamped nerve terminals of squid

    J. Physiol. (Lond.)

    (1985)
  • G.J. Augustine et al.

    Calcium action in synaptic transmitter release

    Annu. Rev. Neurosci.

    (1987)
  • P.H. Axelsen et al.

    Electrochemical ion gradients and the Na/Ca exchange stoichiometry. Measurements of these gradients are thermodynamically consistent with a stoichiometric coefficient greater than or equal to 3

    J. Gen. Physiol.

    (1985)
  • B.A. Bailey et al.

    Sarcoplasmic reticulum-related changes in cytosolic calcium in pressure-overload-Induced feline LV hypertrophy

    Am. J. Physiol.

    (1993)
  • Cited by (35)

    • Sodium-calcium exchanger-3 regulates pain “wind-up”: From human psychophysics to spinal mechanisms

      2022, Neuron
      Citation Excerpt :

      This latter effect was likely a consequence of progressive depolarization on current injection and hence a depolarizing block. The probable cause for this is that NCX is electrogenic and voltage sensitive with a reversal potential around −40 mV (Török, 2007). As membrane potential increases more than −40 mV, NCX3 would switch from forward to reverse mode (extrusion of three Na+ ions and influx of one Ca2+), and this would oppose further membrane depolarization.

    • Cracking the code of sodium/calcium exchanger (NCX) gating: Old and new complexities surfacing from the deep web of secondary regulations

      2020, Cell Calcium
      Citation Excerpt :

      This is to avoid unpredictable (and often deleterious) responses of target and not-intentionally targeted cells. The first evidence that gating reactions impose constrains to the unidirectional modes of the rheogenic exchange processes was clearly provided by landmark works on cardiac and squid preparations [18,19,27]. In both systems, intracellular ions (Na+, Ca2+, H+) and ATP-dependent reactions significantly modify the kinetics of Na+/Ca2+ exchange and the affinities of the carrier for its transported ions [15,19,28–31].

    • Novel coupling between TRPC-like and K<inf>Na</inf> channels modulates low threshold spike-induced afterpotentials in rat thalamic midline neurons

      2014, Neuropharmacology
      Citation Excerpt :

      Previously it has been reported that all three known NCX genes (NCX1, NCX2 and NCX3) are expressed in PVT neurons (Papa et al., 2003). Moreover, SN-6 preferentially inhibits when the NCX operates in the reverse mode (Iwamoto et al., 2004), as could happen after Na+ replacement with either Li+ or choline+ (Török, 2007). Under these conditions, Na+ efflux and Ca2+ influx could lead to compensatory activation of TRPC channels and subsequent prolongation of the LTS-induced sADP.

    • GABA transporters mediate glycine release from cerebellum nerve endings: Roles of Ca<sup>2+</sup>channels, mitochondrial Na<sup>+</sup>/Ca<sup>2+</sup> exchangers, vesicular GABA/glycine transporters and anion channels

      2012, Neurochemistry International
      Citation Excerpt :

      A significant portion (about 25%) of the GABA(10 μM)-evoked glycine release was dependent on external Ca2+ ions. Entry of extracellular Ca2+ into the cytosol can occur by multiple modes including (i) influx through VSCCs activated by depolarizing stimuli and (ii) reversal of plasmalemmal Na+/Ca2+ exchangers (NCXs; Annunziato et al., 2004; Török, 2007) that can occur when the membrane is depolarized and cytosolic Na+ concentrations increase to facilitate entry of Ca2+ by NCX reversal. Because the stoichiometry of GAT1 is 2Na+/1Cl−/GABA (Keynan and Kanner, 1988), GABA uptake is accompanied by an influx of Na+ ions which could produce localized membrane depolarization triggering activation of VSCCs and changing plasmalemmal NCX function.

    View all citing articles on Scopus
    View full text