Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2008 Feb 1.
Published in final edited form as: J Mol Cell Cardiol. 2006 Dec 22;42(2):289–294. doi: 10.1016/j.yjmcc.2006.11.009

Cardiac pacemaker cell failure with preserved If, ICaL, and IKr: A lesson about pacemaker function learned from ischemia-induced bradycardia.

Victor A Maltsev 1, Edward G Lakatta 1
PMCID: PMC1868668  NIHMSID: NIHMS18066  PMID: 17188292

Cardiac ischemia is an important global health problem. It leads to contraction failure, arrhythmias, and cardiac cell death. Sinus bradycardia characterized by a slow heart rate (<60 bpm) is a prominent ischemia-related arrhythmia directly caused by a deficiency of heart beat initiation within the sinoatrial node (SAN), due to a failure of the heart’s primary pacemaker cells (SANC). The study presented by Yi-Mei Du and Richard Nathan in this issue of Journal of Molecular and Cellular Cardiology [1] deals with the ionic basis of ischemia-induced bradycardia in isolated SANC. The results of their study not only contribute to understanding of ischemia-induced bradycardia, but also help to delineate the fundamental mechanisms of cardiac pacemaker cell function.

Study design and hypothesis

In their studies, Du and Nathan simulated ischemia by superfusing SANC with a solution without glucose, at pH 6.6 that contained either 5.4 or 10 mM KCl. This lead, to a 13% or 43% reduction, respectively, in the spontaneous beating rate of freshly isolated SANC. This simulated ischemia is not true ischemia, e.g. there is no associated hypoxia or flow deficiency effects other than elevated [K+]. Nonetheless simulated ischemia of this sort is often applied in studies like the present one. And, in their careful approach to the problem, the authors first demonstrated that their “ischemic” solutions produced bradycardia in the intact rabbit heart. Then, in isolated SANC they measured the effects of this simulated ischemia on membrane potential and ion currents under voltage clamp. Such an approach is not surprising, because it has been believed for almost half a century that the heart rhythm originates on the surface membrane of the cardiac pacemaker cells. This dogma stems from the Nobel prize triumph of the Hodgkin-Huxley neuron membrane excitability theory [2] and subsequent modification of this theory and extrapolation to the heart cells by Noble (1960) [3]. According to this theory, generation of spontaneous action potentials (APs) by cardiac pacemaker cells is portrayed as a membrane delimited process, i.e. by a pure interplay of time- and voltage- dependent ion currents. Pacemaker field researchers have focused mainly on the identification of multiple ion current components in pacemaker cells, and their respective roles in the spontaneous diastolic depolarization of these cells (DD, Fig.1A) that brings the membrane to the excitation threshold (reviews [4, 5]). Du and Nathan thus tested an hypothesis that bradycardia induced by their simulated ischemia is caused by attenuation of inward currents contributing to this most critical DD phase of the pacemaker AP.

Figure 1.

Figure 1

A: A schematic illustration of the fine structure of the diastolic depolarization (DD) of spontaneous action potential in rabbit SANC shown together with inward ion currents assessed in the study by Du and Nathan [1] and related Ca2+ signals (see text for detail). The nonlinear, exponentially rising, late DD part is shown as a gray area between membrane potential curve and an extrapolation of the initial, linear DD part. The action potential shapes are taken from the two first action potentials presented in Fig.1A of the study by Du and Nathan [1]. B: The ‘‘ischemic’’ solution mainly reduces the nonlinear, late DD part (c), rather than the early part (b) of the DD. Shown is a copy of the Figure 1G from [1]. C: Spontaneous beating of rabbit SANC critically depends upon Ca2+ related mechanisms and protein phosphorylation. Bars show a decrease in the beating rate (% control) induced by different drugs that affect Ca2+ cycling [ryanodine (Ry), BAPTA-AM], NCX (Li+ substitution for Na+), protein phosphorylation (PKI, H-89, MDL), or ion currents: If (Cs+) or T-type Ca2+ current (Ni+). From [6]. PKI and H-89 are protein kinase A inhibitors and MDL is an adenylyl cyclase inhibitor.

Although Ca2+ was not measured by Du and Nathan in the present study of ischemic bradycardia, three of the inward currents assessed are directly related to Ca2+ cycling in the cells, i.e. T- and L- type Ca2+ currents and Na+/Ca2+ exchange (NCX) current (ICaT, ICaL INCX, respectively, Fig.1A, “Inward currents”). A plethora of recent data has emerged to conclusively show that intracellular Ca2+ dynamics, in tight cooperation with surface membrane proteins, are critical for the normal spontaneous firing of SANC (reviews [6, 7]). Nathan’s group, in fact, was among the pioneers that discovered the major role of intracellular Ca2+ cycling in pacemaker function, showing that ryanodine, which interferes with Ca2+ release from the sarcoplasmic reticulum (SR), and BAPTA-AM, which chelates intracellular Ca2+, significantly slowed the spontaneous beating rate of cardiac pacemaker cells [810] The forth inward current measured was the “funny” current (If) activated by membrane hyperpolarization, often referred to as “the” pacemaker current [11, 12]. Finally, the authors completed their set of tested currents with an outward current, delayed rectifier K+ current, specifically, its rapid component IKr, deactivation of which has a major contribution into the DD dynamics in rabbit SANC [13]. Thus, all studied currents are critically involved in either membrane-electrophysiology or intracellular Ca2+ dynamics during the DD. Other currents that have been thought to be involved in SANC pacemaker function [14] were not assessed by Du and Nathan.

Three major currents, If, ICaL, and IKr do not fail when pacemaker function fails during ischemia

The results of the study showed that three major ion channels that contribute to the DD, activation of If and deactivation of IKr which both initiate the DD [12, 13] and ICaL, which becomes activated at the termination of the DD and produces the rapid upstroke of the AP [15] (Fig.1A), are not involved in the ischemia-induced rate reduction. ICaL, in fact, increased, whereas IKr remained unchanged and If was either also unchanged or increased in 5.4 or 10 mM KCl, respectively. If these three prominent currents from the standpoint of membrane-delimited concept of the initiation of the heartbeat did not fail under ischemic-like conditions, then some other mechanism that is critical to normal pacemaker function must have failed.

INCX and ICaT fail during ischemia

Du and Nathan have demonstrated that the pacemaker failure in the ischemia-like conditions is accompanied by a reduction of ICaT and INCX, both known to contribute to the late DD [1618]. This result suggests that one or both of these currents are critical to normal pacemaker function. In our opinion, the reduction in ICaT is unlikely to be a major cause of ischemia-induced bradycardia on the basis of the following. The expression of ICaT is strictly species dependent; current density of ICaT in SAN is large in small animals and becomes smaller as body size increases (mouse > guinea-pig > rabbit > pig) [19]. Thus, it has been reasoned that in large mammals (e.g. pigs and humans), Ca2+ may be exclusively carried via ICaL, and not ICaT; the lack of ICaT in porcine SAN cells would be beneficial for slower pacemaker depolarization. Conversely, a high density of ICaT in mouse SAN cells would promote pacemaker depolarization, thereby increasing heart rate [20]. Importantly, since ICaT has never been detected in human heart tissue [19], clinically relevant ischemia-induced bradycardia in human SANC would not likely involve ICaT. Thus, that ICaT is so variable among species, and is even absent in some SANC, questions its fundamental importance to pacemaker function. Further, ICaT has not been considered as a major component of the pacemaker mechanism of SANC, because a blockade of this current by a low concentration of Ni2+ results only in a minor change in the SANC beating rate (14–16% with 40–50 μM Ni2+, respectively, [21, 22]). Du and Nathan did not evaluate the specific contribution of ICaT to ischemic bradycardia. This might have been tested by supplementing their “ischemic” media with a similar low concentration of Ni2+ to block this current.

The key role of the NCX in pacemaker cell failure during ischemia: novel evidence for fundamental importance of NCX function

In our opinion, of the two inward currents that failed under ischemic conditions, the reduction in INCX is likely the major mechanism for ischemic bradycardia. Although Du and Nathan noted a significant effect of the “ischemic” solution on the DD slope, they assessed this as the average slope of the reduction of voltage throughout the entire DD time course. A closer inspection of their AP recordings, however, reveals that the DD is composed of an initial, linear part, followed later by a non-linear, exponentially rising part, which accelerates the change in membrane potential to achieve the AP threshold (See in Fig 1A “Linear DD” and “Nonlinear DD”, or in Fig.1B arrows b and c, respectively). This fine DD structure has been recently analyzed in great detail using a variety of experimental and numerical modeling approaches in rabbit SANC [18]. It was shown that the non-linear, exponentially rising, late DD part is formed by INCX (see in Fig.1A “INCX“ and “Nonlinear DD” coincide). Notice, however, that while the ischemic solution changes the linear part of DD only slightly (if at all) (DD parts b and b’ in Fig.1B), the major effect is clearly on the late, non-linear part (compare parts c and c’ in Fig.1B). So, the non-linear DD part, in fact, becomes “missing” in the ischemic solution (DD part b’-c’ is almost straight in Fig.1B). This major divergence of the DD from its normal shape thus ought to be produced by the reduction of INCX found by Du and Nathan in the ischemic solution. The corollary of the demonstration that INCX is markedly reduced by simulated ischemia is that INCX must be also critical to normal pacemaker function, thus providing a new additional piece of evidence for the fundamental importance of the NCX for SANC function suggested previously by different research groups [7, 17, 23].

How cell Ca2+ cycling links to NCX function and why it is critical for pacemaker function

So, what determinants of INCX render its function to be crucial to normal pacemaker cell beating? In contrast to other currents tested in the study, INCX is carried not by an ordinary ion channel, but an ion transporter, NCX. It is activated by an increase in intracellular [Ca2+], and once activated this way, it works in so called ‘‘forward mode’’ exchanging 1 Ca2+ for 3 Na+ thus generating an inward current [24] at diastolic potentials. A great benefit to have NCX in the plasma membrane is that it can transform and therefore couple “chemical” signals of Ca2+ cycling within pacemaker cells into “electrical” signals of membrane depolarization followed by an excitation.

Ca2+ cycling in cardiac cells is characterized by two major types of Ca2+ releases from SR: global, AP-synchronized Ca2+ transients and spontaneous, local Ca2+ releases (LCRs) (recent reviews [6, 7]). The essence of Ca2+ cycling in normal ventricular muscle cells is the Ca2+ transient, which is initiated and entrained by externally delivered APs, and serves as a high gain signal amplifier within Ca2+-induced-Ca2+-relase (CICR) paradigm. Since NCX function is both Ca2+- and voltage- dependent, at the peak of the AP it operates in the “reverse mode”, transferring more Ca2+ into cells; then, as the membrane potential repolarizes and crosses the reversal potential of the NCX, it operates in the “forward mode” helping to remove the cytosolic Ca2+, thus allowing the muscle to relax. Spontaneous LCRs usually do not occur between beats in ventricular myocytes during their normal operation. Under excessive SR Ca2+ loading conditions, however, diastolic LCRs initiate Ca2+ waves at the diastolic membrane potentials and activate inward NCX current that has been linked to the diastolic membrane depolarizations that have been implicated in ventricular arrhythmias (review [25]).

The normal Ca2+ cycling in SANC is different from that of ventricular cells. During each beat, in the absence of Ca2+ overload, SANC manifest both an AP-induced Ca2+ transient and, importantly, spontaneous Ca2+ releases, i.e. LCRs in the form of Ca2+ sparks and small Ca2+ waves during the late DD phase [17, 26] (Fig.1A, “Ca2+ dynamics”). A recent discovery explained this fundamental difference of Ca2+ cycling in these two cell types. It turned out that in contrast to rabbit ventricular myocytes, Ca2+ cycling proteins, e.g. phospholamban, ryanodine receptors, and likely L-type Ca2+ channels in rabbit SANC, are highly phosphorylated in the basal state [27](for L-type Ca2+ channels see [28]). PKA-dependent phosphorylation of these proteins presents the SR with an increased Ca2+ to be pumped (Ca2+ influx via ICaL), speeds up the SR Ca2+ filling rate (phospholamban), and alters the Ca2+ release threshold (ryanodine receptors). The net effects of this phosphorylation create conditions that are required for LCR spontaneous activation during the DD. The LCR activation in rabbit SANC is indeed truly spontaneous as it does not depend on membrane depolarization (but see [29] for cat latent atrial pacemaker cells), and can be observed under voltage clamp or in permeabilized SANC [27, 30]. Importantly, it is the spontaneous emergence of LCRs during the late DD that activates INCX and thus changes the DD dynamics from a linear to a nonlinear, exponentially rising function, culminating in ICaL activation and membrane excitation (Fig.1A) [17, 18].

Ca2+ cycling within cardiac pacemaker cells is a “Clock”

Spontaneous diastolic LCRs are manifestation of an internal “Clock” that generates rhythmic intracellular Ca2+ signals (Fig.1A, see Ca2+ clock “ticks”, red) that are transformed by NCX to DD acceleration to prompt rhythmic AP’s. The speed at which the Ca2+ clock runs is assessed by the LCR period, defined as the time from the AP-induced Ca2+ transient early in the cycle (Fig.1A, Clock “reset”, blue) to LCR emergence during the subsequent DD (Clock “tick”). This Ca2+ clock’s period is slightly shorter than the cycle length (Fig.1A) and approximately coincides with the period of the intrinsic Ca2+ oscillations under voltage clamp [27, 30]. The speed at which Ca2+ clock ticks is variable, matching the chronotropic demand for a given condition, and is governed by the SR Ca2+ loading and Ca2+ release characteristics, which in turn, are governed by the degree of phosphorylation of its aforementioned Ca2+ cycling proteins. This phosphorylation-graded variability of the Ca2+ clock is thus the essence of basal rate regulation and regulation of chronotropic reserve of SANC by neuromodulators [27].

While the Ca2+ clock keeps the time of the SANC pacemaker, it is tightly integrated with the plasma membrane proteins (including both electrogenic process and Ca2+ transport) resulting in an extremely robust pacemaker operation [6, 7]: Rhythmic LCRs excite the cell surface membrane via activation of inward NCX current during the late DD phase, but AP- induced Ca2+ influx in turn, depletes SR Ca2+ content (Fig.1A, Ca2+ Clock “reset”) via CICR; and it also balances the NCX-related Ca2+ efflux (“refueling” Ca2+ Clock).

It is important to note, although membrane Ca2+ influx and SR Ca2+ cycling are integrated, a failure to generate the critical LCR/NCX signals leads to pacemaker failure, even when Ca2+ influx via ICaL is normal or even enhanced (as in this study by Du and Nathan). A secondary role of ICaL in SANC pacemaker function has also been demonstrated by observations that β-adrenergic receptor (βAR) stimulation still enhances ICaL under ryanodine, but the chronotropic effect of this stimulation is markedly blunted (both in vitro and in vivo) [22, 27].

The study by Du and Nathan provides additional, important evidence for the above idea of Ca2+-integrated pacemaking, i.e., that normal, robust pacemaker function requires the activation of LCRs and INCX during the late DD, that crucial time when If has faded, but ICaL is not yet activated (Fig.1A). In other words, whereas both If and IKr determine the maximum diastolic potential and initiate the early linear part of the DD [12, 13], this later critical INCX prompt is required to steadily and timely bring the membrane potential to the AP threshold. Indeed, while the increase in ICaL found under ischemic-like conditions results in a lower take-off potential, as seen in Fig.1B, this does not compensate for the absence of an exponentially rising, late DD phase resulted from the deficiency of NCX function. That the impact of a greatly increased If under ischemic-like conditions in the study by Du and Nathan fails to drive membrane potential towards the lower take-off potential, similarly questions a major role of this current in the pacemaker function in these cells. A minor role of If in cardiac pacemaker mechanism has been previously suggested based on a variety of considerations, including its activation kinetics, voltage-dependency, a minor change of the beating rate after its blockade, and a minor role in pacemaker rate regulation [22, 31, 32].

Du and Nathan also observed the blocked APs under ischemic-like conditions. Such AP failure also occurs when Ca2+ and membrane interaction via NCX is blocked in Na+-free bath solution or when normal Ca2+ cycling is compromised by ryanodine or protein kinase-A (PKA) inhibition. In such instances, the DD does begin, but then decelerates and ultimately fails to ignite an excitation [17, 27, 32, 33]. This means, generally, that whatever the factor (ischemia-like conditions, Na+-free bath solution, PKI, ryanodine, etc.) that interferes with NCX or Ca2+ cycling, it has the same major consequences: depending on dose and time, beating rate is greatly slowed, become irregular, and may eventually cease (Fig.1C). This is in distinct contrast to relatively minor effects of blockade of ion channels such as If and ICaT (Fig.1C).

What causes the decrease of INCX in ischemia?

Previous experiments showed that interventions that inhibit LCRs markedly reduce the nonlinear DD component and markedly reduce or abolish spontaneous beating; on the other hand, selectively blocking NCX, with LCRs remaining intact, also reduce the late DD and markedly reduce or abolish spontaneous beating [17, 18](Fig.1C). Accordingly, the ischemia-induced reduction of INCX could be due to altered Ca2+ cycling described above and/or a direct effect on NCX function. Evidence that ischemia influences Ca2+ cycling and INCX has accumulated since the late 1970s-early 1980s, when Ca2+ cycling by SR came into focus as the major mechanism of cardiac cell function (review [25]). Under steady-state conditions the direct effect of acidosis on the SR in skinned cardiac cells [34, 35] and in SR vesicles [36] is to decrease Ca2+ uptake and thus subsequent release. Studies in intact hearts showed that ischemia reduced spontaneous Ca2+ release (indexed as scattered light intensity fluctuations generated by Ca2+-dependent myofilament motion) to zero within approximately 30 min [37, 38]. Recent studies in toad pacemaker cells by Ju and Allen 2003 [39] showed that metabolic inhibition, used to simulate ischemia, decreases Ca2+ release from the SR, mainly by reducing SR Ca2+ content due to decreased Ca2+ uptake by the SR pump.

Du and Nathan found that their “ischemic” solution reduced inward current tails that were sensitive to high concentrations of Ni+ (presumably INCX), even in the presence of ryanodine, when SR Ca2+ ought to be markedly reduced. Also, after exposing freely firing SANC to ryanodine for unspecified time, the “ischemic” solution still slowed their firing rate. Based on these finding, Du and Nathan reasoned that this could be due to a direct effect of simulated ischemia on NCX; in other words, a reduction in LCRs is not sufficient to explain full effect of simulated ischemia to reduce INCX. While a direct effect of acidosis [40], presumably due to a change in electrogenicity of the NCX protein [41], seems to be a plausible explanation, the latter interpretation of the further reduction of the beating rate by the “ischemic” solution in the presence of ryanodine needs careful consideration. Based on a recently reported summary of available data of ryanodine effects in SANC (Table 1 in [32]), ryanodine concentrations ranging from 1 to 30 μM produced rate reductions from 19% to 100%. Ryanodine, in a concentration of 40 μM employed by Du and Nathan, is, in fact, a relatively high concentration of the drug. It is puzzling to us why this resulted only in 12% slowing of the beating rate in their experiments. The ryanodine effect on Ca2+ cycling has a complex dynamics: firstly, it increases LCRs from the SR and this is accompanied by an increase in the beating rate; this is followed by SR Ca2+ depletion, cessation of LCRs, and by a slowing of the beating rate, disturbances in rhythm, and ultimately, failure of spontaneous beating. So, the precise times at which the ‘‘ischemic’’ solution is applied and at which the measurements are taken may yield greatly different results. More specifically, after a small initial rate reduction (such as 12%), the subsequent effect of ‘‘ischemic’’ solution might, in fact, overlap with continuing dynamics of the ryanodine effect, thus complicating interpretation of the data. Direct measurement of Ca2+, particularly LCRs, are required in this regard.

Summary and conclusions

The study by Du and Nathan in this issue of Journal of Molecular and Cellular Cardiology [1] suggests that simulated ischemia-induced bradycardia in rabbit SANC is caused, at least in part, by a failure (i.e. reduction) of two ion currents ICaT and INCX. However, the extent to which each of these reductions contributes to bradycardia was not determined. Since previous studies demonstrated a minor role of ICaT in normal pacemaker function of rabbit SANC, the reduction in INCX rather than that in ICaT is likely the major mechanism in the simulated ischemia-induced bradycardia. The finding of that other major currents ICaL, If, and IKr, contributing to the diastolic depolarization, do not fail under ischemic-like conditions, whereas NCX does, suggests that the latter is critical to the normal pacemaker function. This idea is in line with direct, recent demonstrations in other studies of rabbit SANC [17, 18] of the key role of NCX in normal pacemaker function. Furthermore, because NCX is activated by intracellular Ca2+, a broad interpretation of the results of this study is that pacemaker failure during simulated ischemia is not caused solely by ion channel failure, but also by altered SR Ca2+ cycling and/or a compromised intracellular Ca2+ integration with the membrane excitation due to the deficiency of NCX function. This interpretation supports the novel hypothesis [7] that normal cardiac pacemaker function requires functional integration of internal Ca2+ cycling with membrane-delimited electrogenic proteins. According to this Ca2+ integrated pacemaker concept, spontaneous local Ca2+ releases emerging during the late DD activate an ensemble of local inward NCX currents; this accelerates DD and ignites ICaL to generate an AP (Fig.1A). This DD acceleration is clearly demonstrated in the Du and Nathan’s recordings of spontaneous APs, and it is the late, non-linear DD part (associated with normal SANC operation) does fail under ischemia-like conditions (Fig.1B). While previous experiments demonstrated that ischemia interferes with the SR Ca2+ cycling in different cardiac cell types, unfortunately Ca2+ measurements were not performed in this study. A direct evaluation of local spontaneous Ca2+ releases during the late DD of SANC in future studies of ischemia-induced bradycardia merits consideration.

Acknowledgments

This research was supported by the Intramural Research Program of the National Institutes of Health, National Institute on Aging.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  • 1.Du Y-MD, Nathan RD. Ionic basis of ischemia-induced bradycardia in the rabbit sinoatrial node. Journal of Molecular and Cellular Cardiology. 2006 doi: 10.1016/j.yjmcc.2006.10.004. [DOI] [PubMed] [Google Scholar]
  • 2.Hodgkin AL, Huxley AF. A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol. 1952;117:500–44. doi: 10.1113/jphysiol.1952.sp004764. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Noble D. Cardiac action and pacemaker potentials based on the Hodgkin-Huxley equations. Nature. 1960;188:495–7. doi: 10.1038/188495b0. [DOI] [PubMed] [Google Scholar]
  • 4.Noble D. The surprising heart: a review of recent progress in cardiac electrophysiology. J Physiol. 1984;353:1–50. doi: 10.1113/jphysiol.1984.sp015320. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Noma A. Ionic mechanisms of the cardiac pacemaker potential. Jpn Heart J. 1996;37:673–82. doi: 10.1536/ihj.37.673. [DOI] [PubMed] [Google Scholar]
  • 6.Lakatta EG, Vinogradova TM, Lyashkov A, Sirenko S, Zhu W, Maltsev VA. Annals of the New York Academy of Sciences. 2006. The integration of spontaneous Intracellular Ca2+ cycling and surface membrane ion channel activation entrains normal automaticity in cells of the heart’s pacemaker; p. 1080. in print. [DOI] [PubMed] [Google Scholar]
  • 7.Maltsev VA, Vinogradova TM, Lakatta EG. The emergence of a general theory of the initiation and strength of the heartbeat. J Pharmacol Sci. 2006;100:338–69. doi: 10.1254/jphs.cr0060018. [DOI] [PubMed] [Google Scholar]
  • 8.Li J, Qu J, Nathan RD. Ionic basis of ryanodine’s negative chronotropic effect on pacemaker cells isolated from the sinoatrial node. Am J Physiol. 1997;273:H2481–9. doi: 10.1152/ajpheart.1997.273.5.H2481. [DOI] [PubMed] [Google Scholar]
  • 9.Rigg L, Terrar DA. Possible role of calcium release from the sarcoplasmic reticulum in pacemaking in guinea-pig sino-atrial node. Exp Physiol. 1996;81:877–80. doi: 10.1113/expphysiol.1996.sp003983. [DOI] [PubMed] [Google Scholar]
  • 10.Rubenstein DS, Lipsius SL. Mechanisms of automaticity in subsidiary pacemakers from cat right atrium. Circ Res. 1989;64:648–57. doi: 10.1161/01.res.64.4.648. [DOI] [PubMed] [Google Scholar]
  • 11.Baruscotti M, Bucchi A, Difrancesco D. Physiology and pharmacology of the cardiac pacemaker (“funny”) current. Pharmacol Ther. 2005;107:59–79. doi: 10.1016/j.pharmthera.2005.01.005. [DOI] [PubMed] [Google Scholar]
  • 12.DiFrancesco D. The contribution of the ‘pacemaker’ current (if) to generation of spontaneous activity in rabbit sino-atrial node myocytes. J Physiol. 1991;434:23–40. doi: 10.1113/jphysiol.1991.sp018457. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ono K, Ito H. Role of rapidly activating delayed rectifier K+ current in sinoatrial node pacemaker activity. Am J Physiol. 1995;269:H453–62. doi: 10.1152/ajpheart.1995.269.2.H453. [DOI] [PubMed] [Google Scholar]
  • 14.Bers DM. The beat goes on: diastolic noise that just won’t quit. Circ Res. 2006;99:921–3. doi: 10.1161/01.RES.0000249859.10103.a9. [DOI] [PubMed] [Google Scholar]
  • 15.Noma A, Kotake H, Irisawa H. Slow inward current and its role mediating the chronotropic effect of epinephrine in the rabbit sinoatrial node. Pflugers Arch. 1980;388:1–9. doi: 10.1007/BF00582621. [DOI] [PubMed] [Google Scholar]
  • 16.Doerr T, Denger R, Trautwein W. Calcium currents in single SA nodal cells of the rabbit heart studied with action potential clamp. Pflugers Arch. 1989;413:599–603. doi: 10.1007/BF00581808. [DOI] [PubMed] [Google Scholar]
  • 17.Bogdanov KY, Vinogradova TM, Lakatta EG. Sinoatrial nodal cell ryanodine receptor and Na(+)-Ca(2+) exchanger: molecular partners in pacemaker regulation. Circ Res. 2001;88:1254–8. doi: 10.1161/hh1201.092095. [DOI] [PubMed] [Google Scholar]
  • 18.Bogdanov KY, Maltsev VA, Vinogradova TM, Lyashkov AE, Spurgeon HA, Stern MD, Lakatta EG. Membrane potential fluctuations resulting from submembrane Ca2+ releases in rabbit sinoatrial nodal cells impart an exponential phase to the late diastolic depolarization that controls their chronotropic state. Circ Res. 2006;99:979–87. doi: 10.1161/01.RES.0000247933.66532.0b. [DOI] [PubMed] [Google Scholar]
  • 19.Vassort G, Talavera K, Alvarez JL. Role of T-type Ca2+ channels in the heart. Cell Calcium. 2006;40:205–20. doi: 10.1016/j.ceca.2006.04.025. [DOI] [PubMed] [Google Scholar]
  • 20.Ono K, Iijima T. Pathophysiological significance of T-type Ca2+ channels: properties and functional roles of T-type Ca2+ channels in cardiac pacemaking. J Pharmacol Sci. 2005;99:197–204. doi: 10.1254/jphs.fmj05002x2. [DOI] [PubMed] [Google Scholar]
  • 21.Hagiwara N, Irisawa H, Kameyama M. Contribution of two types of calcium currents to the pacemaker potentials of rabbit sino-atrial node cells. J Physiol. 1988;395:233–53. doi: 10.1113/jphysiol.1988.sp016916. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Vinogradova TM, Bogdanov KY, Lakatta EG. beta-Adrenergic stimulation modulates ryanodine receptor Ca(2+) release during diastolic depolarization to accelerate pacemaker activity in rabbit sinoatrial nodal cells. Circ Res. 2002;90:73–9. doi: 10.1161/hh0102.102271. [DOI] [PubMed] [Google Scholar]
  • 23.Sanders L, Rakovic S, Lowe M, Mattick PA, Terrar DA. Fundamental importance of Na+-Ca2+ exchange for the pacemaking mechanism in guinea-pig sino-atrial node. J Physiol. 2006;571:639–49. doi: 10.1113/jphysiol.2005.100305. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Noble D. Modeling the heart--from genes to cells to the whole organ. Science. 2002;295:1678–82. doi: 10.1126/science.1069881. [DOI] [PubMed] [Google Scholar]
  • 25.Lakatta EG. Functional implications of spontaneous sarcoplasmic reticulum Ca2+ release in the heart. Cardiovasc Res. 1992;26:193–214. doi: 10.1093/cvr/26.3.193. [DOI] [PubMed] [Google Scholar]
  • 26.Maltsev VA, Vinogradova TM, Bogdanov KY, Lakatta EG, Stern MD. Diastolic calcium release controls the beating rate of rabbit sinoatrial node cells: numerical modeling of the coupling process. Biophys J. 2004;86:2596–605. doi: 10.1016/S0006-3495(04)74314-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Vinogradova TM, Lyashkov AE, Zhu W, Ruknudin AM, Sirenko S, Yang D, Deo S, Barlow M, Johnson S, Caffrey JL, Zhou YY, Xiao RP, Cheng H, Stern MD, Maltsev VA, Lakatta EG. High basal protein kinase A-dependent phosphorylation drives rhythmic internal Ca2+ store oscillations and spontaneous beating of cardiac pacemaker cells. Circ Res. 2006;98:505–14. doi: 10.1161/01.RES.0000204575.94040.d1. [DOI] [PubMed] [Google Scholar]
  • 28.Petit-Jacques J, Bois P, Bescond J, Lenfant J. Mechanism of muscarinic control of the high-threshold calcium current in rabbit sino-atrial node myocytes. Pflugers Arch. 1993;423:21–7. doi: 10.1007/BF00374956. [DOI] [PubMed] [Google Scholar]
  • 29.Huser J, Blatter LA, Lipsius SL. Intracellular Ca2+ release contributes to automaticity in cat atrial pacemaker cells. J Physiol. 2000;524(Pt 2):415–22. doi: 10.1111/j.1469-7793.2000.00415.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Vinogradova TM, Zhou YY, Maltsev V, Lyashkov A, Stern M, Lakatta EG. Rhythmic ryanodine receptor Ca2+ releases during diastolic depolarization of sinoatrial pacemaker cells do not require membrane depolarization. Circ Res. 2004;94:802–9. doi: 10.1161/01.RES.0000122045.55331.0F. [DOI] [PubMed] [Google Scholar]
  • 31.Vassalle M. The pacemaker current (I(f)) does not play an important role in regulating SA node pacemaker activity. Cardiovasc Res. 1995;30:309–10. [PubMed] [Google Scholar]
  • 32.Lakatta EG, Maltsev VA, Bogdanov KY, Stern MD, Vinogradova TM. Cyclic variation of intracellular calcium: a critical factor for cardiac pacemaker cell dominance. Circ Res. 2003;92:e45–50. doi: 10.1161/01.res.0000055920.64384.fb. [DOI] [PubMed] [Google Scholar]
  • 33.Vinogradova TM, Zhou YY, Bogdanov KY, Yang D, Kuschel M, Cheng H, Xiao RP. Sinoatrial node pacemaker activity requires Ca(2+)/calmodulin-dependent protein kinase II activation. Circ Res. 2000;87:760–7. doi: 10.1161/01.res.87.9.760. [DOI] [PubMed] [Google Scholar]
  • 34.Fabiato A, Fabiato F. Effects of pH on the myofilaments and the sarcoplasmic reticulum of skinned cells from cardiace and skeletal muscles. J Physiol. 1978;276:233–55. doi: 10.1113/jphysiol.1978.sp012231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Fabiato A. Use of aequorin for the appraisal of the hypothesis of the release of calcium from the sarcoplasmic reticulum induced by a change of pH in skinned cardiac cells. Cell Calcium. 1985;6:95–108. doi: 10.1016/0143-4160(85)90037-5. [DOI] [PubMed] [Google Scholar]
  • 36.Inesi G, Hill TL. Calcium and proton dependence of sarcoplasmic reticulum ATPase. Biophys J. 1983;44:271–80. doi: 10.1016/S0006-3495(83)84299-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Stern MD, Weisman HF, Renlund DG, Gerstenblith G, Hano O, Blank PS, Lakatta EG. Laser backscatter studies of intracellular Ca2+ oscillations in isolated hearts. Am J Physiol. 1989;257:H665–73. doi: 10.1152/ajpheart.1989.257.2.H665. [DOI] [PubMed] [Google Scholar]
  • 38.Weiss RG, Gerstenblith G, Lakatta EG. Calcium oscillations index the extent of calcium loading and predict functional recovery during reperfusion in rat myocardium. J Clin Invest. 1990;85:757–65. doi: 10.1172/JCI114501. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Ju YK, Allen DG. Early effects of metabolic inhibition on intracellular Ca2+ in toad pacemaker cells: involvement of Ca2+ stores. Am J Physiol Heart Circ Physiol. 2003;284:H1087–94. doi: 10.1152/ajpheart.00755.2002. [DOI] [PubMed] [Google Scholar]
  • 40.Earm YE, Irisawa H. Effects of pH on the Na-Ca exchange current in single ventricular cells of the guinea pig. Jpn Heart J. 1986;27 (Suppl 1):153–8. [PubMed] [Google Scholar]
  • 41.Egger M, Niggli E. Paradoxical block of the Na+-Ca2+ exchanger by extracellular protons in guinea-pig ventricular myocytes. J Physiol. 2000;523(Pt 2):353–66. doi: 10.1111/j.1469-7793.2000.t01-1-00353.x. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES