Comprehensive Physiology Wiley Online Library

Regulation of Cellular Calcium in Cardiac Myocytes

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 General Aspects of Cellular Calcium Regulation
1.1 Ca Transport Systems
1.2 Cytosolic Volume Conventions
1.3 Ca Binding Sites and Buffering in Cardiac Myocytes
1.4 Ca Requirements for Contractile Activation
1.5 Simplified Kinetic Considerations During a Ca Transient
2 Sarcolemmal Ca Transport Systems
2.1 Ca Channels
2.2 Na/Ca Exchange
2.3 Sarcolemmal Ca‐ATPase
3 Intracellular Ca Transporters
3.1 Sarcoplasmic Reticulum Ca‐ATPase and Sarcoplasmic Reticulum Ca Content
3.2 Sarcoplasmic Reticulum Ca Release Channels
3.3 Mitochondrial Ca Transport
4 Ca Removal from the Cytoplasm During Relaxation
4.1 Relative Contributions of Ca Transporters
4.2 Species, Developmental, and Temperature Dependence
4.3 Ca Recycling from Mitochondria Back to the SR
5 The Ca Supply that Activates Contraction
5.1 Ca Influx vs SR Ca Release
5.2 Fraction of SR Ca Released During a Twitch
6 Perturbations of Cellular Ca Balance
6.1 Rest‐Dependent Changes in Cellular and SR Ca Content
6.2 Refilling Depleted Internal Ca Stores
6.3 Rest‐Decay and Rest‐Potentiation of Twitches
6.4 Force‐Frequency Relationships
Figure 1. Figure 1.

Simplified cardiac myocyte Ca fluxes. Ca enters during the action potential via ICa and Na/Ca exchange (Na/CaX). Ca entry triggers Ca release from the SR and the combination activates the myofilaments (MF). Elevated [Ca]i stimulates Ca removal from the cytosol by the SR Ca‐ATPase (modulated by phospholamban, PLB), the Na/Ca exchange, the sarcolemmal Ca‐ATPase, and the mitochondrial uniporter. As [Ca]i declines Ca dissociates from the myofilaments allowing relaxation. The sarcolemmal (Na + K)ATPase and Na/H exchange are also indicated, along with other key mitochondrial ion transporters.

Modified from Bers 34
Figure 2. Figure 2.

Passive Ca buffering in cardiac myocytes. The six curves represent different estimates of how passive Ca buffering changes as [Ca]i increases from 100 to 1500 nM. These values are indicated as increments above the amount of Ca bound at 100 nM [Ca]i (see also Table 2 and discussion in text). Fabiato‐Orig Fast, Berlin‐Fast and Hove‐Madsen‐Equilibrium are taken directly from values in the relevant papers 29,104,145. Fab/Bers‐New Fast updates some binding constants from the original estimates of Fabiato 104, as described in the text. Berlin‐Fast+MF Ca/Mg adds known myofilament sites at which Ca and Mg compete. Calculated Total‐New refers to values resulting from constants in Table 2.

Figure 3. Figure 3.

Measurement of fast Ca buffering in a voltage clamped rat ventricular myocyte. Cells were exposed to 10 μM thapsigargin to prevent SR Ca uptake and were in Na‐free conditions (inside and out) to prevent Na/Ca exchange. A. Voltage clamp pulses (200 ms from −40 to 0 mV) activated ICa and produced nearly step‐like increases in [Ca]i. B: The integral of the ICa from the fourth pulse in A (∫ ICa follows the kinetics of the rise in [Ca]i. The sag in the [Ca]i trace may represent slow buffering that is not in phase with Ca entry via ICa. The “ S” indicates where [Ca]i would be predicted to fall to eventually if the slow buffering by myofilament Ca/Mg sites are included with the fast buffering measured (see text). The points marked [A] and [B] in panel A are the predicted final settling points for [Ca]i if the slow buffering is included for the total ∫ICa throughout the train (as for S) and when the equilibrium buffering from Fig 2 and Table 2 is assumed, respectively

Modified from Berlin et al. 29
Figure 4. Figure 4.

Free and total Ca requirements for myofilament activation in cardiac myocytes. A: Two different sets of parameters are used for the Hill equation (100/(1 + (Km/[Ca])n) describing the steady‐state relationship between force and [Ca]. The solid curve (Km = 600 nM and n = 4) is more consistent with recent estimates in intact ventricular muscle, whereas the lower dotted curve (Km = 1000 nM and n = 2) is closer to traditional data from skinned muscle fibers (see text). B. The same relationships as in A, but the free [Ca] axis has been altered to account for the cytosolic Ca buffering relationship (Calculated‐Total‐New in Figure 2 and 2). Thus, this predicts the force as a function of the total amount of Ca added to the cytosol from a resting level of 150 nM free [Ca]. Coincidentally, almost the same curves are obtained if resting [Ca]i is assumed to be 100 nM and the intermediate buffering curve from Figure 2 is used (Berlin‐Fast+MF Ca/Mg).

Figure 5. Figure 5.

Cytosolic Ca buffering and transport during a Ca transient. A. A Ca transient is calculated from the product of a rising and declining exponential which rises from 100 to 744 nM and back with time constants (τ) of 30 ms (rising) and 300 ms (falling). This was used as a driving function to calculate the time‐dependent changes in binding to various cytosolic buffers (B) using the rate constants and concentrations in Table 3 (TnC is troponin C, SL is total sarcolemmal site in Table 3, CaM is calmodulin, TnC‐Ca/Mg and My‐Ca/Mg are Ca binding to Ca/Mg sites on TnC and myosin). The change in total cytosolic [Ca] (ΔTotal CaCyt in A) is the sum of the curves in B and the free [Ca]i curve. C: Ca fluxes associated with the Ca transient. ICa activation was calculated by a rising (τ = 3 ms) and falling exponential (τ = 40 ms), peaking at 6.8 pA/pF and bringing in 16 μmol/l cytosol. The SR Ca‐pump rate was 210/(1 + (300 nM/[Ca]i)2) in mol/l cytosol. The SR Ca leak was initiated to counterbalance the SR Ca‐pump rate at rest, and changed linearly as a function of free [Ca] in the SR (based on intra‐SR Ca buffering with a Kd = 600 μM and Bmax equivalent to 180 μmol/l cytosol in an SR volume occupying 3.5% of cell volume). The SR Ca release flux was taken as the residual Ca flux required to produce the driving Ca transient.

Figure 6. Figure 6.

Currents via T‐and L‐type Ca channels. A. Barium currents (with 115 mM Ba) induced by depolarizations to various test potentials from holding potentials of 80 or 30 mV. The Em protocol is shown in the top trace, the currents in the middle trace, and the difference between these currents in the bottom panel. Peak IBa from 30 mV is attributed to L‐type Ca channels and the additional transient difference current activated from 80 mV is attributed to T‐type Ca channels. B: ICa in dog Purkinje fiber cell with 2 mM Ca as charge carrier. The more positive Em required in IBa is due to the higher surface potential in 2 mM Ca vs 115 mM Ba. C: ICa from several species (from 80 to 100 mV), where the hump at ∼−40 mV is due to T‐type current and differs among the tissues studied. Dog Purkinje (ref. 139) and rabbit ventricle (G. M. Briggs and D. M. Bers, unpublished) are in 2 mM Ca. Dog atrium (ref. 27) and guinea‐pig ventricle (from Mitra & Morad, 218) are with 5 mM Ca, but are shifted by 10 mV to compensate for surface potential differences.

From Bean 27, with permission. From Hirano et al. 139, with permission. From Bers 34 with permission)
Figure 7. Figure 7.

Ca channel inactivation with different charge carriers. Normalized current amplitudes measured at 0 mV (except Ins at −30 mV to match activation state). ICa was recorded under both perforated patch (allowing normal SR Ca release and Ca transients) and ruptured patch with cells dialyzed with 10 mM EGTA (to prevent global Ca transients). IBa was also recorded with ruptured patch (with 10 mM EGTA in the pipette). Extracellular [Ca] and [Ba] were both 2 mM and Ins was measured in divalent‐free conditions (10 mM EDTA inside and out) with [Na]o at 20 mM and [Na]i at 10 mM. Peak currents were 1370, 808, 780, and 5200 pA and were attained at 5, 7, 10, and 14 ms for ICa (perforated), ICa (ruptured), IBa and Ins respectively. Halftimes of current decline were 17, 37, 161, and >500 ms, respectively.

Recordings made by Dr. W. Yuan
Figure 8. Figure 8.

Ca‐dependent inactivation by SR Ca release and recovery of ICa. Long (4 sec.) voltage clamp pulses activated ICa and SR Ca release and Ca transient in Na‐free conditions and 5.4 mM [Ca]o. The pulse labeled a was the first pulse after depletion of the SR and b was after 5–10 pulses when the SR was loaded. The smaller Ca transient produces less ICa inactivation. The larger Ca transient produces marked inactivation, but the current partially recovers as [Ca]i declines. The lower difference traces show the similarity between the ICa inactivation and Cai transient

Modified from Sipido et al. 311, with permission
Figure 9. Figure 9.

Ca currents measured during square pulse and action potential clamp in rat and rabbit ventricular myocytes. Command Em waveforms were either the traditional 200 ms square depolarization to 0 mV or an action potential (AP), which was recorded from normal rat and rabbit ventricular myocytes under more physiological conditions (i.e. normal Ca Na and K concentrations, ref 358). ICa was then recorded under conditions where all other currents were blocked (e.g. Na‐free and Cs‐rich inside and out). With the action potential waveform, the peak ICa is smaller in both species and occurs later (middle panel). The ICa integral (lower panels) for the AP clamp vs the square, pulses are smaller for the rat but larger for the rabbit (see ref. 358 for additional information).

Figure 10. Figure 10.

Simulations of guinea‐pig ventricular action potential using Oxsoft Heart (v 4.5). This comprehensive simulation created by Professor Denis Noble and colleagues (at Oxford, UK) calculates the Em and many of the ionic currents that are known to flow during the action potential. Shown here are the Em (top), ICa, INa/Ca (middle panel), [Ca]i, and force (lower panel) for two different initial [Na]i values (5 and 8 mM).

Figure 11. Figure 11.

Na/Ca exchange current in intact ventricular myocytes recorded under controlled conditions. A. Very slow speed recording of Em (top) and INa/Ca where the spikes are from ramp depolarizations that were used to generate the current voltage relationships in B, C, and D at the times indicated (a–h). The bars in A are when [Na]o was applied (replacing Li) to activate INa/Ca ([Ca]o was 1 mM throughout). At the arrow [Ca]i was increased from nominally Ca free to 430 nM with 42 mM EGTA and 140 CsCl in the dialyzing pipette throughout. Ouabain, Ba, Cs, D600 and tetraethylammonium were used to inhibit other ionic currents. Application of [Na]o stimulated INa/Ca only after [Ca]i was raised. The gradual decline in INa/Ca (d–g) was supposed to be due to depletion of [Ca]i.

From Kimura et al. 170, with permission
Figure 12. Figure 12.

Current–voltage relationships for INa/Ca recorded in excised giant patches from guinea pig ventricular myocytes. The patches were treated with chymotrypsin to remove INa/Ca inactivation and all other known currents were blocked 210. A: With constant [Ca]o (2 mM), [Na]o (150 mM) and [Ca]i (1 μM), the [Na]i was varied from 5 to 100 mM. The lower 3–4 curves probably reflect the INa/Ca expected in intact cells when [Ca]i is high (e.g. peak systole). B: The same conditions as A, except [Ca]i is reduced to 100 nM, comparable to diastolic [Ca]i.

From Matsuoka and Hilgemann 210, with permission
Figure 13. Figure 13.

Na/Ca exchange currents and [Ca]i in a guinea pig ventricular myocyte. The Ca transient (using fura‐2) show that Ca influx via Na/Ca exchange can bring in large quantities of Ca during large sustained depolarizations (A and B). The [Na] in the dialyzing pipette was 7.5 mM. Outward currents during the depolarization were off‐scale. C: The [Ca]i‐dependence of the “ tail” current was observed upon repolarization to −80 mV. Other ionic currents are blocked by Cs, tetraethylammonium, verapamil, and ryanodine.

From Barcenas‐Ruiz et al., 13, with permission
Figure 14. Figure 14.

Model of the Na/Ca exchanger based on recent work by Nicoll et al. 236 and Iwamoto et al.157. The structure is consistent with 9 membrane spanning segments and a glycosylation site (CH2O). The large cytoplasmic loop contains the Na/Ca exchange inhibitory peptide (XIP) domain (also associated closely with the region responsible for Na‐dependent inactivation) and also contains the site for allosteric regulation by [Ca]i*. The two homologous α repeats (α‐1 and α‐2) are also indicated.

Figure 15. Figure 15.

Na/Ca exchange reversal potential (ENa/Ca) during the rabbit ventricular action potential. This schematic diagram shows how ENa/Ca is expected to change during the action potential for two different levels of intracellular Na activity (aNai = 5 and 8 mM), where aNai is roughly 0.78 · [Na]i. When Em is positive to ENa/Ca, Ca influx via the Na/Ca exchange is thermodynamically favored (shaded areas). When Em is negative to ENa/Ca, Ca extrusion is favored. Resting [Ca]i =150 nM, [Ca]o = 2 mM and aNao = 110 mM for both traces and aNai and peak [Ca]i are as indicated. The [Ca]i trace reaches a peak 40 msec after the action potential begins.

After Bers 32,34
Figure 16. Figure 16.

Extracellular [Ca] depletion in rat‐rabbit ventricular muscle. Changes in [Ca]o were measured with double‐barreled Ca‐selective microelectrodes during individual contractions in rabbit (A) and rat (B) ventricular muscle (0.5 Hz, 30°C). The traces show [Ca]o (top) and tension (bottom) in the absence and presence of 10 mM citrate (which limits [Ca]o depletion by buffering [Ca]. The bath [Ca]o was 0.5 mM and is indicated by the dotted line.

A is modified from Shattock and Bers 309 and composite from Bers 34, with permission
Figure 17. Figure 17.

Changes in Na/Ca exchange driving force during action potential in rat and rabbit ventricle. ENa/Ca is expected to change during the action potential and Ca transient in rabbit and rat ventricle (top). The estimated changes in the net electrochemical driving force for Na/Ca exchange (ENa/Ca – Em) are shown in the bottom panel. Na/Ca exchange stoichiometry of 3Na:1Ca was assumed, the aNai values were measured 309 and, for simplicity, the Ca transient accompanying the contraction, has been assumed to be the same for both species. Resting [Ca]i was assumed to be 150 nM, rising to a peak of 1 μM, 40 msec after the upstroke of the action potential. Note the similarity between the lower panels and the [Ca]o traces in Figure. 16.

Modified from Shattock and Bers 309 and Bers 34, with permission
Figure 18. Figure 18.

Caffeine‐induced Ca transient and Na/Ca exchange current in a ferret ventricular myocyte. A: After a steady‐state series of voltage clamp pulses, 10 mM caffeine was rapidly and continuously applied to release SR Ca. Experimental conditions blocked most other ionic currents (e.g. Cs inside and out) and indo‐1 was used as the Ca indicator. The inward INa/Ca rises to a peak before the Ca transient. B: Instantaneous INa/Ca from panel A is plotted as a function of the global [Ca]i reported by indo‐1. The hysteresis shows that during the rising phase of the Ca transient, more inward INa/Ca is observed for a given [Ca]i than during relaxation. The relationship between INa/Ca and [Ca]i during relaxation was fit to a linear regression (for [Ca]i between 300 and 450 nM (dashed line; see also Fig. 13C).

Experiment performed by Dr. Li Li. C. INa/Ca and [Ca]i data are reproduced from panel A and the subsarcolemmal [Ca]i sensed by the Na/Ca exchanger ([Ca]Na/CaX) based on the linear regression in panel B is also shown
Figure 19. Figure 19.

Ca transport by the SR Ca‐ATPase based on literature values (see text). Ca pump rates were calculated based on Eq. 3 using Vmax = 200 μmol/liter cytosol, Km = 600 nM and n= 2. Activation of protein kinase A (PKA) decreases the Km to 250 nM, but does not affect Vmax.

Figure 20. Figure 20.

Measurement of SR Ca uptake rate in digitonin permeabilized myocytes. A. Free [Ca] was measured simultaneously with indo‐1 and Ca electrode. Ca was added at time 0 to a suspension of rabbit ventricular myocytes permeabilized with digitonin and incubated with 25 μM ruthenium red, 12.5 mM creatine phosphate, and 10 mM oxalate. B: The change in free [Ca] (Ca, solid line, from A), protein bound Ca (Ca‐Pn), Ca‐oxalate (Ca‐Ox), Ca‐indo (Ca‐In), and the sum of free and bound Ca (ΣCa) after addition of 14 μM total Ca. After an initial increase in free [Ca], Ca uptake by the SR lowered the free [Ca] back to the level before the Ca addition. The Ca bound to protein, oxalate, and indo‐1 was calculated from the passive Ca binding to permeabilized myocytes and indo‐1 144,145. C. The Ca uptake rate was calculated as the rate of change in the total non‐SR Ca with time (i.e. dΣCa/dr). D: [Ca] dependence of Ca uptake by the SR. The Ca uptake rate determined in C was plotted versus the corresponding free [Ca] measured in A. Data were fit with Eq. 3 to determine the maximal Ca uptake rate, Km, and Hill coefficient. Data are shown for addition of two different amounts of Ca to the myocyte suspension (14 μM total Ca = diamonds and 43 μM total Ca = circles.

From Hove‐Madsen and Bers 145, with permission
Figure 21. Figure 21.

Unidirectional SR Ca fluxes at rest. At rest when [Ca]SR is changing only very slowly, net SR Ca flux must be close to zero, implying the forward rate (VFor) is equal to the sum of reverse flux and leak from the SR (VRev + VLeak). The forward rate predicted from the Hill curve is 27 μmol/liter cytosol/sec. For a VLeak of 0.3 μmol/l cytosol/sec 24 the reverse flux must be 26.7 μmol/liter cytosol/sec.

Figure 22. Figure 22.

The influence of SR Ca leak on SR Ca content. The effect of different leak values on steady‐state SR Ca content with [Ca]i = 150 nM was calculated using Eq. 7 and 8 with Vmax = 207 μmol/liter cytosol/sec, Kmf = 300 nM, BmaxSR = 3.95 mM and KdSR = 600 μM. For pump inhibition by thapsigargin (TG) Vmax was decreased by 50%. For Ca‐pump stimulation by isoproterenol, Km was reduced by 50%. The SR Ca leak rate (0.3 μmol/liter cytosol/sec) measured by Bassani and Bers 24 is indicated. (See text and ref 117 for further discussion.)

Figure 23. Figure 23.

Ionic fluxes across the mitochondrial membrane. Ca enters via a uniport, down an electrical gradient established by the proton pump at bottom. Ca can be extruded by a Na/Ca antiport and Na is extruded by Na/H exchange thereby completing the cycle. Elevated cytoplasmic [Ca] can lead to elevated mitochondrial [Ca] and increased activity of mitochondrial dehydrogenases and NADH production.

Figure 24. Figure 24.

Mitochondria free [Ca] ([Ca]m) as a function of cytosolic [Ca] ([Ca]c). Increases of [Ca]c in rat ventricular myocytes were induced by reduction of extracellular [Na] (i.e. via Na/Ca exchange). Mean [Ca]c was measured using indo‐1 (loaded as the salt form) and [Ca]m was measured using indo‐1 (loaded as the AM form) with Mn quench of cytosolic indo‐1. Data are taken from Miyata et al., 220 and have been redrawn (without error bars) and including a broken line corresponding to [Ca]m = [Ca]c (slope = 1).

Figure 25. Figure 25.

Relaxation in rabbit ventricular myocyte with selective inhibition of Ca transporters. Normalized cell relaxations are shown and conditions were either (1) with all Ca transporters functional during relaxation of a steady‐state twitch (Tw), (2) preventing net SR Ca uptake during a caffeine‐induced contracture in NT (Caff), (3) additionally inhibiting Na/Ca exchange in 0Na, 0Ca solution (Caff,0Na,0Ca). To further analyze relaxation during Caff, 0Na, 0Ca, this was coupled with either (4) inhibition of mitochondrial Ca uptake with 1 μM FCCP + 1 μM oligomycin so only sarcolemmal Ca‐ATPase was functional (Caff,0Na,0Ca + FCCP) or (5) inhibition of the sarcolemmal Ca‐pump by elevating [Ca]o to 10 mM after pre‐depletion of [Na]i, so that only mitochondrial Ca uptake was functional (Caff,0Na,10Ca). All four Ca removal systems were also blocked by combining the caffeine application with 0Na, 10Ca and FCCP (after pre‐depletion of [Na]i). The traces are based on mean t1/2 values (shown along the traces without standard error values) for relaxation measured in experiments described by Bassani et al. 20.

Figure 26. Figure 26.

Twitch Ca transients in rabbit and rat ventricular myocytes with SR Ca‐ATPase blocked by thapsigargin. Ca transient were measured using indo‐1 fluorescence during electrically stimulated twitches in rabbit (A) and rat (B) ventricular myocytes before (Control) and after treatment with 2.5 μM thapsigargin (TG). Twitches were evoked 10 sec after switching to control solution (after 5–7 min pre‐perfusion with 0Na,0Ca solution). For the TG twitch the cells were exposed to TG for 2 min during the last part of the pre‐perfusion period. SR Ca load was maintained despite complete block of the SR Ca‐pump. Time constants (τ) of [Ca]i decline are shown

Modified from Bassani et al., 17
Figure 27. Figure 27.

Ca transport functions derived from Ca transients in intact rabbit and rat ventricular myocytes. Ca transport rates as functions of [Ca]i for the SR Ca‐ATPase, Na/Ca exchange and combined slow systems (mitochondrial and sarcolemmal Ca‐ATPase) were determined as described in the text and with respect to Eq. 9 and 10. Independent values obtained for JSR, JNa/CaX and JSlow respectively were: Vmax (in μmol/liter cytosol/sec) = 82, 46 and 3.9 for rabbit and 207, 27 and 4 for rat; Km (in nM) = 264, 316, and 362 in rabbit and 184, 257 and 268 in rat; n = 3.7, 3.7, 3.2 in rabbit and 3.9, 3.4 an 3.5 in rat.

Based on data from Bassani et al. 17
Figure 28. Figure 28.

Integrated Ca fluxes during twitch relaxation in rabbit and rat ventricular myocytes. Free [Ca]i during relaxation of a normal twitch was used as a driving function so that Ca flux via each system could be integrated over time using the [Ca]i dependence as described in Figure 27 and Eq. 9 and 10.

Modified from Bassani et al. 17
Figure 29. Figure 29.

Ca transients ICa during a twitch and INa/Ca used to measure SR Ca content. Ca transients and inward currents (ICa and INa/Ca) in a rat ventricular cardiomyocyte (dialyzed with 50 μM indo‐1) were recorded during the last two (of 8) conditioning voltage clamp pulses to 0 mV and during two rapid applications of 10 mM caffeine (bars, holding potential −70 mV). Amplitude and time scales for ICa (left) and INa/Ca (right) are different.

From Delbridge et al. 90, with permission
Figure 30. Figure 30.

Relative reliance on Ca influx in different cardiac muscle preparations. The effect of caffeine (10 mM) or ryanodine (100 nM) on steady‐state twitch contraction amplitude (0.5 Hz at 30°C or 23°C for frog) in various cardiac muscle preparations.

Data are from refs 33,34; modified from Bers 34
Figure 31. Figure 31.

Fractional SR Ca release during the twitch depends on both ICa trigger and SR Ca load in ferret ventricular myocytes. A. Cells were loaded to a constant level (by stimulation at 0.5 Hz). Extracellular [Ca] was changed just prior to the test contraction where the fraction of SR Ca release was measured. Thus increasing trigger (at constant SR Ca load) increases fractional release. Parallel voltage clamp experiments were done to evaluate how ICa changes over this range of [Ca]o. B. Using a constant trigger (with 2 mM [Ca]o) the SR of cells were Ca loaded to different levels by varying frequency (as indicated) and also increasing [Ca]o at the highest load. SR Ca content was determined by translating the peak and resting [Ca]i during caffeine‐induced contractures to total cytosolic [Ca], using values determined by Hove‐Madsen and Bers 145 for cytosolic Ca buffering.

Modified from Bassani et al. 19, with permission
Figure 32. Figure 32.

Rest decay and rest potentiation of twitches and SR Ca content in rabbit, rat and ferret myocytes. Both post‐rest twitches and caffeine‐induced contractures (Caff) were measured either in a modified normal Tyrode's solution (NT), when Na/Ca exchange was blocked during the rest (0Na,0Ca) or after pre‐depletion of [Na]i and rest in Ca‐free, 140 mM Na solution (0Cao) to stimulate Ca efflux via Na/Ca exchange. Caff data are indicative of SR Ca load. SR Ca content data (Caff) was fit with [Ca]SRC = A · exp(–t/τRD)(1 –exp{–t/τRF}) +B, where t is time, τRD is time constant, of rest decay of SR Ca content and τRF is a rest‐dependent filling time constant used to explain the delay in rest decay of SR Ca content. A and B are constants for scaling and baseline respectively. Twitch data were fit with the expression C[Ca]SRC(1–exp{–t/τFCC}) + D, where τECC: is the time constant for recovery of E‐C coupling and C and D are scaling and baseline constants. A small second exponential factor was also included to explain the slow decline in twitch amplitude in rabbit ventricle after long rests in 0Na, 0Ca.

Data are from Bers et al. 36 and Bassani and Bers 23 and have been combined and reanalyzed
Figure 33. Figure 33.

Refilling of SR Ca stores after depletion in rabbit ventricle. SR Ca content was measured by either caffeine‐induced contracture amplitude (in isolated myocytes at 23°C) or rapid cooling contractures (RCC) in trabeculae from rabbit ventricle (at 30°C). SR Ca was depleted by either a long rest (5 min) in normal solution (muscle) or a 10 sec application of 10 mM caffeine (myocyte). After SR Ca depletion, stimulation was started at 0.5 Hz and the SR Ca content was assessed after the beat number indicated.

Data are from Bers 33 and Bassani et al. 16
Figure 34. Figure 34.

Changes in contraction force with abrupt changes in frequency. Frequency of stimulation of a rabbit ventricular trabecula (at 30°C) was increased from 0.5 to 1.5 Hz and then returned to 0.5 Hz.

Figure 35. Figure 35.

Steady‐state force‐frequency relationships. The effect of frequency (from 0.5 to 2 Hz) on twitch force in rabbit (circles), rat (squares) and guinea pig (triangles) ventricular muscle. Data for rabbit and rat are at 30°C.

From Bers (33 and unpublished.] Data for guinea pig are at 36.5°C and were taken from Kurihara and Sakai 176. RCCs were initiated within 5 sec of the last steady‐state stimulated contraction. [Redrawn from Bers 34


Figure 1.

Simplified cardiac myocyte Ca fluxes. Ca enters during the action potential via ICa and Na/Ca exchange (Na/CaX). Ca entry triggers Ca release from the SR and the combination activates the myofilaments (MF). Elevated [Ca]i stimulates Ca removal from the cytosol by the SR Ca‐ATPase (modulated by phospholamban, PLB), the Na/Ca exchange, the sarcolemmal Ca‐ATPase, and the mitochondrial uniporter. As [Ca]i declines Ca dissociates from the myofilaments allowing relaxation. The sarcolemmal (Na + K)ATPase and Na/H exchange are also indicated, along with other key mitochondrial ion transporters.

Modified from Bers 34


Figure 2.

Passive Ca buffering in cardiac myocytes. The six curves represent different estimates of how passive Ca buffering changes as [Ca]i increases from 100 to 1500 nM. These values are indicated as increments above the amount of Ca bound at 100 nM [Ca]i (see also Table 2 and discussion in text). Fabiato‐Orig Fast, Berlin‐Fast and Hove‐Madsen‐Equilibrium are taken directly from values in the relevant papers 29,104,145. Fab/Bers‐New Fast updates some binding constants from the original estimates of Fabiato 104, as described in the text. Berlin‐Fast+MF Ca/Mg adds known myofilament sites at which Ca and Mg compete. Calculated Total‐New refers to values resulting from constants in Table 2.



Figure 3.

Measurement of fast Ca buffering in a voltage clamped rat ventricular myocyte. Cells were exposed to 10 μM thapsigargin to prevent SR Ca uptake and were in Na‐free conditions (inside and out) to prevent Na/Ca exchange. A. Voltage clamp pulses (200 ms from −40 to 0 mV) activated ICa and produced nearly step‐like increases in [Ca]i. B: The integral of the ICa from the fourth pulse in A (∫ ICa follows the kinetics of the rise in [Ca]i. The sag in the [Ca]i trace may represent slow buffering that is not in phase with Ca entry via ICa. The “ S” indicates where [Ca]i would be predicted to fall to eventually if the slow buffering by myofilament Ca/Mg sites are included with the fast buffering measured (see text). The points marked [A] and [B] in panel A are the predicted final settling points for [Ca]i if the slow buffering is included for the total ∫ICa throughout the train (as for S) and when the equilibrium buffering from Fig 2 and Table 2 is assumed, respectively

Modified from Berlin et al. 29


Figure 4.

Free and total Ca requirements for myofilament activation in cardiac myocytes. A: Two different sets of parameters are used for the Hill equation (100/(1 + (Km/[Ca])n) describing the steady‐state relationship between force and [Ca]. The solid curve (Km = 600 nM and n = 4) is more consistent with recent estimates in intact ventricular muscle, whereas the lower dotted curve (Km = 1000 nM and n = 2) is closer to traditional data from skinned muscle fibers (see text). B. The same relationships as in A, but the free [Ca] axis has been altered to account for the cytosolic Ca buffering relationship (Calculated‐Total‐New in Figure 2 and 2). Thus, this predicts the force as a function of the total amount of Ca added to the cytosol from a resting level of 150 nM free [Ca]. Coincidentally, almost the same curves are obtained if resting [Ca]i is assumed to be 100 nM and the intermediate buffering curve from Figure 2 is used (Berlin‐Fast+MF Ca/Mg).



Figure 5.

Cytosolic Ca buffering and transport during a Ca transient. A. A Ca transient is calculated from the product of a rising and declining exponential which rises from 100 to 744 nM and back with time constants (τ) of 30 ms (rising) and 300 ms (falling). This was used as a driving function to calculate the time‐dependent changes in binding to various cytosolic buffers (B) using the rate constants and concentrations in Table 3 (TnC is troponin C, SL is total sarcolemmal site in Table 3, CaM is calmodulin, TnC‐Ca/Mg and My‐Ca/Mg are Ca binding to Ca/Mg sites on TnC and myosin). The change in total cytosolic [Ca] (ΔTotal CaCyt in A) is the sum of the curves in B and the free [Ca]i curve. C: Ca fluxes associated with the Ca transient. ICa activation was calculated by a rising (τ = 3 ms) and falling exponential (τ = 40 ms), peaking at 6.8 pA/pF and bringing in 16 μmol/l cytosol. The SR Ca‐pump rate was 210/(1 + (300 nM/[Ca]i)2) in mol/l cytosol. The SR Ca leak was initiated to counterbalance the SR Ca‐pump rate at rest, and changed linearly as a function of free [Ca] in the SR (based on intra‐SR Ca buffering with a Kd = 600 μM and Bmax equivalent to 180 μmol/l cytosol in an SR volume occupying 3.5% of cell volume). The SR Ca release flux was taken as the residual Ca flux required to produce the driving Ca transient.



Figure 6.

Currents via T‐and L‐type Ca channels. A. Barium currents (with 115 mM Ba) induced by depolarizations to various test potentials from holding potentials of 80 or 30 mV. The Em protocol is shown in the top trace, the currents in the middle trace, and the difference between these currents in the bottom panel. Peak IBa from 30 mV is attributed to L‐type Ca channels and the additional transient difference current activated from 80 mV is attributed to T‐type Ca channels. B: ICa in dog Purkinje fiber cell with 2 mM Ca as charge carrier. The more positive Em required in IBa is due to the higher surface potential in 2 mM Ca vs 115 mM Ba. C: ICa from several species (from 80 to 100 mV), where the hump at ∼−40 mV is due to T‐type current and differs among the tissues studied. Dog Purkinje (ref. 139) and rabbit ventricle (G. M. Briggs and D. M. Bers, unpublished) are in 2 mM Ca. Dog atrium (ref. 27) and guinea‐pig ventricle (from Mitra & Morad, 218) are with 5 mM Ca, but are shifted by 10 mV to compensate for surface potential differences.

From Bean 27, with permission. From Hirano et al. 139, with permission. From Bers 34 with permission)


Figure 7.

Ca channel inactivation with different charge carriers. Normalized current amplitudes measured at 0 mV (except Ins at −30 mV to match activation state). ICa was recorded under both perforated patch (allowing normal SR Ca release and Ca transients) and ruptured patch with cells dialyzed with 10 mM EGTA (to prevent global Ca transients). IBa was also recorded with ruptured patch (with 10 mM EGTA in the pipette). Extracellular [Ca] and [Ba] were both 2 mM and Ins was measured in divalent‐free conditions (10 mM EDTA inside and out) with [Na]o at 20 mM and [Na]i at 10 mM. Peak currents were 1370, 808, 780, and 5200 pA and were attained at 5, 7, 10, and 14 ms for ICa (perforated), ICa (ruptured), IBa and Ins respectively. Halftimes of current decline were 17, 37, 161, and >500 ms, respectively.

Recordings made by Dr. W. Yuan


Figure 8.

Ca‐dependent inactivation by SR Ca release and recovery of ICa. Long (4 sec.) voltage clamp pulses activated ICa and SR Ca release and Ca transient in Na‐free conditions and 5.4 mM [Ca]o. The pulse labeled a was the first pulse after depletion of the SR and b was after 5–10 pulses when the SR was loaded. The smaller Ca transient produces less ICa inactivation. The larger Ca transient produces marked inactivation, but the current partially recovers as [Ca]i declines. The lower difference traces show the similarity between the ICa inactivation and Cai transient

Modified from Sipido et al. 311, with permission


Figure 9.

Ca currents measured during square pulse and action potential clamp in rat and rabbit ventricular myocytes. Command Em waveforms were either the traditional 200 ms square depolarization to 0 mV or an action potential (AP), which was recorded from normal rat and rabbit ventricular myocytes under more physiological conditions (i.e. normal Ca Na and K concentrations, ref 358). ICa was then recorded under conditions where all other currents were blocked (e.g. Na‐free and Cs‐rich inside and out). With the action potential waveform, the peak ICa is smaller in both species and occurs later (middle panel). The ICa integral (lower panels) for the AP clamp vs the square, pulses are smaller for the rat but larger for the rabbit (see ref. 358 for additional information).



Figure 10.

Simulations of guinea‐pig ventricular action potential using Oxsoft Heart (v 4.5). This comprehensive simulation created by Professor Denis Noble and colleagues (at Oxford, UK) calculates the Em and many of the ionic currents that are known to flow during the action potential. Shown here are the Em (top), ICa, INa/Ca (middle panel), [Ca]i, and force (lower panel) for two different initial [Na]i values (5 and 8 mM).



Figure 11.

Na/Ca exchange current in intact ventricular myocytes recorded under controlled conditions. A. Very slow speed recording of Em (top) and INa/Ca where the spikes are from ramp depolarizations that were used to generate the current voltage relationships in B, C, and D at the times indicated (a–h). The bars in A are when [Na]o was applied (replacing Li) to activate INa/Ca ([Ca]o was 1 mM throughout). At the arrow [Ca]i was increased from nominally Ca free to 430 nM with 42 mM EGTA and 140 CsCl in the dialyzing pipette throughout. Ouabain, Ba, Cs, D600 and tetraethylammonium were used to inhibit other ionic currents. Application of [Na]o stimulated INa/Ca only after [Ca]i was raised. The gradual decline in INa/Ca (d–g) was supposed to be due to depletion of [Ca]i.

From Kimura et al. 170, with permission


Figure 12.

Current–voltage relationships for INa/Ca recorded in excised giant patches from guinea pig ventricular myocytes. The patches were treated with chymotrypsin to remove INa/Ca inactivation and all other known currents were blocked 210. A: With constant [Ca]o (2 mM), [Na]o (150 mM) and [Ca]i (1 μM), the [Na]i was varied from 5 to 100 mM. The lower 3–4 curves probably reflect the INa/Ca expected in intact cells when [Ca]i is high (e.g. peak systole). B: The same conditions as A, except [Ca]i is reduced to 100 nM, comparable to diastolic [Ca]i.

From Matsuoka and Hilgemann 210, with permission


Figure 13.

Na/Ca exchange currents and [Ca]i in a guinea pig ventricular myocyte. The Ca transient (using fura‐2) show that Ca influx via Na/Ca exchange can bring in large quantities of Ca during large sustained depolarizations (A and B). The [Na] in the dialyzing pipette was 7.5 mM. Outward currents during the depolarization were off‐scale. C: The [Ca]i‐dependence of the “ tail” current was observed upon repolarization to −80 mV. Other ionic currents are blocked by Cs, tetraethylammonium, verapamil, and ryanodine.

From Barcenas‐Ruiz et al., 13, with permission


Figure 14.

Model of the Na/Ca exchanger based on recent work by Nicoll et al. 236 and Iwamoto et al.157. The structure is consistent with 9 membrane spanning segments and a glycosylation site (CH2O). The large cytoplasmic loop contains the Na/Ca exchange inhibitory peptide (XIP) domain (also associated closely with the region responsible for Na‐dependent inactivation) and also contains the site for allosteric regulation by [Ca]i*. The two homologous α repeats (α‐1 and α‐2) are also indicated.



Figure 15.

Na/Ca exchange reversal potential (ENa/Ca) during the rabbit ventricular action potential. This schematic diagram shows how ENa/Ca is expected to change during the action potential for two different levels of intracellular Na activity (aNai = 5 and 8 mM), where aNai is roughly 0.78 · [Na]i. When Em is positive to ENa/Ca, Ca influx via the Na/Ca exchange is thermodynamically favored (shaded areas). When Em is negative to ENa/Ca, Ca extrusion is favored. Resting [Ca]i =150 nM, [Ca]o = 2 mM and aNao = 110 mM for both traces and aNai and peak [Ca]i are as indicated. The [Ca]i trace reaches a peak 40 msec after the action potential begins.

After Bers 32,34


Figure 16.

Extracellular [Ca] depletion in rat‐rabbit ventricular muscle. Changes in [Ca]o were measured with double‐barreled Ca‐selective microelectrodes during individual contractions in rabbit (A) and rat (B) ventricular muscle (0.5 Hz, 30°C). The traces show [Ca]o (top) and tension (bottom) in the absence and presence of 10 mM citrate (which limits [Ca]o depletion by buffering [Ca]. The bath [Ca]o was 0.5 mM and is indicated by the dotted line.

A is modified from Shattock and Bers 309 and composite from Bers 34, with permission


Figure 17.

Changes in Na/Ca exchange driving force during action potential in rat and rabbit ventricle. ENa/Ca is expected to change during the action potential and Ca transient in rabbit and rat ventricle (top). The estimated changes in the net electrochemical driving force for Na/Ca exchange (ENa/Ca – Em) are shown in the bottom panel. Na/Ca exchange stoichiometry of 3Na:1Ca was assumed, the aNai values were measured 309 and, for simplicity, the Ca transient accompanying the contraction, has been assumed to be the same for both species. Resting [Ca]i was assumed to be 150 nM, rising to a peak of 1 μM, 40 msec after the upstroke of the action potential. Note the similarity between the lower panels and the [Ca]o traces in Figure. 16.

Modified from Shattock and Bers 309 and Bers 34, with permission


Figure 18.

Caffeine‐induced Ca transient and Na/Ca exchange current in a ferret ventricular myocyte. A: After a steady‐state series of voltage clamp pulses, 10 mM caffeine was rapidly and continuously applied to release SR Ca. Experimental conditions blocked most other ionic currents (e.g. Cs inside and out) and indo‐1 was used as the Ca indicator. The inward INa/Ca rises to a peak before the Ca transient. B: Instantaneous INa/Ca from panel A is plotted as a function of the global [Ca]i reported by indo‐1. The hysteresis shows that during the rising phase of the Ca transient, more inward INa/Ca is observed for a given [Ca]i than during relaxation. The relationship between INa/Ca and [Ca]i during relaxation was fit to a linear regression (for [Ca]i between 300 and 450 nM (dashed line; see also Fig. 13C).

Experiment performed by Dr. Li Li. C. INa/Ca and [Ca]i data are reproduced from panel A and the subsarcolemmal [Ca]i sensed by the Na/Ca exchanger ([Ca]Na/CaX) based on the linear regression in panel B is also shown


Figure 19.

Ca transport by the SR Ca‐ATPase based on literature values (see text). Ca pump rates were calculated based on Eq. 3 using Vmax = 200 μmol/liter cytosol, Km = 600 nM and n= 2. Activation of protein kinase A (PKA) decreases the Km to 250 nM, but does not affect Vmax.



Figure 20.

Measurement of SR Ca uptake rate in digitonin permeabilized myocytes. A. Free [Ca] was measured simultaneously with indo‐1 and Ca electrode. Ca was added at time 0 to a suspension of rabbit ventricular myocytes permeabilized with digitonin and incubated with 25 μM ruthenium red, 12.5 mM creatine phosphate, and 10 mM oxalate. B: The change in free [Ca] (Ca, solid line, from A), protein bound Ca (Ca‐Pn), Ca‐oxalate (Ca‐Ox), Ca‐indo (Ca‐In), and the sum of free and bound Ca (ΣCa) after addition of 14 μM total Ca. After an initial increase in free [Ca], Ca uptake by the SR lowered the free [Ca] back to the level before the Ca addition. The Ca bound to protein, oxalate, and indo‐1 was calculated from the passive Ca binding to permeabilized myocytes and indo‐1 144,145. C. The Ca uptake rate was calculated as the rate of change in the total non‐SR Ca with time (i.e. dΣCa/dr). D: [Ca] dependence of Ca uptake by the SR. The Ca uptake rate determined in C was plotted versus the corresponding free [Ca] measured in A. Data were fit with Eq. 3 to determine the maximal Ca uptake rate, Km, and Hill coefficient. Data are shown for addition of two different amounts of Ca to the myocyte suspension (14 μM total Ca = diamonds and 43 μM total Ca = circles.

From Hove‐Madsen and Bers 145, with permission


Figure 21.

Unidirectional SR Ca fluxes at rest. At rest when [Ca]SR is changing only very slowly, net SR Ca flux must be close to zero, implying the forward rate (VFor) is equal to the sum of reverse flux and leak from the SR (VRev + VLeak). The forward rate predicted from the Hill curve is 27 μmol/liter cytosol/sec. For a VLeak of 0.3 μmol/l cytosol/sec 24 the reverse flux must be 26.7 μmol/liter cytosol/sec.



Figure 22.

The influence of SR Ca leak on SR Ca content. The effect of different leak values on steady‐state SR Ca content with [Ca]i = 150 nM was calculated using Eq. 7 and 8 with Vmax = 207 μmol/liter cytosol/sec, Kmf = 300 nM, BmaxSR = 3.95 mM and KdSR = 600 μM. For pump inhibition by thapsigargin (TG) Vmax was decreased by 50%. For Ca‐pump stimulation by isoproterenol, Km was reduced by 50%. The SR Ca leak rate (0.3 μmol/liter cytosol/sec) measured by Bassani and Bers 24 is indicated. (See text and ref 117 for further discussion.)



Figure 23.

Ionic fluxes across the mitochondrial membrane. Ca enters via a uniport, down an electrical gradient established by the proton pump at bottom. Ca can be extruded by a Na/Ca antiport and Na is extruded by Na/H exchange thereby completing the cycle. Elevated cytoplasmic [Ca] can lead to elevated mitochondrial [Ca] and increased activity of mitochondrial dehydrogenases and NADH production.



Figure 24.

Mitochondria free [Ca] ([Ca]m) as a function of cytosolic [Ca] ([Ca]c). Increases of [Ca]c in rat ventricular myocytes were induced by reduction of extracellular [Na] (i.e. via Na/Ca exchange). Mean [Ca]c was measured using indo‐1 (loaded as the salt form) and [Ca]m was measured using indo‐1 (loaded as the AM form) with Mn quench of cytosolic indo‐1. Data are taken from Miyata et al., 220 and have been redrawn (without error bars) and including a broken line corresponding to [Ca]m = [Ca]c (slope = 1).



Figure 25.

Relaxation in rabbit ventricular myocyte with selective inhibition of Ca transporters. Normalized cell relaxations are shown and conditions were either (1) with all Ca transporters functional during relaxation of a steady‐state twitch (Tw), (2) preventing net SR Ca uptake during a caffeine‐induced contracture in NT (Caff), (3) additionally inhibiting Na/Ca exchange in 0Na, 0Ca solution (Caff,0Na,0Ca). To further analyze relaxation during Caff, 0Na, 0Ca, this was coupled with either (4) inhibition of mitochondrial Ca uptake with 1 μM FCCP + 1 μM oligomycin so only sarcolemmal Ca‐ATPase was functional (Caff,0Na,0Ca + FCCP) or (5) inhibition of the sarcolemmal Ca‐pump by elevating [Ca]o to 10 mM after pre‐depletion of [Na]i, so that only mitochondrial Ca uptake was functional (Caff,0Na,10Ca). All four Ca removal systems were also blocked by combining the caffeine application with 0Na, 10Ca and FCCP (after pre‐depletion of [Na]i). The traces are based on mean t1/2 values (shown along the traces without standard error values) for relaxation measured in experiments described by Bassani et al. 20.



Figure 26.

Twitch Ca transients in rabbit and rat ventricular myocytes with SR Ca‐ATPase blocked by thapsigargin. Ca transient were measured using indo‐1 fluorescence during electrically stimulated twitches in rabbit (A) and rat (B) ventricular myocytes before (Control) and after treatment with 2.5 μM thapsigargin (TG). Twitches were evoked 10 sec after switching to control solution (after 5–7 min pre‐perfusion with 0Na,0Ca solution). For the TG twitch the cells were exposed to TG for 2 min during the last part of the pre‐perfusion period. SR Ca load was maintained despite complete block of the SR Ca‐pump. Time constants (τ) of [Ca]i decline are shown

Modified from Bassani et al., 17


Figure 27.

Ca transport functions derived from Ca transients in intact rabbit and rat ventricular myocytes. Ca transport rates as functions of [Ca]i for the SR Ca‐ATPase, Na/Ca exchange and combined slow systems (mitochondrial and sarcolemmal Ca‐ATPase) were determined as described in the text and with respect to Eq. 9 and 10. Independent values obtained for JSR, JNa/CaX and JSlow respectively were: Vmax (in μmol/liter cytosol/sec) = 82, 46 and 3.9 for rabbit and 207, 27 and 4 for rat; Km (in nM) = 264, 316, and 362 in rabbit and 184, 257 and 268 in rat; n = 3.7, 3.7, 3.2 in rabbit and 3.9, 3.4 an 3.5 in rat.

Based on data from Bassani et al. 17


Figure 28.

Integrated Ca fluxes during twitch relaxation in rabbit and rat ventricular myocytes. Free [Ca]i during relaxation of a normal twitch was used as a driving function so that Ca flux via each system could be integrated over time using the [Ca]i dependence as described in Figure 27 and Eq. 9 and 10.

Modified from Bassani et al. 17


Figure 29.

Ca transients ICa during a twitch and INa/Ca used to measure SR Ca content. Ca transients and inward currents (ICa and INa/Ca) in a rat ventricular cardiomyocyte (dialyzed with 50 μM indo‐1) were recorded during the last two (of 8) conditioning voltage clamp pulses to 0 mV and during two rapid applications of 10 mM caffeine (bars, holding potential −70 mV). Amplitude and time scales for ICa (left) and INa/Ca (right) are different.

From Delbridge et al. 90, with permission


Figure 30.

Relative reliance on Ca influx in different cardiac muscle preparations. The effect of caffeine (10 mM) or ryanodine (100 nM) on steady‐state twitch contraction amplitude (0.5 Hz at 30°C or 23°C for frog) in various cardiac muscle preparations.

Data are from refs 33,34; modified from Bers 34


Figure 31.

Fractional SR Ca release during the twitch depends on both ICa trigger and SR Ca load in ferret ventricular myocytes. A. Cells were loaded to a constant level (by stimulation at 0.5 Hz). Extracellular [Ca] was changed just prior to the test contraction where the fraction of SR Ca release was measured. Thus increasing trigger (at constant SR Ca load) increases fractional release. Parallel voltage clamp experiments were done to evaluate how ICa changes over this range of [Ca]o. B. Using a constant trigger (with 2 mM [Ca]o) the SR of cells were Ca loaded to different levels by varying frequency (as indicated) and also increasing [Ca]o at the highest load. SR Ca content was determined by translating the peak and resting [Ca]i during caffeine‐induced contractures to total cytosolic [Ca], using values determined by Hove‐Madsen and Bers 145 for cytosolic Ca buffering.

Modified from Bassani et al. 19, with permission


Figure 32.

Rest decay and rest potentiation of twitches and SR Ca content in rabbit, rat and ferret myocytes. Both post‐rest twitches and caffeine‐induced contractures (Caff) were measured either in a modified normal Tyrode's solution (NT), when Na/Ca exchange was blocked during the rest (0Na,0Ca) or after pre‐depletion of [Na]i and rest in Ca‐free, 140 mM Na solution (0Cao) to stimulate Ca efflux via Na/Ca exchange. Caff data are indicative of SR Ca load. SR Ca content data (Caff) was fit with [Ca]SRC = A · exp(–t/τRD)(1 –exp{–t/τRF}) +B, where t is time, τRD is time constant, of rest decay of SR Ca content and τRF is a rest‐dependent filling time constant used to explain the delay in rest decay of SR Ca content. A and B are constants for scaling and baseline respectively. Twitch data were fit with the expression C[Ca]SRC(1–exp{–t/τFCC}) + D, where τECC: is the time constant for recovery of E‐C coupling and C and D are scaling and baseline constants. A small second exponential factor was also included to explain the slow decline in twitch amplitude in rabbit ventricle after long rests in 0Na, 0Ca.

Data are from Bers et al. 36 and Bassani and Bers 23 and have been combined and reanalyzed


Figure 33.

Refilling of SR Ca stores after depletion in rabbit ventricle. SR Ca content was measured by either caffeine‐induced contracture amplitude (in isolated myocytes at 23°C) or rapid cooling contractures (RCC) in trabeculae from rabbit ventricle (at 30°C). SR Ca was depleted by either a long rest (5 min) in normal solution (muscle) or a 10 sec application of 10 mM caffeine (myocyte). After SR Ca depletion, stimulation was started at 0.5 Hz and the SR Ca content was assessed after the beat number indicated.

Data are from Bers 33 and Bassani et al. 16


Figure 34.

Changes in contraction force with abrupt changes in frequency. Frequency of stimulation of a rabbit ventricular trabecula (at 30°C) was increased from 0.5 to 1.5 Hz and then returned to 0.5 Hz.



Figure 35.

Steady‐state force‐frequency relationships. The effect of frequency (from 0.5 to 2 Hz) on twitch force in rabbit (circles), rat (squares) and guinea pig (triangles) ventricular muscle. Data for rabbit and rat are at 30°C.

From Bers (33 and unpublished.] Data for guinea pig are at 36.5°C and were taken from Kurihara and Sakai 176. RCCs were initiated within 5 sec of the last steady‐state stimulated contraction. [Redrawn from Bers 34
References
 1. Adachi‐Akahane, S., L. Cleemann, and M. Morad. Cross‐signaling between L‐type Ca2+ channels and ryanodine receptors in rat ventricular myocytes. J. Gen. Physiol. 108: 435–454, 1996.
 2. Allen, D. G., P. G. Morris, C. H. Orchard and J. S. Pirolo. A nuclear magnetic resonance study of metabolism in the ferret heart during hypoxia and inhibition of glycolysis. J. Physiol. 361: 185–204, 1985.
 3. Almers, W. and E. W. McCleskey. Non‐selective conductance in calcium channels of frog muscle: calcium selectivity in a singlefile pore. J. Physiol. Lond. 353: 585–608, 1984.
 4. Antoniu, B., D. H. Kim, M. Morii, and N. Ikemoto. Inhibitors of Ca2+ release from the isolated sarcoplasmic reticulum. I. Ca2+ channel blockers. Biochim. Biophys. Acta 816: 9–17, 1985.
 5. Arreola, J., R. T. Driksen, R‐C. Shieh, D. J. Willford, and S‐S. Sheu. Ca2+ current and Ca2+ transients under action potential clamp in guinea pig ventricular myocytes. Am. J. Physiol. 261 (Cell Physiol. 30): C393–C397, 1991.
 6. Artman, M. Sarcolemmal Na+‐Ca2+ exchange activity and exchanger immunoreactivity in developing rabbit hearts. Am. J. Physiol. 263 (Heart Circ. Physiol. 32): H1506–H1513, 1992.
 7. Artman, M. Developmental Changes in Myocardial Inotropic Responsiveness. Austin, Texas‐R. G. Landes Company, 1994.
 8. Backx, P. H., W‐D. Gao, M. D. Azan‐Backx, and E. Marban. The relationship between contractile force and intracellular [Ca2+] in intact rat cardiac trabeculae. J. Gen. Physiol. 105: 1–19, 1995.
 9. Baker, P. F., M. P. Blaustein, A. L. Hodgkin, and R. A. Steinhardt. The influence of calcium on sodium efflux in squid axons. J. Physiol. (Lond). 200: 431–458, 1969.
 10. Balaguru, D., P. S. Haddock, J. L. Puglisi, D. M. Bers, W. A. Coetzee, and M. Artman. Role of the sarcoplasmic reticulum in contraction and relaxation of immature rabbit ventricular myocytes. J. Mol. Cell. Cardiol. 29: 2747–2757, 1997.
 11. Balke C. W., T. M. Egan, and W. G. Wier. Processes that remove calcium from the cytoplasm during excitation‐contraction coupling in intact rat heart cells. J. Physiol. (Lond). 474: 447–462, 1994.
 12. Banijamali, H. S., W. D. Gao, and H. E. D. J. Ter Keurs. Induction of calcium leak from the sarcoplasmic reticulum of rat cardiac trebeculae by ryanodine. Circulation 82: Supl III abstract‐215, 1990.
 13. Barcenas‐Ruiz L., D. J. Beuckelmann, and W. G. Wier. Sodium‐calcium exchange in heart: membrane currents and changes in [Ca2+],. Science 238: 1720–1722, 1987.
 14. Barth, E., G. Stammler, B. Speiser, and J. Schaper. Ultrastructural quantitation of mitochondria and myofilaments in cardiac muscle from 10 different animal species including man. J. Mol. Cell. Cardiol. 24: 669–681, 1992.
 15. Bassani, J. W. M., R. A. Bassani, and D. M. Bers. Ca2+ cycling between sarcoplasmic reticulum and mitochondria in rabbit cardiac myocytes. J. Physiol. Lond. 460: 603–621, 1993.
 16. Bassani, J. W. M., R. A. Bassani and D. M. Bers. Twitch‐dependent SR Ca accumulation and release in rabbit ventricular myocytes. Am. J. Physiol. 265 (Cell Physiol. 34): C533–C540, 1993.
 17. Bassani, J. W. M., R. A. Bassani, and D. M. Bers. Relaxation in rabbit and rat cardiac cells: species‐dependent differences in cellular mechanisms. J. Physiol. (Lond). 476: 279–293, 1994.
 18. Bassani, J. W. M., M. Qi, A. M. Samarel, and D. M. Bers. Contractile arrest increases SR Ca uptake and SERCa2 gene expression in cultured neonatal rat heart cells. Circ. Res. 74: 991–997, 1994.
 19. Bassani, J. W. M., W. Yuan, and D. M. Bers. Fractional SR Ca release is altered by trigger Ca and SR Ca content in cardiac myocytes. Am. J. Physiol. 268: C1313–C1319, 1995.
 20. Bassani, R. A., J. W. M. Bassani, and D. M. Bers. Mitochondrial and sarcolemmal Ca transport can reduce [Ca]i during caffeine contractures in rabbit cardiac myocytes. J. Physiol. (Lond). 453: 591–608, 1992
 21. Bassani, R. A., J. W. M. Bassani, and D. M. Bers. Relaxation in ferret ventricular myocytes: unusual interplay among calcium transport systems. J. Physiol. 476: 295–308, 1994.
 22. Bassani, R. A., J. W. M. Bassani, and D. M. Bers. Relaxation in ferret ventricular myocytes: role of the sarcolemmal Ca ATPase. Pflugers Arch. 430: 573–579, 1995.
 23. Bassani, R. A. and D. M. Bers. Na‐Ca exchange is required for rest‐decay but not for rest‐potentiation of twitches in rabbit and rat ventricular myocytes. J. Mol. Cell. Cardiol. 26: 1335–1347, 1994.
 24. Bassani, R. A. and D. M. Bers. Rate of diastolic Ca release from the sarcoplasmic reticulum of intact rabbit and rat ventricular myocytes. Biophys. J. 68: 2015–2022, 1995.
 25. Bassani, R. A., A. Mattiazzi, and D. M. Bers. CaMK‐II is responsible for activity‐dependent acceleration of relaxation in rat ventricular myocytes. Am. J. Physiol. 268 (Heart Circ. Physiol. 77): H703–H712, 1995.
 26. Baudet, S., R. Shaoulian, and D. M. Bers. Effects of thapsigargin and cyclopiazonic acid on twitch force and SR Ca content of rabbit ventricular muscle. Circ. Res. 73: 813–819, 1993.
 27. Bean, B. P. Two kinds of calcium channels in canine atrial cells. differences in kinetics, selectivity, and pharmacology. J. Gen Physiol. 86: 1–30, 1985.
 28. Bean, B. P. Classes of calcium channels in vertebrate cells. Annu. Rev. Physiol. 51: 367–384, 1989.
 29. Berlin, J. R., J. W. M. Bassani, and D. M. Bers. Intrinsic cytosolic calcium buffering properties of single rat cardiac myocytes. Biophys. J. 67: 1775–1787, 1994.
 30. Bers, D. M. Early transient depletion of extracellular [Ca] during individual cardiac muscle contractions. Am. J. Physiol. 244 (Heart Circ. Physiol. 13): H462–H468, 1983.
 31. Bers, D. M. Ca influx and SR Ca release in cardiac muscle activation during postrest recovery. Am. J. Physiol. 248 (Heart Circ. Physiol. 17): H366–H381, 1985.
 32. Bers, D. M. Mechanisms contributing to the cardiac inotropic effect of Na‐pump inhibition and reduction of extracellular Na. J. Gen. Physiol. 90: 479–504, 1987.
 33. Bers, D. M. SR Ca loading in cardiac muscle preparations based on rapid cooling contractures. Am. J. Physiol. 256 (Cell Physiol. 25): C109–C120, 1989.
 34. Bers, D. M. Excitation‐contraction coupling and cardiac contractile force. Dordrecht: Netherlands: Kluwer Academic Press, 1991.
 35. Bers D. M. Excitation‐Contraction Coupling and Cardiac Contractile Force, 2nd edition. Dordrecht: Netherlands, Kluwer Academic Press, (in press), 2001.
 36. Bers, D. M., L. A. Allen, and Y. Kim. Calcium binding to cardiac sarcolemma isolated from rabbit ventricular muscle: it's possible role in modifying contractile force. Am. J. Physiol. 251: C861–C871, 1986.
 37. Bers, D. M., R. A. Bassani J. W. M. Bassani, S. Baudet, and L. V. Hryshko. Paradoxical twitch potentiation after rest in cardiac muscle: increased fractional release of SR calcium. J. Mol. Cell. Cardiol. 25: 1047–1057, 1993.
 38. Bers, D. M. and J. H. B. Bridge. The effect of acetylstrophanthidin on twitches, microscopic tension fluctuations and cooling contractures in rabbit ventricular muscle. J. Physiol. (Lond). 404: 53–69, 1988.
 39. Bers, D. M. and J. H. B. Bridge. Relaxation of rabbit ventricular muscle by Na‐Ca exchange and sarcoplasmic reticulum Ca‐pump: ryanodine and voltage sensitivity. Circ. Res. 65: 334–342, 1989.
 40. Bers, D. M., J. H. B. Bridge, and K. T. MacLeod. The mechanism of ryanodine action in cardiac muscle assessed with Ca selective microelectrodes and rapid cooling contractures. Can. J. Physiol. Pharmacol. 65: 610–618, 1987.
 41. Bers, D. M., J. H. B. Bridge, and K. W. Sptizer. Intracellular Ca transients during rapid cooling contractures in guinea‐pig ventricular myocytes. J. Physiol. (Lond). 417: 537–553, 1989.
 42. Bers, D. M., D. M. Christensen, and T. X. Nguyen. Can Ca entry via Na‐Ca exchange directly activate cardiac muscle contraction?. J. Mol. Cell Cardiol. 20: 405–414, 1988.
 43. Bers, D. M., W. J. Lederer and J. R. Berlin. Intracellular Ca transients in rat cardiac myocytes: role of Na/Ca exchange in excitation‐contraction coupling. Am. J. Physiol. 258 (Cell Physiol. 27): C944–C954, 1990.
 44. Bers, D. M., K. D. Philipson, and A. Y. Nishimoto. Sodium‐calcium exchange and sidedness of isolated cardiac sarcolemmal vesicles. Biochim. Biophys. Acta. 601: 358–371, 1980.
 45. Bers, D. M. and V. M. Stiffel. The ratio of ryanodine:dihydropyridine receptors in cardiac and skeletal muscle and implications for E‐C coupling. Am. J. Physiol. 264 (Cell Physiol. 33): C1587–C1593, 1993.
 46. Bers, D. M., L. Li, H. Satoh, and E. McCall. Factors which control SR Ca release in intact ventricular myocytes. Ann. N. Y. Acad. Sci. 853: 157–177, 1998.
 47. Beuckelmann, D. J. Contributions of Ca2+ influx via L‐type Ca2+ current and Ca2+ release from the sarcoplasmic reticulum to [Ca]i transients in human myocytes. Basic Res. Cardiol. 92 (Suppl 1): 105–110, 1997.
 48. Beuckelmann, D. J. and W. G. Wier. Mechanism of release of calcium from sarcoplasmic reticulum of guinea pig cardiac cells. J. Physiol. 405: 233–255, 1988.
 49. Beuckelmann, D. J., and W. G. Wier. Sodium‐calcium exchange in guinea‐pig cardiac cells: exchange current and changes in intracellular Ca2+. J. Physiol. (Lond). 414: 499–520, 1989.
 50. Blater, L. A., and J. A. S. McGuigan. Free intracellular magnesium concentration in ferret ventricular muscle measured with ion selective micro‐electrodes. Q. J. Exp. Physiol. 71: 467–473, 1986.
 51. Blater L. A., J. Hüser, E. Ríos. Sarcoplasmic reticulum Ca2+ release flux underlying Ca2+ sparks in cardiac muscle. Proc Natl Acad Sci U S A. 94: 4176–4181, 1997.
 52. Blaustein, M. P. and W. J. Lederer. Sodium/calcium exchange: its physiological implications. Physiol. Rev. 79: 763–854, 1999.
 53. Blinks, J. R., C. B. Olson, B. R. Jewell, and P. Braveny. Influence of caffeine and other methylxanthines on mechanical properties of isolated mammalian heart muscle: evidence for a dual mechanism of action. Circ. Res. 30: 367–392, 1972.
 54. Block, B. A., T. Imagawa, K. P. Campbell, and C. Franzini‐Armstrong. Structural evidence for direct interaction between the molecular components of the transverse tubule/sarcoplasmic reticulum junction in skeletal muscle. J. Cell Biol. 107: 2587–2600, 1988.
 55. Boerth, S. R., D. B. Zimmer, and M. Artman. Steady‐state mRNA levels of the sarcolemmal Na+‐Ca2+ exchanger peak near birth in developing rabit and rat hearts. Circ. Res. 74: 354–359, 1994.
 56. Bouchard, R. A. and D. Bose. Analysis of the interval‐force relationship in rat and canine ventricular myocardium. Am. J. Physiol. 257 (Heart Circ. Physiol. 26): H2036–H2047, 1989.
 57. Brandes, R. and D. M. Bers. Intracellular Ca2+ increases NADH production and redox potential during elevated work in intact cardiac muscle. Circ. Res. 80: 82–87, 1997.
 58. Bridge, J. H. B. Relationships between the sarcoplasmic reticulum and transarcolemmal Ca transport revealed by rapidly cooling rabbit ventricular muscle. J. Gen. Physiol. 88: 437–473, 1986.
 59. Bridge J. H., P. R. Ershler, M. B. Cannell. Properties of Ca2+ sparks evoked by action potentials in mouse ventricular myocytes. J. Physiol. (Lond). 518: 469–478, 1999.
 60. Briggs, F. N., K. F. Lee, A. W. Wechsler, and L. R. Jones. Phospholamban expressed in slow‐twitch and chronically stimulated fast‐twitch muscles minimally affects calcium affinity of sarcoplasmic reticulum Ca2+‐ATPase. J. Biol. Chem. 267: 26056–26061, 1992.
 61. Brillantes, A. B., S. Bezprozvannaya, and A. R. Marks. Developmental and tissue‐specific regulation of rabbit skeletal and cardiac muscle calcium channels involved in excitation‐contraction coupling. Circ. Res. 75: 503–510, 1994.
 62. Brillantes A. B., K. Ondrias, A. Scott, E. Kobrinsky, E. Ondriasova, M. C. Moschella, T. Jayaraman, M. Landers, B. E. Ehrlich, and A. R. Marks. Stabilization of calcium release channel (ryanodine receptor) function by FK‐506 binding protein. Cell 77: 513–523, 1994.
 63. Butcher, R. W. and E. W. Sutherland. Adenosine 3′, 5′‐phosphate in biological materials. I. Purification and properties of cyclic 3′, 5′‐nucleotide phosphodiesterase and the use of this enzyme to characterize adenosine 3′, 5′‐phosphate in human urine. J. Biol. Chem. 237: 1244–1250, 1962.
 64. Callewaert, G., L. Cleemann, and M. Morad. Caffeine‐induced Ca2+ release activates Ca2+ extrusion via Na+‐Ca2+ exchanger in cardiac myocytes. Am. J. Physiol. 257 (Cell Physiol. 26): C147–C152, 1989.
 65. Campbell, K. P., T. Imagawa, J. S. Smith, and R. Coronado. Purified ryanodine receptor from skeletal muscle sarcoplasmic reticulum is the Ca2+‐permeable pore of the calcium release channel. J. Biol. Chem. 262: 16636–16643, 1987.
 66. Cannell, M. B., J. R. Berlin, and W. J. Lederer. Effect of membrane potential changes on the calcium transient in single rat cardiac muscle cells. Science 238: 1419–1423, 1987.
 67. Cannell M. B., H. Cheng, and W. J. Lederer. Spatial nonuniformities in [Ca2+]i during excitation‐contraction coupling in cardiac myocytes. Biophys. J. 67: 1942–1956, 1994.
 68. Cannell, M. B., H. Cheng, and W. J. Lederer. The control of calcium release in heart muscle. Science 268: 1045–1049, 1995.
 69. Capogrossi, M. C., A. A. Kort, H. A. Spurgeon, and E. G. Lakatta. Single adult rabbit and rat cardiac myocytes retain the Ca2+ and species‐dependent systolic and diastolic contractile properties of intact muscle. J. Gen. Physiol. 88: 589–613, 1986.
 70. Carafoli, E. Mitochondria, Ca2+ transport and the regulation of heart contraction and metabolism. J. Mol. Cell. Cardiol. 7: 83–89, 1975.
 71. Carafoli, E. Intracellular calcium homeostasis. Annu. Rev. Biochem. 56: 395–433, 1987.
 72. Carafoli, E. Biogenesis: plasma membrane calcium ATPase: 15 years of work on the purified enzyme. FASEB J. 8: 993–1002, 1994.
 73. Carafoli, E. and A. L. Lehninger. A survey of the interaction of calcium ions with mitochondria from different tissues and species. Biochem. J. 122: 618–690, 1971.
 74. Carafoli, E. and T. Stauffer. The plasma membrane calcium pump: functional domains, regulation of the activity, and tissue specificity of isoform expression. J. Neurobiol. 25: 312–324, 1994.
 75. Caroni, P. and E. Carafoli. An ATP‐dependent Ca2+‐pumping system in dog heart sarcolemma. Nature 283: 765–767, 1980.
 76. Caroni, P., L. Reinlib, and E. Carafoli. Charge movements during the Na+‐Ca2+ exchange in heart sarcolemmal vesicles. Proc. Natl. Acad. Sci. U.S.A. 77: 6354–6358, 1980.
 77. Caroni, P., and E. Carafoli. The Ca2+‐pumping ATPase of heart sarcolemma. J. Biol. Chem. 256: 3263–3270, 1981.
 78. Caroni, P. and E. Carafoli. Regulation of Ca2+‐pumping ATPase of heart sarcolemma by a phosphorylation‐dephosphorylation process. J. Biol. Chem. 256: 9371–9373, 1981.
 79. Chacon, E., H. Ohata, I. S. Harper, D. R. Trollinger, B. Herman, and J. J. Lemasters. Mitochondrial free calcium transients during excitation‐contraction coupling in rabbit cardiac myocytes. FEBS Lett., 382: 32–36, 1996.
 80. Chen, W., C. Steenbergen, L. A. Levy, J. Vance, R. E. London, and E. Murphy. Measurement of free Ca2+ in sarcoplasmic reticulum in perfused rabbit heart loaded with 1,2‐bis(2‐amino‐5,6‐difluorophenoxy) ethane‐N,N,N'N'‐tetraacetic acid by 19F NMR. J. Biol. Chem. 271: 7398–7403, 1996.
 81. Cheng H., W. J. Lederer, M. B. Cannell. Calcium sparks: elementary events underlying excitation‐contraction coupling in heart muscle. Science 262: 740–744, 1993.
 82. Clarke, D. M., T. W. Loo, G. Inesi, and D. H. MacLennan. Location of high affinity Ca2+‐binding sites within the predicted transmembrane domain of the sarcoplasmic reticulum Ca2+‐ATPase. Nature 339: 476–478, 1989.
 83. Cohen, C. J., H. A. Fozzard, and S.‐S. Sheu. Increase in intracellular sodium ion activity during stimulation in mammalian cardiac muscle. Circ. Res. 50: 651–662, 1982.
 84. Collins, A., A. V. Somlyo, and D. W. Hilgemann. The giant cardiac membrane patch method: stimulation of outward Na+‐Ca2+ exchange current by MgATP. J. Physiol. 454: 27–57, 1992.
 85. Coronado, R., J. Morisette, M. Sukhareva, and M. Vaughan. Structure and function of the ryanodine receptors. Am. J. Physiol. 266 (Cell Physiol. 35): C1485–C1504, 1994.
 86. Crespo, L. M., C. J. Grantham, and M. B. Cannell. Kinetics, stoichiometry and role of the Na‐Ca exchange mechanism in isolated cardiac myocytes. Nature 345: 618–621, 1990.
 87. Crompton, M., M. Capana and E. Carafoli. The sodium‐induced efflux of calcium from heart mitochondria: a possible mechanism for the regulation of mitochondrial calcium. Eur. J. Biochem. 69: 453–462, 1976.
 88. Crompton, M. The regulation of mitochondrial calcium transport in heart. Curr. Top. Memb. Transp. 25: 231–276, 1985.
 89. Crompton, M. The role of Ca2+ in the function and dysfunction of heart mitochondria. In: Calcium and the Heart, edited by G. A. Langer. New York: Raven Press, 1990: 167–198.
 90. Dani, A. M., A. Cittadini, and G. Inesi. Calcium transport and contractile activity in dissociated mammalian heart cells. Am. J. Physiol. 237 (Cell Physiol. 6): C147–C155, 1979.
 91. Danko, S., D. H. Kim, F. A. Sreter and N. Ikemoto. Inhibitors of Ca2+ release from the isolated sarcoplasmic reticulum. II. The effects of dantrolene on Ca2+ release induced by caffeine, Ca2+ and depolarization. Biochim. Biophys. Acta 816: 18–24, 1985.
 92. Delbridge, L. M., J. W. M. Bassani, and D. M. Bers. Steady‐state twitch Ca fluxes and cytosolic Ca buffering in rabbit ventricular myocytes. Am. J. Physiol. 270 (Cell Physiol. 39): C192–C199, 1996.
 93. Delbridge, L. M. D., H. Satoh, W. Yuan, J. W. M. Bassani, M. Qi, K. S. Ginsburg, A. M. Samarel, and D. M. Bers. Cardiac myocyte volume, Ca2+ fluxes and sarcoplasmic reticulum loading in pressure overload hypertrophy. Am. J. Physiol. 272 (Heart Circ. Physiol. 41): H2425–H2435, 1997.
 94. Deleon, M., Y. Wang, L. Jones, E. Perez‐Reyes, X. Wei, T. W. Soong, T. P. Snutch, and D. T. Yue. Essential Ca2+‐binding motif for Ca2+‐sensitive inactivation of L‐type Ca2+ channels. Science. 270: 1502–1506, 1995.
 95. Denton, R. M. and J. G. McCormack. On the role of the calcium transport cycle in heart and other mammalian mitochondria. FEBS Lett. 119: 1–8, 1980.
 96. Denton, R. M. and J. G. McCormack. Ca2+ transport by mammalian mitochondria and its role in hormone action. Am. J. Physiol. 249 (Endocrinol. Metab. 12): E543–E554. 1985.
 97. Denton, R. M. and J. G. McCormack. Ca2+ as a second messenger within mitochondria of the heart and other tissues. Annu. Rev. Physiol. 52: 451–466, 1990.
 98. Díaz, M. E., A. W. Trafford, S. C. O'Neill, and D. A. Eisner. Measurement of sarcoplasmic reticulum Ca2+ content and sarcolemmal Ca2+ fluxes in isolated rat ventricular myocytes during spontaneous Ca2+ release. J. Physiol. 501: 3–16, 1997.
 99. Difrancesco, D., and D. Noble. A model of cardiac electrical activity incorporating ionic pumps and concentration changes. Phil. Trans. R. Soc. Lond. B 307: 353–309, 1985.
 100. Dixon, D. A. and D. H. Haynes. Kinetic characterization of the Ca2+‐pumping ATPase of cardiac sarcolemma in four states of activation. J. Biol. Chem. 264: 13612–13622, 1989.
 101. Durkin, J. T., D. C. Ahrens, Y.‐C. E. Pan, and J. P. Reeves. Purification and amino‐terminal sequence of the bovine cardiac sodium‐calcium exchanger: evidence for the presence of a signal sequence. Arch. Biochem. Biophys. 290: 369–375, 1991.
 102. Edman, K. A. P. and M. Jóhannsson. The contractile state of rabbit papillary muscle in relation to stimulation frequency. J. Physiol. 254: 565–581, 1976.
 103. Egan, T. M., D. Noble, S. J. Noble, T. Powell, A. J. Spindler, and V. W. Twist. Sodium‐calcium exchange during the action potential in guinea‐pig ventricular cells. J. Physiol. 411: 639–661, 1989.
 104. Egger, M. and E. Niggli. Paradoxical block of the Na+‐Ca2+ exchanger by extracellular protons in guinea‐pig ventricular rmyocytes. J. Physiol. 523: 353–366, 2000
 105. El‐Sayed, M. F. and H. Gesser. Sarcoplasmic reticulum, potassium, and cardiac force in rainbow trout and plaice. Am. J. Physiol. 257 (Renal Fluid Electrolyte Physiol. 26): R599–R604, 1989.
 106. Fabiato, A. Calcium release in skinned cardiac cells: variations with species, tissues, and development. Federation Proc. 41: 2238–2244, 1982.
 107. Fabiato, A. Calcium‐induced release of calcium from the cardiac sarcoplasmic reticulum. Am. J. Physiol. 245 (Cell Physiol. 14): C1–C14, 1983.
 108. Fabiato, A. Simulated calcium current can both cause calcium loading in and trigger calcium release from the sarcoplasmic reticulum of a skinned canine cardiac Purkinje cell. J. Gen. Physiol. 85: 291–320, 1985.
 109. Feher, J. J. Unidirectional calcium and nucleotide fluxes in sarcoplasmic reticulum. Biophys. J. 45: 1125–1133, 1984.
 110. Fleischer, S. and M. Inui. Biochemistry and biophysics of excitation‐contraction coupling. Annu. Rev. Biophys. Chem. 18: 333–364, 1989.
 111. Frampton, J. E., S. M. Harrison, M. R. Boyett, and C. H. Orchard. Ca2+ and Na+ in rat ventricular myocytes showing different force frequency relationships. Am. J. Physiol. 261 (Cell Physiol. 30): C739–C750, 1991.
 112. Frank, J. S. and G. A. Langer. The myocardial interstitium: its structure and its role in ionic exchange. J. Cell Biol. 60: 586–601, 1974.
 113. Frankis, M. B. and G. E. Lindenmayer. Sodium‐sensitive calcium binding to sarcolemma‐enriched preparations from canine ventricles. Circ. Res. 55: 676–688, 1984.
 114. Friel, D. D. and B. P. Bean. Two ATP‐activated conductances in bullfrog atrial cells. J. Gen. Physiol. 91: 1–27, 1988.
 115. Fry, C. H., T. Powell, V. W. Twist, and J. P. T. Ward. Net calcium exchange in adult rat ventricular myocytes: an assessment of mitochondrial calcium accumulating capacity. Proc. R. Soc. Lond. 223: 223–238, 1984.
 116. Fry, C. H., T. Powell, V. W. Twist, and J. P. T. Ward. The effects of sodium, hydrogen and magnesium ions on mitochondrial calcium sequestration in adult rat ventricular myocytes. Proc. R. Soc. Lond. 223: 239–254, 1984.
 117. Fujioka Y., M. Komeda, and S. Matsuoka. Stoichiometry of Na+‐Ca2+ exchange in inside‐out patches excised from guinea‐pig ventricular myocytes. J. Physiol. 523: 339–351, 2000.
 118. Gatto, C. and M. A. Milanick. Inhibition of the red blood cell calcium pump by eosin and other fluorescein analogues. Am. J. Physiol. 264 (Cell Physiol. 33): C1577–C1586, 1993.
 119. Gao, W., P. H. Backx, M. Azan‐Backx, and E. Marban. Myofilament Ca2+ sensitivity in intact versus skinned rat ventricular muscle. Circ. Res. 74: 408–415, 1994.
 120. Ginsburg, K. S., C. R. Weber, and D. M. Bers. Control of maximum sarcoplasmic reticulum Ca load in intact ferret ventricular myocytes: effects of thapsigargin and isoproterenol. J. Gen. Physiol. 111: 491–504, 1998.
 121. Goldstein, M. A. and L. Traeger. Ultrastructural changes in postnatal development of the cardiac myocytes. In: The Developing Heart, edited by M. J. Legato. Boston: Martinus Nijhoff Publishing, 1985: 1–20.
 122. Grantham, C. J. and M. B. Cannell. Ca2+ influx during the cardiac action potential in guinea pig ventricular myocytes. Circ. Res. 79: 194–200, 1996.
 123. Griffiths, E. J., M. D. Stern, and H. S. Silverman. Measurement of mitochondrial calcium in single living cardiac myocytes by selective removal of cytosolic indo‐1. Am. J. Physiol. 273 (Cell Physiol. 42): C37–C44, 1997.
 124. Györke, S. and M. Fill. Ryanodine receptor adaptation:control mechanism of Ca2+‐induced Ca2+ release in heart. Science 260: 807–809, 1993.
 125. Györke S., V. Lukyanenko, and I. Györke. Dual effects of tetracaine on spontaneous calcium release in rat ventricular myocytes. J. Physiol. (Lond.) 500: 297–309, 1997.
 126. Hadley, R. W. and J. R. Hume. An intrinsic potential‐dependent inactivation mechanism associated with calcium channels in guinea‐pig myocytes. J. Physiol. 389: 205–222, 1987.
 127. Haddock, P. S., W. A. Coetzee, and M. Artman. Na+/Ca2+ exchange current and contractions measured under Cl–free conditions in developing rabbit hearts. Am. J. Physiol. 273 (Heart Circ. Physiol. 42): H837–H846, 1997.
 128. Hagiwara, N., H. Irisawa, and M. Kameyama. Contribution of two types of calcium currents to the pacemaker potentials of rabbit sino‐atrial node cells. J. Physiol., (Lond.) 359: 233–253, 1988.
 129. Haiech, J., B. Klee, and J. G. Demaille. Effects of cations on affinity of calmodulin for calcium:ordered binding of calcium ions allows the specific activation of calmodulin‐stimulated enzymes. Biochemistry 20: 3890–3897, 1981.
 130. Hansford, R. G. Relation between mitochondrial calcium transport and control of energy metabolism. Rev. Physiol. Biochem. Pharmacol. 102: 1–72, 1985.
 131. Hansford, R. G. Relation between cytosolic free Ca2+ concentration and the control of pyruvate dehydrogenase in isolated cardiac myocytes. Biochem. J. 241: 145–151, 1987.
 132. Harrison, S. M., and D. M. Bers. The influence of temperature on the calcium sensitivity of the myofilaments of skinned ventricular muscle from the rabbit. J. Gen. Physiol. 93: 411–427, 1989.
 133. Hess, P., J. B. Lansman and R. W. Tsien. Different modes of Ca channel gating behavior favored by dihydropyridine Ca agonists and antagonists. Nature 311: 538–544, 1984.
 134. Hess, P., J. B. Lansman, and R. W. Tsien. Calcium channel selectivity for divalent and monovalent cations. Voltage and concentration dependence of single channel current in ventricular heart cells. J. Gen. Physiol. 88: 293–319, 1986.
 135. Hilgemann, D. W. Numerical probes of sodium‐calcium exchange. In: Sodium‐Calcium Exchange, edited by T. J. A. Allen, D. Noble, and H. Reuter. Oxford: Oxford University Press, 1989: 126–152.
 136. Hilgemann, D. W. Regulation and deregulation of cardiac Na+‐Ca2+ exchange in giant excised sarcolemmal membrane patches. Nature 344: 242–245, 1990.
 137. Hilgemann, D. W. and R. Ball. Regulation of cardiac Na+,Ca2+ exchange and KATP potassium channels by PIP2. Science 273: 956–959, 1996.
 138. Hilgemann, D. W. and A. Collins. Mechanism of cardiac Na+‐Ca2+ exchange current stimulation by MgATP:possible involvement of aminophospholipid translocase. J. Physiol. (Lond.) 454: 59–82, 1992.
 139. Hilgemann, D. W., A. Collins, and S. Matsuoka. Steady‐state and dynamic properties of cardiac sodium‐calcium exchange. Secondary modulation by cytoplasmic calcium and ATP. J. Gen. Physiol. 100: 933–961, 1992.
 140. Hilgemann, D. W., S. Matsuoka, G. A. Nagel, and A. Collins. Steady‐state and dynamic properties of cardiac sodium‐calcium exchange. Sodium‐dependent inactivation. J. Gen. Physiol. 100: 905–932, 1992.
 141. Hilgemann, D. W., K. D. Philipson, and G. Vassort. Sodium‐Calcium Exchange:Proceedings of the Third International Conference. Ann. N.Y. Acad. Sci. 1996, vol. 779.
 142. Hirano, Y., H. A. Fozzard, and C. T. January. Characteristics of L‐ and T‐type Ca2+ currents in canine cardiac Purkinje cells. Am. J. Physiol. 256 (Heart Circ. Physiol. 25): H1478–H1492, 1989.
 143. Hoerter, J., F. Mazet, and G. Vassort. Perinatal growth of the rabbit cardiac cell:possible implications for the mechanism of relaxation. J. Mol. Cell. Cardiol. 13: 725–740, 1981.
 144. Holroyde, M. J., E. Howe, and R. J. Solaro. Modification of calcium requirements for activation of cardiac myofibrillar ATPase by cyclic AMP dependent phosphorylation. Biochim. Biophys. Acta 586: 63–69, 1979.
 145. Holroyde, M. J., S. P. Robertson, J. D. Johnson, R. J. Solaro, and J. D. Potter. The calcium and magnesium binding sites on cardiac troponin and their role in the regulation of myofibrillar adenosine triphosphatase. J. Biol. Chem. 255: 11688–11693, 1980.
 146. Horackova, M. and G. Vassort. Sodium‐calcium exchange in regulation of cardiac contractility. Evidence for an electrogenic, voltage‐dependent mechanism. J. Gen. Physiol. 73: 403–424, 1979.
 147. Hove‐Madsen, L. and D. M. Bers. Indo‐1 binding in permeabilized myocytes alters its spectral and Ca binding properties. Biophys. J. 63: 89–97, 1992.
 148. Hove‐Madsen, L. and D. M. Bers. Passive Ca buffering and SR Ca uptake in permeabilized rabbit ventricular myocytes. Am. J. Physiol. 264 (Cell Physiol. 33): C677–C686, 1993.
 149. Hove‐Madsen, L. and D. M. Bers. SR Ca uptake and thapsigargin sensitivity in permeabilized rabbit and rat ventricular myocytes. Circ. Res. 73: 820–828, 1993.
 150. Hryshko, L. V., V. M. Stiffel, and D. M. Bers. Rapid cooling contractures as an index of SR Ca content in rabbit ventricular myocyte. Am. J. Physiol. 257 (Heart Circ. Physiol. 26): H1369–1377, 1989.
 151. Deleted
 152. Hunter, D. R., R. A. Haworth, and H. A. Berkoff. Measurement of rapidly exchangeable cellular calcium in the perfused beating rat heart. Proc. Natl. Acad. Sci. U.S.A. 78: 5665–5668, 1981.
 153. Huynh, T. V., F. H. Chen, G. T. Wetzel, W. F. Friedman, and T. S. Klitzner. Developmental changes in membrane Ca2+ and K+ currents in fetal, neonatal, and adult rabbit ventricular myocytes. Circ. Res. 70: 508–515, 1992.
 154. Hymel, L., M. Inui, S. Fleischer, and H. Schindler. Purified ryanodine receptor of skeletal muscle sarcoplasmic reticulum forms Ca2+‐activated oligomeric Ca2+ channels in planar bilayers. Proc. Natl. Acad. Sci. U.S.A. 85: 441–445, 1988.
 155. Imagawa, T., J. S. Smith, R. Coronado, and K. P. Campbell. Purified ryanodine receptor from skeletal muscle sarcoplasmic reticulum is the Ca2+‐permeable pore of the calcium release channel. J. Biol. Chem. 262: 16636–16643, 1987.
 156. Inesi G. and L. De Meis. Regulation of steady state filling in sarcoplasmic reticulum. J. Biol. Chem. 264: 5929–5936, 1988.
 157. Inui, M., A. Saito, and S. Fleischer. Purification of the ryanodine receptor and identity with feet structures of junctional terminal cisternae of sarcoplasmic reticulum from fast skeletal muscle. J. Biol. Chem. 262: 1740–1747, 1987.
 158. Inui, M., A. Saito, and S. Fleischer. Isolation of the ryanodine receptor from cardiac sarcoplasmic reticulum and identity with the feet structures. J. Biol. Chem. 262: 15637–15642, 1987.
 159. Isenberg, G. and M. F. Wendt‐Gallitelli. Cellular mechanisms of excitation contraction coupling. In: Isolated Adult Cardiomyocytes, Volume II, edited by H. M. Piper and G. Isenberg. Boca Raton, Florida: CRC Press, 1989: 213–248.
 160. Iwamoto, T., T. Y. Nakamura, Y. Pan, A. Uehara, I. Imagawa, and M. Shigekawa. Unique topology of the internal repeats in the cardiac Na+/Ca2+ exchanger. FEBS Lett. 446: 264–268, 1999.
 161. James, P., M. Maeda, R. Fischer, A. K. Verma, J. Krebs, J. T. Penniston, and E. Carafoli. Identification and primary structure of a calmodulin binding domain of the Ca2+ pump of human erythrocytes. J. Biol. Chem. 263: 2905–2910, 1988.
 162. Janczewski, A. M. and E. G. Lakatta. Buffering of calcium influx by sarcoplasmic reticulum during the action potential in guinea‐pig ventricular myocytes. J. Physiol. 471: 343–363, 1993.
 163. Janczewski, A. M. and E. G. Lakatta. Thapsigargin inhibits Ca2+ uptake, and Ca2+ depletes sarcoplasmic reticulum in intact cardiac myocytes. Am. J. Physiol. 265 (Heart Circ. Physiol. 34): H517–H522, 1993.
 164. Jones, L. R., H. R. Besch, J. L. Sutko and J. T. Willerson. Ryanodine‐induced stimulation of net Ca + + uptake by cardiac sarcoplasmic reticulum vesicles. J. Pharmacol. Exp. Ther. 209: 48–55, 1979.
 165. Jung, D. W., K. Baysal, and G. P. Brierley. The sodium‐calcium antiport of heart mitochondria is not electroneutral. J. Biol. Chem. 270: 672–678, 1995.
 166. Kadambi, V. J., S. Ponniah, J. M. Harrer, B. D. Hoit, G. W. Dorn II, R. A. Walsh, and E. G. Kranias. Cardiac‐specific over‐expression of phospholamban alters calcium kinetics and resultant cardiomyocyte mechanics in transgenic mice. J. Clin. Invest. 97: 533–539, 1996.
 167. Kaftan E., A. R. Marks, and B. E. Ehrlich. Effects of rapamycin on ryanodine receptor/calcium release. Circ. Res. 78: 990–997, 1996.
 168. Kass, R. S. and M. C. Sanguinetti. Inactivation of calcium channel current in the calf cardiac Purkinje fiber. Evidence for voltage‐ and calcium‐mediated mechanisms. J. Gen. Physiol. 84: 705–726, 1984.
 169. Deleted
 170. Kawai, M. and M. Konishi. Measurement of sarcoplasmic reticulum calcium content in skinned mammalian cardiac muscle. Cell Calcium 16: 123–136, 1994.
 171. Kijima, Y., E. Ogunbunmi, and S. Fleischer. Drug action of thapsigargin on the Ca2+ pump protein of sarcoplasmic reticulum. J. Biol. Chem. 266: 22912–22918, 1991.
 172. Kimura, J., A. Noma, and H. Irisawa. Na‐Ca exchange current in mammalian heart cells. Nature 319: 596–597, 1986.
 173. Kimura, J., S. Miyamae, and A. Noma. Identification of sodium‐calcium exchange current in single ventricular cells in guinea pig. J. Physiol. 384: 199–222, 1987.
 174. Kirby, M. S., Y. Sagara, S. Gaa, G. Inesi, W. J. Lederer, and T. B. Rogers. Thapsigargin inhibits contraction and Ca transient in cardiac cells by specific inhibition of the sarcoplasmic reticulum Ca pump. J. Biol. Chem. 267: 12545–12551, 1992.
 175. Kirino, Y. and H. Shimizu. Ca2+‐induced Ca2+ release from fragmented sarcoplasmic reticulum: a comparison with skinned muscle fiber studies. J. Biochem. 92: 1287–1296, 1982.
 176. Kiss, E., G. Jakab, E. G. Kranias, and I. Edes. Thyroid hormone‐induced alterations in phospholamban protein expression. Circ. Res. 75: 245–251, 1994.
 177. Koss, K. L. and E. G. Kranias. Phospholamban: a prominent regulator of myocardial contractility. Circ. Res. 79: 1059–1063, 1996.
 178. Kranias, E. G. Regulation of calcium transport by protein phosphatase activity associated with cardiac sarcoplasmic reticulum. J. Biol. Chem. 260: 11006–11010, 1985.
 179. Kurihara, S. and T. Sakai. Effects of rapid cooling on mechanical and electrical responses in ventricular muscle of guinea pig. J. Physiol. 361: 361–378, 1985.
 180. Kuyayama, H. The membrane potential modulates the ATP‐dependent Ca2+ pump of cardiac sarcolemma. Biochim. Biophys. Acta 940: 295–299, 1988.
 181. Lai, F. A., H. Erickson, B. A. Block, and G. Meissner. Evidence for a junctional feet‐ryanodine receptor complex from sarcoplasmic reticulum. Biochem. Biophys. Res. Commun. 143: 704–709, 1987.
 182. Lai, F. A., H. F. Erickson, E. Rousseau, Q.‐Y. Liu, and G. Meissner. Purification and reconstitution of the calcium release channel from skeletal muscle. Nature 331: 315–319, 1988.
 183. Lai, F. A., K. Anderson, E. Rousseau, Q.‐Y. Liu, and G. Meissner. Evidence for a Ca2+ channel within the ryanodine receptor complex from cardiac sarcoplasmic reticulum. Biochem. Biophys. Res. Commun. 151: 441–449, 1988.
 184. Lai, F. A., M. Misra, L. Xu, H. A. Smith, and G. Meissner. The ryanodine receptor‐Ca2+ release channel complex of skeletal muscle sarcoplasmic reticulum. J. Biol. Chem. 264: 16776–16785, 1989.
 185. Langer, G. A. and A. Peskoff. Calcium concentration and movement in the diadic cleft space of the cardiac ventricular cell. Biophys. J. 70: 1169–1182, 1996.
 186. Langer, G. A., T. L. Rich, and F. B. Orner. Ca exchange under non‐perfusion‐limited conditions in rat ventricular cells:identification of subcellular compartments. Am. J. Physiol. 259 (Heart Circ. Physiol. 28): H592–H602, 1990.
 187. Langer, G. A. and T. L. Rich. Further characterization of the Na‐Ca exchange‐dependent Ca compartment in rat ventricular cells. Am. J. Physiol. 265 (Cell Physiol. 34): C556–C561, 1993.
 188. Deleted
 189. Lee, C. O. and H. A. Fozzard. Activities of potassium and sodium ions in rabbit heart muscle. J. Gen. Physiol. 65: 695–708, 1975.
 190. Lee, K. S., E. Marban, and R. W. Tsien. Inactivation of calcium channels in mammalian heart cells:joint dependence on membrane potential and intracellular calcium. J. Physiol. (Lond.) 364: 395–411, 1985.
 191. Legato, M. Cellular mechanisms of normal growth in the mammalian heart. II. A quantitative and qualitative comparison between the right and left ventricular myocytes in the dog from birth to five months of age. Circ. Res. 44: 263–279, 1979.
 192. Lehninger, A. L., E. Carafoli, and C. S. Rossi. Energy linked ion movements in mitochondrial systems. Adv. Enzymol. 29: 259–320, 1967.
 193. Lehninger, A. L. Ca2+ transport by mitochondria and its possible role in the cardiac excitation‐contraction‐relaxation cycle. Circ. Res. 34/35 (Suppl. III): 83–89, 1974.
 194. Levitsky, D. O., D. S. Benevolensky, T. S. Levchenko, V.N. Smirnov, and E. I. Chazov. Calcium‐binding rate and capacity of cardiac sarcoplasmic reticulum. J. Mol. Cell Cardiol. 13: 785–796, 1981.
 195. Lew, W. Y. W., L. V. Hryshko, and D. M. Bers. Dihydropyridine receptors are primarily functional L‐type Ca channels in rabbit ventricular myocytes. Circ. Res. 69: 1139–1145, 1991.
 196. Lewartowski, B. and K. Zdanowski. Net Ca2+ influx and sarcoplasmic reticulum Ca2+ uptake in resting single myocytes of the rat heart:comparison with guinea‐pig. J. Mol. Cell Cardiol. 22: 1221–1229, 1990.
 197. Li, L., G. Chu, E. G. Kranias, and D. M. Bers. Cardiac myocyte calcium transport in phospholamban knockout mouse: Relaxation and endogenous CaMKII effects. Am. J. Physiol. 274 (Heart Circ. Physiol. 43): H1335–H1347, 1998.
 198. Li, L., H. Satoh, K. S. Ginsburg, and D. M. Bers. The effects of CaMKII on cardiac excitation‐contraction coupling in ferret ventricular myocytes. J. Physiol. 501: 17–32, 1997.
 199. Li, Z., D. A. Nicoll, A. Collins, D. W. Hilgemann, A. G. Filoteo, J. T. Penniston, J. N. Weiss J. M. Tomich, and K. D. Philipson. Identification of a peptide inhibitor of the cardiac sarcolemmal Na+‐Ca2+ exchanger. J. Biol. Chem. 266: 1014–1020, 1991.
 200. London, B. and J. W. Krueger. Contraction in voltage‐clamped, internally perfused single heart cells. J. Gen. Physiol. 88: 475–505, 1986.
 201. López‐López, J. R., P. S. Shacklock, C. W. Balke, and W. G. Wier. Local, stochastic release of Ca2+ in voltage‐clamped rat heart cells: visualization with confocal microscopy. J. Physiol. (Lond.). 480: 21–29, 1994.
 202. López‐López JR., P. S. Shacklock, C. W. Balke, and W. G. Wier. Local calcium transients triggered by single L‐type calcium channel currents in cardiac cells. Science. 268: 1042–1045, 1995.
 203. Lu, Y.‐Z. and M. A. Kirchberger. Effects of a nonionic detergent on calcium uptake by cardiac microsomes. Biochemistry 33: 5056–5062, 1994.
 204. Lukyanenko, I. Györke, and S. Györke. Regulation of calcium release by calcium inside the sarcoplasmic reticulum in ventricular myocytes. Pflugers Arch. 432: 1047–1054, 1996.
 205. Lukyanenko V., I. Györke, S. Subramanian, A. Smirnov, T. F. Wiesner, S. Györke. Inhibition of Ca2+ sparks by ruthenium red in permeabilized rat ventricular myocytes. Biophys J. 79: 1273–1284, 2000.
 206. Luo, C.‐H. and Y. Rudy. A dynamic model of the cardiac ventricular action potential. I. Simulations of ionic currents and concentration changes. Circ. Res. 74: 1071–1096, 1994.
 207. Luo, W., B. W. Wolska, I. L. Grupp, J. M. Harrer, K. Haghighi, D. G. Ferguson, J. P. Slack, G. Grupp, T. Doetschman, R. J. Solaro, E. G. Kranias. Phospholamban gene dosage effects in the mammalian heart. Circ. Res. 78: 839–847, 1996.
 208. Lytton, J., M. Westlin, and R. Hanley. Thapsigargin inhibits the sarcoplasmic and endoplasmic reticulum Ca‐ATPase family of calcium pumps. J. Biol. Chem. 266: 17067–17071, 1991.
 209. McCormack, J. G., H. M. Browne, and N. J. Dawes. Studies on mitochondrial Ca2+‐transport and matrix Ca2+ using fura‐2‐loaded rat heart mitochondria. Biochim. Biophys. Acta 973: 420–427, 1989.
 210. Mahony, L. and L. R. Jones. Developmental changes in cardiac sarcoplasmic reticulum in sheep. J. Biol. Chem. 261: 15257–15265, 1986.
 211. Malécot, C. O., D. M. Bers, and B. G. Katzung. Biphasic contractions induced by milrinone at low temperature in ferret ventricular muscle: role of the sarcoplasmic reticulum and transmembrane Ca influx. Circ. Res. 59: 151–162, 1986.
 212. Marks, A. P., P. Tempst, K. S. Hwang, M. B. Taubman, M. Inui, C. Chadwick, S. Fleischer and B. Nadal‐Ginard. Molecular cloning and characterization of the ryanodine receptor/junctional channel complex cDNA from skeletal muscle sarcoplasmic reticulum. Proc. Natl. Acad. Sci. U.S.A. 86: 8683–8687, 1989.
 213. Matsuda, H. Sodium conductance in calcium channels of guinea pig ventricular cells induced by removal of external calcium ions. Pflugers Arch. 407: 465–475, 1986.
 214. Matsuoka, S. and D. W. Hilgemann. Steady‐state and dynamic properties of cardiac sdium‐calcium exchange. Ion and voltage dependencies of the transport cycle. J. Gen. Physiol. 100: 963–1001, 1992.
 215. Matsuoka, S., D. A. Nicoll, L. V. Hryshko, D. O. Levitsky, J. N. Weiss, and K. D. Philipson. Regulation of the cardiac Na+‐Ca2+ exchanger by Ca2+,. J. Gen. Physiol. 105: 403–420, 1995.
 216. Mattiazzi, A., L. Hove‐Madsen, and D. M. Bers. Protein kinase inhibitors reduce SR Ca transport in permeabilized cardiac myocytes. Am. J. Physiol. 267 (Heart Circ. Physiol. 36): H812–H820, 1994.
 217. Mela, L. Inhibition and activation of calcium transport in mitochondria. Effect of lanthanides and local anaesthetic drugs. Biochemistry 8: 2481–2486, 1969.
 218. Meissner, G. Ryanodine activation and inhibition of the Ca2+ release channel of sarcoplasmic reticulum. J. Biol. Chem. 261: 6300–6306, 1986.
 219. Meissner, G. and J. S. Henderson. Rapid calcium release from cardiac sarcoplasmic reticulum vesicles is dependent on Ca2+ and is modulated by Mg2+, adenine nucleotide, and calmodulin. J. Biol. Chem. 262: 3065–3073, 1987.
 220. Mejia‐Alvarez, R., C. Kettlun, E. Rros, M. Stern, M. Fill. Unitary Ca2+ current through cardiac ryanodine receptors in physiological ionic conditions. J. Gen. Physiol. 113: 177–186, 1999.
 221. Mitchell, P. and J. Moyle. Respiration‐driven proton translocation in rat liver mitochondria. Biochem. J. 105: 1147–1162, 1967.
 222. Mitra, R. and M. Morad. Two types of calcium channels in guinea pig ventricular myocytes. Proc. Natl. Acad. Sci. U.S.A. 83: 5340–5344, 1986.
 223. Miura, Y. and J. Kimuara. Sodium‐calcium exchange current. J. Gen. Physiol. 93: 1129–1145, 1989.
 224. Miyata, H., H. S. Silerman, S. J. Sollott, E. G. Lakatta, M. D. Stern, and R. G. Hansford. Measurement of mitochondrial free Ca concentration in living single rat cardiac myocytes. Am. J. Physiol. 261 (Heart Circ. Physiol. 30): H1123–H1134, 1991.
 225. Moore, C. L. Specific inhibition of mitochondrial Ca2+ transport by ruthenium red. Biochem. Biophys. Res. Commun. 42: 298–305, 1971.
 226. Morad, M. and Y. Goldman. Excitation‐contraction coupling in heart muscle: membrane control of development of tension. Prog. Biophys. Mol. Biol. 27: 257–313, 1973.
 227. Moreno‐Sanchez, R. and R. G. Hansford. Dependence of cardiac mitochondrial pyruvate dehydrogenase activity on intramitochondrial free Ca2+ concentration. Biochem. J. 256: 403–412, 1988.
 228. Deleted
 229. Mullins, L. J. The generation of electric currents in cardiac fibers by Na/Ca exchange. Am. J. Physiol. 236 (Cell. Physiol. 5): C103–C110, 1979.
 230. Murphy, E., C. C. Freudenrich, L. A. Levy, R. E. London, and M. Lieberman. Monitoring cytosolic free magnesium in cultured chicken heart cells by use of the fluorescent indicator furaptra. Proc. Natl. Acad. Sci. U.S.A. 86: 2981–2984, 1989.
 231. Murphy, E., C. Steenbergen, L. A. Levy, B. Raju, and R. E. London. Cytosolic free magnesium levels in ischemic rat heart. J. Biol. Chem. 264: 5622–5627, 1989.
 232. Nagasaki, K. and S. Fleischer. Modulation of the calcium release channel of sarcoplasmic reticulum by Adriamycin and other drugs. Cell Calcium 10: 63–70, 1989.
 233. Nakai, J., T. Imagawa, Y. Hakamata, M. Shigekawa, H. Takeshima and S. Numa. Primary structure and functional expression from cDNA of cardiac muscle ryanodine receptor/calcium release channel. FEBS Lett. 271: 169–177, 1990.
 234. Nakamura, Y., J. Kobayashi, J. Gilmore, M. Mascal, K. L. Rinehart, Jr., H. Nakamura, and Y. Ohizumi. Bromoeudistomin D, a novel inducer of calcium release from fragmented sarcoplasmic reticulum that causes contractions of skinned muscle fibers. J. Biol. Chem. 261: 4139–4142, 1986.
 235. Nakanishi, T. and J. M. Jarmakani. Developmental changes in myocardial mechanical function and subcellular organelles. Am. J. Physiol. 246 (Heart Circ. Physiol. 15): H615–H625, 1984.
 236. Nayler, W. G. and E. Fassold. Calcium accumulation and ATPase activity of cardiac sarcoplasmic reticulum before and after birth. Cardiovasc. Res. 11: 231–237, 1977.
 237. Negretti N., S. C. O'Neill, and D. A. Eisner. The relative contributions of different intracellular and sarcolemmal systems to relaxation in rat ventricular myocytes. Cardiovasc. Res. 1993: 27: 1826–1830.
 238. Nicoll, D. A., S. Longoni, and K. D. Philipson. Molecular cloning and functional expression of the cardiac sarcolemmal Na+‐Ca2+ exchanger. Science 250: 562–565, 1990.
 239. Nicoll, D. A., L. V. Hryshko, S. Matsuoka, J. S. Frank, and K. D. Philipson. Mutation of amino acid residues in the putative transmembrane segments of the cardiac sarcolemmal Na+‐Ca2+ exchanger. J. Biol. Chem. 271: 13385–13391, 1996.
 240. Nicoll, D. A., M. Ottolia, L. Lu, Y. Lu, and K. D. Philipson. A new topological model of the cardiac sarcolemmal Na+‐Ca2+ exchanger. J. Biol. Chem. 274: 910–917, 1999.
 241. Nicholls, D. G. and K. E. O. Akerman. Mitochondrial calcium transport. Biochim. Biophys. Acta 683: 57–88, 1982.
 242. Niggli, V., E. S. Adunyah, J. T. Penniston and E. Carafoli. Purified (Ca2+‐Mg2+)‐ATPase of the erythrocyte membrane. J. Biol. Chem. 256: 395–401, 1981.
 243. Nilius, B., P. Hess, J. B. Lansman, and R. W. Tsien. A novel type of cardiac calcium channel in ventricular cells. Nature 316: 443–446, 1985.
 244. Nowycky, M. C., A. P. Fox, and R. W. Tsien. Three types of neuronal calcium channel with different calcium agonist sensitivity. Nature 316: 440–443, 1985.
 245. Nuss, H. B., and S. R. Houser. T‐type Ca2+ current is expressed in hypertrophied adult feline left ventricular myocytes. Circ. Res. 73: 777–782, 1993.
 246. Odermatt A., K. Kurzydlowski, and D. H. MacLennan. The Vmax of the Ca2+‐ATPase of cardiac sarcoplasmic reticulum (SERCA2a) is not altered by Ca2+/calmodulin‐dependent phosphorylation or by interaction with phospholamban. J. Biol. Chem. 271: 14206–14213, 1996.
 247. O'Neill, S. C., J. G. Mill, and D. A. Eisner. Local activation of contraction in isolated rat ventricular myocytes. Am. J. Physiol. 258 (Cell Physiol. 27): C1165–C1168, 1990.
 248. Ohnishi, S. T. A method for studying the depolarization‐induced calcium release from fragmented sarcoplasmic reticulum. J. Biochem. 86: 1147–1150, 1979.
 249. Olivetti, G., P. Anversa, and A. Loud. Morphometric study of early postnatal development in the left and right ventricular myocardium of the rat. II. Tissue composition, capillary growth, and sarcoplasmic alterations. Circ. Res. 46: 503–512, 1980.
 250. Osaka, T. and R. W. Joyner. Developmental changes in calcium currents of rabbit ventricular cells. Circ. Res. 68: 788–796, 1991.
 251. Otsu, K., H. F. Willard, V. J. Khana, F. Zorzato, N. M. Green and D. H. MacLennan. Molecular cloning of cDNA encoding the Ca2+ release channel (ryanodine receptor) of rabbit cardiac muscle sarcoplasmic reticulum. J. Biol. Chem. 265: 13713–13720, 1990.
 252. Overend, C. L., C. S. O'Niell and D. A. Eisner. The effect of tetracaine on stimulated contractions, sarcoplasmic reticulum Ca2+ content and membrane current in isolated rat ventricular myocytes. J. Physiol. 507: 759–769, 1998.
 253. Page, E. Quantitative ultrastructural analysis in cardiac membrane physiology. Am. J. Physiol. 235 (Cell Physiol. 4): C147–C158, 1978.
 254. Page, E. and J. L. Buecker. Development of dyadic junctional complexes between sarcoplasmic reticulum and plasmalemma in rabbit left ventricular myocardial cells. Circ. Res. 48: 519–522, 1981.
 255. Page, E., L. P. McCallister and B. Power. Stereological measurements of cardiac ultrastructures implicated in excitation‐contraction coupling. Proc. Natl. Acad. Sci. U.S.A. 68: 1465–1466, 1971.
 256. Palade, P. Drug‐induced Ca2+ release from isolated sarcoplasmic reticulum. I. Use of pyrophosphate to study caffeine‐induced Ca2+ release. J. Biol. Chem. 262: 6135–6141, 1987.
 257. Palade, P. Drug‐induced Ca2+ release from isolated sarcoplasmic reticulum. II. Releases involving a Ca2+‐induced Ca2+ release channel. J. Biol. Chem. 262: 6142–6148, 1987.
 258. Palade, P. Drug‐induced Ca2+ release from isolated sarcoplasmic reticulum. III. Block of Ca2+‐induced Ca2+ release by inorganic polyamines. J. Biol. Chem. 262: 6149–6154, 1987.
 259. Palade, P., C. Dettbarn, D. Brunder, P. Stein, and G. Hals. Pharmacology of calcium release from sarcoplasmic reticulum. J. Bioenerg. Biomemb. 21: 295–320, 1989.
 260. Pan, B. S. and R. J. Solaro. Calcium‐binding properties of troponin C in detergent‐skinned heart muscle fibers. J. Biol. Chem. 262: 7839–7849, 1987.
 261. Pegg, W. and M. Michalak. Differentiation of sarcoplasmic reticulum during cardiac myogenesis. Am. J. Physiol. 252 (Heart Circ. Physiol. 21): H22–H31, 1987.
 262. Pelzer, D., S. Pelzer and T. F. McDonald. Properties and regulation of Ca channels in muscle cells. Rev. Physiol. Biochem. Pharmacol. 114: 107–207, 1990.
 263. Penefsky, Z. J. Studies on the mechanism of inhibition of cardiac muscle contractile tension by ryanodine. Pflugers Arch. 347: 173–184, 1974.
 264. Penefsky, Z. J. Perinatal development of cardiac mechanisms. In: Perinatal Cardiovascular Function, edited by N. Gootman and P. M. Gootman, NY, NY, 1983: 109–200.
 265. Peterson, B. Z., C. D. Demaria, and D. T. Yue. Calmodulin is the Ca2+ sensor for Ca2+‐dependent inactivation of 1‐type calcium channels. Neuron 22: 549–558, 1999.
 266. Philipson, K. D. Interaction of charged amphiphiles with Na+‐Ca2+ exchange in cardiac sarcolemmal vesicles. J. Biol. Chem. 259: 13999–14002, 1984.
 267. Philipson, K. D. and D. A. Nicoll. Sodium‐calcium exchange. A molecular perspective. Annu. Rev. Physiol. 62: 111–133, 2000.
 268. Philipson, K. D., and D. A. Nicoll. Molecular and kinetic aspects of sodium‐calcium exchange. Int. Rev. Cytol. 137C: 199–227, 1993.
 269. Philipson, K. D. Myocardial ion transporters. In: The Myocardium, 2nd Ed., edited by G. A. Langer. San Diego: Academic Press, 1997: 143–179.
 270. Pierce, G. N., K. D. Philipson, and G. A. Langer. Passive calcium‐buffering capacity of a rabbit ventricular homogenate preparation. Am. J. Physiol. 249 (Cell Physiol. 18): C248–C255, 1985.
 271. Pitts, B. J. R. Stoichiometry of sodium‐calcium exchange in cardiac sarcolemmal vesicles. J. Biol. Chem. 254: 6232–6235, 1979.
 272. Post, J. A., G. A. Langer, J. A. F. Op Den Kamp, and A. J. Verkleij. Phospholipid asymmetry in cardiac sarcolemma. Analysis of intact cells and “gas‐dissected” membranes. Biochim. Biophys. Acta 943: 256–266, 1988.
 273. Post, J. A. and G. A. Langer. Sarcolemmal calcium binding sites in heart: I. Molecular origin in “gas‐dissected” sarcolemma. J. Membr. Biol. 129: 49–57, 1992.
 274. Prabhu, S. D. and G. Salama. The heavy metal ions Ag+ and Hg2+ trigger calcium release from cardiac sarcoplasmic reticulum. Arch. Biochem. Biophys. 277: 47–55, 1990.
 275. Puglisi, J. L., R. A. Bassani, J. W. M. Bassani, J. N. Amin and D. M. Bers. Temperature and the relative contributions of Ca transport systems in cardiac myocyte relaxation. Am. J. Physiol. 270 (Heart Circ. Physiol. 39): H1772–H1778, 1996.
 276. Puglisi, J. L., W. Yuan, J. W. M. Bassani, and D. M. Bers. Ca2+ influx through Ca2+ channels in rabbit ventricular myocytes during action potential clamp: influence of temperature. Circ. Res. 85: e7–e16, 1999.
 277. Pytkowski, B. Rest‐ and stimulation‐dependent changes in exchangeable calcium content in rabbit ventricular myocardium. Basic Res. Cardiol. 84: 22–29, 1989.
 278. Qin N., R. Olcese, M. Bransby, T. Lin, and L. Birnbaumer: Ca2+‐induced inhibition of the cardiac Ca2+ channel depends on calmodulin. Proc. Natl. Acad. Sci. U.S.A. 96: 2435–2438, 1999.
 279. Reed, K. C. and F. L. Bygrave. A kinetic study of mitochondrial calcium transport. Eur. J. Biochem. 55: 497–503, 1975.
 280. Reddy, L. G., L. R. Jones R. C. Pace, and D. L. Stokes. Purified, reconstituted cardiac Ca2+‐ATPase is regulated by phospholamban, but not by direct phosphorylation with by Ca2+/calmodulin‐dependent protein kinase. J. Biol. Chem. 271: 14964–14970, 1996.
 281. Reeves, J. P., and J. L. Sutko. Sodium‐calcium exchange in cardiac membrane vesicles. Proc. Natl. Acad. Sci. U.S.A. 76: 590–594, 1979.
 282. Reeves, J. P., and J. L. Sutko. Sodium‐calcium exchange activity generates a current in cardiac membrane vesicles. Science 208: 1461–1464, 1980.
 283. Reeves, J. P., and C. C. Hale. The stoichiometry of the cardiac sodium‐calcium exchange system. J. Biol. Chem. 259: 7733–7739, 1984.
 284. Reeves, J. P., and P. Poronnik. Modulation of Na+‐Ca2+ exchange in sarcolemmal vesicles by intravesicular Ca2+. Am. J. Physiol. 252 (Cell Physiol. 21): C17–C23, 1987.
 285. Reeves, J. P. Na+/Ca2+ exchange and cellular Ca2+ homeostasis. J. Bioeng. Biomembr. 30: 151–160, 1998.
 286. Rega, A. F. and P. J. Garrahan. The Ca2+‐pump of plasma membranes. Boca Raton, FL: CRC Press, 1986: 173.
 287. Reimer, K. A. and R. B. Jennings. Myocardial ischemia, hypoxia, and infarction. In: The Heart and Cardiovascular System, edited by H. A. Fozzard et al. New York: Raven Press, 1986: 1133–1201.
 288. Reuter, H. and N. Seitz. The dependence of calcium efflux from cardiac muscle on temperature and external ion composition. J. Physiol. Lond.) 195: 45–70, 1968.
 289. Robertson, S. P., J. D. Johnson, and J. D. Potter. The time‐course of Ca2+ exchange with calmodulin, troponin, parvalbumin, and myosin in response to transient increases in Ca2+. Biophys. J. 34: 559–569, 1981.
 290. Rousseau, E., J. S. Smith, J. S. Henderson, and G. Meissner. Single channel and 45Ca2+ flux measurements of the cardiac sarcoplasmic reticulum calcium channel. Biophys. J. 50: 1009–1014, 1986.
 291. Rousseau, E., J. S. Smith, and G. Meissner. Ryanodine modifies conductance and gating behavior of single Ca2+ release channel. Am. J. Physiol. 253 (Heart Circ. Physiol. 22): C364–C368, 1987.
 292. Rousseau, E. and G. Meissner. Single cardiac sarcoplasmic reticulum Ca2+‐release channel: activation by caffeine. Am. J. Physiol. 256 (Heart Circ. Physiol. 25): H328–H333, 1989.
 293. Saito, A., M. Inui, M. Radermacher, J. Frank, and S. Fleischer. Ultrastructure of the calcium release channel of sarcoplasmic reticulum. J. Cell Biol. 107: 211–219, 1988.
 294. Sagara, Y. and G. Inesi. Inhibition of the sarcoplasmic reticulum Ca2+ transport ATPase by thapsigargin at subnanomolar concentrations. J. Biol. Chem. 266: 13503–13506, 1991.
 295. Salama, G. and J. Abramson. Silver ions trigger Ca2+ release by acting at the apparent physiological release site in sarcoplasmic reticulum. J. Biol. Chem. 259: 13363–13360, 1984.
 296. Sasaki, T., M. Inui, Y. Kimura, T. Kuzuya, and M. Tada. Molecular mechanism of regulation of Ca2+ pump ATPase by phospholamban in cardiac sarcoplasmic reticulum. J. Biol. Chem. 267: 1674–1679, 1992.
 297. Satoh H., L. A. Blatter, and D. M. Bers. Effects of [Ca]i, Ca2+ load and rest on Ca2+ spark frequency in ventricular myocytes. Am. J. Physiol. 272 (Heart Circ. Physiol. 41): H657–668, 1997.
 298. Satoh, H., L. M. Delbridge, L. A. Blatter, and D. M. Bers. Surface:volume relationship in cardiac myocytes studied with confocal microscopy and membrane capacitance measurements: species‐dependence and developmental effects. Biophys. J. 70: 1494–1504, 1996.
 299. Schatzmann, H. J. ATP dependent Ca2+ extrusion from human red cells. Experientia 22: 364–368, 1966.
 300. Schatzmann, H. J. The plasma membrane calcium pump of erythrocytes and other animal cells. In: Membrane Transport of Calcium, edited by E. Carafoli. London: Academic Press, 1982: 41–108.
 301. Schatzmann, H. J. The calcium pump of the surface membrane and of the sarcoplasmic reticulum. Annu. Rev. Physiol. 51: 473–485, 1989.
 302. Schiebler, T. and H. H. Wolff. Electronenmikroskopische Untersuchungen am Herzmuskel der Ratte wahrend der Entwicklung. Z. Zellforsch. Mikrosk. Anat. 69: 22–40, 1962.
 303. Schiefer A., G. Meissner, and Isenberg. Ca2+ activation and Ca2+ inactivation of cardiac sarcoplasmic reticulum Ca2+‐release channels. J. Physiol. 489: 337–348, 1995.
 304. Schouten, V. J. A. and H. E. D. J. ter Keurs. The force‐frequency relationship in rat myocardium. Pflugers Arch. 407: 14–17, 1986.
 305. Seguchi, M., J. A. Harding, and J. M. Jarmakani. Developmental change in the function of sarcoplasmic reticulum. J. Mol. Cell. Cardiol. 18: 189–195, 1986.
 306. Seidler, N. W., I. Jona, M. Vegh, and A. Martonosi. Cyclopiazonic acid is a specific inhibitor of the Ca2+‐ATPase of sarcoplasmic reticulum. J. Biol. Chem. 264: 17816–17823, 1989.
 307. Sham, J. S. K., L. R. Jones, and M. Morad. Phospholamban mediates the ã‐adrenergic‐enhanced Ca uptake in mammalian ventricular myocytes. Am. J. Physiol. 261 (Heart Circ, Physiol. 30): H1344–H1349, 1991.
 308. Shanne, F. A. X., A. B. Kane, E. E. Young, and J. L. Farber. Calcium dependence of toxic cell death: a final common pathway. Science 206: 700–702, 1979.
 309. Shannon, T. R. and D. M. Bers. Assessment of intra‐SR free [Ca] and buffering in rat heart. Biophys. J. 73: 1524–1531, 1997.
 310. Shannon, T. R., K. S. Ginsburg, and D. M. Bers. Reverse mode of the SR Ca‐pump and load‐dependent Ca decline in voltage clamped cardiac ventricular myocytes. Biophys. J. 78: 322–333, 2000.
 311. Shannon, T. R., K. S. Ginsburg, and D. M. Bers. Potentiation of fractional SR Ca release by total and free intra‐SR Ca concentration. Biophys. J. 78: 334–343, 2000.
 312. Shattock, M. J. and D. M. Bers. The inotropic response to hypothermia and the temperature‐dependence of ryanodine action in isolated rabbit and rat ventricular muscle: implications for E‐C coupling. Circ. Res. 61: 761–771, 1987.
 313. Shattock, M. J. and D. M. Bers. Rat vs. rabbit ventricle: Ca flux and intracellular Na assessed by ion‐selective microelectrodes. Am. J. Physiol. 256 (Cell Physiol. 25): C813–C822, 1989.
 314. Simmerman, H. K. B., J. H. Collins, J. L. Theiber, A. D. Wegener, and L. R. Jones. Sequence analysis of phospholamban. J. Biol. Chem. 261: 13333–13341, 1986.
 315. Sipido, K. R., G. Callewaert, and E. Carmeliet. Inhibition and rapid recovery of Ca2+ current during Ca2+ release from sarcoplasmic reticulum in guinea pig ventricular myocytes. Circ. Res. 76: 102–109, 1995.
 316. Sipido K. R. and W. G. Weir. Flux of Ca2+ across the sarcoplasmic reticulum of guinea pig cardiac cells during excitation‐contraction coupling. J. Physiol. (Lond.) 435: 605–630, 1991.
 317. Sitsapesan, R., R. A. P. Montgomery, K. T. MacLeod, and A. J. Williams. Sheep cardiac sarcoplasmic reticulum calcium‐release channels: modification of conductance and gating by temperature. J. Physiol. (Lond.) 434: 469–488, 1991.
 318. Sitsapesan R. and A. J. Williams. Regulation of the gating of the sheep cardiac sarcoplasmic reticulum Ca2+‐release channel by luminal Ca2+. J. Membr. Biol. 137: 215–226, 1994.
 319. Smith, J. S., T. Imagawa, J. Ma, M. Foll, K. P. Campbell, and R. Coronado. Purified ryanodine receptor from rabbit skeletal muscle is the calcium‐release channel of sarcoplasmic reticulum. J. Gen. Physiol. 92: 1–26, 1988.
 320. Smith, J. S., E. Rousseau, and G. Meissner. Calmodulin modulation of single sarcoplasmic reticulum Ca release channels from cardiac and skeletal muscle. Circ. Res. 64: 352–359, 1989.
 321. Deleted.
 322. Solaro, R. J. and F. N. Briggs. Estimating the functional capabilities of sarcoplasmic reticulum in cardiac muscle. Circ. Res. 34: 531–540, 1974.
 323. Sordahl, L. A. Effects of magnesium, ruthenium red and the antibiotic ionophore A‐23187 on initial rates of calcium uptake and release by heart mitochondria. Arch. Biochem. Biophys. 167: 104–115, 1975.
 324. Spurgeon, H. A., M. D. Stern, G. Baartz, S. Raffaeli, R. G. Hansford, A Talo, E. G. Lakatta, and M. C. Capogrossi. Simultaneous measurement of Ca2+, contraction and potential in cardiac myocytes. Am. J. Physiol. 258 (Heart Circ. Physiol. 27): H574–H586, 1990.
 325. Su, J. H. and W. G. L. Kerrick. Effects of halothane on caffeine‐induced tension transients in functionally skinned myocardial fibers. Pflugers Arch. 380: 29–34, 1979.
 326. Sutko, J. L., D. M. Bers, and J. P. Reeves. Postrest inotropy in rabbit ventricle: Na+‐Ca2+ exchange determines sarcoplasmic reticulum Ca2+ content. Am. J. Physiol. 250 (Heart Circ. Physiol. 19): H654–H661, 1986.
 327. Sutko, J. L. and J. T. Willerson. Ryanodine alteration of the contractile state of rat ventricular myocardium. Comparison with dog, cat and rabbit ventricular tissues. Circ. Res. 46: 332–343, 1980.
 328. Sutko, J. L. and J. L. Kenyon. Ryanodine modification of cardiac muscle responses to potassium free solutions. Evidence for inhibition of sarcoplasmic reticulum calcium release. J. Gen. Physiol. 82: 385–404, 1983.
 329. Sutko, J. L., K. Ito, and J. L. Kenyon. Ryanodine: a modifier of sarcoplasmic reticulum calcium release. Biochemical and functional consequences of its actions on striated muscle. Federation Proc. 44: 2984–2988, 1985.
 330. Suzuki, T. and J. H. Wang. Stimulation of bovine cardiac sarcoplasmic reticulum Ca2+ pump and blocking of phospholamban phosphorylation and dephosphorylation by a phospholamban monoclonal antibody. J. Biol. Chem. 261: 7018–7023, 1986.
 331. Deleted.
 332. Tada, M., M. A. Kirchberger, D. I. Repke, and A. M. Katz. The stimulation of calcium transport in cardiac sarcoplasmic reticulum by adenosine 3′:5′‐monophosphate‐dependent protein kinase. J. Biol. Chem. 249: 6174–6180, 1974.
 333. Takenaka, H., P. N. Adler, and A. M. Katz. Calcium fluxes across the membrane of sarcoplasmic reticulum vesicles. J. Biol. Chem. 257: 12649–12656, 1982.
 334. Takeshima, H., S. Hishimura, T. Matsumoto, H. Ishida, K. Kangawa, N. Minamino, H. Matsuo, M. Ueda, M. Hanaoka, T. Hirose, and S. Numa. Primary structure and expression from complementary DNA of skeletal muscle ryanodine receptor. Nature 339: 439–445, 1989.
 335. Terracciano, C. M. N. and K. T. MacLeod. Effects of acidosis on Na+/Ca2+ exchange and consequences for relaxation in guinea pig cardiac myocytes. Am. J. Physiol. 267 (Heart Circ. Physiol. 36): H477–H487, 1994.
 336. Terracciano, C. M. N. and K. T. MacLeod. Measurements of Ca2+ entry and sarcoplasmic reticulum Ca2+ content during the cardiac cycle in guinea pig and rat ventricular myocytes. Biophys. J. 72: 1319–1326, 1997.
 337. Terracciano, C. M. N. and K. T. MacLeod. Reloading of Ca2+‐depleted sarcoplasmic reticulum during rest in guinea pig ventricular myocytes. Am. J. Physiol. 271 (Heart Circ. Physiol. 40): H1814–H1822, 1996.
 338. Terracciano, C. M. N., R. U. Naqvi, and K. T. MacLeod. Effects of rest interval on the release of calcium from the sarcoplasmic reticulum in isolated guinea pig ventricular myocytes. Circ. Res. 77: 354–360, 1995.
 339. Thastrup, O., P. J. Cullen, B. K. Drobak, M. R. Hanley, A. P. Dawson. Thapsigargin, a tumor promoter, discharges intracellular Ca stores by specific inhibition of the endoplasmic reticulum Ca‐ATPase. Proc. Natl. Acad. Sci. U.S.A., 87: 2466–2470, 1990.
 340. Tinker A., A. R. G. Lindsay, and A. J. Williams. Cation conductance in the calcium release channel of the cardiac sarcoplasmic reticulum under physiological and pathophysiological conditions. Cardiovasc. Res. 27: 1820–1825, 1993.
 341. Toyofuku, T., K. Kurzydlowski, M. Tada, and D. H. MacLennan. Identification of regions in the Ca2+‐ATPase of sarcoplasmic reticulum that affect functional association with phospholamban. J. Biol. Chem. 268: 2809–2815, 1993.
 342. Toyofuku T., K. Kurzydlowski, N. Narayanan, and D. H. MacLennan. Identification of Ser38 as the site in cardiac sarcoplasmic reticulum Ca2+‐ATPase that is phosphorylated by Ca2+/calmodulin dependent protein kinase. J. Biol. Chem. 269: 26492–26496, 1994.
 343. Trafford, A. W., M. E. Díaz, N. Negretti, and D. A. Eisner. Enhanced Ca2+ current and decreased Ca2+ efflux restore sarcoplasmic reticulum Ca2+ content after depletion. Circ. Res. 81: 477–484, 1997.
 344. Trimm, J. L., G. Salama, and J. Abramson. Sulfhydryl oxidation induces rapid calcium release from sarcoplasmic reticulum vesicles. J. Biol. Chem. 261: 16092–16098, 1986.
 345. Tripathy A. and G. Meissner. Sarcoplasmic reticulum lumenal Ca2+ has access to cytosolic activation and inactivation sites of skeletal muscle Ca2+ release channel. Biophys. J. 70: 2600–2615, 1996.
 346. Tseng, G. Calcium current restitution in mammalian ventricular myocytes is modulated by intracellular calcium. Circ. Res. 63: 468–482, 1988.
 347. Tsien, R. W., P. Hess, E. W. McCleskey, and R. L. Rosenberg. Calcium channels:mechanisms of selectivity, permeation and block. Annu. Rev. Biophys. Chem. 16: 265–290, 1987.
 348. Valdivia, H. H., J. H. Kaplan, G. C. R. Ellis‐Davies, and W. J. Lederer. Rapid adaptation of cardiac ryanodine receptors:modulation by Mg2+ and phosphorylation. Science 267: 1997–1999, 1995.
 349. Varro, A., N. Negretti, S. B. Hester, and D. A. Eisner. An estimate of the calcium content of the sarcoplasmic reticulum in rat ventricular myocytes. Pflugers. Arch. 423: 158–160, 1993.
 350. Verma, A. K., A. Filoteo, D. R. Stanford, E. D. Wieben, J. T. Penniston, E. E. Strehler, R. Fischer, R. Heim, G. Vogel, S. Mathews, M.‐A. Strehler‐Page, P. James, T. Vorherr, J. Krebbs, and E. Carafoli. Complete primary structure of a human plasma membrane Ca2+ pump. J. Biol. Chem. 263: 14152–14159, 1988.
 351. Vercesi, A., B. Reynafarje, and A. L. Lehninger. Stoichiometry of H+ ejection and Ca2+ uptake coupled to electron transfer in rat heart mitochondria. J. Biol. Chem. 253: 6379–6385, 1978.
 352. Wagenknecht, T., R. Grassucci, J. Frank, A. Saito, M. Inui, and S. Fleischer. Three‐dimensional architecture of the calcium channel/foot structure of sarcoplasmic reticulum. Nature 338: 167–170, 1989.
 353. Weber, A. and R. Herz. The relationship between caffeine contracture of intact muscle and the effect of caffeine on reticulum. J. Gen. Physiol. 52: 750–759, 1968.
 354. Weber, C., R. K. S. Ginsburg, K. D. Philipson, T. R. Shannon and D. M. Bers. Allosteric regulation of Na/Ca exchange current by cytosolic Ca in intact cardiac myocytes. J. Gen. Physiol. 117: 119–131, 2001.
 355. Wendt, I. R. and D. G. Stephenson. Effects of caffeine on Ca‐activated force production in skinned cardiac and skeletal muscle fibres of the rat. Pflugers. Arch. 398: 210–216, 1983.
 356. Wier W. G. and C. W. Balke. Ca2+ release mechanisms, Ca2+ sparks, and local control of excitation‐contraction coupling in normal heart muscle. Circ. Res. 85: 770–776, 1999.
 357. Wier, W. G., T. M. Egan, J. R. López‐López, and C. W. Balke. Local control of excitation‐contraction coupling in rat heart cells. J. Physiol. (Lond.) 474: 463–471, 1994.
 358. Williams, A. J. and S. R. M. Holmberg. Sulmazole (AR‐L 115BS) activates the sheep cardiac muscle sarcoplasmic reticulum calcium‐release channel in the presence and absence of calcium. J. Membr. Biol. 115: 167–178, 1990.
 359. Wohlfart, B. Relationship between peak force, action potential duration and stimulus interval in rabbit myocardium. Acta Physiol. Scand. 106: 395–409, 1979.
 360. Wrzosek, A., H. Schneider, S. Grueninger, and M. Chiesi: Effect of thapsigargin on cardiac muscle cells. Cell. Calcium 13: 281–292, 1992
 361. Xu A., C. Hawkins, N. Narayanan. Phosphorylation and activation of the Ca‐pumping ATPase of cardiac sarcoplasmic reticulum by Ca/calmodulin‐dependent protein kinase. J. Biol. Chem. 268: 8394–8397, 1993.
 362. Ying, W. L., J. Emerson, M. J. Clarke, and D. R. Sanadi. Inhibition of mitochondrial calcium ion transport by oxo‐bridged dinuclear ruthenium amine complex. Biochemistry 30: 4949–4952, 1991.
 363. Yuan, W. and D. M. Bers. Ca‐dependent facilitation of cardiac Ca current is due to Ca‐calmodulin dependent protein kinase. Am. J. Physiol. 267 (Heart Circ. Physiol. 36): H982–H993, 1994.
 364. Yuan, W. and D. M. Bers. Protein kinase inhibitor H‐89 reverses forskolin stimulation of cardiac L‐type calcium current. Am. J. Physiol. 267 (Cell Physiol. 36): C651–C659, 1995.
 365. Yuan, W., K. S. Ginsburg, and D. M. Bers. Comparison of sarcolemmal Ca channel current in rabbit and rat ventricular myocytes. J. Physiol. (Lond.) 493: 733–746, 1996.
 366. Yue, D. T., D. Burkhoff, M. R. Franz, W. C. Hunter, and K. Sagawa. Postextrasystolic potentiation of the isolated canine left ventricle. Circ. Res. 56: 340–350, 1985.
 367. Yue, D. T., E. Marban, and W. G. Wier. Relationship between force and intracellular[Ca2+] in tetanized mammalian heart muscle. J. Gen. Physiol. 87: 223–242, 1986.
 368. Zhou, Z., M. A. Matlib, and D. M. Bers. Cytosolic and mitochondrial Ca2+ signals in patch clamped ventricular myocytes. J. Physiol. 507: 379–403, 1998.
 369. Zorzato, F., G. Salviati, T. Facchinetti, and P. Volpe. Doxorubicin induces calcium release from terminal cisternae of skeletal muscle. J. Biol. Chem. 260: 7349–7355, 1985.
 370. Zorzato, F., J. Fujii, K. Otsu, M. Phillips, N. M. Green, F. A. Lai, G. Meissner, and D. H. Maclennan. Molecular cloning of cDNA encoding human and rabbit forms of the Ca2+ release channel (ryanodine receptor) of skeletal muscle sarcoplasmic reticulum. J. Biol. Chem. 265: 2244–2256, 1990.
 371. Zühlke, R. D., G. S. Pitt, K. Deisseroth, R. W. Tsien, and H. Reuter: Calmodulin supports both inactivation and facilitation of L‐type calcium channels. Nature 399: 159–162, 1999.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Donald M. Bers. Regulation of Cellular Calcium in Cardiac Myocytes. Compr Physiol 2011, Supplement 6: Handbook of Physiology, The Cardiovascular System, The Heart: 335-387. First published in print 2002. doi: 10.1002/cphy.cp020109