Comprehensive Physiology Wiley Online Library

Electrical Heterogeneity in the Heart: Physiological, Pharmacological and Clinical Implications

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Action Potential and Ionic Distinctions
1.1 Methodological Considerations in the Assessment of Electrical Heterogeneity
2 Pharmacological Distinctions
2.1 Epicardium versus Endocardium
2.2 M‐cells versus Epicardium and Endocardium
2.3 M‐Cells versus Purkinje Cells
3 Molecular Distinctions
3.1 Potassium Channels
3.2 Sodium Channels
3.3 Gap Junctions
3.4 Chloride Conductances
3.5 Calcium Channels
3.6 Pumps and Exchangers
4 Simulation of Action Potential Heterogeneity
5 Developmental Aspects
6 Physiological and Clinical Implications
6.1 Transmural Distribution of Ito and the J Wave
6.2 Phase 2 Re‐entry as a Mechanism of Extrasystolic Activity
6.3 Phase 2 Re‐entry as a Trigger for VT/VF: The Brugada Syndrome
6.4 Early Repolarization Syndrome
6.5 Ischemia
6.6 Role of Transmural Heterogeneity in Inscription of the Electrocardiographic T Wave
6.7 Role of Transmural Heterogeneity in Inscription of the U Wave
6.8 Role of Transmural Heterogeneity in the Long QT Syndrome
6.9 Torsade de Pointes
6.10 Pharmacological Therapy for LQTS: Reducing Transmural Dispersion of Repolarization
7 Summary
Figure 1. Figure 1.

A: Action potentials recorded from myocytes isolated from the epicardial (Epi), endocardial (Endo), and M regions of the canine left ventricle. B:I‐V relations for IK1 in epicardial, endocardial, and M region myocytes. Values are mean ± S.D. C: Transient outward current (Ito) recorded from the three cell types (current traces recorded during depolarizing steps from a holding potential of −80 mV to test potentials ranging between −20 and +70 mV D: The average peak current–voltage relationship for Ito for each of the three cell types. Values are mean ± S.D. E: Voltage‐dependent activation of the slowly activating component of the delayed rectifier K+ current (IKs) (currents were elicited by the voltage pulse protocol shown in the inset; Na+‐, K+‐ and Ca2+‐free solution). F: Voltage dependence of IKs (current remaining after exposure to E‐4031) and IKr (E‐4031‐sensitive current). Values are mean ± S.E. * p <0.05 compared with Epi or Endo. G: Reverse‐mode sodium–calcium exchange currents recorded in potassium‐ and chloride‐free solutions at a voltage of −80 mV. INa‐Ca was maximally activated by switching to sodium‐free external solution at the time indicated by the arrow. H: Midmyocardial sodium–calcium exchanger density is 30% greater than endocardial density, calculated as the peak outward INa‐Ca normalized by cell capacitance. Endocardial and epicardial densities were not significantly different. I: TTX‐sensitive late sodium current. Cells were held at −80 mV and briefly pulsed to −45 mV to inactivate fast sodium current before stepping to −10 mV. J: Normalized late sodium current measured 300 msec into the test pulse was plotted as a function of test pulse potential.

From references 133,132,257, with permission. Modified from reference 257 with permission
Figure 2. Figure 2.

Transmembrane activity recorded from cells isolated from the epicardial, M, and endocardial regions of the canine left ventricle at basic cycle lengths (BCL) of 300 to 5000 msec (steady‐state conditions). The M‐cells and transitional cells were enzymatically dissociated from the midmyocardial region. Deceleration‐induced prolongation of APD in M‐cells is much greater than in epicardial and endocardial cells. The spike‐and‐dome morphology is also more accentuated in the epicardial cell.

Figure 3. Figure 3.

Transmural distribution of action potential duration and tissue resistivity in the intact ventricular wall. A: Schematic diagram of the arterially perfused canine LV wedge preparation. The wedge is perfused with Tyrode's solution via a small native branch of the left descending coronary artery and stimulated from the endocardial surface. Transmembrane action potentials are recorded simultaneously from epicardial (Epi), M region (M), and endocardial (Endo) sites using three floating microelectrodes. A transmural ECG is recorded along the same transmural axis across the bath, registering the entire field of the wedge. B: Histology of a transmural slice of the left ventricular wall near the epicardial border. The region of sharp transition of cell orientation coincides with the region of high tissue resistivity depicted in D and the region of sharp transition of action potential duration illustrated in C. C: Distribution of conduction time (CT), APD90, and repolarization time (RT = APD90 + CT) in a canine left ventricular wall wedge preparation paced at BCL of 2000 msec. A sharp transition of APD90 is present between epicardium and subepicardium. Epi: epicardium; M: M‐cell; Endo: endocardium; RT: repolarization time; CT: conduction time. D: Distribution of total tissue resistivity (Rt) across the canine left ventricular wall. Transmural distances at 0% and 100% represent epicardium and endocardium, respectively. * p <0.01 compared with Rt at midwall. Tissue resistivity increases most dramatically between deep subepicardium and epicardium. Error bars represent SEM (n = 5).

From Yan et al. 246, with permission
Figure 4. Figure 4.

Effect of d‐sotalol, a specific IKr blocker, on transmembrane activity recorded from epicardial (Epi), endocardial (Endo), and deep subepicardial (M‐cell) sites in a transmural strip of canine left ventricle. A: Each panel shows superimposed action potentials recorded at basic cycle lengths (BCL) of 300–5000 msec, before and after d‐sotalol (100 μM). B: APD‐rate relations.

From Sicouri et al. 198, with permission
Figure 5. Figure 5.

Effects of chronic amiodarone on the rate dependence of action potential characteristics in epicardial (Epi), M, and endocardial (Endo) tissues isolated from the hearts of untreated dogs (left) as well as those receiving chronic amiodarone therapy (right). A: Transmembrane activity recorded simultaneously from Epi, M, and Endo preparations at basic cycle lengths (BCL) of 500, 800, 2000, and 5000 msec (steady‐state conditions). B: Composite data from twelve untreated dogs and five amiodarone treated dogs. The graphs plot APD‐rate relations for Epi (open circles), Endo (closed circles), and M (open triangles) of untreated (left) and amiodarone‐treated animals (right). Each point represents mean ± S.D. * p<0.01 amiodarone vs. control. [K+]o = 4 mM. Chronic amiodarone treatment leads to much more uniform APD‐rate relations in the three cell types.

Figure 6. Figure 6.

Heterogeneous distribution of KCNE1 and KCNQ1 in canine ventricular myocardium. A: Transmural distribution of KCNE1 (filled columns) and KCNQ1 (hatched columns) mRNA as measured by competitive multiplex PCR. RNA was reverse transcribed using standard protocols and 2 μg of DNA from the epi‐, mid‐, and endocardium was amplified against known concentrations of an exogenous external standard (MIMIC). Data are from a pool of acutely dissociated cells from eight dogs. The concentration at which the amplitude of the two signals was equal was used to determine the amount of dsDNA presented. B: Same protocol as in A, with RNA isolated from the right (RV) and left (LV) ventricles. C: Western blot analysis of the distribution of KCNQ1 in canine right and left ventricles using polyclonal antibodies. Two bands are typically observed with a more intense signal in RV. 70 μg of protein was loaded in each well. D: Same experiment as in C with a KCNE1 antibody; the protein is more strongly expressed in LV.

Figure 7. Figure 7.

Age‐related spike‐and‐dome morphology and changes in Ito in canine ventricular epicardium. Each panel depicts transmembrane activity recorded from right ventricular epicardial tissues (upper trace) and transient outward current (lower trace) recorded from left ventricular epicardial cells isolated from a neonate (5 days of age; A), a young dog (3 months old; B), and an adult dog (C). BCL = 2000 msec; [K+]o = 4 mM. The spike‐and‐dome configuration of the epicardial action potential and Ito density are absent in the neonate, relatively small in the young dog, and most prominent in the adult.

Figure 8. Figure 8.

Phase 2 re‐entry. Re‐entrant activity induced by exposure of a canine ventricular epicardial preparation (0.7 cm2) to simulated ischemia. Microelectrode recordings were obtained from four sites as shown in the schematic (upper right). After 35 min of ischemia, the action potential dome develops normally at site 4, but not at sites 1, 2, or 3. The dome then propagates in a clockwise direction re‐exciting sites 3, 2, and 1 with progressive delays, thus generating a closely coupled re‐entrant extrasystole (156 msec) at site 1. In this example of phase 2 re‐entry, propagation of the dome occurs in a direction opposite that of phase 0, a mechanism akin to reflection. BCL = 700 msec.

Modified from Lukas and Antzelevitz 138, with permission
Figure 9. Figure 9.

Phase 2 re‐entrant extrasystole triggers circus movement re‐entry. A: Exposure of a relatively large canine right ventricular epicardial sheet (6.3 cm2) to simulated ischemia results in loss of the dome at sites 3 and 4 but not at sites 1 and 2 (BCL = 1100 msec). Conduction of the basic beat proceeds normally from the stimulation site (site 2; see schematic a). Propagation of the action potential dome from the right half of the preparation caused reexcitation of the left half via a phase 2 re‐entry mechanism (see schematic b). The extrasystolic beat generated by phase 2 re‐entry then initiates a run of tachycardia that is sustained for 4 additional cycles via typical circus movement re‐entry. The proposed re‐entrant path is shown in schematic c. Note that phase 2 re‐entry provides an activation front roughly perpendicular to that of the basic beat. This type of crossfield activation has previously been shown to predispose to the development of vortex‐like re‐entry in isolated epicardial sheets. B: Recorded after addition of 1 mM 4‐aminopyridine (4‐AP), an inhibitor of the transient outward current. In the continued presence of ischemia, 4‐AP restored the dome at all epicardial recording sites within 3 min. Thus electrical heterogeneity was restored and all re‐entrant activity abolished.

Modified from Lukas and Antzelevitz 138, with permission
Figure 10. Figure 10.

ECG and arrhythmias with typical features of the Brugada syndrome recorded from canine right ventricular wedge preparations A: Schematic of arterially perfused right ventricular wedge preparation. B: Pressure—induced phase 2 re‐entry and VT. Shown are transmembrane action potentials simultaneously recorded from two epicardial sites (Epi 1 and Epi 2) and one M region (M) site, together with a transmural ECG. Local application of pressure near Epi 2 results in loss of the action potential dome at that site but not at the Epi 1 or M sites. The dome at Epi 1 then re‐excites Epi 2, giving rise to a phase 2 re‐entrant extrasystole that triggers a short run of ventricular tachycardia. Note the ST segment elevation due to loss of the action potential dome in a segment of epicardium. C: Polymorphic VT/VF induced by local application of the potassium channel opener pinacidil (10 μM) to the epicardial surface of the wedge. Action potentials from two epicardial sites (Epi 1 and Epi 2) and a transmural ECG were simultaneously recorded. Loss of the dome at Epi 1 but not Epi 2 creates a marked dispersion of repolarization, giving rise to a phase 2 re‐entrant extrasystole. The extrasystolic beat then triggers a long episode of ventricular fibrillation (22 sec). Right panel: Addition of 4‐aminopyridine (4‐AP. 2 mM), a specific Ito blocker, to the perfusate restored the action potential dome at Epi 1, thus reducing dispersion of repolarization and suppressing all arrhythmic activity. BCL = 2000 msec. D: Phase 2 re‐entry gives rise to VT following addition of pinacidil (2.5 μM) to the coronary perfusate. Transmembrane action potentials form two epicardial sites (Epi 1 and Epi 2) and one endocardial site (Endo), as well as a transmural ECG were simultaneously recorded. Right panel: 4‐AP (1 mM) markedly reduces the magnitude of the action potential notch in epicardium, thus restoring the action potential dome throughout the preparation and abolishing all arrhythmic activity.

D is from Yan and Antzelevitch 245, with permission
Figure 11. Figure 11.

Proposed mechanism for the Brugada syndrome. A shift in the balance of currents serves to amplify existing heterogeneities by causing loss of the action potential dome at some epicardial sites, but not endocardial sites. A vulnerable window develops as a result of the dispersion of repolarization and refractoriness within epicardium as well as across the wall. Epicardial dispersion leads to the development of phase 2 re‐entry, which provides the extrasystole that captures the vulnerable window and initiates VT/VF via a circus movement re‐entry mechanism.

Figure 12. Figure 12.

Whole‐cell current for wild‐type (WT) and Brugada syndrome mutant (T1620M) in transiently transfected TSA201 cells at room temperature (22°C) and 32°C. A: The cartoon depicts the location of the missense mutations R1232W and T1620M previously described by Chen et al. 52. Current recordings obtained at different test potentials from −70 to −25 mV (32°C) and −65 to −20 (22°C) in increments of 5 mV from a holding potential of −120 mV for four representative cells. B: Current decay of T1620M at 32°C. Representative current recordings from WT and T1620M were elicited by a 20 msec depolarizing pulse to a test potential of 10 mV from a holding potential of −140 mV, normalized to the peak inward current and superimposed. C: WT decay time constant (square) is less sensitive to temperature in the physiological range. Cells were maintained at −80 mV and pulsed at 0 mV for 10 msec at temperatures between 22°C and 42°C. Current decay were fitted by a sum of two exponential functions. The fast time constant was plotted against the temperature (log scale) for WT and T1620M (filled circles). Each symbol represents a different cell.

Modified from Dumaine et al. 75, with permission
Figure 13. Figure 13.

Voltage gradients on either side of the M region are responsible for inscription of the electrocardiographic T wave. Top: Action potentials simultaneously recorded from endocardial, epicardial and M region sites of an arterially perfused canine left ventricular wedge preparation. Middle: ECG recorded across the wedge. Bottom: Computed voltage differences between the epicardium and M region action potentials (ΔVM‐Epi) and between the M region and endocardium responses (ΔVEndo‐M). If these traces are representative of the opposing voltage gradients on either side of the M region, responsible for inscription of the T wave, then the weighted sum of the two traces should yield a trace (middle trace in bottom grouping) resembling the ECG, which it does. The voltage gradients are weighted to account for differences in tissue resistivity between M and Epi and Endo and M regions, thus yielding the opposing currents flowing on either side of the M region. A: Under control conditions, the T wave begins when the plateau of epicardial action potential separates from that of the M‐cell. As epicardium repolarizes, the voltage gradient between epicardium and the M region continues to grow, giving rise to the ascending limb of the T wave. The voltage gradient between the M region and the epicardium (ΔVM‐Epi) reaches a peak when the epicardium is fully repolarized; this marks the peak of the T wave. On the other end of the ventricular wall, the endocardial plateau deviates from that of the M‐cell, generating an opposing voltage gradient (ΔVEndo‐M) and corresponding current that limits the amplitude of the T wave and contributes to the initial part of the descending limb of the T wave. The voltage gradient between the endocardium and the M region reaches a peak when the endocardium is fully repolarized. The gradient continues to decline as the M‐cells repolarize. All gradients are extinguished when the longest M cells are fully repolarized. B: d‐sotalol (100 μM) prolongs the action potential of the M‐cell more than those of the epicardial and endocardial cells, thus widening the T wave and prolonging the QT interval. The greater separation of epicardial and endocardial repolarization times also gives rise to a notch in the descending limb of the T wave. Once again, the T wave begins when the plateau of epicardial action potential diverges from that of the M‐cell. The same relationships as described for panel A are observed during the remainder of the T wave. The d‐sotalol–induced increase in dispersion of repolarization across the wall is accompanied by a corresponding increase in the Tpeak‐Tend interval in the pseudo‐ECG.

Modified from Yan and Antzelevitch 244, with permission
Figure 14. Figure 14.

Transient shift of voltage gradients on either side of the M region results in T wave bifurcation. The format is the same as in Figure 13. All traces were simultaneously recorded form an arterially perfused left ventricular wedge preparation. A: Control. B: In the presence of hypokalemia ([K+]o = 1.5 mM), the IKr blocker dl‐sotalol (100 μM) prolongs the QT interval and produces a bifurcation of the T wave, a morphology some authors refer to as T‐U complex. The rate of repolarization of phase 3 of the action potential is slowed, giving rise to smaller opposing transmural currents that cross over, producing a low‐amplitude bifid T wave. Initially the voltage gradient between the epicardium and M regions (M‐Epi) is greater than that between endocardium and M region (Endo‐M). When endocardium pulls away from the M cell, the opposing gradient (Endo‐M) increases, interrupting the ascending limb of the T wave. Predominance of the M‐Epi gradient is restored as the epicardial response continues to repolarize and the Epi‐M gradients increases, thus resuming the ascending limb of the T wave. Full repolarization of epicardium marks the peak of the T wave. Repolarization of both endocardium and the M region contribute importantly to the descending limb. BCL = 1000 msec.

Modified from Yan and Antzelevitch 244, with permission
Figure 15. Figure 15.

Contribution of transmural vs. apicobasal and anterior–posterior gradients to the registration of the T wave. The four ECG traces were simultaneously recorded at 0°, 45°, −45°, and 90° (apicobasal) angles relative to the transmural axis of an arterially perfused left ventricular wedge preparation. Inscription of the T wave is largely the result of voltage gradients along the transmural axis.

Modified from Yan and Antzelevitch 244 with permission
Figure 16. Figure 16.

Correlation of transmembrane and electrocardiographic activity. Transmembrane potentials and a transmural ECG recorded from two different arterially perfused canine left ventricular wedge preparations. A: Action potentials from epicardial (Epi), midmyocardial (M), and endocardial (Endo) sites were simultaneously recorded using floating glass microelectrodes. A transmural ECG was recorded concurrently across the bath. B: Action potentials from epicardium (Epi), midmyocardium (M), and subendocardial Purkinje were recorded simultaneously together with a transmural ECG. In both cases, repolarization of epicardium is coincident with the peak of the T wave of the ECG, whereas repolarization of the M‐cells is coincident with the end of the T wave. The endocardial APD is intermediate (A). Although repolarization of the Purkinje fiber occurs after that of the M‐cell (B), it does not register on the ECG. BCL = 2000 msec.

Modified from yan and Antzelevitch 244, with permission
Figure 17. Figure 17.

Transmembrane action potentials and transmural electrocardiograms (ECG) in control and LQT1 (A and B), LQT2 (C and D), and LQT3 (E and F) models of LQTS (arterially perfused canine left ventricular wedge preparations), and clinical ECG (lead V5) of patients with LQT1 (KvLQT1 defect) (G), LQT2 (HERG defect) (H), and LQT3 (SCN5A defect) (I) syndromes. Isoproterenol + chromanol 293B—an IKs blocker, d‐sotalol + low [K+]o, and ATX‐II—an agent that slows inactivation of late INa are used to mimic the LQT1, LQT2 and LQT3 syndromes, respectively. A–F depict action potentials simultaneously recorded from endocardial (Endo), M, and epicardial (Epi) sites, together with a transmural ECG. BCL = 2000 msec. In all cases, the peak of the T wave in the ECG is coincident with the repolarization of the epicardial action potential, whereas the end of the T wave is coincident with the repolarization of the M‐cell action potential. Repolarization of the endocardial cell is intermediate between that of the M‐cell and epicardial cell. Transmural dispersion of repolarization across the ventricular wall, defined as the difference in the repolarization time between M‐cells and epicardial cells, is denoted below the ECG traces. B: Isoproterenol (100 nM) in the presence of chromanol 293B (30 μM) produced a preferential prolongation of the APD of the M, resulting in an accentuated transmural dispersion of repolarization and broad‐based T waves as commonly seen in LQT1 patients (G). D: d‐Sotalol (100 μM) in the presence of low potassium (2 mM) gives rise to low‐amplitude T waves with a notched or bifurcated appearance due to a very significant slowing of repolarization as commonly seen in LQT2 patients (H). F: ATX‐II (20 nM) markedly prolongs the QT interval, widens the T wave, and causes a sharp rise in the dispersion of repolarization. ATX‐II also produced a marked delay in onset of the T wave due to relatively large effects of the drug on the APD of epicardium and endocardium, consistent with the late‐appearing T wave pattern observed in LQT3 patients (I).

Modified from Shimizu and Antzelevitch 182,183, with permission
Figure 18. Figure 18.

Spontaneous and stimulation‐induced polymorphic ventricular tachycardia with features of torsade de pointes (TdP). A: Stimulation‐induced TdP in a LV wedge preparation pretreated with dl‐sotalol (100 μmol/liter). S1‐S1 = 2000 msec; S1‐S2 = 250 msec. S2 was applied to epicardium. B: Spontaneous TdP in a preparation pretreated with dl‐sotalol (100 μmol/liter). BCL = 2000 msec. A spontaneous premature beat with a coupling interval of 348 msec, likely originating from subendocardial Purkinje system, initiates an episode of torsade de pointes.

Figure 19. Figure 19.

Proposed cellular mechanism for the development of torsade de pointes in the LQT1, 2, and 3 forms of the long QT syndrome.

Figure 20. Figure 20.

Similarities and differences in mechanisms responsible for the development of arrhythmias in the Brugada and long QT syndromes. Amplification of intrinsic heterogeneities underlies arrhythmogenicity in both syndromes. In the case of the Brugada syndrome, an increase in net outward current amplifies the heterogeneity normally present in the early phases of the action potential, leading to accentuation of the epicardial notch and finally loss of the action potential dome, resulting in marked abbreviation of the potential at some epicardial sites. In the case of the long QT syndrome, a decrease in net outward current amplifies the heterogeneity normally present in the late phases of the action potential, by producing a preferential prolongation of the M‐cell action potential.



Figure 1.

A: Action potentials recorded from myocytes isolated from the epicardial (Epi), endocardial (Endo), and M regions of the canine left ventricle. B:I‐V relations for IK1 in epicardial, endocardial, and M region myocytes. Values are mean ± S.D. C: Transient outward current (Ito) recorded from the three cell types (current traces recorded during depolarizing steps from a holding potential of −80 mV to test potentials ranging between −20 and +70 mV D: The average peak current–voltage relationship for Ito for each of the three cell types. Values are mean ± S.D. E: Voltage‐dependent activation of the slowly activating component of the delayed rectifier K+ current (IKs) (currents were elicited by the voltage pulse protocol shown in the inset; Na+‐, K+‐ and Ca2+‐free solution). F: Voltage dependence of IKs (current remaining after exposure to E‐4031) and IKr (E‐4031‐sensitive current). Values are mean ± S.E. * p <0.05 compared with Epi or Endo. G: Reverse‐mode sodium–calcium exchange currents recorded in potassium‐ and chloride‐free solutions at a voltage of −80 mV. INa‐Ca was maximally activated by switching to sodium‐free external solution at the time indicated by the arrow. H: Midmyocardial sodium–calcium exchanger density is 30% greater than endocardial density, calculated as the peak outward INa‐Ca normalized by cell capacitance. Endocardial and epicardial densities were not significantly different. I: TTX‐sensitive late sodium current. Cells were held at −80 mV and briefly pulsed to −45 mV to inactivate fast sodium current before stepping to −10 mV. J: Normalized late sodium current measured 300 msec into the test pulse was plotted as a function of test pulse potential.

From references 133,132,257, with permission. Modified from reference 257 with permission


Figure 2.

Transmembrane activity recorded from cells isolated from the epicardial, M, and endocardial regions of the canine left ventricle at basic cycle lengths (BCL) of 300 to 5000 msec (steady‐state conditions). The M‐cells and transitional cells were enzymatically dissociated from the midmyocardial region. Deceleration‐induced prolongation of APD in M‐cells is much greater than in epicardial and endocardial cells. The spike‐and‐dome morphology is also more accentuated in the epicardial cell.



Figure 3.

Transmural distribution of action potential duration and tissue resistivity in the intact ventricular wall. A: Schematic diagram of the arterially perfused canine LV wedge preparation. The wedge is perfused with Tyrode's solution via a small native branch of the left descending coronary artery and stimulated from the endocardial surface. Transmembrane action potentials are recorded simultaneously from epicardial (Epi), M region (M), and endocardial (Endo) sites using three floating microelectrodes. A transmural ECG is recorded along the same transmural axis across the bath, registering the entire field of the wedge. B: Histology of a transmural slice of the left ventricular wall near the epicardial border. The region of sharp transition of cell orientation coincides with the region of high tissue resistivity depicted in D and the region of sharp transition of action potential duration illustrated in C. C: Distribution of conduction time (CT), APD90, and repolarization time (RT = APD90 + CT) in a canine left ventricular wall wedge preparation paced at BCL of 2000 msec. A sharp transition of APD90 is present between epicardium and subepicardium. Epi: epicardium; M: M‐cell; Endo: endocardium; RT: repolarization time; CT: conduction time. D: Distribution of total tissue resistivity (Rt) across the canine left ventricular wall. Transmural distances at 0% and 100% represent epicardium and endocardium, respectively. * p <0.01 compared with Rt at midwall. Tissue resistivity increases most dramatically between deep subepicardium and epicardium. Error bars represent SEM (n = 5).

From Yan et al. 246, with permission


Figure 4.

Effect of d‐sotalol, a specific IKr blocker, on transmembrane activity recorded from epicardial (Epi), endocardial (Endo), and deep subepicardial (M‐cell) sites in a transmural strip of canine left ventricle. A: Each panel shows superimposed action potentials recorded at basic cycle lengths (BCL) of 300–5000 msec, before and after d‐sotalol (100 μM). B: APD‐rate relations.

From Sicouri et al. 198, with permission


Figure 5.

Effects of chronic amiodarone on the rate dependence of action potential characteristics in epicardial (Epi), M, and endocardial (Endo) tissues isolated from the hearts of untreated dogs (left) as well as those receiving chronic amiodarone therapy (right). A: Transmembrane activity recorded simultaneously from Epi, M, and Endo preparations at basic cycle lengths (BCL) of 500, 800, 2000, and 5000 msec (steady‐state conditions). B: Composite data from twelve untreated dogs and five amiodarone treated dogs. The graphs plot APD‐rate relations for Epi (open circles), Endo (closed circles), and M (open triangles) of untreated (left) and amiodarone‐treated animals (right). Each point represents mean ± S.D. * p<0.01 amiodarone vs. control. [K+]o = 4 mM. Chronic amiodarone treatment leads to much more uniform APD‐rate relations in the three cell types.



Figure 6.

Heterogeneous distribution of KCNE1 and KCNQ1 in canine ventricular myocardium. A: Transmural distribution of KCNE1 (filled columns) and KCNQ1 (hatched columns) mRNA as measured by competitive multiplex PCR. RNA was reverse transcribed using standard protocols and 2 μg of DNA from the epi‐, mid‐, and endocardium was amplified against known concentrations of an exogenous external standard (MIMIC). Data are from a pool of acutely dissociated cells from eight dogs. The concentration at which the amplitude of the two signals was equal was used to determine the amount of dsDNA presented. B: Same protocol as in A, with RNA isolated from the right (RV) and left (LV) ventricles. C: Western blot analysis of the distribution of KCNQ1 in canine right and left ventricles using polyclonal antibodies. Two bands are typically observed with a more intense signal in RV. 70 μg of protein was loaded in each well. D: Same experiment as in C with a KCNE1 antibody; the protein is more strongly expressed in LV.



Figure 7.

Age‐related spike‐and‐dome morphology and changes in Ito in canine ventricular epicardium. Each panel depicts transmembrane activity recorded from right ventricular epicardial tissues (upper trace) and transient outward current (lower trace) recorded from left ventricular epicardial cells isolated from a neonate (5 days of age; A), a young dog (3 months old; B), and an adult dog (C). BCL = 2000 msec; [K+]o = 4 mM. The spike‐and‐dome configuration of the epicardial action potential and Ito density are absent in the neonate, relatively small in the young dog, and most prominent in the adult.



Figure 8.

Phase 2 re‐entry. Re‐entrant activity induced by exposure of a canine ventricular epicardial preparation (0.7 cm2) to simulated ischemia. Microelectrode recordings were obtained from four sites as shown in the schematic (upper right). After 35 min of ischemia, the action potential dome develops normally at site 4, but not at sites 1, 2, or 3. The dome then propagates in a clockwise direction re‐exciting sites 3, 2, and 1 with progressive delays, thus generating a closely coupled re‐entrant extrasystole (156 msec) at site 1. In this example of phase 2 re‐entry, propagation of the dome occurs in a direction opposite that of phase 0, a mechanism akin to reflection. BCL = 700 msec.

Modified from Lukas and Antzelevitz 138, with permission


Figure 9.

Phase 2 re‐entrant extrasystole triggers circus movement re‐entry. A: Exposure of a relatively large canine right ventricular epicardial sheet (6.3 cm2) to simulated ischemia results in loss of the dome at sites 3 and 4 but not at sites 1 and 2 (BCL = 1100 msec). Conduction of the basic beat proceeds normally from the stimulation site (site 2; see schematic a). Propagation of the action potential dome from the right half of the preparation caused reexcitation of the left half via a phase 2 re‐entry mechanism (see schematic b). The extrasystolic beat generated by phase 2 re‐entry then initiates a run of tachycardia that is sustained for 4 additional cycles via typical circus movement re‐entry. The proposed re‐entrant path is shown in schematic c. Note that phase 2 re‐entry provides an activation front roughly perpendicular to that of the basic beat. This type of crossfield activation has previously been shown to predispose to the development of vortex‐like re‐entry in isolated epicardial sheets. B: Recorded after addition of 1 mM 4‐aminopyridine (4‐AP), an inhibitor of the transient outward current. In the continued presence of ischemia, 4‐AP restored the dome at all epicardial recording sites within 3 min. Thus electrical heterogeneity was restored and all re‐entrant activity abolished.

Modified from Lukas and Antzelevitz 138, with permission


Figure 10.

ECG and arrhythmias with typical features of the Brugada syndrome recorded from canine right ventricular wedge preparations A: Schematic of arterially perfused right ventricular wedge preparation. B: Pressure—induced phase 2 re‐entry and VT. Shown are transmembrane action potentials simultaneously recorded from two epicardial sites (Epi 1 and Epi 2) and one M region (M) site, together with a transmural ECG. Local application of pressure near Epi 2 results in loss of the action potential dome at that site but not at the Epi 1 or M sites. The dome at Epi 1 then re‐excites Epi 2, giving rise to a phase 2 re‐entrant extrasystole that triggers a short run of ventricular tachycardia. Note the ST segment elevation due to loss of the action potential dome in a segment of epicardium. C: Polymorphic VT/VF induced by local application of the potassium channel opener pinacidil (10 μM) to the epicardial surface of the wedge. Action potentials from two epicardial sites (Epi 1 and Epi 2) and a transmural ECG were simultaneously recorded. Loss of the dome at Epi 1 but not Epi 2 creates a marked dispersion of repolarization, giving rise to a phase 2 re‐entrant extrasystole. The extrasystolic beat then triggers a long episode of ventricular fibrillation (22 sec). Right panel: Addition of 4‐aminopyridine (4‐AP. 2 mM), a specific Ito blocker, to the perfusate restored the action potential dome at Epi 1, thus reducing dispersion of repolarization and suppressing all arrhythmic activity. BCL = 2000 msec. D: Phase 2 re‐entry gives rise to VT following addition of pinacidil (2.5 μM) to the coronary perfusate. Transmembrane action potentials form two epicardial sites (Epi 1 and Epi 2) and one endocardial site (Endo), as well as a transmural ECG were simultaneously recorded. Right panel: 4‐AP (1 mM) markedly reduces the magnitude of the action potential notch in epicardium, thus restoring the action potential dome throughout the preparation and abolishing all arrhythmic activity.

D is from Yan and Antzelevitch 245, with permission


Figure 11.

Proposed mechanism for the Brugada syndrome. A shift in the balance of currents serves to amplify existing heterogeneities by causing loss of the action potential dome at some epicardial sites, but not endocardial sites. A vulnerable window develops as a result of the dispersion of repolarization and refractoriness within epicardium as well as across the wall. Epicardial dispersion leads to the development of phase 2 re‐entry, which provides the extrasystole that captures the vulnerable window and initiates VT/VF via a circus movement re‐entry mechanism.



Figure 12.

Whole‐cell current for wild‐type (WT) and Brugada syndrome mutant (T1620M) in transiently transfected TSA201 cells at room temperature (22°C) and 32°C. A: The cartoon depicts the location of the missense mutations R1232W and T1620M previously described by Chen et al. 52. Current recordings obtained at different test potentials from −70 to −25 mV (32°C) and −65 to −20 (22°C) in increments of 5 mV from a holding potential of −120 mV for four representative cells. B: Current decay of T1620M at 32°C. Representative current recordings from WT and T1620M were elicited by a 20 msec depolarizing pulse to a test potential of 10 mV from a holding potential of −140 mV, normalized to the peak inward current and superimposed. C: WT decay time constant (square) is less sensitive to temperature in the physiological range. Cells were maintained at −80 mV and pulsed at 0 mV for 10 msec at temperatures between 22°C and 42°C. Current decay were fitted by a sum of two exponential functions. The fast time constant was plotted against the temperature (log scale) for WT and T1620M (filled circles). Each symbol represents a different cell.

Modified from Dumaine et al. 75, with permission


Figure 13.

Voltage gradients on either side of the M region are responsible for inscription of the electrocardiographic T wave. Top: Action potentials simultaneously recorded from endocardial, epicardial and M region sites of an arterially perfused canine left ventricular wedge preparation. Middle: ECG recorded across the wedge. Bottom: Computed voltage differences between the epicardium and M region action potentials (ΔVM‐Epi) and between the M region and endocardium responses (ΔVEndo‐M). If these traces are representative of the opposing voltage gradients on either side of the M region, responsible for inscription of the T wave, then the weighted sum of the two traces should yield a trace (middle trace in bottom grouping) resembling the ECG, which it does. The voltage gradients are weighted to account for differences in tissue resistivity between M and Epi and Endo and M regions, thus yielding the opposing currents flowing on either side of the M region. A: Under control conditions, the T wave begins when the plateau of epicardial action potential separates from that of the M‐cell. As epicardium repolarizes, the voltage gradient between epicardium and the M region continues to grow, giving rise to the ascending limb of the T wave. The voltage gradient between the M region and the epicardium (ΔVM‐Epi) reaches a peak when the epicardium is fully repolarized; this marks the peak of the T wave. On the other end of the ventricular wall, the endocardial plateau deviates from that of the M‐cell, generating an opposing voltage gradient (ΔVEndo‐M) and corresponding current that limits the amplitude of the T wave and contributes to the initial part of the descending limb of the T wave. The voltage gradient between the endocardium and the M region reaches a peak when the endocardium is fully repolarized. The gradient continues to decline as the M‐cells repolarize. All gradients are extinguished when the longest M cells are fully repolarized. B: d‐sotalol (100 μM) prolongs the action potential of the M‐cell more than those of the epicardial and endocardial cells, thus widening the T wave and prolonging the QT interval. The greater separation of epicardial and endocardial repolarization times also gives rise to a notch in the descending limb of the T wave. Once again, the T wave begins when the plateau of epicardial action potential diverges from that of the M‐cell. The same relationships as described for panel A are observed during the remainder of the T wave. The d‐sotalol–induced increase in dispersion of repolarization across the wall is accompanied by a corresponding increase in the Tpeak‐Tend interval in the pseudo‐ECG.

Modified from Yan and Antzelevitch 244, with permission


Figure 14.

Transient shift of voltage gradients on either side of the M region results in T wave bifurcation. The format is the same as in Figure 13. All traces were simultaneously recorded form an arterially perfused left ventricular wedge preparation. A: Control. B: In the presence of hypokalemia ([K+]o = 1.5 mM), the IKr blocker dl‐sotalol (100 μM) prolongs the QT interval and produces a bifurcation of the T wave, a morphology some authors refer to as T‐U complex. The rate of repolarization of phase 3 of the action potential is slowed, giving rise to smaller opposing transmural currents that cross over, producing a low‐amplitude bifid T wave. Initially the voltage gradient between the epicardium and M regions (M‐Epi) is greater than that between endocardium and M region (Endo‐M). When endocardium pulls away from the M cell, the opposing gradient (Endo‐M) increases, interrupting the ascending limb of the T wave. Predominance of the M‐Epi gradient is restored as the epicardial response continues to repolarize and the Epi‐M gradients increases, thus resuming the ascending limb of the T wave. Full repolarization of epicardium marks the peak of the T wave. Repolarization of both endocardium and the M region contribute importantly to the descending limb. BCL = 1000 msec.

Modified from Yan and Antzelevitch 244, with permission


Figure 15.

Contribution of transmural vs. apicobasal and anterior–posterior gradients to the registration of the T wave. The four ECG traces were simultaneously recorded at 0°, 45°, −45°, and 90° (apicobasal) angles relative to the transmural axis of an arterially perfused left ventricular wedge preparation. Inscription of the T wave is largely the result of voltage gradients along the transmural axis.

Modified from Yan and Antzelevitch 244 with permission


Figure 16.

Correlation of transmembrane and electrocardiographic activity. Transmembrane potentials and a transmural ECG recorded from two different arterially perfused canine left ventricular wedge preparations. A: Action potentials from epicardial (Epi), midmyocardial (M), and endocardial (Endo) sites were simultaneously recorded using floating glass microelectrodes. A transmural ECG was recorded concurrently across the bath. B: Action potentials from epicardium (Epi), midmyocardium (M), and subendocardial Purkinje were recorded simultaneously together with a transmural ECG. In both cases, repolarization of epicardium is coincident with the peak of the T wave of the ECG, whereas repolarization of the M‐cells is coincident with the end of the T wave. The endocardial APD is intermediate (A). Although repolarization of the Purkinje fiber occurs after that of the M‐cell (B), it does not register on the ECG. BCL = 2000 msec.

Modified from yan and Antzelevitch 244, with permission


Figure 17.

Transmembrane action potentials and transmural electrocardiograms (ECG) in control and LQT1 (A and B), LQT2 (C and D), and LQT3 (E and F) models of LQTS (arterially perfused canine left ventricular wedge preparations), and clinical ECG (lead V5) of patients with LQT1 (KvLQT1 defect) (G), LQT2 (HERG defect) (H), and LQT3 (SCN5A defect) (I) syndromes. Isoproterenol + chromanol 293B—an IKs blocker, d‐sotalol + low [K+]o, and ATX‐II—an agent that slows inactivation of late INa are used to mimic the LQT1, LQT2 and LQT3 syndromes, respectively. A–F depict action potentials simultaneously recorded from endocardial (Endo), M, and epicardial (Epi) sites, together with a transmural ECG. BCL = 2000 msec. In all cases, the peak of the T wave in the ECG is coincident with the repolarization of the epicardial action potential, whereas the end of the T wave is coincident with the repolarization of the M‐cell action potential. Repolarization of the endocardial cell is intermediate between that of the M‐cell and epicardial cell. Transmural dispersion of repolarization across the ventricular wall, defined as the difference in the repolarization time between M‐cells and epicardial cells, is denoted below the ECG traces. B: Isoproterenol (100 nM) in the presence of chromanol 293B (30 μM) produced a preferential prolongation of the APD of the M, resulting in an accentuated transmural dispersion of repolarization and broad‐based T waves as commonly seen in LQT1 patients (G). D: d‐Sotalol (100 μM) in the presence of low potassium (2 mM) gives rise to low‐amplitude T waves with a notched or bifurcated appearance due to a very significant slowing of repolarization as commonly seen in LQT2 patients (H). F: ATX‐II (20 nM) markedly prolongs the QT interval, widens the T wave, and causes a sharp rise in the dispersion of repolarization. ATX‐II also produced a marked delay in onset of the T wave due to relatively large effects of the drug on the APD of epicardium and endocardium, consistent with the late‐appearing T wave pattern observed in LQT3 patients (I).

Modified from Shimizu and Antzelevitch 182,183, with permission


Figure 18.

Spontaneous and stimulation‐induced polymorphic ventricular tachycardia with features of torsade de pointes (TdP). A: Stimulation‐induced TdP in a LV wedge preparation pretreated with dl‐sotalol (100 μmol/liter). S1‐S1 = 2000 msec; S1‐S2 = 250 msec. S2 was applied to epicardium. B: Spontaneous TdP in a preparation pretreated with dl‐sotalol (100 μmol/liter). BCL = 2000 msec. A spontaneous premature beat with a coupling interval of 348 msec, likely originating from subendocardial Purkinje system, initiates an episode of torsade de pointes.



Figure 19.

Proposed cellular mechanism for the development of torsade de pointes in the LQT1, 2, and 3 forms of the long QT syndrome.



Figure 20.

Similarities and differences in mechanisms responsible for the development of arrhythmias in the Brugada and long QT syndromes. Amplification of intrinsic heterogeneities underlies arrhythmogenicity in both syndromes. In the case of the Brugada syndrome, an increase in net outward current amplifies the heterogeneity normally present in the early phases of the action potential, leading to accentuation of the epicardial notch and finally loss of the action potential dome, resulting in marked abbreviation of the potential at some epicardial sites. In the case of the long QT syndrome, a decrease in net outward current amplifies the heterogeneity normally present in the late phases of the action potential, by producing a preferential prolongation of the M‐cell action potential.

References
 1. Abbott, G. W., F. Sesti, I. Splawski, M. E. Buck, M. H. Lehmann, K. W. Timothy, M. T. Keating, and S. A. N. Goldstein. MiRP1 forms IKr potassium channels with HERG and is associated with cardiac arrhythmia. Cell 97: 175–187, 1999.
 2. Abildskov, J. A. and R. L. Lux. The mechanism of simulated torsades de pointes in computer model of propagated excitation. J. Cardiovasc. Electrophysiol. 2: 224–237, 1991.
 3. Aizawa, Y., M. Tamura, M. Chinushi, N. Naitoh, H. Uchiyama, Y. Kusano, H. Hosono, and A. Shibata. Idiopathic ventricular fibrillation and bradycardia‐dependent intraventricular block. Am. Heart J. 126: 1473–1474, 1993.
 4. Aizawa, Y., M. Tamura, M. Chinushi, S. Niwano, Y. Kusano, N. Naitoh, A. Shibata, T. Tohjoh, Y. Ueda, and K. Joho. An attempt at electrical catheter ablation of the arrhythmogenic area in idiopathic ventricular fibrillation. Am. Heart J. 123: 257–260, 1992.
 5. Akar, F. G., G. X. Yan, C. Antzelevitch, and D. S. Rosenbaum. Optical maps reveal reentrant mechanism of torsade de pointes based on topography and electrophysiology of mid‐myocardial cells. Circulation 96: 1–355. 1997. (Abstract)
 6. Antzelevitch, C. Are M cells present in the ventricular myocardium of the pig? A question of maturity. Cardiovasc. Res. 36: 127–128, 1997.
 7. Antzelevitch, C. The M cell. Invited Editorial Comment. J. Cardiovasc. Pharmacol. Ther. 2: 73–76, 1997.
 8. Antzelevitch, C., J. M. Davidenko, S. Sicouri, L. Cohen, A. Iodice, R. J. Goodrow, and G. A. Gintant. Electrophysiologic effects of quinidine in canine Purkinje fibers and ventricular myocardium. Slow development of the antiarrhythmic and arrhythmogenic effects of the drug. In Recent Advances in Pharmacology and Therapeutics, edited by M. Velasco, A. Israel, E. Romero, and H. Silva New York: Excerpta Medica, 1989: 259–263.
 9. Antzelevitch, C., J. M. Davidenko, S. Sicouri, L. Cohen, A. Iodice, R. J. Goodrow, and G. A. Gintant. Quinidine‐induced early afterdepolarizations and triggered activity. J. Electrophysiol. 5: 323–338, 1989.
 10. Antzelevitch, C. and J. M. Di Diego. The role of K+ channel activators in cardiac electrophysiology and arrhythmias. Circulation 85: 1627–1629, 1992.
 11. Antzelevitch, C., J. M. Di Diego, S. Sicouri, and A. Lukas. Selective pharmacological modification of repolarizing currents. Antiarrhythmic and proarrhythmic actions of agents that influence repolarization in the heart. In: Antiarrhythmic Drugs: Mechanisms of Antiarrhythmic and Proarrhythmic Actions, edited by J. Breithardt. Berlin: Springer‐Verlag, 1995: 57–80.
 12. Antzelevitch, C., S. H. Litovsky, and A. Lukas. Epicardium vs. endocardium. Electrophysiology and pharmacology. In: Cardiac Electrophysiology, From Cell to Bedside, edited by D. P. Zipes, and J. Jalife. Philadelphia: W. B. Saunders, 1990: 386–395.
 13. Antzelevitch, C., V. V. Nesterenko, and G. X. Yan. The role of M cells in acquired long QT syndrome, U waves and torsade de pointes. J. Electrocardiol. 28 (suppl.): 131–138, 1996.
 14. Antzelevitch, C., W. Shimizu, G. X. Yan, S. Sicouri, J. Weissenburger, V. V. Nesterenko, A. Burashnikov, J. Di Diego, J. Saffitz, and G. P. Thomas. The M cell: its contribution to the ECG and to normal and abnormal electrical function of the heart. J. Cardiovasc Electrophysiol. 10: 1124–1152, 1999.
 15. Antzelevitch, C. and S. Sicouri. Clinical relevance of cardiac arrhythmias generated by afterdepolarizations: the role of M cells in the generation of U waves, triggered activity and torsade de pointes. J. Am. Coll. Cardiol. 23: 259–277, 1994.
 16. Antzelevitch, C., S. Sicouri, S. H. Litovsky, A. Lukas, S. C. Krishnan, J. M. Di Diego, G. A. Gintant, and D. W. Liu. Heterogeneity within the ventricular wall: electrophysiology and pharmacology of epicardial, endocardial and M cells. Circ. Res. 69: 1427–1449, 1991.
 17. Antzelevitch, C., S. Sicouri, A. Lukas, J. M. Di Diego, V. V. Nesterenko, D. W. Liu, J. F. Roubache, A. C. Zygmunt, Z. Q. Zhang, and A. Iodice. Clinical implications of electrical heterogeneity in the heart: the electrophysiology and pharmacology of epicardial, M and endocardial cells. In: Cardiac Arrhythmia: Mechanism, Diagnosis and Management, edited by P. J. Podrid, and P. R. Kowey. Baltimore: William & Wilkins. 1995: 88–107.
 18. Antzelevitch, C., S. Sicouri, A. Lukas, V. V. Nesterenko, D. W. Liu, and J. M. Di Diego. Regional differences in the electrophysiology of ventricular cells: physiological and clinical implications. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. P. Zipes and J. Jalife. Philadelphia: W. B. Saunders, 1995: 228–245.
 19. Antzelevitch, C., Z. Q. Sun, Z. Q. Zhang, and G. X. Yan. Cellular and ionic mechanisms underlying erythromycin‐induced long QT and torsade de pointes. J. Am. Coll. Cardiol. 28: 1836–1848, 1996.
 20. Antzelevitch, C., G. X. Yan, W. Shimizu, and A. Burashnikov. Electrical heterogeneity, the ECG, and cardiac arrhythmias. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. P. Zipes and J. Jalife. Philadelphia: W. B. Saunders, 1999: 222–238.
 21. Anyukhovsky, E. P., E. A. Sosunov, S. J. Feinmark, and M. R. Rosen. Effects of quinidine on repolarization in canine epicardium, midmyocardium, and endocardium. II. In vivo study. Circulation 96: 4019–4026, 1997.
 22. Anyukhovsky, E. P., E. A. Sosunov, R. Z. Gainullin, and M. R. Rosen. The controversial M cell. J. Cardiovasc. Electrophysiol. 10: 244–260, 1999.
 23. Anyukhovsky, E. P., E. A. Sosunov, and M. R. Rosen. Regional differences in electrophysiologic properties of epicardium, midmyocardium and endocardium: in vitro and in vivo correlations. Circulation 94: 1981–1988, 1996.
 24. Asano, Y., J. M. Davidenko, W. T. Baxter, R. A. Gray, and J. Jalife. J. Am. Coll. Cardiol. 29: 831–842, 1997.
 25. Balati, B., A. Varro, and J. G. Papp. Comparison of the cellular electrophysiological characteristics of canine left ventricular epicardium, M cells, endocardium and Purkinje fibres Acta Physiol. Scand. 164: 181–190, 1998.
 26. Balser, J. R., P. B. Bennett, L. M. Hondeghem, and D. M. Roden. Suppression of time‐dependent outward current in guinea‐pig ventricular myocytes. Actions of quinidine and amiodarone. Circ. Res. 69: 519–529, 1991.
 27. Barhanin, J., F. Lesage, E. Guillemare, M. Fink, M. Lazdunski, and G. Romey. KvLQT1 and IsK (minK) proteins associate to form the IKs cardiac potassium current. Nature 384: 78–80, 1996.
 28. Barry, D. M. and J. M. Nerbonne. Myocardial potassium channels: electrophysiological and molecular diversity. Annu. Rev. Physiol. 58: 363–394, 1996.
 29. Barry, D. M., J. S. Trimmer, J. P. Merlie, and J. M. Nerbonne. Differential expression of voltage‐gated K+ channel subunits in adult rat heart. Relation to functional K+ channels? Circ. Res. 77: 361–369, 1995.
 30. Bauer, A., R. Becker, K. D. Freigang, J. C. Senges, F. Voss, A. Hansen, M. Muller, H. J. Lang, U. Gerlach, A. Busch, P. Kraft, W. Kubler, and W. Schols. Rate‐ and site‐dependent effects of propafenone, dofetilide, and the new I(Ks)‐blocking agent chromanol 293b on individual muscle layers of the intact canine heart. Circulation 100: 2184–2190, 1999.
 31. Bjerregaard, P., I. Gussak, S. I. Kotar, and J. E. Gessler. Recurrent synocope in a patient with prominent J‐wave. Am. Heart J. 127: 1426–1430, 1994.
 32. Blair, R. W., T. Shimizu, and V. S. Bishop. The role of vagal afferents in the reflex control of the left ventricular refractory period in the cat. Circ. Res. 46: 378–386, 1980.
 33. Brahmajothi, M. V., D. L. Campbell, R. L. Rasmusson, M. J. Morales, J. S. Trimmer, J. M. Nerbonne, and H. C. Strauss. Distinct transient outward potassium current (Ito) phenotypes and distribution of fast‐inactivating potassium channel alpha subunits in ferret left ventricular myocytes. J. Gen. Physiol 113: 581–600, 1999.
 34. Brahmajothi, M. V., M. J. Morales, R. Liu, R. L. Rasmusson, D. L. Campbell, and H. C. Strauss. In situ hybridization reveals extensive diversity of K+ channel mRNA in isolated ferret cardiac myocytes. Circ. Res. 78: 1083–1089, 1996.
 35. Brahmajothi, M. V., M. J. Morales, R. L. Rasmusson, D. L. Campbell, and H. C. Strauss. Heterogeneity in K+ channel transcript expression detected in isolated ferret cardiac myocytes. Pacing Clin. Electrophysiol. 20: 388–396, 1997.
 36. Brahmajothi, M. V., M. J. Morales, K. A. Reimer, and H. C. Strauss. Regional localization of ERG, the channel protein responsible for the rapid component of the delayed rectifier, K+ current in the ferret heart. Circ. Res. 81: 128–135, 1997.
 37. Brugada, P. and J. Brugada. Right bundle branch block, persistent ST segment elevation and sudden cardiac death: a distinct clinical and electrocardiographic syndrome: a multicenter report. J. Am. Coll. Cardiol. 20: 1391–1396, 1992.
 38. Bryant, S. M., X. Wan, S. J. Shipsey, and G. Hart. Regional differences in the delayed rectifier current (IKr and IKs) contribute to the differences in action potential duration in basal left ventricular myocytes in guinea‐pig. Cardiovasc. Res. 40: 322–331, 1998.
 39. Buchanan, L. V., G. G. Kabell, M. N. Brunden, and J. K. Gibson. Comparative assessment of ibutilide, D‐sotalol, clofilium, E‐4031, and UK‐68,798 in a rabbit model of proarrhythmia. J. Cardiovasc. Pharmacol. 22: 540–549, 1993.
 40. Burashnikov, A. and C. Antzelevitch. Acceleration‐induced action potential prolongation and early afterdepolarizations. J. Cardiovasc. Electrophysiol. 9: 934–948, 1998.
 41. Burashnikov, A. and C. Antzelevitch. Differences in the electrophysiologic response of four canine ventricular cell types to α1‐adrenergic agonists. Cardiovasc. Res. 43: 901–908, 1999.
 42. Burashnikov, A. and Antzelevitch, C. Is the Purkinje system the source of the electrocardiographic U wave? Circulation 100: II–386, 1999. (Abstract)
 43. Burgess, M. J., L. S. Green, K. Millar, R. F. Wyatt, and J. A. Abildskov. The sequence of normal ventricular recovery. Am. Heart J. 84: 660–669, 1972.
 44. Busch, A. E., A. E. Bush, E. Ford, H. Suessbrich, H. J. Lang, R. Greger, K. Kunzelmann, B. Attali, and W. Stümer. The role of the Isk protein in the specific pharmacological properties of the IKs channel complex. Br. J. Pharmacol. 122: 187–189, 1998.
 45. Butler, A., A. G. Wei, K. Baker, and L. Salkoff. A family of putative potassium channel genes in Drosophila. Science 243: 943–947, 1989.
 46. Carl, S. L., K. Felix, A. H. Caswell, N. R. Brandt, W. J. Ball, Jr., P. L. Vaghy, G. Meissner, and D. G. Ferguson. Immunolocalization of sarcolemmal dihydropyridine receptor and sarcoplasmic reticular triadin and ryanodine receptor in rabbit ventricle and atrium. J. Cell Biol. 129: 672–682, 1995.
 47. Carlsson, L., O. Almgren, and G. D. Duker. Qtu‐Prolongation and torsades‐de‐pointes induced by putative class‐III antiarrhythmic agents in the rabbit—etiology and interventions. J. Cardiovasc. Pharmacol. 16: 276–285, 1990.
 48. Catterall, W. A. Structure and function of voltage‐gated ion channels. Trends Neurosci. 16: 500–506, 1993.
 49. Chandy, K. G. Simplified gene nomenclature [letter]. Nature 352: 26, 1991.
 50. Chen, F., G. Mottino, T. S. Klitzner, K. D. Philipson, and J. S. Frank. Distribution of the Na+/Ca2+ exchange protein in developing rabbit myocytes. Am. J. Physiol 268 (Cell Physiol. 35): C1126–C1132, 1995.
 51. Chen, Q., G. E. Kirsch, D. Zhang, R. Brugada, J. Brugada, P. Brugada, D. Potreau, A. Moya, M. Borggrefe, G. Breithardt, M. Ortiz, Z. G. Wang, C. Antzelevitch, R. E. O'Brien, E. Schultz‐Bahr, M. T. Keating, J. A. Towbin, and Q. Wang. Genetic basis and molecular mechanisms for idiopathic ventricular fibrillation. Nature 392: 293–296, 1997.
 52. Choy, A. M., C. C. Lang, D. M. Chomsky, G. H. Rayos, J. R. Wilson, and D. M. Roden. Normalization of acquired QT prolongation in humans by intravenous potassium. Circulation 96: 2149–2154, 1997.
 53. Clark, R. B., R. A. Bouchard, E. Salinas‐Stefanon, J. Sanchez‐Chapula, and W. R. Giles. Heterogeneity of action potential waveforms and potassium currents in rat ventricle. Cardiovasc. Res. 27: 1795–1799, 1993.
 54. Clements, S. D. and J. W. Hurst. Diagnostic value of ECG abnormalities observed in subjects accidentally exposed to cold. Am. J. Cardiol. 29: 729–734, 1972.
 55. Cobbe, S. M., E. P. Hoffman, A. Ritzenhoff, J. Brachmann, W. Kubler, and J. Senges. Action of sotalol on potential reentrant pathways and ventricular tachyarrhythmias in conscious dogs in the late postmyocardial infarction phase. Circulation 68: 865–871, 1983.
 56. Cohen, I. S., W. R. Giles, and D. Noble. Cellular basis for the T wave of the electrocardiogram. Nature 262: 657–661, 1976.
 57. Cohen, S. A. Immunocytochemical localization of rH1 sodium channel in adult rat heart atria and ventricle. Presence in terminal intercalated disks. Circulation 94: 3083–3086, 1996.
 58. Compton, S. J., R. L. Lux, M. R. Ramsey, K. R. Strelich, M. C. Sanguinetti, L. S. Green, M. T. Keating, and J. W. Mason. Genetically defined therapy of inherited long‐QT syndrome. Correction of abnormal repolarization by potassium. Circulation 94: 1018–1022, 1996.
 59. Coumel, P. Early afterdepolarizations and triggered activity in clinical arrhythmias. In: Cardiac Electrophysiology: A Text Book, edited by M. R. Rosen, M. J. Janse, and A. L. Wit. Mount Kisco, NY: Futura Publishing. 1990: 387–411.
 60. Crampton, R. S. Preeminence of the left stellate ganglion in the long Q‐T syndrome. Circulation 59: 769–778, 1979.
 61. Curran, M. E., I. Splawski, K. W. Timothy, G. M. Vincent, E. D. Green, and M. T. Keating. A molecular basis for cardiac arrhythmia: HERG mutations cause long QT syndrome. Cell 80: 795–803, 1995.
 62. Daleau, P. and J. Deleze. Conduction block in Purkinje fibers by homogeneous versus localized decrease of the gap junction conductance. Can. J Physiol Pharmacol. 76: 630–641, 1998.
 63. Davidenko, J. M., L. Cohen, R. J. Goodrow, and C. Antzelevitch. Quinidine‐induced action potential prolongation, early afterdepolarizations, and triggered activity in canine Purkinje fibers. Effects of stimulation rate, potassium, and magnesium. Circulation 79: 674–686, 1989.
 64. De Biasi, M., Z. G. Wang, E. Accili, B. A. Wible, and D. Fedida. Open channel block of human heart hKv1.5 by the β‐subunit hKvβ1.2. Am. J. Physiol. 272 (Heart Circ. Physiol. 41): H2932–H2941, 1997.
 65. Delombe, S., I. Baró, Y. Péréon, J. Bliek, R. Mohammad‐Panah, H. Pollard, S. Morid, M. Mannens, A. A. M. Wilde, J. Barhanin, F. Charpentier, and D. Escande. A dominant negative isoform of the long QT syndrome 1 gene product. J. Biol. Chem. 273: 6837–6843, 1998.
 66. Derakhchan, K., R. Cardinal, S. Brunet, D. Klug, C. Pharand, T. Kus, and B. I. Sasyniuk. Polymorphic ventricular tachycardias induced by d‐sotalol and phenylephrine in canine preparations of atrioventricular block: initiation in the conduction system followed by spatially unstable re‐entry. Cardiovasc. Res. 38: 617–630, 1998.
 67. Di Diego, J. M. and C. Antzelevitch. Pinacidil‐induced electrical heterogeneity and extrasystolic activity in canine ventricular tissues: does activation of ATP‐regulated potassium current promote phase 2 reentry? Circulation 88: 1177–1189, 1993.
 68. Di Diego, J. M. and C. Antzelevitch. High [Ca2+]‐induced electrical heterogeneity and extrasystolic activity in isolated canine ventricular epicardium: phase 2 reentry. Circulation 89: 1839–1850, 1994.
 69. Di Diego, J. M., Z. Q. Sun, and C. Antzelevitch. Ito and action potential notch are smaller in left vs. right canine ventricular epicardium. Am. J. Physiol. 271 (Heart Circ. Physiol. 40): H548–H561, 1996.
 70. Dillon, S. M., M. A. Allessie, P. C. Ursell, and A. L. Wit. Influences of anisotropic tissue structure on reentrant circuits in the epicardial border zone of subacute canine infarcts. Circ. Res. 63: 182–206, 1988.
 71. Dixon, E. J. and D. McKinnon. Quantitative analysis of potassium channel mRNA in atrial and ventricular muscle of rats. Circ. Res. 75: 252–260, 1994.
 72. Dixon, E. J., W. Shi, H.‐S. Wang, C. McDonald, H. Yu, R. S. Wymore, I. S. Cohen, and D. McKinnon. Role of the Kv4.3 K+ channel in ventricular muscle. A molecular correlate for the transient outward current. Circ. Res. 79: 659–668, 1996.
 73. Drouin, E., F. Charpentier, C. Gauthier, K. Laurent, and H. Le Marec. Electrophysiological characteristics of cells spanning the left ventricular wall of human heart: evidence for the presence of M cells. J. Am. Coll. Cardiol. 26: 185–192, 1995.
 74. Dumaine, R., J. A. Towbin, P. Brugada, M. Vatta, V. V. Nesterenko, D. V. Nesterenko, R. Brugada, and C. Antzelevitch. Ionic mechanisms responsible for the electrocardiographic phenotype of the Brugada syndrome are temperature dependent. Circ. Res. 85: 803–809, 1999.
 75. Dumaine, R., Q. Wang, M. T. Keating, H. A. Hartmann, P. J. Schwartz, A. M. Brown, and G. E. Kirsch. Multiple mechanisms of Na+ channel‐linked long‐QT syndrome. Circ. Res. 78: 916–924, 1996.
 76. Dumaine, R., Y. S. Wu, and C. Antzelevitch. Distribution of KvLQT1 but not minK parallels the distribution of IKs in the mid‐myocardium of canine heart. Biophys. J. 76: A366–A366, 2000.
 77. Eagle, K. Images in clinical medicine. Osborn waves of hypothermia. N. Engl. J. Med. 10: 680, 1994.
 78. Eddlestone, G. T., Zygmunt, A. C., and Antzelevitch, C. Larger late sodium current contributes to the longer action potential of the M cell in canine ventricular myocardium. Pacing Clin. Electrophysiol. 19: II569, 1996. (Abstract)
 79. Eisner, D. A. The Na‐K pump and its effectors in cardiac muscle. In: The Heart and Cardiovascular System, edited by H. A. Fozzard, R. B. Jennings, E. Haber, A. M. Katz, and H. E. Morgan. New York: Raven Press, 2000: 863–902.
 80. El‐Sherif, N., E. B. Caref, H. Yin, and M. Restivo. The electrophysiological mechanism of ventricular arrhythmias in the long QT syndrome: tridimensional mapping of activation and recovery patterns. Circ. Res. 79: 474–492, 1996.
 81. El‐Sherif, N., M. Chinushi, E. B. Caref, and M. Restivo. Electrophysiological mechanism of the characteristic electrocardiographic morphology of torsade de pointes tachyarrhythmias in the long‐QT syndrome. Detailed analysis of ventricular tridimensional activation patterns. Circulation 96: 4392–4399, 1997.
 82. Escande, D., D. Loisance, C. Planche, and E. Coraboeuf. Age‐related changes of action potential plateau shape in isolated human atrial fibers. Am. J. Physiol. 249 (Heart Circ. Physiol. 18): H843–H850, 1985.
 83. Fedida, D. and W. R. Giles. Regional variations in action potentials and transient outward current in myocytes isolated from rabbit left ventricle. J. Physiol. (Lond.) 442: 191–209, 1991.
 84. Feng, J., B. A. Wible, G. R. Li, Z. G. Wang, and S. Nattel. Antisence oligodeoxynucleotides directed against Kv1.5 mRNA specifically inhibit ultrarapid delayed rectifier K+ current in cultured adult human atrial myocytes. Circ. Res. 80: 572–579, 1997.
 85. Fontaine, G. A new look at torsades de pointes. In: QT Prolongation and Ventricular Arrhythmias, edited by K. Hashiba, A. J. Moss, and P. J. Schwartz. New York: New York Academy of Science, 1992: 157–177.
 86. Fozzard, H. A. and D. A. Hanck. Structure and function of voltage‐dependent sodium channels: comparison of brain II and cardiac isoforms. Physiol. Rev. 76: 887–926, 1996.
 87. Fozzard, H. A. and G. Lipkind. The guanidinium toxin binding site on the sodium channel. Jpn. Heart J. 37: 683–692, 1996.
 88. Frank, J. S., G. Mottino, D. Reid, R. S. Molday, and K. D. Philipson. Distribution of the Na(+)‐Ca2+ exchange protein in mammalian cardiac myocytes: an immunofluorescence and immunocolloidal gold‐labeling study. J. Cell Biol. 117: 337–345, 1992.
 89. Freigang, K. D., Becker, R., Bauer, A., Voss, F., Senges, J., and Brachmann, J. Electrophysiological properties of individual muscle layers in the in vivo canine heart. J. Amer. Coll. Cardiol. 27 (Suppl A): 124A. 1996. (Abstract)
 90. Furukawa, T., R. J. Myerburg, N. Furukawa, A. L. Bassett, and S. Kimura. Differences in transient outward currents of feline endocardial and epicardial myocytes. Circ. Res. 67: 1287–1291, 1990.
 91. Gao, T., T. S. Puri, B. L. Gerhardstein, A. J. Chien, R. D. Green, and M. M. Hosey. Identification and subcellular localization of the subunits of L‐type calcium channels and adenylyl cyclase in cardiac myocytes. J. Biol. Chem. 272: 19401–19407, 1997.
 92. Gidh‐Jain, M., B. Huang, P. Jain, and N. el Sherif. Differential expression of voltage‐gated K+ channel genes in left ventricular remodeled myocardium after experimental myocardial infarction. Circ. Res 79: 669–675, 1996.
 93. Gilmour, R. F., Jr. and D. P. Zipes. Different electrophysiological responses of canine endocardium and epicardium to combined hyperkalemia, hypoxia, and acidosis. Circ. Res. 46: 814–825, 1980.
 94. Gourdie, R. G., C. R. Green, and N. J. Severs. Gap junction distribution in adult mammalian myocardium revealed by an anti‐peptide antibody and laser scanning confocal microscopy. J. Cell Sci. 99 (Pt 1): 41–55, 1991.
 95. Gourdie, R. G., N. J. Severs, C. R. Green, S. Rothery, P. Germroth, and R. P. Thompson. The spatial distribution and relative abundance of gap‐junctional connexin40 and connexin43 correlate to functional properties of components of the cardiac atrioventricular conduction system. J. Cell Sci. 105 (Pt 4): 985–991, 1993.
 96. Gros, D., T. Jarry‐Guichard, I. Ten Velde, A. de Maziere, M. J. van Kempen, J. Davoust, J. P. Briand, A. F. Moorman, and H. J. Jongsma. Restricted distribution of connexin40, a gap junctional protein, in mammalian heart. Circ. Res. 74: 839–851, 1994.
 97. Gussak, I., P. Bjerregaard, T. M. Egan, and B. R. Chaitman. ECG phenomenon called the J wave. History, pathophysiology, and clinical significance. J. Electrocardiol. 28: 49–58, 1995.
 98. Habbab, M. A. and N. El‐Sherif. TU alternans, long QTU, and torsade de pointes: clinical and experimental observations. Pacing Clin. Electrophysiol. 15: 916–931, 1992.
 99. Harvey, R. D. and J. R. Hume. Isoproterenol activates a chloride current, not the transient outward current, in rabbit ventricular myocytes. Am. J. Physiol. 265 (Cell Physiol. 34): C1177–C1181, 1993.
 100. Haws, C. W. and R. L. Lux. Correlation between in vivo transmembrane action potential durations and action‐recovery intervals from electrograms. Effects of interventions that alter repolarization time. Circulation 81: 281–288, 1990.
 101. Hirao, H., W. Shimizu, T. Kurita, K. Suyama, N. Aihara, S. Kamakura, and K. Shimomura. Frequency‐dependent electrophysiologic properties of ventricular repolarization in patients with congenital long QT syndrome. J. Am. Coll. Cardiol. 28: 1269–1277, 1996.
 102. Hofmann, F., M. Biel, and V. Flockerzi. Molecular basis for Ca2+ channel diversity. Annu. Rev. Neurosci. 17: 399–418, 1994.
 103. Hu, H. and E. Marban. Isoform‐specific inhibition of L‐type calcium channels by dihydropyridines is independent of isoform‐specific gating properties. Mol. Pharmacol. 53: 902–907, 1998.
 104. Hugo, N., I. C. Dormehl, and A. L. Van Gelder. A positive wave at the J‐point of electrocardiograms of anaesthetized baboons. J. Med. Primatol. 17: 347–352, 1988.
 105. Hume, J. R. and R. D. Harvey. Chloride conductance pathways in heart. Am. J. Physiol. 261 (Cell Physiol. 30): C399–C412, 1991.
 106. Imaizumi, Y. and W. R. Giles. Quinidine‐induced inhibition of transient outward current in cardiac muscle. Am. J. Physiol. 253 (Heart Circ. Physiol. 22): H704–H708, 1987.
 107. James, A. F., T. Tominaga, Y. Okada, and M. Tominaga. Distribution of cAMP‐activated chloride current and CFTR mRNA in the guinea pig heart. Circ. Res. 79: 201–207, 1996.
 108. Jeck, C. D. and P. A. Boyden. Age‐related appearance of outward currents may contribute to developmental differences in ventricular repolarization. Circ. Res. 71: 1390–1403, 1992.
 109. Jewell, E. A. and J. B. Lingrel. Comparison of the substrate dependence properties of the rat Na,K‐ATPase alpha 1, alpha 2, and alpha 3 isoforms expressed in HeLa cells. J Biol. Chem. 266: 16925–16930, 1991.
 110. Kaab, S., J. Dixon, J. Duc, D. Ashen, M. Nabauer, D. J. Beuckelmann, G. Steinbeck, D. McKinnon, and G. F. Tomaselli. Molecular basis of transient outward potassium current downregulation in human heart failure: a decrease in Kv4.3 mRNA correlates with a reduction in current density. Circulation 98: 1383–1393, 1998.
 111. Kamiyama, A. and Y. Saeki. Myocardial action potentials of right‐ and left‐subepicardial muscles in the canine ventricle and effects of manganese ions. Proc. Jpn. Acad. 50: 771–774, 1974.
 112. Kamp, T. J., M. Mitas, K. L. Fields, S. Asoh, H. Chin, E. Marban, and M. Nirenberg. Transcriptional regulation of the neuronal L‐type calcium channel alpha 1D subunit gene. Cell Mol. Neurobiol. 15: 307–326, 1995.
 113. Kanter, H. L., J. G. Laing, S. L. Beau, E. C. Beyer, and J. E. Saffitz. Distinct patterns of connexin expression in canine Purkinje fibers and ventricular muscle. Circ. Res. 72: 1124–1131, 1993.
 114. Kaplan, W. D. and W. E. Trout, III. The behavior of four neurological mutants of Drosophila. Genetics 61: 399–409, 1969.
 115. Kieval, R. S., R. J. Bloch, G. E. Lindenmayer, A. Ambesi, and W. J. Lederer. Immunofluorescence localization of the Na‐Ca exchanger in heart cells. Am. J. Physiol. 263 (Cell Physiol. 32): C545–C550, 1992.
 116. Kilborn, M. J. and D. Fedida. A study of the developmental changes in outward currents of rat ventricular myocytes. J. Physiol. (Lond.) 430: 37–60, 1990.
 117. Kimura, S., A. L. Bassett, T. Kohya, P. L. Kozlovskis, and R. J. Myerburg. Regional effects of verapamil on recovery of excitability and conduction time in experimental ischemia. Circulation 76: 1146–1154, 1987.
 118. Kimura, S., H. Nakaya, and M. Kanno. Electrophysiological effects of diltiazem, nifedipine and Ni2+ on the subepicardial muscle cells of canine heart under the condition of combined hypoxia, hyperkalemia and acidosis. Naunyn‐Schmiedebergs Arch. Pharmacol. 324: 228–232, 1983.
 119. Koban, M. U., A. F. Moorman, J. Holtz, M. H. Yacoub, and K. R. Boheler. Expressional analysis of the cardiac Na‐Ca exchanger in rat development and senescence. Cardiovasc. Res. 37: 405–423, 1998.
 120. Kong, W., S. Po, T. Yamagishi, M. D. Ashen, G. Stetten, and G. F. Tomaselli. Isolation and characterization of the human gene encoding Ito: further diversity by alternative mRNA splicing. Am. J. Physiol. 275 (Heart Circ. Physiol. 44): H1963–H1970, 1998.
 121. Kossenjans, W. and M. Ashraf. Localization of sodiumpotassium adenosine triphosphatase in sheep myocardium by immunoelectron microscopy. J. Submicrosc. Cytol. Pathol. 20: 53–58, 1988.
 122. Kraus, F. Ueber die wirkung des kalziums auf den kreislauf. Deutsch. Med. Wochenschr. 46: 201–203, 1920.
 123. Krishnan, S. C. and C. Antzelevitch. Sodium channel blockade produces opposite electrophysiologic effects in canine ventricular epicardium and endocardium. Circ. Res. 69: 277–291, 1991.
 124. Krishnan, S. C. and C. Antzelevitch. Flecainide‐induced arrhythmia in canine ventricular epicardium: Phase 2 Reentry? Circulation 87: 562–572, 1993.
 125. Lehmann, M. H., F. Suzuki, B. S. Fromm, D. Frankovich, P. Elko, R. T. Steinman, J. Fresard, J. J. Baga, and R. T. Taggart. T‐wave “humps” as a potential electrocardiographic marker of the long QT syndrome. J. Am. Coll. Cardiol. 24: 746–754, 1994.
 126. Li, G. R., J. Feng, L. Yue, and M. Carrier. Transmural heterogeneity of action potentials and Itol in myocytes isolated from the human right ventricle. Am. J. Physiol. 275 (Heart Circ Physiol. 44): H369–H377, 1998.
 127. Li, Z., S. Matsuoka, L. V. Hryshko, D. A. Nicoll, M. M. Bersohn, E. P. Burke, R. P. Lifton, and K. D. Philipson. Cloning of the NCX2 isoform of the plasma membrane Na(+)‐Ca2;+ exchanger. J Biol. Chem. 269: 17434–17439, 1994.
 128. Litovsky, S. H. and C. Antzelevitch. Transient outward current prominent in canine ventricular epicardium but not endocardium. Circ. Res. 62: 116–126, 1988.
 129. Litovsky, S. H. and C. Antzelevitch. Rate dependence of action potential duration and refractoriness in canine ventricular endocardium differs from that of epicardium: the role of the transient outward current. J. Am. Coll. Cardiol. 14: 1053–1066, 1989.
 130. Litovsky, S. H. and C. Antzelevitch. Differences in the electrophysiological response of canine ventricular subendocardium and subepicardium to acctylcholine and isoproterenol. A direct effect of acetylcholine in ventricular myocardium. Circ. Res. 67: 615–627, 1990.
 131. Liu, D. W. and C. Antzelevitch. Characteristics of the delayed rectifier current (IKr and IKs) in canine ventricular epicardial, midmyocardial and endocardial myocytes: a weaker IKs contributes to the longer action potential of the M cell. Circ. Res. 76: 351–365, 1995.
 132. Liu, D. W., G. A. Gintant, and C. Antzelevitch. Ionic bases for electrophysiological distinctions among epicardial, midmyocardial, and endocardial myocytes from the free wall of the canine left ventricle. Circ. Res. 72: 671–687, 1993.
 133. Locati, E. H., P. Maison‐Blanche, P. Dejode, B. Cauchemez, and P. Coumel. Spontaneous sequences of onset of torsade de pointes in patients with acquired prolonged repolarization: quantitative analysis of Holter recordings. J. Am. Coll. Cardiol. 25: 1564–1575, 1995.
 134. Lubinski, A., E. Lewicka‐Nowak, M. Kempa, A. M. Baczynska, I. Romanowska, and G. Swiatecka. New insight into repolarization abnormalities in patients with congenital long QT syndrome: the increased transmural dispersion of repolarization. Pacing Clin. Electrophysiol. 21: 172–175, 1998.
 135. Lucchesi, P. A. and K. J. Sweadner. Postnatal changes in Na,K‐ATPase isoform expression in rat cardiac ventricle. Conservation of biphasic ouabain affinity. J. Biol. Chem. 266: 9327–9331, 1991.
 136. Lukas, A. and C. Antzelevitch. Differences in the electrophysiological response of canine ventricular epicardium and endocardium to ischemia: role of the transient outward current. Circulation 88: 2903–2915, 1993.
 137. Lukas, A. and C. Antzelevitch. Phase 2 reentry as a mechanism of initiation of circus movement reentry in canine epicardium exposed to simulated ischemia. The antiarrhythmic effects of 4‐aminopyridine. Cardiovasc. Res. 32: 593–603, 1996.
 138. Lukas, A. and C. Antzelevitch. The contribution of K+ currents to electrical heterogeneity across the canine ventricular wall under normal and ischemic conditions. In: Pathophysiology of Heart Failure, edited by N. S. Dhalla, G. N. Pierce, and V. Panagia. Boston: Academic Publishers, 1996: 440–456.
 139. Lunkenheimer, P. P., K. Redmann, H. H. Scheld, K.‐H. Dietl, C. Cryer, K.‐D. Richter, J. Merker, and W. Whimster. The heart muscle's putative “secondary structure.” Functional implications of a band‐like anisotropy. Technol. Health Care 5: 53–64, 1997.
 140. MacKinnon, R. Determination of the subunit stoichiometry of a voltage‐activated potassium channel. Nature 350: 232–235, 1991.
 141. Martin‐Vasallo, P., W. Dackowski, J. R. Emanuel, and R. Levenson. Identification of a putative isoform of the Na,K‐ATPase beta subunit. Primary structure and tissue‐specific expression. J. Biol. Chem. 264: 4613–4618, 1989.
 142. Matsuo, K., W. Shimizu, T. Kurita, K. Suyama, N. Aihara, S. Kamakura, and K. Shimomura. Increased dispersion of repolarization time determined by monophasic action potentials in two patients with familial idiopathic ventricular fibrillation. J. Cardiovasc. Electrophysiol. 9: 74–83, 1998.
 143. Mays, D. J., J. M. Foose, L. H. Philipson, and M. M. Tamkun. Localization of the Kv1.5 K+ channel protein in explanted cardiac tissue. J. Clin. Invest 96: 282–292, 1995.
 144. Mays, D. J., M. M. Tamkun, and P. A. Boyden. Redistribution of the Kv1.5 K+ channel protein on the surface of myocytes from the epicardial border zone of infarcted canine ventricle. Cardiovasc. Pathobiol. 2: 79–87, 2000.
 145. Miyazaki, T., H. Mitamura, S. Miyoshi, K. Soejima, Y. Aizawa, and S. Ogawa. Autonomic and antiarrhythmic drug modulation of ST segment elevation in patients with Brugada syndrome. J. Am. Coll. Cardiol. 27: 1061–1070, 1996.
 146. Mori, Y., G. Mikala, G. Varadi, T. Kobayashi, S. Koch, M. Wakamori, and A. Schwartz. Molecular pharmacology of voltage‐dependent calcium channels. Jpn. J. Pharmacol. 72: 83–109, 1996.
 147. Moss, A. J. Long QT syndrome. In: Cardiac Arrhythmia: Mechanisms, Diagnosis and Management, edited by P. J. Podrid, and P. R. Kowey. Baltimore, William Wilkins. 1995: 1110–1120.
 148. Moss, A. J., P. J. Schwartz, R. S. Crampton, D. Tzivoni, E. H. Locati, J. W. MacCluer, W. J. Hall, L. R. Weitkamp, G. M. Vincent, A. Garson, J. L. Robinson, J. Benhorin, and S. Choi. The long QT syndrome: prospective longitudinal study of 328 families. Circulation 84: 1136–1144, 1991.
 149. Mubagwa, K. and E. Carmeliet. Effects of acetylcholine on electrophysiological properties of rabbit cardiac Purkinje fibers. Circ. Res. 53: 740–751, 1983.
 150. Nabauer, M., D. J. Beuckelmann, P. Uberfuhr, and G. Steinbeck. Regional differences in current density and ratedependent properties of the transient outward current in subepicardial and subendocardial myocytes of human left ventricle. Circulation 93: 168–177, 1996.
 151. Nakayama, T. and H. A. Fozzard. Adrenergic modulation of the transient outward current in isolated canine Purkinje cells. Circ. Res. 62: 162–172, 1988.
 152. Ng, Y. C. and T. Akera. Relative abundance of two molecular forms of Na+,K+‐ATPase in the ferret heart: developmental changes and associated alterations of digitalis sensitivity. Mol. Pharmacol. 32: 201–205, 1987.
 153. 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.
 154. Nicoll, D. A., B. D. Quednau, Z. Qui, Y. R. Xia, A. J. Lusis, and K. D. Philipson. Cloning of a third mammalian Na+‐Ca2+ exchanger, NCX3. J. Biol. Chem. 271: 24914–24921, 1996.
 155. Noble, D. and I. S. Cohen. The interpretation of the T wave of the electrocardiogram. Cardiovasc. Res. 12: 13–27, 1978.
 156. Orlowski, J. and J. B. Lingrel. Tissue‐specific and developmental regulation of rat Na,K‐ATPase catalytic alpha isoform and beta subunit mRNAs. J. Biol. Chem. 263: 10436–10442, 1988.
 157. Osborn, J. J. Experimental hypothermia: respiratory and blood pH changes in relation to cardiac function. Am. J. Physiol. 175: 389–398, 1953.
 158. Pacioretty, L. M. and R. F. Gilmour, Jr. Developmental changes in the transient outward potassium current in canine epicardium. Am. J. Physiol. 268 (Heart Circ. Physiol. 37): H2513–H2521, 1995.
 159. Pertsov, A. M., J. M. Davidenko, R. Salomonsz, W. T. Baxter, and J. Jalife. Spiral waves of excitation underlie reentrant activity in isolated cardiac muscle. Circ. Res. 72: 631–650, 1993.
 160. Plonsey, R. Action potential sources and their volume conductor fields. Proc. IEEE 65: 601–611, 1977.
 161. Porzig, H., Z. Li, D. A. Nicoll, and K. D. Philipson. Mapping of the cardiac sodium‐calcium exchanger with monoclonal antibodies. Am. J. Physiol. 265 (Cell Physiol. 34): C748–C756, 1993.
 162. Prestle, J., S. Dieterich, M. Preuss, U. Bieligk, and G. Hasenfuss. Heterogeneous transmural gene expression of calciumhandling proteins and natriuretic peptides in the failing human heart Cardiovasc. Res. 43: 323–331, 1999.
 163. Priori, S. G., C. Napolitano, U. Glordano, G. Collisani, and M. Memml. Brugada syndrome and sudden cardiac death in children. Lancet 355: 808–809, 2000.
 164. Prystowsky, E. N., W. M. Jackman, R. L. Rinkenberger, et al. Effect of autonomic blockade on ventricular refractoriness and atrioventricular nodal conduction in man. Evidence supporting a direct cholinergic action on ventricular muscle refractoriness. Circ. Res. 49: 511–518, 1981.
 165. Ramakers, C., Doevendans, P. A., Vos, M. A., Antzelevitch, C., and Dumaine, R. KCNQ1 and KCNE1 expression is reduced in dogs with chronic AV block. Biophys. J. 78: 220A, 2000. (Abstract)
 166. Reder, R. F., D. S. Miura, P. Danilo, and M. R. Rosen. The electrophysiological properties of normal neonatal and adult canine cardiac Purkinje fibers. Circ. Res. 48: 658–668, 1981.
 167. Roberds, S. L. and M. M. Tamkun. Cloning and tissue‐specific expression of five voltage‐gated potassium channel cDNAs expressed in the heart. Proc. Natl. Acad. Sci. U.S.A. 88: 1798–1802, 1991.
 168. Rodriguez‐Sinovas, A., J. Cinca, A. Tapias, L. Armadans, M. Tresanchez, and J. Soler‐Soler. Lack of evidence of M‐cells in porcine left ventricular myocardium. Cardiovasc. Res. 33: 307–313, 1997.
 169. Saeki, Y. and A. Kamiyama. Possible mechanism of ratedependent change of contraction in dog ventricular muscle: relation to calcium movements. In Recent Advances in Studies on Cardiac Structure and Metabolism, Vol. II, edited by T. Kobayashi, R. Sano, and N. S. Dhalla. Baltimore: University Park Press. 1978: 131–135.
 170. Salata, J. J., N. K. Jurkiewicz, J. J. Wang, and H. T. Orme. A novel benzodiazepine that activates cardiac slow delayed rectifier K+ currents. Mol. Pharmacol. 54: 220–230, 1998.
 171. Sanguinetti, M. C., M. E. Curran, P. S. Spector, and M. T. Keating. Spectrum of HERG K+‐channel dysfunction in an inherited cardiac arrhythmia. Proc. Natl. Acad. Sci. U.S.A. 93: 2208–2212, 1996.
 172. Sanguinetti, M. C., M. E. Curran, A. R. Zou, J. Shen, P. S. Spector, D. L. Atkinson, and M. T. Keating. Coassembly of KvLQT1 and minK (IsK) proteins to form cardiac IKs potassium channel. Nature 384: 80–83, 1996.
 173. Sanguinetti, M. C., C. Jiang, M. E. Curran, and M. T. Keating. A mechanistic link between an inherited and an acquired cardiac arrhythmia: HERG encodes the IKr potassium channel. Cell 81: 299–307, 1995.
 174. Santos, E. M. and K. C. Frederick. Electrocardiographic changes in the dog during hypothermia. Am. Heart J. 55: 415–420, 1957.
 175. Schwartz, P. J. The idiopathic long QT syndrome: progress and questions. Am. Heart J. 109: 399–411, 1985.
 176. Schwartz, P. J. The long QT syndrome. In Clinical Approaches to Tachyarrhythmias, edited by A. J. Camm. Armonk, NY: Futura Publishing Company, 1997: 53–53.
 177. Schwartz, P. J., Malteo, P. S., Moss, A. J., Priori, S. G., Wang, Q., Lehmann, M. H., Timothy, K. W., Denjoy, I. F., Haverkamp, W., Guicheney, P., Paganini, V., Scheinman, M. M., and Karnes, P. S. Gene‐specific influence on the triggers for cardiac arrest in the long QT syndrome. Circulation 96: 1–212. 1997. (Abstract)
 178. Schwartz, P. J., S. G. Priori, E. H. Locati, C. Napolitano, F. Cantu, J. A. Towbin, M. T. Keating, H. Hammoude, A. M. Brown, L. S. K. Chen, and T. J. Colatsky. Long QT syndrome patients with mutations of the SCN5A and HERG genes have differential responses to Na+ channel blockade and to increases in heart rate: Implications for gene‐specific therapy. Circulation 92: 3381–3386, 1995.
 179. Schwartz, P. J., M. Stramba‐Badiale, A. Segantini, P. Austoni, G. Bosi, R. Giorgetti, F. Grancini, E. D. Marni, F. Perticone, D. Rosti, and P. Salice. Prolongation of the QT interval and the sudden infant death syndrome N. Engl. J. Med. 338: 1709–1714, 1998.
 180. Shamraj, O. I., D. Melvin, and J. B. Lingrel. Expression of Na,K‐ATPase isoforms in human heart. Biochem. Biophys. Res. Commun. 179: 1434–1440, 1991.
 181. Shimizu, W. and C. Antzelevitch. Sodium channel block with mexiletine is effective in reducing dispersion of repolarization and preventing torsade de pointes in LQT2 and LQT3 models of the long‐QT syndrome. Circulation 96: 2038–2047, 1997.
 182. Shimizu, W. and C. Antzelevitch. Cellular basis for the electrocardiographic features of the LQT1 form of the long QT syndrome: Effects of β‐adrenergic agonists, antagonists and sodium channel blockers on transmural dispersion of repolarization and torsade de pointes. Circulation 98: 2314–2322, 1998.
 183. Shimizu, W. and C. Antzelevitch. Cellular and ionic basis for T wave alternans under long QT conditions. Circulation 99: 1499–1507, 1999.
 184. Shimizu, W. and C. Antzelevitch. Differential response to ‐ adrenergic agonists and antagonists in LQT1, LQT2 and LQT3 models of the long QT syndrome. J. Am. Coll. Cardiol. 35: 778–786, 2000.
 185. Shimizu, W., T. Kurita, K. Matsuo, N. Aihara, S. Kamakura, J. A. Towbin, and K. Shimomura. Improvement of repolarization abnormalities by a K+ channel opener in the LQT1 form of congenital long QT syndrome. Circulation 97: 1581–1588, 1998.
 186. Shimizu, W., B. McMahon, and C. Antzelevitch. Sodium pentobarbital reduces transmural dispersion of repolarization and prevents torsade de pointes in models of acquired and congenital long QT syndromes. J. Cardiovasc. Electrophysiol. 10: 156–164, 1999.
 187. Shimizu, W., T. Ohe, T. Kurita, M. Kawade, Y. Arakaki, N. Aihara, S. Kamakura, T. Kamiya, and K. Shimomura. Effects of verapamil and propranolol on early afterdepolarizations and ventricular arrhythmias induced by epinephrine in congenital long QT syndrome. J. Am. Coll. Cardiol. 26: 1299–1309, 1995.
 188. Shimizu, W., T. Ohe, T. Kurita, H. Takaki, N. Aihara, S. Kamakura, M. Matsuhisa, and K. Shimomura. Early afterdepolarizations induced by isoproterenol in patients with congenital long QT syndrome. Circulation 84: 1915–1923, 1991.
 189. Shipsey, S. J., S. M. Bryant, and G. Hart. Effects of hypertrophy on regional action potential characteristics in the rat left ventricle: a cellular basis for T‐wave inversion? Circulation 96: 2061–2068, 1997.
 190. Shull, G. E., J. Greeb, and J. B. Lingrel. Molecular cloning of three distinct forms of the Na+,K+‐ATPase alpha‐subunit from rat brain. Biochemistry 25: 8125–8132, 1986.
 191. Shyjan, A. W., V. Cena, D. C. Klein, and R. Levenson. Differential expression and enzymatic properties of the Na+,K(+)‐ATPase alpha 3 isoenzyme in rat pineal glands. Proc. Natl. Acad. Sci. U.S.A 87: 1178–1182, 1990.
 192. Sicouri, S. and C. Antzelevitch. A subpopulation of cells with unique electrophysiological properties in the deep subepicardium of the canine ventricle: the M cell. Circ. Res. 68: 1729–1741, 1991.
 193. Sicouri, S. and C. Antzelevitch. Afterdepolarizations and triggered activity develop in a select population of cells (M cells) in canine ventricular myocardium: the effects of acetylstrophanthidin and Bay K 8644. Pacing Clin. Electrophysiol. 14: 1714–1720, 1991.
 194. Sicouri, S. and C. Antzelevitch. Drug‐induced afterdepolarizations and triggered activity occur in a discrete subpopulation of ventricular muscle cell (M cells) in the canine heart: quinidine and digitalis. J. Cardiovasc. Electrophysiol. 4: 48–58, 1993.
 195. Sicouri, S. and C. Antzelevitch. Electrophysiologic characteristics of M cells in the canine left ventricular free wall. J. Cardiovasc. Electrophysiol. 6: 591–603, 1995.
 196. Sicouri, S., J. Fish, and C. Antzelevitch. Distribution of M cells in the canine ventricle. J. Cardiovasc. Electrophysiol. 5: 824–837, 1994.
 197. Sicouri, S., S. Moro, and M. V. Elizari. d‐Sotalol induces marked action potential prolongation and early afterdepolarizations in M but not epicardial or endocardial cells of the canine ventricle. J. Cardiovasc. Pharmacol. Ther. 2: 27–38, 1997.
 198. Sicouri, S., S. Moro, S. H. Litovsky, M. V. Elizari, and C. Antzelevitch. Chronic amiodarone reduces transmural dispersion of repolarization in the canine heart. J. Cardiovasc. Electrophysiol. 8: 1269–1279, 1997.
 199. Sicouri, S., M. Quist, and C. Antzelevitch. Evidence for the presence of M cells in the guinea pig ventricle. J. Cardiovasc. Electrophysiol. 7: 503–511, 1996.
 200. Slezak, J., W. Schulze, L. Okruhlicova, N. Tribulova, and P. K. Singal. Cytochemical and immunocytochemical localization of Na,K‐ATPase alpha subunit isoenzymes in the rat heart. Mol. Cell. Biochem. 176: 107–112, 1997.
 201. Slezak, J., W. Schulze, Z. Stefankova, L. Okruhlicova, L. Danihel and G. Wallukat. Localization of alpha 1,2,3‐subunit isoforms of Na,K‐ATPase in cultured neonatal and adult rat myocardium: the immunofluorescence and immunocytochemical study. Mol. Cell. Biochem. 163–164: 39–45, 1996.
 202. Snyders, D. J. and P. P. Van Bogaert. Effects of 4‐aminopyridine on inward rectifing and pacemaker currents of cardiac purkinje fibers. Pflugers Arch. 394: 230–238, 1982.
 203. Solberg, L. E., D. H. Singer, R. E. Ten Eick, and E. G. Duffin. Glass microelectrode studies on intramural papillary muscle cells. Circ. Res. 34: 783–797, 1974.
 204. Sosunov, E. A., E. P. Anyukhovsky, and M. R. Rosen. Comparison of repolarization of cells from different layers of myocardium in vitro and in vivo. Biophys. J. 70: A276 1996. (Abstract)
 205. Spach, M. S., R. C. Barr, G. A. Serwer, J. M. Kootsey, and E. A. Johnson. Extracellular potentials related to intracellular action potentials in the dog Purkinje system. Circ. Res. 30: 505–519, 1972.
 206. Splawski, I., M. Tristani‐Firouzi, M. H. Lehmann, M. C. Sanguinetti, and M. T. Keating. Mutations in the hminK gene cause long QT syndrome and suppress IKs function. Nat. Genet. 17: 338–340, 1997.
 207. Sridharan, M. R. and L. G. Horan. Electrocardiographic J wave of hypercalcemia. Am. J. Cardiol. 54: 672–673, 1984.
 208. Stankovicova, T., M. Szilard, I. De Scheerder, and K. R. Sipido. M cells and transmural heterogeneity of action potential configuration in myocytes from the left ventricular wall of the pig heart. Cardiovasc. Res. 45: 952–960, 2000.
 209. Steinhaus, B. M. Estimating cardiac transmembrane activation and recovery times from unipolar and bipolar extracellular electrograms: a simulation study. Circ. Res. 64: 449–462, 1989.
 210. Streeter, D. D. Gross morphology and fiber geometry of the heart. In: Handbook of Physiology. Section 2: The Cardiovascular System, edited by R. M. Berne, Baltimore: Waverly Press, 1979: 61–112.
 211. Streeter, D. D., H. M. Spotnitz, D. P. Patel, J. Ross, and E. H. Sonnenblick. Fiber orientation in the canine left ventricle during diastole and systole. Circ. Res. 24: 339–347, 1969.
 212. Studer, R., H. Reinecke, J. Bilger, T. Eschenhagen, M. Bohm, G. Hasenfuss, H. Just, J. Holtz, and H. Drexler. Gene expression of the cardiac Na(+)‐Ca2+ exchanger in end‐stage human heart failure. Circ. Res 75: 443–453, 1994.
 213. Sun, Z. Q., G. T. Eddlestone, and C. Antzelevitch. Ionic mechanisms underlying the effects of sodium pentobarbital to diminish transmural dispersion of repolarization. Pacing Clin. Electrophysiol. 20: 11–1116, 1997. (Abstract)
 214. Surawicz, B. Electrophysiologic substrate of torsade de pointes: dispersion of repolarization or early afterdepolarizations? J. Am. Coll. Cardiol. 14: 172–184, 1989.
 215. Sweadner, K. J. Isozymes of the Na+/K+‐ATPase. Biochim. Biophys. Acta 988: 185–220, 1989.
 216. Takagishi, Y., S. Rothery, J. Issberner, A. Levi, and N. J. Severs. Spatial distribution of dihydropyridine receptors in the plasma membrane of guinea pig cardiac myocytes investigated by correlative confocal microscopy and label‐fracture electron microscopy. J. Electron. Microsc. (Tokyo) 46: 165–170, 2000.
 217. Takano, M. and A. Noma. Distribution of the isoprenalineinduced chloride current in rabbit heart. Pflugers Arch. 420: 223–226, 1992.
 218. Takimoto, K., D. Li, J. M. Nerbonne, and E. S. Levitan. Distribution, splicing and glucocorticoid‐induced expression of cardiac alpha 1C and alpha 1D voltage‐gated Ca2+ channel mRNAs. J. Mol. Cell. Cardiol. 29: 3035–3042, 1997.
 219. Thompson, R., J. Rich, F. Chmelik, and W. L. Nelson. Evolutionary changes in the electrocardiogram of severe progressive hypothermia. J. Electrocardiol. 10: 67–70, 1977.
 220. Toshe, N., N. Haruaki, and M. Kanno. α 1‐Adrenoceptor stimulation enhances the delayed rectifier K+ current of guinea pig ventricular cells through the activation of protein kinase C. Circ. Res. 71: 1441–1446, 1992.
 221. Trautwein, W. and M. Kameyama. Intracellular control of calcium and potassium currents in cardiac cells. Jpn. Heart J. 27 (Suppl 1): 31–50, 1986.
 222. Trudeau, M. C., J. W. Warmke, B. Ganetzky, and G. A. Robertson. HERG, a human inward rectifier in the voltage‐gated potassium channel family. Science 269: 92–95, 1995.
 223. van Kempen, M. J., C. Fromaget, D. Gros, A. F. Moorman, and W. H. Lamers. Spatial distribution of connexin43, the major cardiac gap junction protein, in the developing and adult rat heart. Circ. Res 68: 1638–1651, 1991.
 224. Verheule, S., M. J. van Kempen, P. H. te Welscher, B. R. Kwak, and H. J. Jongsma. Characterization of gap junction channels in adult rabbit atrial and ventricular myocardium. Circ. Res 80: 673–681, 1997.
 225. Viskin, S., S. R. Alla, H. V. Barron, K. Heller, L. A. Saxon, I. Kitzis, G. F. Hare, M. J. Wong, M. D. Lesh, and M. M. Scheinman. Mode of onset of torsade de pointes in congenital long QT syndrome. J. Am. Coll. Cardiol. 28: 1262–1268, 1996.
 226. Viswanathan, P. C., R. M. Shaw, and Y. Rudy. Effects of IKr and IKs heterogeneity on action potential duration and its ratedependence: a simulation study. Circulation 99: 2466–2474, 1999.
 227. Volders, P. G., K. R. Sipido, E. Carmeliet, R. L. Spatjens, H. J. Wellens, and M. A. Vos. Repolarizing K+ currents ITO1 and IKs are larger in right than left canine ventricular midmyocardium. Circulation 99: 206–210, 1999.
 228. Vos, M. A., S. C. Verduyn, A. P. M. Gorgels, G. C. Lipcsei, and H. J. Wellens. Reproducible induction of early afterdepolarizations and torsade de pointes arrhythmias by d‐sotalol and pacing in dogs with chronic atrioventricular block. Circulation 91: 864–872, 1995.
 229. Wang, Q., M. E. Curran, I. Splawski, T. C. Burn, J. M. Millholland, T. J. Van Raay, J. Shen, K. W. Timothy, G. M. Vincent, T. De Jager, P. J. Schwartz, J. A. Towbin, A. J. Moss, D. L. Atkinson, G. M. Landes, T. D. Connors, and M. T. Keating. Positional cloning of a novel potassium channel gene: KVLQT1 mutations cause cardiac arrhythmias. Nat. Genet. 12: 17–23, 1996.
 230. Wang, Q., J. Shen, I. Splawski, D. L. Atkinson, Z. Z. Li, J. L. Robinson, A. J. Moss, J. A. Towbin, and M. T. Keating. SCN5A mutations associated with an inherited cardiac arrhythmia, long QT syndrome. Cell 80: 805–811, 1995.
 231. Wang, Z. G., B. Fermini, and S. Nattel. Sustained depolarization‐induced outward current in human atrial myocytes: evidence for a novel delayed rectifier K+ current similar to Kv1.5 cloned channel currents. Circ. Res. 73: 1061–1076, 1993.
 232. Watanabe, Y. Purkinje repolarization as a possible cause of the U wave in the electrocardiogram. Circulation 51: 1030–1037, 1975.
 233. Weirich, J., Bernhardt, R., Loewen, N., Wenzel, W., and Antoni, H. Regional‐ and species‐dependent effects of K+‐channel blocking agents on subendocardium and mid‐wall slices of human, rabbit and guinea pig myocardium. Pflugers Arch. 431: R 130, 1996. (Abstract)
 234. Weissenburger, J., J. M. Davy, and F. Chezalviel. Experimental models of torsades de pointes. Fundam. Clin. Pharmacol. 7: 29–38, 1993.
 235. Weissenburger, J., J. M. Davy, F. Chezalviel, O. Ertzbischoff, J. M. Poirier, F. Engel, P. Lainee, E. Penin, G. Motte, and G. Cheymol. Arrhythmogenic activities of antiarrhythmic drugs in conscious hypokalemic dogs with atrioventricular block: comparison between quinidine, lidocaine, flecainide, propranolol and sotalol. J. Pharmacol. Exp. Ther. 259: 871–883, 1991.
 236. Weissenburger, J., V. V. Nesterenko, and C. Antzelevitch. Transmural heterogeneity of ventricular repolarization under baseline and long QT conditions in the canine heart in vivo. Torsades de pointes develops with halothane but not pentobarbital anesthesia. J. Cardiovasc. Electrophysiol. 1: 290–304, 2000.
 237. West, T. C., E. L. Frederickson, and D. W. Amory. Single fiber recording of the ventricular response to induced hypothermia in the anesthetized dog. Correlation with multicellular parameters. Circ. Res. 7: 880–888, 1959.
 238. Wettwer, E., G. J. Amos, H. Posival, and U. Ravens. Transient outward current in human ventricular myocytes of subepicardial and subendocardial origin. Circ. Res. 75: 473–482, 1994.
 239. Wilde, A. A. M., R. J. E. Jongbloed, P. A. Doevendans, D. R. Düren, R. N. W. Hauer, I. M. Van Langen, J. P. Van Tintelen, H. J. M. Smeets, H. Meyer, and J. L. M. C. Geelen. Auditory stimuli as a trigger for arrhythmic events differentiate HERG‐related (LQTS2) patients from KVLQT1‐related patients (LQTS1). J. Am. Coll. Cardiol. 33: 327–332, 1999.
 240. Wong, K. R., A. E. Trezise, S. Bryant, G. Hart, and J. I. Vandenberg. Molecular and functional distributions of chloride conductances in rabbit ventricle. Am. J. Physiol. 277 (Heart Circ. Physiol. 46): H1403–H1409, 1999.
 241. Xu, H., J. E. Dixon, D. M. Barry, J. S. Trimmer, J. P. Merlie, D. McKinnon, and J. M. Nerbonne. Developmental analysis reveals mismatches in the expression of K+ channel alpha subunits and voltage‐gated K+ channel currents in rat ventricular myocytes. J. Gen. Physiol 108: 405–419, 1996.
 242. Yan, G. X. and C. Antzelevitch. Cellular basis for the electrocardiographic J wave. Circulation 93: 372–379, 1996.
 243. Yan, G. X. and C. Antzelevitch. Cellular basis for the normal T wave and the electrocardiographic manifestations of the long QT syndrome. Circulation 98: 1928–1936, 1998.
 244. Yan, G. X. and C. Antzelevitch. Cellular basis for the Brugada syndrome and other mechanisms of arrhythmogenesis associated with ST segment elevation. Circulation 100: 1660–1666, 1999.
 245. Yan, G. X., W. Shimizu, and C. Antzelevitch. Characteristics and distribution of M cells in arterially perfused canine left ventricular wedge preparations. Circulation 98: 1921–1927, 1998.
 246. Young, R. M. and J. B. Lingrel. Tissue distribution of mRNAs encoding the alpha isoforms and beta subunit of rat Na+,K+‐ATPase. Biochem. Biophys. Res. Commun. 145: 52–58, 1987.
 247. Zahler, R., M. Brines, M. Kashgarian, E. J. Benz, Jr., and M. Gilmore‐Hebert. The cardiac conduction system in the rat expresses the alpha 2 and alpha 3 isoforms of the Na+,K(+)‐ATPase. Proc. Natl. Acad. Sci. U.S.A 89: 99–103, 1992.
 248. Zahler, R., W. Sun, T. Ardito, M. Brines, and M. Kashgarian. The α3 isoform protein of the Na+/K+‐ATPase is associated with the sites of neuromuscular and cardiac impulse transmission. In The Sodium Pump: Structure, Mechanism, Hormonal Control and Its Role in Disease, edited by E. Bamberg and W. Schoner. New York: Springer‐Verlag, 2000: 714–717.
 249. Zahler, R., W. Sun, T. Ardito, and M. Kashgarian. Na‐K‐ATPase alpha‐isoform expression in heart and vascular endothelia: cellular and developmental regulation. Am. J. Physiol. 270 (Cell Physiol. 39): C361–C371, 1996.
 250. Zhang, J. F., A. D. Randall, P. T. Ellinor, W. A. Horne, W. A. Sather, T. Tanabe, T. L. Schwarz, and R. W. Tsien. Distinctive pharmacology and kinetics of cloned neuronal Ca2+ channels and their possible counterparts in mammalian CNS neurons. Neuropharmacology 32: 1075–1088, 1993.
 251. Zhang, L., S. J. Compton, C. Antzelevitch, K. W. Timothy, G. M. Vincent, and J. W. Mason. Differential response of QT and QU intervals to adrenergic stimulation in long QT patients with IKs defects. J. Am. Coll. Cardiol. 33: 138A, 1999. (Abstract)
 252. Zipes, D. P. The long QT interval syndrome: A Rosetta stone for sympathetic related ventricular tachyarrhythmias. Circulation 84: 1414–1419, 1991.
 253. Zygmunt, A. C. Intracellular calcium activates chloride current in canine ventricular myocytes. Am. J. Physiol. 267 (Heart Circ. Physiol. 36): H1984–H1995, 1994.
 254. Zygmunt, A. C. and W. R. Gibbons. Calcium‐activated chloride current in rabbit ventricular myocytes. Circ. Res. 68: 424–437, 1991.
 255. Zygmunt, A. C. and W. R. Gibbons. Properties of the calciumactivated chloride current in heart. J. Gen. Physiol. 99: 391–414, 1992.
 256. Zygmunt, A. C., R. J. Goodrow, and C. Antzelevitch. INa‐Ca contributes to electrical heterogeneity within the canine ventricle. Am. J. Physiol. Heart Circ. Physiol. 278: H1671–H1678, 2000.
 257. Zygmunt, A. C., G. T. Eddlestone, G. P. Thomas, V. V. Nesterenko, C. Antzelevitch. Larger late sodium conductance in M‐cells contributes to electrical heterogeneity in canine ventricle. Am. J. Physiol. Heart Circ. Physiol., In press, 2001.

Contact Editor

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

* Required Field

How to Cite

Charles Antzelevitch, Robert Dumaine. Electrical Heterogeneity in the Heart: Physiological, Pharmacological and Clinical Implications. Compr Physiol 2011, Supplement 6: Handbook of Physiology, The Cardiovascular System, The Heart: 654-692. First published in print 2002. doi: 10.1002/cphy.cp020117