Comprehensive Physiology Wiley Online Library

Normal and Abnormal Conduction in the Heart

Full Article on Wiley Online Library



Abstract

The sections in this article are:

1 Basic Mechanisms of Cardiac Impulse Propagation
1.1 The Continuous Cable
1.2 Two‐Dimensional Propagation and Wavefront Curvature
1.3 Structural Determinants of Anisotropic and Discontinuous Conduction
1.4 Propagation in Discontinuous Structures
1.5 Discontinuous Propagation in a Cell Chain and a Cell Strand
1.6 Anisotropic Propagation
1.7 The Electrical Resistance of the Extracellular Space
1.8 The Bidomain Behavior of Cardiac Tissue
2 The Activation of the Whole Heart—from the Sinus Node to the Ventricles
2.1 The Sinus Node—Spread of Excitation from the Sinus Node to the Atrium
3 Disturbances of Impulse Conduction and Conduction Block
3.1 The Safety Factor of Propagation
3.2 Effects of Changes in Resting Membrane Potential and Inhibition of Na+ Channels on Conduction Velocity
3.3 Conduction Slowing and Block: The Role of Ca++ Inward Current
3.4 Conduction Slowing and Discontinuous Conduction
3.5 Mechanisms of Unidirectional Block
4 Circulating Excitation, Re‐Entry, and Spiral Waves
4.1 Anatomic Re‐entry
4.2 Functional Re‐entry—the Leading Circle Concept
4.3 Spiral Wave Re‐entry
4.4 Transition from Functional to Anatomic Re‐entry. Anchoring of Spiral Waves
Figure 1. Figure 1.

Schematic presentation of electrical propagation. The scheme depicts an excitable cylindrical structure conducting the action potential from left to right at a velocity of 0.5 m/sec. The change in membrane potential along the axis of the cylinder corresponding to the action potential upstroke is plotted above the cylinder. The inside of the cylinder is negatively charged at its resting potential. The inside of the excited segment is charged positively. This potential difference drives the axial or local circuit current, as symbolized by the closed loop. The local circuit current depolarizes the membrane to the threshold for excitation at the site marked with an asterisk. In such a way a new segment of the membrane gets excited and excitation propagates from left to right.

Figure 2. Figure 2.

Electrical cable. Top: Cylindrical structure of cell membrane enveloping the intracellular medium. Point P marks the site of current injection, as explained in bottom panel. Middle: Equivalent electrical circuit. The extra‐ and intracellular spaces are represented by the resistances ro and ri, respectively. The membrane is represented by a parallel circuit of membrane capacitance, cm, and membrane resistance, rm. Bottom: Decrease of relative membrane voltage, V/Vo, during injection of intracellular current in a cable of infinite length. The voltage drops exponentially from the site of current injection at point P (X = 0), from the initial value Vo. The distance on the abscissa is given in the relative unit X, which corresponds to the distance x scaled by the space constant λ (X = x/λ).

Figure 3. Figure 3.

Relation between the change in transmembrane potential, VM, flow of ionic current, Iion, flow of membrane current, IM, and axial or local circuit current, IA, in a continuous linear structure. The cell membrane is symbolized by a parallel circuit consisting of a capacitance and a changing resistance corresponding to a time‐ and voltage‐dependent ionic conductance. The cell interior is symbolized by an internal resistance. Simulation using the Luo‐Rudy model 7. Note that there is axial or local circuit current flow during the early phase of the action potential, which provides the transmembrane current for excitation, IM. Once the threshold is reached and Na+ channels are activated, the Na+ inward current contributes to axial current (see text).

Figure 4. Figure 4.

Schematic presentation of the effect of wavefront curvature on conduction. Left: A flat wavefront propagates at a basic velocity θ0. Arrows denote direction of flow of local circuit current. Middle: Convex wavefront with dispersion of local current, resulting velocity θ is smaller than θ0. Right: Concave wavefront with conversion of local current, resulting velocity θ is larger than θ0.

Figure 5. Figure 5.

Effect of point stimulation (left panel) versus linear stimulation (right panel) on activation spread. Stimulation with a single electrode (point stimulation) produces a convex excitation front. Stimulation with a line of electrodes (line stimulation) produces an almost flat excitation front. Numbers denote activation times in milliseconds relative to the earliest activation. The interval between isochrones is 3 msec. Average longitudinal velocity of curved wave is 13% lower than of flat wave.

Reproduced from reference 39 with permission
Figure 6. Figure 6.

Effect of the radius of a circular stimulation electrode on current threshold (panel A) and stimulus energy (panel B): Epicardial stimulation of the canine heart. At an electrode size below 0.1–0.4 mm, the current threshold is independent of electrode size; above this radius, which corresponds approximately to the radius of the liminal area, current threshold increases with increasing electrode size. The stimulus energy is lowest at the electrode radius which corresponds to the radius of the liminal area.

Reproduced from references 43 and 45 with permission
Figure 7. Figure 7.

Isolated myocyte: Micrograph of immunostained, paraformaldehyde‐fixed disaggregated canine myocyte. Immunostaining of connexin43 reveals a pattern that conforms precisely to the distribution of intercellular gap junctions.

Reproduced from reference 49 with permission
Figure 8. Figure 8.

Laminar organization of ventricular myocardium. Micrographs of tangential surface of a ventricular specimen showing layered organization of myocytes, branching of layers (arrow) and collagen fibers between adjacent sheets.

Reproduced from reference 25 with permission
Figure 9. Figure 9.

Simulation of the effect of wavefront collision on the upstroke of the transmembrane action potential and the Na+ inward current. The values computed during uniform conduction (solid lines) are compared to the values computed at a collision site (dashed lines). Left top: Change of membrane potential, VM, during action potential upstroke. Left bottom: maximal upstroke velocity of transmembrane action potential in Volts/sec. Right top: Na+ inward current, INa. Right bottom: time course of Na+ conductance, gNa.

Reproduced from reference 20 with permission
Figure 10. Figure 10.

Simulation of the effect of wavefront dispersion on the upstroke of the transmembrane action potential and the Na+ inward current. A: Inset shows simulated two‐dimensional strand of excitable tissue emerging into a large area. Signals on panels A–D are simulated from sites 1–11 shown on the inset. Action potential upstrokes show a double component, which is most prominent at the expansion site. B: First time derivatives dV/dt from action potential upstrokes shown on panel A. C: Time course of Na+ conductance, gNa. D: Time course of Na+ inward current, INa. Note increase of INa at expansion site, associated with a decrease of dV/dtmax.

Reproduced from reference 72 with permission
Figure 11. Figure 11.

Effects of resistive discontinuities on propagation. Top: A row of simulated excitable elements (abscissa denotes element number) is separated by resistors. A number N of elements is connected by resistors of low value (200 Ω cm). Each group of N elements is connected to the next group by a single resistor, , of high value. Discontinuity at a constant value of effective longitudinal resistance can be changed by the simultaneous increase of N and . Bottom: Propagation along the simulated row of excitable elements, as illustrated by the time course of dV/dtmax (upper trace) and the action potential upstroke (lower trace). The degree of discontinuity is increased from panel A to C, while the value of effective or total longitudinal resistance is kept constant. Note increasing delay between the two action potential upstrokes, and the discontinuous upstroke in C.

Reproduced from reference 69 with permission
Figure 12. Figure 12.

Effects of resistive discontinuities on conduction velocity, θ. Propagation velocity is simulated in the model shown in the upper panel of Figure 11 as a function of the overall or effective resistance Ri (expressed as a fraction of the low value resistance of 200 Ωcm shown in FIG. 11, R/200) The solid line depicts the decrease of θ in a continuous cable where θ2 ∼ Ri. In curve A, the value for the high resistor, , is 5000 Ωcm, the numbers on the curve denote the number of elements N. In curve B, the degree of discontinuity is higher, because is 10,000 Ωcm. Note that in curve B, θ decreases above N = 16 and conduction block occurs when N >26 (see text).

Reproduced from reference 69 with permission
Figure 13. Figure 13.

Effects of resistive discontinuities on the maximal upstroke velocity of the transmembrane action potential, dV/dtmax (simulated in the model shown in the upper panel of FIG. 11). As a control, the dashed lines depict the dV/dtmax values for continuous cables (upper line Ri = 200 Ωcm, lower line R, = 4200 Ωcm). In all solid curves shown, the value of the high resistor is set to = 4200 Ωcm, and the curves differ with respect to their numbers of elements N. The curves are shown for N = 5, N = 9, and N = 51. With N = 51, there is dispersion of local current beyond the first resistive obstacle with a decrease of dV/dtmax, and collision before the next resistive obstacle (from N 40 to 51) with an increase in dV/dtmax. With a small number of N 5, the effect of dispersion and collision is minor and all of the dV/dtmax values are above the upper dashed line. This is due to the almost simultaneous excitation of all elements, similar to the situation of a space‐clamped action potential. N = 9 corresponds to an intermediate situation.

Reproduced from reference 69 with permission
Figure 14. Figure 14.

Simulated cell chain. Comparison of continuous with discontinuous conduction with decreasing cell‐to‐cell coupling. Effects of variations in axial (longitudinal) resistance on microscopic velocity (θmic, curve 1) and on average macroscopic velocity (θmac, curve 2). The microscopic velocity corresponds to the velocity inside a cell. The case of a continuous structure is shown for comparison on curve 3 and follows the inverse square root relation of continuous cable theory. The effective longitudinal resistance is changed by varying the disk resistance while the myoplasm resistance is kept constant at 200. Both effective longitudinal resistivity and the corresponding disk resistance are indicated.

Reproduced from reference 26 with permission
Figure 15. Figure 15.

Simulated chain of cells. Changes of maximal upstroke velocity with decreasing cell‐to‐cell coupling. Conduction velocity θ (dashed line) and dV/dtmax (solid line) are plotted as a function of the increasing effective resistance or the specific disk resistance (coupling resistance between simulated cells). There is a transient increase of dV/dtmax with increasing cell‐to‐cell uncoupling and decreasing velocity of propagation.

Reproduced from reference 26 with permission
Figure 16. Figure 16.

Experimental determination of conduction in a single cell chain. A: Reproduction of the microscopic appearance of a cultured cell chain. Dots and numbers denote positions of three light‐sensitive diodes (6.5 μm in diameter) separated by a distance of 30 μm. B: Optical recording of action potential upstrokes from diodes 1–3, measured as fluorescence change ΔF/F of a voltage‐sensitive dye (upper traces) and the first time derivatives (lower traces). Numbers 1–3 denote times of local activation. Note that the conduction delay across the cell border 2,3 is larger than delay within the cytoplasm 1,2. C: Histograms of cytoplasmic (upper graph) and junctional (lower graph) conduction times. The difference between the mean conduction times amounts to approximately 80 μsec and reflects the mean conduction time across the end‐to‐end cell junctions.

Reproduced from reference 83 with permission
Figure 17. Figure 17.

Experimental determination of conduction in a cell strand. A: Reproduction of the microscopic appearance of a cultured cell strand (4–5 cells in width). Dots and numbers denote positions of three light‐sensitive diodes (6.5 μm in diameter) separated by a distance of 30 μm. B: Optical recording of action potential upstrokes from diodes 1–3, measured as fluorescence change ΔF/F of a voltage‐sensitive dye (upper traces) and the first time derivatives (lower traces). Numbers 1–3 denote times of local activation. C: Histograms of cytoplasmic (upper graph) and junctional (lower graph) conduction times. With respect to the measurements shown in Figure 16 the mean cytoplasmic conduction time has increased from 38 μsec to 60 μsec and the mean junctional conduction time has decreased from 118 μsec to 80 μsec. This is explained by electrotonic current flow through lateral gap junctions (“ lateral averaging”; see text).

Reproduced from reference 83 with permission
Figure 18. Figure 18.

Propagation in a cellular network, simulation of intracellular excitation sequences. A, longitudinal conduction: Conduction from left to right is depicted on the upper trace, conduction from right to left is depicted on the lower trace. Isochrone lines are separated by 4 μsec. B, transverse conduction: Conduction from top to bottom is shown on the upper trace, propagation from bottom to top is depicted on the lower trace. Intracellular isochrones are separated by 3 μsec. During longitudinal propagation, there is a crowding of isochrones (slow propagation) at the beginning of propagation in the individual cells and acceleration of propagation toward the end of the cells. During transverse propagation, the arrows indicate preferential longitudinal propagation spread, with microcollisions occurring occasionally (asterisk).

Reproduced from reference 29 with permission
Figure 19. Figure 19.

Simulation of intracellular excitation sequences (A), intracellular distributions of dV/dtmax (B) and inward Na+ charge movements during excitation (C) in an anisotropic cellular network. The left graphs correspond to longitudinal propagation from left to right and the right graphs correspond to transverse propagation from top to bottom. Note the close correspondence between the isochrone spacing, dV/dtmax and inward Na+ charge movement during both transverse and longitudinal propagation: The dV/dtmax is relatively low where excitation is slow and vice versa. By contrast the locations of slow activation and low dV/dtmax correspond to large inward Na+ charge movements and vice versa.

Reproduced from reference 29 with permission
Figure 20. Figure 20.

Ratio of extracellular to intracellular resistance in compact ventricular tissue. The amplitude of the bipolar extracellular electrogram (upper trace, VE) and the amplitude of the action potential upstroke (lower trace, VM) are shown from an isolated, arterially‐perfused papillary muscle. The signals are measured in a muscle which is surrounded by an electrical insulator. The ratio of VE/(VM − VE) which is approximately 1, corresponds to the ratio of extracellular: intracellular resistance, rori. This demonstrates that the extracellular resistance in compact ventricular tissue is of approximately the same magnitude is the resistance of the intracellular space (including the gap junctions).

Reproduced from reference 32 with permission
Figure 21. Figure 21.

Effect of superfusion fluid on conduction velocity, θ, the maximal upstroke velocity of the transmembrane action potential dV/dtmax, and the time constant of the initial rise of the action potential, τfoot. Panels (a) and (b) show action potentials measured at two sites along a cylindrical papillary muscle, between electrodes D and C, (D‐C), and between electrodes A and B, (A‐B). The bipolar extracellular electrogram is measured between the extracellular electrodes c and b, (b‐c). The muscle is either soaked in a large bulk solution (SF, closed circles) or covered only by a thin fluid layer (T, open circles). Panel (b) illustrates the curved wavefronts measured in the presence of the large bulk solution during propagation from left to right. Traces on the right correspond from top to bottom: τa‐b, dV/dtmax, a‐b, τd‐c dV/dtmax d‐c, θ0 (conduction velocity at the surface) and θi (conduction velocity in the core of the fiber). Increasing the thickness of the fluid layer (electrical shunting, transition from T to SF) produces (1) an increase of θ0 and θi; (2) an increase of τa‐b and τd‐c; (3) a decrease of dV/dtmax.

Reproduced from reference 107 with permission
Figure 22. Figure 22.

Three‐dimensional plot of simulated transmembrane voltage (Φm) after application of a point stimulus to two‐dimensional anisotropic tissue. The X‐axis corresponds to the longitudinal direction of the fibers, the Y‐axis to the transversal direction of the fibers. Panel A: Simulation of anisotropy with equal ratios of extracellular to intracellular conductivities. Panel B: Longitudinal conductivity ratio (σexix) 8 × 10−4:2 × 10−4; conductivity ratio (σeyiy) 2 × 10−4:2 × 10−5 Note that with an equal anisotropic ratio there is a drop of Φm with distance from current injection and a elliptical shape of Φm distribution in the x/y plane. In Panel B, which corresponds to the simulation using experimentally determined values of intra‐ and extracellular conductivities, there is a hyperpolarization of Φm in the × (longitudinal) and a depolarization in the Y (transverse) directions.

Reproduced from reference 112 with permission
Figure 23. Figure 23.

Dog‐bone shape of virtual electrode. Plots showing the shape of virtual electrodes caused by stimuli from 1–7 mA. The horizontal axis corresponds to the longitudinal direction of the anisotropic subepicardial layer of a dog, the vertical axis corresponds to the transverse fiber direction. Note that upon application of a central cathodal stimulus, the line where excitation starts to propagate, i.e. the virtual electrode, has a dog bone shape with a very large extension of the electrode in the transverse direction.

Reproduced from reference 113 with permission
Figure 24. Figure 24.

Interaction between an anisotropic medium and a bathing solution. A: Deviation of the isochrones of a wave propagating in the longitudinal direction (left side) and a wave propagating in the transverse direction (right side). Note that due to the anisotropy‐dependent differences in intra‐ and extracellular conductivities, wavefront bending in the longitudinal direction is significantly more expressed. In the absence of a bulk conductor, both longitudinal and transverse wavefronts are flat (not shown). B: Phase plane plots of dV/dtmax versus membrane potential in a model with unequal anisotropy. Superimposed are phase plane plots of action potentials propagating in the longitudinal direction (L), at angles of 30°, 45°, and 60° from the longitudinal direction and in the transverse direction (T). On the left hand side, the bathing fluid was absent. Note that all the traces almost superimpose. On the right hand side, a bathing fluid has been added to the boundary of the tissue. In this case there is a marked direction dependence of both the initial portion of the action potential of and of dV/dtmax.

Reproduced from references 109,120 with permission
Figure 25. Figure 25.

Isochronal map of excitation spread from the sinoatrial node. Tones depict excitation intervals in steps of 5 msec, numbers correspond to activation times in msec. Configuration of action potentials is shown along the pathway of conduction from the sinus node to the atrium. The dashed line indicates the beginning of the atrial electrogram used as time reference. Toward the periphery, action potentials show an increase in amplitude and dV/dtmax and a decrease in rate of diastolic depolarization. The area in which two component action potentials were recorded is hatched in the activation map.

Reproduced with permission from reference 126
Figure 26. Figure 26.

Spread of excitation in the right atrium. Isochronal map of the spread of excitation from the sinoatrial node (S. A. N.) over the epicardial surface of the right atrium made by Thomas Lewis in 1915. Although in general spread of activation is depicted as being radial, the isochrones deviate over the crista terminalis, indicating preferential conduction in that region.

Reproduced with permission from reference 448
Figure 27. Figure 27.

Anatomy of the rabbit right atrium. The contrast of the preparation has been enhanced by supravital staining with methylene blue: ct, cut ends of the crista terminalis; ra, right auricle; fo, fossa ovalis; SVC, superior vena cava; ivc, inferior vena cava; ocs, ostium of the coronary sinus; vc, valve of the coronary sinus; ss, sinus septum; asl, attachment of the septal tricuspid leaflet (marked by a dotted line); SA, sinoatrial node (the letters inside surrounded by a dotted line indicate the approximate area of the compact node consisting of typical nodal cells); CN, cauda of the sinoatrial node; rb, right branch of the sinoatrial right bundle; lb, left (septal) branch of the sinoatrial ring bundle; tc, transitional cell zone of the atrioventricular node; mc, midnodal cells; lc, lower nodal cells; avb, atrioventricular bundle; ao, atrial overlay fibers.

Reproduced with permission from reference 449
Figure 28. Figure 28.

Activation sequence in the right atrium. A: A photograph of the right atrium of a rabbit: the activation sequence has been mapped using a tenfold microelectrode assembly, with which 280 different atrial cells have been impaled during the course of the experiment. The specimen was opened by an incision along the lateral margin of the tricuspid valve, and pinned out in order not to interrupt the internodal tissues. After the experiment, the preparation was fixed before being photographed. The specimen is illuminated from behind to illustrate the thick and thin parts of the myocardium. B: A diagrammatic representation, in which the abbreviations are as follows: SVC, superior vena cava; IVC, inferior vena cava; FO, fossa ovalis; CS, coronary sinus; AVN, atrioventricular node; MS, membranous septum; IVS, interventricular septum; C and D: Activation maps made according to photographs taken while the preparation was in the tissue bath—hence the slight differences in shape when compared to the fixed specimen. The preparation was beating spontaneously for the construction of (C); in (D), the preparation was paced through an electrode placed above the ostium of the SVC. In both instances, the activation sequence, indicated by isochrone lines separating areas activated within 5 msec intervals, follows the thicker muscle bundles. These are the crista terminalis, the septal branch of the crista, and the thick muscle ridge between IVC and FO. The AV node receives a dual input. The localization of the pacemaker determines the dominant input. In C, during spontaneous sinus rhythm, the AV node is reached earliest by the posterior route. In D during driving from the SVC, the anterior limbus activates the AV node earlier.

Reproduced with permission from reference 176
Figure 29. Figure 29.

Anatomy of the AV junction. A: Photograph and B sketch of a normal human heart showing the anatomical landmarks of the triangle of Koch. The approximate site of the compact AV node is indicated by the stippled area adjacent to the central fibrous body.

Reproduced with permission from reference 183
Figure 30. Figure 30.

Propagation toward the AV node. Arrows indicate the four areas from which atrial fibers approach the specialized AV junctional area. The fourth area, indicated by the curved arrow, is from the left atrial aspect of the septum.

Reproduced with permission from reference 183
Figure 31. Figure 31.

Action potential characteristics in the AV node: Action potentials of six types of AV nodal cells during periodic premature stimulation of the right atrium. Each section was obtained by superimposing (in decreasing order of coupling stimulation intervals [numbers at left in msec]) tracings corresponding to last basic and premature beat. Baseline of each subsequent tracing was shifted downward to help distinguish potentials. Action potential after premature potential in lower trace in AN (atrionodal) and H (His) was caused by an atrial re‐entrant beat. Note double components in N (nodal) cell of early premature responses. ANL, late AN cells; ANCO, AN cells with action potential upstroke with two components.

Reproduced with permission from reference 193
Figure 32. Figure 32.

Activation of the AV node. Map showing sequence of normal antegrade conduction of rabbit AV node. Symbols indicate position of AV nodal cells from which action potentials were recorded and also in which 20 msec interval these cells were activated. Note dual input into AV node. CT, crista terminalis; IAS, interatrial septum; CS, ostium of coronary sinus; Tr. V., tricuspid valve; H, position of extracellular electrode on His bundle.

Reproduced with permission from reference 200
Figure 33. Figure 33.

Conduction in the AV node. Two simultaneously recorded transmembrane potentials from the AV node of a Langendorff blood‐perfused canine heart at a superficial and a deep site at the same location. The atrium was paced at a basic cycle length of 600 msec, and a premature stimulus S2 was applied at a coupling interval of 300 msec, either at the posterior input (“ slow pathway,” upper panel) or at the anterior input (“ fast pathway,” lower panel). Note double components, especially during premature stimulation, where the action potential of the superficial cell causes a slow prepotential in the deeper cell which during posterior stimulation is large enough to cause an action potential that is propagated to the His bundle (not shown), but that fails to reach threshold during antegrade stimulation.

Retraced from unpublished recordings by J. M. T. de Bakker
Figure 34. Figure 34.

Abnormal Wenckebach phenomenon. Three simultaneously recorded action potentials in posterior input (cell a), in anterior input (cell b), and in junctional area of these two inputs (cell c). Note the difference in timing and configuration of the action potentials of cells a and c and of the His bundle complex during the first and fifth beats. Double bars indicate block. Numbers are activation times in msec. Atr., recording electrode on crista terminalis from which the electrogram in the upper trace was recorded. His, position of electrode recording electrogram of His bundle (lower trace).

Reproduced with permission from reference 200
Figure 35. Figure 35.

Cycle length dependence of AV‐nodal conduction. Action potentials illustrating dependency of first and second component in N cells upon late AN and early NH potentials. Signals 1 and 2 were recorded from AN cells, signals 3, 4, and 5 from N cells, and signal 6 from an NH cell. Inset shows position of cells. First component is largest in N cells close to AN zone; second component is largest in cells close to NH cells. Note that second component occurs later than upstrokes of action potentials in cell 6. Note also that duration of prepotential in cell 6 increases progressively in successive activation and that level at which prepotential breaks into a fast upstroke remains constant in all cycle lengths. Cells 4, 5, and 6 were recorded simultaneously. Cells 1, 2, and 3 were recorded separately, approximately 3 min earlier.

Reproduced with permission from reference 199
Figure 36. Figure 36.

Cycle length dependence of AV nodal conduction. A: Classical ladder diagram used in electrocardiography to depict cycle length‐dependent conduction delay in the AV junction, in this case during a 4:3 Wenkenbach phenomenon. B: Modification of the ladder diagram to express saltatory nature of the cycle length‐dependent conduction delay.

Figure 37. Figure 37.

Division of the left bundle branch. An illustration of the variation in the structure of the divisions of the left bundle branch in 20 different hearts.

Reproduced with permission from reference 259
Figure 38. Figure 38.

Propagation across the Purkinje–muscle junction. Intracellular and extracellular recordings at Purkinje–ventricular muscle junctional sites. A: Action potential of a Purkinje fiber (Pi), coinciding with the Purkinje (P) deflection preceding the all‐negative ventricular muscle (V) deflection in the extracellular electrogram (e). s, Stimulus artifact. B: Action potential of a transitional cell (Ti), with an early slow component (arrow). C: Simultaneous intracellular recordings of an early (Ti) and a late (T2i) transitional cell. The latter could also be classified as an early ventricular cell that is electrotonically influenced from the transitional cells, giving rise to the long, slow foot (arrow). D: Simultaneous intracellular recordings of a Purkinje fiber and a transitional cell with a slow foot (arrow) and an inherent low amplitude that is heightened on activation of the ventricular mass. E: Intracellular recording from a transitional cell that coincides with a small deflection in the extracellular electrogram during the Purkinje fiber–ventricular muscle delay period. F: Intracellular recording of a transitional cell with multiple components during the upstroke (arrow). Panels A–C derive from rabbit hearts, and panels D–F derive from pig hearts.

Reproduced with permission from reference 272
Figure 39. Figure 39.

Structure of the Purkinje–muscle junction. Schematic representation of the structure of a rabbit Purkinje fiber–ventricular muscle junction. P, Purkinje fibers; T, transitional cells; V, ventricular myocardium.

Reproduced with permission from reference 272
Figure 40. Figure 40.

Safety factor of propagation. Results of computer simulation of propagation in a cell chain. The cells are separated by a simulated gap junction resistor. A: Safety factor of propagation, SF, as a function of propagation velocity. Dashed line: Change of SF with a decrease of excitability. Solid line: Change of SF with decreasing conductance (increasing resistance) between cells. B: Change of SF as a function of propagation velocity in the absence and presence of ICa,L. Note that very low conduction velocities can only be achieved with flow of I Ca,L. (See text.)

Reproduced with permission from reference 3
Figure 41. Figure 41.

Supernormal excitability and conduction. A, B: Transmembrane action potentials. C, D: Strength interval curves. Time scale is identical for both traces. A: Recording from the His bundle. B: Recording from a Purkinje fiber running freely in a false tendon. C: Recording from a transitional type Purkinje fiber. D: Recording from a ventricular cell. The supernormal phase of excitability in B is associated with a supernormal conduction.

Reproduced with permission from reference 279
Figure 42. Figure 42.

Relationship of excitability to conduction velocity during phase‐4 depolarization. Conduction time along a fixed distance of a Purkinje strand and maximal upstroke velocity of the transmembrane action potential, dV/dtmax of a Purkinje fiber are plotted as a function of the “ take‐off” potential during spontaneous phase 4 depolarization ranging from 92.5 to 105 mV. Note that with ongoing spontaneous depolarization (decrease of “ take‐off” potential”) there is a decrease of conduction time corresponding to an increase in conduction velocity and a decrease of dV/dtmax.

Reproduced with permission from reference 281
Figure 43. Figure 43.

Change of longitudinal and transverse conduction velocity with increasing extracellular potassium concentration, [K+]o. Measurements were made in an isolated perfused porcine heart. Note that propagation blocks at a [K+]o of 11 mM.

Reproduced with permission from reference 287
Figure 44. Figure 44.

Recovery of maximal upstroke velocity of the action potential, dV/dtmax. Action potentials were elicited at different times in the wake of the preceding action potential. Time 0 denotes the beginning of the preceding action potential. The curves depict the recovery curves of dV/dtmax with increasing time measured at different resting potentials. Note that recovery from activation becomes delayed with depolarization.

Reproduced with permission from reference 255
Figure 45. Figure 45.

Unidirectional block in the wake of the preceding wavefront. Transmembrane potentials (Vm) and sodium channel conductance (gNa) computed after application of a premature stimulus (cell 1) in the wake of propagating action potential. In the antegrade direction (right panel) recordings were taken from cells 0.5 mm apart. In the retrograde direction (left panel) traces were recorded from cells 1 mm apart. At the time of premature stimulation, membrane excitability at cell 1 was less than 10% of the maximum excitability (compare gNa curves 1 and 5 on the left panel). In the retrograde direction, the action potential propagated a distance of 4 mm before reaching the region of fully excitable membrane. In the antegrade direction, membrane excitability gradually decreased and propagation extinguished. Note different gNa scales in the left and right panels.

Reproduced with permission from reference 312
Figure 46. Figure 46.

Unidirectional block with asymmetric depression of excitability. Top: Injury is produced by a crushing probe. The line spacing on the Purkinje fiber indicates increasing degree of injury. Bottom: The influence of injury on the excitability threshold is illustrated. The amplitudes of the anterograde wavefront (C‐x‐B) are compared to the amplitudes of the retrograde wavefront (A‐x‐C). At y, the transition between normal cells and inexcitable, injured cells is abrupt. C represents decremental or augmental conduction, depending on direction, through a transitional zone of partial injury; x represents the point of transition between partially excitable cells and inexcitable cells (x‐y). A and B represent electrotonic transmission through inexcitable cells. The retrograde wave succeeds in conducting across, and the anterograde wave front fails.

Reproduced with permission from reference 322
Figure 47. Figure 47.

Unidirectional block at a geometrical tissue expansion: The role of Ca++‐inward current. A: Schematic representation of a cultured cell monolayer with geometrical expansion and overlaid photodiodes during antegrade (upper panel) and retrograde (lower panel) conduction. B and C: Action potential upstrokes recorded using a voltage‐sensitive dye in control conditions (B) and after administration of 5 μM nifedipine. In control conditions, the antegrade propagation was characterized by biphasic upstrokes and local slowing of conduction at the expansion (B). The blockage of Ca++ current with nifedipine produced antegrade conduction block (C). The retrograde propagation was successful in both cases.

Reproduced with permission from reference 80
Figure 48. Figure 48.

Unidirectional block in anisotropic tissue. Anisotropic conduction time curves obtained in human and canine atrial bundles. A: Uniform anisotropic pectinate muscle of a 12‐year‐old child. B: Non‐uniform anisotropic pectinate muscle of a 62‐year‐old man. C: Non‐uniform anisotropic muscle (christa terminalis) of an adult dog. In each preparation conduction times (msec per mm interelectrode distance) were obtained from analyzing the unipolar extracellular electrogram of two electrode pairs. As shown in the inset, the two electrode pairs were placed in longitudinal and transverse directions, respectively. Solid circles represent longitudinal propagation, open circles represent transverse propagation. Each preparation was stimulated at a basic rate, and premature action potentials were introduced at variable intervals, A1–A2. In the uniform anisotropic bundle (A) conduction times became longer with the shortening of the A1–A2 intervals, block occurred in both directions at the same interval. In the non‐uniform cases (B and C), block occurred in the longitudinal direction at a premature interval of 325 and 310 msec, respectively. At this prematurity, transverse propagation was still preserved.

Reproduced from reference 74 with permission
Figure 49. Figure 49.

Effect of microscopic resistive barriers on propagation. A: A phase‐contrast image of a cell culture (neonatal rat myocytes) with the overlaid diode array. Action potential upstrokes are measured at each diode location. The numbers 1–10 on the diode array correspond to the locations of the signals shown in D and E. In D and C, the location of these signals is indicated by the gray area. Two clefts in the central area (outlined in white in A) form an narrow isthmus of 40 μm. Activation maps of longitudinal and transverse conduction are shown in B and C respectively. Note slowing and deviation of the wavefront at the isthmus. Numbers denote separation of isochrones by 100 μsec. Selected recordings of action potential upstrokes during longitudinal and transverse conduction are shown in D and E, respectively. Discontinuities in the action potential upstrokes occur at the expansion site during transverse propagation.

Reproduced from reference 34 with permission
Figure 50. Figure 50.

Unidirectional conduction block at an “ isthmus.” Wave propagation across a narrow tissue isthmus in an isolated ventricular preparation of sheep heart. A: Map of activation spread before an isthmus was produced. B: Activation spread in the same preparation with the isthmus 2.26 mm wide. The isthmus was produced by two tissue cuts (gray zones). C: Activation spread after the isthmus was reduced to 0.88 mm. D: Local conduction velocity measured across the isthmus as a function of isthmus width.

Reproduced with permission from reference 332
Figure 51. Figure 51.

Circus movement re‐entry around a large anatomical obstacle.

Reproduced with permission from reference 350
Figure 52. Figure 52.

Effect of a premature impulse entering a re‐entrant circuit. The black and dotted areas show the absolute and relative refractory periods, respectively. In A and B, a premature impulse enters the circuit at the end of the relative refractory period and spreads in two directions. In the retrograde direction, the premature wave is annihilated by the circulating wave; in the anterograde direction, the premature impulse advances, resetting the tachycardia. In C and D, the premature impulse reaches the circuit closer to the state of absolute refractoriness. The impulse annihilates the retrograde wave and fails to propagate in the anterograde direction thereby terminating the tachycardia.

Reproduced with permission from reference 219
Figure 53. Figure 53.

Initiation of functional re‐entry by premature stimulation in an isolated preparation of rabbit atrial muscle. A: Isochronal activation map of basic beat (interval 500 msec). Dots indicate sites of stimulation. Activation times (msec) are given relative to the stimulus onset. B: Map of premature beat (interval 56 msec). T bars indicate conduction block. C: First cycle of tachycardia. D: Refractory periods measured during basic rhythm (in msec).

Reproduced with permission from reference 318
Figure 54. Figure 54.

Functional re‐entry and tachycardia. Activation map (right) and action potential recordings (left) obtained during steady‐state tachycardia. Cells in the central area of the re‐entrant circuit show double potentials of low amplitude (traces 3 and 4). Lower right: Schematic representation of the activation pattern. Double bars indicate conduction block.

Reproduced with permission from reference 350
Figure 55. Figure 55.

Spiral waves. Spiral waves in chemical Belousov‐Zhabotinsky reaction (A) and in an isolated preparation of canine epicardial muscle (B).

Reproduced with permission from references 366 and 376
Figure 56. Figure 56.

Initiation of spiral wave in a simple model of cardiac excitation. A rectangular area R is excited overlapping the absolutely refractory tail of a propagating wave (A). The premature wave propagates in the retrograde direction (right to left) but blocks in the anterograde direction, forming a wave break that turns around the refractory area R (B). When the area R recovers, the excitation wave short‐circuits this area (C) and forms a spiral wave rotating around a linear line of block (D–F).

Reproduced with permission from reference 373)
Figure 57. Figure 57.

Initiation of a spiral wave by cross‐field stimulation in canine right ventricular myocardium. A: Isochronal maps of activation and repolarization during wave propagation induced by stimulation (S1) from a line of eight epicardial pacing sites. Solid lines depict isochronal activation lines; dashed lines depict isorecovery lines. Numbers indicate time in msec. B: Gradients of extracellular potential (in V/cm) produced by a unipolar cathodal shock (S2) of 150 V from a mesh electrode at the bottom. C: Pattern of activation spread following sequential application of S1 stimulus from the right and S2 shock from the bottom. The S1–S2 interval was 191 msec and the S2 strength was 150 V. Activation times (msec) are measured from the start of the 3 msec S2 shock. The heavy solid line represents the transition between successive activation maps. Isochrones are drawn at 10 msec intervals. The hatched line represents a zone of conduction block. The double‐headed arrow indicates the mean epicardial fiber orientation in the area of conduction block. The hatched area indicates the region assumed to be directly excited by the S2 shock field. Earliest post‐shock activation occurs distant from the S2 site, with no early activation wavefronts conducting away from the directly excited region located between the S2 site and the critical point. A counterclockwise re‐entrant circuit is formed around the region containing the critical point and the block line. The potential gradient equals 5.8 V/cm, and the pre‐shock interval equals 171 msec at the critical point (critical refractory period = 169 msec). D: Schematic representation of the re‐entry initiation by cross‐field stimulation. The row of pacing wires (S1) on the right creates parallel isorecovery lines (R7 through R2), with R7 the least refractory and R2 the most refractory. The S2 from the bottom creates parallel isogradient lines (G7 through G3), with G7 the largest potential gradient and G3 the weakest. The S2 shock produces direct excitation (DE), graded response (GR), or neither effect (NE). Activation fronts propagate from only one part of the directly excited area, not from the directly excited region abutting the area of graded response, thus forming a zone of unidirectional conduction block.

Reproduced with permission from reference 379
Figure 58. Figure 58.

Initiation of a spiral wave at a pivoting point. Formation of a free wave break after wavefront detachment from the sharp edge of an inexcitable obstacle. Computer model with Luo‐Rudy ionic kinetics. The maximal sodium conductance was reduced to 6.6 mS/cm2 A: Isochronal map of activation spread with an interval of 5 msec. B: Snapshot of activation at the moment marked by the asterisk in A. Black indicates the excited area defined by the activation of inward Na+ current. Gray indicates the area in the refractory state as defined by Na+ current inactivation. Point P marks the wave tip, defined as a point where excited, refractory, and resting states meet. The dashed line t shows the trajectory of the wave tip with the radius rp

Reproduced with permission from reference 34
Figure 59. Figure 59.

Dynamics of spiral wave rotation in mathematical models of cardiac excitation. A–C: Cellular automata model. Spiral wave rotation changes from circular (A) to meandering (B), and then to Z‐type (C) with increasing excitability. D–G: FitzHugh‐Nagumo model. The same types of rotation are observed when the wavelength of excitation is increased.

Reproduced with permission from references 450 and 383
Figure 60. Figure 60.

Drift of a spiral wave and the Doppler effect. A and B: Isochronal activation maps showing initiation (A) and the first rotation cycle (B) of a spiral wave in an isolated preparation of epicardial muscle. A stepwise inhomogeneity in refractory period was created by separate superfusion of two parts of the preparation with normal and quinidine‐containing solutions. Dashed line shows the border of inhomogeneity with larger refractoriness in the upper part. The asterisk shows the location of the stimulation electrode. C: Trajectory of the spiral wave tip during initiation (S1) and three subsequent cycles of spiral wave rotation (V1−V3). D: Excitation intervals measured along the border of inhomogeneity during spiral wave drift (cycle V2). Because of the drift, excitation intervals in front of the spiral wave are significantly shorter than intervals behind the spiral wave (Doppler effect).

Reproduced with permission from reference 407
Figure 61. Figure 61.

Anchoring of a spiral wave. A: Electrocardiographic recordings showing that premature stimulation (S2) produced polymorphic arrhythmic activity followed by a transition to sustained monomorphic tachycardia. B: Time‐space plot of activation spread obtained from video‐imaging of transmembrane potential (voltage‐sensitive dye). In these plots, the activity from the whole image is projected onto a single direction (vertical axis) and displayed as a function of time. White bands show a planar wave propagation while branching of bands indicates the presence of a spiral wave induced by the S2 stimulus. As detected from the movement of the branching point, which marks the center of the spiral, the spiral drifted during the first seven cycles and became stationary thereafter.

Reproduced with permission from reference 380


Figure 1.

Schematic presentation of electrical propagation. The scheme depicts an excitable cylindrical structure conducting the action potential from left to right at a velocity of 0.5 m/sec. The change in membrane potential along the axis of the cylinder corresponding to the action potential upstroke is plotted above the cylinder. The inside of the cylinder is negatively charged at its resting potential. The inside of the excited segment is charged positively. This potential difference drives the axial or local circuit current, as symbolized by the closed loop. The local circuit current depolarizes the membrane to the threshold for excitation at the site marked with an asterisk. In such a way a new segment of the membrane gets excited and excitation propagates from left to right.



Figure 2.

Electrical cable. Top: Cylindrical structure of cell membrane enveloping the intracellular medium. Point P marks the site of current injection, as explained in bottom panel. Middle: Equivalent electrical circuit. The extra‐ and intracellular spaces are represented by the resistances ro and ri, respectively. The membrane is represented by a parallel circuit of membrane capacitance, cm, and membrane resistance, rm. Bottom: Decrease of relative membrane voltage, V/Vo, during injection of intracellular current in a cable of infinite length. The voltage drops exponentially from the site of current injection at point P (X = 0), from the initial value Vo. The distance on the abscissa is given in the relative unit X, which corresponds to the distance x scaled by the space constant λ (X = x/λ).



Figure 3.

Relation between the change in transmembrane potential, VM, flow of ionic current, Iion, flow of membrane current, IM, and axial or local circuit current, IA, in a continuous linear structure. The cell membrane is symbolized by a parallel circuit consisting of a capacitance and a changing resistance corresponding to a time‐ and voltage‐dependent ionic conductance. The cell interior is symbolized by an internal resistance. Simulation using the Luo‐Rudy model 7. Note that there is axial or local circuit current flow during the early phase of the action potential, which provides the transmembrane current for excitation, IM. Once the threshold is reached and Na+ channels are activated, the Na+ inward current contributes to axial current (see text).



Figure 4.

Schematic presentation of the effect of wavefront curvature on conduction. Left: A flat wavefront propagates at a basic velocity θ0. Arrows denote direction of flow of local circuit current. Middle: Convex wavefront with dispersion of local current, resulting velocity θ is smaller than θ0. Right: Concave wavefront with conversion of local current, resulting velocity θ is larger than θ0.



Figure 5.

Effect of point stimulation (left panel) versus linear stimulation (right panel) on activation spread. Stimulation with a single electrode (point stimulation) produces a convex excitation front. Stimulation with a line of electrodes (line stimulation) produces an almost flat excitation front. Numbers denote activation times in milliseconds relative to the earliest activation. The interval between isochrones is 3 msec. Average longitudinal velocity of curved wave is 13% lower than of flat wave.

Reproduced from reference 39 with permission


Figure 6.

Effect of the radius of a circular stimulation electrode on current threshold (panel A) and stimulus energy (panel B): Epicardial stimulation of the canine heart. At an electrode size below 0.1–0.4 mm, the current threshold is independent of electrode size; above this radius, which corresponds approximately to the radius of the liminal area, current threshold increases with increasing electrode size. The stimulus energy is lowest at the electrode radius which corresponds to the radius of the liminal area.

Reproduced from references 43 and 45 with permission


Figure 7.

Isolated myocyte: Micrograph of immunostained, paraformaldehyde‐fixed disaggregated canine myocyte. Immunostaining of connexin43 reveals a pattern that conforms precisely to the distribution of intercellular gap junctions.

Reproduced from reference 49 with permission


Figure 8.

Laminar organization of ventricular myocardium. Micrographs of tangential surface of a ventricular specimen showing layered organization of myocytes, branching of layers (arrow) and collagen fibers between adjacent sheets.

Reproduced from reference 25 with permission


Figure 9.

Simulation of the effect of wavefront collision on the upstroke of the transmembrane action potential and the Na+ inward current. The values computed during uniform conduction (solid lines) are compared to the values computed at a collision site (dashed lines). Left top: Change of membrane potential, VM, during action potential upstroke. Left bottom: maximal upstroke velocity of transmembrane action potential in Volts/sec. Right top: Na+ inward current, INa. Right bottom: time course of Na+ conductance, gNa.

Reproduced from reference 20 with permission


Figure 10.

Simulation of the effect of wavefront dispersion on the upstroke of the transmembrane action potential and the Na+ inward current. A: Inset shows simulated two‐dimensional strand of excitable tissue emerging into a large area. Signals on panels A–D are simulated from sites 1–11 shown on the inset. Action potential upstrokes show a double component, which is most prominent at the expansion site. B: First time derivatives dV/dt from action potential upstrokes shown on panel A. C: Time course of Na+ conductance, gNa. D: Time course of Na+ inward current, INa. Note increase of INa at expansion site, associated with a decrease of dV/dtmax.

Reproduced from reference 72 with permission


Figure 11.

Effects of resistive discontinuities on propagation. Top: A row of simulated excitable elements (abscissa denotes element number) is separated by resistors. A number N of elements is connected by resistors of low value (200 Ω cm). Each group of N elements is connected to the next group by a single resistor, , of high value. Discontinuity at a constant value of effective longitudinal resistance can be changed by the simultaneous increase of N and . Bottom: Propagation along the simulated row of excitable elements, as illustrated by the time course of dV/dtmax (upper trace) and the action potential upstroke (lower trace). The degree of discontinuity is increased from panel A to C, while the value of effective or total longitudinal resistance is kept constant. Note increasing delay between the two action potential upstrokes, and the discontinuous upstroke in C.

Reproduced from reference 69 with permission


Figure 12.

Effects of resistive discontinuities on conduction velocity, θ. Propagation velocity is simulated in the model shown in the upper panel of Figure 11 as a function of the overall or effective resistance Ri (expressed as a fraction of the low value resistance of 200 Ωcm shown in FIG. 11, R/200) The solid line depicts the decrease of θ in a continuous cable where θ2 ∼ Ri. In curve A, the value for the high resistor, , is 5000 Ωcm, the numbers on the curve denote the number of elements N. In curve B, the degree of discontinuity is higher, because is 10,000 Ωcm. Note that in curve B, θ decreases above N = 16 and conduction block occurs when N >26 (see text).

Reproduced from reference 69 with permission


Figure 13.

Effects of resistive discontinuities on the maximal upstroke velocity of the transmembrane action potential, dV/dtmax (simulated in the model shown in the upper panel of FIG. 11). As a control, the dashed lines depict the dV/dtmax values for continuous cables (upper line Ri = 200 Ωcm, lower line R, = 4200 Ωcm). In all solid curves shown, the value of the high resistor is set to = 4200 Ωcm, and the curves differ with respect to their numbers of elements N. The curves are shown for N = 5, N = 9, and N = 51. With N = 51, there is dispersion of local current beyond the first resistive obstacle with a decrease of dV/dtmax, and collision before the next resistive obstacle (from N 40 to 51) with an increase in dV/dtmax. With a small number of N 5, the effect of dispersion and collision is minor and all of the dV/dtmax values are above the upper dashed line. This is due to the almost simultaneous excitation of all elements, similar to the situation of a space‐clamped action potential. N = 9 corresponds to an intermediate situation.

Reproduced from reference 69 with permission


Figure 14.

Simulated cell chain. Comparison of continuous with discontinuous conduction with decreasing cell‐to‐cell coupling. Effects of variations in axial (longitudinal) resistance on microscopic velocity (θmic, curve 1) and on average macroscopic velocity (θmac, curve 2). The microscopic velocity corresponds to the velocity inside a cell. The case of a continuous structure is shown for comparison on curve 3 and follows the inverse square root relation of continuous cable theory. The effective longitudinal resistance is changed by varying the disk resistance while the myoplasm resistance is kept constant at 200. Both effective longitudinal resistivity and the corresponding disk resistance are indicated.

Reproduced from reference 26 with permission


Figure 15.

Simulated chain of cells. Changes of maximal upstroke velocity with decreasing cell‐to‐cell coupling. Conduction velocity θ (dashed line) and dV/dtmax (solid line) are plotted as a function of the increasing effective resistance or the specific disk resistance (coupling resistance between simulated cells). There is a transient increase of dV/dtmax with increasing cell‐to‐cell uncoupling and decreasing velocity of propagation.

Reproduced from reference 26 with permission


Figure 16.

Experimental determination of conduction in a single cell chain. A: Reproduction of the microscopic appearance of a cultured cell chain. Dots and numbers denote positions of three light‐sensitive diodes (6.5 μm in diameter) separated by a distance of 30 μm. B: Optical recording of action potential upstrokes from diodes 1–3, measured as fluorescence change ΔF/F of a voltage‐sensitive dye (upper traces) and the first time derivatives (lower traces). Numbers 1–3 denote times of local activation. Note that the conduction delay across the cell border 2,3 is larger than delay within the cytoplasm 1,2. C: Histograms of cytoplasmic (upper graph) and junctional (lower graph) conduction times. The difference between the mean conduction times amounts to approximately 80 μsec and reflects the mean conduction time across the end‐to‐end cell junctions.

Reproduced from reference 83 with permission


Figure 17.

Experimental determination of conduction in a cell strand. A: Reproduction of the microscopic appearance of a cultured cell strand (4–5 cells in width). Dots and numbers denote positions of three light‐sensitive diodes (6.5 μm in diameter) separated by a distance of 30 μm. B: Optical recording of action potential upstrokes from diodes 1–3, measured as fluorescence change ΔF/F of a voltage‐sensitive dye (upper traces) and the first time derivatives (lower traces). Numbers 1–3 denote times of local activation. C: Histograms of cytoplasmic (upper graph) and junctional (lower graph) conduction times. With respect to the measurements shown in Figure 16 the mean cytoplasmic conduction time has increased from 38 μsec to 60 μsec and the mean junctional conduction time has decreased from 118 μsec to 80 μsec. This is explained by electrotonic current flow through lateral gap junctions (“ lateral averaging”; see text).

Reproduced from reference 83 with permission


Figure 18.

Propagation in a cellular network, simulation of intracellular excitation sequences. A, longitudinal conduction: Conduction from left to right is depicted on the upper trace, conduction from right to left is depicted on the lower trace. Isochrone lines are separated by 4 μsec. B, transverse conduction: Conduction from top to bottom is shown on the upper trace, propagation from bottom to top is depicted on the lower trace. Intracellular isochrones are separated by 3 μsec. During longitudinal propagation, there is a crowding of isochrones (slow propagation) at the beginning of propagation in the individual cells and acceleration of propagation toward the end of the cells. During transverse propagation, the arrows indicate preferential longitudinal propagation spread, with microcollisions occurring occasionally (asterisk).

Reproduced from reference 29 with permission


Figure 19.

Simulation of intracellular excitation sequences (A), intracellular distributions of dV/dtmax (B) and inward Na+ charge movements during excitation (C) in an anisotropic cellular network. The left graphs correspond to longitudinal propagation from left to right and the right graphs correspond to transverse propagation from top to bottom. Note the close correspondence between the isochrone spacing, dV/dtmax and inward Na+ charge movement during both transverse and longitudinal propagation: The dV/dtmax is relatively low where excitation is slow and vice versa. By contrast the locations of slow activation and low dV/dtmax correspond to large inward Na+ charge movements and vice versa.

Reproduced from reference 29 with permission


Figure 20.

Ratio of extracellular to intracellular resistance in compact ventricular tissue. The amplitude of the bipolar extracellular electrogram (upper trace, VE) and the amplitude of the action potential upstroke (lower trace, VM) are shown from an isolated, arterially‐perfused papillary muscle. The signals are measured in a muscle which is surrounded by an electrical insulator. The ratio of VE/(VM − VE) which is approximately 1, corresponds to the ratio of extracellular: intracellular resistance, rori. This demonstrates that the extracellular resistance in compact ventricular tissue is of approximately the same magnitude is the resistance of the intracellular space (including the gap junctions).

Reproduced from reference 32 with permission


Figure 21.

Effect of superfusion fluid on conduction velocity, θ, the maximal upstroke velocity of the transmembrane action potential dV/dtmax, and the time constant of the initial rise of the action potential, τfoot. Panels (a) and (b) show action potentials measured at two sites along a cylindrical papillary muscle, between electrodes D and C, (D‐C), and between electrodes A and B, (A‐B). The bipolar extracellular electrogram is measured between the extracellular electrodes c and b, (b‐c). The muscle is either soaked in a large bulk solution (SF, closed circles) or covered only by a thin fluid layer (T, open circles). Panel (b) illustrates the curved wavefronts measured in the presence of the large bulk solution during propagation from left to right. Traces on the right correspond from top to bottom: τa‐b, dV/dtmax, a‐b, τd‐c dV/dtmax d‐c, θ0 (conduction velocity at the surface) and θi (conduction velocity in the core of the fiber). Increasing the thickness of the fluid layer (electrical shunting, transition from T to SF) produces (1) an increase of θ0 and θi; (2) an increase of τa‐b and τd‐c; (3) a decrease of dV/dtmax.

Reproduced from reference 107 with permission


Figure 22.

Three‐dimensional plot of simulated transmembrane voltage (Φm) after application of a point stimulus to two‐dimensional anisotropic tissue. The X‐axis corresponds to the longitudinal direction of the fibers, the Y‐axis to the transversal direction of the fibers. Panel A: Simulation of anisotropy with equal ratios of extracellular to intracellular conductivities. Panel B: Longitudinal conductivity ratio (σexix) 8 × 10−4:2 × 10−4; conductivity ratio (σeyiy) 2 × 10−4:2 × 10−5 Note that with an equal anisotropic ratio there is a drop of Φm with distance from current injection and a elliptical shape of Φm distribution in the x/y plane. In Panel B, which corresponds to the simulation using experimentally determined values of intra‐ and extracellular conductivities, there is a hyperpolarization of Φm in the × (longitudinal) and a depolarization in the Y (transverse) directions.

Reproduced from reference 112 with permission


Figure 23.

Dog‐bone shape of virtual electrode. Plots showing the shape of virtual electrodes caused by stimuli from 1–7 mA. The horizontal axis corresponds to the longitudinal direction of the anisotropic subepicardial layer of a dog, the vertical axis corresponds to the transverse fiber direction. Note that upon application of a central cathodal stimulus, the line where excitation starts to propagate, i.e. the virtual electrode, has a dog bone shape with a very large extension of the electrode in the transverse direction.

Reproduced from reference 113 with permission


Figure 24.

Interaction between an anisotropic medium and a bathing solution. A: Deviation of the isochrones of a wave propagating in the longitudinal direction (left side) and a wave propagating in the transverse direction (right side). Note that due to the anisotropy‐dependent differences in intra‐ and extracellular conductivities, wavefront bending in the longitudinal direction is significantly more expressed. In the absence of a bulk conductor, both longitudinal and transverse wavefronts are flat (not shown). B: Phase plane plots of dV/dtmax versus membrane potential in a model with unequal anisotropy. Superimposed are phase plane plots of action potentials propagating in the longitudinal direction (L), at angles of 30°, 45°, and 60° from the longitudinal direction and in the transverse direction (T). On the left hand side, the bathing fluid was absent. Note that all the traces almost superimpose. On the right hand side, a bathing fluid has been added to the boundary of the tissue. In this case there is a marked direction dependence of both the initial portion of the action potential of and of dV/dtmax.

Reproduced from references 109,120 with permission


Figure 25.

Isochronal map of excitation spread from the sinoatrial node. Tones depict excitation intervals in steps of 5 msec, numbers correspond to activation times in msec. Configuration of action potentials is shown along the pathway of conduction from the sinus node to the atrium. The dashed line indicates the beginning of the atrial electrogram used as time reference. Toward the periphery, action potentials show an increase in amplitude and dV/dtmax and a decrease in rate of diastolic depolarization. The area in which two component action potentials were recorded is hatched in the activation map.

Reproduced with permission from reference 126


Figure 26.

Spread of excitation in the right atrium. Isochronal map of the spread of excitation from the sinoatrial node (S. A. N.) over the epicardial surface of the right atrium made by Thomas Lewis in 1915. Although in general spread of activation is depicted as being radial, the isochrones deviate over the crista terminalis, indicating preferential conduction in that region.

Reproduced with permission from reference 448


Figure 27.

Anatomy of the rabbit right atrium. The contrast of the preparation has been enhanced by supravital staining with methylene blue: ct, cut ends of the crista terminalis; ra, right auricle; fo, fossa ovalis; SVC, superior vena cava; ivc, inferior vena cava; ocs, ostium of the coronary sinus; vc, valve of the coronary sinus; ss, sinus septum; asl, attachment of the septal tricuspid leaflet (marked by a dotted line); SA, sinoatrial node (the letters inside surrounded by a dotted line indicate the approximate area of the compact node consisting of typical nodal cells); CN, cauda of the sinoatrial node; rb, right branch of the sinoatrial right bundle; lb, left (septal) branch of the sinoatrial ring bundle; tc, transitional cell zone of the atrioventricular node; mc, midnodal cells; lc, lower nodal cells; avb, atrioventricular bundle; ao, atrial overlay fibers.

Reproduced with permission from reference 449


Figure 28.

Activation sequence in the right atrium. A: A photograph of the right atrium of a rabbit: the activation sequence has been mapped using a tenfold microelectrode assembly, with which 280 different atrial cells have been impaled during the course of the experiment. The specimen was opened by an incision along the lateral margin of the tricuspid valve, and pinned out in order not to interrupt the internodal tissues. After the experiment, the preparation was fixed before being photographed. The specimen is illuminated from behind to illustrate the thick and thin parts of the myocardium. B: A diagrammatic representation, in which the abbreviations are as follows: SVC, superior vena cava; IVC, inferior vena cava; FO, fossa ovalis; CS, coronary sinus; AVN, atrioventricular node; MS, membranous septum; IVS, interventricular septum; C and D: Activation maps made according to photographs taken while the preparation was in the tissue bath—hence the slight differences in shape when compared to the fixed specimen. The preparation was beating spontaneously for the construction of (C); in (D), the preparation was paced through an electrode placed above the ostium of the SVC. In both instances, the activation sequence, indicated by isochrone lines separating areas activated within 5 msec intervals, follows the thicker muscle bundles. These are the crista terminalis, the septal branch of the crista, and the thick muscle ridge between IVC and FO. The AV node receives a dual input. The localization of the pacemaker determines the dominant input. In C, during spontaneous sinus rhythm, the AV node is reached earliest by the posterior route. In D during driving from the SVC, the anterior limbus activates the AV node earlier.

Reproduced with permission from reference 176


Figure 29.

Anatomy of the AV junction. A: Photograph and B sketch of a normal human heart showing the anatomical landmarks of the triangle of Koch. The approximate site of the compact AV node is indicated by the stippled area adjacent to the central fibrous body.

Reproduced with permission from reference 183


Figure 30.

Propagation toward the AV node. Arrows indicate the four areas from which atrial fibers approach the specialized AV junctional area. The fourth area, indicated by the curved arrow, is from the left atrial aspect of the septum.

Reproduced with permission from reference 183


Figure 31.

Action potential characteristics in the AV node: Action potentials of six types of AV nodal cells during periodic premature stimulation of the right atrium. Each section was obtained by superimposing (in decreasing order of coupling stimulation intervals [numbers at left in msec]) tracings corresponding to last basic and premature beat. Baseline of each subsequent tracing was shifted downward to help distinguish potentials. Action potential after premature potential in lower trace in AN (atrionodal) and H (His) was caused by an atrial re‐entrant beat. Note double components in N (nodal) cell of early premature responses. ANL, late AN cells; ANCO, AN cells with action potential upstroke with two components.

Reproduced with permission from reference 193


Figure 32.

Activation of the AV node. Map showing sequence of normal antegrade conduction of rabbit AV node. Symbols indicate position of AV nodal cells from which action potentials were recorded and also in which 20 msec interval these cells were activated. Note dual input into AV node. CT, crista terminalis; IAS, interatrial septum; CS, ostium of coronary sinus; Tr. V., tricuspid valve; H, position of extracellular electrode on His bundle.

Reproduced with permission from reference 200


Figure 33.

Conduction in the AV node. Two simultaneously recorded transmembrane potentials from the AV node of a Langendorff blood‐perfused canine heart at a superficial and a deep site at the same location. The atrium was paced at a basic cycle length of 600 msec, and a premature stimulus S2 was applied at a coupling interval of 300 msec, either at the posterior input (“ slow pathway,” upper panel) or at the anterior input (“ fast pathway,” lower panel). Note double components, especially during premature stimulation, where the action potential of the superficial cell causes a slow prepotential in the deeper cell which during posterior stimulation is large enough to cause an action potential that is propagated to the His bundle (not shown), but that fails to reach threshold during antegrade stimulation.

Retraced from unpublished recordings by J. M. T. de Bakker


Figure 34.

Abnormal Wenckebach phenomenon. Three simultaneously recorded action potentials in posterior input (cell a), in anterior input (cell b), and in junctional area of these two inputs (cell c). Note the difference in timing and configuration of the action potentials of cells a and c and of the His bundle complex during the first and fifth beats. Double bars indicate block. Numbers are activation times in msec. Atr., recording electrode on crista terminalis from which the electrogram in the upper trace was recorded. His, position of electrode recording electrogram of His bundle (lower trace).

Reproduced with permission from reference 200


Figure 35.

Cycle length dependence of AV‐nodal conduction. Action potentials illustrating dependency of first and second component in N cells upon late AN and early NH potentials. Signals 1 and 2 were recorded from AN cells, signals 3, 4, and 5 from N cells, and signal 6 from an NH cell. Inset shows position of cells. First component is largest in N cells close to AN zone; second component is largest in cells close to NH cells. Note that second component occurs later than upstrokes of action potentials in cell 6. Note also that duration of prepotential in cell 6 increases progressively in successive activation and that level at which prepotential breaks into a fast upstroke remains constant in all cycle lengths. Cells 4, 5, and 6 were recorded simultaneously. Cells 1, 2, and 3 were recorded separately, approximately 3 min earlier.

Reproduced with permission from reference 199


Figure 36.

Cycle length dependence of AV nodal conduction. A: Classical ladder diagram used in electrocardiography to depict cycle length‐dependent conduction delay in the AV junction, in this case during a 4:3 Wenkenbach phenomenon. B: Modification of the ladder diagram to express saltatory nature of the cycle length‐dependent conduction delay.



Figure 37.

Division of the left bundle branch. An illustration of the variation in the structure of the divisions of the left bundle branch in 20 different hearts.

Reproduced with permission from reference 259


Figure 38.

Propagation across the Purkinje–muscle junction. Intracellular and extracellular recordings at Purkinje–ventricular muscle junctional sites. A: Action potential of a Purkinje fiber (Pi), coinciding with the Purkinje (P) deflection preceding the all‐negative ventricular muscle (V) deflection in the extracellular electrogram (e). s, Stimulus artifact. B: Action potential of a transitional cell (Ti), with an early slow component (arrow). C: Simultaneous intracellular recordings of an early (Ti) and a late (T2i) transitional cell. The latter could also be classified as an early ventricular cell that is electrotonically influenced from the transitional cells, giving rise to the long, slow foot (arrow). D: Simultaneous intracellular recordings of a Purkinje fiber and a transitional cell with a slow foot (arrow) and an inherent low amplitude that is heightened on activation of the ventricular mass. E: Intracellular recording from a transitional cell that coincides with a small deflection in the extracellular electrogram during the Purkinje fiber–ventricular muscle delay period. F: Intracellular recording of a transitional cell with multiple components during the upstroke (arrow). Panels A–C derive from rabbit hearts, and panels D–F derive from pig hearts.

Reproduced with permission from reference 272


Figure 39.

Structure of the Purkinje–muscle junction. Schematic representation of the structure of a rabbit Purkinje fiber–ventricular muscle junction. P, Purkinje fibers; T, transitional cells; V, ventricular myocardium.

Reproduced with permission from reference 272


Figure 40.

Safety factor of propagation. Results of computer simulation of propagation in a cell chain. The cells are separated by a simulated gap junction resistor. A: Safety factor of propagation, SF, as a function of propagation velocity. Dashed line: Change of SF with a decrease of excitability. Solid line: Change of SF with decreasing conductance (increasing resistance) between cells. B: Change of SF as a function of propagation velocity in the absence and presence of ICa,L. Note that very low conduction velocities can only be achieved with flow of I Ca,L. (See text.)

Reproduced with permission from reference 3


Figure 41.

Supernormal excitability and conduction. A, B: Transmembrane action potentials. C, D: Strength interval curves. Time scale is identical for both traces. A: Recording from the His bundle. B: Recording from a Purkinje fiber running freely in a false tendon. C: Recording from a transitional type Purkinje fiber. D: Recording from a ventricular cell. The supernormal phase of excitability in B is associated with a supernormal conduction.

Reproduced with permission from reference 279


Figure 42.

Relationship of excitability to conduction velocity during phase‐4 depolarization. Conduction time along a fixed distance of a Purkinje strand and maximal upstroke velocity of the transmembrane action potential, dV/dtmax of a Purkinje fiber are plotted as a function of the “ take‐off” potential during spontaneous phase 4 depolarization ranging from 92.5 to 105 mV. Note that with ongoing spontaneous depolarization (decrease of “ take‐off” potential”) there is a decrease of conduction time corresponding to an increase in conduction velocity and a decrease of dV/dtmax.

Reproduced with permission from reference 281


Figure 43.

Change of longitudinal and transverse conduction velocity with increasing extracellular potassium concentration, [K+]o. Measurements were made in an isolated perfused porcine heart. Note that propagation blocks at a [K+]o of 11 mM.

Reproduced with permission from reference 287


Figure 44.

Recovery of maximal upstroke velocity of the action potential, dV/dtmax. Action potentials were elicited at different times in the wake of the preceding action potential. Time 0 denotes the beginning of the preceding action potential. The curves depict the recovery curves of dV/dtmax with increasing time measured at different resting potentials. Note that recovery from activation becomes delayed with depolarization.

Reproduced with permission from reference 255


Figure 45.

Unidirectional block in the wake of the preceding wavefront. Transmembrane potentials (Vm) and sodium channel conductance (gNa) computed after application of a premature stimulus (cell 1) in the wake of propagating action potential. In the antegrade direction (right panel) recordings were taken from cells 0.5 mm apart. In the retrograde direction (left panel) traces were recorded from cells 1 mm apart. At the time of premature stimulation, membrane excitability at cell 1 was less than 10% of the maximum excitability (compare gNa curves 1 and 5 on the left panel). In the retrograde direction, the action potential propagated a distance of 4 mm before reaching the region of fully excitable membrane. In the antegrade direction, membrane excitability gradually decreased and propagation extinguished. Note different gNa scales in the left and right panels.

Reproduced with permission from reference 312


Figure 46.

Unidirectional block with asymmetric depression of excitability. Top: Injury is produced by a crushing probe. The line spacing on the Purkinje fiber indicates increasing degree of injury. Bottom: The influence of injury on the excitability threshold is illustrated. The amplitudes of the anterograde wavefront (C‐x‐B) are compared to the amplitudes of the retrograde wavefront (A‐x‐C). At y, the transition between normal cells and inexcitable, injured cells is abrupt. C represents decremental or augmental conduction, depending on direction, through a transitional zone of partial injury; x represents the point of transition between partially excitable cells and inexcitable cells (x‐y). A and B represent electrotonic transmission through inexcitable cells. The retrograde wave succeeds in conducting across, and the anterograde wave front fails.

Reproduced with permission from reference 322


Figure 47.

Unidirectional block at a geometrical tissue expansion: The role of Ca++‐inward current. A: Schematic representation of a cultured cell monolayer with geometrical expansion and overlaid photodiodes during antegrade (upper panel) and retrograde (lower panel) conduction. B and C: Action potential upstrokes recorded using a voltage‐sensitive dye in control conditions (B) and after administration of 5 μM nifedipine. In control conditions, the antegrade propagation was characterized by biphasic upstrokes and local slowing of conduction at the expansion (B). The blockage of Ca++ current with nifedipine produced antegrade conduction block (C). The retrograde propagation was successful in both cases.

Reproduced with permission from reference 80


Figure 48.

Unidirectional block in anisotropic tissue. Anisotropic conduction time curves obtained in human and canine atrial bundles. A: Uniform anisotropic pectinate muscle of a 12‐year‐old child. B: Non‐uniform anisotropic pectinate muscle of a 62‐year‐old man. C: Non‐uniform anisotropic muscle (christa terminalis) of an adult dog. In each preparation conduction times (msec per mm interelectrode distance) were obtained from analyzing the unipolar extracellular electrogram of two electrode pairs. As shown in the inset, the two electrode pairs were placed in longitudinal and transverse directions, respectively. Solid circles represent longitudinal propagation, open circles represent transverse propagation. Each preparation was stimulated at a basic rate, and premature action potentials were introduced at variable intervals, A1–A2. In the uniform anisotropic bundle (A) conduction times became longer with the shortening of the A1–A2 intervals, block occurred in both directions at the same interval. In the non‐uniform cases (B and C), block occurred in the longitudinal direction at a premature interval of 325 and 310 msec, respectively. At this prematurity, transverse propagation was still preserved.

Reproduced from reference 74 with permission


Figure 49.

Effect of microscopic resistive barriers on propagation. A: A phase‐contrast image of a cell culture (neonatal rat myocytes) with the overlaid diode array. Action potential upstrokes are measured at each diode location. The numbers 1–10 on the diode array correspond to the locations of the signals shown in D and E. In D and C, the location of these signals is indicated by the gray area. Two clefts in the central area (outlined in white in A) form an narrow isthmus of 40 μm. Activation maps of longitudinal and transverse conduction are shown in B and C respectively. Note slowing and deviation of the wavefront at the isthmus. Numbers denote separation of isochrones by 100 μsec. Selected recordings of action potential upstrokes during longitudinal and transverse conduction are shown in D and E, respectively. Discontinuities in the action potential upstrokes occur at the expansion site during transverse propagation.

Reproduced from reference 34 with permission


Figure 50.

Unidirectional conduction block at an “ isthmus.” Wave propagation across a narrow tissue isthmus in an isolated ventricular preparation of sheep heart. A: Map of activation spread before an isthmus was produced. B: Activation spread in the same preparation with the isthmus 2.26 mm wide. The isthmus was produced by two tissue cuts (gray zones). C: Activation spread after the isthmus was reduced to 0.88 mm. D: Local conduction velocity measured across the isthmus as a function of isthmus width.

Reproduced with permission from reference 332


Figure 51.

Circus movement re‐entry around a large anatomical obstacle.

Reproduced with permission from reference 350


Figure 52.

Effect of a premature impulse entering a re‐entrant circuit. The black and dotted areas show the absolute and relative refractory periods, respectively. In A and B, a premature impulse enters the circuit at the end of the relative refractory period and spreads in two directions. In the retrograde direction, the premature wave is annihilated by the circulating wave; in the anterograde direction, the premature impulse advances, resetting the tachycardia. In C and D, the premature impulse reaches the circuit closer to the state of absolute refractoriness. The impulse annihilates the retrograde wave and fails to propagate in the anterograde direction thereby terminating the tachycardia.

Reproduced with permission from reference 219


Figure 53.

Initiation of functional re‐entry by premature stimulation in an isolated preparation of rabbit atrial muscle. A: Isochronal activation map of basic beat (interval 500 msec). Dots indicate sites of stimulation. Activation times (msec) are given relative to the stimulus onset. B: Map of premature beat (interval 56 msec). T bars indicate conduction block. C: First cycle of tachycardia. D: Refractory periods measured during basic rhythm (in msec).

Reproduced with permission from reference 318


Figure 54.

Functional re‐entry and tachycardia. Activation map (right) and action potential recordings (left) obtained during steady‐state tachycardia. Cells in the central area of the re‐entrant circuit show double potentials of low amplitude (traces 3 and 4). Lower right: Schematic representation of the activation pattern. Double bars indicate conduction block.

Reproduced with permission from reference 350


Figure 55.

Spiral waves. Spiral waves in chemical Belousov‐Zhabotinsky reaction (A) and in an isolated preparation of canine epicardial muscle (B).

Reproduced with permission from references 366 and 376


Figure 56.

Initiation of spiral wave in a simple model of cardiac excitation. A rectangular area R is excited overlapping the absolutely refractory tail of a propagating wave (A). The premature wave propagates in the retrograde direction (right to left) but blocks in the anterograde direction, forming a wave break that turns around the refractory area R (B). When the area R recovers, the excitation wave short‐circuits this area (C) and forms a spiral wave rotating around a linear line of block (D–F).

Reproduced with permission from reference 373)


Figure 57.

Initiation of a spiral wave by cross‐field stimulation in canine right ventricular myocardium. A: Isochronal maps of activation and repolarization during wave propagation induced by stimulation (S1) from a line of eight epicardial pacing sites. Solid lines depict isochronal activation lines; dashed lines depict isorecovery lines. Numbers indicate time in msec. B: Gradients of extracellular potential (in V/cm) produced by a unipolar cathodal shock (S2) of 150 V from a mesh electrode at the bottom. C: Pattern of activation spread following sequential application of S1 stimulus from the right and S2 shock from the bottom. The S1–S2 interval was 191 msec and the S2 strength was 150 V. Activation times (msec) are measured from the start of the 3 msec S2 shock. The heavy solid line represents the transition between successive activation maps. Isochrones are drawn at 10 msec intervals. The hatched line represents a zone of conduction block. The double‐headed arrow indicates the mean epicardial fiber orientation in the area of conduction block. The hatched area indicates the region assumed to be directly excited by the S2 shock field. Earliest post‐shock activation occurs distant from the S2 site, with no early activation wavefronts conducting away from the directly excited region located between the S2 site and the critical point. A counterclockwise re‐entrant circuit is formed around the region containing the critical point and the block line. The potential gradient equals 5.8 V/cm, and the pre‐shock interval equals 171 msec at the critical point (critical refractory period = 169 msec). D: Schematic representation of the re‐entry initiation by cross‐field stimulation. The row of pacing wires (S1) on the right creates parallel isorecovery lines (R7 through R2), with R7 the least refractory and R2 the most refractory. The S2 from the bottom creates parallel isogradient lines (G7 through G3), with G7 the largest potential gradient and G3 the weakest. The S2 shock produces direct excitation (DE), graded response (GR), or neither effect (NE). Activation fronts propagate from only one part of the directly excited area, not from the directly excited region abutting the area of graded response, thus forming a zone of unidirectional conduction block.

Reproduced with permission from reference 379


Figure 58.

Initiation of a spiral wave at a pivoting point. Formation of a free wave break after wavefront detachment from the sharp edge of an inexcitable obstacle. Computer model with Luo‐Rudy ionic kinetics. The maximal sodium conductance was reduced to 6.6 mS/cm2 A: Isochronal map of activation spread with an interval of 5 msec. B: Snapshot of activation at the moment marked by the asterisk in A. Black indicates the excited area defined by the activation of inward Na+ current. Gray indicates the area in the refractory state as defined by Na+ current inactivation. Point P marks the wave tip, defined as a point where excited, refractory, and resting states meet. The dashed line t shows the trajectory of the wave tip with the radius rp

Reproduced with permission from reference 34


Figure 59.

Dynamics of spiral wave rotation in mathematical models of cardiac excitation. A–C: Cellular automata model. Spiral wave rotation changes from circular (A) to meandering (B), and then to Z‐type (C) with increasing excitability. D–G: FitzHugh‐Nagumo model. The same types of rotation are observed when the wavelength of excitation is increased.

Reproduced with permission from references 450 and 383


Figure 60.

Drift of a spiral wave and the Doppler effect. A and B: Isochronal activation maps showing initiation (A) and the first rotation cycle (B) of a spiral wave in an isolated preparation of epicardial muscle. A stepwise inhomogeneity in refractory period was created by separate superfusion of two parts of the preparation with normal and quinidine‐containing solutions. Dashed line shows the border of inhomogeneity with larger refractoriness in the upper part. The asterisk shows the location of the stimulation electrode. C: Trajectory of the spiral wave tip during initiation (S1) and three subsequent cycles of spiral wave rotation (V1−V3). D: Excitation intervals measured along the border of inhomogeneity during spiral wave drift (cycle V2). Because of the drift, excitation intervals in front of the spiral wave are significantly shorter than intervals behind the spiral wave (Doppler effect).

Reproduced with permission from reference 407


Figure 61.

Anchoring of a spiral wave. A: Electrocardiographic recordings showing that premature stimulation (S2) produced polymorphic arrhythmic activity followed by a transition to sustained monomorphic tachycardia. B: Time‐space plot of activation spread obtained from video‐imaging of transmembrane potential (voltage‐sensitive dye). In these plots, the activity from the whole image is projected onto a single direction (vertical axis) and displayed as a function of time. White bands show a planar wave propagation while branching of bands indicates the presence of a spiral wave induced by the S2 stimulus. As detected from the movement of the branching point, which marks the center of the spiral, the spiral drifted during the first seven cycles and became stationary thereafter.

Reproduced with permission from reference 380
References
 1. Winfree, A. T. When Time Breaks Down. Princeton, N.J.: Princeton University Press, 1987.
 2. Kucera, J. P., A. G. Kleber, and S. Rohr. Slow conduction in cardiac tissue: II. effects of branching tissue geometry. Circ. Res. 83: 795–805, 1998.
 3. Shaw, R. M., and Y. Rudy. Ionic mechanisms of propagation in cardiac tissue. Roles of the sodium and L‐type calcium currents during reduced excitability and decreased gap junction coupling. Circ. Res. 81: 727–741, 1997.
 4. Joyner, R. W., F. Ramon, and J. W. Moore. Simulation of action potential propagation in an inhomogeneous sheet of coupled excitable cells. Circ. Res. 36: 654–661, 1975.
 5. Beeler, G. W., and H. Reuter. Reconstruction of the action potential of ventricular myocardial fibres. J. Physiol. (Lond.) 268: 177–210, 1977.
 6. Ebihara, L., and E. A. Johnson. Fast sodium current in cardiac muscle. A quantitative description. Biophys. J. 32: 779–790, 1980.
 7. Luo, C. H., and Y. Rudy. A model of the ventricular cardiac action potential—depolarization, repolarization, and their interaction. Circ. Res. 68: 1501–1526, 1991.
 8. Luo, C. H., and Y. Rudy. A dynamic model of the cardiac ventricular action potential. 1. Simulations of ionic currents and concentration changes. Circ. Res. 74: 1071–1096, 1994.
 9. Noble, D. The development of mathematical models of the heart. Chaos Solitons and Fractals. 5: 321–333, 1995.
 10. Jack, J. J. B., D. Noble, and R. W. Tsien. Electric Current Flow in Excitable Cells. Oxford: Clarendon Press, 1975.
 11. Hodgkin, A. L., and W. A. H. Rushton. The electrical constants of crustacean nerve fibers. Proc. R. Soc. (Lond.) B133: 444–479, 1946.
 12. Weidmann, S. The diffusion of radiopotassium across intercalated disks of mammalian cardiac muscle. J. Physiol. (Lond.) 187: 323–342, 1966.
 13. Weidmann, S. The electrical constants of Purkinje fibres. J. Physiol. (Lond.) 118: 348–360, 1952.
 14. Weidmann, S. Electrical constants of trabecular muscle from mammalian heart. J. Physiol. (Lond.) 210: 1041–1054, 1970.
 15. Woodbury, J. W., and W. E. Crill. On the problem of impulse conduction in the atrium. In: Nervous Inhibition, edited by L. Florey. New York: Plenum Press, 124–135, 1961.
 16. Jongsma, H. J. and H. E. van Rijn. Electrotonic spread of current in monolayer cultures of neonatal rat heart cells. J. Membr. Biol. 9: 341–360, 1972.
 17. Pressler, M. L. Cable analysis in quiescent and active sheep Purkinje fibres. J. Physiol. (Lond.) 352: 739–757, 1984.
 18. Fleischhauer, J., L. Lehmann, A. G. Kléber. Electrical resistances of interstitial and microvascular space as determinants of the extracellular electrical field and velocity of propagation in ventricular myocardium. Circulation 92: 587–594, 1995.
 19. Weidmann, S. The effect of the cardiac membrane potential on the rapid availability of the sodium‐carrying system. J. Physiol. (Lond.) 127: 213–224, 1955.
 20. Spach, M. S., and J. M. Kootsey. Relating the sodium current and conductance to the shape of transmembrane and extracellular potentials by simulation: effects of propagation boundaries. IEEE Trans. Biomed. Eng. 32: 743–755, 1985.
 21. Walton, M. K., and H. A. Fozzard. The conducted action potential: models and comparison to experiments. Biophys. J. 44: 9–26, 1983.
 22. Hodgkin A. L. A note on conduction velocity. J. Physiol. (Lond.) 125: 221–224, 1954.
 23. Tasaki I. and S. Hagiwara. Capacity of muscle fiber membrane. Am. J. Physiol. 188: 423–429, 1957.
 24. Sommer, J. R., and B. Scherer. Geometry of cell and bundle appositions in cardiac muscle: light microscopy. Am. J. Physiol. 248 (Heart Circ. Physiol. 17): H792–H803, 1985.
 25. Le Grice, I. J., B. H. Smaill, L. Z. Chai, S. G. Edgar, J. B. Gavin, and P. J. Hunter. Laminar structure of the heart: ventricular myocyte arrangement and connective tissue architecture in the dog. Am. J. Physiol. 269 (Heart Circ. Physiol. 38): H571–H582, 1995.
 26. Rudy, Y., and W. Quan. A model study of the effects of the discrete cellular structure on electrical propagation in cardiac tissue. Circ. Res. 61: 815–823, 1987.
 27. Leon, L. J., and F. A. Roberge. Directional characteristics of action potential propagation in cardiac muscle. A model study. Circ. Res. 69: 378–395, 1991.
 28. Muller‐Borer, B. J., D. J. Erdman, and J. W. Buchanan. Electrical coupling and impulse propagation in anatomically modeled ventricular tissue. IEEE Trans. Biomed. Eng. 41: 445–454, 1994.
 29. Spach, M. S., and J. F. Heidlage. The stochastic nature of cardiac propagation at a microscopic level—electrical description of myocardial architecture and its application to conduction. Circ. Res. 76: 366–380, 1995.
 30. Buchanan, J. W., T. Saito, and L. S. Gettes. The effects of antiarrhythmic drugs, stimulation frequency, and potassium‐induced resting membrane potential changes on conduction velocity and dV/dtmax in guinea pig myocardium. Circ. Res. 56: 696–703, 1985.
 31. Kléber, A. G., C. B. Riegger, and M. J. Janse. Electrical uncoupling and increase of extracellular resistance after induction of ischemia in isolated, arterially perfused rabbit papillary muscle. Circ. Res. 61: 271–279, 1987.
 32. Kléber, A. G., and C. B. Riegger. Electrical constants of arterially perfused rabbit papillary muscle. J. Physiol. (Lond.) 385: 307–324, 1987.
 33. Riegger, C. B., G. Alperovich, and A. G. Kléber. Effect of oxygen withdrawal on active and passive electrical properties of arterially perfused rabbit ventricular muscle. Circ. Res. 64: 532–541, 1989.
 34. Fast, V. G., B. J. Darrow, J. E. Saffitz, and A. G. Kléber. Anisotropic activation spread in heart cell monolayers assessed by high‐resolution optical mapping: role of tissue discontinuities. Circ. Res. 79: 115–127, 1996.
 35. Keener, J. P. A geometrical theory for spiral waves in excitable media. SIAM J. Appl. Math. 46: 1039–1056, 1986.
 36. Zykov, V. S., and O. L. Morozova. Speed of spread of excitation in two‐dimensional excitable medium. Biofizika 24: 739–744, 1979.
 37. Zykov, V. S. Analytical evaluation of the dependence of the speed of an excitation wave in a two‐dimensional excitable medium on the curvature of its front. Biophysics 25: 906–911, 1980.
 38. Zykov, V. S. Simulation of Wave Processes in Excitable Media. Manchester, England: Manchester University Press, 1987.
 39. Knisley, S. B., and B. C. Hill. Effects of bipolar point and line stimulation in anisotropic rabbit epicardium: assessment of the critical radius of curvature for longitudinal block. IEEE Trans. Biomed. Eng. 42: 957–966, 1995.
 40. Noble, D. The relation of Rushton “liminal length” for excitation to the resting and active conductances of excitable cells. J. Physiol. (Lond.) 226: 573–591, 1972.
 41. Fozzard, H. A., and M. Schoenberg. Strength‐duration curves in cardiac Purkinje fibres: effects of liminal length and charge distribution. J. Physiol. (Lond.) 226: 593–618, 1972.
 42. Rushton, W. A. H. Initiation of the propagated disturbance. Proc. R. Soc. B124: 210, 1937.
 43. Lindemans, F. W., and J. J. D. van der Gon. Current thresholds and liminal size in excitations of heart muscle. Cardiovasc. Res. 12: 477–485, 1978.
 44. Ramza, B. M., R. W. Joyner, R. C. Tan, and T. Osaka. Cellular mechanism of the functional refractory period in ventricular muscle. Circ. Res. 66: 147–162, 1990.
 45. Lindemans, F. W., and A. N. E. Zimmerman. Acute voltage, charge, and energy thresholds as functions of electrode size for electrical stimulation of the canine heart. Cardiovasc. Res. 13: 383–391, 1979.
 46. Winfree, A. T. The electrical thresholds of ventricular myocardium. J. Cardiovasc. Electrophysiol. 1: 393–410, 1990.
 47. Hoyt, R. H., M. L. Cohen, and J. E. Saffitz. Distribution and three‐dimensional structure of intercellular junctions in canine myocardium. Circ. Res. 64: 563–574, 1989.
 48. Saffitz, J. E., H. L. Kanter, K. G. Green, T. K. Tolley, and E. C. Beyer. Tissue‐specific determinants of anisotropic conduction velocity in canine atrial and ventricular myocardium. Circ. Res. 74: 1065–1070, 1994.
 49. Luke, R., and J. Saffitz. Remodelling of ventricular conduction pathways in healed canine infarct border zones. J. Clin. Invest. 87: 1594–1602, 1991.
 50. Saffitz, J. E., M. D. Lloyd, B. J. Darrow, H. L. Kanter, J. G. Laing, and E. C. Beyer. The molecular basis of anisotropy: role of gap junctions. J. Cardiovasc. Electrophysiol. 6: 498–510, 1995.
 51. Davis, L. M., H. L. Kanter, E. C. Beyer, and J. E. Saffitz. Distinct gap junction protein phenotypes in cardiac tissues with disparate conduction properties. J. Am. Coll. Cardiol. 24: 1124–1132, 1994.
 52. Kanter, H., J. Saffitz, and E. Beyer. Cardiac myocytes express multiple gap junction proteins. Circ. Res. 70: 438–444, 1992.
 53. Gourdie, R., C. Green, N. Severs, and R. Thompson. Immuno‐labelling patterns of gap junction connexins in the developing and mature rat heart. Anat. Embryol. 185: 163–178, 1992.
 54. Oosthoek, P. W., S. Viragh, A. E. M. Mayen, M. J. A. Vankempen, W. H. Lamers, and A. F. M. Moorman. Immunohistochemical delineation of the conduction system: 1. The sinoatrial node. Circ. Res. 73: 473–481, 1993.
 55. Oosthoek, P. W., S. Viragh, W. H. Lamers, and A. F. M. Moorman. Immunohistochemical delineation of the conduction system. 2. The atrioventricular node and Purkinje fibers. Circ. Res. 73: 482–491, 1993.
 56. Van Kempen, M. J. A., C. Fromaget, D. Gros, A. F. M. Moorman, and W. H. Lamers. Spatial distribution of connexin‐43, the major cardiac gap junction protein‐in the developing and adult rat heart. Circ. Res. 68: 1638–1651, 1991.
 57. Anumonwo, J. M. B., H. Z. Wang, E. Trabkajanik, B. Dunham, R. D. Veenstra, M. Delmar, and J. Jalife. Gap junctional channels in adult mammalian sinus nodal cells—immunolocalization and electrophysiology. Circ. Res. 71: 229–239, 1992.
 58. Trabka, Janik E., W. Coombs, L. F. Lemanski, M. Delmar, and J. Jalife. Immunohistochemical localization of gap junction protein channels in hamster sinoatrial node in correlation with electrophysiologic mapping of the pacemaker region. J. Cardiovasc. Electrophysiol. 5: 125–137, 1994.
 59. Opthof, T. Gap junctions in the sino‐atrial node: immunohistochemical localization and correlation with activation pattern. J. Cardiovasc. Electrophysiol. 5: 138–143, 1994.
 60. Kwong, K. F., R. B. Schuessler, K. G. Green, J. G. Laing, E. C. Beyer, J. P. Boineau, and J. E. Saffitz. Differential expression of gap junction proteins in the canine sinus node. Circ Res. 82: 604–612, 1998.
 61. Chen, S., L. M. Davis, L. M. Westphale, E. C. Beyer, and J. E. Saffitz. Expression of multiple gap junction proteins in human fetal and infant heart. Pediatr. Res. 36: 561–566, 1994.
 62. Bastide, B., L. Neyses, D. Ganten, M. Paul, and K. Willecke. Gap junction protein connexin40 is preferentially expressed in vascular endothelium and conductive bundles of the rat myocardium and is increased under hypertensive conditions. Circ. Res. 73: 1138–1149, 1993.
 63. Gros, D., T. Jarryguichard, I. Tenvelde, A. De Maziere, M. J. A. Van Kempen, J. Davoust, J. P. Briand, A. F. M. Moorman, and H. J. Jongsma. Restricted distribution of connexin40, a gap junctional protein, in mammalian heart. Circ. Res. 74: 839–851, 1994.
 64. Dolber, P. C., E. C. Beyer, J. L. Junker, and M. S. Spach. Distribution of gap junctions in dog and rat ventricle studied with a double‐label technique. J. Mol. Cell. Cardiol. 24: 1443–1457, 1992.
 65. Smith, J. H., C. R. Green, N. S. Peters, S. Rothery, and N. J. Severs. Altered patterns of gap junction distribution in ischemic heart disease—an immunohistochemical study of human myocardium using laser scanning confocal microscopy. Am. J. Pathol. 139: 801–821, 1991.
 66. Peters, N. S. New insights into myocardial arrhythmogenesis: distribution of gap‐junctional coupling in normal, ischaemic and hypertrophied human hearts. Clin. Sci. 90: 447–452, 1996.
 67. Darrow, B. J., V. G. Fast, A. G. Kléber, E. C. Beyer, and J. E. Saffitz. Functional and structural assessment of intercellular communication: increased conduction velocity and enhanced connexin expression in dibutyryl cAMP‐treated cultured cardiac myocytes. Circ. Res. 79: 174–183, 1996.
 68. Zhuang, J., K. A. Yamada, J. E. Saffitz, and A. K. Kléber. Pulsatile stretch remodels cell‐to‐cell communication in cultured myocytes. Circ. Res. 87: 316–322, 2000.
 69. Joyner, R. W. Effects of the discrete pattern of electrical coupling on propagation through an electrical syncytium. Circ. Res. 50: 192–200, 1982.
 70. Spach, M. S., R. C. Barr, G. S. Serwer, E. A. Johnson, and J. M. Kootsey. Collision of excitation waves in the dog Purkinje system: extracellular identification. Circ. Res. 24: 499–511, 1971.
 71. Joyner, R. W., R. Veenstra, D. Rawling, and A. Chorro. Propagation through electrically coupled cells. Effects of a resistive barrier. Biophys. J. 45: 1017–1025, 1984.
 72. Fast, V. G., and A. G. Kléber. Cardiac tissue geometry as a determinant of unidirectional conduction block: assessment of microscopic excitation spread by optical mapping in patterned cell cultures and in a computer model. Cardiovasc. Res. 29: 697–707, 1995.
 73. Fast, V. G., and A. G. Kléber. Block of impulse propagation at an abrupt tissue expansion: evaluation of the critical strand diameter in 2‐ and 3‐dimensional computer models. Cardiovasc. Res. 30: 449–459, 1995.
 74. Spach, M. S, and M. E. Josephson. Initiating reentry: the role of nonuniform anisotropy in small circuits. J. Cardiovasc. Electrophysiol. 5: 182–209, 1994.
 75. Henriquez, C. S., and R. Plonsey. Effects of resistive discontinuities on waveshape and velocity in a single cardiac fibre. Med. Biol. Eng. Comput. 25: 428–438, 1987.
 76. Tan, R. C., and R. W. Joyner. Electrotonic influences on action potentials from isolated ventricular cells. Circ. Res. 67: 1071–1081, 1990.
 77. Joyner, R. W., H. Sugiura, and R. C. Tan. Unidirectional block between isolated rabbit ventricular cells coupled by a variable resistance. Biophys. J. 60: 1038–1045, 1991.
 78. Sugiura, H., and R. W. Joyner. Action potential conduction between guinea pig ventricular cells can be modulated by calcium current. Am. J. Physiol. 263 (Heart Circ. Physiol. 32): H1591–H1604, 1992.
 79. Kumar, R. R., and R. W. Joyner. Calcium currents of ventricular cell pairs during action potential conduction. Am. J. Physiol. 268 (Heart Circ. Physiol. 37): H2476–H2486, 1995.
 80. Rohr, S., A. G. Kléber, and J. P. Kucera. Induction of very slow and discontinuous conduction by palmitoleic acid in linear strands of rat ventricular myocytes. Biophys. J. 70: A279, 1996.
 81. Rudy, Y., and W. Quan. Propagation delays across cardiac gap junctions and their reflection in extracellular potentials: a simulation study. J. Cardiovasc. Electrophys. 2: 299–315, 1991.
 82. Rohr, S., and B. M. Salzberg. Discontinuities in action potential propagation along chains of single ventricular myocytes in culture : multiple site optical recording of transmembrane voltage (MSORTV) suggests propagation delays at the junctional sites between cells. Biol. Bull. Mar. Biol. Lab. 183: 342–343, 1992.
 83. Fast, V. G., and A. G. Kléber. Microscopic conduction in cultured strands of neonatal rat heart cells measured with voltage‐sensitive dyes. Circ. Res. 73: 914–925, 1993.
 84. Purdy, J. E., M. Lieberman, A. E. Roggeveen, and R. G. Kirk. Synthetic strands of cardiac muscle. Formation and ultrastructure. J. Cell Biol. 55: 563–578, 1972.
 85. Lieberman, M., A. E. Roggeveen, J. E. Purdy, and E. A. Johnson. Synthetic strands of cardiac muscle: growth and physiological implication. Science. 175: 909–911, 1972.
 86. Horres, C. R., M. Lieberman, and J. E. Purdy. Growth orientation of heart cells on nylon monofilament: determination of the volume‐to‐surface ratio and intracellular potassium concentration. J. Membr. Biol. 34: 313–329, 1977.
 87. Rohr, S., D. M. Schölly, and A. G. Kléber. Patterned growth of neonatal rat heart cells in culture: morphological and electrophysiological characterization. Circ. Res. 68: 114–130, 1991.
 88. Rohr, S. Determination of impulse conduction characteristics at a microscopic scale in patterned growth heart cell cultures using multisite optical mapping of transmembrane voltage. J. Cardiovasc. Electrophysiol. 6: 551–568, 1995.
 89. Buchanan, J. W., and L. S. Gettes. Ionic environment and propagation. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. P. Zipes and J. Jalife, F. L. Orlando: W. B. Saunders; 149–156, 1990.
 90. Clerc, L. Directional differences of impulse spread in trabecular muscle from mammalian heart. J. Physiol. (Lond.) 255: 335–346, 1976.
 91. Spach, M. S., J. F. Heidlage, P. C. Dolber, and R. C. Barr. Electrophysiological effects of remodeling cardiac gap junctions and cell size: experimental and model studies of normal cardiac growth. Circ. Res. 86: 302–311, 2000.
 92. Spach, M. S., and J. M. Kootsey. The nature of electrical propagation in cardiac muscle. Am. J. Physiol. 244 (Heart Circ. Physiol. 13): H3–H22, 1983.
 93. Spach, M. S., W. T. I. Miller, D. B. Gezelowitz, R. C. Barr, J. M. Kootsey, and E. A. Johnson. The discontinuous nature of propagation in normal canine cardiac muscle. Evidence for recurrent discontinuties of intracellular resistance that affect the membrane currents. Circ. Res. 48: 39–54, 1981.
 94. Spach, M. S., and P. C. Dolber. Relating extracellular potentials and their derivatives to anisotropic propagation at a microscopic level in human cardiac muscle. Evidence for electrical uncoupling of side‐to‐side fiber connections with increasing age. Circ. Res. 58: 356–371, 1986.
 95. Spach, M. S., P. C. Dolber, and J. F. Heidlage. Properties of discontinuous anisotropic propagation at a microscopic level. Ann. N.Y. Acad. Sci. 591: 62–74, 1990.
 96. Cole, W. C., J. B. Picone, and N. Sperelakis. Gap junction uncoupling and discontinuous propagation in the heart. A comparison of experimental data with computer simulation. Biophys. J. 53: 809–818, 1988.
 97. Fast, V. G., and A. G. Kléber. Anisotropic conduction in monolayers of neonatal rat heart cells cultured on collagen substrate. Circ. Res. 75: 591–595, 1994.
 98. Spach, M. S., J. F. Heidlage, E. D. Darken, E. Hofer, K. H. Raines, and C. F. Starmer. Cellular dV/dtmax reflects both membrane properties and the load presented by ajoining cells. Am. J. Physiol. 263 (Heart Circ. Physiol. 32): H1885–H1863, 1992.
 99. Mays, D. J., J. M. Foose, L. H. Philipson, and M. M. Tamkun. Localization of the Kv1.5 K+ channel in explanted cardiac tissue. J. Clin. Invest. 96: 282–292, 1995.
 100. Rohr, S., R. Flückiger, and S. Cohen. Immunocytochemical localization of sodium and calcium channels in cultured neonatal rat ventricular myocytes. Biophys J. 76: A366, 1999 (abstract).
 101. Petrecca, K., F. Amellal, D. W. Laird, S. A. Cohen, and A. Shrier. Sodium channel distribution within the rabbit atrioventricular node as analysed by confocal microscopy. J Physiol (Lond.) 501: 263–274, 1997.
 102. 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.
 103. Kadish, A. H., J. F. Spear, J. H. Levine, and E. N. Moore. The effects of procainamide on conduction in anisotropic canine ventricular myocardium. Circulation 74: 616–625, 1986.
 104. Delgado, C., B. Steinhaus, M. Delmar, D. R. Chialvo, and J. Jalife. Directional differences in excitability and margin of safety for propagation in sheep ventricular epicardial muscle. Circ. Res. 67: 97–110, 1990.
 105. Roberts, D. E., L. T. Hersh, and A. M. Scher. Influence of cardiac fiber orientation on wavefront voltage, conduction velocity, and tissue resistivity in the dog. Circ. Res. 44: 701–712, 1979.
 106. Vander Ark, C. R., and E. W. Reynolds. An experimental study of propagated electrical activity in the canine heart. Circ. Res. 26: 451–460, 1970.
 107. Suenson, M. Interaction between ventricular cells during the early part of excitation in the ferret heart. Acta Physiol. Scand. 125: 81–90, 1985.
 108. Roth, B. J. Action potential propagation in a thick strand of cardiac muscle. Circ. Res. 68: 162–173, 1991.
 109. Henriquez, C. S., A. L. Muzikant, and C. K. Smoak. Anisotropy, fiber curvature, and bath loading effects on activation in thin and thick cardiac tissue preparations: Simulations in a three‐dimensional bidomain model. J. Cardiovasc. Electrophysiol. 7: 424–444, 1996.
 110. Henriquez, C. S. Structure and volume conductor effects on propagation in cardiac tissue. In: Durham, North Carolina, USA: Department of Biomedical Engeneering, Duke University; 1988.
 111. Roth, B. J. The effect of a perfusing bath on the rate of rise of an action potential propagating through a slab of cardiac tissue. Ann. Biomed. Eng. 24: 639–646, 1996.
 112. Sepulveda, N. G., B. J. Roth, and J. P. Wikswo. Current injection into a two‐dimensional anisotropic bidomain. Biophys. J. 55: 987–999, 1989.
 113. Wikswo, J. P., T. A. Wisialowski, W. A. Altemeier, J. R. Balser, H. A. Kopelman, and D. M. Roden. Virtual electrode effects during stimulation of cardiac muscle. Two‐dimensional in vivo experiments. Circ. Res. 68: 513–530, 1991.
 114. Wikswo, J. P., S.‐F. Lin, and R. A. Abbas. Virtual electrode effect in cardiac tissue: a common mechanism for anodal and cathodal stimulation. Biophys. J. 69: 2195–2210, 1995.
 115. Knisley, S. B. Transmembrane voltage changes during unipolar stimulation of rabbit ventricle. Circ. Res. 77: 1229–1239, 1995.
 116. Neunlist, M., and L. Tung. Optical recordings of ventricular excitability of frog heart by an extracellular stimulating point electrode. PACE 17: 1641–1654, 1994.
 117. Wikswo, J. P. Tissue ansiotropy, the cardiac bidomain, and the virtual cathode effect. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. Zipes and J. Jalife. Orlando, FL: W. B. Saunders; 348–361, 1990.
 118. Plonsey, R., C. Henriquez, and N. Trayanova. Extracellular (volume conductor) effect on adjoining cardiac muscle electrophysiology. Med. Biol. Eng. Comp. 26: 126–129, 1987.
 119. Henriquez, C. S. Simulating the electrical behavior of cardiac tissue using the bidomain model. Crit. Rev. Biomed. Eng. 21: 1–77, 1993.
 120. Wu, J. The anatomical basis of anisotropic propagation in cardiac muscle. In: Durham, North Carolina, USA: Department of Biomedical Engineering, Duke University, 1993.
 121. Spach, M. S., J. F. Heidlage, P. C. Dolber, and R. C. Barr. Extracellular discontinuities in cardiac muscle: evidence for capillary effects on the action potential foot. Circ. Res. 83: 1144–1164, 1998.
 122. Spach, M. S., and R. C. Barr. Effects of cardiac microstructure on propagating electrical waveforms. [In Process Citation]. Circ. Res. 86: E23–E28, 2000.
 123. Keith, A., and M. Flack. The form and nature of the muscular connections between the primary dividions of the vertebrate heart. J. Anat. Physiol. 41: 172–189, 1907.
 124. Tranum‐Jensen, J. The fine structure of the sinus node: a survey. In: The sinus node, edited by F. J. M. Bonke. The Hague: Nijhoff; 149–165, 1978.
 125. Masson‐Pévet, M., W. K. Bleeker, A. J. C. Mackaay, L. N. Bouman, and J. M. Houtkooper. Sinus node and atrial cells from the rabbit heart: a quantitative electron microscopic description after electrophysiological localization. J. Mol. Cell. Cardiol. 11: 555–568, 1979.
 126. Bleeker, W. K., A. J. Mackaay, M. Masson‐Pévet, L. N. Bouman, and A. E. Becker. Functional and morphological organization of the rabbit sinus node. Circ. Res. 46: 11–22, 1980.
 127. Opthof, T. The mammalian sinoatrial node. Cardiovasc. Drugs Ther. 1: 573–597, 1988.
 128. James, T. N. The sinus node. Am. J. Cardiol. 40: 965–986, 1977.
 129. Opthof, T., B. de Jonge, A. J. Mackaay, W. K. Bleeker, M. Masson‐Pévet, H. J. Jongsma, and L. N. Bouman. Functional and morphological organization of the guinea‐pig sinoatrial node compared with the rabbit sinoatrial node. J. Mol. Cell Cardiol. 17: 549–564, 1985.
 130. Opthof, T., B. de Jonge, M. Masson‐Pévet, H. J. Jongsma, and L. N. Bouman. Functional and morphological organization of the cat sinoatrial node. J. Mol. Cell Cardiol. 18: 1015–1031, 1986.
 131. Opthof, T., B. de Jonge, H. J. Jongsma, and L. N. Bouman. Functional morphology of the pig sinoatrial node. J. Mol. Cell Cardiol. 19: 1221–1236, 1987.
 132. Davies, M. J. Pathology of atrial arrhythmias. In: M. J. Davies, R. H. Anderson, and A. E. Becker eds. The Conduction System of the Heart. London: Butterworths; 203–215, 1983.
 133. Alings, A. M. W. The aging sino‐atrial node. In: University of Amsterdam. Amsterdam, The Netherlands: University of Amsterdam; 1993.
 134. Wybauw, R. Sur le point d'origine de la systole cardiaque dans l'oreillette droite. Arch. Int. Physiol. 10: 78–89, 1910.
 135. Lewis, T., B. S. Oppenheimer, and A. Oppenheimer. The site of origin of the mammalian heart beat: the pacemaker in the dog heart. Heart 2: 147–169, 1910.
 136. Trautwein, W., and K. Zink. Ueber Membran‐und Aktionspotentiale einzelner Muskelfasern des Kalt‐und Warmblüterherzens. Pflugers Arch. 256: 68–84, 1952.
 137. West, T. C. Ultramicroelectrode recording from the cardiac pacemaker. J. Pharmacol. Exp. Ther. 115: 283–290, 1955.
 138. Yanagihara, K., and H. Irisawa. Inward current activated during hyperpolarization in the rabbit sinoatrial node cell. Pflugers Arch. 385: 11–19, 1980.
 139. DiFrancesco, D., and C. Ojeda. Properties of the current if in the sino‐atrial node of the rabbit compared with those of the current iK, in Purkinje fibres. J. Physiol. (Lond.) 308: 353–367, 1980.
 140. Reuter, H. Ion channels in cardiac cell membranes. Annu. Rev. Physiol. 46: 473–484, 1984.
 141. DiFrancesco, D., A. Ferroni, M. Mazzanti, and C. Tromba. Properties of the hyperpolarizing‐activated current (if) in cells isolated from the rabbit sino‐atrial node. J. Physiol. (Lond.) 377: 61–88, 1986.
 142. Trautwein, W., and K. Uchizono. Electrophysiologic study of the pacemaker in the sino‐atrial node of the rabbit heart. Z. Zellforsch. 61: 96–109, 1963.
 143. Janse, M. J., J. Tranum‐Jensen, A. G. Kléber, and F. J. L. Van Cappelle. Techniques and problems in correlating cellular electrophysiology and morphology in cardiac nodal tissue. In: The Sinus Node, edited by F. J. M. Bonke. The Hague: Nijhoff; 183–194, 1978.
 144. Sano, T., and S. Yamagishi. Spread of excitation from the sinus node. Circ. Res. 16: 423–431, 1965.
 145. Steinbeck, G., M. A. Allessie, F. I. M. Bonke, and W. E. J. P. Lammers. The response of the sinus node to premature stimulation of the atrium studied with microelectrodes in isolated atrial preparations of the rabbit heart. In: The Sinus Node, edited by F. I. M. Bonke. The Hague: Nijhoff; 245–257, 1978.
 146. Bouman, L. N., A. J. C. Mackaay, W. K. Bleeker, and A. E. Becker. Pacemaker shifts in the sinus node. Effects of vagal stimulation, temperature and reduction of extracellular calcium. In: The Sinus Node, edited by F. I. M. Bonke. The Hague: Nijhoff; 245–257, 1978.
 147. Bouman, L. N., and H. J. Jongsma. Structure and function of the SA node: a review. Europ. Heart J. 7: 94–104, 1986.
 148. Noble, D. Discussion on models of entrainment of cardiac cells by R. L. de Haan. In: Cardiac Rate and Rhythm, edited by L. N. Bouman and H. J. Jongsma. The Hague, Boston, New York: Martinus Nijhoff Publishers; 359–361, 1982.
 149. Rook, M. B., B. de Jonge, and H. L. Jongsma. Gap junction formation and functional intercation between neonatal rat cardiocytes in culture. J. Membr. Biol. 118: 179–192, 1990.
 150. Kodama, I., and M. R. Boyett. Regional differences in the electrical activity of the rabbit sinus node. Pflugers Arch. 404: 214–226, 1985.
 151. Kirchhof, C. J., F. I. M. Bonke, M. A. Allessie, and W. E. J. P. Lammers. The influence of the atrial myocardium on impulse formation in the rabbit sinus node. Pflugers Arch. 410: 198–203, 1987.
 152. Joyner, R. W., and F. J. L. van Capelle. Propagation through electrically coupled cells. How a small SA node drives a large atrium. Biophys. J. 50: 1157–1164, 1986.
 153. Optho, T., W. K. Bleeker, M. Masson‐Pévet, H. J. Jongsma, and L. N. Bouman. Little‐excitable transitional cells in the rabbit sinoatrial node: a statistical, morphological and electrophysiological study. Experientia 39: 1099–1101, 1983.
 154. Meek, W. J., and J. A. E. Eyster. Experiments on the origin and propagation of the impulse in the heart. IV. The effect of vagal stimulation and cooling on the location of the pacemaker within the sino‐atrial node. Am. J. Physiol. 34: 368–383, 1914.
 155. Bouman, L. N., E. D. Gerlings, P. A. Biersteker, and F. I. M. Bonke. Pacemaker shift in the sino‐atrial node during vagal stimulation. Pflugers Arch. 302: 255–267, 1968.
 156. Mackaay, A. J. C., T. Opthof, W. K. Bleeker, H. J. Jongsma, and L. N. Bouman. Interaction of adrenaline and acetylcholine on sinus node function. In: Cardiac Rate and Rhythm, edited by L. N. Bouman and H. J. Jongsma. The Hague: Nijhoff; 507–523, 1982.
 157. Cramer, M., M. Siegal, J. T. J. Bigger, and B. F. Hoffman. Characteristics of extracellular potentials recorded from the sinoatrial pacemaker of the rabbit. Circ. Res. 41: 292–300, 1977.
 158. Cramer, M., R. J. Hariman, R. Boxer, and B. F. Hoffman. Electrograms from the canine sinoatrial pacemaker recorded in vitro and in situ. Am. J. Cardiol. 42: 939–946, 1978.
 159. Hariman, R. J., B. F. Hoffman, and R. E. Naylor. Electrical activity from the sinus node region in conscious dogs. Circ. Res. 47: 775–791, 1980.
 160. Hariman, R. J., E. Krongrad, R. A. Boxer, F. O. Bowman, J. R. Malm, and B. F. Hoffman. Methods for recording electrograms from the sino‐atrial node during cardiac surgery in man. Circulation 61: 1024–1029, 1980.
 161. Rijlant, P. The pacemaker of the mammalian heart. J. Physiol. (Lond.) 75: 28P–29P, 1932.
 162. Van der Kooi, M. W., D. Durrer, R. T. Van Dam, and L. H. Van der Tweel. Electrical activity in the sinus node and atrioventricular node. Am. Heart J. 51: 684–700, 1956.
 163. Masuda, M. O., and A. Paes de Carvalho. Sinoatrial transmission and atrial invasion during normal rhythm in the rabbit heart. Circ. Res. 37: 414–421, 1975.
 164. Boineau, J. P., R. B. Schuessler, C. R. Mooney, A. C. Wylds, C. B. Miller, R. D. Hudson, J. M. Borremans, and C. W. Brockus. Multicentric origin of the atrial depolarization wave: the pacemaker complex. Relation to dynamics of atrial conduction, P‐wave changes and heart rate control. Circulation 58: 1036–1048, 1978.
 165. Boineau, J. P., C. B. Miller, R. B. Schuessler, W. R. Roeske, L. J. Autry, A. C. Wylds, and D. A. Hill. Activation sequence and potential distribution maps demonstrating multicentric atrial impulse origin in dogs. Circ. Res. 54: 332–347, 1984.
 166. Schuessler, R. B., J. P. Boineau, and B. I. Bromberg. Origin of the sinus impulse. J. Cardiovasc. Electrophysiol. 7: 263–274, 1996.
 167. Lewis T. Lectures on the Heart: New York: Paul H. Hoeber, London: Shaw and Sons, 1915.
 168. Rothberger, C. G., and D. Scherf. Zur Kenntnis der Erregungsausbreitung vom Sinusknoten auf den Vorhof. Z. Ges. Exp. Med. 53: 792–835, 1926.
 169. Puech, P., M. Esclavissat, D. Sodi‐Pallares, and F. Cineros. Normal auricular activation in the dog's heart. Am. Heart J. 47: 174–191, 1954.
 170. Yamada, K., M. Horiba, Y. Sakaida, M. Okajima, H. Horibe, H. Muraki, T. Kobayashi, A. Miyauchi, H. Oishi, A. Nonogawa, K. Ishikawa, and J. Toyama. Origination and transmission of impulse in the right auricle. Jpn. Heart J. 6: 71–97, 1965.
 171. Spach, M. S., T. D. King, R. C. Barr, D. E. Boaz, M. N. Morrow, and S. Herman‐Giddens. Electrical potential distribution surrounding the atria during depolarization and repolarization in the dog. Circ. Res. 24: 857–873, 1969.
 172. Hamlin, R. L., D. L. Smetzer, T. Senta, and C. R. Smith. Atrial activation paths and P waves in horses. Am. J. Physiol. 219: 306–313, 1970.
 173. Durrer, D., R. T. Van Dam, G. E. Freud, M. J. Janse, F. L. Meijler, and R. C. Arzbaecher. Total excitation of the isolated humane heart. Circulation 41: 895–912, 1970.
 174. Goodman, D., A. B. M. van der Steen, and R. T. van Dam. Endocardial and epicardial activation pathways of the canine right atrium. Am. J. Physiol. 220: 1–11, 1971.
 175. Spach, M. S., M. Lieberman, J. G. Scott, R. C. Barr, E. A. Johnson, and J. M. Kootsey. Excitation sequences of the atrial septum and the AV node in isolated hearts of the dog and rabbit. Circ. Res. 29: 156–172, 1971.
 176. Janse, M. J., and R. H. Anderson. Specialized internodal atrial pathways: fact or fiction?. Eur. J. Cardiol. 2: 117–136, 1974.
 177. Wittig, J. H., M. R. de Leval, and G. Stark. Intraoperative mapping of atrial activation before, during and after the Mustard operation. J. Thorac. Cardiovasc. Surg. 73: 1–13, 1977.
 178. Sherf, L., and T. N. James. Fine structure of cells and their histologic organization within internodal pathways of the heart: clinical and electrocardiographic implications. Am. J. Cardiol. 44: 345–369, 1979.
 179. Tranum‐Jensen, J., and M. J. Janse. Fine structural identification of individual cells subjected to microelectrode recording in perfused cardiac preparations. J. Mol. Cell. Cardiol. 14: 233–247, 1982.
 180. Moore, E. N., S. L. Jomain, J. H. Stuckey, J. W. Buchanan, and B. F. Hoffman. Studies on ectopic atrial rhythms in dogs. Am. J. Cardiol. 19: 676–685, 1967.
 181. Moore, E. N., J. Melbin, J. F. Spear, and J. D. Hill. Sequence of atrial excitation in the dog during antegrade and retrograde activation. J. Electrocardiol. 4: 283–290, 1971.
 182. Waldo, A. L., K. J. Vittikainen, and B. F. Hoffman. The sequence of retrograde atrial activation in the canine heart: correlation with positive and negative retrograde p waves. Circ. Res. 37: 156–163, 1975.
 183. Janse, M. J., R. H. Anderson, M. A. McGuire, and S. Y. Ho. “AV‐nodal” reentry: Part I: “AV‐nodal” reentry revisited. J. Cardiovasc. Electrophysiol. 4: 561–572, 1993.
 184. Scherf, D., and J. Cohen. The Atrioventricular Node and Selected Cardiac Arrhythmias. New York: Grune and Stratton; 1964.
 185. Tawara, S. Das Reizleitungssystem des Säugetierherzens. Eine anatomisch‐histologissche Studie über das Atrioventrikularbündel und die Purkinjeschen Fäden. Jena: Fischer; 1906.
 186. Anderson, R. H. Histologic and histochemical evidence concerning the presence of morphologically distinct cellular zones within the rabbit atrioventricular node. Anat. Rec. 173: 7–23, 1972.
 187. Woods, W. T., L. Sherf, and T. N. James. Structure and function of specific regions in the canine atrioventricular node. Am. J. Physiol. 243 (Heart Circ. Physiol. 12): H41–H50, 1982.
 188. Paes de Carvalho, A. and D. F. de Almeida. Spread of activity through the atrioventricular node. Circ. Res. 8: 801–809, 1960.
 189. Becker, A. E., and R. H. Anderson. Morphologyy of the human atrioventricular junctional area. In: H. J. J. Wellens, K. I. Lie M. J. Janse, eds. The Conduction System of the Heart: Structure, Function and Clinical Implication, edited by H. J. J. Wellens, K. I. Lie, and M. J. Janse. Philadelphia: Lea and Febiger; 263–286, 1976.
 190. Kawamura, K., and T. N. James. Comparative ultrastructure of cellular junctions in working myocardium and the conduction system under normal and pathological conditions. J. Mol. Cell. Cardiol. 3: 31–60, 1972.
 191. Marino, T. A. The atrioventricular bundle in the ferret heart. A light and quantitative electron microscopic study. Am. J. Anat. 154: 365–392, 1979.
 192. Thaemert, J. C. Fine structure of the atrioventricular node as viewed in serial sections. Am. J. Anat. 136: 43–66, 1973.
 193. Billette, J. Atrioventricular nodal activation during premature stimulation of the atrium. Am. J. Physiol. 252 (Heart Circ. Physiol. 21): H163–H177, 1987.
 194. Anderson, R. H., M. J. Janse, F. J. L. Van Capelle Billette, J., A. E. Becker, and D. Durrer. A combined morphological and electrophysiological study of the atrioventricular node of the rabbit heart. Circ. Res. 35: 909–922, 1974.
 195. Nagata, F. An experimental study on the conduction of excitation in the A‐V nodal region. Jpn. Circ. J. 30: 1507–1527, 1966.
 196. Takayasu, M., Y. Tateishi, H. Tamai, J. Kanazu, T. Nagata, and K. Kawamura. Conduction of excitation in the A‐V nodal region. In: T. Sano, V. Mizuhira, and K. Matsuda, eds. Electrophysiology and Ultrastructure of the Heart, edited by J. Jans, V. Mizuhira, and K. Matsuda. New York: Grune and Stratton; 143–152, 1967.
 197. McGuire, M. A., J. M. T. De Bakker, J. T. Vermeulen, A. F. Moorman, P. Loh, B. Thibault, J. L. M. Vermeulen, A. E. Becker, and M. J. Janse. Atrioventricular junctional tissue. Discrepancy between histological and electrophysiological characteristics. Circulation 94: 571–577, 1996.
 198. McGuire, M. A., J. M. T. De Bakker, J. T. Vermeulen, T. Opthof, A. E. Becker, and M. J. Janse. Origin and significance of double potentials near the atrioventricular node. Correlation of extracellular potentials, intracellular potentials, and histology. Circulation 89: 2351–2360, 1994.
 199. Billette, J., M. J. Janse, F. J. L. Van Capelle, R. H. Anderson, P. Touboul, and D. Durrer. Cycle‐length‐dependent properties of AV nodal activation in rabbit hearts. Am. J. Physiol. 231: 1129–1139, 1976.
 200. Janse, M. J., F. J. L. Van Capelle, R. H. Anderson, P. Touboul, and J. Billette. Electrophysiology and structure of the atrioventricular node of the rabbit heart. In: The Conduction System of the Heart, edited by H. J. J. Wellens, K. I. Lie, and M. J. Janse. Leiden: Stenfert Kroese; 296–315, 1976.
 201. Van Capelle, F. J. L., M. J. Janse, P. J. Varghese, G. E. Freud, C. Mater, and D. Durrer. Spread of excitation in the atrioventricular node of isolated rabbit hearts studied by multiple microelectrode recording. Circ. Res. 31: 602–616, 1972.
 202. Watanabe, Y., and L. S. Dreifus. Sites of impulse formation within the atrioventricular junction of the rabbit. Circ. Res. 22: 717–727, 1968.
 203. McGuire, M. A., M. J. Janse, and D. L. Ross. “AV nodal” re‐entry: Part II: AV nodal, AV junctional, or atrionodal reentry?. J. Cardiovasc. Electrophysiol. 4: 573–586, 1993.
 204. Akhtar, M., M. R. Jazayeri, J. Sra, Z. Blanck, S. Deshpande, and A. Dhala. Atrioventricular nodal reentry: clinical, electrophysiological, and therapeutic considerations. Circulation 88: 282–295, 1993.
 205. Mines, G. R. On dynamic equilibrium in the heart. J. Physiol. (Lond.) 46: 349–382, 1913.
 206. White, P. D. A study of atrioventricular rhythm following auricular flutter. Arch. Intern. Med. 16: 517–535, 1915.
 207. Scherf, D., and C. Shookhoff. Experimentelle Untersuchungen ueber die “Umkehr‐Extrasystole” (reciprocating beats). Wien. Arch. Inn. Med. 12: 501–529, 1926.
 208. Moe, G. K., J. B. Preston, and H. J. Burlington. Physiologic evidence for a dual A‐V transmission system. Circ. Res. 4: 357–375, 1956.
 209. Rosenblueth A. Ventricular “echoes”. Am. J. Physiol. 195: 53–60, 1958.
 210. Kistin, A. D. Atrial reciprocating rhythm. Circulation 32: 687–707, 1965.
 211. Puech, P. La conduction reciproque par le noeud de Tawara. Bases experimentales et aspects cliniques. Ann. Cardiol. Angiol. 19: 21–40, 1970.
 212. Schuilenburg, R. M., and D. Durrer. Atrial echo beats in the human heart elicited by induced atrial premature beats. Circulation 37: 680–693, 1968.
 213. Schuilenburg, R. M., and D. Durrer. Ventricular echo beats in the human heart elicited by induced ventricular premature beats. Circulation 40: 337–347, 1969.
 214. Bigger J. T. and B. N. Goldreyer. The mechanism of supraventricular tachycardia. Circulation 42: 673–688, 1970.
 215. Mendez, C., J. Han, P. D. Garcia de Jalon, and G. K. Moe. Some characteristics of ventricular echoes. Circ. Res. 16: 562–581, 1965.
 216. Mignone, R. J., and A. G. Wallace. Ventricular echoes. Evidence for dissociation of conduction and reentry within the A‐V node. Circ. Res. 19: 638–649, 1966.
 217. Mendez, C., and G. K. Moe. Demonstration of a dual AV nodal conduction system in the isolated rabbit heart. Circ. Res. 19: 378–393, 1966.
 218. Moe, G. K., W. Cohen, and R. L. Vick. Experimentally induced paroxysmal A‐V nodal tachycardia in the dog. Am. Heart J. 65: 87–92, 1963.
 219. Janse, M. J., F. J. L. Van Capelle, G. E. Freud, and D. Durrer. Circus movement within the A‐V node as a basis of supraventricular tachycardia as shown by multiple microelectrode recording in the isolated rabbit heart. Circ. Res. 28: 403–414, 1971.
 220. Wit, A. L., B. N. Goldreyer, and A. N. Damato. An in vitro model of paroxysmal supraventricular tachycardia. Circulation 43: 862–875, 1971.
 221. Coumel, P., C. Cabrol, A. Fabiato, R. Gourgon, and R. Slama. Tachycardie permanente par rythme reciproque. Arch. Mal. Coeur Vaiss. 60: 1830–1864, 1967.
 222. Casta, A., G. Wolff, A. Mehta, D. Tamer, O. L. Garcia, A. S. Pickoff, P. L. Ferrer, R. J. Sung, and H. Gelband. Dual atrioventricular nodal pathways. A benign finding in arrhythmiafree children with heart disease. Am. J. Cardiol. 46: 1013–1018, 1980.
 223. Denes, P., D. Wu, R. C. Dhingra, E. AmatyLeon, C. R. C. Wyndham, and K. M. Rosen. Dual A‐V nodal pathways. A common electrophysiological response. Br. Heart J. 37: 1069–1076, 1975.
 224. Denes, P., D. Wu, R. C. Dhingra, R. Chuquimia, and K. Rosen. Demonstration of dual A‐V nodal pathways in patients with paroxysmal supraventricular tachycardia. Circulation 48: 549–555, 1973.
 225. McGuire, M. A., J. P. Bourke, M. C. Robotin, I. C. Johnson, W. Meldrum‐Hanna, G. R. Nunn, J. B. Uther, and D. L. Ross. High resolution mapping in Koch's triangle using sixty electrodes in humans with atrioventricular junctional (AV nodal) reentrant tachycardia. Circulation 88: 2315–2328, 1993.
 226. Sung, R. J., H. L. Waxman, S. Saksena, Z. Juma. Sequence of retrograde atrial activation in patients with dual atrioventricular nodal pathways. Circulation 64: 1059–1067, 1981.
 227. Ross, D. L., D. C. Johnson, A. R. Denniss, M. J. Cooper, D. A. Richards, and J. B. Uther. Curative surgery for atrioventricular junctional (“AV nodal”) reentrant tachycardia. J. Am. Coll. Cardiol. 6: 1383–1392, 1985.
 228. Jackman, W. M., K. J. Beckman, J. H. McClelland, X. Wang, K. J. Friday, C. A. Roman, K. P. Moulton, N. Twidale, A. Hazlitt, M. I. Prior, J. Oren, E. D. Overholt, and R. Lazzara. Treatment of supraventricular tachycardia due to atrioventricular nodal reentry by radiofrequency ablation of slow‐pathway conduction. N. Engl. J. Med. 327: 313–318, 1992.
 229. Haissaguerre, M., F. Gaita, B. Fischer, D. Commenges, P. Montserrat, P. d'lvernois, P. Lemetayer, and J. Warin. Elimination of atrioventricular nodal reentrant tachycardia using discrete slow potentials to guide application of radiofrequency energy. Circulation 85: 2162–2175, 1992.
 230. Cox, J. L., W. L. Holman, and M. E. Cain. Cryosurgical treatment of atrioventricular node reentrant tachycardia. Circulation 76: 1329–1336, 1987.
 231. Ho, S. Y., J. M. McComb, C. D. Scott, and R. H. Anderson. Morphology of the cardiac conduction system in patients with electrophysiologically proven dual atrioventricular pathways. J. Cardiovasc. Electrophysiol. 4: 504–512, 1993.
 232. Loh, P., J. M. de Bakker, M. Hocini, B. Thibault, R. N. Hauer, and M. J. Janse. Reentrant pathway during ventricular echoes is confined to the atrioventricular node: high‐resolution mapping and dissection of the triangle of Koch in isolated, perfused canine hearts. Circulation 100: 1346–1353, 1999.
 233. Medkour, D., A. E. Becker, K. Khalife, and J. Billette. Anatomic and functional characteristics of a slow posterior AV nodal pathway: role in dual‐pathway physiology and reentry. Circulation 98: 164–174, 1998.
 234. Lin, L. J., J. Billette, K. Khalife, K. Martel, J. Wang, and D. Medkour. Characteristics, circuit, mechanism, and ablation of reentry in the rabbit atrioventricular node. J. Cardiovasc. Electrophysiol. 10: p954–964, 1999.
 235. Mazgalev, T., and P. Tschou. Atrial‐AV nodal electrophysiology: a view from the millenium. Armonk, NY: Futura Publishing Company; 2000.
 236. Zipes, D. The atrioventricular node: a riddle wrapped in a mistery inside an enigma. In: T. Mazgalev, P. Tschou eds. Atrial‐AV Nodal Electrophysiology: A View from the Millennium. Armonk, NY: Futura Publishing Company; 2000.
 237. Mazgalev, T., and P. Tschou. The AV nodal dual pathway electrophysiology: still a controversial concept. In: Atrial‐AV Nodal Electrophysiology: A View from the Millennium, edited by J. Mazgalev and P. Tschou. Armonk, NY: Futura Publishing Company; 2000.
 238. Bukauskas F. F., and R. P. Veteikis. Passive electrical properties of the atrioventricular region of the rabbit heart. Biofizika 22: 499–504, 1977.
 239. De Mello, W. C. Passive electrical properties of the atrioventricular node. Pflugers Arch. 371: 135–139, 1977.
 240. Ikeda, N., J. Toyama, T. Shimizu, I. Kodama, and K. Yamada. The role of electrical uncoupling in the genesis of atrioventricular conduction disturbance. J. Mol. Cell Cardiol. 12: 809–826, 1980.
 241. Kokubun, S., M. Nishimura, A. Noma, and H. Irisawa. Membrane currents in the rabbit atrioventricular node cell. Pflugers Arch. 393: 15–22, 1982.
 242. Cranefield, P. E., B. F. Hoffman, and A. Paes de Carvalho. Effects of acetylcholine on single fibers of the atrio‐ventricular node. Circ. Res. 7: 19–23, 1959.
 243. Janse, M. J. Influence of the direction of the atrial wave front on A‐V nodal transmission in isolated hearts of rabbits. Circ. Res. 25: 439–449, 1969.
 244. Mazgalev, T., L. S. Dreifus, H. Iinuma, and E. L. Michelson. Effects of the site and timing of atrio‐ventricular nodal input on atrio‐ventricular conduction in the isolated perfused rabbit heart. Circulation 70: 748–759, 1984.
 245. Zipes, D. P., C. Mendez, and G. K. Moe. Evidence for summation and voltage dependency in rabbit atrioventricular nodal fibers. Circ. Res. 32: 170–177, 1973.
 246. Hoffman, B. F. Physiology of atrioventricular transmission. Circulation 24: 506–517, 1961.
 247. Mendez, C. Characteristics of impüulse propagation in the mammalian atrioventricular node. In: Normal and Abnormal Conduction in the Heart, edited by A. Paes de Carvalho, B. F. Hoffman, and M. Lieberman. Mount Kisco, NY: Futura; 363–377, 1982.
 248. Noma, A., H. Irisawa, S. Kokobun, H. Kotake, M. Nishimura, and Y. Watanabe. Slow current systems in the A‐V node of the rabbit heart. Nature 285: 228–229, 1980.
 249. Wit, A. L., and P. F. Cranefield. Effect of verapamil on the sinoatrial and atrioventricular nodes of the rabbit and the mechanism by which it arrests reentrant atrioventricular tachycardias. Circ. Res. 35: 413–425, 1974.
 250. Zipes, D. P. and C. Mendez. Action of manganese ions and tetrodotoxin on atrioventricular nodal transmembrane potentials in isolated rabbit hearts. Circ. Res. 32: 447–454, 1973.
 251. Akiyama, T., and H. A. Fozzard. Ca and Na selectivity of the active membrane of rabbit AV nodal cells. Am. J. Physiol. 236 (Cell Physiol. 5): C1–C8, 1979.
 252. Van Capelle, F. J., and M. J. Janse. Influences of geometry on the shape of the propagated action potential. In: The Conduction System of the Heart, edited by H. J. J. Wellens, K. I. Lie, and M. J. Janse. Leiden: Stenfert Kroese; 316–335, 1976.
 253. Shigeto, N., and H. Irisawa. Slow conduction in the atrioventricular node of the cat: a possible explanation. Experientia 28: 1442–1443, 1972.
 254. Ruiz‐Ceretti, I., and A. Ponce Zumino. Action potential changes under varied (Na+) and (Ca2+) indicating the existence of two inward currents in cells of the rabbit atrioventricular node. Circ. Res. 39: 326–336, 1976.
 255. Gettes, L. S. and H. Reuter. Slow recovery from inactivation of inward currents in mammalian myocardial fibres. J. Physiol. (Lond.) 240: 703–724, 1974.
 256. Maglaveras, N., F. J. L. van Cappelle, J. M. T. de Bakker, C. Pappas, and M. J. Janse. Activation delay in healed myocardial infarction: a comparison between model and experiment. Am. J. Physiol. 269 (Heart Circ. Physiol. 38): H1441–1449, 1995.
 257. Truex, R. C. Comparative anatomy and functional considerations of the cardiac conduction system. In: A. Paes de Carvalho, W. C. de Mello, and B. F. Hoffman, eds. The Specialized Tissues of the Heart, edited by A. Paes de Carvalho, W. C. de Mello, and B. F. Hoffman. Amsterdam: Elsevier; 22–43, 1961.
 258. Hoffman, B. F., and P. F. Cranefield. Electrophysiology of the Heart. New York: McGraw‐Hill; 1960.
 259. Demoulin, G. C., and H. E. Kulbertus. Histopathological examination of concept of left hemiblock. Br. Heart J. 34: 807–814, 1972.
 260. Truex, R. C., and M. Q. Smythe. Comparative morphology of the cardiac conduction tissue in animals. Ann. N.Y. Acad. Sci. 127: 19–33, 1965.
 261. Myerburg, R. J., K. Nilsson, and H. Gelband. Physiology of canine intraventricular conduction and endocardial excitation. Circ. Res. 30: 217–243, 1972.
 262. Van Dam, T. T., and M. J. Janse. Activation of the heart. In: Comprehensive Electrocardiology, edited by P. W. MacFarlane, and T. D. Veitch Lawrie. New York: Pergamon Press; 101–128, 1988.
 263. Rosenbaum, M. B., M. V. Elizari, and J. O. Lazzari. The Hemiblocks: New Concepts of Intraventricular Conduction Based on Human Anatomical, Physiological and Clinical Studies. Oldsmar: Tampa Tracings; 1970.
 264. Veenstram, R. D., R. W. Joyner, and D. A. Rawling. Purkinje and ventricular activation sequences of canine papillary muscle. Effects of quinidine and calcium on the Purkinje‐ventricular conduction delay. Circ. Res. 54: 500–515, 1984.
 265. Overholt, E. D., R. W. Joyner, R. D. Veenstra, D. Rawling, and R. Wiedmann. Unidirectional block between Purkinje and ventricular layers of papillary muscle. Am. J. Physiol. 247 (Heart Circ. Physiol. 16): H584–H595, 1984.
 266. Mendez, C., W. J. Mueller, and X. Urguiaga. Propagation of impulses across the Purkinje fiber‐muscle junctions in the dog heart. Circ. Res. 36: 135–150, 1970.
 267. Rawling, D. A., R. W. Joyner, and E. D. Overholt. Variations in the functional electrical coupling between the subendocardial Purkinje and ventricular layers of the canine left ventricle. Circ. Res. 57: 252–261, 1985.
 268. Alanis, J., D. Benitez, and G. Pilar. A functional discontinuity between the Purkinje and ventricular muscle cells. Acta Physiol. Latino Am. 11: 171–183, 1961.
 269. Alanis, J., and D. Benitez. Transitional potentials and the propagation of impulses through different cardiac cells. In: Electrophysiology and Ultrastructure of the Heart, edited by T. Sano, V. Misuhira, and K. Matsuda. New York: Grune & Stratton; 153–175, 1967.
 270. Matsuda, K., A. Kamiyama, and T. Hoshi. Configuration of the transmembrane action potential of the Purkinje‐ventricular fiber junction and its analysis. In: Electrophysiology and Ultrastructure of the Heart, edited by T. Sano, V. Misuhira, and K. Matsuda. New York: Grune & Stratton; 177–187, 1967.
 271. Martinez‐Palomo, A., J. Alanis, and D. Benitez. Transitional cardiac cells of the conductive system of the dog heart. Distinguishing morphological and electrophysiological features. J. Cell Biol. 47: 1–17, 1970.
 272. Tranum Jensen, J., A. A. Wilde, J. T. Vermeulen, and M. J. Janse. Morphology of electrophysiologically identified junctions between Purkinje fibers and ventricular muscle in rabbit and pig hearts. Circ. Res. 69: 429–437, 1991.
 273. Scher, A. M. and A. C. Young. Ventricular depolarization and the genesis of QRS. Ann. N. Y. Acad. Sci. 65: 766–778, 1957.
 274. Nagao, K., J. Toyama, I. Kodama, and K. Yamada. Role of the conduction system in the endocardial excitation spread in the right ventricle. Am. J. Cardiol. 48: 864–870, 1981.
 275. Wang, Y., Y. Rudy Action potential propagation in inhomogeneous cardiac tissue: safety factor considerations and ionic mechanism. Am. J. Physiol. 278 (Heart Circ. Physiol. 47): H1019–H1029, 2000.
 276. Dominguez, G., and H. A. Fozzard. Influence of extracellular K+ concentration on cable properties and excitability of sheep cardiac Purkinje fibers. Circ. Res. 26: 565–574, 1970.
 277. Han, J., A. M. Malozzi, and G. K. Moe. Transient ventricular conduction disturbances produced by intra‐atrial injection of single doses of KCl. Circ. Res. 21: 3–8, 1967.
 278. Antoni, H., and T. Zerweck. Besitzen die sympathischen Uberträgerstoffe einen direkten Einfluss auf die Leitungsgeschwindigkeit des Säugetiermyokards. Pflugers Arch. 293: 310–330, 1967.
 279. Spear, J. E., and E. N. Moore. Supernormal excitability and conduction in the His‐Purkinje system of the dog. Circ. Res. 35: 782–792, 1974.
 280. Spear, J. F., and E. N. Moore. Supernormal conduction in the canine bundle of His and proximal bundle branches. Am. J. Physiol. 238 (Heart Circ. Physiol. 7): H300–H306, 1980.
 281. Peon, J., G. R. Ferrier, and G. K. Moe. The relationship of excitability to conduction velocity in canine Purkinje tissue. Circ. Res. 43: 125–135, 1978.
 282. Weidmann, S. Effects of calcium ions and local anaesthetics on electrical properties of Purkinje fibers. J. Physiol. (Lond.) 129: 568–582, 1955.
 283. Corrado, G., R. J. Levi, G. J. Nau, and M. B. Rosenbaum. Paroxysmal atrioventriculasr block related to phase 4 bilateral bundle branch block. Am. J. Cardiol. 33: 553–556, 1974.
 284. Elizari, M. V., G. J. Nau, R. J. Levi, J. O. Lazzari, M. S. Halpern, and M. B. Rosenbaum. Experimental production of rate‐dependent bundle branch block in the canine heart. Circ. Res. 34: 730–742, 1974.
 285. Singer, D. H., R. Lazzara, and B. F. Hoffman. Interrelationships between automaticity and conduction in Purkinje fibers. Circ. Res. 21: 537–558, 1967.
 286. Jalife, J., C. Antzelevitch, V. Lamanna, and G. K. Moe. Rate‐dependent changes in excitability of depressed cardiac Purkinje fibers as a mechanism of intermittent bundle branch block. Circulation 67: 912–922, 1983.
 287. Kléber, A. G., M. J. Janse, F. J. G. Wilms‐Schopmann, A. A. M. Wilde, and R. Coronel. Changes in conduction velocity during acute ischemia in ventricular myocardium of the isolated porcine heart. Circ. Res. 73: 189–198, 1986.
 288. Kishida, H., B. Surawicz, and L. T. Fu. Effects of K+ and K+‐ induced depolarization on (dV/dt)max, threshold potential, and membrane input resistance in guinea pig and cat ventricular myocardium. Circ. Res. 44: 800–814, 1979.
 289. Whalley, D. W., D. J. Wendt, C. F. Starmer, Y. Rudy, and A. O. Grant. Voltage‐independent effects of extracellular K+ on the Na+ current and phase 0 of the action potential in isolated cardiac myocytes. Circ. Res. 75: 491–502, 1994.
 290. Noma, A. ATP‐regulated K channels in cardiac muscle. Nature 305: 147–148, 1983.
 291. Grant, A. O., C. F. Starmer, and H. C. Strauss. Antiarrhythmic drug action: blockade of inward sodium current. Circ. Res. 55: 427–439, 1984.
 292. Janse, M. J., and A. L. Wit. Electrophysiological mechanisms of ventricular arrhythmias resulting from ischemia and infarction. Physiol. Rev. 69: 1049–1169, 1989.
 293. Janse, M. J., and A. G. Kléber. Electrophysiological changes and ventricular arrhythmias in the early phase of regional myocardial ischemia. Circ. Res. 49: 1069–1081, 1981.
 294. Schütz, E. Electrophysiologie des Herzens bei einphasischer Ableitung. Ergebn. Physiol. 38: 493–620, 1936.
 295. Moréna, H., M. J. Janse, J. W. T. Fiolet, W. J. G. Krieger, H. Crijns, and D. Durrer. Comparison of the effects of regional ischemia, hypoxia, hyperkalemia, and acidosis on intracellular and extracellular potentials and metabolism in the isolated porcine heart. Circ. Res. 46: 635–646, 1980.
 296. Downar, E., M. J. Janse, and D. Durrer. The effect of acute coronary artery occlusion on subepicardial transmembrane potentials in the intact porcine heart. Circulation 56: 217–224, 1977.
 297. Kodama, I., A. A. M. Wilde, M. J. Janse, D. Durrer, and K. Yamada. Combined effects of hypoxia, hyperkalemia and acidosis on membrane action potential and excitability of guinea‐pig ventricular muscle. J. Mol. Cell Cardiol. 16: 247–259, 1984.
 298. Nakayama, T., Y. Kurachi, A. Noma, and H. Irisawa. Action potential and membrane currents of single pacemaker cells of the rabbit heart. Pflugers Arch. 402: 248–257, 1984.
 299. Gomez, J. P., J. E. Potreau, J. E. Branka, and G. Raymond. Developmental changes in Ca2+ current from newborn rat cardiomyocytes in primary culture. Pflugers Arch. 428: 241–249, 1994.
 300. Cranefield, P. F., H. O. Klein, and B. . Hoffman. Conduction of the cardiac impulse. 1. Delay, block, and one‐way block in depressed Purkinje fibers. Circ. Res. 28: 199–219, 1971.
 301. Cranefield, P. E., A. L. Wit, and B. F. Hoffman. Conduction of the cardiac impulse. 3. Characteristics of very slow conduction. J. Gen. Physiol. 59: 227–246, 1972.
 302. Wit, A. L., B. F. Hoffman, and P. F. Cranefield. Slow conduction and reentry in the ventricular conducting system. I. Return extrasystole in canine Purkinje fibers. Circ. Res. 30: 1–10, 1972.
 303. Weingart, R. Electrical properties of the nexal membrane studied in rat ventricular cell pairs. J. Physiol. (Lond.) 370: 267–284, 1986.
 304. Weingart, R., and P. Maurer. Action potential transfer in cell pairs isolated from adult rat and guinea pig ventricles. Circ. Res. 63: 72–80, 1988.
 305. Rohr, S., J. P. Kucera, and A. G. Kleber. Slow conduction in cardiac tissue: I. Effects of a reduction of excitability vs. a reduction in cell‐to‐cell coupling on microconduction. Circ. Res. 83: 781–794, 1998.
 306. Streit, J. Effects of hypoxia and glycolytic inhibition on electrical properties of sheep cardiac Purkinje fibers. J. Mol. Cell. Cardiol. 19: 875–885, 1987.
 307. Wojtczak, J. Contractures and increase in internal longitudinal resistance of cow ventricular muscle induced by hypoxia. Circ. Res. 44: 88–95, 1979.
 308. Kaplinsky, E., S. Ogawa, C. W. Balke, and L. S. Dreifus. Two periods of early ventricular arrhythmia in the canine acute myocardial infarction model. Circulation 60: 397–403, 1979.
 309. Smith, W. T., W. F. Fleet, T. A. Johnson, C. L. Engle, and W. E. Cascio. The 1b phase of ventricular arrhythmias in ischemic in situ porcine heart is related to changes in cell‐to‐cell coupling. Circulation 92: 3051–3060, 1995.
 310. Starmer, C. F., V. N. Biktashev, D. N. Romashko, M. R. Stepanov, O. N. Makarova, and V. I. Krinsky. Vulnerability in an excitable medium: analytical and numerical studies of initiating unidirectional propagation. Biophys. J. 65: 1775–1787, 1993.
 311. Quan, W., and Y. Rudy. Induced unidirectional block and reentry of cardiac excitation. Proc. 9th Annu. Conf. IEEE Eng. 1: 210–211, 1987.
 312. Quan, W., and Y. Rudy. Unidirectional block and reentry of cardiac excitation: a model study. Circ. Res. 66: 367–382, 1990.
 313. Van, Capelle F. J. L. and D. Durrer. Computer simulation of arrhythmias in a network of coupled excitable elements. Circ. Res. 47: 453–466, 1980.
 314. Janse, M. J. The effects of changes of heart rate on the refractory period of the heart. In:. Amsterdam: University of Amsterdam; 1971.
 315. Han, J., and G. K. Moe. Nonuniform recovery of excitability of ventricular muscle. Circ. Res. 14: 44–60, 1964.
 316. Kuo, C. S., K. Munakata, C. P. Reddy, and B. Surawicz. Characteristics and possible mechanism of ventricular arrhythmia dependent on the dispersion of action potential durations. Circulation 67: 1356–1367, 1983.
 317. Wallace, A. G., and R. J. Mignone. Physiologic evidence concerning the re‐entry hypothesis for ectopic beats. Am. Heart J. 72: 60–70, 1966.
 318. Allessie, M. A., F. I. M. Bonke, and F. J. C. Schopman. Circus movement in rabbit atrial muscle as a mechanism of tachycardia. II. The role of nonuniform recovery of excitability in the occurrence of unidirectional block as studied with multiple microelectrodes. Circ. Res. 39: 168–177, 1976.
 319. Engelmann, T. W. Über die reziproke und irreziproke Reizleitung mit besonderer Beziehung auf das Herz. Pflugers Arch. 61: 272–284, 1895.
 320. Schmitt, F. O., and J. Erlanger. Directional differences in the conduction of the impulse through heart muscle and their possible relation to extrasystolic anf fibrillatory contractions. Am. J. Physiol. 87: 326–347, 1928.
 321. Downar, E., and M. B. Waxman. Depressed conduction and unidirectional block in Purkinje fibers. In: H. J. J. Wellens, K. I. Lie, and M. J. Janse. eds. The Conduction System of the Heart, Philadelphia: Lea & Febiger; 393–409, 1976.
 322. Waxman, M. B., E. Downar, and R. W. Wald. Unidirectional block in Purkinje fibers. Can. J. Physiol. Pharmacol. 58: 925–933, 1980.
 323. Engelmann T. W. Versuche über die irreziproke Reizleitung in Muskelfasern. Pflugers Arch. 62: 400–414, 1896.
 324. Spach, M. S., W. T. Miller, P. C. Dolber, J. M. Kootsey, J. R. Sommer, and C. E. J. Mosher. The functional role of structural complexities in the propagation of depolarization in the atrium of the dog. Cardiac conduction disturbances due to discontinuities of effective axial resistivity. Circ. Res. 50: 175–191, 1982.
 325. De Bakker, J. M. T., F. J. L. Van Capelle, M. J. Janse, S. Tasseron, J. T. Vermeulen, N. Dejonge, and J. R. Lahpor. Slow conduction in the infarcted human heart—zigzag course of activation. Circulation 88: 915–926, 1993.
 326. Rohr, S., and B. M. Salzberg. Characterization of impulse propagation at the microscopic level across geometrically defined expansions of excitable tissue: multiple site optical recording of transmembrane voltage (MSORTV) in patterned growth heart cell cultures. J. Gen. Physiol. 104: 287–309, 1994.
 327. Rohr, S., J. P. Kucera, V. G. Fast, and A. G. Kleber. Paradoxical improvement of impulse conduction in cardiac tissue by partial cellular uncoupling. Science 275: 841–844, 1997.
 328. Spach, M. S., P. C. Dolber, and J. F. Heidlage. Influence of the passive anisotropic properties on directional differences in propagation following modification of the sodium conductance in human atrial muscle. A model of reentry based on anisotropic discontinuous propagation. Circ. Res. 62: 811–832, 1988.
 329. Delmar, M., D. C. Michaels, T. Johnson, and J. Jalife. Effects of increasing intercellular resistance on transverse and longitudinal propagation in sheep epicardial muscle. Circ. Res. 60: 780–785, 1987.
 330. Schalij, M. J. Anisotropic conduction and ventricular tachycardia. In:. Maastricht, The Netherlands: Rijksuniversiteit Limburg; 1988.
 331. Girouard, S. D., J. M. Pastore, K. R. Laurita, K. W. Gregory, and D. S. Rosenbaum. Optical mapping in a new guinea pig model of ventricular trachycardia reveals mechanisms for multiple wavelengths in a single reentrant circuit. Circulation 93: 603–613, 1996.
 332. Cabo, C. A. M. Pertsov, W. T. Baxter, J. M. Davidenko, R. A. Gray, and J. Jalife. Wave‐front curvature as a cause of slow conduction and block in isolated cardiac muscle. Circ. Res. 75: 1014–1028, 1994.
 333. De la Fuente, D., B. Sasyniuk, and G. K. Moe. Conduction through a narrow isthmus in isolated canine atrial tissue. A model of the W‐P‐W syndrome. Circulation 44: 803–809, 1971.
 334. Lin, S. F., B. J. Roth, D. S. Echt, and J. P. Wikswo. Complex dynamics following unipolar stimulation during the vulnerable phase. Circulation 94: I–174, 1996.
 335. Saypol, J. M., and B. J. Roth. A mechanism for anisotropic reentry in electrically active tissue. J. Cardiovasc. Electrophysiol. 3: 558–566, 1992.
 336. Efimov, I. R., Y. N. Cheng, W. D. Van, T. N. Mazgalev, and P. J. Tchou. Virtual electrodes induced phase singularity: a basic mechanism of defibrillation failure. Circ Res. 82: 918–925, 1998.
 337. Mayer, A. G. Nerve conduction and other reactions in Cassiopea. Am. J. Physiol. 39: 375–393, 1916.
 338. Mayer, A. G. Rhythmical pulsation in scyphomedusae. II. In: Papers from the Marine Biological Laboratory at Tortugas. Washington; 115–131, 1908.
 339. Mines, G. R. On circulating excitations in heart muscles and their possible relation to tachycardia and fibrillation. Trans. R. Soc. Can. Sect. IV: 43–52, 1914.
 340. Durrer, D., and J. P. Roos. Epicardial excitation of the ventricles in a patient with Wolff‐Parkinson‐White syndrome (type B). Circulation 35: 15–21, 1967.
 341. Durrer, D., L. Schoo, R. M. Schuilenburg, and H. J. J. Wellens. The role of premature beats in the initiation and termination of supraventricular tachycardia in Wolff‐Parkinsin‐White syndrome. Circulation 36: 644–662, 1967.
 342. Holzmann, M., and D. Scherf. Über Elektrokardiogramme und verkürzter Vorhof‐Kammerdistanz und positiven P‐Zacken. Z. Klin. Med. 121: 404–410, 1932.
 343. Wolff, L., J. Parkinson, and P. D. White. Bundle branch block with short PR‐interval in healthy young people prone to paroxysmal tachycardia. Am. Heart J. 5: 685–692, 1930.
 344. Janse, M. J. Reentrant arrhythmias. In: The Heart and Cardiovascular System, 2nd Edition, edited by H. A. Fozzard. New York: Raven Press; 2055–2094, 1992.
 345. MacLean, W. A. H., V. J. Plumb, and A. L. Waldo. Transient entrainment and interruption of ventricular tachycardia. PACE 4: 358–365, 1981.
 346. Arenal, A., J. Almendral, D. San Román, J. L. Delcan, and M. E. Josephson. Frequency and implications of resetting and entrainment with right atrial stimulation in atrial flutter. Am. J. Cardiol. 70: 1292–1298, 1992.
 347. Boersma, L., J. Brugada, C. Kirchhof, and M. Allessie. Entrainment of reentrant ventricular tachycardia in anisotropic rings of rabbit myocardium—mechanisms of termination, changes in morphology, and acceleration. Circulation 88: 1852–1865, 1993.
 348. Frazier, D. W. and M. S. Stanton. Resetting and transient entrainment of ventricular tachycardia. Pacing Clin. Electrophysiol. 18: 1919–1946, 1995.
 349. Waldecker, B., J. Coromilas, A. E. Saltman, S. M. Dillon, and A. L. Wit. Overdrive stimulation of functional reentrant circuits causing ventricular tachycardia in the infarcted canine heart—resetting and entrainment. Circulation 87: 1286–1305, 1993.
 350. Allessie, M. A., F. I. M. Bonke, and F. J. C. Schopman. Circus movement in rabbit atrial muscle as a mechanism of tachycardia. III. The “leading circle” concept: a new model of circus movement in cardiac tissue without the involvement of an anatomical obstacle. Circ. Res. 41: 9–18, 1977.
 351. Smeets, J. L. R. M., M. A. Allessie, L. W. J. E. P. F. I. M. Bonke, and J. Hollen. The wavelength of the cardiac impulse and reentrant arrhythmias in isolated rabbit atrium. The role of heart rate, autonomic transmitters, temperature, and potassium. Circ. Res. 58: 96–108, 1986.
 352. Frame, L. H., R. L. Page, and B. F. Hoffman. Atrial reentry around an anatomic barrier with a partially refractory excitable gap. A canine model of atrial flutter. Circ. Res. 58: 495–511, 1986.
 353. Baeriswyl, G., M. Zimmermann, and R. Adamec. Efficacy of rapid atrial pacing for conversion of atrial flutter in medically treated patients. Clin. Cardiol. 17: 246–250, 1994.
 354. Simson, M. B., J. F. Spear, E. N. Moore. Stability of an experimental atrioventricular reentrant tachycardia in dogs. Am. J. Physiol. 240 (Heart Circ. Physiol. 9): H947–H953, 1981.
 355. Vinet, A., and F. A. Roberge. The dynamics of sustained reentry in a ring model of cardiac tissue. Ann. Biomed. Eng. 22: 568–591, 1994.
 356. Ito, H., and L. Glass. Theory of reentrant excitation in a ring of cardiac tissue. Physica D. 56: 84–106, 1992.
 357. Frame, L. H., and M. B. Simson. Oscillations of conduction, action potential duration, and refractoriness. A mechanism for spontaneous termination of reentrant tachycardias. Circulation 78: 1277–1287, 1988.
 358. Garrey, W. E. Auricular fibrillation. Physiol. Rev. 4: 215–250, 1924.
 359. Allessie, M. A., F. I. M. Bonke, and F. J. C. Schopman. Circus movement in rabbit atrial muscle as a mechanism of tachycardia. Circ. Res. 33: 54–62, 1973.
 360. El‐Sherif, N., R. Mehra, W. B. Gough, and R. H. Zeiler. Reentrant ventricular arrhythmias in the late myocardial infarction period. Interruption of reentrant circuits by cryothermal techniques. Circulation 68: 644–656, 1983.
 361. El‐Sherif, N., A. Smith, and K. Evans. Canine ventricular arrhythmias in the late myocardial infarction period: epicardial mapping of reentrant circuits. Circ. Res. 1981: 255–265, 1981.
 362. Gough, W. B., R. Mehra, M. Restivo, R. H. Zeiler, and N. El‐Sherif. Reentrant ventricular arrhythmias in the late myocardial infarction period in the dog. 13. Correlation of activation and refractory maps. Circ. Res. 57: 432–442, 1985.
 363. Hoffman, B. F., and M. R. Rosen. Cellular mechanisms for cardiac arrhythmias. Circ. Res. 49: 1–15, 1981.
 364. Allessie, M. A., W. J. E. P. Lammers, F. I. M. Bonke, and J. Hollen. Experimental evaluation of Moe's multiple wavelet hypotheis of atrial fibrillation. In: Cardiac Arrhythmias, edited by D. P. Zipes and J. Jalife. New York: Grune & Stratton; 265–276, 1985.
 365. Janse, M. J., F. J. L. Van Capelle, H. Morsink, A. G. Kléber, F. J. G. Wilms‐Schopman, R. Cardinal, C. Naumann d'Alnoncourt, and D. Durrer. Flow of “injury” current and patterns of excitation during early ventricular arrhythmias in acute regional myocardial ischemia in isolated porcine and canine hearts. Evidence for 2 different arrhythmogenic mechanisms. Circ. Res. 47: 151–165, 1980.
 366. Müller, S. C., T. Plesser, B. Hess. The structure of the core of the spiral wave in the Belousov‐Zhabotinskii reaction. Science 230: 661–663, 1985.
 367. Winfree, A. T. Spiral waves of chemical activity. Science 175: 634–636, 1972.
 368. Gorelova, N. A. and J. Bures. Spiral waves of spreading depression in the isolated chicken retina. J. Neurobiol. 14: 353–363, 1983.
 369. Shibata, J. and J. Bures. Optimum topographical conditions for reverberating cortical spreading depression in rats. J. Neurobiol. 5: 107–118, 1974.
 370. Lechleiter, J., S. Girard, E. Peralta, and D. Clapham. Spiral calcium wave propagation and annihilation in. Xenopus laevis oocytes. Science 252: 123–126, 1991.
 371. Lipp, P. and E. Niggli. Microscopic spiral waves reveal positive feedback in subcellular calcium signaling. Biophys. J. 65: 2272–2276, 1993.
 372. Tomchik, K. J. and P. N. Devreotes. Adenosine 3′,5′‐monophosphate waves in. Dictyostelium discoideum: a demonstration by isotope dilution‐fluorography. Science 212: 443–446, 1981.
 373. Selfridge, O. Studies of flutter and fibrillation. Arch. Inst. Cardiologia de Mexico. 18: 177–187, 1948.
 374. Balakhovsky, I. S. Several modes of excitation movement in ideal excitable tissue. Biophysics 10: 1175–1179, 1965.
 375. Gul'ko, F. B. and A. A. Petrov. Mechanism of the formation of closed pathways of conduction in excitable media. Biophysics 17: 271–282, 1972.
 376. Davidenko, J. M., A. V. Pertsov, R. Salomonsz, W. Baxter, and J. Jalife. Stationary and drifting spiral waves of excitation in isolated cardiac muscle. Nature 355: 349–351, 1992.
 377. Gray, R., A. Pertsov, and J. Jalife. Spatial and temporal organization during cardiac fibrillation. Nature 392: 75–78, 1998.
 378. Fast, V. G. and A. M. Pertsov. Shift and termination of functional reentry in isolated ventricular preparations with quinidine‐induced inhomogeneity on refractory period. J. Cardiovasc. Electrophysiol. 3: 255–265, 1992.
 379. Frazier, D. W., P. D. Wolf, J. M. Wharton, A. S. L. Tang, W. M. Smith, and R. E. Ideker. Stimulus‐induced critical point. Mechanism for electrical initiation of reentry in normal canine myocardium. J. Clin. Invest. 83: 1039–1052, 1989.
 380. Davidenko, J. M. Spiral wave activity: a possible common mechanism for polymorphic and monomorphic ventricular tachycardias. J. Cardiovasc. Electrophysiol. 4: 730–746, 1993.
 381. Pertsov, A. M., A. V. Panfilov, and F. U. Medvedeva. Instabilities of autowaves in excitable media associated with critical curvature phenomenon. Biofizika 28: 100–102, 1983.
 382. Agladze, K., J. P. Keener, S. C. Muller, and A. Panfilov. Rotating spiral waves created by geometry. Science 264: 1746–1748, 1994.
 383. Fast, V. G., I. R. Efimov, and V. I. Krinsky. Transition from circular to linear rotation of a vortex in an excitable cellular medium. Physics Lett. A 151: 157–161, 1990.
 384. Krinsky, V. I., I. R. Efimov, and J. Jalife. Vortices with linear cores in excitable media. Proc. R. Soc. London Ser. A. 437: 645–655, 1992.
 385. Winfree, A. T. Scroll‐shaped waves of chemical activity in three dimensions. Science 181: 937–939, 1973.
 386. Zykov, V. S. Cycloid circulation of spiral waves in excitable medium. Biophysics 31: 940–944, 1986.
 387. Lugosi, E. Analysis of meandering in Zykov kinetics. Physica D 40: 331–337, 1989.
 388. Gerhardt, M., H. Schuster, and J. J. Tyson. A cellular automaton model of excitable media. II. Curvature, dispersion, rotating waves and meandering waves. Physica D 46: 392–415, 1990.
 389. Efimov, I., V. Krinsky, and J. Jalife. Dynamics of rotating vortices in the Beeler‐Reuter model of cardiac tissue. Chaos, Solitons & Fractals 5: 513–526, 1995.
 390. Holden, A. V. and H. Zhang. Characteristics of atrial re‐entry and meander computed from a model of a rabbit single atrial cell. J. Theor. Biol. 175: 545–551, 1995.
 391. Starmer, C. F., D. N. Romashko, R. S. Reddy, Y. I. Zilberter, J. Starobin, A. O. Grant, and V. I. Krinsky. Proarrhythmic response to potassium channel blockade: numerical studies of polymorphic tachyarrhythmias. Circulation 92: 595–605, 1995.
 392. El‐Sherif, N. Reentrant mechanisms in ventricular arrhythmias. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. P. Zipes and J. Jalife. Philadelphia: W. B. Sounders; 567–582, 1995.
 393. Gray, R. A., J. Jalife, A. Panfilov, W. T. Baxter, C. Cabo, J. M. Davidenko, and A. M. Pertsov. Nonstationary vortexlike reentrant activity as a mechanism of polymorphic ventricular tachycardia in the isolated rabbit heart. Circulation 91: 2454–2469, 1995.
 394. Courtemanche, M. and A. T. Winfree. Two dimensional rotating depolarization waves in a modified Beeler‐Reuter model of cardiac cell activity. In: Science at the John von Neumann National Supercomputer Center, edited by G. Cook. Princeton, NJ: Consortium for Scientific Computing, 79–86, 1990.
 395. Leon, L. J., F. A. Roberge, and A. Vinet. Simulation of two‐dimensional anisotropic cardiac reentry: effects of the wavelength on the reentry characteristics. Ann. Biomed. Eng. 22: 592–609, 1994.
 396. Courtemanche, M. Complex spiral wave dynamics in a spatially distributed ionic model of cardiac activity. Chaos 6: 579–600, 1996.
 397. Qu, Z., J. N. Weiss, A. Garfinkel. Cardiac electrical restitution properties and stability of reentrant spiral waves: a simulation study. Am J Physiol. 276 (Heart Circ. Physiol. 45): H269–283, 1999.
 398. Cao, J. M., Z. Qu, Y. H. Kim, T. J. Wu, A. Garfinkel, J. N. Weiss, H. S. Karagueuzian, and P. S. Chen. Spatiotemporal heterogeneity in the induction of ventricular fibrillation by rapid pacing: importance of cardiac restitution properties. Circ. Res. 84: 1318–1331, 1999.
 399. Ito, H. and L. Glass. Spiral breakup in a new model of discrete excitable media. Phys. Rev. Lett. 66: 671–674, 1991.
 400. Karma, A. Spiral breakup in model equations of action potential propagation in cardiac tissue. Phys. Rev. Lett. 71: 1103–1106, 1993.
 401. Panfilov, A. and P. Hogeweg. Spiral breakup in a modified Fitzhugh‐Nagumo model. Phys. Lett. A. 176: 295–299, 1993.
 402. Karma A. Electrical alternans and spiral wave breakup in cardiac tissue. Chaos 4: 461–472, 1994.
 403. Koller, M. L., M. L. Riccio, and R. F. Gilmour, Jr. Dynamic restitution of action potential duration during electrical alternans and ventricular fibrillation. Am J Physiol. 275 (Heart Circ. Physiol. 44): H1635–1642, 1998.
 404. Riccio, M. L., M. L. Koller, and R. F. Gilmour, Jr. Electrical restitution and spatiotemporal organization during ventricular fibrillation. Circ. Res. 84: 955–963, 1999.
 405. Panfilov, A. V. and B. N. Vasiev. Vortex initiation in a heterogeneous excitable medium. Physica D 49: 107–113, 1991.
 406. Fast, V. G. and I. R. Efimov. Stability of vortex rotation in an excitable cellular medium. Physica D 49: 75–81, 1991.
 407. Fast, V. G. and A. M. Pertsov. Drift of a vortex in the myocardium. Biophysics 35: 489–494, 1990.
 408. 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.
 409. Abildskov, J. A. and R. L. Lux. The mechanism of simulated torsade de pointes in a computer model of propagated excitation. J. Cardiovasc. Electrophysiol. 2: 224–237, 1991.
 410. Jalife, J. and R. Gray. Drifting vortices of electrical waves underlie ventricular fibrillation in the rabbit heart. Acta. Physiol. Scand. 157: 123–131, 1996.
 411. Schalij, M. J., W. E. J. P. Lammers, P. L. Rensma, and M. A. Allessie. Anisotropic conduction and reentry in perfused epicardium of rabbit left ventricle. Am. J. Physiol. 263 (Heart Circ. Physiol. 32): H1466–H1478, 1992.
 412. Wit, A. L. and S. M. Dillon. Anisotropic reentry. In: Cardiac Electrophysiology. From Cell to Bedside, edited by D. P. Zipes and J. Jalife. Philadelphia: W. B. Saunders, 353–364, 1990.
 413. Pertsov, A. M., J. Jalife. Three‐dimensional vortex‐like reentry. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. P. Zipes and J. Jalife. Philadelphia: W. B. Saunders; 403–409, 1995.
 414. Panfilov, A. V. and A. M. Pertsov. Vortex rings in a three‐dimensional medium described by reaction‐diffusion equations. Doklady AN SSSR 274: 58–60, 1984.
 415. Winfree, A. T. Electrical turbulence in three‐dimensional heart muscle. Science 266: 1003–1006, 1994.
 416. 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.
 417. Ursell, P. C., P. I. Gardner, A. Albala, J. J. J. Fenoglio, and A. L. Wit. Structural and electrophysiological changes in the epicardial border zone of canine myocardial infarcts during infarct healing. Circ. Res. 56: 436–451, 1985.
 418. Restivo, M., H. Yin, E. B. Caref, A. I. Patel, G. Ndrepepa, M. J. Avitable, M. A. Assadi, N. Isber and N. Elsherif. Reentrant arrhythmias in the subacute infarction period: the proarrhythmic effect of flecainide acetate on functional reentrant circuits. Circulation 91: 1236–1246, 1995.
 419. El‐Sherif, N., H. Yin, E. B. Caref, and M. Restivo. Electrophysiological mechanisms of spontaneous termination of sustained monomorphic reentrant ventricular tachycardia in the canine postinfarction heart. Circulation 93: 1567–1578, 1996.
 420. Boersma, L., J. Brugada, M. J. Schalij, C. Kirchhof, and M. Allessie. The effects of K+ on anisotropic conduction in sheets of perfused rabbit ventricular epicardium. J. Cardiovasc. Electrophysiol. 2: 492–502, 1991.
 421. Wit, A. L., S. M. Dillon, and J. Coromilas. Anisotropic reentry as a cause of ventricular tachyarrhythmias in myocardial infarction. In: Cardiac Electrophysiology: From Cell to Bedside, edited by D. P. Zipis and J. Jalife. Philadelphia: W. B. Saunders; 511–526, 1995.
 422. Kadish, A., M. Shinnar, E. N. Moore, J. H. Levine, C. W. Balke, and J. F. Spear. Interaction of fiber orientation and direction of impulse propagation with anatomic barriers in anisotropic canine myocardium. Circulation 78: 1478–1494, 1988.
 423. Bonke, F. I. M. Electrotonic spread in the sinoatrial node of the rabbit heart. Pflugers Arch. 339: 17–23, 1973.
 424. Bonke, F. I. M. Passive electrical properties of atrial fibers of the rabbit heart. Pflugers Arch. 339: 1–15, 1973.
 425. Sakamoto, Y., and M. Goto. A study of the membrane constant in the dog myocardium. Jpn. J. Physiol. 20: 30–41, 1970.
 426. Tille, J. Electronic interaction between muscle fibers in the rabbit ventricle. J. Gen. Physiol. 50: 189–202, 1966.
 427. Daut, J. The passive electrical properties of Guinea‐pig ventricular muscle as examined with a voltage‐clamp technique. J. Physiol. (Lond.) 330: 221–242, 1982.
 428. Horiba, M. Stimulus conduction in atria studied by means of intracellular microelectrode. Part I. That in Bachmann's bundle. Jpn. Heart. J. 4: 333–345, 1963.
 429. Wagner, M. L., R. Lazzara, R. M. Weiss, and B. F. Hoffman. Specialized conducting fibers in the interatrial band. Circ. Res. 18: 502–518, 1966.
 430. Spach, M., and P. C. Dolber. The relation between discontinuous propagation in anisotropic cardiac muscle and the vulnerable period of reentry. In: Cardiac Electrophysiology, edited by D. P. Zipes and J. Jalife. New York: Grune and Stratton; 241–252, 1985.
 431. Kirchhof, C., M. Wijffels, J. Brugada, J. Planellas, and M. Allessie. Mode of action of a new class IC drug (ORG 7797) against atrial fibrillation in conscious dogs. J. Cardiovasc. Pharmacol. 17: 116–124, 1991.
 432. Horibe, H. Studies on the spread of the right atrial activation by means of intracellular microelectrode. Jpn. Circ. J. 25: 583–593, 1961.
 433. Hogan, P. M. and L. D. Davis. Electrophysiological characteristics of canine atrial plateau fibers. Circ. Res. 28: 62–73, 1971.
 434. Brendel, W., W. Raule and W. Trautwein. Leitungsgeschwindigkeitund Erregungsausbreitung in den Vörhofendes Hundes. Pflugers Arch. 253: 106–113, 1950.
 435. Frame, L. H., R. L. Page, P. A. Boyden, J. J. Fenoglio, and B. F. Hoffman. Circus movement in the canine atrium around the tricuspid ring during experimental atrial flutter and during reentry in vitro. Circulation 76: 1155–1175, 1987.
 436. Draper, M. H., and S. Weidmann. Cardiac resting and action potentials recorded with an intracellular electrode. J. Physiol. (Lond.) 115: 74–94, 1951.
 437. Trautwein, W., U. Gottstein, and K. Federschmidt. Der Einfluss der Temperatur auf den Actionsstrom des excidierten Purkinje‐Fadens, gemessen mit einer intracellulären Electrode. Pflugers Arch. 258: 243–260, 1953.
 438. Hoffman, B. F., P. F. Cranefield, and J. H. Stuckey, et al. Direct measurement of conduction velocity in. in situ specialized conduction system of mammalian heart. Proc. Soc. Exp. Biol. Med. 102: 55–57, 1959.
 439. Draper, M. H., and M. Mya‐Tu. A comparison of the conduction velocity in cardiac tissues of various mammals. Q. J. Exp. Physiol. 44: 91–109, 1959.
 440. Rosen, M. R., M. J. Legato, and R. M. Weiss. Developmental changes in impulse conduction in the canine heart. Am. J. Physiol. 240 (Heart Circ. Physiol. 9): H546–H554, 1981.
 441. Pressler, M. L., V. Elharrar, and J. C. Bailey. Effects of extracellular calcium ions, verapamil, and lanthanum on active and passive properties of canine cardiac Purkinje fibers. Circ. Res. 51: 637–651, 1982.
 442. Sano, T., N. Takayama, and T. Shimamoto. Directional difference of conduction velocity in the cardiac ventricular syncytium studied by microelectrodes. Circ. Res. VII: 262–267, 1959.
 443. Roberts, D. E., and A. M. Scher. Effect of tissue anisotropy on extracellular potential fields in canine myocardium in situ. Circ. Res. 50: 342–351, 1982.
 444. Spear, J. F., E. L. Michelson, and E. N. Moore. Cellular electrophysiologic characteristics of chronically infarcted myocardium in dogs susceptible to sustained ventricular tachyarrhythmias. J. Am. Coll. Cardiol. 1: 1099–1110, 1983.
 445. Tsuboi, N., I. Kodama, J. Tayama, and K. Yamada. Anisotropic conduction properties of canine ventricular muscles. Jpn. Circ. J. 49: 487–498, 1985.
 446. Balke, C. W., M. D. Lesh, J. F. Spear, A. Kadish, J. H. Levine, and E. N. Moore. Effects of cellular uncoupling on conduction in anisotropic canine ventricular myocardium. Circ. Res. 63: 879–892, 1988.
 447. Quinteiro, R. A., M. O. Biagetti, E. de Forteza. Relationship between Vmax and conduction velocity in uniform anisotropic canine ventricular muscle: differences between the effects of lidocaine and amiodarone. J. Cardiovasc. Pharmacol. 16: 931–939, 1990.
 448. Lewis, T. The Mechanism and Graphic Registration of the Heart Beat. London: Shaws & Sons, 1920.
 449. Tranum‐Jensen, J. The fine structure of the atrial and atrioventricular (AV) junctional specialized tissues of the rabbit heart. In: The Conduction System of the Heart, edited H. J. J. Wellens, K. I. Lie, and M. J. Janse. Leiden: Stenfert Kroese, 1976.
 450. Krinsky, V. I., I. R. Efimov. Vortices with linear cores in mathematical models of excitable media. Physica A 188: 55–60, 1992.

Contact Editor

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

* Required Field

How to Cite

André G. Kléber, Michiel J. Janse, Vladimir G. Fast. Normal and Abnormal Conduction in the Heart. Compr Physiol 2011, Supplement 6: Handbook of Physiology, The Cardiovascular System, The Heart: 455-530. First published in print 2002. doi: 10.1002/cphy.cp020112