What provides a graphic representation of all the action potentials occurring in the heart?

Recommended textbook solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Clinical Reasoning Cases in Nursing

7th EditionJulie S Snyder, Mariann M Harding

2,512 solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Clinical Reasoning Cases in Nursing

7th EditionJulie S Snyder, Mariann M Harding

2,512 solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Pathophysiology for the Health Professions

6th EditionKarin VanMeter, Robert Hubert

1,226 solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Pharmacology: An Introduction

8th EditionBarbara T Nagle, Hannah Ariel, Henry Hitner, Michele B. Kaufman, Yael Peimani-Lalehzarzadeh

1,355 solutions

Recommended textbook solutions

What provides a graphic representation of all the action potentials occurring in the heart?

The Human Body in Health and Disease

7th EditionGary A. Thibodeau, Kevin T. Patton

1,505 solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Clinical Reasoning Cases in Nursing

7th EditionJulie S Snyder, Mariann M Harding

2,512 solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Exercise Physiology: Theory and Application to Fitness and Performance

11th EditionEdward Howley, John Quindry, Scott Powers

593 solutions

What provides a graphic representation of all the action potentials occurring in the heart?

Medical Assisting Review: Passing the CMA, RMA, and CCMA Exams

7th EditionJahangir Moini

1,975 solutions

Principles of Electrophysiology

Lee Goldman MD, in Goldman-Cecil Medicine, 2020

The Cardiac Action Potential

The cardiac action potential (Fig. 55-1) is a recording of a cell’s membrane potential,Vm, versus time. During each cardiac cycle, ions move back and forth across the cardiomyocyte cell membrane, thereby changingVm. The cardiac action potential, which reflects the integrated behavior of numerous individual ionic currents, is largely dominated by the movement of Na+, Ca2+, and K+ ions. These ions traverse the cell membrane through ion-selective pores formed by assemblies of integral membrane-spanning proteins and accessory proteins. The behavior of these ionic pathways is highly regulated, and permeation of specific ions is influenced by multiple factors, the most prominent of which are changes in membrane potential (i.e., voltage gating), ligand binding, second messengers such as cyclic adenosine monophosphate, and post-translational modification. Channel function and, by extension, action potential behavior are dynamically tuned in response to normal physiologic factors, especially heart rate. However, a number of pathologic stressors influence channel activity, including acquired syndromes that are associated with cardiac hypertrophy and failure, as well as an ever-growing number of congenital diseases. Regardless of the underlying pathology, the effects on action potential behavior may trigger arrhythmic activity.

The cardiac action potential is divided into phases, each reflecting the major ionic movements that take place. In working cardiomyocytes, such as ventricular or atrial myocytes, theresting membrane potential during diastole, or phase 4 of the cardiac action potential, is determined by the baseline ionic and charge gradients that exist across the sarcolemmal membrane. These gradients are generated by pumps and transporters, the most important of which is the Na+, K+-ATPase. This energy-requiring electrogenic pump, which is the major target of ouabain-like compounds such as digoxin, extrudes three Na+ ions from the intracellular compartment in exchange for two K+ ions, thereby resulting in directionally opposite gradients of Na+ ions (outside > inside) and K+ ions (inside > outside). Under resting conditions, a subset of membrane channels highly permeable to K+ is open, but those that allow for the passage of other ions such as Na+ or Ca2+ are only minimally permeable. As a consequence, the concentration gradient promotes the movement of potassium ions from inside to outside of the cell, until the resulting excess of negative charge within the cell balances the diffusional forces and an electrochemical equilibrium is established. The equilibrium potential for a given ion is calculated by theNernst equation, where Eeq is the equilibrium potential, R is the universal gas constant, T is the absolute temperature, z is the valence of the ionic species, and F is Faraday constant:

Eeq=RTzFln([X] out[X]in)

If the cell membrane wereonly permeable to K+ ions, at the measured concentrations of intracellular and extracellular K+, the resting membrane potential would be approximately −100 mV. However, because of the slight but measurable permeability to other ionic species, which have Nernst potentials that are less negative than that for K+, the actual resting membrane potential in a typical ventricular cardiac myocyte is closer to −85 mV.

Cardiac Excitability and Heritable Arrhythmias

DAVID E. CLAPHAM, MARK T. KEATING, in Nadas' Pediatric Cardiology (Second Edition), 2006

Phase 2—The Action Potential Plateau

Cardiac action potentials have uniquely long periods of time during which the potential remains depolarized near 0 mV (plateau). Relatively few channels are open during the plateau, and thus the total membrane conductance is low. The high resistance of the membrane during the plateau acts like thick electrical insulation around a wire; it allows rapid propagation of the action potential with little dissipation. The plateau phase is maintained by a finely tuned balance between two types of inward Ca2+ currents and at least four types of outward K+ currents. Ultimately, K+ currents dominate, and the membrane potential is driven back toward EK.

Two Ca2+ currents, the low voltage-activated, transient Ca2+ current (ICa.T, T-type, ICaV3.2) and the high voltage-activated, long-lasting Ca2+ current (ICa.L, L-type, ICaV1.2) admit the Ca2+ needed to initiate contraction. In the normal heart, ICaV3.2 is found predominantly in atrial pacemaker cells, Purkinje fibers, and coronary artery smooth muscle. These Ca2+ channels rapidly activate at about –50 mV, peak at –20 mV, inactivate with time, are blocked by Ni2+ and a novel Ca2+ channel blocker, mibefradil, but are insensitive to dihydropyridines. ICaV3.2 is only one fifth the size of ICaV1.2 and thus contributes relatively little to the Ca2+ influx of excitation–contraction coupling.11

In contrast to ICaV3.2, ICaV1.2 is the dominant Ca2+ current found in virtually all cardiac cells in all species.12 ICa.L is activated at –30 mV, reaches its peak conductance at +10 mV within 3 to 5 msec, inactivates over hundreds of milliseconds, and is sensitive to block by dihydropyridines (nifedipine), benzothiazepines (diltiazem), and phenylalkylamines (verapamil).6 These channels carry inward current throughout the plateau phase and are required for coupling membrane excitability to myocardial contraction.

The terminal portion of the plateau is sustained by inward current through the electrogenic Na+/Ca2+ exchanger as Ca2+ is transported out of the cell at a ratio of 1 Ca2+ ion per 3 Na+ ions.13 This electrogenic current is capable of providing inward current up to 50% of the size of ICaV1.2. During the action potential upstroke, the exchanger brings in Ca2+ ions, increasing contractility. As the chief means of extruding Ca2+ ions, the exchanger also modulates the Ca2+ content of the sarcoplasmic reticulum.14

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9781416023906500660

Mechanisms of Cardiac Arrhythmias

Douglas P. Zipes MD, in Braunwald's Heart Disease: A Textbook of Cardiovascular Medicine, 2019

Phases of the Cardiac Action Potential

The cardiac transmembrane action potential consists of five phases:phase 0, upstroke or rapid depolarization;phase 1, early rapid repolarization;phase 2, plateau;phase 3, final rapid repolarization; andphase 4, resting membrane potential and diastolic depolarization (Fig. 34.2andeFig. 34.1). These phases are the result of passive ion fluxes moving down their electrochemical gradients established by active ion pumps and exchange mechanisms. Each ion moves primarily through its own ion-specific channel. The following discussion explains the electrogenesis of each of these phases.

EFIGURE 34.1. Demonstration of action potentials recorded during impalement of a cardiac cell.Upper row, Shown are a cell (circle), two microelectrodes, and stages during impalement of the cell and its activation and recovery. Both microelectrodes are extracellular (A), and no difference in potential exists between them (0 potential). The environment inside the cell is negative, and the outside is positive, because the cell is polarized. One microelectrode has pierced the cell membrane (B) to record the intracellular resting membrane potential, which is −90 mV with respect to the outside of the cell. The cell has depolarized (C), and the upstroke of the action potential is recorded. At its peak voltage, the inside of the cell is approximately +30 mV with respect to the outside of the cell. The repolarization phase (D) is shown, with the membrane returning to its former resting potential (E).

General Considerations.

Ionic fluxes regulate membrane potential in cardiac myocytes in the following fashion. When only one type of ion channel opens, assuming that this channel is perfectly selective for that ion, the membrane potential of the entire cell would equal the Nernst potential of the permeant ion. By solving the Nernst equation for the four major ions across the plasma membrane, the following equilibrium potentials are obtained: sodium, +60 mV; potassium, −94 mV; calcium, +129 mV; and chloride, −83 to −36 mV (Table 34.2). Therefore, if K+-selective channels open, such as the inwardly rectifying K+ (Kir) channel (see later), the membrane potential approaches EK (−94 mV). If Na+-selective channels open, the transmembrane potential becomes ENa (+60 mV). A quiescent cardiac myocyte (phase 4) has many more open potassium than sodium channels, and the cell's transmembrane potential is close to EK. When two or more types of ion channels open simultaneously, each channel moves the membrane potential to the equilibrium potential of their respective permeant ions. The contribution of each ion type to the overall membrane potential at any given moment is determined by the instantaneous permeability of the plasma membrane to that ion. For example, deviation of the measured resting membrane potential from EK (seeTable 34.2) would predict that other ion types with equilibrium potentials positive to EK are contributing to the resting membrane potential in cardiac myocytes. If it is assumed that Na+, K+, and Cl− are the permeant ions at resting potential, their individual contributions to the resting membrane potential (V) can be quantified by the Goldman-Hodgkin-Katz (GHK) voltage equation:

V=(RT/F)ln[( PK[Na] o+PCl[Cl]i)/ PK[K]i+PNa[Na ]PCl[Cl]o

where the symbols have the meanings outlined previously. With only one permeant ion, V approximates the Nernst potential for that ion. With several permeant ion types, V is a weighted mean of all the Nernst potentials.

Resting Membrane Potential.

The intracellular potential during electrical quiescence in diastole is −50 to −95 mV, depending on the type of cell (Table 34.2). Therefore the inside of the cell is 50 to 95 mV negative relative to the outside of the cell because of the transmembrane gradients of ions such as K+, Na+, and Cl−.

Because cardiac myocytes have an abundance of open K+ channels at rest, the cardiac transmembrane potential (in phase 4) is close to EK. Outward potassium current through open, inwardly rectifying K+ channels (IK1) under normal conditions contributes to the resting membrane potential mainly in atrial and ventricular myocytes, as well as in Purkinje cells. Deviation of the resting membrane potential from EK is the result of movement of ions with an equilibrium potential greater than the EK, for example, Cl− efflux through activated chloride channels, such as ICl.cAMP, ICl.Ca, and ICl.swell. Calcium does not contribute directly to the resting membrane potential, but changes in intracellular free calcium concentration [Ca2+]i can affect other membrane conductance values. For example, an increase in sarcoplasmic reticulum (SR) Ca2+ load can cause spontaneous intracellular Ca2+ waves, which in turn activate the Ca2+ -dependent chloride conductance ICl.Ca and thereby lead to spontaneous transient inward currents and concomitant membrane depolarization. Increases in [Ca2+]i can also stimulate the Na+/Ca2+ exchanger INa/Ca. This protein exchanges three Na+ ions for one Ca2+ ion and thus generates a current; the direction depends on the [Na+] and [Ca2+] on the two sides of the membrane and the transmembrane potential difference (seeElectrogenic Transporters). At the resting membrane potential and during a spontaneous SR Ca2+-release event, this exchanger would generate a net Na+ influx, possibly causing transient membrane depolarization.2 Another transporter, the Na-K pump, electrogenically pumps Na+ out of the cell and simultaneously pumps K+ into the cell (three Na+ outward and two K+ inward) against their respective chemical gradients, keeping the intracellular K+ concentration high and the intracellular Na+ concentration low. The rate of Na+-K+ pumping to maintain the same ionic gradients must increase as the heart rate increases because the cell gains a small amount of Na+ and loses a small amount of K+ with each depolarization. Cardiac glycoside block of Na+,K+-ATPase increases contractility through an increase in intracellular Na+ concentration [Na+]i, which in turn reduces Ca2+ extrusion through the Na+/Ca2+ exchanger and thereby increases myocyte contractility.3

Phase 0: Upstroke or Rapid Depolarization.

A stimulus delivered to excitable tissues can evoke an action potential characterized by a sudden change in voltage caused by transient depolarization followed by repolarization. The action potential is conducted throughout the heart and is responsible for initiating each heartbeat. Electrical changes in the action potential follow a relatively fixed time and voltage relationship that differs according to specific cell types (Fig. 34.3). In neurons, the entire process takes several milliseconds, whereas action potentials in human cardiac fibers last several hundred milliseconds. Normally, the action potential is independent of the size of the depolarizing stimulus if the latter exceeds a certain threshold potential. Small, subthreshold depolarizing stimuli depolarize the membrane in proportion to the strength of the stimulus. However, when the stimulus is sufficiently intense to reduce membrane potential to a threshold value in the range of −70 to −65 mV for normal Purkinje fibers, an “all-or-none” response results. More intense depolarizing stimuli do not produce larger action potential responses; in contrast, hyperpolarizing pulses, or stimuli that render the membrane potential more negative, elicit a response proportional to the strength of the stimulus.

Mechanism of Phase 0.

The upstroke of the cardiac action potential in atrial and ventricular muscle and His-Purkinje fibers is the result of a sudden increase in membrane conductance of Na+. An externally applied stimulus or a spontaneously generated local membrane circuit current in advance of a propagating action potential depolarizes a sufficiently large area of membrane at an adequately rapid rate to open the Na+ channels and depolarize the membrane further. When the stimulus activates enough Na+ channels, Na+ ions enter the cell down their electrochemical gradient. The excited membrane no longer behaves like a K+ electrode, that is, exclusively permeable to K+, but more closely approximates an Na+ electrode, and the membrane voltage moves toward the Na+ equilibrium potential (+60 mV).

The rate at which depolarization occurs during phase 0, that is, the maximum rate of change in voltage over time, is indicated by the expression dV/dtmax or V̇max (seeTable 34.2), which is an approximation of the rate and magnitude of Na+ entry into the cell and a determinant of conduction velocity for the propagated action potential. The transient increase in sodium conductance lasts 1 to 2 milliseconds. The action potential, or more properly the Na+ current (INa), is said to be regenerative; that is, intracellular movement of a little Na+ depolarizes the membrane more, which increases conductance of Na+ more and allows more Na+ to enter, and so on. As this process is occurring, however, [Na+]i and positive intracellular charges increase and reduce the driving force for Na+ flux into the cell. When the equilibrium potential for Na+ (ENa) is reached, the driving force acting on the ion to enter the cell balances the driving force acting on the ion to exit the cell, and no current flows. Importantly, Na+ conductance is time dependent, so when the membrane spends some time at voltages less negative than the resting potential, Na+ conductance decreases (inactivation). Therefore an intervention that reduces membrane potential for a time (acute myocardial ischemia), but not to threshold, partially inactivates Na+ channels, and if the threshold is now achieved, the magnitude and rate of Na+ influx are reduced, which causes conduction velocity to slow.

In cardiac Purkinje fibers, sinoatrial cells, and to a lesser extent, ventricular muscle, different populations of Na+ channels exist: the tetrodotoxin (TTX)-sensitive, neuronal Na+ channel isoforms and the TTX-resistant Nav1.5 isoform, the latter being the predominant isoform in cardiac muscle.4 Although the precise roles of TTX-sensitive Na+ channels in ventricular or atrial cardiomyocytes have not been defined, these channels may be important modulators of sinoatrial node pacemaking, Purkinje myocyte action potential duration, and in arrhythmia production in some situations.5 Neuronal Nav channels in the heart have been identified as regulators of contractility.6

Normal atrial and ventricular muscle cells and fibers in the His-Purkinje system exhibit action potentials with very rapid, large-amplitude upstrokes calledfast responses. Action potentials in the normal sinoatrial (SA) and atrioventricular (AV) nodes and many types of diseased tissue have very slow, reduced-amplitude upstrokes and are calledslow responses (seeTable 34.1andFigs. 34.2and34.3). Upstrokes of slow responses are mediated by a slow inward, predominantly L-type voltage-gated (Cav) Ca2+ current (ICa.L) rather than by the fast inward INa and are referred to asslow response potentials because the time required for activation and inactivation of ICa.L is approximately an order of magnitude slower than that for the fast INa. The recovery of slow responses is delayed because of slow recovery of ICa,L from inactivation. Recovery of ICa,L slow-response channel requires establishment of the maximal diastolic potential (i.e., is voltage dependent) and more time before the channel can be activated again (i.e., time dependent), a phenomenon termedpostrepolarization refractoriness. Moreover, calcium entry and [Ca2+]i promote inactivation and delay recovery of slow-response channels.

The prolonged time for reactivation of ICa.L probably accounts for the fact that SA and AV nodal cells remain refractory longer than the time that it takes for full voltage repolarization to occur. Thus, premature stimulation immediately after the membrane potential reaches full repolarization leads to action potentials with reduced amplitudes and upstroke velocities. Therefore, slow conduction and prolonged refractoriness are characteristic features of nodal cells. These cells also have a reduced “safety factor for conduction,” which means that the stimulating efficacy of the propagating impulse is low, and conduction block occurs easily. The electrophysiologic changes accompanying acute myocardial ischemia may represent a depressed form of a fast response in the center of the ischemic zone and a slow response in the border area.

The threshold for activation of ICa.L is about −30 to −40 mV. In fibers of the fast-response type, ICa.L is normally activated during phase 0 by the regenerative depolarization caused by the fast INa. Current flows through both fast and slow channels during the latter part of the action potential upstroke. However, ICa.L is much smaller than the peak INa and therefore contributes little to the action potential until the fast INa is inactivated after completion of phase 0. Thus, ICa.L affects mainly the plateau of action potentials recorded in atrial and ventricular muscle and His-Purkinje fibers. In addition, ICa.L may play a prominent role in partially depolarized cells in which fast INa has been inactivated, if conditions are appropriate for slow-channel activation.

Ca2+ entry through activated L-type Cav channels triggers release of Ca2+ from SR stores and is an essential component of cardiac excitation-contraction coupling in atrial and ventricular myocardium (seeChapter 22). L-type Cav channels are expressed in SA and AV nodal cells, where they play a role in controlling automaticity and action potential propagation, respectively. Although T-type Cav channels have not been detected in human myocardium, experimental evidence in animals has suggested that these channels play an important role in determining SA node automaticity and AV nodal conduction.7

Other significant differences exist between the fast and slow channels. Drugs that elevate cyclic adenosine monophosphate (cAMP) levels, such as beta-adrenoceptor agonists, phosphodiesterase inhibitors such as theophylline, and the lipid-soluble derivative of cAMP, dibutyryl cAMP, increase ICa.L. Although Nav channels are sensitive to increases in cAMP, the net effect (decrease versus increase) appears to be species and condition dependent. Acetylcholine reduces ICa.L by decreasing adenylate cyclase activity. However, acetylcholine stimulates the accumulation of cyclic guanosine monophosphate (cGMP). cGMP has negligible effects on basal ICa.L but decreases the ICa.L levels that have been elevated by beta-adrenoceptor agonists. This effect is mediated by cAMP hydrolysis through a cGMP-stimulated cyclic nucleotide phosphodiesterase.

Fast and slow channels can be differentiated on the basis of their pharmacologic sensitivity. Calcium channel antagonists that block the slow channel with a fair degree of specificity include verapamil, nifedipine, diltiazem, and D-600 (a methoxy derivative of verapamil). Antiarrhythmic agents such as lidocaine, quinidine, procainamide, and disopyramide affect the fast channel and not the slow channel (seeChapter 36).

Phase 1: Early Rapid Repolarization.

Following phase 0, the membrane repolarizes rapidly and transiently to almost 0 mV (early notch), partly because of inactivation of INa and concomitant activation of several outward currents.

The 4-aminopyridine–sensitive transient outward K+ current, commonly termed Ito (or Ito1), is turned on rapidly by depolarization and then rapidly inactivates. Both the density and the recovery of Ito from inactivation exhibit transmural gradients in the left and right ventricular free wall, with the density decreasing and reactivation becoming progressively prolonged from epicardium to endocardium. Transmural differences in the expression of KChIP2, the auxiliary subunit to Kv4.3 pore-forming alpha subunits, may also contribute to the transmural gradient in Ito properties and densities in the human heart.8 This gradient gives rise to regional differences in action potential shape, with increasingly slower phase 1 restitution kinetics and diminution of the notch along the transmural axis (eFig. 34.2).

EFIGURE 34.2. Action potential recordings demonstrating differences in the action potential shape of human ventricular myocytes of subepicardial(A) and subendocardial(B) origin. Subepicardial myocytes present a prominent notch during phase 1 repolarization of the action potential, most likely caused by a larger Ito in these cells. The notch is absent in subendocardial cells. The peak plateau potential is higher in subendocardial than in subepicardial myocytes, and the action potential duration tends to be shorter in subepicardial cells.C, Transmembrane action potential in a human ventricular cardiomyocyte from a failing heart. Note loss of the prominent phase 1 notch and delayed repolarization. Recording temperature = 35°C;Vm, membrane potential.

(A, B, From Näbauer M et al. Regional differences in current density and rate-dependent properties of the transient outward current in subepicardial and subendocardial myocytes of human left ventricle. Circulation 1996;93:168;C, from Priebe L, Beuckelmann DJ. Simulation studies of cellular electrical properties in heart failure. Circ Res 1998;82:1206.)

These regional differences might create transmural voltage gradients, thereby increasing dispersion of repolarization, a putative arrhythmogenic factor (Brugada syndrome;seeChapters 33and39). However, elimination of the physiologic repolarization gradient appears to be similarly arrhythmogenic. Downregulation of Ito is at least partially responsible for slowing of phase 1 repolarization in failing human myocytes. Studies have demonstrated that these changes in the phase 1 notch of the cardiac action potential cause a reduction in the kinetics and peak amplitude of the action potential–evoked intracellular Ca2+ transient because of failed recruitment and synchronization of SR Ca2+ release through ICa.L (eFig. 34.3). Thus, modulation of Ito appears to play a significant physiologic role in controlling cardiac excitation-contraction coupling, and it remains to be determined whether transmural differences in phase 1 repolarization translate into similar differences in regional contractility.

EFIGURE 34.3. Diminution of phase 1 amplitude (“notch”) causes asynchronous sarcoplasmic reticulum (SR) Ca2+ release. Normal cardiomyocytes were voltage-clamped, with action potential profiles having a normal or heart failure wave shape(top), and local changes in intracellular calcium were recorded simultaneously. When the myocyte was clamped with a normal action potential profile having the early phase 1 repolarization notch(left), there was uniform Ca2+ release, reflected in the rapid and synchronous increase in fluorescence. However, when a congestive heart failure action potential profile without early rapid phase 1 repolarization was used(right), Ca2+ release was dyssynchronous. This dyssynchrony causes slowing in the rate of rise of the Ca2+ transient and loss of spatial and temporal release uniformity. F/F0, Fluorescence of the Ca2+ indicator normalized to its baseline fluorescence.

(From Harris DM et al. Alterations in early action potential repolarization causes localized failure of sarcoplasmic reticulum Ca2+ release. Circ Res 2005;96:543. By permission of the American Heart Association.)

The 4-aminopyridine–resistant, Ca2+ -activated chloride current ICl.Ca (or Ito2) also contributes a significant outward current during phase 1 repolarization.1 This current is activated by the action potential–evoked intracellular Ca2+ transient. Therefore, interventions that augment the amplitude of the Ca2+ transient associated with the twitch (e.g., beta-adrenergic receptor stimulation) also enhance outward ICl.Ca. It is not currently known whether human cardiac myocytes express Ca2+-activated chloride channels. Other, time-independent chloride currents may also play a role in determining the time course of early repolarization, such as the cAMP- or swelling-activated chloride conductances ICl.cAMP and ICl.swell.

A third current contributing to early repolarization is outward Na+ movement through the Na+/Ca2+ exchanger operating in reverse mode. Sometimes, a transient depolarization follows phase 1 repolarization altering the initial voltage of the plateau (seeeFig. 34.2).

Phase 2: Plateau.

During the plateau phase, which may last several hundred milliseconds, membrane conductance of all ions falls to rather low values; this is a time of high membrane resistance. Less change in current is required near plateau voltages than near resting potential levels to produce the same changes in transmembrane potential. The plateau is maintained by competition between the outward current carried by K+ and Cl− ions and the inward current carried by Ca2+ moving through ICa,L and Na+ being exchanged for internal Ca2+ by the Na+/Ca2+ exchanger operating in forward mode. After depolarization, IK1 conductance falls to plateau levels as a result of inward rectification, despite the large electrochemical driving force on K+ ions.

Several potassium currents are activated during the plateau phase, including the rapid (IKr) and slow (IKs) delayed rectifier currents (seeVoltage-Gated K+ Channels). The mechanism underlying rectification of the rapid component of the delayed rectifier K+ current (IKr) in cardiac cells is rapid inactivation that occurs during depolarizing pulses. More IKr channels enter the inactivated state with stronger depolarizations, thereby causing inward rectification. This fast inactivation mechanism is sensitive to changes in extracellular K+ in the physiologic range, with inactivation being more accentuated at low extracellular K+ concentrations. Thus, hypokalemia would decrease outward IKr, thereby prolonging the action potential duration (APD).

Outward K+ movement carried by IKs also contributes to plateau duration. Mutations in theKvLQT1 subunit, which in combination with the IKs ancillary subunit (KCNE1 encoding minK) reconstitutes the cardiac IKs current, are associated with abnormally prolonged ventricular repolarization (LQTS type 1;seeChapters 33and39). Although IKs activates slowly compared to the APD, it is only slowly inactivated. Therefore, increases in heart rate can cause this activation to accumulate during successive depolarizations, increasing K+ currents that are active during the plateau of the action potential and thus shortening the APD appropriately at higher heart rates.

In conditions of reduced intracellular adenosine triphosphate (ATP) concentration (e.g., hypoxia, ischemia), K+ efflux through activated KATP channels is enhanced, thereby shortening the plateau phase of the action potential. Other ionic mechanisms that control plateau potential and duration include the kinetics of inactivation of the L-type Ca2+ current. Reduced efficiency of intracellular free Ca2+ in inducing Ca2+-dependent inactivation, such as in myocytes from hypertrophic hearts, can result in delayed repolarization. Steady-state components of both INa and ICa.L (window currents) also shape the plateau phase. Na+,K+ -ATPase generates a net outward current by electrogenic ion exchange. Noninactivating chloride currents, such as ICl.swell and ICl.cAMP, may produce significant outward currents during the plateau phase under certain conditions, thereby significantly shortening the APD. A nonselective, swelling-induced cation current has been shown to cause prolongation of action potentials in myocytes from failing ventricles.1

Phase 3: Final Rapid Repolarization.

Repolarization of the terminal portion of the action potential proceeds rapidly in part because of two currents: time-dependent inactivation of ICaL, with a decrease in the intracellular movement of positive charges, and activation of repolarizing K+ currents, including IKs and IKr and the inwardly rectifying K+ currents IK1 and IKACh, which all cause an increase in the movement of positive charges out of the cell. The net membrane current becomes more outward, and the membrane potential moves to the resting potential. A small-conductance Ca2+ -activated K+ current, IKCa, expressed in human atrial myocytes, controls the time course of phase 3 repolarization.9

Loss-of-function mutations in the human ether-a-go-go–related or hERG gene (KCNH2), which encodes the pore-forming subunit of IKr, prolong phase 3 repolarization, thereby predisposing to the development of torsades de pointes. Macrolide antibiotics such as erythromycin, antihistamines such as terfenadine, several neurologically active agents, and antifungal drugs such as ketoconazole inhibit IKr and have been implicated in acquired forms of LQTS (seeChapters 33and39). Similarly, mutations inKVLQT1, which encodes the pore-forming subunit of IKs, will prolong repolarization and predispose to lethal ventricular arrhythmias. A decrease in IK1 activity, as is the case in left ventricular myocytes from failing hearts, causes prolongation of the action potential by slowing of phase 3 repolarization and resting membrane depolarization. A reduction in the outward potassium current through open inwardly rectifying K+ channels renders the failing cardiomyocyte more susceptible to the induction of delayed afterdepolarizations triggered by spontaneous intracellular Ca2+-release events and therefore plays a major role in arrhythmogenesis in the failing heart.1

Phase 4: Diastolic Depolarization.

Under normal conditions, the membrane potential of atrial and ventricular muscle cells remains steady throughout diastole. IK1 is the current responsible for maintaining the resting potential near the K+ equilibrium potential and shuts off during depolarization in atrial, His-Purkinje, and ventricular cells. In other fibers found in certain parts of the atria, in the muscle of the mitral and tricuspid valves, in His-Purkinje fibers, and in the SA node and portions of the AV nodal tract, the resting membrane potential does not remain constant in diastole but gradually depolarizes (seeFigs. 34.2and34.3). The property possessed by spontaneously discharging cells is called phase 4 diastolic depolarization, which leads to initiation of action potentials resulting in automaticity. The discharge rate of the SA node normally exceeds the discharge rate of other potentially automatic pacemaker sites and thus maintains dominance of the cardiac rhythm. The discharge rate of the SA node is usually more sensitive than the discharge rate of other pacemaker sites to the effects of norepinephrine and acetylcholine. Normal or abnormal automaticity at other sites can cause discharge at rates faster than the SA nodal discharge rate and can thus usurp control of the cardiac rhythm for one cycle or many (seeChapter 35).

Cardiac rhythm diseases

Anika Niambi Al-Shura BSc, MSOM, PhD, in Medical Empathy, Pharmacological Systems, and Treatment Strategies in Integrative Cardiovascular Chinese Medicine, 2020

Pharmaceutical drugs


Cardiac action potentials
Phase 0 Influx of sodium ions
Phase 1 Inactivation of the sodium-channel
Phase 2 Voltage plateau causes the opening of calcium channels
Phase 3 Repolarizing of potassium ions

Antidysrhythmic drugs are often categorized using the Vaughan–Williams classification system, which is based on their mechanism of activity. The Vaughn–Williams classes are as follows:

Class I: Sodium channel blockers

Class II: Beta-adrenergic blockers

Class III: Potassium channel blockers

Class IV: Calcium channel blockers

Class V: Other or unknown mechanism of action: These include the actions of magnesium, digoxin, and adenosine.

Class I: Sodium channel blockers

Reduce depolarization rate, which slows and reduces phase 0

Provide local anesthesia by inhibiting neuronal cells

Inhibit depolarization in atrial, ventricular, and Purkinje myocytes

Decrease conduction velocity and automaticity

Are categorized into A, B, or C subclasses according to the degree and effects of blocking of sodium channels and repolarization

Class IA agents

Prolong both the QRS and QTc intervals

Prolong the repolarization and action potential phases through blocking potassium channels

Slow cardiac conduction

Class IB agents

Do not prolong the QRS interval; shorten the duration of the action potential phases

Depress cardiac conduction in ischemic cells

Bind to the sodium channel in its inactivated state

Class IC agents

Prolong the QRS interval; decreases the rise rate of phase 0 of the action potential

Have little effect on action potential duration

Depress cardiac conduction

Bind to sodium channels in the active state

Are slow to release from sodium channels

Class II: Beta-adrenergic blockers

Block the opening of calcium channels

Block the proarrhythmic effects of catecholamines

Class III: Potassium channel blockers

Lead to prolonged QTc intervals

Delay repolarization by blocking potassium channels

Have very little effect on sodium channels

Class IV: Calcium channel blockers

Slow sinoatrial node pacemaker cell

Slow atrioventricular conduction by blocking L-type voltage-gated calcium channels

Class IA antidysrhythmics

Disopyramide

Sodium and potassium channel blocker and muscarinic antagonist

Indications: Atrial and ventricular dysrhythmias

Metabolism: Metabolized by the liver (CYP3A4) with 40%–60% excreted by the kidneys; caution in renal failure

Procainamide

Sodium and potassium channel blocker and prolong the action potential duration of ventricular myocytes and Purkinje fibers

Indications: Supraventricular or ventricular dysrhythmias

Metabolism: Metabolized in the liver by acetylation into a metabolite that prolongs the action potential with drug and metabolite excreted by the kidneys

Quinidine

Sodium, potassium channel, alpha adrenergic, and muscarininc receptor blocker

Indications: Atrial and ventricular dysrhythmias and Brugada syndrome

Metabolism: Hepatic elimination is 60%–80% and renal elimination is 20%–40%; caution in hepatic or renal diseases

Class IB antidysrhythmics

Lidocaine

Sodium channel blocker and leads to phase 0 depolarization reduction rate and shortens action duration of Purkinje fibers

Indications: For ventricular dysrhythmias and as a local anesthetic; previously prevented dysrhythmias after heart attack (amiodarone is now for this purpose and should not be taken with lidocaine).

Metabolism: Hepatic metabolism by CYP3A4 into an active metabolite

Mexiletine

Sodium channel blocker; leads to phase 0 depolarization reduction rate and shortens action duration of Purkinje fibers and prevents delayed ventricular repolarization and torsades de pointes in LQTS.

Indications: Ventricular dysrhythmias and pain in peripheral neuropathy

Metabolism: After absorption in the small intestine, it is metabolized by liver CYP2D6 at 90%, eliminated unchanged by the kidneys at 10%; caution in hepatic and renal disease

Class IC antidysrhythmics

Flecainide

Sodium channel blocker has negative inotropic effects and slows conduction in all cardiac fibers

Indications: Paroxysmal atrial fibrillation and ventricular dysrhythmias, used to help diagnose and treat congenital Brugada and LQT3 syndromes

Metabolism: Hepatic metabolism CYP2D6 at 75% and renal metabolism at 25%; caution in hepatic and renal diseases

Propafenone

Sodium and calcium channel blocker

Indications: Atrial fibrillation and life-threatening ventricular dysrhythmias

Metabolism: Hepatic metabolism by CYP2D6, CYP3A4, and CYP1A2

Class III antidysrhythmics

Amiodarone

Sodium, L-type calcium potassium channel or beta receptor delayer or blocker; prolongs refractory periods of cardiac tissue

Indications: Supraventricular and life-threatening ventricular dysrhythmias

Metabolism: Hepatic metabolism by CYP3A4 to an active metabolite and mostly excreted in bile

Dronedarone

Sodium, L-type calcium potassium channel and beta receptor delayer or blocker; prolongs refractory periods of cardiac tissue and inhibits alpha1 receptors

Indications: Atrial and ventricular dysrhythmias and sinus rhythm for atrial flutter or fibrillation

Metabolism: Hepatic metabolism by CYP3A4 to active and inactive metabolites

Sotalol

Nonselective beta-adrenergic antagonist, potassium channel blocker; prolongs the action potential and effective refractory period

Indications: Ventricular dysrhythmias, atrial fibrillation, AV tachycardia

Metabolism: 90%–100% bioavailability and absorption rate, no metabolism, caution in renal disease so creatinine clearance is necessary and excreted unchanged by kidneys

Class V antidysrhythmics

Adenosine

Induces a short-duration heart block and atrial action potential, polarizes myocyte membrane potential and slows AV node conduction, and increases potassium conduction

Indications: Given for supraventricular tachycardias after failure of vagal maneuver

Metabolism: Intracellular metabolism

Alpha/beta-adrenergic agonists

Norepinephrine

Increases cardiac output, blood pressure, and heart rate

Strong beta 1 and alpha-adrenergic effects with moderate beta 2 effects

Decreases renal perfusion

Decreases pulmonary vascular resistance

Beta 1/beta 2 adrenergic agonists

Isoproterenol

Treats torsades de pointes if treatment with magnesium fails

Treats ventricular tachycardia or fibrillation with Brugada syndrome

Anticonvulsants

These drugs include benzodiazepines; CNS (limbic and reticular formation) depressants by increasing the action of gamma aminobenzoic acid which is a major inhibitory neurotransmitter

Diazepam

Treats emotional irritability or seizures

Accumulates active metabolites that may prolong sedation

Lorazepam

Remains in the central nervous system longer than diazepam

Treatment of status epilepticus

Administered intramuscularly if vascular access cannot be obtained

Midazolam

Treatment of status epilepticus

Takes three times longer than diazepam to achieve results

Administered intramuscularly if vascular access cannot be obtained

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128175743000096

Cardiac Muscle; The Heart as a Pump and Function of the Heart Valves

John E. Hall PhD, in Guyton and Hall Textbook of Medical Physiology, 2021

What Causes the Long Action Potential and Plateau in Cardiac Muscle?

At least two major differences between the membrane properties of cardiac and skeletal muscle account for the prolonged action potential and the plateau in cardiac muscle. First, theaction potential of skeletal muscle is caused almost entirely by the sudden opening of large numbers offast sodium channels that allow tremendous numbers of sodium ions to enter the skeletal muscle fiber from the extracellular fluid. These channels are calledfast channels because they remain open for only a few thousandths of a second and then abruptly close. At the end of this closure, repolarization occurs, and the action potential is over within about another thousandth of a second.

In cardiac muscle, the action potential is caused by opening of two types of channels: (1) the samevoltage-activated fast sodium channels as those in skeletal muscle; and (2) another entirely different population ofL-type calcium channels (slow calcium channels), which are also calledcalcium-sodium channels. This second population of channels differs from the fast sodium channels in that they are slower to open and, even more importantly, remain open for several tenths of a second. During this time, a large quantity of both calcium and sodium ions flows through these channels to the interior of the cardiac muscle fiber, and this activity maintains a prolonged period of depolarization,causing the plateau in the action potential. Furthermore, the calcium ions that enter during this plateau phase activate the muscle contractile process, whereas the calcium ions that cause skeletal muscle contraction are derived from the intracellular sarcoplasmic reticulum.

The second major functional difference between cardiac muscle and skeletal muscle that helps account for both the prolonged action potential and its plateau is that immediately after the onset of the action potential, the permeability of the cardiac muscle membrane for potassium ionsdecreases about fivefold, an effect that does not occur in skeletal muscle. This decreased potassium permeability may result from the excess calcium influx through the calcium channels just noted. Regardless of the cause, the decreased potassium permeability greatly decreases the efflux of positively charged potassium ions during the action potential plateau and thereby prevents early return of the action potential voltage to its resting level. When the slow calcium-sodium channels do close at the end of 0.2 to 0.3 second, and the influx of calcium and sodium ions ceases, the membrane permeability for potassium ions also increases rapidly. This rapid loss of potassium from the fiber immediately returns the membrane potential to its resting level, thus ending the action potential.

Channelopathies in clinical medicine—cardiac arrhythmias

Julian O.M. Ormerod, Elijah R. Behr, in Clinical Molecular Medicine, 2020

9.3.2 Relation of the action potential to the surface electrocardiogram

The cardiac action potential has been investigated elucidated using traditional patch clamp and sharp electrode techniques in single cells and myocardial wedge preparations. However, the main clinical tool used to investigate patients is the 12-lead surface ECG. The power of the ECG to diagnose arrhythmia conditions is augmented by various tools including provocation with drugs (adrenaline or sodium-channel blockers such as ajmaline) or exercise, signal averaging, and prolonged recording (e.g., Holter monitoring) (Fig. 9.4).

What provides a graphic representation of all the action potentials occurring in the heart?

Figure 9.4. Correlation of ECG segments with repolarization and depolarization phases.

Source: Reprinted from Morita H, Wu J, Zipes DP. The QT syndromes: long and short. Lancet 2008;372(9640):750–63. https://doi.org/10.1016/S0140-6736(08)61307-0 [1] with permission.

Each vector (or lead) of the surface ECG reflects the summation of many individual action potentials. It is predominantly influenced by atrial and ventricular myocardium, as the specialized conduction tissue is comparatively very small in mass. The p wave represents atrial depolarization, the QRS complex begins with depolarization of ventricular muscle, and its width reflects the time taken for all muscle to depolarize. The ST segment corresponds to the plateau phase and shifts from the baseline (i.e., ST elevation or depression) reflect voltage gradients from epicardium to endocardium (with the largest contribution from the middle or M cells) and are commonly caused by ischemia (which tends to affect endocardium first due to the nature of coronary blood supply) or infarction. The T wave represents the repolarization phase.

Importantly, the QT interval (measured between the beginning of the earliest Q wave and the end of the latest T wave) does not solely measure the repolarization phase of the action potential, but in fact it measures the entire of the depolarization–repolarization period. Slowed conduction in cardiomyopathy, for example, leads to QRS prolongation (e.g., left bundle branch block), which also therefore prolongs the QT interval. LQTS can only be diagnosed in the absence of structural heart disease or coronary artery disease.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128093566000095

Basic Physiology and Hemodynamics of Cardiac Pacing

Frits W. Prinzen, ... Angelo Auricchio, in Clinical Cardiac Pacing, Defibrillation and Resynchronization Therapy (Fourth Edition), 2011

Electrical Activation During Sinus Rhythm

The cardiac action potential originates from the sinus node, located high in the right atrium (Fig. 9-1). Its cells depolarize spontaneously and initiate the spontaneous depolarization of action potentials at a regular rate from the sinus node. This rate depends on various conditions, such as atrial stretch and sympathetic activation, but is usually between 60 and 100 beats per minute (bpm) at rest. Myocytes are electrically coupled to each other through gap junctions. These structures consist of connexin molecules and allow direct intercellular communication.1 Gap junctions do not have a preferential direction of conduction, but because the action potential starts in the sinus node, it spreads from there through the atria. Evidence suggests specialized conduction pathways in the atrium, but their (patho) physiologic relevance is still disputed.

In the human heart, spreading of the action potential through the atria takes approximately 100 msec, after which the impulse reaches the AV node (see Fig. 9-1). In the normal heart the AV node is the only electrical connection between the atria and ventricles, because a fibrous ring (anulus fibrosus) is present between the remaining parts of the atria and ventricles. The AV nodal tissue conducts the electrical impulses very slowly; indeed, it takes approximately 80 msec for these impulses to travel from the atrial side to the ventricular side of the AV node. This delay between atrial activation and ventricular activation has functional importance because it allows optimal ventricular filling. Its slow conduction renders the AV node sensitive to impaired conduction and even complete conduction block, an important indication for ventricular pacing. As with other locations in the heart, conduction in the AV node has no preferential direction. Consequently, impulses can also be conducted retrogradely through the AV node, a condition that can occur when the ventricles are electrically stimulated.

From the AV node, the electrical impulse reaches the His bundle, the first part of the specialized conduction system of the ventricles called the Purkinje system. Within this system the electrical impulse is conducted approximately four times faster (3-4 m/sec) than in the working myocardium (0.3-1 m/sec). This difference results because Purkinje cells are longer and have a higher content of gap junctions.1-5

The intraventricular conduction system can be regarded as a trifascicular conduction system consisting of the right bundle branch (RBB) and two divisions of the left bundle branch (LBB). The RBB proceeds subendocardially along the right side of the interventricular septum until it terminates in the Purkinje plexuses of the right ventricle; the LBB also has a short subendocardial route.

It seems that the bifascicular vision of structure of the LBB is oversimplified.6 The general picture that now emerges is that the left ventricular Purkinje network is composed of three main, widely interconnected parts, depending on the anterior subdivision, the posterior subdivision, and a centroseptal subdivision of the left main bundle. This third medial or centroseptal division supplies the midseptal area of the left ventricle and arises from the LBB, from its anterior or posterior subdivision, or both. The fascicles continue in a network of subendocardially located Purkinje fibers. In the left ventricle, Purkinje fibers form a network in the lower third of the septum and free wall, which also covers the papillary muscles.4,7 In humans the bundles are present only underneath the endocardium, whereas other species (e.g., ox, sheep, goat) have networks of Purkinje fibers across the entire ventricular wall.8,9

It is important to note that the His bundle as well as the right and left bundle branches and their major tributaries are electrically isolated from the adjacent myocardium. The only sites where the Purkinje system and the normal working myocardium are electrically coupled are the Purkinje-myocardial junctions. The exits of the Purkinje system are located in the subendocardium of the anterolateral wall of the right ventricular (RV) and the inferolateral left ventricular (LV) wall4,10 (see Fig. 9-1). This area of exit corresponds with the area of the ventricular muscle that is activated earliest.2-5,11,12 The distribution of the Purkinje-myocardial junctions is spatially inhomogeneous, and the junctions themselves have varying degrees of electrical coupling.13

Endocardial activation of the right ventricle starts near the insertion of the anterior papillary muscle 10 msec after onset of LV activation11 (Fig. 9-2). After activation of the more apical regions, the activation of the ventricular working myocardium occurs predominantly from apex to base, both in the septum and in the LV and RV free wall. Further depolarization occurs centrifugally from endocardium to epicardium as well as tangentially.11,12 The earliest epicardial breakthrough occurs in the pretrabecular area in the right ventricle, from which there is an overall radial spread toward the apex and base. The last part of the right ventricle that becomes activated is the AV sulcus and pulmonary conus. Overall, the posterobasal area of the left ventricle (or an area more lateral) is the last part of the heart to be depolarized (see Fig. 9-2).

The time between arrival of the impulse in the His bundle and the first ventricular muscle activation is approximately 20 msec,4 whereas total ventricular activation lasts 60 to 80 msec, corresponding to a QRS duration of 70 to 80 msec.11 These numbers illustrate the important role of the Purkinje fiber system in the synchronization of myocardial activity. This role results from the system's unique propagation properties and its geometrically widespread distribution. During normal orthodromic excitation, fast propagation over long fibers, together with wide distribution of Purkinje-myocardial junctions, induces a high degree of coordination between distant regions of the myocardium.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9781437716160000096

Acquired long QT syndrome and sex hormones

Katja E. Odening, Alessandro Castiglione, in Sex and Cardiac Electrophysiology, 2020

Estrogen

Estrogen affects cardiac action potential duration and the associated long QT-related arrhythmogenesis through numerous different mechanisms. Several studies have shown that 17β-estradiol can inhibit the inward rectifier potassium current (IK1) and the delayed and rapid rectifier potassium currents IKs and IKr [113–115]. Estradiol inhibits IKr both directly by blocking the KCNH2/HERG channel and indirectly by increasing the transcription of the β-subunit KCNE2 [116]. Thereby, estradiol exerts a synergistic effect when combined with IKr-blocking drugs [117]. Similarly, the inhibition of IKs by estradiol is caused by a downregulation of the mRNA levels of KCNE1, the β-subunit to the ion channel conducting IKs [35].

Beyond the potassium currents, estradiol affects the regional expression and function of the L-type calcium channel (ICa,L) and the sodium–calcium exchanger (NCX). Studies in rabbits revealed a higher density of ICa,L and INCX in cardiomyocytes from the ventricular base than from the apex in female—but not in male—hearts, thus enhancing a proarrhythmic dispersion of repolarization. This regional difference could be reproduced by incubation of isolated ventricular cardiomyocytes from female hearts with estrogen: Here, an estrogen-induced upregulation of mRNA, proteins, and current densities was observed in basal but not in apical cardiomyocytes [118,119]. The increase of ICa,L was mediated by a genomic mechanism, e.g., an estrogen receptor binding to the promoter region of the gene encoding for the L-type Ca2+ channel [119]. By increasing ICa,L, estradiol enhances the propensity to develop EADs [51]. In summary, estradiol prolongs cardiac repolarization (particularly at slow heart rates) and increases the likelihood of drug-induced arrhythmias, exerting a proarrhythmic synergic effect with concomitant QT-prolonging drugs (Figs. 73.4 and 73.5).

What provides a graphic representation of all the action potentials occurring in the heart?

Figure 73.4. Sex hormone effects on cardiac ion channels.

−, decrease; +, increase; CaV1.2/CACNA1C, α-subunit of L-type calcium channels conducting ICa,L; DHT, testosterone; EST, estradiol; HERG/KCNH2, channel conducting the rapid delayed rectifier repolarizing potassium current IKr; Kir2.1/KCNJ2, channel conducting the inward rectifier potassium current IK1; Kv7.1/KCNQ1, channel conducting the slow delayed rectifier repolarizing potassium current IKs; PROG, progesterone; RyR2, ryanodine receptor 2; SERCA, sarcoplasmic reticulum calcium ATPase.

What provides a graphic representation of all the action potentials occurring in the heart?

Figure 73.5. Sex hormone effects on ion currents and action potential duration.

↑, increase; IKr, rapid delayed rectifier repolarizing potassium current; ↓, decrease; DHT;EST, estradiol; ICa,L, L-type calcium current; IK1, inward rectifier potassium current. Estrogen prolongs the ventricular action potential duration by inhibiting the repolarizing potassium currents IKr, IK, and IK1 and by enhancing the depolarizing ICa,L calcium current. Testosterone exerts opposite effects on the mentioned ion currents, causing a shortening of the myocardial action potential duration. A slighter shortening effect on the ventricular action potential duration is caused by progesterone, which inhibits ICa,L and enhances IKs.; IKs, slow delayed rectifier repolarizing potassium current; PROG, progesterone; testosterone.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128177280000735

Cardiac Function and Circulatory Control

Andrew R. Marks, in Goldman's Cecil Medicine (Twenty Fourth Edition), 2012

Electrical Cells

The heart is a muscular pump controlled by regular electrical discharges from specialized muscle cells in the conduction system (Chapter 61). The molecular basis for the electrical activity of the heart is the activation of specific ion-conducting channels (Fig. 52-1). Coordinated activation and inactivation of cardiac ion channels regulate the membrane potential of the cardiac cells, thereby resulting in a rapid sequence of depolarization followed by repolarization. This electrical activity, which is manifested on the body surface as the electrocardiogram (ECG), is known as the action potential, and it is responsible for activating the contraction of the cardiac muscle. At a typical heart rate of 70 beats per minute, the heart beats about 100,000 times per day, or 37 million beats a year, corresponding to 3 billion beats over a lifespan of 80 years. Failure to propagate the signal throughout the heart (e.g., heart block), or abnormal rhythms (arrhythmias) that are either too slow (bradycardia) or too fast (tachycardia), can result in death (Chapter 62).

FIGURE 52-1. Cardiac action potential and ion channels.

Myocardial contraction begins when sodium channels open and positively charged sodium ions flow into the cell and cause membrane depolarization (phase 0). During phases 1, 2, and 3, calcium ions flow into the cell through l-type calcium channels, while potassium flows out of the cell through voltage-gated potassium channels. These three phases correspond to the myocardial contraction, which corresponds to the QRS complex on the surface electrocardiogram (ECG). The sodium-potassium adenosine triphosphatase (NKA) helps return the system to its resting state.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9781437716047005522

Cardiac Action Potentials

GORDON M. WAHLER, in Heart Physiology and Pathophysiology (Fourth Edition), 2001

C. Phase 1—Transient Outward Current

Phase 1 of the cardiac AP is the transient and relatively small repolarization phase that immediately follows the upstroke of the AP. The size of phase 1 repolarization varies between species and also between different regions of the heart within a given species. Thus, APs recorded from the outer (epicardial) layer of ventricular cells display a more prominent phase 1, whereas APs recorded from the inner (endocardial) layer of ventricular cells display a small phase 1 repolarization (Liu et al., 1993). Phase 1 is also very large in Purkinje fibers and in atrial cells, but is largely absent in nodal cells.

Identification of the specific current or currents responsible for phase 1 repolarization has been controversial. It is now clear that the primary ion responsible for most of the phase 1 repolarization is K+ (Kenyon and Gibbons, 1979), although a Cl− component does appear to contribute to phase 1 repolarization (Bouron et al., 1991; see later).

Phase 1 repolarization is largely due to a transient outward current (Ito). Ito turns on rapidly with depolarization (i.e., beginning during the final portion of the AP upstroke) and is only active at very depolarized potentials; the threshold for activation is approximately –30 mV (Fig. 3). Thus, Ito has a characteristic transient shape—the rapid activation of this current is followed by inactivation during the AP plateau. Because of its voltage dependence and time course, Ito significantly overlaps (and opposes) the inward Ca2+ current (which is the primary depolarizing current during the plateau phase, see later).

What provides a graphic representation of all the action potentials occurring in the heart?

FIGURE 3. Transient outward current (Ito) recorded from a 21-day rat ventricular cell. (A) Original currents obtained upon stepwise depolarization from –80 mV. Ito activates rapidly and then inactivates. Currents were recorded in the presence of tetrodotoxin to eliminate INa and cadmium to eliminate overlapping ICa. (B) The current–voltage curve shows increasing activation of this current at voltages above approximately –30 mV (○). This current is the Ca2+-independent, 4-aminopyridine-sensitive component of Ito (i.e., Ito1) and is therefore blocked readily by 4-aminopyridine (4-AP, •).

This current (Ito) is actually composed of at least two separate currents (Ito1 and Ito2), which are carried through two physically distinct channels (Tseng and Hoffman, 1989). One of the currents (Ito1) is a K+ current that is independent of the internal Ca2+ concentration ([Ca2+]i) and is sensitive to the K+ channel blocker 4-aminopyridine (4-AP). This component of Ito is very similar to the IA current recorded in nerve fibers. The second component of Ito (Ito2) is Ca2+ dependent and less sensitive to 4-AP, but is more sensitive to another K+ channel blocker, tetraethylammonium ion (TEA+). It is thought that Ito2 is, at least in part, a Ca2+ -activated Cl− channel (Harvey, 1996).

Under physiological conditions, the first Ito component, Ito1, is by far the larger of the two components. Thus, irrespective of the exact nature of Ito2, the efflux of K+ (through Ito1 channels) appears responsible for the vast majority of phase 1 repolarization. The second component (Ito2) may become more important when intracellular levels of Ca2+ become too high. Thus, under Ca2+ overload conditions, Ito2 would be activated and shorten the AP duration, thereby indirectly abbreviating the duration of Ca2+ current, resulting in a reduced Ca2+ influx. Thus, activation of Ito2 by intracellular Ca2+ likely acts as a negative feedback mechanism to reduce calcium overload.

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780126569759500122

Which electrocardiogram finding can be used to measure heart rate?

For regular heart rhythms, heart rate can easily be estimated using the large squares (0.2s) on an ECG. Simply identify two consecutive R waves and count the number of large squares between them. By dividing this number into 300 (remember, this number represents 1 minute) we are able to calculate a person's heart rate.

What is the action potential of the heart?

The cardiac action potential is a measurement of the membrane potential waveform of the cardiac myocytes signifying the electrical activity of the cell during the contraction and relaxation of the heart. Specific ionic currents contribute to each phase of the cardiac action potential (see Fig.

What does the ECG wave tracing represent?

The recorded tracing is called an electrocardiogram (ECG, or EKG). A "typical" ECG tracing is shown to the right. The different waves that comprise the ECG represent the sequence of depolarization and repolarization of the atria and ventricles.

Where do action potentials originate in the heart?

The cardiac action potential originates from the sinus node, located high in the right atrium (Fig. 9-1). Its cells depolarize spontaneously and initiate the spontaneous depolarization of action potentials at a regular rate from the sinus node.